This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
0
0 to be allowed to set hO > 0; in the second, she "pays"
,(q) for i = 1, ... ,1 (whereas in the case of a private good, the market demand curve is identified by adding the individual demand curves horizontally). The inefficiency of private provision is often remedied by governmental intervention in the provision of public goods. Just as with externalities, this can happen not only through quantity-based intervention (such as direct governmental provision) but also through "price-based" intervention in the form of taxes or subsidies. For example, suppose that there are two consumers with benefit functions 4> I (x I + x 2) and 4>2(X I + X2)' where x, is the amount of the public good purchased by consumer i, and that qO > O. By analogy with the analysis in Section II.B, a subsidy to each consumer i per unit purchased of 5, = 4>'-/(qO) [or, equivalently, a tax of -4>'-,(qO) per unit that consumer i's purchases of the public good fall below some specified 10. The conclusion follows immediately if q' = O. So suppose instead that q' > O. Then since L, ;(q' )-c'(q') > 0 and L, <1>;(')- c'(·) is decreasing, any solution to (ll.e I) must have a larger value than q'. Note that, in contrast, if we are dealing with a public bad, so that <1>;(') < 0 and c'(') < 0, then the inequalities reverse and qO < q •.
SECTION
PUBLIC
level] faces each consumer with the marginal external effect of her actions and so generates an optimal level of public good provision by consumer i. Formally, if (XI' x2 ) are the competitive equilibrium levels of the public good purchased by the two consumers given these subsidies, and if p is the equilbrium price, then consumer i's purchases of the public good, X" must solve Max.,;,o 4>,(x, + XI) + s,x l - px" and so Xi must satisfy the necessary and sufficient first-order condition
Private provision leads to an insufficientltve! of a desirable public
good.
II.C:
GOODS
363
,--------------------------------------------------~~~~~~::
4>;U,
+ x2) + 5i S p, with equality
of Xi> O.
Substituting for 5 i , and using both condition (1I.C.4) and the market-clearing condition that x, + x2 = ii, we conclude that ii is the total amount of the public good in the competitive equilibrium given these subsidies if and only if 4>M)
+ 4>'-i(qO) S c'(ii),
with equality for some i if ii > O. Recalling (II.C.I) we see that ii = qQ. (Exercise II.C.1 asks you to extend this argument to the case where 1 > 2; formally, we then have a multilateral externality of the sort studied in Section 11.0.) Note that both optimal direct public provision and this subsidy scheme require that the government know the benefits derived by consumers from the public good (i.e., their willingness to pay in terms of private goods). In Section I I.E, we study the case in which this is not so.
Lindahl Equilibria Although private provision of the sort studied above results in an inefficient level of the public good, there is in principle a market institution that can achieve optimality. Suppose that, for each consumer i, we have a market for the public good "as experienced by consumer i." That is, we think of each consumer's consumption of the public good as a distinct commodity with its own market. We denote the price of this personalized good by Pi' Note that P, may differ across consumers. Suppose also that, given the equilibrium price each consumer i sees herself as deciding on the total amoullt of the public good she will COllsume, x" so as to solve
p,o.,
4>,(x,) - p,.·x,.
Max Xj~O
Her equilibrium consumption level sufficient first-order condition
p,..,
x,.· must therefore satisfy the necessary and with equality if x,.' > O.
(11.C.6)
The firm is now viewed as producing a bundle of 1 goods with a fixed-proportions technology (i.e., the level of production of each personalized good is necessarily the same). Thus, the firm solves Max q~O
(±
p:.q) - c(q).
iEt
The firm's equilibrium level of output q"
therefore satisfies the necessary and
364
CHAPTER
11:
EXTERNALITIES
AND
PUBLIC
SECTION
GOODS
~~~~~~~~~----------------------------------sufficient first-order condition I
L pro ~ c'(q**),
with equality if q** > O.
(11.C.?)
i='l
Together, (11.C.6), (ll.c.?), and the market-clearing condition that all i imply that
, L ¢,(q*') ~ c'(q*'),
j=
with equality if q" > O.
xr' = q"
for
(11.C.S)
1
Comparing (11.e.S) with (II.C.I), we see that the equilibrium level of the public good consumed by each consumer is exactly the efficient level: q" = q'. This type of equilibrium in personalized markets for the public good is known as a Lilldahl equilibrium, after Lindahl (1919). [See also Milleron (1972) for a further discussion.] To understand why we obtain efficiency, note that once we have defined personalized markets for the public good, each consumer, taking the pr~ce in her personalized market as given, fully determines her own level of consumption of the public good; externalities are eliminated. Yet, despite the attractive properties of Lindahl equilibria, their realism is questionable. Note, first, that the ability to exclude a consumer from use of the public good is essential if this equilibrium concept is to make sense; otherwise a consumer would have no reason to believe that in the absence of making any purchases of the public good she would get to consume none of it. II Moreover, even if exclusion is possible, these are markets with only a single agent on the demand side. As a result, price-taking behavior of the sort presumed is unlikely to occur. The idea that inefficiencies can in principle be corrected by introducing the right kind of markets, encountered here and in Section II.B, is a very general one. In particular cases, however, this "solution" mayor may not be a reali.s~ic ~ossibil~ty. We encounter this issue again in our study of multilateral externahtles m Section 11.0. As we shall see, these types of externalities often share many of the features of public goods.
11.0:
MULTILATERAL
piece of property, that much less is left to be dumped on others. 1l Depletable externalities therefore share the characteristics of our usual (private) sort of commodity. In contrast, air pollution is a nondepletable externality; the amount of air pollution experienced by one agent is not affected by the fact that others are also experiencing it. Nondepletable externalities therefore have the characteristics of public goods (or bads). In this section we argue that a decentralized market solution can be expected to work well for multilateral depletable externalities as long as well-defined and enforceable property rights can be created. In contrast, market-based solutions are unlikely to work in the nondepletable case, in parallel to our conclusions regarding public goods in Section 11.C. We shall assume throughout this section that the agents who generate externalities are distinct from those who experience them. This simplification is inessential but eases the exposition and facilitates comparison with the previous sections (Exercise 11.0.2 asks you to consider the general case). For ease of reference, we assume here that the generators of the externality are firms and that those experiencing the externality are consumers. We also focus on the special, but central, case in which the externality generated by the firms is homogeneous (i.e., consumers are indifferent to the source of the externality). {Exercise 11.0.4 asks you to consider the case in which the source matters.} We again adopt a partial equilibrium approach and assume that agents take as given the price vector p of L traded goods. There are J firms that generate the externality in the process of production. As discussed in Section 11.B, given price vector p, we can determine firm j's derived profit function over the level of the externality it generates,llj ;:: 0, which we denote by nJ(hJ). There are also I consumers, who have quasilinear utility functions with respect to a numeraire, traded commodity. Given price vector p, we denote by 4J,(h,) consumer i's derived utility function over the amount of the externality h, she experiences. We assume that nJ(') and 4J,{') are twice differentiable with nj(') < 0 and
11. D M ul tilateral Externalities Depletable Externalities In most cases, externalities are felt and generated by numerous parties. This is particularly true of those externalities, such as industrial pollution, smog caused by automobile use, or congestion, that are widely considered to be "important" policy problems. In this section, we extend our analysis of externalities to these multilateral settings. An important distinction can be made in the case of multilateral externalities according to whether the externality is depletable (or private, or rivalrous) or IlOlIdeplerable (or public, or lIonrivalrous). Depletable externalities have the feature that experience of the externality by one agent reduces the amount that will be felt by other agents. For example, if the externality takes the form of the dumping of garbage on people's property, if an additional unit of garbage is dumped on one II. Thus. the possibility of exclusion can be important for efficient supply of the public good, even though the use of an exclusion technology is itself inefficient (a Pareto opumal allocation cannot involve any exclusion).
EXTERNALITIES
We begin by examining the case of depletable externalities. As in Section 11.B, it is easy to see that the level of the (negative) externality is excessive at an unfettered competitive equilibrium. Indeed, at any competitive equilibrium, each firmj will wish to set the externality-generating activity at the level hj satisfying the condition nj(hn ~ 0,
with equality if hj > O.D
(11.0.1)
In contrast, any Pareto optimal allocation involves the levels (h~, ... , iiI' h;, ... , h~) 12. A distinction can also be made as to whether a depletable externality is allocable. For example, acid rain is depletable in the sense that the total amount of chemicals put into the air will fall somewhere. but it is not readily allocable because where it falls is determined by weather patterns. Throughout this section. we take depletable externalities to be allocable. The analytical implications of nonallocable depletable externalities parallel those of nondepletable ones. 13. The firms are indifferent about which consumer is affected by their externality. Therefore, the particular values of the individual ii/s are indeterminate. apart from the fact that L, hi = L;
h7.
365
366
CHAPTER
11:
AND
EXTERNALITIES
PUBLIC
SECTION
GOODS
that solve 14 I
J
I (Mii,) + I
Max I,.I ....• "JI~O
1=1
(li, ..... ~,)~o
J
S.t.
1!)(h)
):1
I i""l
(11.0.2)
/
h) =
Iii,. i""t
The constraint in (11.0.2) reflects the depletability of the externality: If ii, is increased by one unit, there is one unit less of the externality that needs to be experienced by others. Letting /1 be the multiplier on this constraint, the necessary and sufficient first-order conditions to problem (11.0.2) are
~ /1, with equality if ii~ > 0, i = I, ... , I,
and /1 ~ -1!j(h'j),
with equality if hi > 0, j = I, ... , J.
(11.0.3) (11.0.4)
Conditions (11.0.3) and (11.0.4), along with the constraint in problem (11.0.2), characterize the optimalleveis of externality generation and consumption. Note that they exactly parallel the efficiency conditions for a private good derived in Chapter 10, conditions (10.0.3) to (10.0.5), where we interpret - nj(') as firm j's marginal cost of producing more of the externality. If well-defined and enforceable property rights can be specified over the externality, and if I and J are large numbers so that price taking is a reasonable hypothesis, then by analogy with the analysis of competitive markets for private goods in Chapter 10, a market for the externality can be expected to lead to the optimal levels of externality production and consumption
Nondepletable Externalities We now move to the case in which the externality is nondepletable. To be specific, assume that the level of the externality experienced by each consumer is L) II), the total amount of the externality produced by the firms. In an unfettered competitive equilibrium, each firm j's externality generation hj again satisfies condition (I 1.0.1). In contrast, any Pareto optimal allocation involves externality generation levels (h~, ... , h'J) that solve Max ("I ..... "JI~O
I
J
i=t
I
1!)(h j ).
(11.0.5)
i""l
This problem has necessary and sufficient first-order conditions for each firm j's optimal level of externality generation, of
h;,
/
I <paL) h'j) ~
-1!j(hj),
with equality if hj > 0.
MULTtLATERAL
EXTERNALITIES
Condition (11.0.6) is exactly analogous to the optimality condition for a public good, condition (\ 1.e.1), where -1!j(') is firmj's marginal cost of externality production. 1S By analogy with our discussion of private provision of public goods in Section II.C, the introduction of a standard sort of market for the externality will not lead here, as it did in the bilateral case of Section II.B, to an optimal outcome. The free-rider problem reappears, and the equilibrium level of the (negative) externality will exceed its optimal level. Instead, in the case of a multilateral nondepletable externality, a market-based solution would require personalized markets for the externality, as in the Lindahl equilibrium concept. However, all the problems with Lindahl equilibrium discussed in Section 1I.C will similarly afflict these markets. As a result, purely market-based solutions, personalized or not, are unlikely to work in the case of a depletable externality.lo In contrast, given adequate information (a strong assumption!), the government can achieve optimality using quotas or taxes. With quotas, the government simply sets an upper bound on each firm j's level of externality generation equal to its optimal level Iti. On the other hand, as in Section II.B, optimality-restoring taxes face each firm with the marginal social cost of their externality. Here the optimal tax is identical for each firm and is equal to t. = - L, <paL} h)) per unit of the externality generated. Given this tax, each firm j solves
which has the necessary and sufficient first-order condition nj(/I) ~ t., Given t.
in the depletable case.
/
11.0:
367
----------------------------------------------~~~
(11.0.6)
i= 1
14. The objective function in (1I.D.2) amounts to the usual difference between benefits and costs arising in the aggregate surplus measure. Note, to this effect, that -1tA') can be viewed as firm j's cost function for producing the externality.
with equality if h] > 0.
= - L,
A partial market-based approach that can achieve optimality with a nondepletable multilateral e.xternality involves specification of a quota on the lotallevel of the externality and distributIon of that number of tradeable externality permits (each permit grants a firm the right to ge~erate one unit of th~ externality). Suppose that h' = LJ hi permits are given to the firms, with firm j receiving hi of them. Let denote the equilibrium price of these permIts. Then each firmj's demand for permits, hi' solves Max'J~o (nJ(h J) + p:(hJ - hJ» and so satlsfi~" the necessary and sufficient first-order condition n;(hj) oS P:, with equality if hJ > O. In addttlOn, market clearing in the permits market requires that LJ hJ = h'. The competitive eqUlltbrlum In the market for permits then has price = - L, t/>;(h') and each firm j using hj permtts and so yields an optimal allocation. The advantage of this scheme relative to a strict quota method arises when the government has limited information about the n (.) functIOns and cannot tell which particular firms can efficiently bear the burden of extern:lity reductIOn, although it has enough information, perhaps of a statistical sort, to allow the computation of the optimal aggregate level of the externality, h'.
p:
p:
15. Recallihat the single firm's cost function c(·) in Section I I.e could be viewed as the aggregate cost functIOn of J separate profit-maximizing firms. Were we to explicitly model these J firms in SectIon t I.e, the optimality conditions for public good production would take exactly the form in (II.D.6) with c;{hj) replacing -nj{hj). I~. The public nature of the externality leads to similar free-rider problems in any bargaining Soiutlon. (See ExerCIse 11.0.6 for an illustration.)
368
CHAPTER
11:
EXTERNALITIES
AND
PUBLIC
GOODS
l1.E Private Information and Second-Best Solutions In practice, the degree to which an agent is affected by an externality or benefits from a public good will orten be known only to her. The presence of privately held (or asymmetrically held) information can confound both centralized (e.g., quotas and taxes) and decentralized (e.g., bargaining) allempts to achieve optimality. In this section, we provide an introduction to these issues, focusing for the sake of specificity on the case of a bilateral externality such as that studied in Section I LB. Following the convention adopted in Section I I.D, we shall assume here that the externalitygenerating agent is a firm and the affected agent is a consumer. (For a more general treatment of some of the topics covered in this section, see Chapter 23.) Suppose, then, that we can write the consumer's derived utility function from externality level h (see Section II.B for more on this construction) as (h, 'I), where 'I E R is a parameter, to be called the consumer's type, that affects the consumer's costs from the externality. Similarly, we let 7[(h, 8) denote the firm's derived profit given its type 8 E R. The actual values of 0 and 'I are privately ohsert'ell: Only the consumer knows her type 'I, and only the firm observes its type 0. The ex ante likelihoods (probability distributions) of various values of 8 and 'I are, however, publicly known. For convenience, we assume that 0 and 'I are independently distributed. As previously, we assume that 7[(h, 0) and (h, III are strictly concave in h for any given values of 0 and 'I.
Decell!ralized Bargaining Consider the decentralized approach to the externality problem first. In general, bargaining in the presence of bilateral asymmetric information will not lead to an efficient level of the externality. To see this, consider again the case in which the consumer has the right to an externality-free environment, and the simple bargaining process in which the consumer makes a take-it-or-Ieave-it offer to the firm. For simplicity, we assume that there are only two possible levels of the externality, 0 and > 0, and we focus on the case of a negative externality in which externality level [" relative to the level 0, is detrimental for the consumer and beneficial for the firm (the analysis is readily applied to the case of a positive externality). It is convenient to define b(O) = 7[(ii, 0) - 7[(0,0) > 0 as the measure of the firm's benefit from the externality-generating activity when its type is O. Similarly, we let c(11l = I/J(O, 'I) -I/J(ij, III > 0 give the consumer's cost from externality level ii. In this simplified selling, the only aspects of the consumer's and firm's types that matter are the values of band c that these types generate. Hence, we can focus directly on the various possible values of band c that the two agents might have. Denote by G(b) and F(c) the distribution functions of these two variables induced by the underlying probability distributions of 0 and 'I (note that, given the independence of 0 and 'I, b and C are independent). For simplicity, we assume that these distributions have associated density functions g(b) and ftc), with g(b) > 0 and ftc) > 0 for all b> 0 and (' > O. Since the consumer has the right to an externality-free environment, in the absence of any agreement with the firm she will always insist that the firm set h = 0 (recall that c > 0). However, in any arrangement that guarantees Pareto optimal outcomes for all values of hand c, the firm should be allowed to set h = ii whenever b > c.
I.
---
SECTION
11.E:
PRIVA.TE
INFORMATION
AND
SECOND-BEST
SOLUTIONS
369
---------------------------~~~~~~~~~~~~:: Now consider the amount that the consumer will demand from the firm when her cost is c in exchange for permission to engage in the externality-generating activity. Since the firm knows that the consumer will insist on h = 0 if there is no agreement, the firm will agree to pay the amount T if and only if b ;;" T. Hence, the consumer knows that if she demands a payment of T, the probability that the firm will accept her offer equals the probability that b ;;" T; that is, it is equal to I - G(T). Given her cost c > 0 (and assuming risk neutrality), the consumer optimally chooses the value of T she demands to solve Max
(I - G(T»(T - c).
(II.E.I)
T
The objective function of problem (I I.E.I) is the probability that the firm accepts the demand, multiplied by the net gain to the consumer when this happens (T - c). Under our assumptions, the objective function in (\I.E.I) is strictly positive for all T> c and equal to zero when T = c. Therefore, the solution, say P,', is such that P,' > c. But this implies that this bargaining process must result in a strictly positive probability of an inefficient outcome, since whenever the firm's benefit b satisfies c < b < P,', the firm will reject the consumer's offer, resulting in an externality level of zero, even though optimality requires that h = ii."·'·
Quotas alld Taxes Just as decentralized bargaining will involve inefficiencies in the presence of privately held information, so too will the use of quotas and taxes. Moreover, as originally noted by Weitzman (1974), the presence of asymmetrically held information causes these two policy instruments to no longer be perfect substitutes for one another, as they were in the model of Section 11.8.'· To begin, note that given 8 and 'I, the aggregate surplus resulting from externality level h (we return to a continuum of possible externality levels here) is I/J(h, ~) + 7[(h,O). Thus, the externality level that maximizes aggregate surplus depends in general on the realized values of (0, Ill. We denote this optimal value by the function hO(O, 'I). Figure II.E.I depicts this optimum value for two different pairs of parameters, (0', ~') and (0",
,n.
Suppose, first, that a quota level of h is fixed. The firm will then choose the level of the externality to solve Max
7[(h, (I)
';;'0
s.t.
h~
ft.
Denote its optimal choice by h«h, 8). The typical effect of the quota will he to make 17. No.e .he similarity between problem (II.E.I) and the monopolist's problem studied in Section 12.B. Here the consumer's inability to discriminate among firms of different types leads her optImal oITer to be one that yields an inefficient outcome.
18. We could, of course, also consider the outcomes from other, perhaps more elaborate, bargaining procedures. In Chapter 23, however, we shall study a result due to Myerson and Salterthwai.e (1983) that implies that no bargaining procedure can lead to an efficient outcome ror all values of band c in this setting. 19. T~e discussion that follows also has implications for the relative advantages of quantity-
versus pTlce-based control mechanisms in organizations.
370
C HAP T E R
1 1:
EXT ERN ALIT I E SAN D
PUB Lie
SEC T ION
GOO D S
1 1 . E:
P R I V ATE
I N FOR MAT ION
AND
SEC 0 N D - 8 EST
SOL UTI 0 N S
371
----------------------------------------------------------~ ~-----------------------------------------------------~~~ _ ilq,(h,
Loss in / Aggregate Surplus ,/ ilq,(h, ~') --il-hFigure II.E.l
The surplusmaximizing aggregale externality level for two different pairs or parameters, (0', ~') and
iJh
~)
ilq,(h,~)
----ah
,/ /""
"
iJq,(h, ilh
r,
+ n(h'(h, 0), 0) -
r·'c';." (on(h, 0) + iJ
J"•.•,
iJh
iJh
This loss is represented by the shaded region in Figure II.E.2 for a case in which the quota is set equal to = hO(9,1i), the externality level that maximizes social surplus when 0 and" each take their mean values, 9 and Ii [the dashed lines in the figure are the graphs of iJn(h,9)/iJh and -iJ
r,
,:.0 Denote its optimal choice by h'(t,O). The loss in aggregate surplus from the tax relative to the optimal outcome for types (0, II) is therefore given by
"c,.•, (iJn(h, 0) iJ
i'c',.,
iJh
iJh
Its value is depicted by the shaded region in Figure II.E.3 for the same situation as in Figure II.E.2, but now assuming that a tax is set at t = -iJq,(hO(9, Ii), ii)/iJh, the value that results in the maximization of aggregate surplus when (0, II) = (9, Ii). Note that under a tax, as under a quota, the level of the externality is responsive to changes in the marginal benefits of the firm but not to changes in the marginal costs of the consumer. Which of these instruments, quota or tax, performs better? The answer is that it depends. Imagine, for example, that" is a constant, say equal to Ii. Then, for 0 such that the benefits of the externality'S use to the firm are high, a quota will typically miss the optimal externality level by not allowing the externality to increase above the quota level. On the other hand, because a fixed tax rate t does not reflect any
ilh_ on(h,O)
iJh
Figure II.E.3 (right) h'(I,0)
h'(O, ~)
increasing marginal costs of the externality to the consumer at higher externality levels, for such values of 0 the tax may result in excess production of the externality. Intuitively, when the optimal externality level varies little with 0, we expect a quota to be better. Figure II.E.4(a), for example, depicts a case in which the marginal cost to the consumer of the externality is zero up to some point h- and infinite thereafter. In this case, by setting a quota of h = h-, we can maximize aggregate surplus for any value of (0, ,,), but no tax can accomplish this. A tax would have to be very high to guarantee that with probability one the externality level fixed by the firm is not larger than h·. But if so, the resulting externality level would be too low most of the time. In contrast, in Figure !I.E.4(b) we depict a case in which the marginal cost to the consumer of the externality is independent of the level of h. In this case, a tax equal to this marginal cost (t = t·) achieves the surplus-maximizing externality level for all (0,11), but no quota can do so. lf we take the expected value of aggregate surplus as our welfare measure, we therefore see from these two examples that either policy instrument may be preferable, depending on the circumstances. 2o (Exercise !I.E.I asks you to provide a full analysis for a linear-quadratic example.) Note also that the bargaining procedure we have discussed will not result in optimality in either case depicted in Figure I I. E.4. 21 Thus, we have here two cases in which either a quota or a tax performs better than a particular decentralized outcome. 22 20. In Chapter 13, we discuss in greater detail some of the issues that arise in making welfare
comparisons in sellings with privately held information. There we shall justify the maximization of expected aggregate surplus in this partial equilibrium selling as a requirement of a notion of ex ante Pareto optimality for the two agents. See also the discussion in Section 23.f. 21. Striclly speaking, our previous discussion of bargaining assumed only two possible levels of the externality, while here we have a continuum of levels. This difference is not important. The inefficiency of the bargaining procedure previously studied would hold in this continuous environment as well.
22. We should emphasize that in these two examples other bargaining procedures will perform beller than che procedure involving a take-it-or-Ieave-it offer by the consumer. for example, if a take-or-Ieave offer is made by the firm, then full optimality results in both of these cases because the lype of the consumer is known with certainty. The conclusion of our discussion is therefore a qualitative one: With asymmetric information, it is difficult to make very general assertions about the relative performance of centralized versus decentralized approaches.
Figure 11.E.2 (Iell)
The loss in aggregate surplus under a quota for types (0, ~).
iln(h,0) /
(O·,~·).
the actual level of the externality much less sensitive to the values of 0 and" than is required by optimality. The firm's externality level will be completely insensitive to ~. Moreover, if the level of the quota h is such that iJn(h,O)/iJh > 0 for all 0, we will have It'(fr, 0) = for every O. The loss in aggregate surplus arising under the quota for types (0, II) is given by
~)
The loss in aggregate surplus under a tax for types (0, ~).
372
C HAP T
eR
1 1:
eXT ERN A LIT I e SAN 0
P U 8 Lie
SEC T ION
GOODS
::~~~~~~~~~~~-------------------------------------o(h,~)
P R I V ATE
I N FOR MAT ION
AND
SEC 0 NO· B EST
M ore General Policy Mechanisms
iih Figure 11.E.4
ch on(h,O")
-?-,,-
"n(h,O")
,1h ,)n(h,O')
iJh h* = h"(O",~)
1 1 • E:
= h'(O',~)
(b)
In Exercise II.E.2, you are asked to extend the analysis just given to a case with two firms (j = 1,2) generating an externality, where the tWO firms are identical except possibly for their realized levels of OJ. The exercise illustrates the importance of the degree of correlation between
the O;'s for the relative performance of quotas versus taxes. In comparing a uniform quota policy versus a uniform tax policy (" uniform" here means that the two firms face the same quota or tax rate), the less correlated the shocks across the firms, the belter the tax looks. The reason is not difficult to discern. With imperfect correlation, a uniform tax has a benefit that is not achieved with a uniform quota: It allows for the individual levels of externalities generated to be responsive to the realized values of the O/s, Indeed, with a uniform tax, the production of the total amount of externality generated is always efficiently distributed across the two firms, The presence of multiple generators of an externality also raises the possibility that a market for tradeable emissions permits could be created, as discussed at the end of Secuon 11.0. This simple addition to the quota policy can potentially eliminate the inefficient distribution of externality generation across different generators that is often a feature of a quota policy. In particular, suppose that instead of simply giving each firm a quota level, we now give them tradeable externality permits entitling them to generate the same number of units of the externality as in the quota, Suppose also that each firm would always fully use its quota if no trade was possible. Then trade must result in aC leasc as large a value of aggregate surplus as the simple quota scheme for any realization of the firms' and consumer's types, because we still get the same total level of emissions and we can never get a trade between firms that lowers aggregate profits." Of course, the same bargaining problems that we studied above can prevent a fully efficient distribution of externality generation from arising; but if the firms know each others' values of OJ or are numerous enough to act competiliv~ly in the market for these rights, then we can expect a distribution of the total externahty generation that is efficient across generators. In fact, in the case where the statistical distribution of costs among the firms is known but the particular realizations for individual firms are not known, this type of policy can achieve a fully optimal outcome.
23. Note, however, that the assumption that the externalities generated by the different firms are perfect substitutes to the consumer is crucial to this conclusion. If this is not true, then the reallocation of externality generation can reduce aggregate surplus by lowering the well·being of the agents affected by the externality.
Two cases in which a quota or tax maximizes aggregale surplus for every realization of O. (a) Quota" = h* maximizes aggregate surplus for all O. (b) Tax c = c· maximizes aggregate surplus for all O.
SOL UTI 0 N
S
373
------------------------------------------------------------The tax and quota schemes considered above are, as we have seen, completely unresponsive to changes in the marginal costs of the externality to the affected agent (the consumer in this case). It is natural to wonder whether any other sorts of schemes can do better, perhaps by making the level of the externality responsive to the consumer's costs. The problem in doing so is that these benefits and costs are unobservable, and the parties involved may not have incentives to reveal them truthfully if asked. For example, suppose that the government simply asks the consumer and the firm to report their benefits and costs from the externality and then enforces whatever appears to be the optimal outcome based on these reports. In this case, the consumer will have an incentive to exaggerate her costs when asked in order to prevent the firm from being allowed to generate the externality. The question, then, is how to design mechanisms that control these incentives for misreporting and, as a consequence, enable the government to achieve an efficient outcome. This problem is studied in a very general form in Chapter 23; here we confine ourselves to a brief examination of one well-known scheme. Return to the case in which there are only two possible levels of the externality, and ii. Can we design a scheme that achieves the optimal level of externality generation for every realization of b (the firm's benefit from the externality) and c (the consumer's cost)? We now verify that the answer is "yes." Imagine the government setting up the following revelation mechanism: The firm and the consumer are each asked to report their values of band c, respectively. Let 6 and c denote these announcements. For each possible pair of announcements (6, c), the government sets an allowed level of the externality as well as a tax or subsidy payment for each of the two agents. Suppose, in particular, that the government declares that it will set the allowed externality level h to maximize aggregate surplus given the announcements. That is, h = ii if and only if 6 > c. In addition, if externality generation is allowed (i.e., if h = ii), the government will tax the firm an amount equal to c and will subsidize the consumer with a payment equal to 6. That is, if the firm wants to generate the externality (which it indicates by reporting a large value of b), it is asked to pay the externality's cost as declared by the consumer; and if the consumer allows the externality (by reporting a low value of c) she receives a payment equal to the externality's benefit as declared by the firm. In fact, under this scheme both the firm and the consumer will tell the truth, so that an optimal level of externality generation will, indeed, result for every possible (b, c) pair. To see this, consider the consumer's optimal announcement when her cost level is c. If the firm announces some 6 > c, then the consumer prefers to have the externality-generating activity allowed (she does 6 - c better than if it is prevented). Hence, her optimal announcement satisfies {: < 6; moreover, because any such announcement will give her the same payoff, she might as well announce the truth, that is, c = c < 6. On the other hand, if the firm announces 6 S c, the consumer prefers to have the externality level set to zero. Hence, she would like announce C~ 6; and again, because any of these announcements will give her the same payoff, she may as well announce the truth, that is, {: = c ~ 6. Thus, whatever the firm's announcement, truth-telling is an optimal strategy for the consumer. (Formally, telling the truth is a weakly dominant strategy for the consumer in the sense studied
o
A P PEN D I X
A:
NON CON V E X I TIE SAN D
THE
THE 0 R Y
0 F
EXT ERN A LIT I E S
375
---------------------------------------------------------------- ,---------------------------------------------------------------374
CHAPTER
11:
EXTERNALITIES
AND
PUBLIC
GOODS
in Section 8.B. In fact, it is the consumer's only weakly dominant strategy; see Exercise 11.E.3.) A parallel analysis yields the same conclusion for the firm.
.,(h)
= ,h"" -"
Exercise 11.E.4: Show that in the tax-subsidy part of the mechanism above we could add, without affecting the mechanism's truth-telling or optimality properties, an additional payment to each agent that depends in an arbitrary way on the other agent's announcement. The scheme we have described here is an example of the Groves-Clarke mechanism [due to Groves (1973) and Clarke (1971); see also Section 23.C] and was originally proposed as a mechanism for deciding whether to carry out public good projects. Some examples for the public goods context are contained in the exercises at the end of lhe chapter. The Groves-Clarke mechanism has two very attractive features: it implements the optimal level of the externality for every (b, c) pair, and it induces truth-telling in a very strong (i.e., dominanl strategy) sense. But the mechanism has some unattractive features as well. In particular, it does not result in a balanced budget for the government: The government has a deficit equal to (b - c) whenever b > c. We could use the flexibility offered by Exercise II.E.4 to eliminale this deficit for all possible (b, c), but then we would necessarily create a budget surplus and therefore a Pareto inefficient outcome for some values of (b, c) (not all units of the numeraire will be left in the hands of the firm or the consumer). In fact, this problem is unavoidable with this type of mechanism: If we want to preserve the properties that, for every (b, c), truth-telling is a dominant strategy and the optimal level of exlernalily is implemented, then we generally cannot achieve budget balance for every (b, c). In Chapler 23 we discuss this issue in greater detail and also consider other mechanisms that can, under certain circumstances, get around the problem. (Sec also Exercise II.E.5 for an analysis in which budget balance is required only on average.)
APPENDIX A: NONCONVEXITIES AND THE THEORY OF EXTERNALITIES
Throughout this chapter, we have maintained the assumption that preferences and production sets are convex, leading the derived utility and profit functions we have considered to be concave. With these assumptions, all the decision problems we have studied have been well behaved; they had unique solutions (or, more generally, convex-valued solutions) that varied continously with the underlying parameters of the problems (e.g., the prices of the L traded commodities or the price of the externality if a market existed for it). Yet, this is not a completely innocent assumption. In this appendix, we present some simple examples designed to illustrate that externalities may themselves generate nonconvexities, and we comment on some of the implications of this fact. We consider here a bilatenll externality situation involving two firms. We suppose that firm I may engage in an externality-generating activity that affects firm 2's production. The level of externality generated by firm I is denoted by h, and firm j's profits conditional on the production of externality level hare ItJ(h) for j = 1,2. It is perfectly natural to assume that It l (·) is concave: The level h could, for example,
Flgur.".AA.' The derived profit function of firm 2 (the externality recipienl) in Example II.AA.I when a + fJ > I.
be equal to firm I's output. 24 As Examples II.AA.I and II.AA.2 illustrate, however, this may not be true of firm 2's profit function. Example 11.AA.1: Positive Externalities as a Source of Increasing Returns. Suppose that firm 2 produces an output whose price is I, using an input whose price, for simplicity, we also take to equal I. Firm 2's production function is q = h' z', where IX, PE [0, I). Thus, the externality is a positive one.2S Note that, for fixed h, the problem of firm 2 is concave and perfectly well behaved. Given a level of h, the maximized profits of firm 2 can be calculated to be 1t2(h) = yh"11 -", where y > 0 is a constant. In Figure II.AA.I, we represent 1t 2(h) for P> I - IX. We see there that firm 2's derived profit function is not concave in /1; in fact, it is convex. This reflects the fact that if we think of the externality h as an input to firm 2's production process, then firm 2's overall production function exhibits increasing returns to scale because IX + P> I. • Example 11.AA,2: Negative Externalities as a Source of Nonconvexities. In Example II.AA.I, the nonconvexity in firm 2's production set, and the resulting failure of concavity in its derived profit function, were caused by a positive externality. In this example the failure of concavity of firm 2's derived profit function is the result of a negative externality. Suppose, in particular, that Iti(h) $ 0 for all h, with strict inequality for some h, and that firm 2 has the option of shutting down when experiencing externality level h and receiving profits of zero. 26 In this case, the function 1t2(') can never be concave 24. Note also Ihal we may well have n,(Ir) < 0 for some levels h ~ 0 because n,(h) is firm I's maximized profit conditional on producing eXlernality level h (and so shulling down is not possible if h > 0). 25. More generally, we could think that there is an industry composed of many firms and that Ihe externalily is produced and felt by all firms in Ihe industry (e.g., h could be an index, correia led with outpul, of accumulated know-how in the industry). Externalities were first studied by Marshall (1920) in Ihis context. See also Chipman (1970) and Romer (1986). 26. In the more typical case of a multilaleral externality, the abilily of affected parties to shut down in this manner orten depends Oil whelher the externalilY is deplelable. In the case of a nondcplctable externality, such as air pollution. affected firms can always shut down and receive zero profits. In contrasl, in Ihe case of a depletable externality (such as garbage), where ,,)(h) reHects firmj's profits when it individually absorbs h units of the externality, the absorption of the externality may itself require the usc of some inpuls (e.g., land to absorb garbage). Indeed, were this not the case for a deplelable eXlernality, the exlernality could always be absorbed in a manner that creates no social costs by allocating all or the externality to a firm that shuts down.
----
376_ _ C_ HAP T E_ R _ 1': ERN A_ LIT E SAN 0 _PUB Lie _ __ _ _EXT __ __ _I _ ___ __ _ _GOODS __________________
",
A E F E A E N C ES
~
Flgur. 11 .AA.2
(\
concave Over he [0, 00].
over all Ii E [0, if_), a point originally noted by Starrett (1972). The reason can be seen in Figure II.AA.2: If nz<-) were a strictly decreasing concave function, then it would have to become negative at some level of h (see the dashed curve), but n2(') must be nonnegative if firm 2 can always choose to shut down. _ The failure of n,(·) to be concave can create problems for both centralized and decentralized solutions to the externality problem. For example, if property rights over the externality are defined and a market for the externality is introduced in either Example II.AA.I or Example II.AA.2, a competitive equilibrium may fail to exist (even assuming that the two agents act as price takers). Firm 2's objective function will not be concave, and so its optimal demand may fail to be well defined and continuous (recall our discussion in Section IO.e of the equilibrium existence problems caused by nonconvexities in firms' cost functions). In contrast, taxes and quotas can, in principle, still implement the optimai outcome despite the failure of firm 2's profit function to be concave because their use depends only on the profit function of the externality generator (here, firm I) being well behaved. In practice, however, nonconvexities in firm 2's profit function may create problems for these centralized solutions as well. Example II.AAJ illustrates this point. Example I1.AA.3: Externalities as a Source of Multiple Local Social Optima. It is, in principle, true that if the decision problem of the generator of an externality is concave, then the optimum can be sustained by means of quotas or taxes. But if n,{') is not concave, then the aggregate surplus function n, (h) + n2 (h) may not be concave and, as a result, the first-order conditions for aggregate surplus maximization may suffice only for determining local optima. In fact, as emphasized by Baumol and Oates (1988), the nonconvexities created by externalities may easily generate situations with mUltiple local social optima, so that identifying a global optimum may be a formidable task. Suppose, for example, that the profit functions of the two firms are n,(h) =
{~
n2(h) =
{~(I
and
for h :s; I for h > I
-
h)2
'v--{('" "): " + "
earn zero profits for any level of the externality, then its derived profit function 1t,(h) cannot be
for h :s; I for h > 1.
" , __ (n.(O), niO») = (0,2)
"-,
If the recipient o[ a negative externality can shut down and
",(h)
377
----------------------------------------------------~~~~~~
{('" ",): "J:S "J(h) [or j = I, Hor some h
"''''''
=
",(0) + ",(O)} Flgur.ll.AA.3
(.,(1),
",(I»
",
= (1,0)
, ", The function 1t2(') is not concave, something that the two previous examples have shown us can easily happen with externalities. The profit levels for the two firms that are attainable for different levels of h are depicted in Figure II.AA.3 by the shaded set {(It" 1t2): Itj:S; ItJ(h) for j = 1,2 for some h > 0) (note that this definition allows for free disposal of profits). The social optimum has h = 0 (joint profits are then equal to 2), in which case firm 2 is able to operate in an environment free from the externality. This can be implemented by setting a tax rate on firm I of t > I per unit of the externality. But note that the outcome h = I (implemented by setting a tax rate on firm I of t = 0) is a local social optimum: As we decrease h, it is not until h < ! that we get an aggregate surplus level higher than that at h = I. Hence, this latter outcome satisfies both the first-order and second-order conditions for the maximization of aggregate surplus (e.g., at this point, the marginal benefits of the externality exactly equal its marginal costs), and it will be easy for a social planner to be misled into thinking that she is at a welfare maximum. _
REFERENCES Arrow, K. J. (1969). The organization of economic activity: Issues pertinent to the choice of market versus non-market allocation. In Collected Papers of K. J. Arrow, Vol. 2 Cambridge, Mass.: Harvard University Press, 1983.
Baumol, W. J. (1964). External economies and second-order optimality conditions. American Economic Review 54: 368-72. Baumol. W. J.. and W. E. Oa'es. (1988). The Theory of Environmental PolJcy, 2nd. ed. New York: Cambridge University Press. Chipman, J. S. (1970). External economies of scale and competitive equilibrium. Quarterly Journal of Economics 84: 347-85. aarke, E. H. (1971). Multipart pricing of public goods. PublJc Choice 11: 17-33. Coase, R. (t960). The problem of social cosl. Journal of Law and Economics 1: 1-44. Groves, T. (1973). Incentives in 'earns. Econometrica 41: 617-31. Holmstrom, B., and J. Tirole. (1989). The theory of the firm. In Handbook of Induslrial Organization, edtled by R. Schmalensee and R. D. Willig. Amsterdam: North-Holland. !.afron', J.-J. (1988). Fundamenlals of Public Economics. Cambridge. Mass.: MIT Press. Lindahl. E. (1919). Die Gerechtigkeit der BeSleuring. Lund: Gleerup. [English translation: Just 'axation-a positive solution. In Classics in the Theory of Public Finance, edited by R. A. Musgrave and A. T. Peacock. London: Macmillon, 1958.] Marshall, A. (1920). Principles of Economics. London: Macmillan.
The set of possible profit pairs (7[10 ~,) in Example II.AAJ exhibits mUltiple local maxima of aggregate surplus 1t,(h) + .,(h).
378
C HAP T E R
1 1:
EXT ERN .. LIT I E S
.. N D
PU8 L IC
GOO D S
-------------------------------------------------------------------~ Meade, J. (1952). External economies and diseconomies in a competitive situation. Economic )our"aI62: 54-67. Milleron. J..c. (1972). Theory of value with public goods: A survey article. Journal of Economic Theory 5: 419-77. Myerson, R.o and M. Satterthwaite. (1983). Efficient mechanisms for bilateral trading. Journal of Economic T/uory 29: 265-81. Pigou, A. C. (1932). nfe Economics of Welfare. London: Macmillan. Romer. P. (1986). Increasing returns and long-run growth. Journal of Political Economy 94: 1002-36. Samuelson, P. A. (1954). The pure theory of public expenditure. Review of Economics and Statistics 36: 387-89. Samuelson, P. A. (1955). Diagrammatic exposition of a pure theory of public expenditure. Review of £('onomics and Statistics 37: 350-56. Starrett. D. A. (1972). Fundamental non-convexities in the theory of externalities. Journal of Economic Theory 4: 180-99. Viner. J. (1931). Cost curves and supply curves. ZeilschriJt fur NaJionaliikonomie 111: 23-46. Weitzman, M. (1974). Prices vs. quantities. Review of Economic Studies 41: 477-91.
EXERCISES
II.B.1 8 (M. Weitzman) On Farmer Jones' farm, only honey is produced. There are two ways to make honey: with and without bees. A bucket full of artificial honey, absolutely indistin· guishable from the real thing, is made out of I gallon of maple syrup with one unit of labor. If the same honey is made the old·fashioned way (with bees), k total units of labor are required (including bee·keeping) and b bees are required per bucket. Either way, Farmer Jones has the capacity to produce up to H buckets of honey on his farm. The neighboring farm, belonging to Smith, produces apples. If bees are present, less labor is needed because bees pollinate the blossoms instead of workers doing it. For this reason, c bees replace one worker in the task of pollinating. Up to A bushels of apples can be grown on Smith's farm. Suppose that the market wage rate is w, bees cost P. per bee, and maple syrup costs Pm per gallon. If each farmer produces her maximal output at the cheapest cost to her (assume the output prices they face make maximal production efficient), is the resulting outcome efficient? How does the answer depend on k, b, c, w, p" and Pm? Give an intuitive explanation of your result. Up to how much would Smith be willing to bribe Jones to produce honey with bees? What would happen to efficiency if both farms belonged to the same owner? How could the government achieve efficient production through taxes? II.B.2C
Consider the two-consumer externality problem studied in Section II.B, but now assume that Consumer 2's derived utility function over the externality level h and her wealth available for commodity purchases w, takes the form ¢,(h, w,). Assume that ¢,(h, w,) is a twice-differentiable, strictly quasiconcave function with o¢,(h, w,)/ow, > 0 and, for simplicity, that we have a positive externality so that o¢,(h, w,)/oh > O. (a) Set up the Pareto optimality problem as one of choosing h and a wealth transfer T to maximize consumer I's welfare subject to giving consumer 2 a utility level of at least ii,. Derive the (necessary and sufficient) first-order condition characterizing the optimal levels of hand T, say h' and T'. (b) Imagine that consumer I could purchase h on an externality market. Let P. be the price per unit, and let h(p., w,) be consumer 2's demand function for h. Express the wealth effect iJh(p., w,)/ow, in terms of first-order and second-order partial derivatives of consumer 2's utility function.
EXERCISES
379
-------------------------------------------------------------(c) Derive the comparative statics change in the Pareto optimal level of the externality h' (for a given ii,) with respect to a differential increase dw, > 0 in consumer 2's wealth. Show that if consumer 2's demand for the externality, derived in (b), is normal at price fi, = [iJ¢,(h', w, - r)/oh]/ [o¢,(h', r)/ow,] and wealth level w, = 1" [i.e., if oh(fi., w,)/ow, > 0], then a marginal increase in consumer 2's wealth w, causes the Pareto optimal level of the externality II' to increase. (Similarly, in the case of a negative externality, if consumer 2's demand for reductions in the externality is a normal good, then when consumer 2 becomes wealthier, the Pareto optimal level of the externality declines.)
w, -
w, -
II.B.3" Consider the optimal Pigouvian tax identified in Section II.B for the two-consumer externality problem studied there. What happens if, given this tax, the two consumers are able to bargain with each other? Will the efficient level of the externality still result? What about with the optimal quota? II.B.4 8 Consider again the two-consumer externality problem studied in Section I LB. Suppose that consumer 2 can take some action, sayee R, that affects the degree to which she is affected by the externality, so that we now write her derived utility function as ¢,(h, e) + K',. To fix ideas,let h be a negative externality, and suppose that o'¢,(h, e)/ohoe > 0, so that increases in e reduce the negative effect of the externality on the margin. Suppose that both hand e can in principle be taxed or subsidized. Should e be taxed or subsidized in the optimal tax scheme? Why or why not? II.B.5" Suppose that at fixed input prices of wa firm produces output with the differentiable and strictly convex cost function c(q, h), where q ;;, 0 is its output level (whose price is P > 0) and h is the level of a negative externality generated by the firm. The externality affects a single consumer, whose derived utility function takes the form ¢(h) + w. The actions of the firm and consumer do not affect any market prices. (a) Derive the first-order condition for the firm's choice of q and h. (b) Derive the first-order conditions characterizing the Pareto optimal levels of q and h. (c) Suppose that the government taxes the firm's output level. Show that this cannot restore efficiency. Show that a direct tax on the externality can restore efficiency. (d) Show, however, that in the limiting case where h is necessarily produced in fixed proportions with q, so that h(q) = ~q for some ~ > 0, a tax on the firm's output can restore efficiency. What is the efficiency-restoring tax? 11.e.1 A Consider the model discussed in Section II.C, in which J consumers privately purchase a public good. Identify per-unit subsidies 5" .. . ,5" such that when each consumer i faces subsidy rate 5" the total level of the public good provided is optimal. 11.e.2A Consider the model discussed in Section II.C, in which J consumers privately purchase a public good. Show that a per-unit subsidy on the firm's output (paid to the firm) can also restore efficiency. II.C.3c Reconsider the Ramsey tax problem from Exercise 10.E.3, but now suppose that the government can also provide a public good Xo that can be produced from good I at cost c(xo). However, the government must still balance its budget (including any expenditures on the public good). Consumer i's utility function now takes the form Xli + LJ., ¢,,(Xlh xo). Derive and interpret the conditions characterizing the optimal commodity taxes and the optimal level of the public good. How do the two problems of Ramsey taxation and provision of the public good interact? 11.0.1 8 (M. Weitzman) First-year graduate students are a hard-working group. Consider a typical class of J students. Suppose that each student i puts in hi hours of work on her classes. This effort involves a dis utility of hfl2. Her benefits depend on how she performs relative to her peers and take the form ¢(h;/h) for all i, where h = (1/l)LI hi is the average number of hours put in by all students in the class and ¢(.) is a differentiable concave function, with
380
CHAPTER
11:
EXTERNALITIES
AND
PUBLIC
GOODS
=-~~~~~~~----------------------q,'(') > 0 and lim, _ 0 q,'(h) = 00. Characterize the symmetric (Nash) equilibrium. Compare it
with the Pareto optimal symmetric outcome. Interpret. II.D.2" Consider a setting with 1 consumers. Each consumer i chooses an action h, E R •. Consumer i's derived utility function over her choice of h and the choices of other consumers takes the form q"(h,, L, h,) + w" whereI identical firms that act as price takers. The price of their output is p, and the prices of their inputs are unaffected by their acti~ns. Suppo~e that partial equilibrium analysis is valid and that the aggregate demand for their product IS given by the function x(p). The industry is characterized by "learning by doing." in that each firm's total cost of producing a given level of output is declining in the level of total Industry output; that is, each firm j has a twice-differentiable cost function of the form c(~J' Q.) f~r Q = Lj
O and (I/")c.. + 2c,Q+ I/cQQ > 0 for 1/ = I and J. Compare the equilibrium and optimal industry output levels. Interpret. What tax or subsidy restores efficiency? 11.1).4" Reconsider the nondeplctable externality example discussed in Section 11.0, but now assume that the externalities produced by the J firms are not homogeneous, In particular, suppose that if h" ... ,II, are the firms' externality levels, then con.'um.er j's derive~ utility is given by "',{h" ... , II,) + W, for each j = I, .. , ,I. Compare the eqUlhbrium and effiCient levels of II" . .. , h,. What tax/subsidy scheme can restore efficiency? Under what condition should each firm face the same tax/subsidy rate?
11.058 (The problem of the commons) Lake Ec can be freely accessed by fishermen. The cost of sending a boat out on the lake is r > O. When b boats are sent out onto the lake, f(b) fish are caught in total [so each boat catches f(b)/b fish], where f'(b) > 0 and j"(b) < 0 at all b 0, which is unaffected by the level of the catch from Lake Ec. (a) Characterize the equilibrium number of boats that are sent out on the lake. (b) Characterize the optimal number of boats that should be sent out on the lake. Compare
EXERCISES
(b) Show that as 1 increases, the equilibrium level of 0 declines. Also show that 0 = O.
Iim,_~
t1.D.7 C Individuals can build their houses in one of two neighborhoods, A or B. It costs to build a house in neighborhood A and c. < C A to build in neighborhood B. Individuals care about the prestige of the people living in their neighborhood. Individuals have varying levels of prestige, denoted by the parameter O. Prestige varies between 0 and I and is uniformly distributed across the population. The prestige of neighborhood k (k = A, B) is a function of the average value of 0 in that neighborhood, denoted by 0,. If individual i has prestige parameter 0 and builds her house in neighborhood k, her derived utility net of building costs is (I + O)(! + 0,) - c,. Thus, individuals with more prestige value a prestigous neighborhood more. Assume that C A and c. are less than I and that fA - f. E(!, I).
CA
(8) Show that in any building-choice equilibrium (technically, the Nash equilibrium of the simultaneous-move game in which individuals simultaneously choose where to build their house) both neighborhoods must be oceupied.
(b) Show that in any equilibrium in which the prestige levels of the two neighborhoods differ, every resident of neighborhood A must have at least as high a prestige level as every resident of neighborhood B; that is, there is a cutoff level of 0, say 0, such that all types (/
r..
(a) Identify the best quota for a planner who wants to maximize the expected value of aggregate surplus. (Assume the lirm must produoe an amount exactly equal to the quota.) (b) Identify the best tax t· for this same planner. (c) Compare the two instruments: Which is beller and when?
this with your answer to (a). (c) What per-boat fishing tax would restore efficiency? (d) Suppose that the lake is instead owned by a single individual who can choose how many boats to send out. What level would this owner choose? II.D,68 Suppose that there is a piece or land that is affected adversely by a~ e~ternality produced by a single firm, The firm's derived profit runction ror the externahty IS ,,(h) = , + IIh - II we have 0 < I in equilibrium and too much of the externality is generated.
381
----------------------------------------------------------------------
t1.E.2 C Extend the model in Exercise II.E.t to the case of two producers. Now let a"l(h" O,)/vh = P- bit, + 0, for i = 1,2. Let "" = E[O, 0,]. Calculate and compare the optimal quotas and taxes. How does the choice depend on ",,? II,E.3 B Show that truth-telling is the consumer's only weakly dominant strategy in the (Groves-Clarke) revelation mechanism studied in Section I I.E. II.E.4A In text. tt.E.5 B Suppose that the government is considering building a public project. The cost is K. There are 1 individuals indexed by i. Individual i's privately known benelit from the project is bi' The government's objective is to maximize the expected value of aggregate surplus. Derive the extension of the Groves-Clarke mechanism discussed in Section I I.E for this case. Can you construct your scheme so that the government balances its budget on average (over all realizations of the b,'s)? tt.E.6 B Extend Exercise I t.E.5 to the case in which there are N possible projects," = I, ... , N, with individual i deriving a (privately known) benelit of b,(n) from project n.
382
CHAPTER
11:
EXTERNALITIES
AND
PUBLIC
C
GOODS
~~~~--------------------------------------------------1l.E.7" Suppose that in the model of Section II.E the consumer's type ~ takes only one possible value, ~. We have seen in the text that in this case neither a quota nor a tax will maximize aggregate surplus for all realizations of 0 when the derived utility function 4>(h, ~) for the consumer has a4>(h, ~)Iah E (0, - 00). Show, however, that a variable tax per unit in which the total tax collected from the firm is 4>(h,~) when the level of the externality is h will maximize aggregate surplus for all values of 0 for any derived utility function 4>(h, ~).
Market Power
HAP
T
E
R
12
12.A Introduction In the competitive model, all consumers and producers are assumed to act as price takers, in effect behaving as if the demand or supply functions that they face are infinitely elastic at going market prices. However, this assumption may not be a good one when there are only a few agents on one side of a market, for these agents will often possess market power-the ability to alter profitably prices away from competitive levels. The simplest example of market power arises when there is only a single seller, a monopolist, of some good. If this good's market demand is a continuous decreasing function of price, then the monopolist, recognizing that a small increase in its price above the competitive level leads to only a small reduction in its sales, will find it worthwhile to raise its price above the competitive level. Similar effects can occur when there is more than one agent, but still not many, on one side of a market. Most often, these agents with market power are firms, whose fewness arises from nonconvexities in production technologies (recall the discussion of entry in Section 10. F). In this chapter, we study the functioning of markets in which market power is present. We begin, in Section 12.B, by considering the case in which there is a monopolist seller of some good. We review the theory of monopoly pricing and identify the welfare loss that it creates. The remaining sections focus on situations of oligopoly, in which a number of firms compete in a market. In Sections 12.C and 12.0, we discuss several models of oligopolistic pricing. Each incorporates different assumptions about the underlying structure of the market and behavior of firms. The discussion highlights the implicatons of these differing assumptions for market outcomes. In Section I2.C, we focus on static models of oligopolistic pricing, where competition is viewed as a one-shot, simultaneous event. In contrast, in Section 12.0, we study how repeated interaction among firms may affect pricing in oligopolistic markets. This discussion constitutes an application of the theory of repeated games, a subject that we discuss in greater generality in Appendix A. The analysis in Sections 12.B to 12.0 treats the number of firms in the market as 383
384
CHAPTER
12:
MARKET
exogeneously given. In reality, however, the number of active firms in a market is likely to be affected by factors such as the size of market demand and the nature of competition within the market. Sections 12.E and 12.F consider issues that arise when the number of active firms in a market is determined endogenously. Section 12.E specifies a simple model of entry into an oligopolistic market and studies the determinants of the number of active firms. It offers an analysis that parallels that considered in Section 10.F for competitive markets. Section 12.F returns to a theme raised in Chapter 10. We illustrate how the competitive (price-taking) model can be viewed as a limiting case of oligopoly in which the size of the market, and hence the number of firms that can profitably operate in it, grows large. In the model we study, an active firm's market power diminishes as the market size expands; in the limit, the equilibrium market price comes to approximate the competitive level. In Section 12.G, we briefly consider how firms in oligopolistic markets can make strategic precommitments to affect the conditions of future competition in a manner favorable to themselves. This issue nicely illustrates the importance of credible commitments in strategic settings, an issue we studied extensively in Chapter 9. In Appendix B, we consider in greater detail a particularly striking example of strategic precommitment to affect future market conditions, the case of entry deterrence through capacity choice. If you have not done so already, you should review the game theory chapters in Part II before studying Sections 12.C to 12.G (in particular, review all of Chapter 7, Sections 8.A to 8.0, and Sections 9.A and 9.B). An excellent source for further study of the topics covered in this chapter is Tirole (1988).1
12.B Monopoly Pricing In this section, we study the pricing behavior of a profit-maximizing monopolist, a firm that is the only producer of a good. The demand for this good at price p is given by the function x(p), which we take to be continuous and strictly decreasing at all p such that x(p) > 0. 1 For convenience, we also assume that there exists a price ji < OC! such that x(p) = 0 for all p ~ p.3 Throughout, we suppose that the monopolist knows the demand function for its product and can produce output level q at a cost of c(q). The monopolist's decision problem consists of choosing its price p so as to maximize its profits (in terms of the numeraire), or formally, of solving Max
- --
_--------------------------------------------~S~E~C~T~I:O~N~~1:2~.8~:-=M~O::N~O~P~O::L~Y~P~R~I:C~I~N~G~~3~8~5
POWER
px(p) - c(x(p».
(12.B.I)
I. See also Ihe survey by Shapiro (1989) for the topics covered in Sections I2.C, 12.0,
and 12.G. 2. Throughout this chapter we take a partial equilibrium approach; see Chapter 10 for a discussion of this approach. 3. This assumption helps to insure that an optimal solution to the monopolist's problem exists. (Sec Exercise 12.B.2 for an example in which the failure of this condition leads to nonexistence.)
An equivalent formulation in terms of quantity choices can be derived by thinking instead of the monopolist as deciding on the level of output that it desires to sell, q ~ 0, lettmg the price at which it can sell this output be given by the inverse demand junction pC-) = x- I (. ).4 Using this inverse demand function, the monopolist's problem can then be stated as Max
p(q)q - c(q).
(12.B.2)
.>:0
We shall focus our analysis on this quantity formulation of the monopolist's problem [identical conclusions could equally well be developed from problem (12.B.I)]. We assume throughout that p(.) and c(') are continuous and twice differe~tiable at all q ~ 0, that p(O) > c'(O), and that there exists a unique output level q E (0, (0) such that p(qO) = c'(qO). Thus, qO is the unique socially optimal (competitive) output level in this market (see Chapter 10). Under these assumptions, a solution to problem (12.8.2) can be shown to exist.' Given the differentiability assumed, the monopolist's optimal quantity, which we denote by qm, must satisfy the first-order condition 6
p'(qm) qm
+ p(qm)
~
c'(qm), with equality if qm > O.
(12.B.3)
The left-hand side of (12.B.3) is the marginal revenue from a differential increase in q. at ~m, which is equal to the derivative of revenue d[p(q)q]jdq, while the right-hand Side tS the. corresponding marginal cost at q"'. Since p(O) > c'(O), condition (12.B.3) can be satisfied only at qm > O. Hence, under our assumptions, marginal revenue must equal marginal cost at the monopolist's optimal output level: (l2.B.4) For the typical case in which p'(q) < 0 at all q ~ 0, condition (12.B.4) implies that we must have p(q") > c'(qm), and so the price under monopoly exceeds marginal cos~. Correspond~ngly, the monopolist's optimal output qm must be below the socially opumal (compettt.lve) output level qQ. The cause of this quantity distortion is the monopolist's recognition that a reduction in the quantity it sells allows it to increase the price charged on its remaining sales, an increase whose effect on profits is captured by the term p'(q")qm in condition (12.B.4). The welfare loss from this quantity distortion, known as the deadweight loss oj mOllopoly, can be measured using the change in Marshallian aggregate surplus
4. More precisely, 10 lake account of the fact that x(p) = 0 for more than one value of p, we take p(q) = Min {p: x(p) = q} at all q 2: O. Thus, P(O) = p, the lowest price at which x(p) = O. 5. In par,ticular, it follows from condition (12.B.3) and from the facts that p'(q) sO for all q 2: 0 and f(q) < c (q) for all q ~ qO, that the monopolist's optimal choice must lie in the compact set [0: q ]. Becau~ the objective function In problem (12.B.2) is continuous, a solution must therefore eXist (see Section M.F of the Mathematical Appendix). 6. Satisfaction .of first-order condition (12.B.3) is sufficient for q" to be an optimal choice if the obJ:ctl.ve funct~on of problem (12.B.2) is concave on [0, q0]. Note, however, that concavity of thiS objective funchon depends not only on the technology of the firm, as in Ihe competitive model, but also on the shape of the Inverse demand function. In particular, even with a convex cost Cunction the ~onopolist's profit function can violate this concavity condition if demand is a convex functio~ of prtce.
386
CHAPTER
12:
MARKET
POWER
~~~~~~~~~~~----------------------------------[p(q)
+ p'(q)q] c'(q) Flgur. 12.B.1 (teft)
\----"""7f-'-- p(q) : p for all q
The monopoly solution and welfare loss when p'(') < O. Figure 12.B.2 (right)
q": q'
(sec Section 10.E),
r
The monopoly solution when p'(q) : 0 for all q
--
SECT tON
[pes) - c'(s)]
ds > 0,
where q" is the socially optimal (competitive) output level. Figure 12.B.1 illustrates the monopoly outcome in this case. The ~onopolist's quantity qm is determined by the intersection of the graphs of marginal revenue p'(q)q + p(q) and marginal cost c'(q). The monopoly. price p(qm) can then be determined from the inverse demand curve. The deadweight welfare loss IS equal to the area of the shaded region. Note from condition (12.B.4) that the monopoly quantity distortion is absent in the special case in which p'(q) = 0 for all q. In this case, wh~re p(q) eq~als so~e constant p at all q > 0, the monopolist sells the same quantity as a price-taking competitive firm because it perceives that any increase in i~s pri~ ab~ve the competitive price pcauses it to lose all its sales. 7 Figure 12.B.2 depicts thiS special case. Example 12.B.1: Monopoly Pricing with a Linear Inverse Dem~nd . Function and Constant Returns to Scale. Suppose that the inverse demand function In a monopolized market is p(q) = a - bq and that the monopolist's cost function is c(q) = cq, where (/ > c 2': 0 [so that p(O) > c'(O)) and b > O. In this case, the objective function of the monopolist's problem (12.B.2) is concave, and so condition (12.B.4) is b?th necessary and sufficient for a solution to the monopolist's pro~lem. From ,condition (12.B.4), we can calculate the monopolist's optimal quantity. and price t? be qm : (0 _ c)(2b and pm = (a + c)(2. In contrast, the socially optimal (competitive) output level and price are q' = (a - c)(b and po = p(qO) = c, • Although we do not discuss these issues here, we point out that the behavi~ral distortions arising under monopoly are not limited to pricing decisions. (ExerCises 12.B.9 and 12.B.10 ask you to investigate two examples.) The monopoly quantity distortion is fundamentally linked to the fact that if the monopolist wants to increase the quantity it sells, it must lower its price on all its existing sales. In fact, 7. This inverse demand function arises, for example. when each consumer i has quasilinear
preferences of the form u,(q,) + m, with u,(q,) = pq" where q, is consu,:,er j's ~onsumpti~n of the good under study and m, is his consumption of the numeraire commodity. [Stnctly speakong, wIth these preferences we now have a multi valued demand correspondence rather than a demand function, but p(') is nevertheless a function as before.]
L
STATIC
MODELS
OF
if the monopolist were able to perfeclly discriminate among its customers in the sense that it could make a distinct offer to each consumer, knowing the consumer's preferences for its product, then the monopoly quantity distortion would disappear. To see this formally, let each consumer i have a quasilinear utility function of the form u;(q;) + m; over the amount q, of the monopolist's good that he consumes and the amount m; that he Consumes of the numeraire good, and normalize .,(0) = O. Suppose that the monopolist makes a take-it-or-leave-it offer to each consumer i of the form (q;, 7;), where q; is the quantity offered to consumer i and 7; is the total payment that the consumer must make in return. Given offer (q;, 7;), consumer i will accept the monopolist's offer if and only ifu;(q;) - 7; 2': O. As a result, the monopolist can extract a payment of exactly u;(q;) from consumer i in return for q, units of its product, leaving the consumer with a surplus of exactly zero from consumption of the good. Given this fact, the monopolist will choose the quantities it sells to the I consumers (q" ... , q,) to solve
,
q
Jq~
12,C:
Max
L
14'10 ...••,12:0
'''''I
u,(q;) - c(Lq,)·
(12.8.5)
Note, however, that any solution to problem (12.8.5) maximizes the aggregate surplus in the market, and so the monopolist will sell each consumer exactly the socially optimal (competitive) quantity. Of course, the distributional properties of this outcome would not be terribly attractive in the absence of wealth redistribution: The monopolist would get all the aggregate surplus generated by its product, and each consumer i would receive a surplus of zero (i.e., each consumer i's welfare would be exactly equal to the level he would achieve if he consumed none afthe monopolist's product). But in principle, these distributional problems can be corrected through lump-sum redistribution of the numeraire. Thus, the welfare loss from monopoly pricing can be seen as arising from constraints that prevent the monopolist from charging fully discriminatory prices. In practice, however, these constraints can be significant. They may include the costs of assessing separate charges for different consumers, the monopolist's lack of information about consumer preferences, and the possibility of consumer resale. Exercise 12.B.5 explores some of these factors. It provides conditions under which the best the monopolist can do is to name a single per-unit price, as we assumed at the beginning of Ihis section.
12,C Static Models of Oligopoly We now turn to cases in which more than one, but still not many, firms compete in a market. These are known as situations of oligopoly. Competition among firms in an oligopolistic market is inherently a setting of strategic interaction. For this reason, the appropriate tool for its analysis is game theory. Because this discussion constitutes our first application of the theory of games, we focus on relatively simple static models of oligopoly, in which there is only one period of competitive interaction and firms take their actions simultaneously. We begin by studying a model of simultaneous price choices by firms with constant returns to scale technologies, known as the Bertrand model. This model displays a striking feature: With just two firms in a market, we obtain a perfectly competitive outcome. Motivated by this finding, we then consider three alterations of this model that weaken its strong and ollen implausible conclusion: a change in the firm's strategy from choosing its price to choosing its quantity of output
OLtGOPOLY
387
388
CHAPTER
12:
MARKET
POWER
SECTION
12.C:
STATIC
MODELS
OF
OLIGOPOLY
389
----------------------------------------------------------------------- ---------------------------------------------------------------------(the Cournot model); the introduction of capacity constraints (or, more generally, decreasing returns to scale); and the presence of product differentiation.· One lesson of this analysis is that a critical part of game-theoretic modeling goes into choosing the strategies and payoff functions of the players. In the context of oligopolistic markets, this choice requires that considerable thought be given both to the demand and technological features of the market and to the underlying processes of competition. Unless otherwise noted, we restrict our attention to pure strategy equilibria of the models we study.
Tile Bertrand Model of Price Competition We begin by considering the model of oligcipolistic competition proposed by Bertrand (1883). There are two profit-maximizing firms, firms I and 2 (a duopoly), in a market whose demand function is given by x(p). As in Section IO.B, we assume that x(·) is continuous and strictly decreasing at all P such that x(p) > 0 and that there exists a ;; < 00 such that x(p) = 0 for all P 0, per unit produced. We assume that x(c) E (0, 00), which implies that the socially optimal (competitive) output level in this market is strictly positive and finite (see Chapter 10). Competition takes place as follows: The two firms simultaneously name their prices P, and P2' Sales for firm j are then given by
Xj(Pj, P.) =
l
X(Pi)
ifpj
tX(Pj)
if Pj = P.
o
ifpj> P•.
The firms produce to order and so they incur production costs only for an output level equal to their actual sales. Given prices Pi and P., firm j's profits are therefore equal to (Pj - c)xj(Pj, P.)· The Bertrand model constitutes a well-defined simultaneous-move game to which we can apply the concepts developed in Chapter 8. In fact, the Nash equilibrium outcome of this model, presented in Proposition l2.C.I, is relatively simple to discern. Proposition 12.C.1: There is a unique Nash equilibrium (pr, pn in the Bertrand
duopoly model. In this equilibrium, both firms set their prices equal to cost:
pr = p~ = c.
incurs losses. But by raising its price above c, the worst it can do is earn zero. Thus, these price choices could not constitute a Nash equilibrium. Now suppose that one firm's price is equal to c and that the other's price is strictly greater than c: Pj = c, P. > c. In this case, firm j is selling to the entire market but making zero profits. By raising its price a little, say to Pj = C + (p, - c)/2, firm j would still make all the sales in the market, but at a strictly positive profit. Thus, these price choices also could not constitute an equilibrium. Finally, suppose that both price choices are strictly greater than c: Pj > c, p, > c. Without loss of generality, assume that Pj :s; P•. In this case, firm k can be earning at most !(Pj - c)x(Pj)' But by setting its price equal to Pj - 0 for 0> 0, that is, by undercutting firm j's price, firm k will get the entire market and earn (Pj - " - c)x(Pj - 0). Since (Pj - C - c)x(Pj - r.) > HpJ - c)x(Pj) for small-enough c > 0, firm k can strictly increase its profits by doing so. Thus, these price choices are also not an equilibrium. The three types of price configurations that we have just ruled out constitute all the possible price configurations other than PI = P2 = c, and so we are done. _ The striking implication of Proposition 12.C.1 is that with only two firms we get the perfectly competitive outcome. In effect, competition between the two firms makes each firm face an infinitely elastic demand curve at the price charged by its rival. The basic idea of Proposition 12.C.1 can also be readily extended to any number of firms greater than two. [In this case, if firm j names the lowest price in the market, say p, along with J - I other firms, it earns (1/J)x(p).] You are asked to show this in Exercise 12.C.1. Exercise l2.C.I: Show that in any Nash equilibrium of the Bertrand model with J > 2 firms, all sales take place at a price equal to cost. Thus, the Bertrand model predicts that the distortions arising from the exercise of market power are limited to the special case of monopoly. Notable as this result is, it also seems an unrealistic conclusion in many (although not all) settings. In the remainder of this section, we examine three changes in the Bertrand model that considerably weaken this strong conclusion: First, we make quantit), the firms' strategic variable. Second, we introduce capacity constraillCS (or, more generally, decreasing returns to scale). Third, we allow for product differentiation.
Qualltity Competitioll (The Cournot Model)
Proof: To begin, note that both firms setting their prices equal to C is indeed a Nash equilibrium. At these prices, both firms earn zero profits. Neither firm can gain by raising its price because it will then make no sales (thereby still earning zero); and by lowering its price below c a firm increases its sales but incurs losses. What remains is to show that there can be no other Nash equilibrium." Suppose, first, that the lower of the two prices named is less than c. In this case, the firm naming this price 8. Section 12.0 studies a rourth variation that involves repeated interaction among firms. 9. Recall that we restrict attention to pure strategy equilibria here. See Exercise l2.C.2 for a consideration of mixed strategy equilibria. There you are asked to show that under the conditions assumed here, Proposition l2.C.1 continues to hold: p! = = c is the unique Nash equilibrium, pure or mixed, of the Bertrand model.
P;
Suppose now that competition between the two firms takes a somewhat different form: The two firms simultaneously decide how much to produce, ql and q2' Given these quantity choices, price adjusts to the level that clears the market, p(ql + q2), where p(.) = x- I (-) is the inverse demand function. This model is known as the Coumot model, after Cournot (1838). You can imagine farmers deciding how much of a perishable crop to pick each morning and send to a market. Once they have done so, the price at the market ends up being the level at which all the crops that have been sent are sold. IO In this discussion, we assume that p(.) is differentiable 10. One scenario that will lead to this outcome arises when buyers bid ror the crops sent that day (very much like sellers in the Bertrand model; see Exercise 12.C.S).
SEC T ION
390
C HAP T E R
1 2:
MAR K E T
POW E R
~~~~~--------------------------------------------------------with p'(q) < 0 at al1 q ~ O. As before, both firms produce output at a cost of c > 0 per unit. We also assume that p(O) > c and that there exists a unique output level qO E (0, OC!) such that p(qO) = c [in terms of the demand function x(·), qO = x(c)]. Quantity qO is therefore the social1y optimal (competitive) output level in this market. To find a (pure strategy) Nash equilibrium of this model, consider firm j's maximization problem given an output level ii, of the other firm, k # j: Max
(12.C.1)
p(qj + ii,)qj - cqj.
4};<:0
In solving problem (12.C.1), firmj acts exactly like a monopolist who faces inverse demand function p(qj) = p(qj + ii,). An optimal quantity choice for firm j given its rival's output ii, must therefore satisfy the first-order condition
p'(qj
+ ii,) qj + p(qj + ii,)!5: c,
with equality if q} >
o.
(12.C.2)
For each ii.. we let b/ii,) denote firmj's set of optimal quantity choices; b j(') is firm j's best-response correspondence (or function if it is single-valued). A pair of quantity choices (q!, is a Nash equilibrium if and only if qj E bj(q:) for k # j and j = 1,2. Hence, if (q!, is a Nash equilibrium, these quantities
qn
qn
must satisfy"
p'(q!
+ q!lq! + p(q! + q~) !5: c,
with equality if q! > 0
(I2.C.3)
p'(q!
+ q!lq! + p(q! + q!) !5: c,
with equality if q! > O.
(12.C.4)
and
It can be shown that under our assumptions we must have (q!, q!) » 0, and so conditions (12.C.3) and (12.C.4) must both hold with equality in any Nash equilibrium.' 2 Adding these two equalities tel1s us that in any Nash equilibrium we
must have
p'(q!
+ qn(q! ; q!) + p(q! + qn
= c.
(12.C.S)
Condition (12.C.S) al10ws us to reach the conclusion presented in Proposition 12.C.2. Proposition 12.C.2: In any Nash equilibrium of the Cournot duopoly model with cost c > a per unit for the two firms and an inverse demand function p(.) satisfying p'(q) < a for all q ~ a and pta) > c, the market price is greater than c (the competitive price) and smaller than the monopoly price. II. Nole Ihal Ihis melhod of analysis, which relies on Ihe use of first-order condilions 10 calculale best responses, differs from the method used in the analysis of the Berlrand model. The reason is that in the Bertrand model each firm's objective runction is discontinuous in its decision
variable. so that differential optimization techniques cannot be used. Fortunately, the determination of Ihe Nash equilibrium in the Berlrand modellurned out, nevertheless, 10 be quite simple. 12. To see this, suppose Ihat q~ = O. Condition (12C.3) then implies Ihal P(qi) S c. By condilion (12.CA) and the fact that p'(') < 0, Ihis implies Ihat were q; > 0 we would have p'(q;)q; + p(q;) < c, and so q; = O. Bullhis means that p(O) s c, contradicting the assumption that p(O) > c. Hence, we must have q! > O. A similar argument shows that q; > O. Note, however, Ihal this conclusion depends on our assumption of equal costs for the two firms. For example, a firm might set its output equal to zero if it is much less efficient than its rival. Exercise 12.C.9 considers some of the issues that arise when firms have differing costs.
1 2 • C:
5 TAT I C
MOO E L S
0 F
0 L I GOP 0 L y
391
------------------------------------------~~~~~~~= Proof: That the equilibrium price is above c (the competitive price) follows immediately from condition (I2.C.S) and the facts that q! + q! > 0 and p'(q) < 0 at all q ~ O. We next argue that (q! + q!l > q"', that is, that the equilibrium duopoly price p(q! + q!) is strictly less than the monopoly price p(q"'). The argument is in two parts. First, we argue that (q! + q!) ~ q"'. To see this, suppose that q'" > (q! + q!), By increasing its quantity to qj = q"' - q:, firmj would (weakly) increase the joint profit of the two firms (the firms' joint profit then equals the monopoly profit level, its largest possible level). In addition, because aggregate quantity increases, price must fall, and so firm k is strictly worse off. This implies that firm j is strictly better off, and so firm j would have a profitable deviation if q"' > (q! + q!), We conclude that we must have (q! + ~ q"'. Second, condition (l2.C.S) implies that we cannot have (q! + = q"' because then
qn
qn
in violation of the monopoly first-order condition (12.B.4). Thus, we must in fact have (q! + q!) > q"' . • Proposition 12.C.2 tells us that the presence of two firms is not sufficient to obtain a competitive outcome in the Cournot model, in contrast with the prediction of the Bertrand model. The reason is straightforward. In this model, a firm no longer sees itself as facing an infinitely elastic demand. Rather, if the firm reduces its quantity by a (differential) unit, it increases the market price by - p'(ql + q2)' If the firms found themselves jointly producing the competitive quantity and consequently earning zero profits, either one could do strictly better by reducing its output slightly. At the same time, competition does lower the price below the monopoly level, the price that would maximize the firms' joint profit. This occurs because when each firm determines the profitability of selling an additional unit it fails to consider the reduction in its rival's profit that is caused by the ensuing decrease in the market price [note that in firm j's first-order condition (12.C.2), only qj multiplies the term p'('), whereas in the first-order condition for joint profit maximization (q, + q2) does]. Example 12.C.I: Cournot Duopoly wit It a Linear Inverse Demand Function and COOlstant Returns to Scale. Consider a Cournot duopoly in which the firms have a cost per unit produced of c and the inverse demand function is p(q) = a - bq, with a > c ~ 0 and b > O. Recall that the monopoly quantity and price are q"' = (a - c)/2b and p" = (il + c)/2 and that the socially optimal (competitive) output and price are qO = (a - c)/b and po = p(qO) = c. Using the first-order condition (12.C.2), we find that firm j's best-response function in this Cournot model is given by
bj(q,) = Max {O, (a - c - bq,)/2b}. Firm I's best-response function b,(q2) is depicted graphically in Figure 12.c'1. Since b,(O) = (a - c)/2b, its graph hits the q, axis at the monopoly output level (a - c)/2b. This makes sense: Firm I's best response to firm 2 producing no output IS to produce exactly its monopoly output level. Similarly, since b,(q2) = 0 for all
392
C HAP T E R
1 2:
MAR K E T
-
POW E R
a-c
+ q,)q, {(q,. q,): p(q, + q,)q,
------------ {(q,.q,): p(q,
- cq, = n') - ("q, = W); W>
n'
2b
SEC T ,
0N
12•
c:
S TAT I C
MOO E L S
Flgur. 12.C.3
Firm. 1'5 ~st-responsc funct,on 10 the Cournot duopoly model of Example I 2.C. I.
equilibrium in the Cournot model.
Symmelric Joinl Monopoly Poinl Figure 12.C.2
3b (a)
q,
(b)
q2 ~ (a - c)/b, the graph of firm I's best-response function hits the q2 axis at the socially optimal (competitive) output level (a - c)/b, Again, this makes sense: If firm 2 chooses an output level of at least (a - c)/b, any attempt by firm I to make sales results in a price below c. Two isoprofit loci of firm I are also drawn in the figure; these are sets of the form (q" q2): p(q, + q2)q, - cq, = n} for some profit level n. The profit levels associated with these loci increase as we move toward firm 1's monopoly point (q" q2) = «a - c)/2b, 0). Observe that firm I's isoprofit loci have a zero slope where they cross the graph of firm I's best-response function. This is because the best response b,(ih) identifies firm I's maximal profit point on the line q, = ii2 and must therefore correspond to a point of tangency between this line and an isoprofit locus. Firm 2's best-response function can be depicted similarly; given the symmetry of the firms, it is located symmetrically with respect to firm I's best-response function in (q" q2)-space [i.e., it hits the q2 axis at (a - c)/2b and hits the q, axis at (a - c)/b]. The Nash equilibrium, which in this example is unique, can be computed by finding the output pair (q!, qt) at which the graphs of the two bestresponse functions intersect, that is, at which q! = b,(qf) and q! = b,(q!). It is depicted in Figure 12.C.2(a) and corresponds to individual outputs of qT = q! = H{a - c)/b], total output of Wa - c)/b], and a market price of p(qT + q!) = !(a + 2c) E (c, pM). Also shown in Figure 12.C.2(b) is the symmetric joint monopoly point (qm/2, qm/2) = «a _ c)/4b, (a - c)/4b). It can be seen that this point, at which each
Nash equilibrium in the Cournot duopoly model of Example 12.C.1.
393
Nonexistence of (pure strategy) Nash
Exercise 12.C.6: Verify the computations and other claims in Example l2.C.1.
a-c
L Y
q,
firm produces half of Ihe monopoly output of (a - c)/2b, is each firm's most profitable point on the q, = q, ray. _
4b
0 L I GOP 0
Figure 12.C.1
q,
4b
0f
---------------------------~~~~~~~~~~~~
Up 10 Ihis poinl we have nol made any assumptions aboul the quasiconcavilY in q. of each firmj's objeclive funclion in problem (12.C.1). Withoul quasiconcavilY of Ihese funcli~lls. however. a pure slralegy Nash equilibrium of Ihis quantilY game may nol exisl. For example. as happens in Figure 12.C.3. Ihe besl-response funclion of a firm lacking a quasiconcavc objeclive fUllclion may "jump." leading 10 Ihe possibilily of nonexislence. (Striclly speaking, for a silualion like Ihe one depicled in Figure 12.C.310 arise.lhe Iwo firms must have differenl cosl funclions; see Exercise 12.C.8.) Wilh quasiconcavity. we can use Proposilion 8.0.3 10 show Ihal a pure slralegy Nash equilibrium necessarily ex iSIs. Suppose now that we have J > 2 identical firms facing the same cost and demand functions as above. Letting QJ be aggregate output at equilibrium, an argument parallel to that above leads to the following generalization of condition (l2.C.S):
P'(QJ)~! + p(Q1) =
c.
(l2.C.6)
At one exlreme, when J = I, condition (12.C.6) coincides with the monopoly first-order condition that we have seen in Section 12.B. At the other extreme, we must have p(Q1J .... c as J .... 00. To see this, note that since QJ is always less than the socially optimal (competitive) quantity qO, it must be the case that p'(QJ )(QJ /J) .... 0 as J .... oc;. Hence, condition (12.C.6) implies that price must approach marginal cost as the number of firms grows infinitely large. This provides us with our first taste of a "competitive limit" result, a topic we shall return to in Section 12.F. Exercise 12.C.7 asks you to verify these claims for the model of Example 12.C.!. Exercise 12.C.7: Derive the Nash equilibrium price and quantity levels in the Cournot model with J firms where each firm has a constant unit production cost of c and the inverse demand function in the market is p(q) = a - bq, with a > c ~ 0 and b > O. Verify that when J = I, we get the monopoly outcome; that output rises and price falls as J increases; and that as J .... DC the price and aggregate output in the market approach their competitive levels. In contrast with the Bertrand model, the Cournot model displays a gradual reduction in market power as the number of firms increases. Yet, the "farmer sending
394
CHAPTER
12:
MARKET
---
POWER
crops to market" scenario may not seem relevant to a wide class of situations. After all, most firms seem to choose their prices, not their quantities. For this reason, many economists have thought that the Cournot model gives the right answer for the wrong reason. Fortunately, the departure from the Bertrand model that we study next offers an alternative interpretation of the Cournot model. The basic idea is that we can think of the quantity choices in the Cournot model as long-run choices of capacity, with the determination of price from the inverse demand function being a proxy for the outcome of short-run price competition given these capacity choices.
In many settings, it is natural to suppose that firms operate under conditions of eventual decreasing returns to scale, at least in the short run when capital is fixed. One special case of decreasing returns occurs when a firm has a capacity constraint that prevents it from producing more than some maximal amount, say Ii. Here we consider, somewhat informally, how the introduction of capacity constraints affects the prediction of the Bertrand model. With capacity constraints (or, for that matter, costs that exhibit decreasing returns to scale in a smoother way), it is no longer sensible to assume that a price announcement represents a commitment to provide any demanded quantity, since the costs of an order larger than capacity are infinite. We therefore make a minimal adjustment to the rules of the Bertrand model by taking price announcements to be a commitment to supply demand only up to capacity. We also assume that capacities are commonly known among the firms. To see how capacity constraints can affect the outcome of the duopoly pricing game, suppose that each of the two firms has a constant marginal cost of c > 0 and a capacity constraint of ii = ~x(c). As before, the market demand function x(·) is continuous, is strictly decreasing at all p such that x(p) > 0, and has x(c) > O. In this case, the Bertrand outcome p~ = p! = c is no longer an equilibrium. To see this, note that because firm 2 cannot supply all demand at price p! = c, firm I can anticipate making a strictly positive level of sales if it raises PI slightly above c. As a result, it has an incentive to deviate from p~ = c. In fact, whenever the capacity level ii satisfies ii < x(c), each firm can assure itself of a strictly positive level of sales at a strictly positive profit margin by setting its price below p(ij) but above c. This is illustrated in Figure 12.CA. In the figure, we assume that the lower-priced firm 2 fills the highest-valuation demands. By charging
p, = c
MODELS
a price PI E (c, p(ii)), firm I sells to the remaining demand at price p" making sales of x(I'.J - ii > O. Hence, with capacity constraints, competition will not generally drive price down to cost, a point originally noted by Edgeworth (1897).
Product Differentiation Calculation of demand in the presence of capacity constraints when the low-priced firm satisfies high-valuation demands first.
p(ti)
STATIC
Up to this point in our discussion, we have taken a firm's capacity level as exogenous. Typically, however, we think of firms as c/lOosing their capacity levels. This raises a natural question: What is the outcome in a model in which firms first choose their capacity levels and then compete in prices? Kreps and Scheinkman (1983) address this question and show that under certain conditions (among these is the assumption that high-valuation demands get served first when demand for a low-priced firm outstrips its capacity), the unique subgame perfect Nash equilibrium in this two-stage model is the Cournol outcome. This result is natural: the computation of price from the inverse demand curve in the Cournot model can be thought of as a proxy for this second-stage price competition. Indeed, for a wide range of capacity choices (iiI' ii 2), the unique equilibrium of the pricing subgame involves both firms setting their prices equal to p(iil + ii,) (see Exercise 12.C.II). Thus, this two-stage model of capacity choice/price competition gives us the promised reinterpretation of the Cournot model: We can think of Cournot quantity competition as capturing long-run competition through capacity choice, with price competition occurring in the short run given the chosen levels of capacity.
Figure 12.C.4
Demand Satisfied
12.C,
Determining the equilibrium outcome in situations in which capacity constraints are presenl can be tricky because knowledge of prices is no longer enough to determine each firm's sales. When the prices quoted are such that the low-priced firm cannot supply all demand at its quoted price, the demand for the higher-priced firm will generally depend on precisely who manages to buy from the low-priced firm. The high'priced firm will typically have greater sales if consumers with low valuations buy from the low· priced firm (in contrast with the assumption made in Figure 12.C.4) than if high-valuation consumers do. Thus, to determine demand functions for the firms, we now need to state a rarioning rule specifying which consumers manage to buy from the low-priced firm when demand exceeds its capacity. In fact, the choice of a rationing rule can have important effects on equilibrium behavior. Exercise 12.C.11 asks you to explore some of the features of the equilibrium outcome when the highest valuation demands are served first, as in Figure 12.C.4. This is the rationing rule that tends to give the nicest results. Yct, it is neither more nor less plausible than other rules, such as a queue system or a random allocation of available units among possible buyers.
Capacity Constraints alld Decreasing Returns to Scale
~YFirm2
SECTION
In the Bertrand model, firms faced an infinitely elastic demand curve in equilibrium: With an arbitrarily small price differential, every consumer would prefer to buy from the lowest-priced firm. Often, however, consumers perceive differences among the products of different firms. When product differentiation exists, each firm will possess some market power as a result of the uniqueness of its product. Suppose, for example, that there are J > I firms. Each firm produces at a constant marginal cost of c > O. The demand for firm j's product is given by the continuous function xj(Pj' P_j), where P _ j is a vector of prices of firm j's rivals. \3 In a setting of simultaneous price
p(' ) ij
X(PI)
'---y-l
Firm I's Sales
x,q
OF
OLIGOPOLY
395
------------------------------------------------------------
13. Note the departure from the Bertrand model: In the Bertrand model, X/Pl' P_j) is discontinuous at Pj = Minl#i Pl.
5 E C T ION
396
C HAP T E R
1 2:
MAR K E T
POW E R
.~~~------------------------------------Firm 2 "..
/Firml
.
•
o
Flgur. 12.C.S
\
The linear city.
M Consumers Uniformly Distributed on Segment
choices, each firmj takes its rivals' price choices P-i as given and chooses Pi to solve
buy from firm I. At these locations, p, + IZ < p, + c(l - z} (purchasing from firm I is better than purchasing from firm 2), and v - p, - Iz > 0 (purchasing from firm I is better than not purchasing at all). At location z" a consumer is indifferent between purchasing from firm I and not purchasing at all; that is, z, satisfies v - p, - tz, = O. In Figure 12.C.6(a}, consumers in the interval (z" Z2) do not purchase from either firm, while those in the interval (Z" I] buy from firm 2. Figure 12.C.6(b), by contrast, depicts a case in which, given prices PI and p" all consumers can obtain a strictly positive surplus by purchasing the good from one of the firms. The location of the consumer who is indifferent between the two firms is the point such that
p, + Ii = P2 + 1(1 - i)
i = I
'----y-''----y----'~
Buy From Firm I (a)
p,
PI
(12.0)
.
In general, the analysis of this model is complicated by the fact that depending on the parameters (v, c, c), the equilibria may involve market areas for the firms that do not touch [as in Figure 12.C.6(a)], or may have the firms battling for consumers in the middle of the market [as in Figure 12.C.6(b}]. To keep things as simple as possible here, we shall assume that consumers' value from a widget is large relative to production and travel costs, or more precisely, that v > c + 3c. In this case, it can be shown that a firm never wants to set its price at a level that causes some consumers not to purchase from either firm (see Exercise 12.CI3). In what follows, we shall therefore ignore the possibility of non purchase. Given p, and P2' let be defined as in (12.C.7). Then firm I's demand, given a pair of prices (p" P2), equals Mi when i E [0, I], M when > I, and 0 when i < 0.14 Substituting for i from (12.C. 7), we have
z
z
if p, > P2 if p,
E
+c
[p, - C, P2
+ c]
(12.C.8)
if p, < P, - r. By the symmetry of the two firms, the demand function of firm 2, x 2(p" P2), is if P, > p, Consumer purchase decisions given PI and P"
O~----~~----~l
+ p, 21
if P2
Flgur. 12.C.6
I
MOD E L S
or
Example I2.C2: The Lillear City Model oj Produci Differentiation. Consider a city that can be represented as lying on a line segment of length I, as shown in Figure 12.C.S. There is a continuum of consumers whose total number (or, more precisely, measure) is M and who are assumed to be located uniformly along this line segment. A consumer's location is indexed by Z E [0, I], the distance from the left end of the city. At each end of the city is located one supplier of widgets: Firm I is at the left end; firm 2, at the right. Widgets are produced at a constant unit cost of c > O. Every consumer wants at most I widget and derives a gross benefit of v from its consumption. The total cost of buying from firm j for a consumer located a distance d from firm j is Pi + td, where 1/2 > 0 can be thought of as the cost or disutility per unit of distance traveled by the consumer in going to and from firmj's location. The presence oftravel costs introduces differentiation between the two firms' products because various consumers may now strictly prefer purchasing from one of the two firms even when the goods sell at the same price. Figure 12.C.6(a) illustrates the purchase decisions of consumers located at various points in the city for a given pair of prices p, and P" Consumers at locations [0, ztl
z,
S TAT I C
z
p,
Note that as long as xi(c, p-;l > 0, firmj's best response necessarily involves a price in excess of its costs (Pi> c) because it can assure itself of strictly positive profits by setting its price slightly above c. Thus, in the presence of product differentiation, equilibrium prices will be above the competitive level. As with quantity competition and capacity constraints, the presence of product differentiation softens the strongly competitive result of the Bertrand model. A number of models of product differentiation are popular in the applied literature. Example 12.C.2 describes one in some detail.
0"
1 2 . C;
E
+c
[PI - C, PI
+ c]
(12.C.9)
if P2 < PI - r. Note from (12.C.8) and (12.C.9) that each firm j, in searching for its best response to any price choice P_i by its rival. can restrict itself to prices in the interval [P-i - C, p-J + c]. Any price Pi > P-i + c yields the same profits as setting Pi = P-i + c (namely. zero), and any price Pi < P-i - t yields lower profits than setting Pi = P-i - r (all such prices result in sales of M units). Thus, firm j's best
~ '------y--'
Buy From Firm I
No Buy From Purchase Firm 2 (b)
Buy From Firm 2
0 F
0 l I GOP 0 L
Y
397
----------------------------------------------------------------
14. Recall that the M consumers are unirormly distributed on the line segment, so i is the
rraction who buy from firm I.
398
CHAPTER
12:
MARKET
POWER
~~~~~~---------------------------------------------------response to
P_j
solves
(Pj - C)(I
Max
+ P-j -
M
Pj)
s.t. PjE [p_j - I,P_j
---
SEC T , 0 N
C
<0
ifpj=p_j-I
=0
if pjE(p_j - I, P_j
;::0
if Pj= P-j+!'
l
P-j+1
ifp_j:$c-I
~
if p_jE (c -I, C + 31)
+ P_j + c)/2
P _j
-
if
I
P_j
(12.C.11)
;::
C
(t2.C.12)
+ 31.
When P_j < C - I, firm j prices in a manner that leads its sales to equal zero (it cannot make profits because it cannot make sales at any price above c). When P_. > c + 31, firm j prices in a manner that captures the entire market. In the int~rmediate case, firmj's best response to P_j leaves both firms with strictly positive sales levels. Given the symmetry of the model, we look for a symmetric equilibrium, that is, an equilibrium in which pf = p! = p'. In any symmetric equilibriu~, p' = b.(p*). Examining (12.C.12), we see that this condition can be satisfied only on the middle case (note also that this is the only case in which both firms can have strictly positive sales, as they must in any symmetric equilibrium). Thus, P* must satisfy
p' =
c
c +2
(' +'
C
+ 3c
(right) The circular city model when J = 5.
p,
Firm 4
The essential features of the linear cily model can be extended to the case in which J > 2. In doing so, it is often most convenient for analytical purposes to consider instead a model of a circular cit)', so that firms can be kept in symmetric positions." In this model, which is due to Salop (1979), consumers are uniformly distributed along a circle of circumference I, and the firms are positioned at equal intervals from one another. Figure l2.C.8 depicts a case where J = 5. Models like the linear and circular city models are known as spalial models of product differentiation because each firm is identified with an "address" in product space. More generally, we can imagine firms' products located in some N-dimensional characteristics space, with consumers' "addresses" (their ideal points of consumption) distributed over this space. Spatial models share the characteristic that each firm competes for customers only locally, that is. solely with the firms offering similar products. A commonly used alternative to spatial formulations, in which each product competes instead for sales with all other products, is the representati", COIlsumer model introduced by Spence (1976) and Dixit and Stiglitz (1977). In this model, a representative consumer is postulated whose preferences over consumption of the J products (x" ... , xJ) and a numeraire good m take the quasilinear form u(m, X" ... , XJ) =
G( t
f(X j
») + m,
i<'
+ I.
In this Nash equilibrium, each firm has sales of M /2 and a profit of 1M /2. Note that as I approaches zero, the firms' products become completely undifferentiated and the equilibrium prices approach c, as in the Bertrand model. In the other direction, as the travel cost I becomes greater, thereby increasing the differentiation between the firms' products, equilibrium prices and profits increase. Figure 12.C.7 depicts the best-response functions for the two firms (for prices greater than or equal to c) and the Nash equilibrium. As usual, the Nash equilibrium lies at the intersection of the graphs of these best-response functions. Note that there are no asymmetric equilibria here. Matters become more complicated when v < c + 31 because firms may wish to set prices at which some consumers do not want to purchase from either firm. One can show, however, that the equilibrium just derived remains valid as long as v", c + it. In contrast, when v < C + I, in equilibrium the firms' market areas do not touch (the firms are like "loc~1 monopolists"). In the intermediate case where v E [c + I, C + il], firms are at a "kink" in the~r demand functions and the consumer at the indifferent location f receives no surplus from h,s purchase in the equilibrium. Exercise 12.C.14 asks you to investigate these cases.
399
Best-response
+C
1(/ + p' + c),
P* = c
0 l , GOP 0 l Y
functions and Nash equilibrium in the linear city model when v> C + 31.
and so
L
0 F
Figure 12.C.8
+ I)
Solving (12.C.11), we find that each firm j's best-response function is
b(p_j) =
MOD E l S
+ I].
problem is
1
S TAT' C
b(p,)
(12.C.10)
The necessary and suflicient (Kuhn-Tucker) first-order condition for this
l+p_j+c-2pj
c:
Figure 12.C.7 (left)
c + 3c
2t
12.
where both G(') and f(·) are concave.'· Normalizing the price of the numeraire to be I, the first-order conditions for the representative consumer's maximization problem are
G'( t
i< ,
f(Xi»)J'(X j ) = Pj
for j = I.... , J.
(I2.C.13)
These first-order conditions can be inverted to yield demand functions xi(p" ... , PJ) for j = I .... , J, which can then be used to specify a game of simultaneous price choices." An important variant of this representative consumer model arises in the limiting case where we have many products, each of which constitutes a small fraction of the sales in the overall market. In the limit, we can write the representative consumer's utility function as G(Jf(xj) dj) + m. where Xj is now viewed as a function of the continuous index variable j. 15. In the segment [0,1], only with two firms can we have symmetric positioning. With more than two firms, the two firms closest to the endpoints of the segment would have only one nearest neighbor but the firms in the interior would have two.
16. Dixit and Stiglitz (1977) actually consider more general utility functions of the form U(G(LJ f(xi))' no).
of
17. It is also common in the literature to study games simultaneous quantity choices, using the expression in (12.C.13) directly as the inverse demand functions for the firms.
400
CHAPTER
12,
MARKET
POWER
-----------------------------------------------------------This leads to a considerable simplification because each firm j, in deciding on its price choice, can take the value of x = lJ(x j ) dj, called the index of aggregate output, as given; its own production has no effect on the value of this index. Given the value of x, firm j faces the demand function Xj(Pj,
.x) = I/t(G~J.xJ
where I/t(.) = f' "('). Its optimal choice can then be viewed as a function p;(x) of the index .x. Thus, the equilibrium value of the aggregate output index, say x', satisfies .X' = Jf(xj(pJ(x'), .x')) dj. This limiting case is known as the monopolistic competition model. It originates in
Chamberlin (1933); sec Hart (1985) for a modern treatment. In markets characterized by monopolistic competition, market power is accompanied by a low level of strategic interaction, in that the strategies of any particular firm do not affect the payoff of any other firm."
12.D Repeated Interaction One unrealistic assumption in the models presented in Section 12.C was their static, one-shot nature. In these models, a firm never had to consider the reaction of its competitors to its price or quantity choice. In the Bertrand model, for instance, a firm could undercut its rival's price by a penny and steal all the rival's customers. In practice, however, a firm in this circumstance may well worry that if it does undercut its rival in this manner, the rival will respond by cutting its own price, ultimately leading to only a short-run gain in sales but a long-run reduction in the price level in the market. In this section, we consider the simplest type of dynamic model in which these concerns arise. Two identical firms complete for sales repeatedly, with competition in each period I described by the Bertrand model. When they do so, the two firms know all the prices that have been chosen (by both firms) previously. There is a discount factor (j < I, and each firm j attempts to maximize the discounted value of profits, :[''';., (j'-lltj" where Itj' is firm j's profit in period I. The game that this situation gives rise to is a dynamic game (see Chapter 9) of a special kind: it is obtained by repeated play of the same static simultaneous-move game and is known as a repeared game. In this repeated Bertrand game, firm j's strategy specifies what price Pj' it will charge in each period t as a function of the history of all past price choices by the two firms, H,_ 1 = {Ph' Pz,}:: \. Strategies of this form allow for a range of interesting behaviors. For example, a firm's strategy could call for retaliation if the firm's rival ever lowers its price below some "threshold price." This retaliation could be brief, calling for the firm to lower its price for only a few periods after the rival "crosses the line," or it could be unrelenting. The retaliation could be tailored to the amount by which the firm's rival undercut it, or it could be severe no matter how minor the rival's transgression. The firm could also respond with increasingly cooperative
18. In contrast, in spatial models, even in the limit of a continuum of firms, strategic interaction remains. In lhat case, firms interact locally. and neighbors count, no matter how large the economy is.
--
SEC T ION
1 2 . 0:
REP EAT E 0
behavior in return for its rival acting cooperatively in the past. And, of course, the firm's strategy could also make the firm's behavior in any period t independent of past history (a strategy involving no retaliation or rewards). Of particular interest to us is the possibility that these types of behavioral responses could allow firms, in settings of repeated interaction, to sustain behavior more cooperative than the outcome predicted by the simple one-shot Bertrand mode\. We explore this possibility in the remainder of this section. We begin by considering the case in which the firms compete only a finite number of times T (this is known as a finilely repeated game). Can the rich set of possible behaviors just described actually arise in a subgame perfect Nash equilibrium of this model? Recalling Proposition 9.BA, we see the answer is "no." The unique subgamc perfect Nash equilibrium of the finitely repeated Bertrand game simply involves T repetitions of the static Bertrand equilibrium in which prices equal cost. This is a simple consequence of backward induction: [n the last period, T, we must be at the Bertrand solution, and therefore profits are zero in that period regardless of what has happened earlier. But, then, in period T - I we are, strategically speaking, at the last period, and the Bertrand solution must arise again. And so on, until we get to the first period. In summary, backward induction rules out the possibility of morc cooperative behavior in the finitely repeated Bertrand game. Things can changc dramatically, however, when the horizon is extended to an infinite number of periods (this is known as an infinitely repealed game). To sec this, consider the following strategies for firms j = I, 2: if all elements of H,., equal (pm, pm) or t = I otherwise.
(12.0.1)
In words, firm j's strategy calls for it to initially play the monopoly price pm in period I. Then, in each period t > I, firm j plays pm if in every previous period both firms have charged price pm and otherwise charges a price eq ual to cost. This type of strategy is called a Nash reversion stralegy: Firms cooperate until someone deviates, and any deviation triggers a permanent retaliation in which both firms thereafter set their prices equal to cost, the one-period Nash strategy. Note that if both firms follow the strategies in (12.0.1), then both firms will end up charging the monopoly price in every period. They start by charging pm, and therefore no deviation from pm will ever be triggered. For the strategies in (12.0.1), we have the result presented in Proposition 12.0.1. Proposition 12.0.1: The strategies described in (12.0.1) constitute a subgame perfect Nash equilibrium (SPNE) of the infinitely repeated Bertrand duopoly game if and only if c5 ~ ~. Proof: Recall that a set of strategies is an SPNE of an infinite horizon game if and only if it specifies Nash equilibrium play in every subgame (see Section 9.B). To start, note that although each subgame of this repeated game has a distinct history of play leading to it, all of these subgames have an identical structure; Each is an infinitely repeated Bertrand duopoly game exactly like the game as a whole. Thus, to establish that the strategies in (12.0.1) constitute an SPNE, we need to show that after any previous history of play, the strategies specified for the remainder of the game constitute a Nash equilibrium of an infinitely repeated Bertrand game.
IN T ERA C T ION
401
402
C HAP T E R
1 2:
MAR K E T
POW E R
In fact, given the form of the strategies in (12.0.1), we need to be concerned with only two types of previous histories: those in which there has been a previous deviation (a price not equal to pm) and those in which there has been no deviation. Consider, first, a subgame arising after a deviation has occurred. The strategies call for each firm to set its price equal to c in every future period regardless of its riva!"s behavior. This pair of strategies is a Nash equilibrium of an infinitely repeated Bertrand game because each firm j can earn at most zero when its opponent always sets its price equal to c, and it earns exactly this amount by itself setting its price equal to (' in every remaining period. Now consider a subgame starting in, say, period t after no previous deviation has occurred. Each firm j knows that its rival's strategy calls for it to charge pm until it encounters a deviation from pm and to charge c thereafter. Is it in firm j's interest to usc this strategy itself given that its rival does? That is, do these strategies constitute a Nash equilibrium in this subgame? Suppose that firm j contemplates deviating from price pm in period t 2: t of the subgame if no deviation has occurred prior to period r'9 From period t through period r - I, firm j will cam !(pm - c)x(pm) in each period, exactly as it does if it never deviates. Starting in period r, however, its payoffs will differ from those that would arise ifit does not deviate. In periods after it deviates (periods r + I, r + 2, ... ), firmj's rival charges a price of c regardless of the form offirmj's deviation in period r, and so firm j can earn at most zero in each of these periods. In period r, firm j optimally deviates in a manner that maximizes its payoff in that period (note that the payoffs firm j receives in later periods are the same for any deviation from pm that it makes). It will therefore charge pm - £ for some arbitrarily small £ > 0, make all sales in the market, and earn a one-period payoff of (pm - C - £)x(pm). Thus, its overall discounted payoff from period r onward as a result of following this deviation strategy, discounted to period r, can be made arbitrarily close to (pm - c)x(pm). On the other hand, if firm j never deviates, it earns a discounted payoff from period r onward, discounted to period r, of [l(pm - c)x(pm)]j(l - b). Hence, for any t and r 2: t, firm j will prefer no deviation to deviation in period r if and only if
--- ---
(12.0.2) Thus. the strategies in (12.0.1) constitute an SPNE if and only if .I 2:
! .•
19. From our previous argument, we know that once a deviation has occurred within this stlbgame. firm j can do no bener than to play c in every period given that its rival will do so. Hence, to check whether these strategies form a Nash equilibrium in this subgame, we need only check whether firm j will wish to deviate from I'M if no such deviation has yet occurred.
REPEATED
The discount factor need not be interpreted literally. For example. in a model in which market tiL-mand is growing at ratc~' [i.e .• .'(,(1') = ),'.\:(p)J, larger values of y make the model behave as if there is a larger discount factor because demand growth increases the size of any future losses caused by a current deviation. Alternatively, we can imagine that in each period there is a probability)' that the firms' interaction might end. The larger y is, the more firms
will effectively discount the future. (This interpretation makes clear that the infinitely repeated game framcwork can be relevant even when the firms may cease their interaction within some
finite amount of time: what is needed to fil Ihe analysis into the framework above is a strictly positive probability of continuing upon having reached any period.) Finally, the value of b can rellect how long il takes to deleCI a deviation. These interpretalions are developed in Exercise Il.D.!.
1.
Although the strategies in (12.0.1) constitute an SPNE when .I 2: they are nor the only SPNE of the repeated Bertrand model. In particular, we can obtain the result presented in Proposition 12.0.2. ProposItion 12.0.2: In the infinitely repeated Bertand duopoly game, when ,j 2: ~ repeated choice of any price p € [C, pm] can be supported as a subgame perfect Nash equilibrium outcome path using Nash reversion strategies. By contrast, when ,I < ~, any subgame perfect Nash equilibrium outcome path must have all sales occurring at a price equal to c in every period. Proof: For the first part of the result, we have already shown in Proposition 12.0.1 that repeated choice of price pm can be sustained as an SPNE outcome when rI 2: 1. The proof for any price p € [c, pm) follows exactly the same lines; simply change price I'm in the strategies of (12.0.1) to p€ [c, pm). The proof of the second part of the result is presented in small type. We now show that all sales must occur at a price equal to c when ~ <~. To begin, let -I Tt , denote firmj's profits, discounted to period t, when the equilibrium strategies J are played from pcriod ( onward. Also define 7r, = nit + 7r21Observe thaI, because every firm j finds il oplimal to conform to the equilibrium stralegies in every rcriod t, it must be that
= L, -", (i'
1t, ~ t'jf
The implication of Proposition 12.0.1 is that the perfectly competitive outcome of the static Bertrand game may be avoided if the firms foresee infinitely repeated interaction. The reason is that, in contemplating a deviation, each firm takes into accollnt not only the one-period gain it earns from undercutting its rival but also the profits forgone by triggering retaliation. The size of the discount factor b is
12.D:
i mportant here because it affects the relative weights put on the future losses versus the present gain from a deviation. The monopoly price is sustainable if and only if the present value of these future losses is large enough relative to the possible current gain from deviation to keep the firms from going for short-run profits.
l'J'
or
SECTION
for j = 1.2 and every t,
( 12.D.3)
sincl! each firm j can obtain a payoff arbitrarily close to 1t, in period t by deviating and undcrcutting the lov.'cst price in the market by an arbitrarily small amount and can assure itself a nonnegative payoff in any period thereafter. Suppose that there exists at least one period I in \'v'hich contradiction. There arc two cases to consider:
(i) Suppl..)se. first. that there is a period r with T over j = 1,2, we have
11:1
> 0 such that
adding (12.D.3) for { =
211: r ~ (l:1t
+
V2r)'
But (,." + r,,) S [1/(1 - ,liJn" and so this is impossible if b < j.
Tt f
> O. \Ve will derive a
11:, ~ Tt,
for all r. If so. thell
INTERACTION
403
404
C HAP T E R
1 2:
MAR K E T
POW E R
(ii) Suppose. instead. that no such period exists; that is. for any period I. there is a period n, > n,. Define t(l) for I <: I recursively as follows: Let t( I) = I and for 1<: 2 define t{t) = Min {t > t(1 - I): n, > n,U-II}' Note that, for all I, n, is bounded above by the monopoly profit level "m = (pM - c)x(p') and that the sequence {n,u,},";" is monotonically increasing. lien<.:e, as t - ~..), 1I: (lJ must converge to some it E (0, 11:"'] such that 1I: r < it for all t. Now. adding (12.D.3) over j = 1.2. we see that we must have t > t such that
I
(12.0.4) for alit. Moreover.
1'111Il
+
[1 2110 ::;
[1/(1 -
(~))It
for alit. and so we must have
(12.0.5) for alit. llut when
,j
<
l. condition
(12.0.5) must be violated for t sutlicicntly large.
This completes the proof or the proposition. _ The presence of mUltiple equilibria identified in Proposition 12.0.2 ror ii ;:: ! is common in infinitely repeated games. Typically. a range or cooperative equilibria is possible ror a given level of J, as is a complete lack or cooperation in the rorm of the static Nash equilibrium outcome repeated rorever. Proposition 12.0.2 also tells us that the set of SPNE or the repeated Bertrand game grows as ,) gets Iarger. 20 The discontinuous behavior as a runction or ii of the set or SPNE displayed in Proposition 12.0.2 is, however, a special reature or the repeated Bertrand model. The repeated Coumot model and models of repeated price competition with difTerentiated products generally display a smoother increase in the maximal level or joint profits that can be sustained as ii increases (sec Exercise 12.0.3). In fact. a general result in the theory or repeated games, known as thefolk theorem, tells us the rollowing: In an infinitely repeated game, any feasible discounted payoffs
that give eaclt player, Oil a per-period basis, more tltall tire 10IVest payoff that Ire could guaralltee himself ill a sillgle play of tire simultaneous-move componellt game call be suscailled as the payoff' of WI SPN E if players discoullt the future to a sufficiently small degree. In Appendix A, we provide a more precise statement and extended discussion of the rolk theorem for general repeated games. Its message is clear: Although infinitely repeated games allow ror cooperative behavior, they also allow ror an eXlremely wide rallge or possible behavior. The wide range or equilibria in repeated game models or oligopoly is somewhat disconcerting. From a practical point or view, how do we know which equilibrium behavior will arise? Can "anything happen" in oligopolistic markets? To get around this problem, researchers orten assume that symmetrically placed firms will find the symmetric profit-maximizing equilibrium rocal (sec Section 8.0). However, even restricting attention to the case of symmetric firms, the validity or this assumption is likely to depend on the setting. For example, the history or an industry could make olher equilibria rocal: An industry that has historically been very noncooperative (maybe because c5 has always been low) may find noncooperative outcomes more rocal. The assumption that the symmetric profit-maximizing equilibrium arises is 20. Strictly speaking. Proposition 12.0.2 shows this only for the class of stationary. symmetric
equilibria (i.e .. equilibria in which the firms adopt identical strategies and in which. on the equilibrium path. the actions taken arc the same in every period).
----
SECTION
12.E:
ENTRY
405
-----------------------------------------------------------------more natural when the selr-enrorcing agreement interpretation or these equilibria is relevant. as when oligopolists secretly meet to discuss their pricing plans. Because antitrust laws preclude oligopolists from writing a rormal contract specirying their behavior, any secret collusive agreement among them must be selr-enrorcing and so must constitute an SI'NE. It seems reasonable to think that, in such circumstances, identical lirms will therefore agree to the most profitable symmetric SPNE. (If the finns arc not identical. similar logic suggests that the firms would agree to an SI'NE corresponding to a point on the rrontier or their set or SPNE payofTs.) Finally, just as with the static models discussed in Section 12.C, it is or interest to investigate how the number of firms in a market afTccts ils competitiveness. You arc asked to do so in Exercise 12.0.2. Exercise 12.1).2: Show thai with} firms, repeated choice of any price p E (c, pm] can be sustained as a stationary SPNE outcome path or the infinitely repeated Bertrand game using Nash reversion strategies ir and only if i5;:: (J - I)I). What docs this say ahout the effect or having more firms in a market on the difficulty or sustaining collusion'!
In
pra\.:li<':l!.
an important feature of many settings of oligopolistic collusion (as well as other
settings of <.:oopcration) is thaI firms are likely to be able to observe their rivals' behavior only
imperfeclly. For e,ample, as emphasized by Sligier (1960). an oligopolist's rivals may make secret pri<.:c <.:uts 10 consumers. If the market demand is stochastic. a firm will be unable to tell with Cerlalilly whether there have been any deviations from collusive pricing simply from observatil)n of its own demand. This possibility lends formally to study of rep('ateJ O(ll"e.~ willI imp,·,}"", "h"arC/hili,)'; sec. for example. Green and Porter (1984) and Abreu. Pearce, and Slachclli (1990). A fcalure of this class of models is Ihat they are able to explain observed breakdowns of cooperation as being an inevilable result of allempts to cooperate in environments characterized by imperfect observability. This is so because equilibrium strategies must be such that some negative realizations of demand result in a breakdown of cooperation if firms arc to be prl.!vented from secretly deviating from a collusive scheme.
12.E Entry In Sections 12.B to 12.0, we analyzed monopolistic and oligopolistic market outcomes, keeping the number or active firms exogenously fixed. In most cases, however. we wish to view the number of firms that will be operating in an industry as an endogene()us variable. Doing so also raises a new question regarding the welfare properties or siluations in which market power is present: Is the equilibrium number or firms thai enter the market socially efficient" In Seclion IO.F, we saw that Ihe answer to this question is "yes" in the case of competitive markets as long as an equilibrium exists. In this section, however, we shall sec that this is no longer true when market power is present. We now take the view that there is an infinite (or finite but very large) number of potential firms, each or which could enter and produce the good under consideration if it were profitable to do so. As in Section IO.F, we rocus on the case in which all potential firms are identical. (See Exercise 12.E.1 ror a case in which they arc not.)
406
CHAPTER
12:
MARKET
SEC T ION
POWER
We illustrate the determination of the equilibrium number of firms with two examples in which the stage 2 oligopoly games correspond, respectively, to the Coumot and Bertrand models discussed in Section 12.C.
A natural way to conceptualize entry in oligopolistic settings is as a two-stage process in which a firm first incurs some setup cost K > 0 in entering the industry and then, once this cost is sunk, competes for business. The simplest sort of model that captures this idea has the following structure:
Example I2.E.I: Elluilihrilll1l Entry with Coumol Compelilion. Suppose that competition in stage 2 of the two-stage entry game corresponds to the Cournot model studied in Section 12.C, with e(l/) = el/, p(q) = 0 - "if. 0 > e ~ 0, and h > O. The stage 2 output per firm. if}. and profit per firm, tr}. arc given (see Exercise 12.C.?) by
Sraye I: All potential firms simultaneously decide "in" or "out." If a firm decides "in," it pays a setup cost K > O. Staye 2: All firms that have entered play some oligopolistic game.
The oligopoly game in stage 2 could be any of those considered in Sections 12.C and 12.D. Formally, this two-stage entry model defines a dynamic game (see Chapter 9). Note that its stage 2 subgames are exactly like the games we have analyzed in the previous sections because, at that stage, the number of firms is fixed. Throughout our discussion we shall assume that for each possible number of active firms, there is a unique, symmetric (across firms) equilibrium in stage 2, and we let tr} denote the profits of a firm in this stage 2 equilibrium when J firms have entered (tr) does not include the entry cost K). This two-stage entry model provides a very simple representation of the entry process. There is very little dynamic structure, and no firm has any "first-mover" advantage that enables it to deter entry or lessen competition from other firms (see Section 12.G and Appendix B for a discussion of these possibilities). Consider now the (pure strategy) subgame perfect Nash equilibria (SPNEs) of this model. In any SPNE of this game, no firm must want to change its entry decision given the entry decisions of the other firms. For expositional purposes, we shall also adopt the convention that a firm chooses to enter the market when it is indifferent. With this assumption, there is an equilibrium with J* firms choosing to enter the market if and only if (12.E.I)
(12.E.3) (12.E.4) Note that11} is strictly decreasing in J and thattr} - 0 as J - CD. Also, Ji/} - (a - e)/" as J - 'lO, so that aggregate quantity approaches the competitive level. Solving for the real number J E !;l at which trj = K gives
(J +
, (0 - e)' 1)- = hK
or
J=
(0 :- e) _ I.
/hK
The equilibrium number of entrants J * is the largest integer that is less than or equal to 1. Note that as K decreases. the number of firms active in the market (weakly) increases. and that as more firms become active. aggregate output increases and price decreases. Indeed, J * - CD as K - O. and output and price approach their competitive levels. Note also that II proportional increase in demand at every price, captured by a reduction in b, changes the equilibrium number of firms and price in a manner that is identical to a decrcase in K. _
and Example 12.E.2: E'IlIilihrilll1l flllrr will, Berlrallli Competilion. Suppose now that competition in stage 2 of the two-stage entry game takes the form of the Bertrand model studied in Section 10.C. Once again, c(if) = e'l, p(q) = a - hq, a > C ~ 0, and b> O. Now tr, = tr", the monopoly profit level. and tr} = 0 for all J ~ 2. Thus, assuming that tr" > K, the SPNE must have J* = I and result in the monopoly price and quantity levels. Comparing this result with the result in Example 12.E.l for the Cournat model. we sec that the presence of more intense stage 2 competition here actually lowers the ultimate level of competition in the market' _
(12.£0.2) Condition (12.E.I) says that a firm that has chosen to enter does at least as well by doing so as it would do if it were to change its decision to "out," given the anticipated result of competition with J* firms. Condition (I2.E.2) says that a firm that has decided to remain out of the market does strictly worse by changing its decision to "in," given the anticipated result of competition with J* + I firms. Typically, we expect that tr} is decreasing in J and that tr} - 0 and J - :£. In this case, there is a unique integer j such that tr} ~ K for all J S j and tr} < K for all J> j, and so J* = j is the unique equilibrium number of firms 2 1.22 21. NOle, however, that although there is a unique number of entrants. there arc many equilibria, in each of which the particular firms choosing to enter differ. 22. Without the assumption that firms enter when indifferent, condition (12.1.:":.2) would be a weak inequality. This change in (12.E.2) matters for the idcntinc~ltIon ofthc equilihrium numher or firms only in the case in which (here is an integer number of firms J such that "Ttj = K (so that with J firms in the market each firm earns exactly zero net of its entry cost K). When this is so, this change allows both J and J - 1 to be equilibria. With minor adaptations but some loss of exrosilional simplicity, all the points made in this section can be extended to cover this case.
Elltry {(lid WeI/1m'
J
Consider now how the number of firms entering an oligopolistic market compares with the number that would maximize social welfare given the presence of oligopolistic competition in the market. We begin by considering this issue for the case of a homogeneous-good industry. Let 'I) be the symmetric equilibrium output per firm when there are J firms in the market. As usual, the inverse demand function is denoted by pC). Thus, p(Jq}) is the price when there are J active firms; and so tr} = p(Jq})q} - e(q}), where c(') is the cost function of a firm after entry. We assume that e(O) = O.
1 2 . E:
E N TRY
407
408
CHAPTER
12:
MARKET
POWER
We measure welfare here by means of Marshallian aggregate surplus (see Section IO.E). In this case, social welfare when there are J active firms is given by J .,
W(J) =
f
[>(s) ds - Je(qJ) - JK.
0
(12.E.5)
Example I2.E,3: Consider the Cournot model of Example 12.E.1. For the moment, ignore the requirement that the number of firms is an integer, and solve for the number of firms I at which w'(I) = O. This gives (J
+
(a - e)2 I) = J
-"K-
proposition 12.E,1: Suppose that conditions (Al) to (A3) are satisfied by the post-entry oligopoly game. that p'(') < 0, and that c ~ O. Then the equilibrium number of entrants. J*. is at least J O - 1. where J O is the socially optimal number of 23 entrants. H
_
+ 1)2
I) =
Q,
f
Q,
bK
(I +
pis) ds - l"c(qJ )
+ (J" -
I)c(qr-,) ~ K,
,
where we let QJ = JlJJ. We can rearrange this expression to yield Q,
nJ
-,
-
K ~ p(QJ -, )qJ -, -
f
Q,
pIs) ds
+ J"[r(qJ')
- r(qJ _,)].
,
Given 1"(') < 0 and condition (A I), this implies that 1tJ-,-I\.~p(QJ-I)[(JJ'-,+QJ-,-QJ]+r[c(qJ)-(qJ-,)].
(a _ 1')2
= - ---
the equilibrium number of firms is the largest integer less than or equal to (12.E.6) and (12.E.7), we see that
(J +
Proof: The result is trivial for J 0 = I, so suppose that r > I. Under the assumptions of the proposition, 1t J is decreasing in J (Exercise 12.E.2 asks you to show this). To establish the result, we therefore need only show that 1tJ '_I ~ K. To prove this, note first that by the definition of r we must have W(J°)W(}" - I) ~ 0, or
(12.E.6)
If I turns out to be an integer, then the socially optimal number of firms is r = 1. Otherwise, J '" is one of the two integers on either side of I [recall that W(·) is concave]. Now, recall from (12.E.4) that 1t J ~ (I/h)[(a - C)/(J + 1)]2. As noted in Example 12.E.I, if we let I be the real number such that (J
For markets satisfying these three conditions we have the result shown in Proposition 12.E.1.
(.)
The socially optimal number of active firms in this oligopolistic industry, which we denote by J , is any integer number that solves Max J W(J). Example 12.E.3 illustrates that in contrast with the conclusion arising in the case of a competitive market, the equilibrium number of firms here need not be socially optimal.
_
---
SECTION
(12.E.7)
1.
From
1)J!2.
Thus, when the demand and cost parameters arc such that the optimal number of firms is exactly two (F = I = 2), four firms actually enter this market (J* = 4, since J;; 4.2); when the social optimum is for exactly three firms to enter (J" = I = 3), seven firms actually do (J * = 7, since J = 7); when the social optimum is for exactly eight firms to enter (J' = I = 8), 26 actually enter (J * = 26, since J = 26) . • Can we say anything general about the nature of the entry bias? It turns out that we can as long as stage 2 competition satisfies three weak conditions [we follow Mankiw and Whinston (1986) here]: (A I) JqJ ~ J''1J' whenever J > J'; (A2) '1J $: '1J' whenever J > J'; (A3) p(J'lJ) - e'('1J) ~ 0 for all J. Conditions (A I) and (A3) arc straightforward: (A I) requires that aggregate output increases (price falls) when more firms enter the industry, and (A3) says that price is not below marginal cost regardless of the number of firms entering the industry. Condition (A2) is more interesting. It is the assumption of bl/siness stealing. It says that when an additional firm enters the market, the sales of existing firms fall (weakly). Hence, part of the new firm's sales come at the expense of existing firms. These conditions are satisfied by most, although not all, oligopoly models. [In the Bertrand model. for example, condition (A3) does not hold.]
12.E,
ENTRY
409
,---------------------------------------------------------------------
(12.E.8)
But since 1'''(-) ~ 0, we know that (,'(qJ _, )[qr - qJ _,] $: (qJ) - c(qJ _,). Using this inequality with (12.E.8) and the faet that q, _, + QJ _ I - Qr = J"(qr- , - qJ') yields 1t J -, - K ~ [p(QJ-,) - ('(q, -,)]J"(qr-, - qJ ).
Conditions (A2) and (A3) then imply that 1t J
_,
~ K.24 •
The idea behind the proof of Proposition 12.E.1 is illustrated in Figure 12.E.1 for the case where ('I) = 0 for all q. In the figure, the incremental welfare benefit of the rth firm, before taking its entry cost into account, is represented by the shaded area (ahcd). Since entry of this firm is socially efficient, this area must be at least K. But area (ahed) is less than area (aha), which equals p(QJ -,)(QJ" - QJ -,i. Moreover, business stealing implies that (Qr - QJ _,) = J'qJ - (J" - 1)'1, _, $: q, _I' and so we see that area (abce) $: p(QJ' -,)qr-I ~ 1CJ ,-, [the value of1CJ"_' is represented in Figure 12.E.1 by area (abfg)]. Hence 1tJ-' ~ K. The tendency for excess entry in the presence of market power is fundamentally driven by the business-stealing effect. When business stealing accompanies new entry and price exceeds marginal cost, part of a new entrant's profit comes at the expense of existing firms, creating an excess incentive for the new firm to enter. Of course, as Proposition 12.E.1 indicates, we may also see too few lirms in an industry. The classic example concerns a situation in which the socially optimal number of lirms is one. A single firm deciding whether to enter a market as a
1I J
21 Iflhere is more than one maximizer of W(J), say p;, .. .,J~}, then J* ~ Max{J; .... ,J~} - t. 24. Note that if (A I) holds with strict inequality. then this conclusion can be strengthened to O _ I > K [a strict inequality appears in (12.E.8»). In this case, J* ~ J I even if firms do not enter
when indifferent.
410
CHAPTER
12:
MARKET
SECTION
POWER
------------------------------------------------------------------- ,---
AC(q) = K
Ftgure 12.E.l
/
(left)
12.F:
THE
COMPETITIVE
+ cq q
Diagrammatic
p(Q,
I Q
12E.4. Figure 12.E.2 (right) An insufficient entn incentive.
monopolist compares its monopoly profit-the hatched area (abde) in Figure 12.E.2-with the entry cost K. However, the firm fails to capture, and therefore ignores. the increase in consumer surplus that its entry generates-the shaded area (file). As a result, the firm may find entry unprofitable even though it is socially desirable. Proposition 12.E.1 tells us, however, that if we have too little entry in a homogeneous-good market, this can be at most by a single firm. What happens when product differentiation is present? It turns out that we can then say very little of a general nature. The reason is that the sort of problem illustrated in Figure 12.E.2 can now happen for many products, leading to many "too few by one" conclusions. An additional issue is that. with product differentiation. the number of firms is not all that matters. We may also fail to have the right selection of prod ucts. 25 An alternative approach to the two-stage entry game models the actions of entry and quantity/price choice as simultaneous. In this one-stage tntr}' game, a firm incurs its setup cost
.
several firms at price p*, none of them could cover their cost).2b In this equilibrium, all lirms
make zero prolits. The equilibrium outcome
IS
depicted in Figure 12.E.3. Observe that it is
strictly superior in welfare terms to the outcome that arises in the two-stage entry process considered in Example 12.E.2, where there is also a single firm active but it quotes a monopoly pricc. 27 •
What is the critical ditTerence between the one-stage and two-stage entry processes" In the two-stage model an cntrant must sink its fixed costs prior to competing, whereas in the one-stage model it can compete for sales while retaining the option not to sink these costs if it does not make any sales. We can think of the two-stage case as a model of a firm incurring a once-and·for-all sunk entry cost that allows for many later periods of competitive interaction, whereas the one-stage case cHptures a setting in which ·'hit·and-run" entry is possible (i.e.,
entry for just one period while paying only the one-period rental price of capital). When a firm must incur a sunk cost in entering it must consider the reaction of other firms to its entry. In the Bertrand model with constant costs this reaction is severe: price falls to cost and the firm loses money by entering. In contrast, in the one-stage game the firm can enter and undercut active firms' prices without fearing their reactions. This makes entry more aggressive
and leads to a lower equilibrium price. This one-stage entry model with price competition providcs olle formalization of what Baumol. Panzar. and Willig (1982) call a co"lesllll>/e market.
only if it sells a positive amount. For example. the one-stage versions of Examples 12.E.1 and 12.E.2 are Cournot and Bertrand games. respectively. with cost functions C(q) = { :
+ c(q)
if q > 0 if q = 0
and an infinite (or very large) number of firms. For models of price competition. this change can have dramatic consequences. Consider the effect on the result of Example 12.E.2 that is illustrated in Example 12.E.4. Example 12_E.4: Tire Olle-Stage Elltry Model witlr Bertralld Competi/ioll. Suppose that + cx(p)J/x(p) for some p (the parameter c > 0 is the cost per unit); that is, suppose
12.F The Competitive Limit In Chapter 10. we introduced the idea that a competitive market might usefully be thought of as a limiting case of an oligopolistic market in which firms' market power grows increasingly small (see Section 10.B). We also noted that this view could provide a framework for reconciling cases in which competitive equilibria fail to exist in the presence of frcc entry and average costs that exhibit a strictly positive etlicient
p> [K
there is some price level at which a monopolist can earn strictly positive profits after paying its set up cost K. Assume that many firms simultaneously name prices and that a firm incurs the setup cost K only if it actually makes sales. Any equilibrium of this game has all sales occurring at price p' = Min{p: p '" [K + cx(plJjx(p)} (if price is above p', some firm could
gain by setting a price p' - r.; if price is below p', some firm must be making strictly negative prolits). and one firm satisfying all demand at this price (if the demand were split among 25. See Spence (1976). Dixit and Stiglitz (t977), Salop (1979). and Mankiw and Whinston (1986) for more on the case of product differentiation.
Figure 12.E.3
Equilibrium in the one-stage entry game discussed in Example
explanation of Proposition 12.E.1.
II----I-~
LIMIT
26. Note that we now allow consumer demand to be given entirely to one firm when several firms name the same price (before, we had taken the division of demand in this case to be exogenously given). This is the only division of demand that is compatible with equilibrium in this example. It can he formally juslificd as the limit of the equilibria thaI arise when prices must be quoted in di"crctc units as the size of these units grows small. 27. in fact. this equilibrium outcome is the solution to the problem faced by a welfaremaximizing planner who can control the outputs qj of the firms but must guarantee a nonnegative profit to all active firms. that is. who faces the constraint that p(L.1t qk)qj 2: cqj + K for every j with qj > o.
411
412
CHAPTER
12:
MARKET
POWER
----------------------------------------------------------scale (see Section to. F). In this situation, we argued, as long as many firms could fit into the market, the market outcome ought to be close to the competitive outcome that would arise if industry average costs were actually constant at the level of minimum average cost. In this section, we elaborate on these points and develop, in a setting of free entry, the theme that if the size of individual firms is small relative to the size of the market, then the equilibrium will be nearly competitive. We have already seen one example of this phenomenon in Example l2.E.1. Here we establish the point in a more general way. We now let market demand be x,(p) = rn(p), where x(p) is differentiable and x'(') < O. Increases in IX correspond to proportional increases in demand at all prices. Letting p(q) be the inverse demand function associated with x(p), the inverse demand function associated with x,(p) is then p,({j} = p(q/~). All potential firms have a strictly convex cost function c(q) and entry cost K > O. We denote the level of minimum average cost for a firm by <' = Min.>o [K + c(q))/q, and we let ii > 0 denote a firm's (unique) efficient scale. As in Example l2.E.l, we focus on the case of a two-stage entry model with Cournot competition in the second stage, in which the cost K is incurred only if the firm decides to enter in stage I. We let b(Q _ j) denote active firm j's optimal output level for any given level of aggregate output by its rivals, Q _ j ' and we assume that this best response is unique for all Q_j' Finally, we let p, and Q, denote the price and aggregate output in a subgame perfect Nash equilibrium (SPNE) of the two-stage Coumot entry model when the market size is ~. We denote by P, the set of all SPNE prices for market size IX. Proposition 12.F.l: As the market size grows, the price in any subgame perfect Nash equilibrium of the two-stage Cournot entry model converges to the level of minimum average cost (the "competitive" price). Formally, Max
Ip, - cl-+
0
as
IX
-+
00.
p,EI'.
Proof: The argument consists of three steps: (i) First, you are asked in Exercise l2.F.l to show that for large enough IX, an active firm's best-response function b(Q _j) is (weakly) decreasing in Q_j' (ii) Second, we argue that if b(Q _j) is decreasing, then we must have Q, ~ H(i') - in any SPNE of the two-stage entry game with market size IX. To sec why this is so, suppose that with market size IX we had an SPNE with J. firms entering and an aggregate output level Q, < IXX(C) - ij. Consider any firmj whose equilibrium entry choice is "out" in this equilibrium, and suppose that firm j instead decided to enter and produce quantity Ii. Because b(') is decreasing, it is intuitively plausible that the aggregate output level of the original J, active firms cannot increase when firm j enters in this way (see the small-type paragraph that follows for the formal argument behind this claim). As a result, aggregate output in the market following firm j"s entry is no more than (Q, + iii; and since (Q. + ii) < IXX(C), the resulting (post-entry) price is above C. Hence, firm j would earn strictly positive profits by entering in this fashion, contradicting the hypothesis that we were at an SPNE to start with. The argument that the output of the existing J, firms cannot increase following entry of firm j is as follows: Let Q_j be the initial equilibrium level of these firms' aggregate output,
SECTION
12.F:
THE
COMPETITIVE
----
and let Q_J be their post-entry aggregate output. Suppose that Q_j > Q_j' Then at least one of these firms, say firm k, must have increased its output level in response to firm j's entry, say from q. to ii. > q•. Because b(·) is decreasing, it must be that Q_. < Q_.: that is, the post-entry output Q_. of active firms other than k (which includes firm j) must be less than their pre-entry output, Q_•. By part (c) of Exercise 12.C.8, this implies that q. + Q_.;:' ij, + Q_ •. But Q_; = q. + Q_. (since firm j initially produces nothing), and ii, + Q-. ~ Q-j (because firmj's post-entry output is nonnegative). Hence, Q-J;:' Q_j, which is a contradiction.
(iii) Finally, we argue that the conclusion of (ii) implies the result. To see this, consider how much above c the price can be if aggregate output is no more than ii below exx(c). This is given by dp, = P.(exx(C) -
ii) -
P.(o:x(c»
exX(C) - ii) «_) . =p ( - - -pxc 0:
But as
(X
-+
00, [o:x(c) -
iiJ/o: -+
x(c), so that dP.
-+
O. •
There are two forces driving Proposition 12.F.1. First, the entry process ensures that firms will enter if there is too much "room" left in the market. Second, in a market that is very large relative to the minimum efficient scale, a reduction of output equal to the level of minimum efficient scale has very little effect on price. The consequence of these two facts is that as the market size grows large, firms' market power is dissipated and price approaches the level of minimum average cost (the competitive level). In this limiting outcome, welfare approaches its optimal level. 2. In Example 12.E.2, we saw that in a two-stage Bertrand market, no such limiting result holds.29 Because price drops to marginal cost if even two firms enter, the market is always monopolized, no matter what its size. However, the two-stage Bertrand model's limiting properties are quite special. As long as, for any market size, price is above marginal cost for any finite number of firms that enter the market, and approaches marginal cost as the number of firms grows large, a limiting result like that in Proposition 12.F.1 holds. Finally, Proposition 12.F.l applies only for the case of homogeneous-good markets. With product differentiation, we must be careful. Firms may be small relative to the size of the entire set of interrelated markets, but they may still be large relative to their own particular niche. In this case, each firm may maintain substantial market power even in the limit, and the limiting equilibrium can be far from efficient (see Exercise 12.FA).
28. The sense of approximation is relative to the size parameter of the market ex. Assuming that a is a proxy for the number of consumers, this means that the welfare loss per consumer relative to the social optimum goes to zero. 29. Strictly speaking, firms' cost functions in Example 12.E.2 differ from the cost functions assumed in Proposition 12.F.l (average costs including K are declining everywhere in Example
t2.E.2J. Nevertheless, for the two-stage Cournot model, Proposition 12.F.l can be shown to be valid for the cost function of Example 12.E.2 (letting c in the statement of the proposition now be the limiting value of average cost as a firm's output grows large).
LIMIT
413
414
CHAPTER
12:
MARKET
---
POWER
12.0 Strategic Precommitments to Affect Future Competition
SECTION
12.G:
,...---
STRATEGIC
PRECOMMITMENTS
Strategic Substitutes:
Strategic Complements
db,(') < 0 ds,
db,(') > 0 ds,
ds!(k) < 0 dk
ds;(k) > 0 dk
d.. ;(k) > 0 dk
ds;(k) < 0 dk
An important feature of many oligopolistic settings is that firms attempt to make strategic precommitments in order to alter the conditions of future competition in a manner that is favorable to them. Examples of strategic precommitments abound. For example, investments in cost reduction, capacity, and new-product development all lead to long·lasting changes that can affect the nature of future competition. In practice, these types of decisions can be among the most important competitive decisions that firms make. Some general features of these types of strategic precommitments can be usefully illuminated through examination of the following simple two-stage duopoly model:
dS!(k) _ db,(sf(k)) ( __ ~ ____ ob,(s!(k), kJ!...~ _____ ) . dk ds, I - [ob, (si(k), k)/os,] [db,(sr(k))/ds,]
(12.G.I)
The denominator of the second term on the right-hand side of (12.G.I) being nonegativc is oftcn called the stahility cOlldilion. It implies that the simple dynamic adjustment process in which the firms take turns myopically playing a best response toeach others' current strategies converges to the Nash equilibrium from any strategy pair in a neighborhood of the equilibrium. We shall maintain this assumption for the remainder of our discussion. Thus, the effect of k on s, can be seen to depend on two factors: (i) Does k make firm I more or less "aggressive" in stage 2 competition [i.e., what is the sign of obi (si(k), k)/ilk]? and (ii) Does firm 2 respond to the anticipation of more aggressive play by firm I with more aggression itself or with less [i.e., what is the sign of db,(sr(k))/ds ,]?
L
AFFECT
FutURE
COMPETITION
FIgure 12.G.l
Determinants of the sign of ds!(k)/dk.
I/p,
Staae I: Firm I has the option to make a strategic investment, whose level we denote by k E IR. This choice is observable. Staae 2: Firms I and 2 play some oligopoly game, choosing strategies s, E S, c G;l and s, E S, c R, respectively. Given investment level k and strategy choices (s" s,), profits for firms I and 2 are given by 1I,(s" s,' k) and 11,(-", -',), respectively. For example, k might be an investment that reduces firm I's marginal cost of production with the stage 2 game being Cournot competition (so Sj = qj' firm j's quantity choice). Alternatively, stage 2 competition could be differentiated products price competition. We suppose that there is a unique Nash equilibrium in stage 2 given any choice of k, (s!(k), silk)), and we assume for convenience that it is differentiable in k. We also assume for purposes of our discussion that 01l,(s" s,' k)/cJs, < 0 and cJ1I,(s" s,)/os, < 0, that is, that stage 2 actions are "aggressive" in the sense that a higher level of L j by firm j's rival lowers firm j's profit. Hence, firm I would be better off, all else being equal, if it could induce firm 2 to lower its choice of s,. When can investment by firm I cause firm 2 to lower s,? Letting b,(s" k) and b,(s,) denote firm I's and firm 2's stage 2 best-response functions (note that firm I's best response depends on k), we can differentiate the equilibrium condition s! = h,(h,(s!, k)) to get
TO
Nash Equilibrium wilh Cost e(k') / / ,
(a)
Nash Equilibrium , / with Cost dk")
FIgure 12.G.2
(b)
When firm 2 responds in kind to more aggressive choices of St by firm I [i.e., when dh,(si(k))/d.,·, > 0], we say that s, is a strategic complement of s,; and if firm 2 becomes less aggressive in the face of more aggressive play by firm I [i.e., if db,(sf(k»/ds, < 0], s, is a strategic suhstifllte ofs,. [This terminology is derived from Bulow, Geanakoplos, and Klemperer (1985); see also Fudenberg and Tirole (1984) for a related taxonomy.] Figure 12.G.1 summarizes these two determinants offirm 2's response, ds!(k)/dk. Example 12.G.I: The Strategic Effects /rom IIIt:estmellf ill Marginal Cost Reduction. The importance for strategic behavior of the distinction between cases of strategic complements and strategic substitutes is nicely illustrated by examining the strategic effects of investments in marginal cost reduction for models of quantity versus price competition. Suppose that if firm I invests k then its (constant) per-unit production costs are c(k), where c'(k) < O. Consider, first, the case in which stage 2 competition takes the form of the Cournot model of Example 12.C.I, so that the stage 2 strategic variable is Sj = ql' firm j's quantity choice. In this model, we have a situation of strategic substitutes because firm 2's best-response function in stage 2 is downward sloping [db,(q,)/llq, < 0 at all 'I, such that b,(qtl > 0]. As shown in Figure 12.G.2(a), the lowering of firm I's marginal cost because of an increase in k from, say, k' to k" > k', shifts firm I's best-response function outward from b,(q" k') to b,(q" k"); with lower marginal costs, firm I will wish to produce more for any quantity choice of its rival
Strategic effects of a reduction in marginal cost from c(k') to elk") < elk'). (a) Quantity model. (b) Price model.
415
416
CHAPTER
12:
MARKET
APPENDIX
POWER
[and so, in terms of our earlier analysis, iJb,(q!(k), k)/iJk > 0). Thus, in this model, investment in cost reduction leads to a reduction in firm 2's output level, an effect that is beneficial for flrm I [sec Figure 12.G.2(a»). In contrast, suppose that stage 2 competition takes the form of the differentiated price competition model of Example 12.C.2. Here we take Sj = (I/p;) to conform with the interpretation of Sj as an "aggressive" variable [i.e., ,)1!,(s" 52' k)/DS 2 < 0]. In this model. we have a situation of strategic complements: an anticipated reduction in lirm 1'5 pri<:e causes firm 2 to reduce its price also [i.e., dh 2(I/p,)/d(l/p,) > 0]. As depicted in Figure 12.G.2(b), a reduction in firm I's marginal cost because of an increase in k from k' to k" > k' once again makes flrm I more aggressive, leading it to choose a lower price given any price choice of its rival: its best-response function shirts to the right from h,(I/P2,k') to h,(I/P2,k") [hence, in terms of our earlier analysis. ,oh,(I/p!(k), k)/,'k > 0]. With strategic complements, the result of the reduction in firm I's marginal cost is therefore to lower flrm 2's equilibrium price, an elTect that is undesirable for flrm I. Thus, the strategic ciTects of a reduction in flrm I's marginal cost differ between the two models, being beneficial to firm I in the quantity model and detrimental in the price model.·'o Which model more accurately captures the nature of competitive intera<:tion depends on the particulars of an industry's situation. For example, if flrms in a mature industry have excess capacity, the price model is likely to be more descriptive, and the strategic effect will be detrimental. On the other hand, in a new market where firms arc investing in capacity, the strategic effect is likely to be better captured by the quantity model (recall our interpretation of the Cournot model in terms of capacity choices in Section 12.C>. • In dcciding on its level of investment, firm I must therefore consider not only the direct elTects of its investment (say, the direct benefit of lower costs), but also the strategic effects that arise through induced changes in its its rival's behavior. Formally, the derivative of flrm l's proflts with respect to a change in k can be written as tl1!,(sW), sj(k), k) = "::,(~l'Sk}~!(~), k) tlk ,?k
+ ';ll,<:!~~)~~!(~~kl ~-"-E~)
,'5,
tlk
+ ~1!,-(:l'(k), :t(~~kl ~s!(k).
A:
INFINITELY
REPEATED
GAMES
AND
THE
FOLK
In the above discussion, we have considered situations in which a firm makes a strategic precommitment to affect future competition with another firm who is (or will be) in the market. A particularly striking example of strategic precommitment to aft'ect future market conditions, however, arises when one firm is the first into an industry and seeks to usc its flrst-mover advantage to deter further entry into its market. We can analyze this case formally by introducing a stage between stages I and 2, say stage 1.5, at which flrm 2 decides whether to be in the market and by supposing that if firm 2 chooses "in" then it must pay a set-up cost F > O. Firm 2 will therefore choose "out" given firm l's stage I choice of k if its anticipated profit in stage 3, 1!2(s1'(k), s!(k», is less than F. Given this fact, the incumbent would, of course, like simply to announce that in response to any entry it will engage in predatory pricing (i.e., it will choose a very high level of s, in stage 3). The problem, however, is that this threat must be credible (recall the discussion in Chapter 9). Thus, what the incumbent needs to do to deter entry is choose a level of k that preeommits it to sufliciently aggressive behavior that flrm 2 chooses not to enter. In any particular problem, this mayor may not be possible, and it mayor may not be profitable. As a general matter, there arc many potential mechanisms (i.e., many types of variables k) by which such precolllmitments can be made. In Appendix B, we examine in some detail the classic mechanism of entry deterrence through capacity expansion first studied hy Spence (1977) and Dixit (1980).
APPENDtX A: tNFINITELY REPEATED GAMES AND THE FOLK THEOREM
In this appendix, we extend the discussion in Section 12.D of infinitely repeated games to a more general setting. Our primary aim is to develop a formal statement of a version of the jiJlk the()rem of infinitely repeated games. Infinitely repeated games have a very rich theoretical structure and we shall only touch on a limited number of their properties. Fudenberg and Tirolc (1992) and Osborne and Rubinstein (1994) provide Illore extended discussions.
The Model
The IIrst term on the right-hand side of (12.G.2) is the direct effect on firm l's proflts from changing k: the second term is the strateyic effect that arises because of flrm 2's equilibrium response to the change in k. Since e", (sf(k), s!
An infinitely repeated game consists of an infinite sequence of repetitions of a one-period simultaneous-move game, known as the stage game. For expositional simplicity, we focus here on the case in which there are two players. In the one-period stage game, each player i has a compact strategy set Si: qi E Si is a particular feasible action for player i. Denote q = (q" q2) and S = S, X S2- Player i's payolT function is 7[i(Cfi, Cfj)' We restrict our attention throughout to pure strategies. It will be convenient to define player i's one-period best-response payoff given that his rival plays Cfj by "i(Cfj) = Max.,s, 1!i(q, qj).31 We assume that the stage game has a unique pure strategy Nash equilibrium q* = (ql', q!l (the assumption of uniqueness is for expositional simplicity only). In the infinitely repeated game, actions are taken and payoffs arc earned at the beginning of each period. The players discount payoffs with discount factor b < I.
30. Best-response functions need not always slope thiS way in the price and quantity models, but the particular examples considered here represent the "normal" cases; see Exercise 12.C.12.
31. We assume that conditions on the sets Sj and functions 1[i(Qj. q) hold such that this function exists (i.e .. slich that each player's best response is always well defined).
C5 2
dk
Since at a Nash equilibrium in stage 2 given investment level k we have ,'",(s1'(k), 5!(k), k)/,'s, = 0, this simplifies to d" ,(51'(/.;), s!(/.;), k) ...
d/';
<'",(s1'(/';), s!(k), k)
=--
.--.- .. -- .".
,'k
+
<,,,,(51'(/';), sj(k), k) ds!(k)
--.... -.----- - - . (12.G.2) ':S2
dk
THEOREM
417
418
CHAPTER
12:
MARKET
POWER
Players observe each other's action choices in each period and have perfect recall. A pure strategy in this game for player i, s" is a sequence of functions {s,,(' )},~, mapping from the history of previous action choices (denoted H,_,) to his action choice in period I, s,,(H, _,) E S,. The set of all such pure strategies for player i is denoted by L" and s = (5,,5,) E L, X L, is a profile of pure strategies for the two players. Any pure strategy profile s = (5,,5,) induces an oUlcome palh Q(5), an infinite sequence of actions {q, = (q,,, q,,)},'~, that will actually be played when the players follow strategies 5, and 5,. Player i's discounted payoff from outcome path Q is given by v,(Q) = I:~o O'It,(q, +,). We also define player i's average payoff from outcome path Q to be (I - o)v'(Q); this is the per-period payoff that, if infinitely repeated, would give player i a discounted payoff of v,(Q). Finally, it is also useful to define the discounted continuation payoff from outcome path Q from some period 1 onward (discounted to period I) by V,(Q, I) = I;". 0 (j'It,(q,+,). We can note immediately the following fact: The strategies that call for each player i to play his stage game Nash equilibrium action in every period, regardless of the prior history of play, constitute an SPNE for any value of (j < I. In the discussion that follows, we are interested in determining to what extent repetition allows other outcomes to emerge as SPNEs.
qr
Nash Revel'sion and Ihe Nash Reversion Folk Theorem We begin by considering strategies with the Nash reversion form that we considered for the Bertrand pricing game in Section 12.0. Definition 12.AA.1: A strategy profile s = (5 , ,52) in an infinitely repeated game is one of Nash reversion if each player's strategy calls for playing some outcome path Q until someone defects and playing the stage game Nash equilibrium q' = (qf, q~) thereafter. What outcome paths Q can be supported as outcome paths of an SPNE using Nash reversion strategies? Following logic similar to that discussed in Section 12.0, we can derive the test in Lemma 12.AA.1. Lemma 12.AA.1: A Nash reversion strategy profile that calls for playing outcome path Q = {q,P q2'};';.' prior to any deviation is an SPNE if and only if • "i(qi')
(wherej # i) for all
+ 1 _(j 0 "i (* q"
*)
q2
~ vi ( Q ,t)
(12.AA.l)
---
APPENDIX
A:
INFINITELY
REPEATED
GAMES
AND
THE
FOLK
In the other direction, suppose that condition (12.AA.I) is satisfied for all i and 1 but that we do not have an SPNE. Then there must be some period t in which some player i finds it worthwhile to deviate from outcome path Q if no previous deviation has occurred. Now, when his opponent follows a Nash revision strategy, player i's optimal deviation will involve deviating in a manner that maximizes his payoff in period I and then playing q1 thereafter. But his payoff from this deviation is exactly that on the left side of condition (12.AA.I), and so this deviation cannot raise his payoff. _ Condition (12.AA.I) can be written to emphasize the trade-off between one-period gains and future losses as follows: •
)
«
Q
",(qj' - It,(q,,, q2,) ~ u v,( ,I
+
I)
q!l)
7t,(qt, - --
l-lJ
(12.AA.2)
for all 1 and j = I, 2. The left-hand side of condition (12.AA.2) gives player i's one-period gain from deviating in period I, and the right-hand side gives player i's discounted future losses from reversion to the Nash equilibrium starting in period 1 + I. For stationary outcome paths of the sort considered in Section 12.0 [where each player i takes the same action q, in every period, so that Q = (q" q2), (q,. q2), ... ], the infinite set of inequalities that must be checked in condition (12.AA.2) reduce to just two: infinite repetition of (q" q,) is an outcome path of an SPNE that uses Nash revcrsion if and only if, for i = I and 2, (j ft,(qj) - It,(q" q,) ~ I _ (j [",(q" q2) - It,(qt, q!)J.
(12.AA.3)
How much better than the static Nash equilibrium outcome q* = (qT, q!) can the players do using Nash reversion? First, under relatively mild conditions (which the Bertrand game considered in Section 12.0 does not satisfy), the players can sustain a stationary outcome path that has strictly higher discounted payoffs than docs infinite repetition of q* = (qt, q!) as long as fJ > O. This fact is developed formally in Proposition 12.AA.1. Proposition 12.AA.1: Consider an infinitely repeated game with (j > 0 and S, c IR for i = 1,2. Suppose also that 7t i (q) is differentiable at q* = (qT, qn, with {)7ti(q;, qJ* )/iJqi # 0 for j # i and i = 1,2. Then there is some q' = (q;, q,), with ["I(q'). 1t2(q')] » [",(q*). 7t 2 (q*)] whose infinite repetition is the outcome path of an SPNE that uses Nash reversion.
t and i = 1,2.
Proof: As discussed in Section 12.0. the prescribed play after any deviation is a Nash equilibrium in the continuation subgame; so we need only check whether these strategies induce a Nash equilibrium in the subgame starting in any period 1 when there has been no previous deviation. Note first that if for some i and t condition (12.AA.I) did not hold, then we could not have an SPNE. That is, if no deviation had occurred prior to period I, then in the continuation subgame, player i would not find following path Q to be his best response to player j's doing so (in particular, a deviation by player i in period t that maximizes his payoff in that period, followed by his playing q1 thereafter, would be superior for him).
THEOREM
419
~----------------------------------------------------------
Proof: At q = (qT, qn, condition (l2.AA.3) holds with equality. Consider a differential change in l/, (dq" dq,), such that [Dni(q~, qil/cqJ dqi > 0 for i = 1,2. The differential change in firm i's profits from this change is
1(*
{ Tr.
*)_on,(q;,qil .+t1n,(q,',qn - - - . - - d q, - - . - - dq, cq, aqj
q• . qj
cn,(q;, qn d ..
cqj
qj.
(t2.AA.4)
since q't is a best response to qj. Thus, dn,(q~,
qn > O.
(t2.AA.5)
420
C HAP T E R
1 2:
MAR K E T
A P PEN 0 I X
POW E R
A:
I NFl NIT ELY
REP EAT E 0
GAM E 5
AND
THE
F0 LK
THE 0 REM
421
-------------------------------------------------------------------------- ,-----------------------------------------------------------------------"" (I - b)v,
On the other hand, the envelope theorem (see Section M.L of the Mathematical Appendix) tells us that at any 4j • (;rr,(h,(4j), 4) tirr,(qj)
= .... -:-_._.
,,, ,,
J4"
cqj
where 11,(') is player i's best response to 'II in the stage game. Hence,
.
•
tirr,(4j)
tn, ('I: . qj)
= .
....
~
.,(q') J4 r
(12.AA.6)
----+-------,
,
"irq')
Ftgur.12.AA.1 The Nash reversion folk theorem.
rr,,(1 - b)v,
+ 6.41,42 + lltlz) is sustainable as players from deviating from any given outcome path. In general, Nash reversion is not the most severe credible punishment that is possible. Just as players can be induced to cooperate through the use of threatened punishments, they can also be induced to punish each other. To consider this issue, it is useful to let ,!, = Min., [Max., ",(q;, qj)) denote player i's minimax po)'uj):'.1 Payoff ,!; is the lowest payoff that player i's rival can hold him to in the stage game if player i anticipates the action that his rival will play. Note, first, that player i's payoff in the stage game Nash equilibrium q. = ('Ii, cannot be below'!,. More importantly, regardless of the strategies played by his rival, player i's average payoff in the infinitely repeated game or in any subgame within it cannot be below'!,. Thus, no punishment following a deviation can give player i an average payolT below'!,. Payoffs that strictly exceed ,!, for each player i are known as illdivitillol/y rational pu)'offs. Note that for a punishment to be credible we must be sure that after an initial deviation occurs and the punishment is called for, no player wants to deviate from the prescribed punishment path. This means that a punishment is credible if and only if it itself constitutes an SPNE outcome path. Proposition 12.AA.4 tells us that as long as " > 0 and conditions similar to those in Proposition 12.AA.1 hold, SPNEs that yield more severe punishments than Nash reversion can be constructed whenever each player i's stage game Nash equilibrium payoff strictly exceeds ,!;. (You are asked to prove this result in Exercise 12.AA.2.)
the outcome path of an S I'N E using Nash reversion strategies and. by (12.AA.5), yields strictly higher discounted payolfs to the two players than does inHnite repetition of 4· = ('If. 'In. • Proposition 12.AA.I tells us that with continuous strategy sets and differentiable payolT functions, as long as there is some possibility for a joint improvement in payolTs around the stage game Nash equilibrium, some cooperation can be sustained. Going further, examination of condition (12.AA.2) tells us that cooperation becomes easier as t) grows.
qn
Proposition 12.AA.2: Suppose that outcome path Q can be sustained as an SPNE outcome path using Nash reversion when the discount rate is ,5. Then it can be so sustained for any (\' ;::: ii. In fact, as ,\ gets very large. a great number of outcomes become sustainable. The result presented in Proposition 12.AA.3, a version of the Nash reversioll/olk rheorem [originally due to Friedman (1971)], shows that allY stationary outcome path that gives «Ich player a discounted payoff that exceeds that arising from infinite repetition of the stage game Nash equilibrium q. = ('Ii, q!) can be sustained as an SPNE if ,\ is sulficiently close to I. Proposition 12.AA.3: For any pair of actions q = (q" q2) such that ";(q" q2) > ,,;(qj, qn for i = 1,2, there exists a ~ < 1 such that, for all c5 > Q, infinite repetition of q = (q" q2) is the outcome path of an SPNE using Nash reversion strategies. The proof of Proposition 12.AA.3 follows immediately from condition (12.AA.3) letting J -+ I. In fact, with a more sophisticated argument, the logic of Proposition 12.AAJ can be extended to nonstationary outcome paths. By doing so, it is possible to convcxify the set of possible payoffs identified in Proposition 12.AA.3 by alternating between various action pairs (If" 'I,). In this way, we can support any payolTs in the shaded region of Figure 12.AA.I as the average payoffs of an SPN E. .12
Proposition 12.AA.4: Consider an infinitely repeated game with [) > 0 and S; c il for i = 1, 2. Suppose also that ,,;(q) is differentiable at q. = (qf, qn. with 01!;(qj, qtJ/ciq/ i' 0 for i i' i and i = 1,2, and that ,,;(qf, q~) > ,!; for i = 1,2. Then there is some SPNE with discounted payoffs to the two players of (v;, v,,) such that (1 - il) vi < ,,;(qj, q~) for i = 1,2. Under the conditions of Proposition 12.AA.4, for any" E (0, I), more severe punishments than Nash reversion can credibly be threatened. We should therefore expect that more cooperative outcomes can be sustained than those sustainable through the threat of Nash reversion whenever a fully cooperative outcome is not already achievable using Nash reversion strategies.
Exercise I2.AA.I: Argue that no pair of actions 'I such that ",(If" 'I,) < 1!,(qj, '1j) for some i can be sustained as a stationary SPNE outcome path using Nash reversion.
More SCI'ere PUllishmellts alld the Folk Theorem It is intuitively clear that, for a given level of b < I, the more severe the punishments that can be credibly threatened in response to a deviation, the easier it is to prevent 32. See Fudcnbcrg and Maskin (1991) for details.
Possible Payoffs
,,
,,
rq, Together. (12.AA.4) and (12.AA.6) imply that. to first order, the value of the left-hand side of condition (12.AA.3) is unalfected by this change. However, (12.AA.5) implies that the right-hand side of (12.AA.3). to lirst order, increases. Hence, for a small enough change (/\t/1' Ll'!.!) in din:ction (tit/I' dt/~). infinite repetition of (tit
Supportable as SPNE Average Payoffs as b ~ I with Nash Reversion
33. In general. a player's minimax payoff will be lower if mixed strategies are allowed. In this case, the statement of the folk theorem given in Proposition 12.AA.S remains unchanged. but with
these (potentially) lower levels of ~,.
1
422
C HAP T E R
1 2:
MAR K E T
POWE
R
AP PEN 0 I X
8:
5 T RAT E G I C E N TRY
0 E T ERR ENe E
AND
Ace 0 M MOO A T ION
423
---------------------------------------------------------------,-------------------------------------------------------------6)v, .,,(1 -
,,
,,, ,,
n,(q')
!l.,
Supportable as SPNE Average Payoffs as J ~ I
--t-.,
,I· ,I --+-+---------, , !l.. ,,(q')
(i) Both finns play quantity ii in period I followed by the monopoly quantity '1 m in every period I > I as long as no one deviates, where quantity ij satisfies ",(1 - 8)1,
For arbitrary 0 < I, constructing the full set ofSPNEs is a delicate process. Each SPNE, whether collusive or punishing, uses other SPNEs as threatened punishments. For details on how this is done, see the original contributions by Abreu (1986) and (1988) and the presentation in Fudenberg and Tirole (1992). As with SPNEs using Nash reversion strategies, the full set of SPNEs grows as {) increases, making possible both more cooperation and more severe punishments. In fact, the result presented in Proposition 12.AA.S, known as the folk theorem, tells us that lilly feasible individually rational payoffs can be supported as the average payoffs in an SPNE as long as players discount the future to a sufficiently small degree. 34 (Feasibility simply means that there is some outcome path Q that generates these average payoffs.) Proposition 12.AA,5: (The Folk Theorem) For any feasible pair of individually rational payoffs (n" n 2 ) » (,!" '!2)' there exists a ~ < 1 such thaI, for all /j > §, (n" n 2 ) are the average payoffs arising in an SPNE. In comparison with Proposition 12.AA.3, Proposition 12.AA.S tells us that as I we can support any average payoffs that exceed each player's minimax payoff. 35 This limiting set of SPNE average payoffs is shown in Figure 12.AA.2. Example 12.AA.I gives some idea of how this can be done. c5
chooses quantity 'I as n(q).36 Note that '!J = 0 for j = 1,2 here; if firm j's rival chooses a quantity at least as large as the competitive quantity q, satisfying p(q,) = c, then the best firm j can do is to produce nothing and earn zero, and firm j can never be forced to a payoff worse than zero. Consider strategies for the players that take the following form:
--+
Example 12,AA,\: Sustaining an Al'erage Payoff of Zero in the Infinitely Repealed Game. In this example, we construct an SPNE in which both firms earn an average payoff of zero in an infinitely repeated Cournot game. In particular, let the stage game be a symmetric Cournot duopoly game with cost function c(q) = cq, where c > 0, and a continuous inverse demand function p(.) such that p(x) --+ 0 as x --+ Yo. It will be convenient to write a firm's profit when both firms choose quantity 'I as n(q) = [p(2q) - e]q and, as before, a firm's best-response profits when its rival
Ftgure 12.AA.2 The folk theorem.
n«i)
o
+1-"':-3 n(qm) = o.
(I2.AA.7)
(ii) If anyone deviates when ij is meant to be played, the outcome path described in (i) is restarted. (iii) If anyone deviates when '1 m is meant to be played, Nash reversion OCCurs. Note that the outcome path described in (i), if followed by both players, gives both players an average payoff of zero by construction [recall (12.AA.7)). By Proposition 12.AA.3. we know that for some ~ < I we ClIO sustain infinite m repetition of '1 through Nash reversion for all c5 > ,j. Thus, for ,) > ,), neither firm will deviate from the above strategies when '1 m -is supposed to b~ played. Will they deviate when ,j is supposed to be played" Consider firm j's payoff from deviating from ,j in a single period and conforming with the prescribed strategy thereaf1er. Firm j earns n(,7) + (0)(0) because it plays a best response when deviating, and then the original path is restarted. Thus, this deviation does not improve firm j's payoff if n(ij) = 0 (it cannot be less than zero because ,!, = 0). This is so if ij;:>: 'I,. But examining condition (12.AA.7), we see that as ,) approaches I, 11('7) must get increasingly negative for (12.AA.7) to hold and, in particular, that there exists a D, < I such that ,j will exceed q, for all /j > 0,. Thus, for ,j > Max {J" ~}, these strategies constitute an SPNE that gives both firms an average payoff of OJ7 •
COUf/lOl
34. The theorem's name refers to the fact that some version of the result was known in game theory "folk wisdom" well before its formal appearance in the literature. See Fudenberg and Maskin
APPENDIX B: STRATEGIC ENTRY DETERRENCE AND ACCOMMODATION
In this appendix, we discuss an important example of credible precommitments to affect future market conditions in which an incumbent firm engages in pre-entry capacity expansion to gain a strategic advantage over a potential entrant and possibly to deter this firm's entry altogether [the original analyses of this issue are due to Spence (1977) and Dixit (1980)). In what follows, we study the following three-stage game that is adapted from Dixit (1980)
(1986) and t1991) for a proof of the result. When there are more than two players, the result requires that the set of feasible payoffs satisfy an additional "dimensionality" condition. The original appearances of the result in the litemture actually analyzed infinitely repeated games wirluml
discounting [see, for example, Rubinstein (1979)]. 35. We may also be able in some cases to give each player exactly his minimax payoff. This is the case, for ex.ample, in the repeated Bertrand game, where (he stage game's Nash equilibrium yields the minimax payofTs. In Example 12.AA.1. we show that we can also do this for large enough i) in the repeated Cournot duopoly game.
36. We can make the strategy sets compact by noting that in no period will any firm ever
choose a quantity larger than the level q such that "(4) + [,,/(1 - ,')](Max, n(q)) = 0, because it would do better setting its quantity equal to zero forcvcr. Then, Without loss, we can let each firm choose its output from the compact set [0, ii 1
37. We have not considered any multiperiod deviations, but it can be shown that if no single-period deviation followed by conformity with the strategies is worthwhile, then neither is any multiperiod deviation (this is a general principle of dynamic programming).
424
C HAP T E R
1 2;
MAR K E T
POW E R
---------------------------------------------------------------Stage I: An incumbent, firm I, chooses the capacity level of its plant, denoted by k/. Capacity costs r per unit. Stage 2: A potential entrant, firm E, decides whether to enter the market. If it does, it pays an entry cost of F. Slaye 3: If firm E enters, the two firms choose their output levels, q/ and q" simultaneously. The resulting price is p(q/ + q£). For firm E, output costs (IV + r) per unit: for each unit of output produced, firm E incurs both a capacity cost of r and a labor cost of w. For firm I, production must not exceed its previously chosen capacity level. Its production cost, however, is only w per unit because it has already built its capacity. If, on the other hand, firm E does not enter, then firm I acts as a monopolist who can produce up to k/ units of output at cost IV per unit.
APPENDIX
B:
STRATEGIC
ENTRY
DETERRENCE
AND
ACCOMMODATION
425
~-------------------------------------------------------------
h(4 ... 1 .. + r) /
Flgur. 12.BB.2 (leH)
h,(q, I k,)
Nash Equilibrium b(q,lw) b(q,1 w t r) /'
k,
4,
4,
Firm I's stage 3 best-response function after entry.
Flgur. 12.BB.3 (right)
Stage 3 Nash equilibrium after entry.
'II.
To determine the subgame perfect Nash equilibrium (SPNE) of this game, we begin by analyzing behavior in the stage 3 sub games and then work backward. Slllfje 3: QlIalltit)' Competitiol1 Nash
The subgames in stage 3 are distinguished by two previous events: whether firm E has entered and the previous capacity choice of firm I. We first consider the outcome of stage 3 competition following entry and then discuss firm l's behavior in stage 3 if entry does not occur. For simplicity, we assume throughout that firms' profit functions are strictly concave in own quantity; a sufficient condition for this is for p(') to be concave. The concavity of p(.) also implies that firms' best-response functions are downward sloping.
,/b(q,I.·tr) 4,
Staye 3 competition after ell try. Figure 12.BB.l depicts firm E's best-response function in stage 3, which we denote by b(qlw + r) to emphasize that it is the best-response function for a firm with marginal cost w + r. Firm E's stage 3 profits decline as we move along this curve to the right (involving higher levels of q/) and, at some point, denoted Z in the figure, they fall below the entry cost F. Now consider firm l's optimal behavior. The key difference between firm I and firm E is that firm I has already built its capacity. Hence, firm I's expenditure on this capacity is sunk (it cannot recover it by reducing its capacity), its capacity level is fixed, and its marginal cost is only w. Suppose we let b(ql IV) denote the best-response function of a firm with marginal cost IV. Then firm f's best-response function in stage 3 is b,(qElk/) = Min{b('/rlll'),k,}. Figure 12.66.1
'If
Firm Ts SI:!I.!C: J Profits = F
/
-
4,
Firm E's stage J best-response function after entry.
That is, firm f's best response to an output choice of qE by firm E is the same as that for a firm with marginal cost level was long as this output level does not exceed its previously chosen capacity. Figure 12.BB.2 illustrates firm I's best-response function. We can now put together the best-response functions for the two firms to determine the equilibrium in stage 3 following firm E's decision to enter, for any given level of k,. This equilibrium is shown in Figure 12.BB.3. In Figure 12.B8.3, point A is the outcome that would arise if there were no first-mover advantage for firm I, that is, if the two firms chose both their capacity and output levels simultaneously. However, when firm I is able to choose its capacity level first, by choosing an appropriate level of k" it can get the post-entry equilibrium to lie anywhere on firm E's best-response function up to point B. Firm I is able to induce points to the right of point A because its ability to incur its capacity costs prior to stage 3 competition allows it to have a marginal cost in stage 3 of only w, rather than IV + r. Note, however, that firm I cannot induce a point on firm 2's best-response function beyond point B, even though it might want to; if it built a capacity greater than level k., it would not have an incentive to actually use all of it. Figure 12.BB.4 depicts this situation. A threat to produce up to capacity following entry would in this case not be credible. Staye 3 olltcomes if firm E does not elller. If firm E decides not to enter, then firm I will be a monopolist in stage 3. Its optimal monopoly output is then the point where its best-response function hits the q£ = 0 axis, b,(Olk,).
------_.-----------------
Flgur. 12.BB.4
A stage 3 equilibrium in which firm I does not use all of its capacity.
426
C HAP T E R
1 2:
MAR K E T
POW E R
REF ERE NeE s
--------------------------------------------------------------------- ,---
Flgur. 12.88.7 (left)
b(q, I w + r)
Entry deterrence is possible but not
b(q, I w)
Figure 12.88.5 (left)
Blockaded entry. h(q, I II'
h(OI'\'
+ r)
+ r)
inevitable.
Figure 12.88.6 (right)
b(q,I.' + r)
Strategic entry accom,m?dation When q,
q,
Sraye 2: Firm E's Encr), Decision Firm E's entry decision is straightforward: Given the level of capacity k, chosen by firm I in stage I, firm E will enter if it expects nonnegative profits net of its entry cost F. This means that firm E will enter when it expects that the postentry equilibrium will lie to the left of point Z on its best-response function in Figure 12.BB.1.
Stayl' 1: Firm /'5 Stage 1 Capacity Illvestment Now consider firm I's optimal capacity choice in stage I. There are three situations in which firm I could find itself: Entry could be blockaded, entry could be inevitable, or entry deterrence could be possible but not inevitable. Let us consider each in turn.
Elllry is blockaded. One possibility is that the entry cost F is large enough that firm E does not find it worthwhile to enter even if firm I ignores the possibility of entry and simply builds the same capacity that it would if it were an uncontested monopolist, h(OI \I" + r). This situation, in which we say that entry is blockaded, is shown in Figure 12.BB.5. [n this case, firm [ achieves its best possible outcome: it builds a capacity of h(Olw + r), no entry occurs, and then it sells b(Olw + r) units of output. £111ry delerrellce is impossible: strategic entry accommodalion. Suppose that point Z is to the right of point B. [n this case, entry deterrencc is impossible; firm E will find it profitable to enter regardless of k,. What is firm I's optimal choice of k, in this casc') In Figure 12.BB.6, we have drawn isoprofit curves for firm [; note that because these include the cost of capacity, they are the isoprofit curves corresponding to those of a firm with marginal cost (w + r). Now recall that firm [ can induce any point on firm E's best-response function up to point B through an appropriate choice of capacity. It will choose the point that maximizes its profit. In Figure 12.BB.6, this point, which involves a tangency between firm E's best-response function and firm I's isoprofit curves, is denoted as point S. This outcome corresponds to exactly the outcome that would emerge in a model of sequential quantity choice, known as a Sracklebery leadership model (see Exercise 12.C.IS). Note that firm I's first-mover advantage allows it to earn higher profits than the otherwise identical firm E. The point of tangency, S, could also lie to the right of point B. [n this case, the optimal capacity choice will be k, = k., and the outcome will not be as desirable for firm I as the Stackleberg point. Here firm I is unable to credibly
entry
IS
q,
lOevitable.
b(q,l.'
+ r)
q,
commit to produce the output associated with point S, even if it builds sufficient capacity in stage I.
Emr.\" cit'lerrl!llCc is possihll! bIll nOI ineL'ilable. Suppose now that point Z lies to the left of point B but not so far that entry is blockaded, as shown in Figure 12.BB.7. Firm I can deter firm E's entry by picking a capacity level at least as large as point kl in the figure. The only question is whether this will be optimal for firm I, or whether firm [ is better off accommodating firm E's entry. To judge this, firm [ will compare its profits at point (kz,O) to those at point S (or at point B if point S lies to the right of B). This can be done by comparing the capacity level k, in Figure 12.BB.S, the output level under monopoly that gives the same profit as the optimal accommodation point S, with k z . If k. > k z , then firm I prefers to deter entry because its profits are higher in this case; but if k, < kl' then it will prefer accommodation. Note that if deterrence is optimal, then even though entry does not occur its threat nevertheless has an effect on the market outcome, raising the level of output and welfare relative to a situation in which no entry is possible. Exercise 12.IIB.I: Show that when entry deterrence is possible but not inevitable, if point S lies to the right of point Z, then entry deterrence is better than entry accommodation.
REFERENCES Abrell, D. (19S6). Extremal equilibria 191-225.
or
oligopolistic supl!rgamcs. J(lurtla/ of Economic Theory 39:
Abreu, D. (198H)' On the theory of infinitely repeated games with discounting. Econometrica 56: 383-96.
Abreu. D .. D. Pearce. and E. Stachctti. (1990). Toward a theory of discounted repeated games with imperfect monitoring. ECrHlOmt'trica 58; 1041-64. Baumol, W .• J. Panzar. and R. Willig. {I982}. Contestahft> Markers mid the Theory of Industry Structure. San Diego: Harcourt. Brace. Jovanovich. Bertrafld, 1. (tS~U). Thcorie mathematique de la nchesse sociale. Journal des Sal'Qllts 67: 499-508. Bulow. J., J. Geanakoplos. and P. Klcmpcrer. (1985). Mullimarket oligopoly: strategic substitutes and complements. Journal of Political Economy ~3: 488-511. Chamberlin. E. (1933). The Thear,\' oj Monopolistic Competition. Cambridge. Mass.: Harvard University Press.
Figure 12.88.8 (right)
Entry deterrence versus entry accommodation.
427
428
C HAP T E R
1 2:
MAR K E T
POW E R
Cournot, A. (IS)!S). R"dlt'rcill':i .\ur It's Prindpf!.\ Mml!i'mmlqul's df! la J"IIi'orit' dt's Riches.H's. [English edition. Rt':it'tlrdll's illto llll' Mar/lt'fPlulicu/ Pri/ldpies of Ihe Tllt'orr of Weulth. edited by N. Bacon London: Macmillan, 1X97.J Di.\it. A. (19S0). The role of investment in entry dctt:rrence. £("(lIIomic" Jounwl90: 95-106. Diut. A .. and J. E. Stiglitz. (1977). Monopolistic competition and optimal product diversity. AIII..,inllJ Economic Rt't"it'\I' 67; 297-308. Edg.('worth. F. (1~97). Me tc.:oria pura del nlOnoroitO. Gion",h- clt'f/Ii L/"{}//omi\(i 40. 13-31. lEnglish tr.ln:.l:ilillll: Th(' pure theory of monopoly. In PtI/'l'r.\ ReI,""I!/ '" 1'"lakul t:C/lllllfllj', Vol I. edited by F. Edgl.'worth. London: Mal.'millan. 1925.J I-"ricdman. J. (1971). A non-cooj1\:rativc equilibrium fllr supcrgamcs. Rt,!·it,\\· oj" EClJtlomic Slucllt'J 28: 1-12. Fudcnlxrg. D .. and E. Maskin. (19X6). The folk theorem in ~cpcated games with discounting or with il1l:ompkh! information. £collomt'trim 52; 533-54. FudL·n~rg. J),
l-\rcps. D. M .. and J. Schcinkman. (19M). QU,lI1l1ly prcc.:ommilmcnl and Bcrtr.tnd c()mpctltion yield Cournot oull.:omcs. ReUle/ jOlmw/ (!r EnUiomic.\ 14: 326-37. Mall~iw. N. G .. and M. D. Whinslon. (19~6). Free entry and social inefiiclency. RUlld Journul of [COlli/miD
17: 4X 5X. M. J.. and A. Rubinstein. (1994). A Count' ill GUlllt' 71't'ory. Cambridge. Mass.: MIT Press. Rotcmbcrg. J., and G. Saloner. (1986). A supergame-thcoretic model of business cycles and price wan. during booms. Amt'rinm Enllwmk Rfrh'w 76: 390-407. RlIbin~tein. A. (1979). Equilibrium in supergames with (he overtaking criterion. Jourllu/ oj £("(1110"11("
O~borne.
Tllt'ory 21: 1-9. S;t1op. S. (1979). Monopolistic competition with outside goods. Bell Jvuma/ of /:",;onomics 10; 141-56. Shapiro. C. (19X9). Theories of oligopolY behavior. In lIul1dhook of IlIduwriul Orymli:atiml, edited by R. SI.:hmalenscc and R. D. Willig. Amsterdam: North-Holland. Spcncc. A. M. (1976). Product selection. fix.ed cosh. and monopoli~tic competition. Refit'\\· if [("(molll/(' Silnlit's 43: 217-35. Srx:nc.:c. A. M. (1977). Entry. capacity investment. and oligopolistic pricing. Bef{ Jorm",1 of cnmomics 8' 5.34-44. Stigler. G. (1960). A Iheor) of oligopoly. JOt/mal of Political Economy 72: 44-6J. Tirole. J. (19X~). nil' Ttlt·flr.\" (lr/Helustria! Or!JclIIi:llIiOlI. Cambridge. Mass.: MIT Prc.:ss.
---
EXERCtSES
12.11.2" Consider a monopolist with cost function ('(q) = eq, with c > 0, facing demand function x(p) = ~p-', where r. > O. (a) Show that if r. :5 I, then the monopolist's optimal price is not well defined. (b) Assume that r. > I. Derive the monopolist's optimal price, quantily, and price-cost margin (p. - t")/pm. Calculate Ihe resuiting deadweighl welfare loss.
(c) (Harder) Consider a sequence of demand functions thai differ in their levels of r. and ~ but that all involve the same compelitive quantity x(e) [i.e., for each level of 1:, a is adjusted to keep x(c) the same]. How does the deadweighl loss vary with F.? (If you eannol derive an analytic answer. try calculating some values on a computer.) 12.8.3" Suppose Ihat we consider a monopolist facing demand function x(p,O) with cost fUIH.:tion c«(/. lfJ). where () and. When will each lead 10 a price increase? 12.11.4" Consider a monopolist with a cosl of c per unil. Use a "revealed preference" proof 10 show Ihat the monopoly price is nondecreasing in c. Then extend your argument 10 Ihe case in which the monopolisl's cosl function is e(q,4», with [<"('1",4» - e(q', 4>)] increasing in 4> for all 'I" > 'I', by showing thai Ihe monopoly price is nondecreasing in 4>. (If you did Exercise 12.B.3, also rehlle this condilion 10 Ihe one you derived Ihere.) 12.S.5" Suppose Ihat a monopolist faees many consumers. Argue that in each of the following two cases. the monopolist can do no beller Ihan il does by reslricling itself to simply charging a price per unit, say p. (a) Suppose that each consumer i wants eilher one or no units of Ihe monopolist's good and that the monopolist is unable to discern any particular consumer's preferences. (b) Suppose that consumers way desire to consume multiple unils of Ihe good. The monopolist cannot discern any parlicular consumer's preferences. In addition, resale of Ihe good is cost less and after the monopolisl has made ils sales to consumers a compelitive market develops among consumers for the good. 12.B.6'\ Suppose that the government can tax or subsidize a monopolist who faces inverse demand funclion p(q) and has cost function c(q) [assume bOlh are differentiable and Ihat p(q)({ - c(q) is concave in 'I]. What tax or subsidy per unit of output would lead the monopolist to act elliciently? 12.8.7" Consider the widget markel. The total demand by men for widgets is given by = II - (J.P, and the total demand by women is given by x •. (p) = a - O•. P, where O. < 0•. Tht!' cost or production is c per widget. xm(p)
(a) Suppose the widget market is competitive. Find the equilibrium price and quantity sold.
EXERCISES 12.8.1" The expression [p. - ("(i/"')]/p., where p. and q. arc the monopolist's price and output level. respectively. is known as the monopolist's price-cost marg;" (or as the Lt'l'II!'r ;1/,1£'.'( of monopoly pOl\'('r). It measures the distortion of the monopolist's price above its 1l1~lrginal cost as a proportion of its price.
(b) Suppose, instead, that firm A is a monopolist of widgets [also make this assumption in (e) and (d)]. If linn A is prohibited from "discriminating" (i.e., charging different prices to men and women), what is its profit-maximizing price? Under what conditions do both men and women consume a positive level of widgets in this solution'!
the inverse of the price
(c) If firm A has produced some 10lallevel of OUlpul X, what is Ihe welfare·maximizing way to distribute it between the men and the women? (Assume here and below thai Marshallian aggregrate surplus is a valid measure of welfare.)
(b) Also argue that if the monopolist's marginal cost is positive at every outpullevel, then demand must be ellIsl;(' (i.e., the price elaslicity of demand is greater Ihan I) al the monopolist's optimal price.
(d) Suppose that firm A is allowed to discriminate. What prices does it charge? In Ihe case where the nondiscriminatory solution in (b) has posilive consumption of widgets by bOlh men and women, does aggregate welfare as measured by the Marshallian aggregate surplus rise or
(a) Show the monopolist's price-cost margin is always equal elasticity of demand at price pm.
to
429
~-------------------------------------------------------------------
430
CHAPTER
12:
MARKET
POWER
fall relative to when discrimination is allowed? Relate your conclusion to your answer in (c). What if the nondiscriminatory solution in (b) has only one type of consumers being served? I2.B.S" Consider the following two·period model: A firm is a monopolist in a market with an inverse demand function (in each period) of p(q) = a - bq. The cost per unit in period I is c,. In period 2, however, the monopolist has "learned by doing," and so its constant cost per unit of output is ('2 = C 1 - mql. where ql is the monopolist's period I output level. Assume u > c and h > m. Also assume that the monopolist does not discount future earnings. (a) What is the monopolist's level of output in each of the periods? (b) What outcome would be implemented by a benevolent social planner who fully controlled the monopolist? Is there any sense in which the planner's period I output is sciected so that" price equals marginal cost "? (c) Given that the monopolist will be selecting the period 2 output level, would the planner like the monopolist to slightly increase the level of period I output above that identified in (a)? Can you give any intuition for this?
--
EXERCISES
(b) Examine a model in which prices must be named in discrete units, as in Exercise 12.C.3. What are the pure strategy Nash equilibria of such a game? Which do not involve the play of weakly dominated strategies? As the grid becomes finer, what is the limit of these equilibria in undominated strategies? 12.C.S" Suppose that we have a market with J buyers, each of whom wants at most one unit of the good. Buyer i is willing to pay up to v, for his unit, and v, > v, > ... > v,. There are a total of q < J units available. Suppose that buyers simultaneously submit bids for a unit of the output and that the output goes to the q highest bidders, who pay the amounts of their bids. Show that every buyer making a bid of vf . , and the good being assigned to buyers I, ... , q is a Nash equilibrium of this game. Argue that this is a competitive equilibrium price. Also show that in all)' pure strategy Nash equilibrium of this game, buyers I through q receive a unit and buyers q + I through J do not. 12.C.6 A In text. 12.C.7" In tex t.
I2.R.9< Consider a situation in which there is a monopolist in a market with inverse demand function p(q). The monopolist makes two choices: How much to invest in cost reduction, I, and how much to sell, q. If the monopolist invests I in cost reduction, his (constant) per· unit cost of production is c(l). Assume that <'(I) < 0 and that <"(I) > o. Assume throughout that the monopolist's objective runction is concave in q and I. (a) Derive the first·order conditions for the monopolist's choices.
C
12.C.8 Consider a homogeneous-good J·firm Cournot model in which the demand function x(p) is downward sloping but otherwise arbitrary. The firms all have an identical Cost function c(q) that is increasing in q and convex. Denote by Q the aggregate output of the J firms, and let Q-, = L. •• i q•. (a) Show that firm j's best response can be written as b(Q _ j).
(b) Compare the monopolist's choices with those of a benevolent social planner who can control both q and 1 (a "first-best" comparison).
(b) Show that h(Q _ j) need not be unique (i.e., that it is in general a correspondence, not a function).
(e) Compare the monopolist's choices with those of a benevolent social planner who can cantrall but not q (a "second·best" comparison). Suppose that the planner chooses 1 and then the monopolist chooses q.
(e) Show that if Q_j > Q- i , q, E h(Q_j), and qjE b(Q -i)' then (4 j + Q-i) ~ (qj + Q_j). Deduce from this that h(') can jump only upward and that b'(Q _i) ~ -I whenever this derivative is defined.
12.B.10· Consider a monopolist that can choose both its product's price p and its quality q. The demand for its product is given by x(p, q), which is increasing in q and decreasing in p. Given the price chosen by the monopolist, does the monopolist choose the socially efficient quality level?
(d) Use you result in (e) to prove that a symmetric pure strategy Nash equilibrium exists in this model. (e) Show that multiple equilibria are possible. (f) Give sufficient conditions (they are very weak) for the symmetric equilibrium to be the only equilibrium in pure strategies.
12.C.IA In text. 12.C.2C Extend the argument of Proposition 12.C.1 to show that under the assumptions made in the text [in particular, the assumption that there is a price p < co such that x(p) = 0 for all p ~ p], both firms setting their price equal to c with certainty is the unique Nash equilibrium of the Rertrand duopoly model even when we allow for mixed strategies.
(a) Derive the Nash equilibrium of this model. Under what conditions does it involve only one firm producing? Which will this be' (b) When the equilibrium involves both firms producing. how do equilibrium outputs and profits vary when firm I's cost changes?
(a) Show that both firms naming prices equal to the smallest multiple of A that is strictly greater than c is a pure strategy equilibrium of this game. Argue that it does not involve either firm playing a weakly dominated strategy.
(e) Now consider the general case of J firms. Show that the ratio of industry profits divided by industry revenue in any (pure strategy) Nash equilibrium is exactly H 1£, where £ is the elasticity of the market demand curve at the equilibrium price and H, the Herfindahl index of concelltration, is equal to the sum of the firms' squared market shares L.i(q;lQ*)'. (Note: This result depends on the assumption of constant returns to scale.)
12.C.4" Consider altering the Bertrand duopoly model to a case in which each firm j's cost per unit is ('j and C 1 < C2' (a) What are the pure strategy Nash equilibria of this game?
-
12.C.9" Consider a two·firm Cournot model with constant returns to scale but in which firms' costs may differ. Let c j denote firm j's cost per unit of output produced, and assume that c, > c,. Assume also that the inverse demand function is p(q) = a - bq, with a > c,.
12.C.3" Note that the unique Nash equilibrium of the Rertrand duopoly model has each firm playing a weakly dominated strategy. Consider an alteration of the model in which prices must be named in some discrete unit of account (e.g., pennies) of size A.
(b) Argue that as A _ 0, this equilibrium converges to both firms charging prices equal to c.
431
,-------------------------------------------------------------~
12.C.10· Consider a J·firm Cournot model in which firms' costs differ. Let Cj(qi) = ajc(qj) denote firm j's cost function, and asSUme that c(·) is strictly increasing and convex. Assume that 17 1 >'" > (XJ.
432
CHAPTER
12:
MARKET
POWER
~~~--------------------------------------------------------(a) Show that if more than one firm is making positive sales in a Nash equilibrium of this model, then we cannot have productive efficiency; that is, the equilibrium aggregate output Q* is produced inefticiently.
--
(c) Provide an example in which wclfare decreases when a firm becomes more productive (i.e .. when ':X J falls for somej). [Him: Consider an improvement in cost for IIrm I in the model of Exercise 12.C.9.] Why can this happen?
12_C.16" Derive the Nash equilibrium prices and profits in the circular city model with J firms when travel costs are quadratic, as in Exercise 12.C.IS. Restrict attention to the case in which v is large enough that the possibility of non purchase can be ignored. What happens as J grows large! As I falls?
12.C.11 (' Consider a capacity-constrained duopoly pricing game. Firm j's capacity is 'Ij for j = 1,2, and it has a constant cost per unit of output of (' 0 and that there exists a price p such that x(f» = 'I, + 'I,. Suppose also that x(p) is concave. Let p(.) = x - '(.) denote the inverse demand function. Given a pair of prices charged. sales are dt:termined as follows: consumers try to buy at the low·priccd firm first. If demand exceeds this firm's capacity, consumers are servcd in order of their valuations. starting with high· valuation consumers. If prices are the same, demand is split evenly unless one firm's demand exceeds its capacity, in which case the extra demand spills over to the other firm. Formally, the firms' sales arc given by the functions x,(p" p,) and x,(p"p,) satisfying [x i(·) gives the amount firm i sells taking account of its capacity limitation in fullilling demand]
12,C.17" Consider the linear city model in which the two firms may have different constant unit production costs c, > 0 and c, > O. Without loss of generality, take c, OS; c, and suppose that I' is large enough that non purchase can be ignored. Determine the Nash equilibrium prices and sales levcls for equilibria in which both firms make strictly positive sales. How do local changes in c, affect the equilibrium prices and profits of firms I and 2? For what values of c, and c, docs the equilibrium involve one firm making no sales? 12.C.IH" (The SI"ckl"herli leaderxhip model) There are two firms in a market. Firm I is the -'leader" and picks its quantity lirst. Firm 2, the "follower," observes firm I's choi<.:c and then chooses its quantity. Profits for each firm i given quantity choices q. and q2 are p('1, + '1')'li - (''Ii' where P'('/) < 0 and p'(q) + p"(q)q < 0 at all q ~ O.
X,(p" p,) = Min!'I" x(p,): Xj(p"p,) = Min{qj' Max!x(pj) - q"O}}
If p,
= p, = p:
xi(p" p,)
= Min{qi' Max{x(p)/2, x(p)
- 'Ij}}
for i
(e) Compare the change in equilibrium prices and profits from a reduction in t in the case studied in (d) with that in the equilibria of (a) and (b). 12_C.15" Derive the Nash equilibrium prices of the linear city model where a consumer's travel cost is quadratic in distance, that is, where the total cost of purchasing from firm j is Pi + rd', where d is the consumer's distance from firm j. Restrict attention to the case in which v is large enough that the possibility of nonpurchase can be ignored.
(b) If so, what is the correct measure of welfare loss relative to a fully ellicient (competitive) outcome') [/lilll: Rcconsider the discussion in Section 10.E.]
If P, > Pi:
EXERCISES
= 1,2.
(a) Prove formally that firm I's quantity choice is larger than its quantity choice would
be if the firms chose quantities simultaneously and that its profits are larger as well. Also show
(a) Suppose that 'I, < b,(q,) and 'I, < h,(q,), where h,(-) is the best-response function for a firm with cons!;)nt marginal costs of c. Show that pT = pl = p(q, + q,) is a Nash equilibrium of this gamc.
that aggregate output is larger and that lirm 2's profits are smaller. (b) Draw a picture of this outcome using best-response functions and isoprofit contours.
(b) Argue that ifeither ,/, > h,(q,) or ,/, > b,('/,), then no pure strategy Nash equilibrium exists. 12.C.121\ Consider two strictly concave and differentiable profit functions ",('1" 'I.), j defined on 'Ij E [0, 'I].
=
12.C.19 C Do Exercise 8.B.S. 12.C.20" Prove Proposition 12.C.2 for the case of a general convex cost function c(q).
1,2,
12_D_1 II Consider an infinitely repeated Bertrand duopoly with discount factor 0 < 1. Determine the conditions under which strategies of the form in (12.D.l) sustain the monopoly price in each of the following cases:
(a) Give sulfident conditions for the best·rcsponsc functions bj(qj} to be increasing or decreasing.
(b) Specialize to the Coumot model. Argue that a decreasing (downward-sloping) bestresponse function is the "norma'" casco
(a) Market demand in period t is x,(p) =
i x(p)
where y > O.
(b) At the end of each period, the market ceases to exist with probability y. 12.C.13" Show that when I' > I' + 31 in the linear city model discussed in Example 12.C.2, a Jirm j's best response to any price of its rival p _i always results in all consumers purchasing
(e) It takes K periods to respond to a deviation.
frol11 one of the two firms. 12.C.l4
C
12.0.2" In text.
Consider the linear city model discussed in Example 11.C.2.
(a) Derivc the bcst·rcsponsc functions when equilibrium in this case is = p! = c + l.
l' E
I'r
(b) Repeat (a) for the case in which v E (c
+
(c
+
2f, c
I' -
+ 1')/2
(a) Under what conditions can the symmetric joint monopoly outputs (q" q,) = (q~/2, q~/2)
l/, c + 21).
l' < f + t. the unique Nash equilibrium involves prices of and some consumers not purchasing from either firm.
(e) Show that when (v
12.D.3" Consider an infinitely repeated Coumot duopoly with discount factor b < I, unit costs of c > 0, and inverse demand function p(q) = a - bq, with a > c and b > O.
+ 31). Show that the unique Nash
(d) Show that when v E (c + I, C + 11), the unique symmetric equilibrium is 112. Are there asymmetric equilibria in this case?
be sustained with strategies that call for (q~12, q~ 12) to be played if no one has yet deviated
pi = pi =
and for the single-period Cournot (Nash) equilibrium to be played otherwise?
pt
(b) Derive the minimal level of <5 such that output levels (q" q,) = (q, q) with q E [«a - c)/2b), «a - c)lb)] are sustainable through Nash reversion strategies. Show that this level of 0,0('1), is an increasing, differentiable function of q.
=
pl
=
j
433
434
CHAPTER
12:
MARKET
POWER
---------------------------------------------------------------------12.0.4" Consider an infinitely repeated Bertrand oligopoly with discount factor"
E
n, I).
(a) If the cost of production changes, what happens to the most profitable price that can
--
EXERCISES
cost of i' reached at ij. Show that if there exists a J' such that J 0ij = Xli'), then any equilibrium of this model produces the perfectly competitive outcome and, hence, the outcome is (first-best) enicient.
be sustained?
12.F.I" Show that in the Cournot model discussed in Section 12.F with demand function
(b) Suppose, instead, that the cost of production will increase permanently in period 2 (i.e., from period 2 on, it will be higher than in period I). What effect does this have on the maximal
n(p), a firm's best-response function b(Q-;l is (weakly) decreasing in Q_j provided, is
price that can be sustained in period I?
large enough.
12.D.SC [Based on Rotemberg and Saloner (1986)] Consider a model of infinitely repeated Bertrand interaction where in each period there is a probability i. E (0, I) of a "high-demand" state in which demand is x(p) and a probability (I - i.) of a "low-demand" state in which demand is n(p), where, E (0, I). The cost of production is c > per unit. Consider Nash reversion strategies of the following form: charge price p" in a high-demand state ifno previoLis deviation has occurred, charge PL in a low-demand state if no previous deviation has occurred, and set price equal to c if a deviation has previously occurred.
t2.F.2" Suppose each of the J consumers in the economy has quasilinear preferences and a demand function for good ( of -'hlP) = a - bp.
°
(a) Derive the market inverse demand function. (b) Now consid~r a COllrnat entry model with this market inverse demand function. technology c(q) = cq, and entry cost K. Analyze what happens to the equilibrium prices and output levels, as well as what happens to consumer welfare (measured by consumer surplus), as I -- Cf.., for both a one·stagc and a two-stage entry model.
(a) Show that if " is sufficiently high, then there is an SPNE in which the firms set
p"
= p, = pm,
the monopoly price.
(b) Show that for some ~ above ~, a firm will want to deviate from price P'" in the high-demand state whenever J < §. Identify the highest price p" that the firms can sustain when ,j E [l, ,j) (verify that they can still sustain price P,. = pm in the low-demand state). Notice that this equi-librium may involve "countercyclical" pricing; that is, P,. > PII' Intuitively, what drives this resuh? (c) Show that when (5 <
! we
must have PII = PI. =
C.
12.F:.l" Suppose that we have a two-stage model of entry into a homogeneous-good market characterized by price competition. If potential firms differ in efficiency, need the equilibrium have the most efficient firm being active? 12.E.2" Prove that nJ is decreasing in J under assumptions (A I) to (A3) of Proposition 12.E.I. 12.E.3" Calculate the welfare loss from the free-entry equilibrium number of firms relative to the socially optimal number of firms in the models discussed in Examples 12.E.I and 12.E.2.
12.F.3" Analyze the two-stage Cournot entry model discussed in Section 12.F when ::r remains flxcd hut 1\ __ o. Show. in particulur, that the welfare loss goes to zero. 12.F.48 Consider the following two-stage entry model with differentiated products and price competition following entry: All potential firms have zero marginal costs and an entry cost of K > o. In stage 2. the demand function for firm j as a function of the price vector p = (p, •. .. PJ) of the J active firms is Xj(p) = '[i' -/I(Jp/LI P.)]. Analyze the welfare properties as the size (:;l) and the substitution (fl) parameters change. 12.G.l" Consider the linear inverse demand Cournot duopoly model and the linear city dilTerentiated-price duopoly model with differing unit costs that you examined in Exercises 12.C9 and 12.CI7. Find the derivative, with respect to a change in firm J's unit cost, affirm 2's equilibrium quantity in the Cournot model and equilibrium price in the linear city model. In which model is this change in firm 2's behavior beneficial to firm I? 12.AA.IA In text.
What happens to this loss as K ~ 0" 12.E,4" Consider a two-stage model of entry in which all potential entrants have a cost per unit of (" (in additional to an entry cost of K) and in which, whatever number of firms enter, a perfect cartel is formed. What is the socially optimal number of firms for a planner who cannot control this cartel behavior? What are the welfare consequences if the planner cannot
12.AA.2C Prove Proposition 12.AAA. [I/illl: Consider a strategy profile of the following form: the players arc to play an outcome path involving some pair (qt' q,) in period I and «Ii, qj) in every period thereafter. If either player deviates, this outcome path is restarted.] 12.BB.1A In text.
control entry? I2.E.SC Consider a two-stage entry model with a market that looks like the market in Exercise 12.CI6. The entry cost is K. Compare the equilibrium number of firms to the number that a planner would pick who can control (a) entry and pricing and (b) only entry. I2.E.6" Compare a one-stage and a two-stage model of entry with Cournot competition [all potential entrants arc identical and production costs arc c(q) = cq]. Argue that any (SPNE) equilibrium outcome of the two-stage game is also an outcome of the one-stage game. Show by example thal the reverse is not true. Argue that we cannot, however, have more firms active in the one-stage game than in the two-stage game. 12,E.7" Consider a one-stage entry model in which firms announce prices and all potential firms have average costs of AC('l) (including their. fixed setup costs) with a minimum average
=
12.1IB,2" Show that if the incumbent in the entry deterrence model discussed in Appendix II is indifTercnt between deterring entry and accommodating it, social wclfare is strictly greater if he chooses deterrence. Discuss generally why we might not be too surprised if entry deterrence could in somc cases raise social welfare. I2,BII.3 C Consider the linear city model of Exercise 12.C2 with v> (. + 31. Suppose that firm I enters the market first and can choose to set up either one plant at onc end of the city or two plants, one at each end. Each plant costs F. Then firm E decides whether to enter (for simplicity, restrict it to building one plant) and at which end it wants to locate its plant. Determine the equilibrium of this model. How is it affected by the underlying parameter values? Compare the welfare of this outcome with the welfare if there Were no entrant. Compare with the case where there is an entrant but firm I is allowed to build only one plant.
435
C
Adverse Selection, Signaling,
HAP
T
E
R
SEC T ION
13
13.A Introduction One of the implicit assumptions of the fundamental welfare theorems is that the characteristics of all commodities are observable to all market participants. Without this condition, distinct markets cannot exist for goods having differing characteristics, and so the complete markets assumption cannot hold. In reality, however, this kind of information is orten asymmetrically held by market participants. Consider the following three examples: (i) When a firm hires a worker, the firm may know less than the worker does about the worker's innate ability. (ii) When an automobile insurance company insures an individual, the individual may know more than the company about her inherent driving skill and hence about her probability of having an accident. (iii) In the used-car market, the seller of a car may have much better information about her car's quality than a prospective buyer does.
436
I N FOR MAT ION A l A S Y M MET R I E SAN 0
A 0 V E R S ESE LEe T ION
will be low. Moreover, this fact may even further exacerbate the adverse selection problem: If the price that can be received by selling a used car is very low, only sellers with really bad cars will ofTer them for sale. As a result, we may see little trade in markets in which adverse selection is present, even if a great deal of trade would occur were information symmetrically held by all market participants. We also introduce and study in Section 13.B an important concept for the analysis of market intervention in settings of asymmetric information: the notion of a coIISrmilled Pareto oprimal allocarioll. These are allocations that cannot be Pareto improved upon by a central authority who, like market participants, cannot observe individuals' privately held information. A Pareto-improving market intervention can be achieved by such an authority only when the equilibrium allocation fails to be a constrained Pareto optimum. In general, the central authority's inability to observe individuals' privately held information leads to a more stringent test for Paretoimproving market intervention. I n Sections 13.C and 13.0, we study how market behavior may adapt in response to these informational asymmetries. In Section \3.C, we consider the possibility that informed individuals may find ways to sigllill information about their unobservable knowledge through observable actions. For example, a seller of a used car could ofTer to allow a prospective buyer to take the car to a mechanic. Because sellers who have good cars are more likely to be willing to take such an action, this offer can serve as a signal of quality. In Section 13.0, we consider the possibility that uninformed parties may develop mechanisms to distinguish, or screen, informed individuals who have dilTcring information. For example, an insurance company may ofTer two policies: one with no deductible at a high premium and another with a significant deductible at a much lower premium. Potential insureds then self-selecr, with high-ability drivers choosing the policy with a deductible and low-ability drivers choosing the no-deductible policy. In both sections, we consider the welfare characteristics of the resulting market equilibria and the potential for Pareto-improving market intervention. For expositional purposes, we present all the analysis that follows in terms of the labor market example (i). We should nevertheless emphasize the wide range of settings and fields within economics in which these issues arise. Some of these examples are developed in the exercises at the end of the chapter.
and Screening
A number of questions immediately arise about these settings of asymmetric illjomlllrioll: How do we characterize market equilibria in the presence of asymmetric information? What are the properties of these equilibria? Are there possibilities for welfare-improving market intervention? In this chapter, we study these questions, which have been among the most active areas of research in microeconomic theory during the last twenty years. We begin, in Section 13.B, by introducing asymmetric information into a simple competitive market model. We see that in the presence of asymmetric information, market equilibria often fail to be Pareto optimal. The tendency for inefficiency in these settings can be strikingly exacerbated by the phenomenon known as adverse selecrioll. Adverse selection arises when an informed individual's trading decisions depend on her privately held information in a manner that adversely affects uninformed market participants. In the used-car market, for example, an individual is more likely to decide to sell her car when she knows that it is not very good. When adverse selection is present, uninformed traders will be wary of any informed trader who wishes to trade with them, and their willingness to pay for the product offered
1 3 • B:
437
----------------------------------------------------~----~~~
!
13.B Informational Asymmetries and Adverse Selection Consider the following simple labor market model adapted from Akerlof's (1970) pioneering work: I there arc many identical potential firms that can hire workers. Each produces the same output using an identical constant returns to scale teChnology in which labor is the only input. The firms arc risk neutral, seek to maximize their expccted prolits, and act as price takers. For simplicity, we take the price of the firms' output to equal I (in units of a numeraire good). Workers difTer in the number of units or output they produce if hired by a firm, I. Akcrlof (1970) used the example of a used-car market in which only the setler of a used car knows if the car is a "lemon." for this reason, (his type: of model is sometimes referred to as a It'mon.~ model.
I
1
438
C HAP T E R
1 3:
A 0 V E R 5 ESE L E C T ION
I
S I G N A LIN G,
AND
S CREE NI N G
SECTION
---------------------------------------------------------------which we denote by 0.' We let [Q, 0] c R denote the set of possible worker productivity levels, where 0 ~ Q< Ii < 00. The proportion of workers with productivity of 0 or less is given by the distribution function F(O), and we assume that F(') is nondegenerate, so that there are at least two types of workers. The total number (or, more precisely, measure) of workers is N. Workers seek to maximize the amount that they earn from their labor (in units of the numeraire good). A worker can choose to work either at a firm or at home, and we suppose that a worker of type 0 can earn r(O) on her own through home production. Thus, r(O) is the opportunity cost to a worker of type 0 of accepting employment; she will accept employment at a firm if and only if she receives a wage of at least r(O) (for convenience, we assume that she accepts if she is indifferen!).' As a point of comparison, consider first the competitive equilibrium arising in this model when workers' productivity levels are publicly observable. Because the labor of each different type of worker is a distinct good, there is a distinct equilibrium wage 11"(0) for each type O. Given the competitive, constant returns nature of the firms, in a competitive equilibrium we have IV'(O) = 0 for all 0 (recall that the price of their output is I), and the set of workers accepting employment in a firm is
13.8:
INFORMATIONAL
ASYMMETRIES
e(l\') =
~
{II: r(O)
Consider, next, the demand for labor as a function of IV. If a firm believes that the average productivity of workers who accept employment is fl, its demand for labor is given by if" < if I'
I\'
= I\'
if I' >
(13.B.3)
1\'.
Now, if worker types in set e' are accepting employment offers in a competitive equilibrium, and if firms' beliefs about the productivity of potential employees correctly renect the actual average productivity of the workers hired in this equilibrium, then we must have I' = £[0 I 0 E eo]. Hence, (I3.B.3) implies that the demand for labor can equal its supply in an equilibrium with a positive level of employment if and only if I\' = £[0 10 E eo]. This leads to the notion of a competitive equilibrium presented in Definition 13.B.1.
(I3.B.I)
(This is simply the total revenue generated by the workers'labor.)5 Aggregate surplus is therefore maximized by setting 1(0) = I for those () with r(O) ~ 0 and 1(0) = 0 otherwise (we again resolve indifference in favor of working at a firm). Put simply,
Definition 13.B.1: In the competitive labor market model with unobservable worker productivity levels, a competitive equilibrium is a wage rate w' and a set e' of worker types who accept employment such that
2. A worker's productivity could be random without requiring any change in the analysis th<1t follows; in this case, 0 is her expected (in a statistical sense) level of productivity. 3. An equivalent model arises from instead specifying r(O) as the disutility of labor. In this alternative model, a worker of type 0 has quasilinear preferences of the form u(m,1) = m - r(O)/, where In is the worker's consumption of the numeraire good and 1 E {O,I} is a binary variable with 1 = I if the worker works and 1 = 0 if not. With these preferences, a worker again accepts employment if and only if she receives a wage of at least r(O), and the rest of our analysis remains unaltered. 4. More precisely, there are also competitive equilibria in which ",*(0) = 0 for all types of workers
e' =
{o: r(O)
~
w'}
(13.B.4)
e'],
(13.B.5)
and w' = £[0 I 0 E
Condition (11B.5) involves ratio/lal <'xp<'cClllio/ls on the part of the firms. That is, firms correctly anticipate the average productivity of those workers who accept employment in the equilibrium. Note, however, that the expectation in (13.B.5) is not well defined when /10 workers arc accepting employment in an equilibrium (i.e., when e' = 0). In the discussion that follows. we assume for simplicity that in this circumstance each firm's expectation of potential employees' average productivity is simply the unconditional expectation £[11], and we take 1\" = £[0] in any such equilibrium. (As discussed in footnote 4, we restrict attention to wages that equal workers' expected productivity in any no-trade equilibrium. See Exercise 13.B.5 for the consequences of altering the assumption that expected productivity is £[11] when e' = 0.)
who are employed in the equilibrium [those with reO) " 0] and IV'(O) ;, 0 for those types who are not [those with r(O) > 0]. However, for the sake of expositional simplicity, when discussing competitive equilibria that involve no trade in this section we shall restrict attention to equilibrium wages that are equal to workers' (expected) productivity. 5. In Section to.E, the aggregate surplus from an allocation in a product market (where firms produce output) was written as consumers' direct benefits from consumption of the good less firms' total costs of production. Here, in a labor market setting, a firm's "cost" of employing a worker is the positive revenue it earns, and a worker receives a direct utility (exclusive of any wage payments) of 0 if she works for a firm and r(O) if she does not. Hence, aggregate surptus in these markets is equal to firms' total revenues, N 1(0)0 dF(O), plus consumers' total revenue from home production, N(t - /(O))r(O) dF(O).
J
(I3.B.2)
I\'}.
As would be expected from the first fundamental welfare theorem, this competitive outcome is Pareto optimal. To verify this, recall that any Pareto optimal allocation of labor must maximize aggregate surplus (see Section IO.E). Letting 1(0) be a binary variable that equals I if a worker of type Ii works for a firm and 0 otherwise, the sum of the aggregate surplus in these labor markets is equal to
N[l(O)O + (I -1(O))r(O)] dF(O).
ADVERSE
since a type () worker produces at least as much at a firm as at home if and only if r(O) ~ 0, in any Pareto optimal allocation the set of workers who are employed by the firms must be {O: r(O) ~ O}. We now investigate the nature of competitive equilibrium when workers' productivity levels are ullobservable by the firms. We begin by developing a notion of competitive equilibrium for this environment with asymmetric information. To do so, note first that when workers' types are not observable, the wage rate must be independent of a worker's type, and so we will have a single wage rate IV for all workers. Consider, then, the supply of labor as a function of the wage rate IV. A worker of type II is willing to work for a firm if and only if r(O) ~ w. Hence, the set of worker types who are willing to accept employment at wage rate IV is
{O: r(O) ~ O}.'
r
AND
J
i
J
--~----~----
~~~-~----~------~~
SELECTION
439
440
CHAPTER
13:
ADVERSE
SELECTION.
SIGNALING,
AND
SCREENING
Asymmelric Informalion and Pareto Inefficiency
Typically, a competitive equilibrium as defined in Definition 13.B.1 will fail to be Pareto optimal. To see this point in the simplest-possible setting, consider the case where r(O) = r for all 0 (every worker is equally productive at home) and suppose that F(r) E (0, I), so that there are some workers with 0> r and some with 0 < r. In this setting, the Pareto optimal allocation of labor has workers with 0 ~ r accepting employment at a firm and those with 0 < r not doing so. Now consider the competitive equilibrium. When r(O) = r for all 0, the set of workers who are willing to accept employment at a .given wage, e(w), is either [Q, 0] (if w ~ r) or 0 (if w < rl. Thus, £[010 E e(wl] = £[0] for all wand so by (I3.B.5) the equilibrium wage rate must be w' = £[0]. If £[0] ~ r, then al/ workers accept employment at a firm; if £[/1] < r, then none do. Which type of equilibrium arises depends on the relative fractions of good and bad workers. For example, if there is a high fraction of low-productivity workers then, because firms cannot distinguish good workers from bad, they will be unwilling to hire any workers at a wage rate that is sunicicnt to have them accept employment (i.e., a wage of at least r). On the other hand, if there arc very few low-productivity workers, then the average productivity of the workforce will be above r, and so the firms will be willing to hire workers at a wage that they arc willing to accept. In one case, too many workers arc employed relative to the Pareto optimal allocation, and in the other too few. The cause of this failure of the competitive allocation to be Pareto optimal is simple to see: because firms are unable to distinguish among workers of differing productivities, the market is unable to allocate workers efficiently between firms and home production"
--
SECTION
13.8:
INfORMATIONAL
ASYMMETRIES
AND
ADVERSE
SELECTION
441
I. . - - - - - - - - - - 45'
!
I I I I
£[/1]
~ !
£[Olr(O):5"]
I
Figure 13.B.1
I I I I
r@
A compelitive equilibrium wilh adverse selection.
'I'
not so; indeed, the market may fail completely despite the fact that every worker type should work at a lirm. To see the power of adverse selection, suppose that r(O)!> 0 for all 0 E [Q, 0] and that r(') is a strictly increasing function. The first of these assumptions implies that the Pareto optimal labor allocation has every worker type employed by a firm. The second assumption says that workers who are more productive at a firm arc also more productive at home. It is this assumption that generates adverse selection: Because the payolT of home production is greater for more capable workers, only less capable workers accept employment at any given wage "' [i.e., those with r(O) !> "l The expected value of worker productivity in condition (13.B.5) now depends on the wage rate. As the wage rate increases, more productive workers become willing to accept employment at a firm, and the average productivity of those workers accepting employment rises. For simplicity, from this point on, we assume that F(') has an associated density function [(.), with [(0) > for all 0 E [Q, 0]. This insures that the average productivity of those workers willing to accept employment, E[O 1 1'(0) !> w], varies continuously with the wage rate on the set"' E [r@,oo]. To determine the equilibrium wage, we use conditions (13.B.4) and (11B.5). Together they imply that the competitive equilibrium wage IV' must satisfy
°
Adverse Seieclioll and Markel Ullravelillg
A particularly striking breakdown in efficiency can arise when r(O) varies with O. In this case, the average productivity of those workers who arc willing to accept employment in a firm depends on the wage, and a phenomenon known as adverse select;oll may arise. Adverse selection is said to occur when an informed individual's trading decision depends on her unobservable characteristics in a manner that adversely alTects the uninformed agents in the market. In the present context, adverse selection arises when only relatively less capable workers are willing to accept a firm's employment offer at any given wage. Adverse selection can have a striking effect on market equilibrium. For example, it may seem from our discussion of the case in which r(O) = r for all 0 that problems arise for the Pareto optimality of competitive equilibrium in the presence of asymmetric information only if there are some workers who should work for a firm and some who should not (since when either 0 < r or Q> r the competitive equilibrium outcome is Pareto optimal). In fact, because of adverse selection, this is
w' = £[01 r(O) !> 11"].
(13.8.6)
We can use Figure I3.B.1 to study the determination of the equilibrium wage w·. There we graph the values of £[0 1r(O) !> w] as a function of w. This function gives the expected value of Ii for workers who would choose to work for a firm when the prevailing wage is II'. It is increasing in the level IV for wages between r(O) and r(ii), has a minimum value of Q when II' = r(q), and attains a maximum value ~f £[0] for IV ~ 1'(1)).7 The competitive equilibrium wage 11" is found by locating the wage rate at which this function crosses the 45-degree line; at this point, condition (13.B.6) is satisfied. The set of workers accepting employment at a firm is then e' = {o: r(O)!> w<}. Their average productivity is exactly w·. 8
6. Anolht:r way (0 understand the diHiculty here is that asymmetric inrormation leads to a situation with missing markets and thereby creates externalities (recall Chapter 11). When a worker of Iype /I > £[0] = ... marginally reduces her supply of tabor 10 a firm here, Ihe firm is made worse olT, in contrast with the situation in a competitive market with perfect information, where the wage exactly equals a worker's marg.inal productivity.
7. The figure does nol depici Ihis funclion for wages below r@. Because £[0] > r(q) in this model. no wage below r(q) can be an equilibrium wage under our assumption that £[/118(",) = 0] = £[0]. 8. For another diagrammatic determination of equilibrium, see Exercise 118.1.
J
442
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING.
AND
SECTION
SCREENING
13.B:
INFORMATIONAL
ASYMMETRIES
AND
ADVERSE
SELECTION
443
--------------------------------------------------------------------- ,--------------------------------------------------------------------45" 45'
I\"*
£[0]
,, ,,
£[/1]
:----r, , ,, ,,
= 0
£[Olr(O),; .. ]
£[Ulr(O),; w]
Complete market failure.
,, ,, ,,
r(lI) = ~
r(li)
,,
.
r(Q)
t
.
"'~ w~ w~ .'
We can see immediately from Figure I3.B.1 that the market equilibrium need not be ellicient. The problem is that to get the best workers to accept employment at a firm, we need the wage to be at least r(O). But in the case depicted, firms cannot break evcn at this wage because their inability to distinguish among different types of workers leaves them receiving only an expected output of £[0] < r(O) from each worker that they hire. The presence of enough low-productivity workers therefore forces thc wage down below r(ih, which in turn drives the best workers out of the market. But once the best workers are driven out of the market, the average productivity of the workforce falls, thereby further lowering the wage that firms are willing to pay. As a result, once the best workers are driven out of the market, the next-best may follow; the good may then be driven out by the mediocre. How far can this process go? Potentially very far. To see this, consider the case depicted in Figure 13.B.2, where we have r@ = and r(O) < 0 for all other O. There the equilibrium wage rate is w' = ~, and only type workers accept employment in the equilibrium. Because of adverse selection, essentially no workers are hired by firms (more precisely, a set of measure zero) even though the social optimum calls for all
q
q
to be hired,9 Example 13.8.1: To see an explicit example in which the market completely unravels let r(O) = 0:0, where 2 < 1, and let 0 be distributed uniformly on [0, 2]. Thus, r(q) = q (since (J = 0), and r(O) < 0 for 0 > 0. In this case, £[0 \ r(O) ,;; w] = (w/20:). For 0: > !, £[0 \ r(-O) ,;; 0] = 0 and £[0 \ r(O) ,;; w] < w for all w > 0, as in Figure 13.B.2.10 The competitive equilibrium defined in Definition 13.8.1 need not be unique. Figure 13.B.3, for example, depicts a case in whieh there are three equilibria with strictly positive employment levels. Multiple competitive equilibria can arise because there is virtually no restriction on the slope of the function £[0\ r(O) ,;; w). At any wage II'. this slope depends on the density of workers who are just indifferent about accepting employment and so it can vary greatly if this density varies.
9. In this equilibrium, every agent receives the same payoff as if the market were abolished: every firm earns zero and a worker of type 0 earns rIO) for all 0 (including 0 = q). 10. This example is essentially the one developed in Akeriof (1970). His example corresponds to the case): = ~.
Flgur. 13.B.2 (left)
Figure 13.B.3 (right)
Multiple eompetitiv, equilibria .
Note that the equilibria in Figure I3.B.3 can be Pareto ranked. Firms earn zero profits in any equilibrium, and workers are better off if the wage rate is higher (those workers who do not accept employment are indifferent; all other workers are strictly better off). Thus, the equilibrium with the highest wage Pareto dominates all the others. The low-wage, Pareto-dominated equilibria arise because of a coordination failure: the wage is too low because firms expect that the productivity of workers accepting employment is poor and, at the same time, only bad workers accept employment precisely because the wage is low.
A
GllI1lC- Theoret ic
Approach
The notion of competitive equilibrium that we have employed above is that used by Akerlof (1970). We might ask whether these competitive equilibria can be viewed as the outcome of a richer model in which firms could change their offered wages but choose not to in equilibrium. The situation depicted in Figure 13.B.3 might give you some concern in this regard. For example, consider the equilibrium with wage rate \\'!. In this equilibrium, a firm that experimented with small changes in its wage offer would find that a small increase in its wage, say to the level w' > \I'! depicted in the figure, would raise its profits because it would then attract workers with an average productivity of £[0\ r(O) S \\"] > 11". Hence, it seems unlikely that a model in which firms could change their offered wages would ever lead to this equilibrium outcome. Similarly, at the equilibrium involving wage II'f, a firm that understood the structure of the market would realize that it could earn a strictly positive profit by raising its offered wage to ",'. To be more formal about this idea, consider the following game-theoretic model: The underlying structure of the market [e.g., the distribution of worker productivities F(') and the reservation wage function r(')] is assumed to be common knowledge. Market behavior is captured in the following two-stage game: In stage 1, two firms simultaneously announce their wage offers (the restriction to two firms is without loss of generality). Then, in stage 2, workers decide whether to work for a firm and, if so, which one. (We suppose that if they are indifferent among some set of firms, then thcy randomize among them with equal probabilities.)" Proposition 13.B.1 characterizes the subgame perfect Nash equilibria (SPNEs) of this game for the adverse selection model in which r(') is strictly increasing with 1'(0) S () for all 0 E [Q, 0] and F(') has an associated density f(·) with f(O) > 0 for all () E [Q, 0]. Proposition 13.B.1: Let W' denote the set of competitive equilibrium wages for the adverse selection labor market model, and let w· = Max {w: WE W·}. (i) If w· > r(Q) and there is an r. > 0 such that E[O \ r(O) ,;; w'] > w' for all w' E (w* - r., w*), then there is a unique pure strategy SPNE of the two-stage game-theoretic mode\. In this SPNE, employed workers receive It. Note that if there is a single type of worker with productivity 0, this model is simply the labor market version of the Bertrand model of Section 12.C and has an equilibrium wage equal to 0, the competitive wage.
444
CHAPTER
13:
ADVERSE
SELECTION.
SIGNALING,
AND
SCREENING
a wage of w', and workers with types in the set 0(w') = (II; r(lI) :s w'} accept employment in firms. (ii) If w' = rIO), then there are multiple pure strategy SPNEs. However, in every pur~ strategy SPNE each agent's payoff exactly equals her payoff in the highest-wage competitive equilibrium. Proof: To begin, note that in any SPNE a worker of type II must follow the strategy of accepting employment only at one of the highest-wage firms, and of doing so if and only if its wage is at least r(O)." Using this fact, we can determine the equilibrium behavior of the firms. We do so for each of the two cases in turn. (i) 11" > rIO); Note, first, that in any SPNE both firms must earn exactly zero. To see this, supp;se that there is an SPNE in which a total of M workers arc hired at a wage IV and in which the aggregate profits of the two firms are
n=
M(£[Olr(O):s IV] - IV) > O.
Note that n > 0 implies that M > 0, which in turn implies that IV ~ r@. In this case, the (weakly) less-profitable firm, say firm j, must be earning no more than n/2. But firm j can earn profits of at least M(£[O I r(O) :s IV + IX] - IV - IX) by instead offering wage Ii' + ex for IX> O. Sinoe £[Olr(O):s 11'] is continuous in 11', these profits can be made arbitrarily close to by choosing IX small enough. Thus, firm j would be better off deviating, which yields a contradiction; we must therefore have :s O. Because neither firm can have strictly negative profits in an SPNE (a firm can always offer a wage of zero), we conclude that both firms must be earning exactly zero in any SPNE. From this fact, we know that if IV is the highest wage rate offered by either of the two firms in an SPNE, then either IV E W' (i.e., it must be a competitive equilibrium wage rate) or Ii' < r(O) (it must be so low that no workers accept employment). But suppose that IV < w,-= Max {II'; WE W'}. Then either firm can earn strictly positive expected profits by deviating and offering any wage rate 11" E (II" - e, 11"). We conclude that the highest wage rate offered must equal IV' in any SPNE. Finally, we argue that both firms naming 11" as their wage, plus the strategies for workers described above, constitute an SPNE. With these strategies, both firms earn zero. Neither firm can earn a positive profit by unilaterally lowering its wage because it gets no workers if it does so. To complete the argument, we show that £[0 I rIO) :s IV] < IV at every IV > 11", so that no unilateral deviation to a higher wage can yield a firm positive profits either. By hypothesis, IV" is the highest competitive wage. Hence, there is no IV> w' at which £[0 I rIO) :s 11'] = IV. Therefore, because £[0 I rIO) :s 11'] is continuous in IV, £[0 I riO) :s 11'] - IV must have the same sign for all 1\' > IV'. But we cannot have £[0 I rIO) :s IV] > II' for all II' > 11'" because, as W .... 00, £[(11 rIO) :s 11'] .... £[0], which, under our assumptions, is finite. We must therefore have £[0 I rIO) :s 11'] < II' at all IV> 11". This completes the argument for case (i). The assumption that there exists an e > 0 such that £[111 r(lI) :s 11"] > 11" for all IV' E (IV' - t, IV') rules out pathological cases such as that depicted in Figure 13.BA.
n
n
(ii) IV' = rIO); In this case, £[Olr(O):s IV] < '" for all IV> IV', so that any firm attracting workers at a wage in excess of "," incurs losses. Moreover, a firm must 12. Recall that we assume that a worker accepts employment whenever she is indilTerent.
---
SECTION
13.8:
INFORMATIONAL
ASYMMETRIES
AND
ADVERSE
SELECTION
445
,-----------------------------------------------------------, i
E[O)
~!
i, E[Olr(O) ~ .. ) ,,i!~ ,, ,,, ,, ,,
r@
r(ii)
earn exaclly zero by announcing any IV S; 11'". Hence, the set of wage offers (w" "'2) that can arise in an SPNE is {(w" w2 ): Wj:S 11'" for j = I, 2}. In everyone of these SPNEs, all agents earn exactly what they earn at the competitive equilibrium involving wage rate IV"; both firms earn zero, and a worker of type 0 earns rIO) for all liE [q, /1]. • One difference between this game-theoretical model and the notion of competitive equilibrium specified in Definition I3.B.1 involves the level of firms' sophistication. In the competitive equilibria of Definition I3.B.I, firms can be fairly unsophisticated. They need know only the average productivity level of the workers who acoept employment at the going equilibrium wage; they need not have any idea of the underlying market mechanism. In contrast, in the game-theoretic model, firms understand the entire structure of the market, including the full relationship that exists bet ween the wage rate and the quality of employed workers. The game-theoretic model tells us that if sophisticated firms have the ability to make wage offers, then we break the coordination problem described above. If the wage is too low, some firm will find it in its interest to offer a higher wage and attract better workers; the highest-wage competitive outcome must then arise.')
COllstrained Pareto Optima alld Market intervention We have seen that the presence of asymmetric information often results in market equilibria that fail to be Pareto optimal. As a consequence, a central authority who knows all agents' private information (e.g., worker types in the models above), and can engage in lump-sum transfers among agents in the economy, can achieve a Pareto improvement over these outcomes. In practice, however, a central authority may be no more able to observe agents' private information than are market participants. Without this information, the authority will face additional constraints in trying to achieve a Pareto improvement. For example, arranging lump-sum transfers among workers of different types will be impossible because the authority cannot observe workers' types directly. For Pareto-improving market intervention to be possible in this case, a more stringent test must therefore be passed. An allocation that cannot be Pareto improved by an J 3. See Exercise 13.8.6, however, for an example of a model of adverse selection in which, for some parameter values, the highest-wage competitive equilibrium is not an SPNE of our gametheoretic model.
Figure 13.8.4
A pathologicat exampte.
446
CHAPTER
'3:
ADVERSE
SELECTION.
SIGNALING.
AND
SCREENING
5 E C T ION
, 3 • B:
I N FOR MAT ION A L A . V M MET A IE.
AND
A D V E A S ESE l E C T ION
447
---------------------------------------------------------------------- ------------------------------------------------------------------authority who is unable to observe agents' private information is known as a constrained (or second-best) Pareto optimum. Because it is more difficult to generate a Pareto improvement in the absence of an ability to observe agents' types, a constrained Pareto optimal allocation need not be (fully) Pareto optimal [however, a (full) Pareto optimum is necessarily a constrained Pareto optimum]. Here, as an example, we shall study whether Pareto-improving market intervention is possible in the context of our adverse selection model (where r(') is strictly increasing with r(O) S 0 for all 0 E [Q, Ii] and F(') has an associated density f(·) with f(O) > 0 for all 0 E [~, 0]) when the central authority cannot observe worker types. That is, we study whether the competitive equilibria of this adverse selection model are constrained Pareto optima. In general, the formal analysis of this problem uses tools that we develop in Section 14.C in our study of principal-agent models with hidden information (see, in particular, the discussion of monopolistic screening). As these techniques have yet to be introduced, we shall not analyze this problem fully here. (Once you have studied Section 14.C, however, refer back to the discussion in small type at the end of this section.) Nevertheless, we can convey much of the analysis here. By way of motivation, note first that in examining whether a Pareto improvement relative to a market equilibrium is possible, we might as well simply think of intervention schemes in which the authority runs the firms herself and tries to achieve a Pareto improvement for the workers (the firms' owners will then earn exactly what they were earning in the equilibrium, namely zero profits). Second, because the authority cannot distinguish directly among different types of workers, any differences in lump-sum transfers to or from a worker can depend only on whether the worker is employed (the workers otherwise appear identical). Thus, intuitively, there should be no loss of generality in restricting attention to interventions in which the authority runs the firms herself, offers a wage of w. to those accepting employment, an unemployment benefit of w. to those who do not [these workers also receive r(O)]. leaves the workers free to choose whether to accept employment in a firm, and balances her budget. (In the small-type discussion at the end of this section, we show formally that this is the case.) Given this background, can the competitive equilibria of our adverse selection model be Pareto-improved upon in this way? Consider, first, dominated competitive equilibria, that is, competitive equilibria that are Pareto dominated by some other competitive equilibrium (e.g., the equilibrium with wage rate wf shown in Figure 13.B.3). A central authority who is unable to observe worker types can always implement the best (highest-wage) competitive equilibrium outcome. She need only set w, = w', the highest competitive equilibrium wage, and w. = O. All workers in set 0( w') then accept employment in a firm and, since w' = E[O 1r(/I) S w'], the authority exactly balances her budget." Thus, the outcome in such an equilibrium is not a constrained Pareto optimum. In this case, the planner is essentially able to step in and solve the coordination failure that is keeping the market at the low-wage equilibrium. 14. An equivalent but less heavy·handed intervention would have the authority simply require any operating firm to pay a wage rate equal to w*. Firms will be willing to remain
operational because they break even at this wage rate, and a Pareto improvement results.
What about the highest-wage competitive eqUilibrium (i.e., the SPNE outcome in the game-theoretic model of Proposition 13.B.1)? As Proposition 13.B.2 shows, any such equilibrium is a constrained Pareto optimum in this model. proposition 13.B.2: In the adverse selection labor market model (where r(') is strictly increasing with rIO) S /I for all /I E [0, Ii] and F(') has an associated density (.) with (0) > 0 for all 0 E [Q.O]). the- highest-wage competitive equilibrium is a constrained Pareto optimum. Proof: If all workers are employed in the highest wage competitive equilibrium then the outcome is fully (and, hence, constrained) Pareto optimal. So suppose some are not employed. Note, first, that for any wage w. and unemployment benefit w. offered by the central authority the set of worker types accepting employment has the form [Q, 0] for some 0 [it is {O: .... + r(O) S IV.}]. Suppose, then,that the authority attempts to implement an outcome in which worker types 0 s ~ for bE [Q, Ii] accept employment. To do so, she must choose .... and w. so that
"'. + r(O) = w•. In addition, to balance her budget, IV, and IV. must also satisfy" IV,F(O) + IV.(I - F(O» =
f
Of(O) dO.
(13.B.7)
(I3.B.8)
Substituting into (13.B.7) from (13.B.8), we find that. given the choice of 0, the values of w .. and WI' must be ",.(0) =
and
"',(0) =
r r
Of(O) dO - r(O)F(O)
(l3.B.9)
Of(O) dO + r(O)(1 - F(O)),
(13.B.10)
or, equivalently, IV.(O) = F(O)(E[OIO
s
/I] -
r(9»
w,(O) = F(O)(E[O lOs 9] - riO»~
(13.B.11)
+ r(O).
(l3.B.12) Now, let 0* denote the highest worker type who accepts employment in the highest-wage competitive equilibrium. We know that r(O') = E[/il 0 SO']. Hence, from conditions (13.B.II) and (I3.B.12), we see that w.(O') = 0 and w.(O') = r(O'). Thus, the outcome when the authority sets 0 = 0' is exactly the same as in the highest-wage competitive equilibrium. We now examine whether a Pareto improvement can be achieved by setting 6 ¥ 0*. Note that for any 0 E [Q, 0] with 6 ¥ 0', type Q workers are worse off than in the equilibrium if .... (0) < r(O') [r(O') is their wage in the equilibrium] and type workers are worse off if "',(Ii) < O. Consider 0 < 0* first. Since r(O*) > r(O), condition (l3.B.lO) implies that
o
w.(O)
s
r
Of(O) dO
+ r(O')(1
- F({J)),
15. The authority will never wish to run a budget surplus. If "". and w .. lead to a budget surplus, then setting w = "' .. +,; and wf' = WI' + (; for some I: > 0 is budget feasible and is Pareto superior. (Note that the set of workers accepting employment would be unchanged.) lI
448
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
SCREENING
--------------------------------------------------------------------------and so
11',(0) - ,(0*) :s; F(O)(E[OI 0 :s; 0] - ,(0*») = F(O)(E[OIO:s;
0] -
E[OIO:s; 0*])
< O. Thus, type 0 workers must be made worse off by any such intervention. Now cO;lsider 0 > 0*. We know that E[O I,(0) :s; 11'] < II' for all II' > 11'* (see the proof of Proposition 13.8.1). Thus, since ,(0*) = 11'* and ,(-) is strictly increasing, we have £[0 I ,(0) :s; ,(0)] < ,(ti) for all Ii> 0*. Moreover,
E[OI'(O):s; ,(0)]
= E[OIO:s; 0],
and so £[0 I 0 :s; 0] - ,(0) < 0 for all 0> 0*. But condition (l3.B.II) then implies that 11',(6) < 0 for all 0 > 0*, and so type ii workers are made worse off by any such intervention. _ Hence, when a central authority cannot observe worker types, her options may be severely limited. Indeed, in the adverse selection model just considered, the authority is unable to create a Pareto improvement as long as the highestwage competitive equilibrium (the SPNE outcome of the game-theoretic model of Proposition 13.B.I) is the market outcome. " More generally, whether Paretoimproving market intervention is possible in situations of asymmetric information depends on the specifics of the market under study (and as we have already seen, possibly on which equilibria result). Exercises 13.B.8 and 13.B.9 provide two examples of models in which the highest-wage competitive equilibrium may fail to be a constrained Pareto optimum. Although it is impossible to Pareto improve a constrained Pareto optimal allocation, market inlcrvcntion could still be justified in the pursuit of distributional aims. For example, if social welfare is given by the sum of weighted worker utilities
r
[1(0)0
SECTION
13.8:
INFORMATIONAL
ASYMMETRIES
- I(O»,(O)]i.(O) dF(O),
(13.B.13)
where i.(0) > 0 for all 0, then social welfare may be increased even though some worker types end up worse off. In the applied literature, for example, it is common to see aggregate surplus used as the social welfare function, which is equivalent to the choice of i.(O) = N for all 0." When society has this social welfare function, social welfare can be raised relative to the competitive equilibrium in Figure 13.B.1 (which, by Proposition 13.B.2, is a constrained Pareto optimum) simply by mandating that all workers must work for a firm and that all firms must
16. Proposition 13.8.2 Can also be readily generalized to allow r(O) > 0 for some O. (See Exercise I3.B.IO.) 17. Note that when types cannot be observed. aggregate surplus is no longer a valid weHare measure for any social welfare function because, unlike the case of perfect information, lump~sum
transfers across worker types are infeasible. (See Section 1O.E for a discussion of the need for lump-sum transfers to justify aggregate surplus as a welfare measure for any social welfare function,)
ADVERSE
pay workers a wage of £(0). Although workers of type ii are made worse off by this intervention, welfare as measured by aggregate surplus increases.'· An interesting interpretation of the choice of aggregate surplus as a social welfare function is in tcrms of an unborn worker's ex ante expected utility. In particular, imagine that each worker originally has a probability /(0) of ending up a type 0 worker. If this unborn worker is risk neutral, then her ex ante expected utility is exactly equal to expression (13.B.13) with i.(O) = I for all O. Thus, maximization of aggregate surplus is equivalent to maximization of this unborn worker's expected utility. We might then say that an allocation is an ex ante COII.wrailJeti Pareto optimum in this model ir, in the absence of an ability to observe worker types, it is impossible to devise a market intervention that raises aggregate surplus. We see, therefore. thai whether an allocation is a constrained optimum (and. thus, whether a planned intervention leads to a Pareto improvement) can depend on the point at which the welrare evaluation is conducted (i.e., before the workers know their types, or after)'· Lei us now use the techniques of Section 14.C 10 show formally that we can restrict attention in searching for a Pareto improvement to interventions of the type considered above. We shall look for a Pareto improvement for the workers keeping the profits of the firms' owners nonnegative. For notational simplicity, we shall treat the firms as a single aggregate firm. By the revelation principle (see Section 14.C), we know that we can restrict attention to direct rcvcl;.ttion mechanisms in which every worker type tells the truth, Here a direct revelation mechanism assigns, for each worker type () E [q, 0], a payment from the authority to the worker of ",(Ii) E R, a lax t(O) paid by the firm to the authority, and an employment decision /(Ii) E 10, I:. The sci of feasible mechanisms here are those that satisfy the illdividual ratiollality c0I1s1raillf for the firm.
r
[/(0)0 - t(O)] dF(O)
(13.B.14)
the hwly£'[ hulcmce i'olldilion for the central authority,
f
[t(O) - ...(0)] dF(O)
(13.B.15)
and the lrur/Hellillf} (or ;'Icentivt! compatibilil)', or seltse/(-ftiml) constraints that say that for
all 0 and
iJ 1\'(0) + (I - 1(0»,.(0)
+ (1
AND
SELECTION
449
~----------------------------------------------------------------------
(13.8.16)
Note, first, that mechanism [w('), /(.), 1(')] is feasible only if [IV('), 1(')] satisfies both condition (13.B.16) and
r
[/(0)0 - \1'(0)] dF(O) ;;, 0.
(I3.B.l7)
I R. Moreover. bec
450
c HAP T E R
'3:
A D V E R S ESE L E C T ION,
S I G N A LIN G ,
AND
SCREEN I NG
-------------------------------------------------------------------
SECTION
'3.C:
SIGNALING
451
~---------------------------------------------------------------
Moreover, if [w{-l, 1(·)] satisfies (I3.B.16) and (13.B.17), then there exists a 1(') such that [w('), I( . ), I( .)] satisfies (13.B.14)-( 13.B.16). Condition (13.B.17), however, is exactly the budget constraint faced by a central authority who runs the firms herself. Hence, we can restrict attention to schemes in which the authority runs the firms herself and uses a direct revelation mechanism [w('), 1(')] satisfying (13.B.16) and (13.B.17). Now consider any two types 0' and O· for which 1(0') = 1(0"). Setting 0 = O' and 6 = O· in condition (13.B.16), we see that we must have w(O') ~ w(O"). Likewise, letting 0= o· and () = 0', we must have w(O") ~ w(O'). Together, this implies that w(0') = w(0"). Since 1(0) E {O,l}. we see that any feasible mechanism [w(' ),/(')] can be viewed as a scheme that gives each worker a choice between two outcomes. (w .. 1 = I) and (W., I = 0) and satisfies the budget balance condition (13.B.17). This is exactly the class of mechanisms studied above.
In the analysis that follows, we shall see that this otherwise useless education may serve as a signal of unobservable worker productivity. In particular, equilibria emerge in which high-productivity workers choose to get more education than lowproductivity workers and firms correctly take differences in education levels as a signal of ability. The welfare effects of signaling activities are generally ambiguous. By revealing information about worker types, signaling can lead to a more efficient allocation of workers' labor, and in some instances to a Pareto improvement. At the same time, because signaling activity is costly, workers' welfare may be reduced if they are compelled to engage in a high level of signaling activity to distinguish t hemscl ves. To keep things simple, throughout most of this section we concentrate on the special case in which r(O,,) = r(Od = O. Note that under this assumption the unique equilibrium that arises in the absence of the ability to signal (analyzed in Section I3.B) has all workers employed by firms at a wage of IV· = £[0] and is Pareto efficient. Hence, our study of this case emphasizes the potential inefficiencies created by signaling. After studying this case in detail, we briefly illustrate (in small type) how, with alternative assumptions about the function r('), signaling may instead generate a Pareto improvement. A portion of the game tree for this model is shown in Figure 13.C.1. Initially, a random move of nature determines whether a worker is of high or low ability. Then, conditional on her type, the worker chooses how much education to obtain. After obtaining her chosen education level, the worker enters the job market. Conditional on the observed education level of the worker, two firms simultaneously make wage ofTers to her. Finally, the worker decides whether to work for a firm and, if so, which one.
13.C Signaling Given the problems observed in Section I3.B, one might expect mechanisms to develop in the marketplace to help firms distinguish among workers. This seems plausible because both the firms and the high-ability workers have incentives to try to accomplish this objective. The mechanism that we examine in this section is that of sil//laling. which was first investigated by Spence (1973, 1974). The basic idea is that high-ability workers may have actions they can take to distinguish themselves from their low-ability counterparts. The simplest example of such a signal occurs when workers can submit to some costless test that reliably reveals their type. It is relatively straightforward to show that in any subgame perfect Nash equilibrium all workers with ability greater than o will submit to the test and the market will achieve the full information outcome (see Exercise 13.C.1). Any worker who chooses not to take the test will be correctly treated as being no better than the worst type of worker. However, in many instances, no procedure exists that directly reveals a worker's type. Nevertheless, as the analysis in this section reveals, the potential for signaling may still exist. Consider the following adaptation of the model discussed in Section I3.B. For simplicity, we restrict attention to the case of two types of workers with productivities 011 and OL' where Oil > IJ L > 0 and ;. = Prob (0 = ON) E (0,1). The important extension of our previous model is that before entering the job market a worker can get some education, and the amount of education that a worker receives is observable. To make matters particularly stark, we assume that education does nothing for a worker's productivity (see Exercise 13.C.2 for the case of productive signaling). The cost of obtaining education level e for a type 0 worker (the cost may be of either monetary or psychic origin) is given by the twice continuously differentiable function c(e, 0). with e(O, 0) = 0, c.(e, 0) > 0, c.. (e, 0) > 0, c.(e,O) < 0 for all e > 0, and c.. (e, 0) < 0 (subscripts denote partial derivatives). Thus, both the cost and the marginal cost of education are assumed to be lower for high-ability workers; for example, the work required to obtain a degree might be easier for a high-ability individual. Letting U(IV, e 10) denote the utility of a type 0 worker who chooses education level e and receives wage IV, we take U(IV, e I0) to equal her wage less any educational costs incurred: U(IV, e 10) = w - c(e, 0). As in Section 13.B, a worker of type 0 can earn r(O) by working at home.
Random mOve of
Flgure13.C.l
worker Iype.
The ex.tensive form of the education signaling
,.A,,""------------ nature determines
game. , ; : : . - - - - - - - - - - - - - - - " " : I l.....I ' - - - - Worker chooses
education level contingent on her Iype (really a continuous choice),
Conditional on seeing a level of t. say".
firms make wage offers simultaneously (really a continuous choice),
Worker decides which olTer 10 accept. if any. }
J
452
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
SECTION
SCREENING
--------------------------------------------------------------------
13.C:
SIGNALING
453
~---------------------------------------------------------------
Note that, in contrast with the model of Section 13.B, here we explicitly model only a single worker of unknown type; the model with many workers can be thought of as simply having many of these single-worker games going on simultaneously, with the fraction of high-ability workers in the market being).. In discussing the equilibria of this game, we often speak of the "high-ability workers" and "low-ability workers," having the many-workers case in mind. The equilibrium concept we employ is that of a weak perfect Bayesian equilibrium (see Definition 9.C.3), but with an added condition. Put formally, we require that, in the game tree depicted in Figure 13.C.l, the firms' beliefs have the property that, for each possible choice of e, there exists a number /lie) E [0, 1] such that: (i) firm I's belief that the worker is of type 011 after seeing her choose e is /lie) and (ii) after the worker has chosen e, firm 2's belief that the worker is of type 011 and that firm I has chosen wage offer w is precisely /l(e)ur(w I e), where ur(w I e) is firm I's equilibrium probability of choosing wage offer IV after observing education level e. This extra condition adds an element of commonality to the firms' beliefs about the type of worker who has chosen e, and requires that the firms' beliefs about each others' wage offers following e are consistent with the equilibrium strategies both on and off the equilibrium path. We refer to a weak perfect Bayesian equilibrium satisfying this extra condition on beliefs as a perfect Bayesian equilibrium (PBE). Fortunately, this PBE notion can more easily, and equivalently, be stated as follows: A set of strategies and a belief function /,(e) E [0, 1] giving the firms' common probability assessment that the worker is of high ability after observing education level e is a PBE if
011 ------------------------
Flgur. 13.C.2 (left)
Indifference curves for high- and low·ability workers: the single-crossing
-----------------------
()I.
property. Flgur. 13.C.3 (rlghl)
o Knowing this fact, we turn to the issue of the worker's equilibrium strategy, her choice of an education level contingent on her type. As a first step in this analysis, it is useful to examine the worker's preferences over (wage rate, education level) pairs. Figure IlC.2 depicts an indifference curve for each of the two types of workers (with wages measured on the vertical axis and education levels measured on the horizontal axis). Note that these indilTerence curves cross only once and that, where they do, the indifference curve of the high-ability worker has a smaller slope. This property of preferences, known as the single-crossing property, plays an important role in the analysis of signaling models and in models of asymmetric information more generally. It arises here because the worker's marginal rate of substitution between wages and education at any given (IV, e) pair is (dIVide)" = e.(e, 0), which is decreasing in 0 because (" ... (e, II) < O. We can also graph a function giving the equilibrium wage offer that results for each education level, which we denote by wee). Note that since in any PBE wee) = /,(e)1I1I + (I - /,(e»Ol for the equilibrium belief function /lie), the equilibrium wage offer resulting from any choice of e must lie in the interval [0,.,0,,]. A possible wage offer function w(e) is shown in Figure 13.C.3. We are now ready to determine the equilibrium education choices for the two types of workers. It is useful to consider separately two different types of equilibria that might arise: separating equilibria, in which the two types of workers choose different education levels, and pooling equilibria, in which the two types choose the same education level.
(i) The worker's strategy is optimal given the firm's strategies. (ii) The belief function /lie) is derived from the worker's strategy using Bayes' rule where possible. (iii) The firms' wage offers following each choice e constitute a Nash equilibrium of the simultaneous-move wage offer game in which the probability that the worker is of high ability is /l(e).20 In the context of the model studied here, this notion of a PBE is equivalent to the sequential equilibrium concept discussed in Section 9.C. We also restrict our attention throughout to pure strategy equilibria. We begin our analysis at the end of the game. Suppose that after seeing some education level e, the firms attach a probability of /,(e) that the worker is type 0Il' lf so, the expected productivity of the worker is /l(e)OI/ + (1 - /l(e»Ol' In a simultaneous-move wage offer game, the firms' (pure strategy) Nash equilibrium wage offers equal the worker's expected productivity (this game is very much like the Bertrand pricing game discussed in Section 12.C). Thus, in any (pure strategy) PBE, we must have both firms offering a wage exactly equal to the worker's expected productivity, /l(e)OIl + (I - /,(e»Ol'
Separating Equilihria To analyze separating equilibria, let e*(tI) be the worker's equilibrium education choice as a function of her type, and let 1V*(e) be the firms' equilibrium wage offer as a function of the worker's education level. We first establish two useful lemmas. Lemma 13.C.1: In any separating perfect Bayesian equilibrium, w*(e*(OH» = 0H and w*(e*(OLl) = 0L: that is, each worker type receives a wage equal to her productivity level. Proof: In any PBE, beliefs on the equilibrium path must be correctly derived from the equilibrium strategies using Bayes' rule. Here this implies that upon seeing education level e*(Ol)' rlrms must assign probability one to the worker being type 0,.. Likewise, upon seeing education level e*(OIl)' firms must assign probability one
20. Thus, the extra condition we add imposes equilibrium-like play in parts of the tree off the equilibrium path. See Section 9.C for a discussion of the need to augment the weak perfect Bayesian equilibrium concept to achieve this end.
J
A wage schedule.
454
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
--- --
SCREENING
Type 0" II'
Type Ot
qJ ,// i
,, , ,,
. I
/
0" ""(to)
.--
I
(/,.
._._._._._.--;-----------
I
I
I I
I
,,
I
e
11
11
e'(O,.)
e'(O,,)
(h
and 0",
Figure 13.C.4 (left)
Low-ability worker's outcome in a
Lemma 13.C.2: In any separating perfect Bayesian equilibrium, e'(Otl = 0; that is, a low-ability worker chooses to get no education. Proof: Suppose not, that i~, that when the worker is type OL' she chooses some strictly positive education level e ::. O. According to Lemma 13.C.I, by doing so, the worker receives a wage equal to 0L' However, she would receive a wage of at least 0,_ if she instead chose e = O. Since choosing e = 0 would have save her the cost of education, she would be strictly better off by doing so, which is a contradiction to the assumption that > 0 is her equilibrium education level. _
e
Lemma 13.C.2 implies that, in any separating equilibrium, type O,.'s indifference curve through her equilibrium level of education and wage must look as depicted in Figure 13.CA. Using Figure 13.CA, we can construct a separating equilibrium as follows: Let c'(O,,) = ii, let e'(Od = 0, and let the schedule ",'(e) be as drawn in Figure 13.C.5. The firms' equilibrium beliefs following education choice e are JI'(e) = (w'(e) - Od/(Oll - Od. Note that they satisfy I,'(e) E [0, I] for all e;::o; 0, since \\,'(e) E [OL' 0,,]. To verify that this is indeed a PBE, note that we are completely free to let firms have any beliefs when e is neither 0 nor On the other hand, we must have JI(O) = 0 and Il(e) = 1. The wage offers drawn, which have ",'(0) = 0,- and ""(e) = 0", renect exactly these beliefs. What about the worker's strategy? It is not hard to see that, given the wage function IV'(e), the worker is maximizing her utility by choosing e = 0 when she is type 0L and by choosing e = when she is type 0". This can be seen in Figure 13.C.S by noting that, for each type that she may be, the worker's indifference curve is at its highest-possible level along the schedule ""(e). Thus, strategies [e'(O), ""(e)] and the associated beliefs JI(e) of the firms do in fact constitute a PBE. Note that this is not the only PBE involving these education choices by the two types of workers. Because we have so much freedom to choose the firms' beliefs off the equilibrium path, many wage schedules can arise that support these education
e.
0,.
/~r'\" I
"'(e)
L/ I --·---r : -~:'~------i--------t----I I I I I
I I , , ,
I
o
to the worker being type 0". The resulting wages are then exactly respectively. _
/
'-.
----------.----------,
"'(e'(O,.)) = 01.
,,
e
A separating equilibrium with the
"'(e)
------------~----------
separating equilibrium.
Figure 13.C.S (right)
A separating equilibrium: Type is inferred from education level.
SIGNALING
Figure 13.C.6 (Iell)
(/"
. . . . . ·*'
I
13.C:
Type O.
'I"~---------
8" ------------ ,----------
,, ,, ,, ,
SECTION
"'(0,.)
11
"
r'(O,,)
.'«(/,,)
11
choices. Figure 13.C.6 depicts another one; in this PBE, firms believe that the worker is certain to be of high quality if c ;::0; i' and is certain to be of low quality if e < e. The resulting wage schedule has ""(e) = 0" if e ;::0; ii and ""(e) = OL if e < ii. In these separating equilibria, high-ability workers arc willing to get otherwise useless education simply because it allows them to distinguish themselves from low-ability workers and receive higher wages. The fundamental reason that education can serve as a signal here is that the marginal cost of education depends on a worker's type. Because the marginal cost of education is higher for a low-ability worker [since c,.• (e, 0) < 0], a type 0" worker may find it worthwhile to get some positive level of education e' > 0 to raise her wage by some amount "'"' > 0, whereas a type OL worker may be unwilling to get this same level of education in return for the same wage increase. As a result, firms can reasonably come to regard education level as a signal of worker quality. The education level for the high-ability type observed above is not the only one that can arise in a separating equilibrium in this model. Indeed, many education levels for the high-ability type arc possible. In particular, any education level between i! and el in Figure 13.C.7 can be the equilibrium education level of the high-ability workers. A wage schedule that supports education level e'(O,,) = e, is depicted in the figure. Note that the education level of the high-ability worker cannot be below in a separating equilibrium because, if it were, the low-ability worker would deviate and pretend to be of high ability by choosing the high-ability education level. On the other hand, the education level of the high-ability worker cannot be above €, because, ifit were, the high-ability worker would prefer to get no education, even if this resulted in her being thought to be of low ability. Note that these various separating equilibria can be Pareto ranked. In all of them, firms cam zero profits. and a low-ability worker's utility is 0,.. However, a high-ability worker does strictly better in equilibria in which she gets a lower level of education. Thus, separating equilibria in which the high-ability worker gets education level e (e.g., the equilibria depicted in Figures 13.C.S and 13.C.6) Pareto dominate all the others. The Pareto-dominated equilibria are sustained because of the high-ability worker's fear that if she chooses a lower level of education than that prescribed in the equilibrium firms will believe that she is not a 'high-ability worker. These beliefs can be maintained because in equilibrium they are never disconfirmed.
e
same education choices as in Figure 13.C.S but different
off-equilibriumpath beliefs. Figure 13.C.7 (right)
A separating equilibrium with an
education choice
> e by high-ability workers.
e"(OH)
455
456
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
SCREENING
.
w
Figure 13.C.8
~TypeO.
E[O) --- --------------------
0, ------------------------
(a)
SEC T ION
1 3 • C:
S I G N A LIN G
457
----- ,------------------------------------------------------------------------
0, -------------------------
(b)
\I'
Separating equilibria may be Pareto dominated by the no·signaling OUlcom (a) A. separating e equlhbnum that is nOI Pareto domInated by the no,s'gnaling
w
Type 0.
0" ------ ------~~~~:-~~
£[11]
,,--.~
I I I
/:
I
V~
outcome.
-----~--------------- I
I
(b) A separating equilibrium that is Pareto dominated by the nO'signaling
0,
.-.,,/
.. '(e)
\
------r---------------
I
I I
t"
I I I
e'
"
outcome.
"~(Oil; j
Figure 13.C,10 (right)
= L, H
The only remaining issue therefore concerns what levels of education can arise in a pooling equilibrium, It turns out that any education level between 0 and the level e' depicted in Figure 13.C.9 can be sustained. Figure 13.C.1O shows an equilibrium supporting education level e', Given the wage schedule depicted, each type of worker maximizes her payoff by choosing education level e', This wage schedule is consistent with Bayesian updating on the equilibrium path because it gives a wage offer of £[0] when education level e' is observed, Education levels between 0 and e' can be supported in a similar manner. Education levels greater than e' cannot be sustained because a low-ability worker would rather set e = 0 than e > e' even if this results in a wage payment of OL' Note that a pooling equilibrium in which both types of worker get no education Pareto dominates any pooling equilibrium with a positive education leveL Once again, the Pareto-dominated pooling equilibria are sustained by the worker's fear that a deviation will lead firms to have an unfavorable impression of her ability. Note also that a pooling equilibrium in which both types of worker obtain no education results in exactly the same outcome as that which arises in the absence of an ability to signal. Thus, pooling equilibria are (weakly) Pareto dominated by the no-signaling outcome,
It is of interest to compare welfare in these equilibria with that arising when worker types arc unobservable but no opportunity for signaling is available. When education is not available as a signal (so workers also incur no education costs), we arc back in the situation studied in Section I3.B. In both cases, firms earn expected profits of zero. However, low-ability workers are striclly worse off when signaling is possible. I n both cases they incur no education costs, but when signaling is possible they receive a wage of OL rather than £(0). What about high-ability workers~ The somewhat surprising answer is that high-ability workers may be either better or worse off when signaling is possible. In Figure 13.C.8(a), the high-ability workers are better off because of the increase in their wages arising through signaling. However, in Figure I3.C.8(b), even though high-ability workers seek to take advantage of the signaling mechanism to distinguish themselves, they are worse off than when signaling is impossible! Although this may seem paradoxical (if high-ability workers choose to signal, how can they be worse olP), its cause lies in the fact that in a separating signaling equilibrium firms' expectations are such that the wage-education outcome from the no-signaling situation, (w, e) = (£[0],0), is no longer available to the high-ability workers; if they get no education in the separating signaling equilibrium, they are thought to be of low ability and offered a wage of 0L' Thus, they can be worse off when signaling is possible, even though they are choosing to signal. Note that because the set of separating equilibria is completely unaffected by the fraction i. of high-ability workers, as this fraction grows it becomes more likely that the high-ability workers are made worse off by the possibility of signaling [compare Figures 13.C.8(a) and 13.C.8(b)]. In fact, as this fraction gets close to I, nearly every worker is getting costly education just to avoid being thought to be one of the handful of bad workers!
Multiple Equilibria alld Equilibrium Refinemelll The multiplicity of equilibria observed here is somewhat disconcerting, As we have seen, we can have separating equilibria in which firms learn the worker's type, but we can also have pooling equilibria where they do not; and within each type of equilibrium, many different equilibrium levels of education can arise, In large part, this multiplicity stems from the great freedom that we have to choose beliefs off the equilibrium path, Recently, a great deal of research has investigated the implications of pulling "reasonable" restrictions on such beliefs along the lines we discussed in Section 9.D, To see a simple example of this kind of reasoning, consider the separating equilibrium depicted in Figure 13,C.7, To sustain e l as the equilibrium education level of high-ability workers, firms must believe that any worker with an education level below e I has a positive probability of being of type OL' But consider any education level eE (e, el)' A type OL worker could never be made better off choosing such an education level than she is getting education level e = 0 regardless of what
Poolillg Equilibria Consider now pooling equilibria, in which the two types of workers choose the same level of education, e'(OL) = e'(O/l) = eO. Since the firms' beliefs must be correctly derived from the equilibrium strategies and Bayes' rule when possible, their beliefs when they see education level e' must assign probability). to the worker being type 0/1' Thus, in any pooling equilibrium, we must have w'(e') = ).0/1 + (I - i.)OL = £[0]. I
J
(Ienl The highest-possible education level in a pooling equilibrium.
Figure 13.C.9
A pooling equilibrium.
458
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
SCREENING
firms believe about her as a result. Hence, any belief by firms upon seeing education level e > e other than !l(e) = I seems unreasonable, But if this is so, then we must have w(e) = 9", and so the high-ability worker would deviate to e. In fact, by this logic, the only education level that can be chosen by type 0" workers in a separating
---
SECTtON
"'II'
wage of If so, low-ability workers would choose e = 0 and high-ability workers would choose e = ell' This alternative outcome involves firms incurring losses on low-ability workers and making profits on high-ability workers. However, as long as the firms break even on average, they are no worse off than before and a Pareto improvement has been achieved. The key to this Pareto improvement is that the central authority introduces cross-subsidization, where high-ability workers are paid less than their productivity level while low-ability workers are paid more than theirs, an outcome that cannot occur in a separating signaling equilibrium. (Note that the outcome when signaling is banned is an extreme case of cross-subsidization.)
equilibrium involving reasonable beliefs is e. In Appendix A we discuss in greater detail the use of these types of reasonablebeliefs refinements. One refinement proposed by Cho and Kreps (1987), known as the illluirive criterion, extends the idea discussed in the previous paragraph to rule out not only the dominated separating equilibria but also all pooling equilibria. Thus, If we accept the Cho and Kreps (1987) argument, we predict a unique outcome to this two-type signaling game: the best separating equilibrium outcome, which is shown in Figures 13.C.S and 13.C.6.
Exercise l3,C.3: In the signaling model discussed in Section I3.C with r(O,,) = r(O,,) = 0, construct an example in which a central authority who does not observe worker types can achieve a Pareto improvement over the best separating equilibrium through a policy that involves cross-subsidization, but cannot achieve a Pareto improvement by simply banning the signaling activity. [Hint: Consider first a case with linear inditTerence curves.J
Secolld-Best Market Intervention I n contrast with the market outcome predicted by the game-theoretic model studied in Section I3.B (the highest-wage competitive equilibrium), in the presence of signaling a central authority who cannot observe worker types may be able to achieve a Pareto improvement relative to the market outcome. To see this in the simplest manner, suppose that the Cho and Kreps (1987) argument predicting the best separating equilibrium outcome is correct. We have already seen that the best separating equilibrium can be Pareto dominated by the outcome that arises when signaling is impossible. When it is, a Pareto improvement can be achieved simply by banning the signaling activity, In fact, it may be possible to achieve a Pareto improvement even when the no-signaling outcome does not Pareto dominate the best separating equilibrium. To see how, consider Figure l3.C.1 I. In the figure, the best separating equilibrium has low-ability workers at point (OL' 0) and high-ability workers at point (01/, e). Note that the high-ability workers would be worse off if signaling were banned, since the point (£[OJ, O) gives them less than their equilibrium level of utility. Nevertheless, note that if we gave the low- and high-ability workers outcomes of (IVL,O) and ("'", ell), respectively, both types would be better off. The central authority can achieve this outcome by mandating that workers with education levels below ell receive a wage of"'L and that workers with education levels of at least ell receive a
,, ,I
13.C:
The case with r(OIl) = r(O"l = 0 studied above, in which the market outcome in the absence of signaling is Pareto optimal. illustrates how the use of costly signaling can reduce welfare. Yet, when the market outcome in the absence of signaling is not efficient, signaling's ability to reveal information about worker types may instead create a Pareto improvement by leading to a more efficient allocation of labor. To see this point, suppose that we have r = ,«(lL) = '(011)' with 0L < r < 0" and £[0] <,. In this case, the equilibrium outcome without signaling has no workers employed. In contrast, any Pareto efficient outcome must have the high-ability workers employed by firms. We now study the equilibrium outcome when signaling is possible. Consider, first, the wage and employment outcome that results after educational choice • by the worker. Following the worker's choice of educational level " equilibrium behavior involves a wage of w'(e) = 1,(e)O" + (I -11(ellOL.1f ""(e);;' r, then both types of workers would accept employment; if w'(e) < r, then neither type would do so. We now determine the equilibrium education choices of the two types of workers. Note first that any pooling equilibrium must have bOlh types choosing' = 0 and neither type accepting employment. To sec this, suppose that both types are choosing education level e. Then II(e) = i. and ",'(.) = £[0] < r, and so neither type accepts employment. Hence, if e > 0, both types would be better ofT choosing e = 0 instead. Thus, only an education level of zero is possible in a pooling equilibrium. In this zero education pooling equilibrium, the outcome is identical to the equilibrium outcome arising in the absence of the opportunity to signal. The set of separating equilibria. on the other hand, is illustrated in Figure 13.C.12. In any separating equilibrium, a low·ability worker sets e = 0, is ofTered a wage of 0,., and chooses to work at home, thereby achieving a utility of r. High-ability workers, on the other hand, select an education level in the interval [e. e,] depicted in the figure, are ofTered a wage of 0,,_ and accept employment. Note thai no separating equilibrium can have e*(OIl) < t. since then
Figure 13.C.11 Achieving a Parelo improvement through cross-subsidization.
low·ability workers would deviate and set e e'(OIl)
I I ----------t-------------,I
= ('*(0,,); also, no separating equilibrium can have
> "" since high-ability workers would then be better off setting. = 0 and working at
home.
Note that in all these equilibria. both pooling and separating, the high·ability workers arc weakly better ofT compared with the equilibrium arising without signaling opportunities and are strictly better ofT in separating equilibria with ,'(0,,) < ',. Moreover, both the low-ability workers and the firms are equally well ofT. Thus, in the case with OL < r < 0" and £[0] < r,
,, ,,
any pooling or separating signaling equilibrium weakly Pareto dominates the outcome arising
j
SIGNALING
459
,-------------------------------------------------------------
I
460
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
---
SCREENING
SECTION
We assume that the utility of a type 0 worker who receives wage wand faces task level I ;:: 0 is u(w, 110) = w - C{I, 0),
Type OH
-
where e(I, 0) has all the properties assumed of the function c(e, 0) in Section 13.e. In particular, c(O, 0) = 0, c,(I, 0) > 0, c,,(I, 0) > 0, C,{I, 0) < 0 for aliI> 0, and c,,(I, 0) < O. As will be clear shortly, the task level I serves to distinguish among types here in a manner that parallels the role of education in the signaling model discussed in Section 13.e. Here we study the pure strategy subgame perfect Nash equilibria (SPNEs) of the following two-stage game:23
I I I I
I
I I I I
I I
UL -----~----------t-----I I I
I
I
I
:
I
f
t'2
Ftgure 13.C.12 Separating equilibria when r(Od = r(OH) ~ re(OL'OU)'
'----y--'
('(/lL)
Possible Values of ,'(OH)
Slage I:
Slage 2: in the absence of signaling, and this ParelO dominance is equilibria.
Slr;CI
for (essentially) all separating
13,0 Screening In Section 13.C, we considered how signaling may develop in the marketplace as a response to the problem of asymmetric information about a good to be traded. There, individuals on the more illformed side of the market (workers) chose their level of education in an attempt to signal information about their abilities to uninformed parties (the firms). In this section, we consider an alternative market response to the problem of unobservable worker productivity in which the uninformed parties take steps to try to distinguish, or screell, the various types of individuals on the other side of the market,2' This possibility was first studied by Rothschild and Stiglitz (1976) and Wilson (1977) in the context of insurance markets (see Exercise 13.0.2). As in Section 13.C, we focus on the case in which there are two types of workers, OL and 0", with 0" > OL > 0 and where the fraction of workers who are of type 0" is ;, E (0, I). In addition, workers earn nothing if they do not accept employment in a firm [in the notation used in Section 13.B, r(Od = reO,,) = 0). However, we now suppose that jobs may differ in the "task level" required of the worker. For example, jobs could differ in the number of hours per week that the worker is required to work. Or the task level might represent the speed at which a production line is run in a factory. To make matters particularly simple, and to make the model parallel that in Section IlC, we suppose that higher task levels add 1I0lilillg to the output of the worker; rather, their ollly effect is to lower the utility of the worker.n The output of a type 0 worker is therefore 0 regardless of the worker's task level.
13.0,
Two firms simultaneously announce sets of offered contracts. A contract is a pair (w, I). Each firm may announce any finite number of contracts. Given the offers made by the firms, workers of each type choose whether to accept a contract and, if so, which one. For simplicity, we assume that if a worker is indifferent between two contracts, she always chooses the one with the lower task level and that she accepts employment if she is indifferent about doing so. If a worker's most preferred contract is offered by both firms, she accepts each firm's offer with probability 1.
Thus, a firm can offer a variety of contracts; for example, it might have several production lines, each running at a different speed. Different types of workers may then end up choosing different contracts. 2' It is helpful to start by considering what the outcome of this game would be if worker types were observable. To address this case, we allow firms to condition their offer on a worker's type (so that a firm can offer a contract (WL' ILl solely to type OL workers and another contract (WII' I,,) solely to type 011 workers). Proposition 13.0,1: In any SPNE of the screening game with observable worker types, a type 0, worker accepts contract (wi, til = (6" 0). and firms earn zero profits. Proof: We first argue that any contract (w1, til that workers of type 0i accept in equilibrium must produce exactly zero profits; that is, it must involve a wage w~ = 0i' To see this, note that if wi > Oi' then some firm is making a loss offering this contract and it would do better by not offering any contract to type 0i workers. Suppose, on the other hand, that wi < 0i. and let n > 0 be the aggregate profits earned by the two firms on type 0i workers. One of the two firms must be earning no more than n/2 from these workers. If it deviates by offering a contract (IV~ + c, til for any 23. For this game. the set or subgame perfect Nash equilibria is identical to the sets or strategy profiles in weak perfect Bayesian equilibria or sequential equilibria. 24. The models in the original Rothschild and Stiglitz (1976) and Wilson (1977) analyses differ from our model in two respects. First. firms in those papers were restricted to offering only a single contract. This could make sense in the production line interpretation, ror example, if each firm had only a single production line. Second, those authors allowed for "free entry," so that an additional firm could always enter ir a profitable contracting opportunity existed. In ract, making these two changes has little effect on our conclusions. The only difference is in the precise conditions under which an equilibrium exists. (For more on this, see Exercise 13.D.4.)
21. The setting analyzed here is one of comperilire sCfec'lillg of workers, since we assume that there ;He several competing firms. See Section t4.C ror a discussion of the monopolistic screfmillg case, where a single firm screens workers. 22. As was true in the case or educational signaling. the assumption that higher task levels do not raise productivity is made purely for expositional purposes. Exercise 13.D.I considers the case in which the firms' profits are increasing in the task level.
J
SCREENING
461
462
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
SECTION
SCREENING
--------------------------------------------------------------------High-Ability Break-Even Line 0" __________ L ___________ _ Pooled Break·Even Line E[O) _________ L ____________ _ Low·Ability Break·Even Line ./
0,. ----------"---------------
> 0, it will attract all type (I, workers. Since I: can be made arbitrarily small, its profits from type 0i workers can be made arbitrarily close to n, and so this deviation will increase its profits. Thus, we must have w7 = 0,. Now suppose that (wt, In = (Oi' I') for some I' > O. Then, as shown in Figure 13.0.1 (where the wage is measured on the vertical axis and the task level is measured on the horizontal axis). either firm could deviate and earn strictly positive profits by oITering a contract in the shaded area of the figure, such as (w, i). The only contract at which there arc no profitable deviations is (w7. = (Oi' 0), the contract that maximizes a type (I, worker's utility subject to the constraint that the firms oITering the contract break even. _
In
We now turn to the situation in which worker types are nor observable. In this case, each contract oITered by a firm may in principle be accepted by either type of worker. We can note immediately that the complete information outcome identified in Proposition 13.0.1 cannot arise when worker types are unobservable: Because every low-ability worker prefers the high-ability contract (0 11 ,0) to contract (01.,0), if these were the two contracts being offered by the firms then all workers would accept contract (0 11 ,0) and the firms would end up losing money. To determine the equilibrium outcome with unobservable worker types, it is useful to begin by drawing three break-even lines: the zero-profit lines for productivity levels (/,., £[0], and 011, respectively. These three break-even lines arc depicted by the dashed lines in Figure 13.0.2. The middle line represents the break-even line for a contract that attracts both types of workers, and we therefore refer to it as the pooled break-even line. As in Section I3.C, we can in principle have two types of (pure strategy) equilibria: separarillg equilibria, in which the two types of workers accept different contracts, and poolillg equilibria, in which both types of workers sign the same contract. (It can be shown that in any equilibrium both types of workers will accept some contract; we assume that this is so in the discussion that follows.) We proceed with a series of lemmas. Lemma 13.0.1 applies to both pooling and separating equilibria.
Figure 13.0.1 (leH)
The equilibrium contract (wi!
463
An important implication of Lemma IlD.1 is that, in any equilibrium, no firm can have a deviation that allows it to earn strictly positive profits. We shall use this fact repeatedly in the discussion that follows. Using it, we immediately get the result given in Lemma 13.0.2 regarding pooling equilibria.
o I:
SCREENING
(IV" + e, I,,) for e > O. Contract (wI. + e, I,J will attract all type 01. workers. and contract (IV II + e, 11/) will attract all type 01/ workers. [Note that since type 0, initially prefers contract (w" I,) to (WI' tl)' we have w, - C(I" 0,) ~ WI - C(II' 0,), and so (Wi + e) - C(ti' Oil ~ (wj + e) - c(tjo 0,).] Since e can be chosen to be arbitrarily small. this deviation yields this firm profits arbitrarily close to n, and so the firm has a profitable deviation. Thus. we must have n :5 O. Because no firm can incur a loss in any equilibrium (it could always earn zero by oITering no contracts), both firms must in fact earn a profit of zero. _
w
Type (I, :..,....-- Indifference Curve
13,0:
~
til
lemma 13,0.2: No pooling equilibria exist.
for type 0, with perfttt observability.
Proof: Suppose that there is a pooling equilibrium contract (WP.I P). By Lemma 13.0.1, it lies on the pooled break-even line, as shown in Figure 13.0.3. Suppose that firm j is oITering contract (w p • I P ). Then firm k ¥ j has a deviation that yields it a strictly positive profit: It ofTers a single contract (1\', i) that lies somewhere in the shaded region in Figure IlD.3 and has Ii· < 0". This contract attracts all the type 0" workers and none of the type {lL workers, who prefer (w p• I P) over (IV, i). Moreover, since II- < 0", firm k makes strictly positive profits from this contract when the high-ability workers accept it. _
Flgur. 13.0.2 (right)
Break-even lines.
We now consider the possibilities for separating equilibria. Lemma 13.0.3 shows that all contracts accepted in a separating equilibrium must yield zero profits. Lemma 13.0.3: If (WL' ttl and (WH' tH ) are the contracts signed by the low- and high-ability workers in a separating equilibrium. then both contracts yield zero profits; that is, w L = OL and w H = 0H' Proof: Suppose first that 11'1. < 0,.. Then either firm could earn strictly positive profits by instead oITering only contract (w,., rd, where 01. > wL > IVI.' All low-ability workers would accept this contract: moreover, the deviating firm earns strictly positive profits from any worker (of low or high ability) who accepts it. Since Lemma 13.0.1 implies that no such deviation can exist in an equilibrium, we must have II"L ~ OL in any separating equilibrium. Suppose, instead, that 11'11 < 011' as in Figure 13.0.4. If we have a separating Flgur. 13.0.3 (left)
Type OL
No pooling equilibria Type 0"
exist. Figure 13.0.4 (right)
Lemma 13,0.1: In any equilibrium, whether pooling or separating, both firms must earn zero profits.
The high·ability contract in a
Proof: Let (wI.' t,J and (IV II • til) be the contracts chosen by the low- and high-ability workers, respectively (these could be the same contract), and suppose that the two firms' aggregate profits are n > O. Then one firm must be making no more than n/2. Consider a deviation by this firm in which it alTers contracts (WI. + e, tl.) and
°L ----------------------
1
separating equilibrium cannot have wli < 0Il'
464
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
SCREENING
equilibrium, then the type Ii, contract (w"ILl must lie in the hatched region of the figure (by Lemma 13.0.1, it must also have 11', > lid. To see this, note that since type 011 workers choose contract (11'1/, 11/), contract (11'" Id must lie on or below the type 011 indifference curve through (11'1/, I,,), and since type 0, workers choose (11'" I,) ovcr (11'", I,,), contract (w" Id must lie on or above the type 0, indifference curve through (11'11' III)' Suppose that firm j is offering the low-ability contract (11',,1 1.1. Then firm k # j could earn strictly positive profits by deviating and offering only a contract lying in the shaded region of the figure with a wage strictly less than 011' sllch as (Ii', i). This contract, which has 11'/1 < 011' will be accepted by all the type 011 workers and by none of the type 0, workers [since firm j will still be offering contract (11'1., I,J]. SO we must have 11'11 ~ 0" in any separating equilibrium. Since, by Lemma 13.0.1, firms break even in any equilibrium, we must in fact have 11'1. = III. and 11'11 = 011' •
---- ...--Oc (~,;,,;)------------------
Proposition 13.0.2 summarizes the discussion so far. Proposition 13.0.2: In any subgame perfect Nash equilibrium of the screening game, low-ability workers accept contract (OL' 0), and high-ability workers accept contract (OH' iH), where iH satisfies OH - e(iH. Ot! = OL - e(O, Ot!. Proposition 13.0.2 does not complete our analysis, however. Although we have established what any equilibrium must look like, we have not established that one exists. In fact, we now show that one may nOI exist. Suppose that both firms are offering the two contracts identified in Proposition 13.0.2 and illustrated in Figure 13.0.7(a). Does either firm have an incenlive to deviate? No firm can earn strictly positive profits by deviating in a manner that attracts either only high-ability or only low-ability workers (just try to find such a deviation). But what about a deviation that attracts 0/1 workers? Consider a deviation in which the deviating firm attracts all workers 10 a single pooling contract. In Figure 13.0.7(a), a contract can attract both types of workers if and only if it lies in the shaded region. There is no profitable deviation of this type if, as depicted in the figure, Ihis shaded area lies completely above the pooled break-even line. However, when some of the shaded area lies strictly below the pooled break-even line, as in Figure 13.D.7(b), a profitable deviation to a pooling contract such as (I", i) exists. In this case, 110 eql/i1ibrium exists. Even when no single pooling contract breaks the separating equilibrium, it is possible that a profitable deviation involving a pair of contracts may do so. For example, a firm can attract both types of workers by offering the contracts (IV" Id and ("'1/,1/1) depicted in Figure 13.0.8. When it does so, type 0, workers accept contract (1"/., I,) and type 0/1 workers accept (WI/,1I/)' If this pair of contracts yields the firm a positive profit, then this deviation breaks the separating contracts identified
Proof: Consider Figure 13.0.6. By Lemmas 13.0.3 and 13.0.4, we know that = (0,,0) and that 11'11 = 01/' In addition, if the type 0, workers are willing to acccpt contract (0/.,0), III must be at least as large as the level ill depicted in the
(11'/., I,J
Figure 13.0.5 (leH)
The low-ability workers must recci\c
contract (Oc,O) in an! separating equilibrium (Ii',i)
I I I
I
I I I
I
o
ill
(b)
(0/1, i/l)' •
Wc can now derive the high-ability workers' contract.
Of. (;~.t~)-----r------------
o
(al
posilive profit by also offering, in addition to its current contracts, a contract lying in the shaded region of the figure with "'/I < 0/1, such as (w, I). This contract attracts all the high-ability workers and does not change the choice of the low-ability workers. Thus, in any separating equilibrium, the high-ability contract must be
Lemma 13.0.5: In any separating equilibrium, the high-ability workers accept contract (OH' i H ), where iH satisfies OH - c(iH , Ii L ) = OL - c(O, Ot!.
I
o
figure. Note that low-ability workers are indifferent between contracts (OL' 0) and (0/1' i/l), and so 0/1 - e(il/' Od = 0, - e(O, 0,). Suppose, then, that the high-ability contract (0/1' III) has 1/1 > il/' as in the figure. Then either firm can earn a strictly
I'roof: By Lemma 13.0.3,11'1. = 0, in any separating equilibrium. Suppose that the low-ahility workers' contract is instead some point (0/., li.1 with Ii. > 0, as in Figure 13.0.5. (Although it is not important for the proof, the high-ability contract must then lie on the segment of the high-ability break-even line lying in the hatched region of the figure, as shown.) If so, then a firm can make strictly positive profits by offering only a contract lying in the shaded region of the figure, such as (w, I). All low-ability workers accept this contract, and the contract yields the firm strictly positive profits from any worker (of low or high ability) who accepts it. •
I
SCREENING
465
Figure 13.0.7
Lemma 13.0.4: In any separating equilibrium, the low-ability workers accept contract (OL'O); that is, they receive the same contract as when no informational imperfections are present in the market.
Typ< Ii,.
13.0:
£[0] -----
Lemma 13.0.4 identifics the contract that must be accepted by low-ability workers in any separating equilibrium.
\I'
SECTION
Flgur. 13.0.6 (rig hi)
The high·ability workers must receivc
contract (Oil' ill) in any separating equilibrium.
J
An equilibrium may not exist. (a) No pooling contract breaks the separating equilibrium. (b) The pooling contracl (Ii>, i) breaks the separating equilibrium.
466
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
--- --
SCREENING
Type 0.
W
Flgur. 13.0.8 ~----------------------
(wL.td
in Proposition I3.D.2 and no equilibrium exists. More generally, an equilibrium exists only ir there is no such profitable deviation.
Welfare Properties of Screening Equilibria Restricting attention to cases in which an equilibrium does exist, the screening equilibrium has welrare properties parallel to those or the signaling model's best separating equilibrium [with r(Od = r(O.) = 0). First, as in the earlier model, asymmetric inrormation leads to Pareto inefficient outcomes. Here high-ability workers end up signing contracts that make them engage in completely unproductive and disutility-producing tasks merely to distinguish themselves rrom their less able counterparts. As in the signaling model, the low-ability workers are always worse ofT here when screening is possible than when it is not. One difference rrom the signaling model, however, is that in cases where an equilibrium exists, screening must make the high-ability workers better off; it is precisely in those cases where it would not that a move to a pooling contract breaks the separating equilibrium [see Figure 13.D.7(b)]. Indeed, when an equilibrium does exist, it is a constrained Pareto optimal outcome; ir no firm has a deviation that can attract both types or workers and yield it a positive profit, then a central authority who is unable to observe worker types cannot achieve a Pareto improvement either."
A profitable deviation uSlOg a pair of
contracts may eXist that breaks the separating equilibrium.
A P PEN 0 I X
A:
A E A SON A B L E • BEL I E F 8
A E FIN E MEN T SIN
S I G N A LIN 0
money. But if (w', t') is withdrawn as a result, then low·ability workers will accept (w, i) and this deviation ends up being unprofitable. Hellwig (1986) examines sequential equilibria and their refinements in a game that explicitly allows for such withdrawals. By introducing such reactions, these papers establish the existence of pure strategy equilibria. Introducing reactions of this sort does not simply eliminate the nonexistence problem. however, but also yields somewhat different predictions regarding the character· istics of market equilibria and their welfare properties. For example, when firms can make multiple offers as we have allowed here, cross·subsidization can arise in Wilson equilibria. Indeed, Miyazaki (1977) shows that in the case in which multiple offers are possible, a Wilson equilibrium always exists and is necessarily a constrained Pareto optimum. In the screening model examined above. we took the view that the uninformed firms made employment offers to the informed workers. Vet we could equally well imagine a model in which informed workers instead make contract offers to the firms. For example, each worker might propose a task level at which she is willing to work, and firms might then offer a wage for that task level. Note, however, that this alternative model exactly parallels the signaling model in Section 13.C and. as we have seen, yields quite different predictions. For example. the signaling model has numerous equilibria, but here we have at most a single equilibrium. This is somewhat disturbing. Given that our models are inevitably simplifications of actllal market processes, if market outcomes are really very sensitive to issues such as this our models may provide us with little predictive ability. One approach to this problem is offered by Maskin and Tirole (1992). They note that contracts like those we have allowed firms to offer in the screening model discussed in this section are still somewhat restricted. In particular, we could imagine a firm offering a worker a contract that involved an ex post (after signing) choice among a set of wage-task pairs (you will sec more about contracts of this type in Section 14.C). Similarly, in considering the counterpart model in which workers make offers. we could allow a worker to propose such a contract. Maskin and Tirole (1992) show that with this enrichment or the allowed contracts (and a weak additional assumption) the sets of sequential equilibria of the two models coincide (there may be mUltiple equilibria in both cases).
APPENDIX A: REASONABLE·BELIEFS REFINEMENTS IN SIGNALING GAMES
What can be said about the potential nonexistence of equilibrium in this model? Two paths have been followed in the literature. One approach is to establish existence of equilibria in the larger strategy space that allows for mixed strategies; on this, see Dasgupta and Maskin (1986). The other is to take the position that the lack of equilibria indicates that, in some important way. the model is incompletely specified. The aspect the literature has emphasized in this regard is the lack of any dynamic reactions to new contract offers [see Wilson (1977), Riley (1979), and Hellwig (1986»). Wilson (1977), for example, uses a definition of equilibrium that captures the idea that firms are able to withdraw unprofitable contracts from the market. A set of contracts is a Wilson equilibrium if no firm has a profitable deviation that remains profitable once existing contracts that lose money after the deviation are withdrawn. This extra requirement may make deviations less attractive. In the deviation considered in Figure 13.0.3, for example, once contract (w, i) is introduced, the original contract (w', t') loses 25. Actually, there is a small gap: An equilibrium may exist when there is another pair of Contracts that would give higher utility to both types of workers and that would yield the firm deviating to it exactly zero profits. In this case, the equilibrium is not a constrained Pareto optimum.
In this appendix, we describe several commonly used reasonable-beliefs refinements or the perrect Bayesian and sequential equilibrium concepts for signaling games, and we apply them to the education signaling model discussed in Section 13.c. Excellent sources for rurther details and discussion are Cho and Kreps (1987) and Fudenberg and Tirole (1992). Consider the rollowing class or signaling games: There are I players plus nature. The first move or the game is nature's, who picks a "type" for player I, E 0 = {O" ... , ON}' The probability ortype is [(0), and this is common knowledge among the players. However, only player I observes O. The second move is player I's, who picks an action a rrom set A aner observing O. Then, aner seeing player I's action choice (but not her type), each player i = 2, ... , I simultaneously chooses an action s, rrom set S,. We define S = S2 X . . . x St. Ir player I is of type 0, her utility rrom choosing action a and having players 2, ... , I choose s = (S2" .. ,St) is ",(a, s, 0). Player ii'I receives payoff ",(a, s, 0) in this event. A perrect Bayesian
e
a
0 A M ES
467
468
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
ANO
SCREENING
equilibrium (PBE) in the sense used in Section 13.C is a profile of strategies
(a(O), s2(a), ... ,s,(a)), combined with a common belief function 1'(0 Ia) for players 2, ... ,I that assigns a probability 1'(0 I a) to type 0 of player I conditional on
---- --
observing action a E A, such that (i) Player l's strategy is optimal given the strategies of players 2, ... , I. (ii) The belief function }leO I a) is derived from player l's strategy using Bayes' rule where possible. (iii) The strategies of players 2, ... , 1 specify actions following each choice a E A that constitute a Nash equilibrium of the simultaneous-move game in which the probability that player I is of type II is 11(11 I a) for all II E 0. In the context of the model under study here, this notion of a PBE is equivalent to the sequential equilibrium notion. The education signaling model in Section 13.C falls into this category of signaling games if we do not explicitly model the worker's choice between the firms' offers and instead simply incorporate into the payoff functions the implications of her optimal choice (she chooses from among the firms offering the highest wage if this wage is positive and refuses both firms' offers otherwise). In that model, 1 = 3,0 = {II,., II/I}, the set A = :r: r;e, O} contains the possible education choices of the worker, and the set Si = : II': \\' E IR) contains the possible wage offers by firm i.
Domillatioll-Based Refillemellts of Beliefs The simplest reasonable-belief refinement of the PBE notion arises from the idea (discussed in Section 9.0) that reasonable beliefs should not assign positive probability to a player taking an action that is strictly dominated for her. In a signaling game, this problem can arise when players 2, ... ,I (the firms in the education signaling model) assign a probability }leO I a) > 0 to player I (the worker) being of type 0 after observing action a, even though action a is a strictly dominated choice for player I when she is of type O. Formally, we say that action a E A is a strictly dominated choice for type 0 if there is an action a' E A such that Min IIt(a', s', 0) > Max " t (a, s, 0). 2"
REASONABLE.BELIEFS
REFINEMENTS
IN
SIGNALING
Unfortunately, in the education signaling model discussed in Section l3.C, this refinement does not narrow down our predictions at all. The set 0(e) equals {OL' Ou} for all education levels e because either worker type will find e to be her optimal choice if the wage offered in response to e is sufficiently in excess of the wage offered at other education levels. Thus, no beliefs are ruled out, and all PBEs of the signaling game pass this test. If we want to narrow down our predictions for this model, we need to go beyond the use of refinements based only on notions of strict dominance. 28 Recall the argument we made in Section 13.C for eliminating all separating equilibria but the best one. We argued that since, in Figure 13.C.7, a worker of type 0,. would be better off choosing e = 0 than she would choosing an education level above i' for any beliefs and resulling equilibrium wage Ihal mighl follow Ihese Iwo edllcalion lel'e/s, no reasonable belief should assign a positive probability to a worker of type II,. choosing any e > e. This is close to an argument that education levels e > e are dominated choices for a type OL worker, but with the critical difference reflected in the italicized phrase: Only equilibrium responses of the firms are considered, rather than all conceivable responses. That is, we take a backwardinduction-like view that the worker should only concern herself with possible equilibrium reactions to her education choices. To be more formal about this idea, for any nonempty set 0 c 0, let S*(0, a) c S, x ... x S, denote the set of possible equilibrium responses that can arise after action a is observed for some beliefs satisfying the property that 1'(11 I a) > 0 only if II E 0. The set S*(0, a) contains the set of equilibrium responses by players 2, ... ,I that can follow action choice a for some beliefs that assign positive probability only to types in 0. When 0 = 0, the set of all conceivable types of player I, this construction allows for all possible beliefs. 29 We can now say that action a E A is strictly dominated for type 0 in this stronger sense if there exists an action a' with Min
u,(a', s', 0)
> Max
u,(a, s, 0).
(13.AA.2)
nS·(9.4I'
Using this stronger notion of dominance, we can define the set 0*(a) = {O: there is no a'
E
A satisfying (13.AA.2)j,
containing those types of player I for whom action a is not strictly dominated in the sense of (\ 3.AA.2). We can now say that a PBE has reasonable beliefs if for all a E A with 0*(a) i' 0, I/(a, II) > 0 only if 0 E 0*(a). Using this reasonable-beliefs refinement significantly reduces the set of possible outcomes in the educational signaling model, sometimes even to a unique prediction. In that model, S*(0, e) = [OL' 011] for all education choices e because, for any belief I' E [0, I], the resulting Nash equilibrium wage must lie between OL and 0/1' As a
(13.AA.I)
For each action a E A, it is useful to define the set 0(a) = [0: there is no a' E A satisfying (13.AA.I»).
This is the set of types of player I for whom action a is not a strictly dominated choice. We can then say that a PBE has reasonable beliefs if, for all a E A with 0(a) i' 0, onlyif
A:
s'£S·(9.o',
Sf,S
II(Ola»O
APPENDIX
2S. We could, in principle, go further with this identification of strictly dominated strategies for player I by also eliminating any strictly dominated strategies for players 2, ... ,I, then looking to see whether we have any more strictly dominated actions for any of player I's types, and so on. However, in the educational signaling model, this does not help us because the firms have no strictly dominated strategies. 29. Note that when there is only one player responding (so I = 2), the set S'(0, a) is exactly the set of responses that are not strictly dominated for player 2 conditional on following action a. Note also that in this case a strategy s,ta) is weakly dominated for player 2 if, for any a e A, it involves play of some Sf S'(0, a).
OE0(a)
and we consider a PBE to be a sensible prediction only if it has reasonable beliefs."
26. Note that a strategy a(O) is strictly dominated for player I if and only if it involves play of a strictly dominated action for some type O. 27. Doing this is equivalent to first eliminating each type (J's dominated actions from the game and then identifying the PBEs of this simplified game.
J
GAMES
469
470
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING.
ANO
SCREENING
~~~---------------------------------------------------Type
(I,.
<.----- Type
(/"
Figure 13.AA.l (left)
..
Type 0"
A pooling equilibrium
that is eliminated using the dominance test in (I3.AA.2).
0"
I _._._.-.- "'(e)
·C-
--
.::.~~----------
(lL
I , , :
o
e
consequence, an education choice in excess of in Figure 13.C.7 is dominated for a type 0,. worker according to the test in (I3.AA.2) by the education choice e = O. Hence, in any PRE with reasonable beliefs, 1,(0,,1 e) = I for all e> e. But if this is so. then no separating equilibrIum with e'(O,,) > e can survive because, as we argued in Section D.C. the high-ability worker will do better by deviating to an education level slightly in excess of e. Furthermore, we can also eliminate any pooling equilibrium in which the equilibrium outcome is worse for a high-ability worker than outcome (Ii", e), such as in the equilibrium depicted in Figure 13.AA.I, since any such equilibrium must involvc unreasonable beliefs: If 1,(0" 1e) = I for all e > thcn a type Ii" worker could do better deviating to an education level just above where she would receive a wage of 0". In fact, when the high-ability worker prefers outcome (0", e) to (£[0],0), this argument rules out all pooling equilibria, and so we get the unique prediction of the best separating equilibrium.
e, e
E(/UiliiJl'iulIl Domination and the Intuitive Criterion We now consider a further strengthening of the notion of dominance, known as <''1uililJriulll dOlllinance. This leads to a refinement known as the intuitive criteriol! [Cho and Kreps (1987)] that always gives us the unique prediction of the best separating equilibrium in the two-type education signaling model studied in Section IlC. The idea behind this refinement can be seen by considering the pooling equilibrium of the education signaling model that is shown in Figure 13.AA.2, an equilibrium that is not eliminated by our previous refinements. Note that, as illustrated in the figure, to support education choice e' as a pooling equilibrium outcome we must have beliefs for the firms satisfying 1,(0" 1e) < I for all e E (e', e"). Indeed, if 11(0" 1e) = I at any such education level, then the wage offered would be 0" and the type 0" worker would find it optimal to deviate. Suppose. however, that a firm is confronted with a deviation to some education level r: E (e', e") when it was expecting the equilibrium level of education e' to be chosen. It might reason as follows: .. Either type of worker could be sure of getting outcome (II'. e) = (£[0], eO) by choosing the equilibrium education level eO. But a low-ability worker would be worse off deviating to education level e' regardless of what beliefs firms have after this choice, while a high-ability worker might be made better off by doing this. Thus, this must not be a low-ability worker." In this case, the choice of e' by the low-ability worker is dominated by her equilibrium payoff.
REF I N E MEN T S I N
S I G N A liN G
To formalize this idea in terms of our general specification, denote the equilibrium payoff to type 0 in PBE (a"(O), s"(a), p) by u~(O) = u,(a"(O), s"(a"(O), 0). We then say that action a is equilibrium dominated for type 0 in PBE (a·(O), s'(a), p) if
> Max u,(a, s, 0).
(I3.AA.3)
Using this notion of dominance, define for each a E A the set e"(a) = (0: condition (I3.AAJ) does not hold). We can nOw say that a PBE has reasonable beliefs if for all actions a with e"(a) # 0, p(OI a) > 0 only if 0 E e'"(a), and we can restrict attention to those PBEs that have reasonable beliefs. Note that any action a that is dominated in the sense of (13.AA.2) for type 0 must also be equilibrium dominated for this type because u~(O) = u~(a'(O), s'(a'(O)), 0) > Min,·.s"" .•·, u,(a', s', 0) by the definition ofa PRE. Thus, this equilibrium dominancebased procedure must rule out all the PREs that were ruled out by our earlier procedure and may rule out more. Consider the use of this refinement in the education signaling model of Section 13.c. Since it is stronger than the refinement based On (13.AA.2), this refinement also eliminates ail but the best separating equilibrium. However, unlike our earlier dominance-based refinements, the equilibrium dominance-based refinement also eliminates all pooling equilibria. For example, in the pooling equilibrium depicted in Figure 13.AA.2, any education choice i! E (e', e") is equilibrium dominated for the low-ability worker. Moreover, once the firms' beliefs following this education choice are restricted to assigning probability I to the worker being type 01/, the high-ability worker wishes to deviate to this education level. Thus, we get a unique prediction for the outcome in this game: the best separating equilibrium. In signaling games with two types, this equilibrium dominance-based refinement is equivalent to the intuitive criterion proposed in Cho and Kreps (1987). Formally, a PBE is said to violate the intuitive criterion if there exists a type 0 and an action a E A such that
A pooling eqUilibrium that is eliminated using the dominance test in (l3.AA.3).
I I I I
REA. SOH A B L E ~ 8 ELI E F S
seS-(8 .• )
Figure 13.AA.2 (right)
---f--------+-----------
A:
u~(O)
I
I I
A P PEN 0 I X
Min
u,(a, s, 0) > "t(O).
(I3.AAA)
Thus, we eliminate a PBE using the intuitive criterion if there is some type 0 who has a deviation that is assured of yielding her a payoff above her equilibrium payoff as long as players 2, ... , I do not assign a positive probability to the deviation having been made by any type 0 for whom this action is equilibrium dominated. We can think of the intuitive criterion as saying that to eliminate a PBE we must find a type of player I who wants to deviate even if she is not sure what exact belief of players 2, ... , I will result, she is only sure that they will not think she is a type who would find the deviation to be an equilibrium-dominated action. In general, the intuitive criterion is a more conservative elimination procedure than just insisting on PBEs involving reasonable beliefs using set e"(a) because any PBE with reasonable beliefs using set e"(a) passes the intuitive criterion's test, but as Example I3.AA.1 illustrates, a PBE could satisfy the intuitive criterion's test but fail to have reasonable beliefs. However. when there are only two types of player I, the two notions are equivalent. i
I
1
Example I3.AA.I: Suppose that there are three types of player I, (0"0,, OJ), and
GAM E
S
471
472
c
HAP T E R
1 3:
A 0 V E R S ESE L E C T
I0
H.
S
I G HAL I H G.
A H0
SC REEH
IN G
that in some PBE the out-of-equilibrium action ais equilibrium dominated for type 0 1 only. so that e""(a) = {O,. OJ}. Suppose also that type Ii, strictly prefers to deviate to action if and only if beliefs over types Ii, and IiJ have /J(Ii,1 d)
*
a
--- --
EX ERe I S E S
EXERCISES 13.B.I A Consider three functions of 0: r(0), £[010:s; 0], and O. Graph these three functions over the domain [Q, 0], assuming that the first two functions are continuous in 0 but allowing them to be otherwise quite arbitrary. Identify the competitive equilibria of the adverse selection model of Section I3.B using this diagram. What about the Pareto optimal labor allocation? Now produce a diagram to depict each of the situations in Figures 13.B.1 to 13.B.3.
a;
13.B.2" Suppose that r(') is a continuous and strictly increasing function and that there exists {j E (q, til such thai reO) > 0 for 0 > 0 and r(0) < 0 for 0 < O. Let the densilY of workers of type 0 be f(O), with f(O) > 0 for all 0 E [Q, 0]. Show that a competitive equilibrium with unobservable worker types necessarily involves a Pareto inefficient outcome.
Although the use of either equilibrium domination or the intuitive criterion yields a unique prediction in the education signaling model when there are two types of workers, they do not accomplish this when there are three or more possible worker types (see Exercise 13.AA.l). 'Stronger refinements such as Banks and Sobel's (1987) notions of divilliry and universal diviniry, Cho and Kreps' (1987) related notion called D I, and Kohlberg and Mertens' (1986) srabiliry do yield the unique prediction of the best separating equilibrium in these games with many worker types. See Cho and Kreps (1987) and Fudenberg and Tirole (1992) for further details.
13.B.3" Consider a positive selection version of the model discussed in Section 13.B in which r(') is a continuous, strictly decreasing function of O. Let the density of workers of type Ii be f(O), wilh f(O) > 0 for all 0 E [Q, 0].
(a) Show that the more capable workers are the ones choosing to work at any given wage. (b) Show that if reO) > 0 for all 0, then the resulting competitive equilibrium is Pareto efficient.
(e) Suppose that there exists a (; such Ihat reO) < 0 for 0 > () and reO) > 0 for 0 < 0. Show that any competitive equilibrium with strictly positive employment necessarily involves too
ml/eh employment relative to the Pareto optimal allocation of workers. 13.B.4" Suppose two individuals, I and 2, are considering a trade at price p of an asset that they bOlh use only as a store of wealth. Ms. I is currently the owner. Each individual i has a privately observed signal of the asset's worth YI' In addition, each cares only about the expected value of the asset one year from now. Assume that a trade at price p takes place only if both parties think they are being made strictly better off. Prove that the probability of trade occurring is zero. [Hilll: Study the following trading game: The two individuals simullaneously say either "trade" or "no trade," and a trade at price p takes place only if they bOlh say" trade."]
REFERENCES Akcrlof. G. (1970). The market for lemons: Quality uncertainly and the market mechanism. Quarterly JOU"'QI of Economics 89: 488-500. Banks. J.. and J. Sobel. (1987). Equilibrium selection in signaling games. £wnomt'lr;ca 55: 647-62. Cho. I·K .. and D. M. Kreps. (1987). Signaling games and stable equilibria. Quarurly Journal of Economics 102: 179-221. Dasgupta. P., and E. Maskin. (1986). The existence of equilibrium in discontinuous economic games. Rt'lliew of Economic Studies 46: 1-41. fudcnbcrg. D .. and 1. Tirole. (1992). Game Thear),. Cambridge, Mass.: M IT Press. Hellwig, M. (1986). Some recent developments in the theory of competition in markets with adverse selection. (University or Bonn. mimeographed). Holmstrom, B., and R. B. Myerson. (1983). Efficient and durable decision rules with incomplete information. Econometrica 51: 1799-819. Kohlberg. E.. and J.-F. Mertens. (1986). On the strategic stability of equilibria. ECotlOmetrica 54: 1003-38. Maskin, E., and J. Tirole. (1992). The principal.agcnt relationship with an informed principal,ll: Common values. Econometrica 60: 1-42. Miyazaki. H. (1977). The rat race and internal labor markets. Bell JOUr/wi of Economics 8: 394-418. Riley, 1. (1979). Informational equilibrium. Econometrica 47: 331-59. Rothschild, M .. and 1. E. Stiglitz. (1976). Equilibrium in competitive insurance mark.ets: An essay in the economics of imperfect information. Quarterly Journal of Economics 80: 629-49. Spence. A. M. (1973). Job mark.l signaling. Quarl.,ly Journal of Economics 87: 355-74. Spence, A. M. (1974). Markel Signaling. Cambridge. Mass.: Harvard University Press. Wilson, C. (1977). A model of insurance markets with incomplete information. Journal of Economic Theory 16: 167-207. Wilson, C. (1980). The nature or equilibrium in markets with adverse selection. Bell Journal of Economics 11: 108-30.
I3.B.5" Reconsider the case where reO) = r for all 0, but now assume thai when the wage is such thai no workers are accepting employment firms believe that any worker who might accept would be of the lowest quality, that is, £[0 Ie = 0] = Q. Maintain the assumption that all workers accept employment when indifferent. (a) Argue that when £[0] :2: r > Q, there are now two competitive equilibria: one with e" = [Q,O] and one with w" = Q and e" = 0. Also show that when Q :2: r Ihe unique competitive equilibrium is w" = £[0] and e" = [Q,O], and when r > £[0] the unique competitive equilibrium is w" = Q and e" = 0.
1\""
= £[0] and
(b) Show that when £[0] > r and there are two equilibria, the full-employment equilibrium Pareto dominates the no-employment one. (e) Argue that when £[0] :2: r the unique SPNE of the game-theoretic model in which two firms simultaneously make wage offers is the competitive equilibrium when this equilibrium is unique, and is the full-employment (highest-wage) competitive equilibrium when the competitive equilibrium is not unique and £[0] > r. What happens when £[0] = r? What about the case where £[0] < r? (d) Argue that the highest-wage competitive equilibrium is a constrained Pareto optimum.
J
473
474
CHAPTER
13:
ADVERSE
SELECTION,
SIGNALING,
AND
SCREENING
13.B.6C [Based On Wilson (1980)] Consider the following change in the adverse selection model of Section 13.B. Now there are N firms. each of which wants to hire at most I worker. The N firms differ in their productivity: In a firm of type y a worker of type 0 produces yO units of output. The parameter y is distributed with density function g(') on [0,00], and Y(i') > 0 for all y E [0, 00].
---
E X E R CIS E S
13.C2C Reconsider the two-type signaling model with r(Od = r(8 H ) = 0, assuming a worker's productivity is 0(1 + I'e) with I' > O. Identify the separating and pooling perfect Bayesian equilibria, and relate them to the perfect information competitive outcome. 13.C3" In text.
(a) Let :(11', II) denote the aggregate demand for labor when the wage is II' and the average productivity of workers accepting employment at that wage is II. Derive an expression for this function in terms of the density function g(').
I3.C4" Reconsider the signaling model discussed in Section I3.C, now assuming that worker types are drawn from the interval [Q,O] with a density function f(O) that is strictly positive everywhere on this interval. Let the cost function be c(e, 0) = (e' /0). Derive the (unique) perfect
(b) Let 11(\\') = £[Olr(O),;; w], and define the aggregate demalld fUlletiall for labor by z*(\\') = :(\\',11(\\')). Show that :*(11') is strictly increasing in II' at wage IV if and only if the
Bayesian equilibrium.
elasticity of It with respect to w exceeds 1 at wage \\. (assume that all relevant functions arc uifferentiable).
13.CS" Assume a single firm and a single consumer. The firm's product may be either high or low qualily and is of high quality with probability i .. The Consumer cannot observe quality before purchase and is risk neutral. The consumer's valuation of a high-quality product is v,,; her valuation of a low-quality product is VL' The costs of production for high (H) and low (L) quality are <'/I and f,., respectively. The consumer desires at most one unit of the product. Finally, the firm's price is regulated and is set at p. Assume that v" > P > vL > fH > fL'
f;' ', •.,
(0) Let "(11') = f(O) dO denote the aggregate supply fum·tioll of labar, and define a competitive equilibrium wage IV* as one where z*(w*) = 5(11'*). Show that if there are multiple competitive equilibria, then the one with the highest wage I'areto dominates all the others.
(d) Consider a game-theoretic model in which the firms make simultaneous wage offers, (lnu denote the highest !,ompetitive equilibrium wage by 11'*. Show that (i) only the highest-wage competitive equilibrium can arise as an SPNE, and (ii) the highest-wage competitive equilibrium is (In SPNE if and only if z*(\\') ,;; z*(\\'*) for all II' > \\'*.
(a) Given the level of p, under what conditions will the consumer buy the product? (b) Suppose that before the consumer decides whether tq buy, the firm (which knows its type) can advertise. Advertising conveys no information directly, but consumers can observe the total amount of money that the firm is spending on advertising, denoted by A. Can there be a separating perfect Bayesian equilibrium, that is, an equilibrium in which the consumer rationally expects firms with different quality levels to pick ditTerent levels of advertising?
13.8.7" Suppose that it is impossible to observe worker types and consider a competitive equilibrium with wage rate \\'*. Show that there is a Pareto-improving market intervention (Ii". Ii',) that reduces employment if and only if there is one of the form (11',., \\',) = (\\'*, IV,) with Ii', > O. Similarly, argue that there is a Pareto·improving market intervention (Ii'" "',) that increases employment ifand only if there is one ofthcform (W., IV,) = (IV., 0) with Ii', > "'*. Can you use these facts to give a simple proof of Proposition 13.8.2?
13.C6' Consider a market for loans to finance investment projects. All investment projects require an outlay of I dollar. There are two types of projects: good and bad. A good project > 0 and a probability (I - Po) of yielding has a probability of Po of yielding profits of profits of zero. For a bad project. the relative probabilities are P. and (I - P.), respectively, where Po > P" The fraction of projects that are good is i, E (0, I). Entrepreneurs go to banks to borrow the cash to make the initial outlay (assume for now that they borrow the entire amount). A loan contract specifies an amount R that is supposed to be repaid to the bank. Entrepreneurs know the type of project they have, but the banks do not. In the event that a project yields profits of zero, the entrepreneur defaults on her loan contract. and the bank receives nothing. Banks are competitive and risk neutral. The risk-free rate of interest (the rate the banks pay to borrow funds) is r. Assume that
n
I3.B.S" Consider the following alteration to the adverse selection model in Section I3.B. Imagine that when workers engage in home production, they use product x. Suppose that the amount consumed is related to a worker's type, with the relation given by the increasing function x(O). Show that if a central authority can observe purchases of good x but not worker types, then there is a market intervention that results in a Pareto improvement even if the market is at the highest·wage competitive equilibrium. 13.8.9" Consider a model of positive seleetioll in which r(') is strictly decreasing and there are two types of workers, 0" and 01., with 00 > 0" > 01• > O. Let i. = Prob (0 = 0,,) E (0,1). Assume that r(O,,) < 0" and that r(Od> 0L' Show that the highest-wage competitive equilibrium need not be a constrained Pareto optimum. [Hint: Consider introducing a small unemployment benefit for a case in which £[0] = r(Od. Can you use the result in Exercise 13.8.7 to give an exact condition for when a competitive equilibrium involving full employment is a constrained Pareto optimum?]
Pc; n
on
-
(I + r) > 0 > p. n
-
(I
+ r).
(a) Find the equilibrium level of R and the set of projects financed. How does this depend i., n, and r'~
{'G' PH'
(b) Now suppose that the entrepreneur can offer to contribute some fraction x of the I dollar initial outlay from her own funds (x E [0, I]). The entrepreneur is liquidity constrained, howe,.r. so that the effective cost of doing so is (I + I')x, where I' > r. (i) What is an entrepreneur's payoff as a function of her project type. her loan-repayment amount R, and her contribution x? (ii) Describe the best (from a welfare perspective) separating perfect Bayesian equifibrium of a game in which the entrepreneur first makes an offer that specifies the level of x she is willing to put into a project, banks then respond by making offers specifying the fevel of R they would require, and finally the entrepreneur accepts a bank's offer or decides not to go ahead with the project. How does the amount contributed by entrepreneurs with good projects change with small changes in P" PG, A, n, and r?
13.B.10" Show that Proposition 13.8.2 continues to hold when r(O) > 0 for some O. l3.CI" Consider a game in which, first, nature draws a worker's type from some continuous distribution on [Q, 0]. Once the worker observes her type, she can choose whether to submit to a cost less test that reveals her ability perfectly. Finally, after observing whether the worker has taken the test and its outcome if she has, two firms bid for the worker's services. Prove that in any subgame perfect Nash equilibrium of this model all worker types submit to the test, and firms offer a wage no greater than Q to any worker not doing so.
J
475
476
C HAP T E R
1 3:
A 0 V E R S ESe LEe T ION.
5 I G N A LIN G.
AND
C HAP T E R
S eRe e N I N G
----------------------------------------------------------------------(iii) How do the two types of entrepreneurs do in the separating equilibrium of (b)(ii) compared with the equilibrium in (o)?
The Principal-Agent Problem
\3.0.1" Extend the screening model to a case in which tasks are productive. Assume that a type 0 worker produces O( I + 1'1) units of output when her task level is r where II > O. Identify the subgame perfect Nash equilibria of this model.
14
I3.D.2" Consider the following model of the insurance market. There are two types of individuals: high risk and low risk. Each starts with initial wealth W but has a chance that "n accident (c.g., a firc) will reduce her wealth by L. The probability of this happening is PL for low· risk types and PH for high-risk types, where PH > Pl.. Both types are expected utility m"ximizers with a Bernoulli utility function over wealth of u(w), with u'(w) > 0 and u"(I\') < 0 at all
1\'.
There arc two risk-neutral insurance companies. An insurance policy consists of a
premium payment M made by the insured individual to her insurance firm and a paymcnt R from the insurance company to the insured individual in the event of a loss. (0) Suppose that individuals are prohibited from buying more than one insurancc policy.
Arguc that a policy can be thought of as specifying the wcalth levels of the insured individual in the two states "no loss'" and "loss."
i
I 14,A Introduction In Chapter 13, we considered situations in which asymmetries of information exist between individuals at the time of contracting. In this chapter, we shift our attention to asymmetries of information that develop subsequem to the signing of a contract. Even when informational asymmetries do not exist at the time of contracting, the parties to a contract often anticipate that asymmetries will develop sometime after the contract is signed. For example, after an owner of a firm hires a manager, the owner may be unable to observe how much effort the manager puts into the job. Similarly, the manager will often end up having better information than the owner about the opportunities available to the firm. Anticipating the development of such informational asymmetries. the contracting parties seek to design a contract that mitigates the difficulties they cause. These problems arc endemic to situations in which one individual hires another to take some action for him as his "agent." For this reason. this contract design problem has come to be known as the principal-agent problem. The literature has traditionally distinguished between two types of informational problems that can arise in these settings: those resulting from hidden actions and those resulting from hidden in/ormation. The hidden action case. also known as moral hazard. is illustrated by the owner's inability to observe how hard his manager is working; the manager's coming to possess superior information about the firm's opportunities, on the other hand. is an example of hidden information.' Although many economic situations (and some of the literature) contain elements of both types of problems. it is useful to begin by studying each in isolation. In Section 14.B, we introduce and study a model of hidden actions. Section 14.C analyzes
(b) Assumc that the insuranc~ companies simultaneously olTer policies; as in Section 13.0, they can each olTer any finitc number of policics. What are the subgamc perfcct Nash equilibrium outcomes of the model'! Docs an equilibrium necessarily exist'!
I3.D.3 c Consider the following extension of the model you developed in Exercise 13.0.1. Suppose that therc is a fixed task level T that all workers face. The monetary equivalent cost of accepting cmployment at this task level is c > 0, which is independent of worker type. However, now a worker's actual output is observable and verifiable. and so contracts can base I;ompensation on the worker's ex post observed output level. (0) What is the subgame perfect Nash equilibrium outcome of this model?
(b) Now suppose that the output realization is random. It can be either good (qG) or bad ('I.). Thc probability th"t it is good is PH for a high-ability worker and PI. for a low-ability worker (Pu > Pt.>. If workers are risk-neutral expected utility maximizers with a Bernoulli utility fUllction over wealth of u(w)
= "',
what is the subgame perfect Nash equilibrium
outcomc? (\:) \Vltat ir workers are strictly risk averse with II"(\\,) < 0 at all w?
13.\).4" Reconsider the scrcening model in Section 13.0, but assume that (i) there is an infinite nllmber of firms that could potentially enter the industry and (ii) firms can each offer at most one contract. [The implication of (i) is that, in any SPNE, no firm can have a profitable entry opportunity.] Characterize the equilibria for this casc. I3.AA.l c Consider the extension of the signaling model discussed in Section \3.C to thc OiISC of three types. Assumc all thrce types have rIO) = O. Provide an example in which more than olle perfect Bayesian equilihrium satisfies the intuitive critcrion.
1. The literature's use of the term moral hazard is not entirely uniform. The term originates in the insurance literature. which first focused attention on two types of informational imperfections: the "moral hazard" that arises when an insurance company cannot observe whether the insured exerts effort to prevent a loss and the "adverse selection" (see Section 13. B) that occurs when the
insured knows more than the company at the time he purchases a policy about his likelihood of an accident. Some authors use moral hazard to refer to either of the hidden action or hidden
information variants of the principal-agent problem [see, for example, Hart and Holmstrom (1987)]. Here, however, we use the term in the original sense.
j
477
478
C HAP T E R
1.:
THE
P R INC I PAL. AGE N T
PRO B L E M
---------------------------------------------------------------------a hidden information model. Then, in Section 14.0, we provide a brief discussion of hybrid models that contain both of these features. We shall see that the presence of postcontractual asymmetric information often leads to welfare losses for the contracting parties relative to what would be achievable in the absence of these informational imperfections. It is important to emphasize the broad range of economic relationships that fit into the general framework of the principal-agent problem. The owner-manager relationship is only one example; others include insurance companies and insured individuals (the insurance company cannot observe how much care is exercised by the insured), manufacturers and their distributors (the manufacturer may not be able to observe the market conditions faced by the distributor), a firm and its workforce (the firm may have more information than its workers about the true state of demand for its products and therefore about the value of the workers' product), and banks and borrowers (the bank may have difficulty observing whether the borrower uses the loaned funds for the purpose for which the loan was granted). As would be expected given this diversity of examples, the principal-agent framework has found application in a broad range 'of applied fields in economics. Our discussion will focus on the owner-manager problem. The analysis in this chapter, particularly that in Section 14.C, is closely related to that in two other chapters. First, the techniques developed in Section 14.C can be applicd to the analysis of screening problems in which, in contrast with the case studied in Section 13.0, only one uninformed party screens informed individuals. We discuss the analysis of this monopolistic screening problem in small type at the end of Section 14.C. Second, the principal-agent problem is actually a special case of "mechanism design," the topic of Chapter 23. Thus, the material here constitutes a first pass at this more general issue. Mastery of the fundamentals of the principalagent problem, particularly the material in Section 14.C, will be helpful when you study Chapter 23. A good source for further reading on topics of this chapter is Hart and Holmstrom
--
(1987).
14.B Hidden Actions (Moral Hazard) Imagine that the owner of a firm (the principal) wishes to hire a manager (the agent) for a one-time project. The project's profits are affected, at least in part, by the manager's actions. If these actions were observable, the contracting problem between the owner and the manager would be relatively straightforward; the contract would simply specify the exact actions to be taken by the manager and the compensation (wage payment) that the owner is to provide in return. 2 When the manager's actions are not observable, however, the contract can no longer specify them in an effective manner, because there is simply no way to verify whether the manager has fulfilled his obligations. In this circumstance, the owner must design the manager's compensation scheme in a way that indirectly gives him the incentive to take the correct
SEC T ION
1 4 • 8:
HID DEN
ACT ION S
t
M 0 RA L
actions (those that would be contracted for if his actions were observable). In this section, we study this contract design problem. To be more specific,let It denote the project's (observable) profits, and let e denote the manager's action choice. The set of possible actions is denoted by E. We interpret e as measuring managerial effort. In the simplest case that is widely studied in the literature, e is a one-dimensional measure of how "hard" the manager works, and so E c R More generally, however, managerial effort can have many dimensionshow hard the manager works to reduce costs, how much time he spends soliciting customers, and so on-and so e could be a vector with each of its elements measuring managerial effort in a distinct activity. In this case, E c AM for some M.l In our discussion, we shall refer to e as the manager's effort choice or effort level. For the nonobservability of managerial effort to have any consequence, the manager's effort must not be perfectly deducible from observation of It. Hence, to make things interesting (and realistic), we assume that although the project's profits are affected bye, they are not fully determined by it. In particular, we assume that the firm's profit can take values in ['.!' x] and that it is stochastically related to e in a manner described by the conditional density function f(lt I e), with f(lt Ie) > 0 for all e E E and all 11 E ['.!, xl Thus, any potential realization of 11 can arise following any given effort choice by the manager. In the discussion that follows, we restrict our attention to the case in which the manager has only two possible effort choices, ell and eL (see Appendix A for a discussion of the case in which the manager has many possible actions), and we make assumptions implying that ell is a "high-effort" choice that leads to a higher profit level for the firm than eL but entails greater difficulty for the manager. This fact will mean that there is a conflict between the interests of the owner and those of the manager. More specifically, we assume that the distribution of It conditional on ell first-order stochastically dominates the distribution conditional on e/.; that is, the distribution functions F(lt Ied and F(1I I ell) satisfy F(lt I ell) ::; F(lt I ed at alllt E ['.!, x], with strict inequality on some open set n c ['.!, x] (see Section 6.0). This implies that the level of expected profits when the manager chooses ell is larger than that from eL: J1If(1I I ell) d1l > J1If(1I I eLl d1l. The manager is an expected utility maximizer with a Bernoulli utility function U(IV, e) over his wage IV and effort level e. This function satisfies u.. (w, e) > 0 and u.... (w, e) ~ 0 at all (w, e) (subscripts here denote partial derivatives) and u(\\', ell) < II(IV, "L) at all w; that is, the manager prefers more income to less, is weakly risk averse over income lotteries, and dislikes a high level of effort! In what follows, we focus on a special case of this utility function that has attracted much of the
1 In fact, more general interpretations are possible. For example. e could include non-effortrelated managerial decisions such as what kind of inputs are purchased or the strategies that are adopted for appealing to buyers. We stick to the effort interpretation largely because it helps wilh intuition. 4. Note that in the multidimensional-effort case, it need not be that eH has higher effort in every dimension; the only important thing for our analysis is that it leads to higher profits and entails a larger managerial disutility than does iL.
2. Note that this requires not only that the manager's actions be observable to the owner but also that they be observable to any court that might be called upon to enforce the contract.
1
HA Z AR DI
479
480
CHAPTER
14:
THE
PRINCIPAL-AGENT
PROBLEM
attention in the literature: u(w, e) = v(w) - g(e).' For this case, our assumptions on u(w, e) imply that v'(w) > 0, v"(w) ~ 0, and g(eH) > geed. The owner receives the project's profits less any wage payments made to the manager. We assume that the owner is risk neutral and therefore that his objective is to maximize his expected return. The idea behind this simplifying assumption is that the owner may hold a well-diversified portfolio that allows him to diversify away the risk from this project. (Exercise 14.B.2 asks you to consider the case of a risk-averse owner.)
--- --
f
(11 - W(lI» [(111 e) dll
S.t.
f
f
v(w(n»[(nl e) dn -gee)
-[(lIle)
+ yv'(w{lI))[(nle)
,,(w:> -
f
481
= 0,
(14.B.3)
gee) = ii.
(14.B.4)
n[(l1l e) dll - v-'(ii
+ gee)).
(14.B.5)
The first term in (14.8.5) represents the gross profit when the manager puts forth effort e; the second term represents the wages that must be paid to compensate the manager for this effort [derived from condition (14.B.4)]. Whether ell or eL is optimal depends on the incremental increase in expected profits from ell over eL compared with the monetary cost of the incremental disutility it causes the manager. This is summarized in Proposition 14.B.1.
to offer the manager? Second, what is the best choice of e? Given that the contract specifies effort level e, choosing w(n) to maximize S(n - w(nll[(nl e) dn = (S lI[(nl e) dn) - (S w(n)[(nle) dn) is equivalent to minimizing the expected value of the owner's compensation costs, S w(n)[(l1l e) dll, so (14.B.l) tells us that the optimal compensation scheme in this case solves
S.t.
I
or
~ ii.
Proposition 14.B.1: In the principal-agent model with observable managerial effort, an optimal contract specifies that the manager choose the effort e* that maximizes n f (nl e) dn - v- '(D + gee»~] and pays the manager a fixed wage w* = v- '(D + g(e*)). This is the uniquely optimal contract if v"(w) < 0 at all w.
(14.B.2)
w(n)[(nle)dll
H A Z A R0
Note that since Ilk,,) > g(e,,), the manager's wage will be higher if the contract calls for effort ell than if it calls for eL . On the other hand, when the manager is risk neutral, say with v(w) = w, condition (14.B.3) is necessarily satisfied for any compensation function. In this case, because there is no need for insurance, a fixed wage scheme is merely one of many possible optimal compensation schemes. Any compensation function w(n) that gives the manager an expected wage payment equal to ii + gee) [the level derived from condition (14.B.4) when v(\\') = IV] is also optimal. Now consider the optimal choice of e. The owner optimally specifies the effort level e E :e,., ell} that maximizes his expected profits less wage payments,
It is convenient to think of this problem in two stages. First, for each choice of
f
(M 0 R A L
w:
e that might be specified in the contract, what is the best compensation scheme w(n)
Min
ACT ION S
If the manager is strictly risk averse [so that v'(w) is strictly decreasing in w], the implication of condition (14.B.3) is that the optimal compensation scheme W{lI) is a constant; that is, the owner should provide the manager with a fixed wage payment. This finding is just a risk-sharing result: Given that the contract explicitly dictates the manager's effort choice and that there is no problem with providing incentives, the risk-neutral owner should fully insure the risk-averse manager against any risk in his income stream (in a manner similar to that in Example 6.C.1). Hence, such given the contract's specification of e, the owner offers a fixed wage payment that the manager receives exactly his reservation utility level:
(l4.B.I)
V(W(lI)) [(111 e) dn - I/(e)
HID 0 E N
condition 6
I
The Optimal Contract when EjJort is Observable
Max
1 4. B:
V'(W{lI» = y.
It is uscful to begin our analysis by looking at the optimal contracting problem when effort is observable. Suppose that the owner chooses a contract to offer the manager that the manager can then either accept or rcject. A contract here specifies the manager's effort e E {e,., ell} and his wage payment as a function of observed profits W(lI). We assume that a competitive market.for managers dictates that the owner must provide the manager with an expected utility level of at least ii if he is to accept the owner's contract offer ('I is the manager's reser"",ioll lIIility le,'e/). If the manager rejects the owner's contract offer, the owner receives a payoff of zero. We assume throughout that the owner finds it worthwhile to make the manager an offer that he will accept. The optimal contract for the owner then solves the following problem (for notational simplicity, we suppress the lower and upper limits of integration ~ and x):
,'EI"/.t'"I. ".. _I
SEC T ION
0
~ 'I.
6. The first-order condition ror w{tt) is derived by taking the derivative with respect to the manager's wage at each level or 1t separately. To see this point, consider a discrete version or the model in which there is a finite number or possible profit levels (7[1 •...• 1t N ) and associated wage levels (Wi •...• wN ). The first-order condition (14.8.3) is analogous to the condition onc gels in the discrete model by examining the first-order conditions ror each W II , n = I..... N (note that we allow the wage payment to be negative). To be rigorous. we should add that when we have a continuum of possible levels or 7[, an optimal compensation scheme need only satisry condition
The constraint in (14.B.2) always binds at a solution to this problem; otherwise, the owner could lower the manager's wages while still getting him to accept the contract. Letting y denote the multiplier on this constraint, at a solution to problem (14.B.2) the manager's wage W(lI) at each level of n E [~, x] must satisfy the first-order
(14.B.3) aL a seL of profiL levels LhaL is of full measure.
5. Exercise 14.B.1 considers one implication or relaxing this assumption.
.1
------"
482
CHAPTER
14:
THE
PRINCIPAL·AGENT
PROBLEM
-------------------------------------------------------------------------The Optimal Colltract whell Effort is Not Observable
-
The optimal contract described in Proposition 14.B.1 accomplishes two goals: it specifics an efficient effort choice by the manager, and it fully insures him against income risk. When effort is not observable, however, these two goals often come into conflict because the only way to get the manager to work hard is to relate his pay to the realization of profits, which is random. When these goals come into conflict, the nonobservability of effort leads to inefficiencies. To highlight this point, we first study the case in which the manager is risk neutral. We show that in this case, where the risk-bearing concern is absent, the owner can still achievc the same outcome as when effort is observable. We then study the optimal contract when the manager is risk averse. In this case, whenever the first-best (full ohscrvability) contract would involve the high-efTort level, eflicient risk bearing and ellieient incentive provision come into conflict, and the presence of nonobservable actions leads to a welfare loss.
14.B:
HIDDEN
ACTIONS
(MORAL
The manager is willing to accept this contract as long as it gives him an expected utility of at least ii, that is, as long as
f
rrf(rr Ie') drr - ex - g(e') ;;:: ii.
(14.B.8)
Let (I.' be the level of ex at which (14.B.8) holds with equality. Note that the owner's payoff if the compensation scheme is w(rr) = rr - IX' is exactly 0:' (the manager gets all of" except for the fixed payment (I.'). Rearranging (14.B.8), we see that 0:' = rrf(rr Ie') drr - g(e') - U. Hence, with compensation scheme "~rr) = rr - (I.', both the owner and the manager get exactly the same payoff as when effort is observable. _
J
The basic idea behind Proposition 14.B.2 is straightforward. If the manager is risk neutral, the problem of risk sharing disappears. Efficient incentives can be provided without incurring any risk-bearing losses by having the manager receive the full marginal returns from his effort.
A ri ....{-II('lIt,.al HllIIUlfWr
A risk-averse manager
Surpose that 1'(\\') = \\'. Applying Proposition 14.B.I, the optimal efTort level (" when elTort is observable solves Max ('eh',.(',d
f
l£f(l£ I e) dl£ - y(e) - U.
When the manager is strictly risk averse over income lotteries, matters become more complicated. Now incentives for high efTort can be provided only at the cost of having the manager face risk. To characterize the optimal contract in these circumstances, we again consider the contract design problem in two steps: first, we characterize the optimal incentive scheme for each efTort level that the owner might want the manager to select; second, we consider which efTort level the owner should induce. The optimal incentive scheme for implementing a specific effort level e minimizes the owner's expected wage payment subject to two constraints. As before, the manager must receive an expected utility of at least II if he is to accept the contract. When the manager's effort is unobservable, however, the owner also faces a second constraint: The manager must actually desire to choose effort e when facing the incentive scheme. Formally, the optimal incentive scheme for implementing e must therefore solve
(14.B.6)
The owner's profit in this case is the value of expression (l4.B.6), and the manager receives an expected utility of exactly II. Now consider the owner's payoff when the manager's efTort is not observable. In Proposition 14.B.2, we establish that the owner can still achieve his full-information payofT. Proposition 14.B.2: In the principal-agent model with unobservable managerial effort and a risk-neutral manager, an optimal contract generates the same effort choice and expected utilities for the manager and the owner as when effort is observable. Proof: We show explicitly that there is a contract the owner can ofTer that gives him the same payoff that he receives under full information. This contract must therefore be an optimal contract for the owner because the owner can never do better when effort is not observable than when it is (when efTort is observable, the owner is always free to offer the optimal nonobservability contract and simply leave the choice of an effort level up to the manager). Suppose that the owner offers a compensation schedule of the form w(l£) = 1£ - (I., where 1 is some constant. This compensation schedule can be interpreted as "selling the project to the manager" because it gives the manager the full return rr except for the fixed payment ~ (the "sales price"). If the manager accepts this contract, he chooses e to maximize his expected utility,
f
SECTION
I\'(rr) I(rr I e)
f
rrf(rr I e) drr - (I. - y(e).
Min W{lt)
f
W (l£)f(rr
S.t. (i)
f
, e)drr
(14.B.9)
o(w(rr» f(" Ie) d" - g(e) ;;:: ii
(ii) e solves M,ax
f
v(w(rr)) f(rr I i') drr - g(e).
Constraint (ii) is known as the illcelltil'e cOllstraillt: it insures that under compensation scheme w(1[) the manager's optimal effort choice is e. How does the owner optimally implement each of the two possible levels of e? We consider each in turn.
(14.B.7)
Implemelltillg eL : Suppose, first, that the owner wishes to implement effort level
Comparing (14.B.7) with (14.B.6), we see that e' maximizes (14.B.7). Thus, this contract induces the first-best (full observability) effort level eO.
In this case, the owner optimally offers the manager the fixed wage payment + y(ed), the same payment he would offer if contractually specifying effort c,. when effort is observable. To see this, note that with this compensation el.'
w:
J
= V-'(II
HAZARD)
483
484
CHAPTER
14:
THE
PRINCIPAL·AGEHT
PROBLEM
SECTION
14.B:
HIDDEN
ACTIONS
(MORAL
scheme the manager selects eL: His wage payment is unaffected by his effort, and so he will choose the effort level that involves the lowest disutility, namely eL • Doing so, he earns exactly ii. Hence, this contract implements eL at exactly the same cost as when effort is observable. But, as we noted in the proof of Proposition 14.B.2, the owner can never do better when effort is unobservable than when effort is observable [formally, in problem (14.B.9), the owner faces the additional constraint (ii) relative to problem (14.8.2)]; therefore, this must be a solution to problem (14.B.9).
Lemma 14.8.1 tells us that both constraints in problem (14.B.9) bind when e = el/ 8 Moreover, given Lemma 14.B.I, condition (14.B.l0) can be used to derive some useful insights into the shape of the optimal compensation schedule. Consider. for example, the fixed wage payment wsuch that (I/v'(w)) = y. According to condition (14.B.10),
/mplememinlj el/: The more interesting case arises when the owner decides to induce effort level ell' In this case, constraint (ii) of (14.B.9) can be written as
and
(iil/)
f
v(l\'(n») f(n I el/) dn - lI(e,,);::
f
v(l\'(n») f(n I ed dn - y(ed·
+ i'r'(I\'(n» f(n I ell) + J1[f(n I el/) - f(n led] v'(I\'(n))
= 0
(14.B.10)
We first establish that in any solution to problem (14.B.9), where e = el/' both i' and J1 are strictly positive. Lemma 14.B.1: In any solution to problem (14.8.9) with e = eH , both y > 0 and J1 > O. Proof: Suppose that i' = O. Because F(n Iel/) first-order stochastically dominates F(n I eLl. there must exist an open set of profit levels Ii c [,!, if] such that [f(n!eL)/f(n!el/)] > I at alinE Ii. But ify = O,condition (14.B.IO) then implies that "'(II'(n» :$ 0 at any such n (recall that JJ :<: 0), which is impossible. Hence, 7 > O. On the other hand, if I' = 0 in the solution to problem (14.B.9) then, by condition (14.B.10), the optimal compensation schedule gives a fixed wage payment for every profit realization. But we know that this would lead the manager to choose eL rather than ell' violating constraint (iil/) of problem (14.B.9). Hence, J1 > O. • 7. Although problem (l4.B.9) may not appear to be a convcx programming problem. a simple
transformation of the problem shows lhat (14.B.1O) is both a necessary and a sufficient condition for a solution. To see this. reformulate (I4.B.9) as a problem of choosing the manager's level of utility for each profit outcome n. say ii(n). Lelling 4>(') = .. -1(.), the objective function becomes
reformulated problem (see Section M.K of the Mathematical Appendix). The first-order condition
for this problem is
"~(wen»~ = v(n).
[(nle..) < I f(nlel/)
'v
if
f(nle,J --~~-~ > I. I(nlel/)
X. A more direct argument for constmint (i) being binding goes as follows: Suppose that w(n) is a solution to (14.B.9) in which constraint (i) is not binding. Consider a change in the compensation function that lowers the wage paid at each level of 7t in such a way that the resulting decrease in utility is equal at all n. that is, to a new function "~'In) with [v(w(n)) - v(,,'(n))] = t.v > 0 at all 7r E [~. ill This change does not affect the satisfaction of the incentive constraint (ii H ) since if the manager wa~ willing to pick ell when faced with w(n), he will do so when faced with \v(n). Furthermore, because constraint (i) is not binding, the manager will still accept this new contract if j,r is small enough. Lastly. the owner's expected wage payments will be lower than under w(n). This yields a contradiction.
which is convex in ii(n). and the constraints are then all linear in ti(n).
Thus. (Kuhn-Tucker) first·order conditions are both necessary and sufficient for a maximum of this
Defining wen) by
if
This relationship is fairly intuitive. The optimal compensation scheme pays more than IV for outcomes that are statistically relatively more likely to occur under ell than under (',. in the sense of having a likelihood ratio [f(nl e,J/f(n I el/)] less than I. Similarly, it ofTers less compensation for outcomes that are relatively more likely when "" is chosen. We should stress, however, that while this condition evokes a statistical interpretation, there is no actual statistical inference going on here; the owner kllows whal level of effort will be chosen given the compensation schedule he ofTers. Rather, the compensation package has this form because of its illcenril'e effects. That is, by structuring compensation in this way, it provides the manager with an incentive for choosing el/ instead of "t. This point leads to what may at first seem a somewhat surprising implication: in an optimal incentive scheme, compensation is not necessarily monotonically increasing in profits. As is clear from examination of condition (14.8.10), for the optimal compensation scheme to be monotonically increasing, it must be that the likelihood ratio [/(nle..)/I(nle,,)] is decreasing in n; that is, as n increases, the likelihood of getting profit levelrr if effort is ell relative to the likelihood if effort is eL must increase. This property, known as the monotolle likelihood ratio property [see Milgrom (1981)], is lIot implied by first-order stochastic dominance. Figures 14.B.l(a) and (b), for example, depict a case in which the distribution of n conditional on ell stochastically dominates the distribution of n conditional on e L but the monotone likelihood ratio property does not hold. In Ihis example, increases in effort serve to convert low profit realizations into intermediate ones but have no effect on the likelihood of very high profit realizations. Condition (14.8.10) tells us that in this case. we should have higher wages at intermediate levels of profit than at very high ones because it is the likelihood of intermediate profit levels that is sensitive to increases in effort. The optimal compensation function for this example is shown in Figure 14.8.I(c).
or
S4>(l'(nli/(n I",,) dn.
w(n) > ,(,
wIn} <
Letting y;:: 0 and J1:<: 0 denote the multipliers on constraints (i) and (iil/)' respectively. I\'(n) must satisfy the following Kuhn-Tucker first-order condition at every n E [,!, n]: 7
- f(n!el/)
HAZARD)
485
-----------------------------------------------------------------------------
---------------------------------------------------------------------------------------
and noting that 4>'(v(w(n))) = tlv'(w(n)). this gives (14.B.IO).
j
486
CHAPTER
U:
THE
PRINCIPAL·AGENT
SEC T ION
PROBLEM
1 •• B:
HID DEN
ACT ION S
(M 0 R A L
H A Z A RD )
487
-------------------------------------------------------------------- ,------------------------------------------------------------------------f
.,,'
F(·I'.)
7:
F('letl \
\
/
'/
.......-1' / no (a)
simply implement 1".' In either case, nonobservability causes a welfare loss to the owner (the manager's expected utility is u in either case)." These observations arc summarized in Proposition 14.B.3.
w(.)
F
I
i:
:
Figure 14.B.1
:
A violation of the monotone likelihoOd ratio property . (a) Densities. (b) Distribution functions. (c) Optimal
: :' '0
n, (b)
"
(e)
PropOSition 14.B.3: In the principal-agent model with unobservable manager effort, a risk-averse manager, and two possible effort choices, the optimal compensation scheme for implementing eH satisfies condition (14.B.10), gives the manager expected utility D, and involves a larger expected wage payment than is required when effort is observable. The optimal compensation scheme for implementing eL involves the same fixed wage payment as if effort were observable. Whenever the optimal effort level with observable effort would be eH , nonobservability causes a welfare loss.
wage scheme.
The fact that nonobservability leads in this model only to dowllward distortions in the manager's effort level is a special feature of the two-effort-Ievel specification. With many possible effort choices, nonobservability may still alter the level of managerial effort induced in an optimal contract from its level under full obscrvability, but the direction of the bias can be upward as well as downward. (Sec Exercise 14.8.4 for an illustration.)
Condition (14.B.1O) also implies that the optimal contract is not likely to take a simple (e.g., linear) form. The optimal shape of w(rr) is a function of the informational content of various profit levels (through the likelihood ratio), and this is unlikely to vary with rr in a simple manner in most problems. Finally, note that given the variability that is optimally introduced into the manager's compensation, the expected value of the manager's wage payment must be strictly greater than his (fixed) wage payment in the observable case, 11";" = r"(II + lI(e,,». Intuitively, because the manager must be assured an ex· pected utility level of ii, the owner must compensate him through a higher average wage payment for any risk he bears. To see this point formally, note that since £[v(lI'(rr)) Ie,,] = u + g(e,,) and v"(-) < 0, Jensen's inequality (see Section M.C of the Mathematical Appendix) tells us that v(£[w(")le,,]) > u + g(e,,). But we know that v(w:,.) = u + g(e,,), and so £[w(rr) Ie,,] > 11':", As a result, nonobservability increases the owner's expected compensation costs of implementing effort
Ima,ginc: that another statistical signal
or effort, say J" is available to the owner in addition
Ihe rcali/alion of profus, and Ihat Ihe joinl densilY of nand J' given e is given by fin, J'I <'I. In this case. Ihe manager's compensation can, in principle, be made to depend on both nand y. When should compensation be made a function of Ihis variable as well? That is, when docs the optimal compensation runction w(n. r) actually depend on J'?
10
To answer this question, suppose that the owner wishes to implement (~". Following along the same: lines as above, we can derive a condition analogous to condition (14.B.IO):
1 r'(I1'(n, y)) =)'
level c".
[f(n,rie/'>]
+ I' I -
I(",)"-Ie ll )
.
(14.0.11)
Consider. Hrst, the case in which y is simply a noisy random variable th~ll is unrelated to e. Then we can write the density f(n,.I'le) as the product of two densities, f,(nl") "nd J;( y): J(n. y I<'I = I, (n Ie) r,( r)· Subslituting into (l4.B.II), Ihe f,(-) terms cancel out, and so the optim," compensation package is independent of )'. The inluilion behind this result is straightforward. Suppose that the owner is initially olTcring a conlract that has wage payments dependent on y. Intuitively, this contract induces a randomness in the manager's wage that is unrelated to e and therefore makes the manager
Given the preceding analysis, which effort level should the owner induce? As before, the owner compares the incremental change in expected profits from the two effort levels [f "f(rr I e,,) d" - J"f(rr I ed drr] with the difference in expected wage payments in the contracts that optimally implement each of them, that is, with the difference in the value of problem (14.B.9) for I' = 1'" compared with C = C L • From the preceding analysis, we know that the wage payment when implementing c /. is exactly the same as when effort is observable, whereas the expected wage payment when the owner implements ell under nonobservability is strictly larger than his payment in the observable case. Thus, in this model, nonobservability raises the cost of implementing I'll and does not change the cost of implementing eL' The implication of this fact is that nonobservability of effort can lead to an inefficiently low level of effort being implemented. When eL would be the optimal effort level if effort were observable, then it still is when effort is nonobservable. I n this case, nonobservability causes no losses. In contrast, when ell would be optimal if effort were observable, then one of two things may happen: it may be optimal to implement ell using an incentive scheme that faces the manager with risk; alternatively, the risk·bearing costs may be high enough that the owner decides that it is beller to
face risk without achieving any beneficial incentive effect. If the owner instead offers, for each rl.!alization of n. the certain payment \\'(71:) stich that ,.(,i·(n))
= E[r( ...(n, y)) I n] =
f,·( . .
(n, y»f,(.I') dr,
9. Notc. howcver, that although nonobservability leads to a welfare loss, the outcome here is a constrained Pareto optimum in the sense introduced in Section 13.B. To see this, note that the owner maximizes his profit subject to giving the manager an expected utility level no less than 17 and subject to constraints deriving from his inability to observe the manager's effort choice. As a restlll, no allocation that Pareto dominates this outcome can be achieved by a central authority who cannot observe the manager's effort choice. For market intervention by such an authority to generate:'1 Pareto improvement, there must be externalities among the contracts signed by different pairs of individuals.
J
88
CHAPTER
1":
THE
PRINCIPAL~AGENT
PROBLEM
SECTION
1".C:
HIDDEN
INFORMATION
(AND
MONOPOLISTIC
SCREENING)
489
------------------------------------------------------- ,-----------------------------------------------------------------then the manager gets exactly the same expected utility under Wen) as under wen, y) for any level of effort he chooses. Thus, the manager's effort choice will be unchanged, and he will still accept the contract. However, because the manager faces less risk, the expected wage payments are lower and the owner is better off (this again follows from Jensen's inequality: for all n, v(E[w(n, Y)ln) > E[v(w(n, y))ln), and so Ii~n) < E[w(n, Y)ln). This point can be pushed further. Note that we can always write f(n, Y I e) = f,(n Ie)f,(y I n, e).
If f, (y I n, e) does not depend on
f, then the f,(·) terms in condition (14.B.II) again cancel out and the optimal compensation package does not depend on y. This condition on f,(y I n, e) is equivalent to the statistical concept that n is a sufficient statistic for y with respect to e. The
converse is also true: As long as 1t is not a sufficient statistic ror y, then wages should be made to depend on y, at least to some degree. See Holmstrom (1979) for further details.
A number of extensions of this basic analysis have been studied in the literature. For example, Holmstrom (1982), Nalebuff and Stiglitz (1983), and Green and Stokey (1983) examine cases in which many managers are being hired and consider the use of relative performance evaluation in such sellings; Bernheim and Whinston (1986), on the other hand, extend the model in the other direction, examining sellings in which a single agent is hired simultaneously by several principals; Dye (1986) considers cases in which effort may be observed through costly monitoring; Rogerson (1985a), Allen (1985), and Fudenberg, Holmstrom, and Milgrom (1990) examine situations in which the agency relationship is repeated over many periods, with a particular focus on the extent to which long-term contracts are more effective at resolving agency problems than is a sequence of short-term contracts of the type we analyzed in this section. (This list of extensions is hardly exhaustive.) Many of these analyses focus on the case in which effort is single-dimensional; Holmstrom and M ilgrom (1991) discuss some interesting aspects of the more realistic case of multidimensional effort. Holmstrom and Milgrom (1987) have pursued another interesting extension. Bothered by the simplicity of real-world compensation schemes relative to the optimal contracts derived in models like the one we have studied here, they investigate a model in which profits accrue incrementally over time and the manager is able to adjust his effort during the course of the project in response to early profit realizations. They idenlify conditions under which the owner can restrict himself without loss to the use of compensation schemes that are linear functions of the project's total profit. The optimality of linear compensation schemes arises because of the need to offer incentives that are "robust" in the sense that they continue to provide incentives regardless of how early profit realizations turn out. Their analysis illustrates a more general idea, namely, that complicating the nature of the incentive problem can actually lead to simpler forms for optimal contracts. For illustrations of this point, see Exercises 14.B.5 and 14.B.6. The exercises at the end of the chapter explore some of these extensions.
'4,C Hidden Information (and Monopolistic Screening) I n this section, we shift our focus to a selling in which the postcontractual informational asymmetry takes the form of hidden information.
Once again, an owner wishes to hire a manager to run a one-time project. Now, however, the manager's effort level, denoted bye, is fully observable. What is not observable after the contract is signed is the random realization of the manager's disutility from elTort. For example, the manager may come to find himself well suited to the tasks required at the firm, in which case high elTort has a relatively low disutility associated wilh it, or the opposite may be true. However, only the manager comes to know which case obtains' o Before proceeding, we note that the techniques we develop here can also be applied to models of mOllopolistic screenillg where, in a selling characterized by preuilltractual informational asymmetries, a single uninformed individual olTers a menu of contracts in order to distinguish, or screell, informed agents who have dilTering information at the time of contracting (see Section 13.0 for an analysis of a competitive screening model). We discuss this connection further in small type at the end of this section. To formulate our hidden information principal-agent model, we suppose that effort can be measured by a one-dimensional variable e E [0, (fj). Gross profits (excluding any wage payments to the manager) are a simple deterministic function of eITOrl, n(e), with n(O) = 0, n'(e) > 0, and n"(1') < 0 for all e. The manager is an cxpected utility maximizer whose Bernoulli utility function over wages and effort, 11(11', e, 0), depends on a state of nature II that is realized after the contract is signed and that only the manager observes. We assume that II E R and we focus on a special form of II(W, e, 0) that is widely used in the literature:" u(w, r, 0) = v(w -
gee, 0)).
The function ,,(e,O) measures the disutility of effort in monctary units. We assume that y(O, 1/) ~ 0 for all II and, letting subscripts denote partial derivatives, that ",.(r,O) {>o = 0
for r > 0 for e = 0
q,.,.(e,O)
>0
for all e
y.(r, O)
<0
for all e
!I,.,,(r, O) {
for e > 0 for e = o.
Thus, the manager is aversc to increases in effort, and this aversion is larger the greater the current level of eITorl. In addition, higher values of 0 are more productive states in the sense that both the manager's total disutility from eITort, g(e,O), and his marginal disutility from elTort at any current effort level, g,.(e, 0), are lower when 0
10. A ~ccnllngly more important source or hidden information between managers and owners is thaI the manager of a firm orten comes to know more about the potential profitability or various actions than docs the owner. In Section 14.D, we discuss one hybrid hidden action~hidden inform
490
C HAP T E R
1 .. :
THE
P R INC I PAL· AGE N T
PRO B L E M
is greater. We also assume that the manager is strictly risk averse, with v"(-) < 0." As in Section 14.B, the manager's reservation utility level, the level of expected utility he must receive if he is to accept the owner's contract offer, is denoted by U. Note that our assumptions about gee, 8) imply that the manager's indifference curves have the single-crossing property discussed in Section I3.C Finally, for expositional purposes, we focus on the simple case in which 0 can take only one of two values, 0" and 01.' with 0" > 01. and Prob (0,,) = J. E (0, I). (Exercise 14.CI asks you to consider the case of an arbitrary finite number of states.) A contract must try to accomplish two objectives here: first, as in Section 14.B, the risk-neutral owner should insure the manager against fluctuations in his income: second, although there is no problem here in insuring that the manager puts in elTort (because the contract can explicitly state the elTort level required), a contract that maximizes the surplus available in the relationship (and hence, the owner's payoff) must make the level of managerial effort responsive to the disutility incurred by the manager, that is, to the state O. To fix ideas, we first illustrate how these goals are accomplished when is observable: we then turn to an analysis of the problems that arise when II is observed only by the manager.
--- --
1 ... C:
HID 0 E N
IN FOR MAT ION
(A NOM 0 N 0 POL 1ST I
These conditions indicate how the two objectives of insuring the manager and making elTort scnsitive to the state are handled. First, rearranging and combining conditions (14.C2) and (14.C3), we sec that
rr'(en = (I,.(e1, 0,)
for i = L, H.
(14.C7)
This condition says that the optimal level of elTort in state 01 equates the marginal benefit of elTort in terms of increased profit with its marginal disutility cost. The pair (11'1. is illustrated in Figure 14.CI (note that the wage is depicted on the vertical axis and the elTort level on the horizontal axis). As shown, the manager is beller off as we move to the northwest (higher wages and less elTort), and the owner is beller 01T as we move toward the southeast. Because the manager receives utility level II in state 0" the owner seeks to find the most profitable point on the manager's state 0, indilTerence curve with utility level U. This is a point of tangency between the manager's indilTerence curve and one of the owner's isoprofit curves. At this point, the marginal benefit to additional effort in terms of increased profit is exactly equal to the marginal cost borne by the manager. The owner's profit level in state 0, is n1 = rr(e1) - 1'-'(11) - l/(e1, 0,). As shown in Figure 14.C1. this profit is exactly equal to the distance from the origin to the point at which the owner's isoprofit curve through point (\\'1, hits the vertical
en
°
(14.CI)
Max w/ •• '/2!O
>\·/I.""~O
i. v(w" - g(e", 0,,»
+ (I
- i.)v(w l •
-
g(e l., Otl);:> II.
en
In any solution [(11';, en. (wr" er,)] to problem (14.CI) the reservation utility constraint must bind; otherwise, the owner could lower the level of wages olTered and still have the manager accept the contract. In addition, letting y ;:> denote the multiplier on this constraint, the solution must satisfy the following first-order conditions: (14.C2) -i. + }·i.v'(wr, - g(er" 0,,)) = o.
°
. '( ell.)
I.n
-
. '(. yJ.V W II -
491
°°
If () is observable, a contract can directly specify the elTort level and remuneration of the manager contingent on each realization of (note that these variables fully determine the economic outcomes for the two parties). Thus, a complete information contract consists of two wage-elTort pairs: (w", ell) E JR x JR+ for state 0" and (11"/., erJ E JR x JR+ for state 01.' The owner optimally chooses these pairs to solve the following problem:
+ y( I
R E E N I N G)
so the manager's marginal utility of income is equalized across states. This is the usual condition for a risk-neutral party optimally insuring a risk-averse individual. Condition (14.C6) implies that ",r, - y(er" 0Il) = wt - y(et, 01,), which in turn implies that 1'(II'r, - y(er" 011)) = 1'(11'; - y(et, 01.)); that is, the manager's utility is equalized across states. Given the rescrvation utility constraint in (14.CI), the managcr therefore has utility level Ii in each state. Now consider the optimal elTort Icvels in the two states. Since 1/,.(0, 0) = and rr'(O) > 0, conditions (14.C4) and (14.CS) must hold with equality and e1 > for i = 1,2. Combining condition (14.C2) with (14.C4), and condition (14.C3) with (14.CS), we see that the optimallevcl of elTort in state 0" e1, satisfies
TIle State (] is Observable
-( I - i.)
esc
(14.C6)
°
s.t.
SEC T ION
- i.) D'(W! -
gee!, 0tl) = 0 .
~ 0, 9 (* ell. 0)) II ge (* ell' 0II ){ = 0
(I - i.)rr'(erJ - y(1 - i.)v'(II'! - gee!, 0tl)g,(e!,
OIJ{ ~~'
Figure 14.C.1
The optimal wage-effort pair for state OJ when states are observable. n(e) - w
(14.C3)
if
eil > 0.
if et > O.
(14.CA)
(14.CS)
Profits of { 0 Owner in Slalc 0,
m:)
12. As with the case of hidden actions studied in Section 14.B, nonobservabililY causes no welfare loss In the case of managerial risk neutralily. As there. a "sellout" contract that races the manager with the full marginal returns from his actions can generate the first-best outcome. (See
Exercise 14.C.2.)
J
= n7':
492
CHAPTER
14:
THE
PRINCIPAL·AGENT
v(.' - y(e, Od)
PROBLEM
="
SECTION
v( .. - y(r,
0.» = U
.(e) - w =
.(
II
nr
w = n~
.,'
•
e'
•
HIDDEN
INFORMATION
(AND
MONOPOLISTIC
choice in each state to the manager's discretion. Alternatively, the owner could olTer a compensation schedule w(n) but restrict the possible effort choices by the manager to some degree. Another possibility is that the owner could olTer compensation as a function of the observable effort level chosen by the manager, possibly again with some restriction on the allowable choices. Finally, more complicated arrangements might be imagined. For example, the manager might be required to make an announcement about what the state is and then be free to choose his elTort level while facing a compensation function w(n I 0) that depends on his announcement 0. Although finding an optimal contract from among all these possibilities may seem a daunting task, an important result known as the revelation principle greatly simplifies the analysis of these types of contracting problems:') Proposition 14.C.2: (The Revelation Principle) Denote the set of possible states by 0. In searching for an optimal contract, the owner can without loss restrict himself to contracts of the following form:
Flgur. 14.C.2
The optimal contraci with full observability of O.
(i) After the state 0 is realized, the manager is required to announce which state has occurred. (ii) The contract specifies an outcome [w(ti), e(Ii)] for each possible announcement liE 0. (iii) In every state Ii E 0, the manager finds it optimal to report the state truthfully.
axis [since n(O) = 0, if the wage payment at this point on the vertical axis is IV < 0, the owner's profit at (11',*, is exactly -IV). From condition (14.C7), we see that g..,(e, 0) < 0, n"(e) < 0, and g,,(e, 0) > imply that e1, > e!- Figure 14.C2 depicts the optimal contract, [(w1" e1,), (wt, These observations are summarized in Proposition 14.CI.
en
14.C:
SCREENING)
493
------------
1
°
em.
i\ contract that asks the manager to announce the state aand associates outcomes with the various possible announcements is known as a revelation mechanism, The revelation principle tells us that the owner can restrict himself to using a revelation mechanism for which the manager always responds truthfully; revelation mechanisms with this truthfulness property arc known as incentive compatible (or trlllhfll/) revelation mechanisms. The revelation principle holds in an extremely wide array of incentive problems. Although we defer its formal (and very general) proof to Chapter 23 (sec Sections 23.C and 23.D), its basic idea is relatively straightforward. For example, imagine that the owner is offering a contract with a compensation schedule w(n) that leaves the choice of effort up to the manager. Let the resulting levels of elTort in states and 0" be eL and ell' respectively. We can now show that there is a truthful revelation mechanism that generates exactly the same outcome as this contract. In particular, suppose that the owner uses a revelation mechanism that assigns outcome [w(n(eLl), ed if the manager announces that the state is OL and outcome [1I'(n(e,,)), 1',,] if the manager announces that the state is 0". Consider the manager's incentives for truth telling when facing this revelation mechanism, Suppose, first, that the state is II,.. Under the initial contract with compensation schedule w(n), the manager could have achieved outcome [1I'(n(e,,», ell] in state OL by choosing elTort level <'", Since he instead chose eL , it must be that in state 0,. outcome [lI"(n(c,.l), e,.l is at least as good for the manager as outcome [I\'(n(e,,», 1',,). Thus, under the proposed revelation mechanism, the manager will find telling the truth to A similar argument applies for state 0". be an optimal response when the state is We see therefore that this revelation mechanism results in truthful announcements
Proposition 14.C.1: In the principal-agent model with an observable state variable 0, the optimal contract involves an elfort level e7 in state Ojsuch that n'(en = g.(e7, OJ) and fully insures the manager, setting his wage in each state OJ at the level w7 such that v(w7 - g(e7, OJ)) = D. Thus. with a strictly risk-averse manager, the first-best contract is characterized by two basic features: first, the owner fully insures the manager against risk; second, he requires the manager to work to the point at which the marginal benefit of effort exactly equals its marginal cost. Because the marginal cost of effort is lower in state 0" than in state ilL' the contract calls for more effort in state Ou.
a,.
Tile State 0 is Observed Ollly by the Mallager As in Section 14.B, the desire both to insure the risk-averse manager and to elicit the proper levels of effort come into conflict when informational asymmetries arc present. Suppose, for example, that the owner offers a risk-averse manager the contract depicted in Figure 14.C2 and relies on the manager to reveal the state voluntarily. If so, the owner will run into problems. As is evident in the figure, in state 0", the manager prefers point (\\'t, eV to point (\\'1" e1,). Consequently, in state he will lie to the owner, claiming that it is actually state 0L' As is also evident in the figure, this misrepresentation lowers the owner's profit. Given this problem, what is the optimal contract for the owner to offer? To answer this question, it is necessary to start by identifying the set of possible contracts that the owner can offer. One can imagine many different forms that a contract could conceivably take. For example, the owner might olTer a compensation function w(n) that pays the manager as a function of realized profit and that leaves the effort
a"
a,.,
11 Two early discussions of the revelation principle are Myerson (1979) and Oasgupta,
/lammond. and Maskin (t979) .
.J
494
C HAP T E R
1.:
THE
P R INC I PAL· AGE N T
PRO B L E M
~~~~~~~~~~~-----------------------------------by the manager and yields exactly the same outcome as the initial contract. In fact, a similar argument can be constructed for any initial contract (see Chapter ~3), and so the owner can restrict his attention without loss to truthful revelation
--
mechanisms." To simplify the characterization of the optimal contract, we restrict attention from this point on to a specific and extreme case of managerial risk aversion: infinite risk aversion. In particular, we take the expected utility of the manager to equal the manager's lowest utility level across the two states. Thus, for the manager to accept the owner's contract, it must be that the manager receives a utility of at least ii In each state." As above, efficient risk sharing requires that an infinitely risk-averse manager have a utility level equal to ii in each state. If, for example, his utility is ii in one state and 1/' > 'i.in the other, then the owner's expected wage payment is larger
"'/,f'/,
I N FOR MAT ION
(A N D
M 0 N 0 POL 1ST I
. (or indw,dual rattonallCY) (ii) 11'" - g(el/' Ol/);?: v"(ii) constraint incentive compatibility (iii) 11'" - g(ell' Oil);?: w L - g(eL' Oil)} (or truth-telling
Lemma 14.C.1 is iIlustraled in Figure 14.C.3. By constraint (i), (IV L , cd must lie in the shaded region of the figure. But by constraint (iii), (w", ell) must lie on or above the state 0" indifference curve through point (IV L , ed. As can be seen, this implies that the manager's state 0" utility is at least ii, the utility he gets at point (II', e) = (1"'(,i), 0). Therefore, from this point on we can ignore constraint (ii).
2:: 0
s.t. (i) 11", - g(e, , Od ;?: V"(ii)} .'
(iv)
W
L
-
g(eL'
OLl
;?:
reserva~i~n utility
11'11 - g(e", 0L)
or self-~e1ection)
Lemma 14.C.2: An optimal contract in problem (14.C.8) must have w L
constraints.
,.,. 1'(11' - g(e.O,,)) =
V(WL -
OLl
=
Ii
Figure 14.C.3 Constraint (ii) in problem (I4.C.S) is satisfied by any constraints (i) and (iii).
..
g(e ll , 011)),
g(e L •
contract satisfying
15. This can be thought of as the limiting case in which. starting from the concave utility function l'{x), we take the concave transformation v,(v) = -v(x)P for p < 0 as th~, manager's Bernoulli utility runction and let p -+ -C(). To see this, note that the manager's expected ,utIlity over, t.he random oulcome giving (WH - g(e", On)) with probability i. and (w" - g(e", 0,,)) w,th probab,lIly (I _ i.) is then EU = -[i,v~ + (I - i.)v[), where v, = v(w, - gte"~ 0,)) for i = L, H. This expected utility is correctly ordered by (-EU)'/P = [i.v~ + (1 - i.)V£]l/p. Now as p - -00, [Avr, + (I -:- A)VrJI/PMin {VII' v,J (see Exercise lC.6). Hence, a contract gives the manager an expected uttllty greate: -
-
Proof: Suppose not. that is. that there is an optimal solution [(WL' ed, ("'II' e,,)] in which \I'" - I/(e,., Ii,.l > ,,' '(Ii). Now, consider an alteration to the owner's contract
14, One restriction that we have imposed here for expositional purposes is to limit the outcomes specified following the manager's announcement to being nonstochastic (in fact~ much of th.e literature does so as well), Randomization can sometimes be desirable in these settings because It can aid in satisfying the incentive compatibility constraints that we introduce in problem (l4.C.8).
Ihan his (cerlain) reservation utililY if and only if Min {v(w II
495
v' ,(0).
The pairs (1\'", ell) and (WL' eLl that the contract specifies are now the wage and effort levels that result from different announcements of the state by the manager; that is, the outcome if the manager announces that the state is 0, is (11'" e,). Constraints (i) and (ii) make up the reservation utility (or individual rationality) constraint for the infinitely risk-averse manager; if he is to accept the contract, he must be guaranteed a utility of at least ii in each state. Hence, we must have V(IV, - g(e" 0;) ;?: ii for i = L, H or, equivalently, WI - g(e l , 0,) ;?: v'l(ii) for i = L, H. Constraints (iii) and (iv) are the incentive compatibility (or truth-telling or self-selection) constraints for the manager in states 01/ and OL' respectively. Consider, for example, constraint (iii). The
Sec Maskin and Riley (1984a) for an example.
esc R E E N I N G )
manager's utility in state 011 is v(w" - g(el/' 011» if he tells the truth, but it is v(wl. - g(e,., 011» if he instead claims that it is state 0L' Thus, he will tell the truth if W ll - g(e" , 0,,) ;?: W L - g(eL' 011)' Constraint (iv) follows similarly. Note that the first-best (full observability) contract depicted in Figure 14.C.2 does not satisfy the constraints of problem (14.C.8) because it violates constraint (iii). We analyze problem (l4.C.8) through a sequence of lemmas. Our arguments for these results make extensive use of graphical analysis to build intuition. An analysis of this problem using Kuhn-Tucker conditions is presented in Appendix B.
(14.C.8)
;.[rr(e,,) - 11',,] + (I - ;.)[rr(ed- wL]
Max
HID 0 E N
Proof: Whenever both constraints (i) and (iii) are satisfied, it must be that \I'" - 1/(<,,,,0,,) ;?: w,. - {/(eL, 0,,) ;?: w L - Y(<'", Od ;?: v' '(ii), and so constraint (ii) is also satisfied. This implies that the set of feasible contracts in the problem derived from (14.C.8) by dropping constraint (ii) is exactly the same as the set of feasible contracts in problem (14.C.8). •
allows us to write the owner's problem as follows:
2:: 0,
1", C:
Lemma 14.C.l: We can ignore constraint (ii). That is, a contract is a solution to problem (14.C.8) if and only if it is the solution to the problem derived from (14.C.8) by dropping constraint (ii).
than necessary for giving the manager an expected utility of Ii. Given this assumption about managerial risk preferences, the revelation principle
"'//.('/1
SEC T ION
g(e,., O,,))} 2: u.
J
496
C HAP T E R
"W -
g(e,
Vd)
1 4:
THE
P R INC I PAL. AGE N T
PRO B L E M
'"
SEC T ION
v(w - gte,OL))
(wH,e H)
="
"'
./ \ "-.\
1 4 • C:
HID 0 E N
I N FOR MAT ION
(A NOM 0 HOP 0 LIS TIC
S eRE E N I N G )
497
,---------------------------------------------~~~~~~ = Ii
___ v(w - gte, 0,,))
"..
= v(Ii'L -
g(e/" 0,,))
Isoprofit Curves
State eH Indifference Curve Through (w" It!
Isoprofit Curves
('~
,,' "
iL
State 0Il
Profit {
in which the owner pays wages in the two states of IV,. = W L - c and IV II = IVII - c, where £ > 0 (i,e" the owner lowers the wage payments in both states by c), This new contract still satisfies constr'!int (i) as long as c is chosen small enough, In addition, the incentive compatibility constraints are still satisfied because this change just subtracts a constant, c, from each side of these constraints, But if this new contract satisfies all the constraints, the original contract could not have been optimal because the owner now has higher profits, which is a contradiction. _ Lemma 14.C.3: In any optimal contract: (i) e L :5 et; that is, the manager's effort level in state OL is no more than the level that would arise if 0 were observable, (ii) eH = e;,; that is, the manager's effort level in state OH is exactly equal to the level that would arise if 0 were observable,
Stale { ()
"
/,
I;~olil
and, as is evident in Figure I 4.C.S, the truth-telling constraints are still satisfied, Thus, > e;' cannot be optimal. a contract with Now consider part (ii), Given any wage -elTort pair (Ii'/" "/,) with ,"- :5 such as that shown in Figure 14,C.6. the owner's problem is to find the location for (1\'11'"11) in the shaded region that maximizes his profit in state 11 11 , The solution occurs at a point of tangency between the manager's state 1111 indifTerence curve through point (Ii',.. ell) and an isoprofit curve for the owner. This tangency occurs at point (lVII' cr,) in the figure, and necessarily involves efTort level er, because all points of tangency between the manager's state 0" indifTerence curves and the owner's isoprofit curves occur at efTort level er, [they are characterized by condition (14.C7) for i = If]. Note that this point of tangency occurs strictly to the right of efTort level e,. because
"I.
Ftgur. 14.C.4 (teft)
I n a feasible contract alTering (wL' eel for state OL' the pair ("'H,eH) must lie in the shaded region, Figure 14.C,5 (right)
An optimal contract has eL. :s; ei.
el. :s ei.
<
et..
e'1i. -
A secondary point emerging from the proof of Lemma 14,C3 is that only the truth-telling constraint for state 0" is binding in the optimal contract. This property is common to many of the other applications in the literature'·
Proof: Lemma 14,C3 can best be seen graphically, By Lemma 14,C2. (11'", cd lies on the locus {(II', c): r(1I' - y(e, 0L» = Ii} in any optimal contract. Figure 14,C4 depicts one possible pair (IVL , eL).ln addition, the truth-telling constraints imply that the outcome for state 01/, (11'", el/), must lie in the shaded region of Figure 14,C4, To see this. note that by constraint (iv), (lVII' ell) must lie on or below the state 01. indilTerence curve through (IV,., ed. In addition, by constraint (iii), (lVII' ell) must lie on or above the state 011 indifference curve through (WL' ell. To see part (i), suppose that we have a contract with t?L > Figure I4,CS depicts such a contract offer: (li'L' ell lies on the manager's state 0,. indifference curve with utility level Ii, and (\I'll' ell) lies in the shaded region defined by the truth-telling constraints, The state OL indifference curve for the manager and the isoprofit curve for the owner which go through point (IV L, e,,) have the relation depicted at point (WL' e,,) because eL > As can be seen in the figure, the owner can raise his profit level in state OL by moving the state 0L wage-effort pair down the manager's indifference curve from (Ii'", it,,) to its first-best point ("'t, This change continues to satisfy all the constraints in problem (I 4,C8): The manager's utility in each state is unchanged,
Lemma 14.C.4: In any optimal contract, e L < ei; that is, the effort level in state ilL is necessarily strictly below the level that would arise in state ilL if 0 were observable, Proof: Again, this point can be seen graphically, Suppose we start with (w L , cd = as in Figure 14.C7. By Lemma 14.C3, this determines the state 011 outcome, denoted by (Ii'll' ettl in the figure. Note that by the definition of (w;', the isoprofit curve through this point is tangent to the manager's state OL indifference curve. Recall that the absolute distance between the origin and the point where each state's isoprofit curve hits the vertical axis represents the profit the owner earns in that state, The owner's overall expected profit with this contract ofTer is therefore
et.
(1Vt..
et-
en.
en.
en.
16. In models with more than two lypes, this properly takes the form thai only the incentive constraints between adjacent types bind, and they do so only in one direction. (See Exercise t4.C.1.)
J
("
"
Flgur. 14,C,6 (left)
An optimal contract h~ls ('II = e;'. Figur. 14,C.7 (right)
The best contract wilh (!L
=
It!-
498
CHAPTER
14:
THE
PRINCIPAL.AGENT
v(w - gte, 8,,))
PROBLEM
="" > "
5 E C T ION
Siale 0" Indifferenee
Isoprofil Curves
1 4 . C:
HID 0 E N
I N FOR MAT ION
011 [note that once we move away from (wi,
(A NOM 0 N 0 POL 1ST I
en.
the envelope result no longer holds and the marginal reduction in state OL'S profit is strictly positive). It should not be surprising that the extent to which the owner wants to make this trade-off depends on the relative probabilities of the two states. In particular, the greater the likelihood of state Oil' the more the owner is willing to distort the state 0, outcome to increase profit in statc 0/1' In the extreme case in which the probability of state 0, gets close to zero, the owner may set "I. = 0 and hire the manager to work only in state Oil' I7 The analysis in Appendix B confirms this intuition. There we show that the optimal levcl of satisfies the following first-order condition:
Curves ~
"I.
[n'h) - O,(e l., aLl]
"
1-
(a)
Profil Gain in Siale 0" -
(b)
Figure 14.C.B (a) The change in profils in slale 0,. [rom lowering eL slighlly below et. (b) The change in profits in state 0 [rom lowering ',. slightly below
).
1=;' [o,.(e l., (/11) -
o,.(el., 0,.)] =
o.
(14.C.9)
The first term of this expression is zero at ", = ei and is strictly positive at e,. < et; the second term is always strictly negative. Thus, we must have e, < ei to satisfy this condition, confirming our finding in Lemma 14.C.4. DifTerentiating this expression reveals that the optimal level of e, falls as i./( I - ;.) rises. These findings are summarized in Proposition 14.C.3.
V'
Profil Loss in Siale 0, L
+
et and optimally adjusling
"'/I'
equal to the average of these two profit levels (with weights equal to the relative probabilities of the two states). We now argue that a change in the state 0, outcome that lowers this state's effort level to one slightly below necessarily raises the owner's expected profit. To see this, start by moving the state 01• outcome to a slightly lower point, (w" ell, on the manager's state 0, indifference curve. This change is illustrated in Figure 14.C.8, along with the owner's isoprofit curve through this new point. As is evident in Figure 14.C.8(a), this change lowers the profit that the owner earns in state (h. However, it also relaxes the incentive constraint on the state all outcome and, by doing so, it allows the owner to offer a lower wage in that state. Figure 14.C.8(b) shows the new state all outcome, say (WII' eft), and the new (higher-profit) isoprofit curve through this point. Overall, this change results in a lower profit for the owner in state 0, and a higher profit for the owner in state Oil' Note, however, that because we started at a point of tangency at (\\'t, the profit loss in state a, is small relative to the gain in state Oil' Indeed, if we were to look at the derivative of the owner's profit in state 0, with respect to an il1fillitesimal change in that state's outcome, we would find that it is zero. In contrast, the derivative of profit in state 011 with respect to this infinitesimal is an envelope change would be strictly positive. The zero derivative in state theorem result: because we started out at the first-best level of effort in state (J" a small change in (11'" eLl that keeps the manager's state 0, utility at Ii has no first-order effect on the owner's profit in that state; but because it relaxes the state (JII incentive constraint, for a small-enough change the owner's expected profit is increased. _
et
en.
a,
How far should the owner go in lowering e? In answering this question, the owner must weigh the marginal loss in profit in state 01. against the marginal gain in state
H
Proposition 14.C.3: In the hidden information principal-agent model with an infinitely risk-averse manager the optimal contract sets the level of effort in state OH at its first-best (full observability) level e~. The effort level in state Ot is distorted downward from its first-best level In addition, the manager is inefficiently insured, receiving a utility greater than 0 in state OH and a utility equal to 0 in state ()t. The owner's expected payoff is strictly lower than the expected payoff he receives when is observable, while the infinitely risk-averse manager's expected utility is the same as when 0 is observable (it equals 0).18.19
et.
°
A basic, and very general, point that emerges from this analysis is that the optimal for the owner in this setting of hidden information necessarily distorts the efTort choice of the manager in order to ameliorate the costs of asymmetric information, which here take the form of the higher expected wage payment that the owner makes because the manager has a utility in state 011 in excess of ii. Note that nothing would change if the profit level n were not publicly observable (and so could not be contracted on), since our analysis relied only on the fact that the efTort level e was observable. Moreover, in the case in which 11 is not publicly observable, we can extend the model to allow the relationship between profits and efTort to depend on the state; that is, the owner's profits in states 0, and 011 given efTort level e might be given by the functions n,.(e) and lIH(e).20 As long as contra~t
17. In [acl, Ihis can happen only i[ ~,(O. 0,) > O. I X. Recall that an infinitely risk-averse manager's expected utility is equal to his lowest utility level across the two states. 19. Note. however. that while the outcome here is Pareto inefficient. it is a constrained Pareto
optimum in the sense introduced in Section 13.8; the reasons paraliel those given in footnote that the authority cannol observe rather than ('). 20. The nonobservability of profits is important for this extension because if n could be contracted upon, the manager could be punished for misrepresenting the state by simply comparing the realized profit level with the profit level that should have been realized in the announced state for the specified level of effort.
9 of Seclion 14.R for Ihe hidden aClion model (although here il is
°
esc R E E N I N G )
499
500
CHAPTER
14:
THE
PRINCIPAL-AGENT
PROBLEM
5
n;,(e) ;?: n~(e) > 0 for all e;?: 0, the analysis of this model follows exactly along the lines of the analysis we have just conducted (see Exercise 14.C.5). As in the case of hidden action models, a number of extensions of this basic hidden information model have been explored in the literature. Some of the most general treatments appear in the context of the "mechanism design" literature associated with social choice theory. A discussion of these models can be found in Chapter 23.
T ION
1 4 . 0:
HID 0
eN
ACT ION SAN 0
HID 0 E N
I N FOR MAT ION:
H Y 8 RID
and seeks to offer a menu of (x;, 7;) pairs to maximize its profit. The monopolist's problem then takes the form in (14.C.IO) where we take I, = X;, IV; = - 7;, Ii = O. g(I;. 0;) = - v(x;, 0;), and 11;(1;) = -ex;. Baron and Myerson's (1982) analysis of optimal regulation of a monopolist with unknown costs provides another example. There, a regulated firm faces market demand function x(p) and has unobservable unit costs of O. The regulator, who seeks to design a regulatory policy that maximizes consumer surplus, faces the monopolist with a choice among a set of pairs (p;, 7;), where p; is the allowed retail price and 7; is a transfer payment from the regulator to the firm. The regulated firm is able to shut down if it cannot earn profits of at least zero from any of the regulator'S offerings. The regulator'S problem then corresponds to (14.C.10) with I; = p;, \\'; = 7;, II = 0, 1/(1;,0;) = -(p; - O;)x(p;), and n;(I;) = X(5) ds."
The MOllopolistic Screellillg Model
J;:
In Section 13.D, we studied a model of competitive suet!lIi/lg in which firms try to design their employment contracts in a manner that distinguishes among workers who, at the time of contracting. have different unobservable productivity levels (i.e., there is precontractual asym· metric information). The techniques that we have developed in our study of the principal-agent model with hidden information enable us to formulate and solve a model of monopolislic s(''''''lIiIl9 in which. in contrast with the analysis in Section 13.0. only a single firm offers employment contracts (actually, this might more properly be called a monopsonislic screening model because the single firm is on the demand side of the market). To sec this, suppose that, as in the model in Section 13.0, there are two possible types of workers who differ in their productivity. A worker of type a has utility u(II', II 0) = II' - g(I, 0) when he receives a wage of 1\' and faces task level I. His reservation utility level is Ii. The productivities of the two types of workers are a" and OL' with a" > OL > O. The fraction of workers of type 0" is i. E (0, I). We assume that the firm's profits. which are not publicly observable, are given by the function 111/(1) for a type 0" worker and by nL(I) for a type OL worker. and that 11;,(1) ~ n,,(I) > 0 for all I ~ 0 [e.g .• as in Exercise 13.0.1. we could have n;(I) = 0;( I - Ill) for I' > 0]." The firm's problem is to offer a set of contracts that maximizes its profits given worker self-selection among, and behavior within. its offered contracts. Once again. the revelation principle can be invoked to greatly simplify the firm's problem. Here the firm can restrict its attention to offering a menu of wage-task pairs [(11'1/,1,,). (IVL,I L )] to solve (14.C.IO)
Max
(ii)
11'" -
ad
(iii)
11'" -
g(l". 0,,)
s.t. (i) WL - g(IL'
(iv) WL - g(IL'
ad
~ WI/ - g(I",
ad·
This problem has exactly the same structure as (14.C.8) but with the principal's (here the firm's) profit being a function of the state. As noted above, the analysis of this problem follows exactly the same lines as our analysis of problem (14.C.8). This class of models has seen wide application in the literature (although often with a continuum of types assumed). Maskin and Riley (1984b), for example, apply this model to the study of monopolistic price discrimination. In their model, a consumer of type 0 has utility t~x. 0) - T when he consumes x units of a monopolist'S good and makes a total payment of T to the monopolist, and can earn a reservation utility level of V(0. 0) = 0 by not purchasing from the monopolist. The monopolist has a constant unit cost of production equal to c > 0 21. The model studied in Section 13.D with ltj(t) = 0, corresponds to the limiting case where
I' -0.
eC
MOD E L 5
501
,---------------------------------------------------------------------
Exercises 14.C.7 to 14.C.9 ask you to study some examples of monopolistic screening models.
14,D Hidden Actions and Hidden Information: Hybrid Models Although the hidden action - hidden information dichotomization serves as a useful starting point for understanding principal-agent models, many real-world situations (and some of the literature as well) involve elements of both problems. To consider an example of such a model, suppose that we augment the simple hidden information model considered in Section 14.C in the following manner: let the level of efTort e now be unobservable, and let profits be a stochastic function of efTort, described by conditional density function f(n I e). I n essence, what we now have is a hidden action model, but one in which the owner also docs not know something about the disutility of the manager (which is captured in the state variable 0). Formal analysis of this model is beyond the scope of this chapter. but the basic thrust of the revelation principle extends to the analysis of these types of hybrid problems. In particular, as Myerson (1982) shows, the owner can now restrict attention to contracts of the following form: (i) After the state (/ is realized, the manager announces which state has occurred. (ii) The contract specifies, for each possible announcement 0 E e, the efTort level dO) that the manager should take and a compensation scheme 1\'(11 I 0). (iii) In every state II, the manager is willing to be both Irulirful in stage (i) and obedient following stage (ii) [i.e., he finds it optimal to choose effort level e(O) in state OJ, This contract can be thought of as a revelation game, but one in which the outcome of the manager's announcement about the state is a hidden action-style contract, that is, a compensation scheme and a "recommended action." The requirement of "obedience" amounts to an incentive constraint that is like that in the hidden action
22. The regulator's objective function can be generalized to allow a weighted average of consumer and producer surplus, with greater weight on consumers. In this case, the runction 7r i (·) will depend on 0;.
502
APPENDIX
A:
MULTIPLE
EFFORT
LEVELS
IN
THE
HIDDEN
ACTION
MODEL
503
-------------------------------------------------------------------------- ,---------------------------------------------------------------CHAPTER
14:
THE
PRINCIPAL.AGENT
PROBLEM
model considered in Section 14.B; the "truthfulness" constraints are generalizations of those considered in our hidden information model. See Myerson (1982) for details. One special case of this hybrid model deserves particular mention because its analysis reduces to that of the pure hidden information model considered in Section 14.C. In particular, suppose that effort is unobservable but that the relationship between effort and profits is determillistic, given by the function n(e). In that case, for any particular announcement 8, it is possible to induce any wage-effort pair that is desired, say (w, e), by use of a simple "forcing" compensation scheme: Just reward the manager with a wage payment of w if profits are n(e), and give him a wage payment of - OCJ otherwise. Thus, the combination of the observability of n and the one-to-one relationship between nand e effectively allows the contract to specify e. The analysis of this model is therefore identical to that of the hidden information model considered in Section 14.C, where wage-effort pairs could be specified directly as functions of the manager's announcement. To see this point in a slightly different way, note first that because of the ability to write forcing contracts, in this model an optimal contract can be thought of as specifying, for each announcement 6, a wage-profit pair (w(O), n(O)). Now, for any required profit level n, the effort level necessary to achieve a profit of n is e such that nee) = n. Let the function e(n) describe this etTort level. We can now think of the manager as having a disutility function defined directly over the profit level which is given by y(n, 0) = g(e(n), II). But this model looks just like a model with observable etTort where the effort variable is n, the disutility function over this etTort is y(n, II), and the profit function is n(7I) = 71. Thus, the analysis of this model is identical to that in a pure hidden information model. A similar point applies to a closely related hybrid model in which, instead of the manager's disutility of effort, it is the relation between profit and effort that depends on the state. In particular, suppose that the disutility of effort is given by the function g(e) and profits are given by the function 7I(e, Ii), where 71.(') > 0, 71 .. (') < 0, 71,(') > 0, and n.. (·) > 0. Effort is not observable, but profits are. The idea is that the manager knows more than the owner does about the true profit opportunities facing the firm (e.g., the marginal productivity of effort). Again, we can think of a contract as specifying, for each announcement by the manager, a wage-profit pair (implicitly using forcing contracts). In this context, the effort needed to achieve any given level of profit 71 in state II is given by some function 1'(71, Ii), and the disutility associated with this effort is then g(n, 0) = g(';(7I, II». But this model is also equivalent to our basic hidden information model with observable etTort: just let the etTort variable be n, the disutility of this etTort be g(n, 0), and the profit function be n(n) = 71. Again, our results from Section 14.C apply.
Figure 14.AA.1
Density functions for E = {eL , eM' ell}: effort choice eAt may not be
implementable.
Profit Realization
to the more general specification initially introduced in Section 14.B in which E is the feasible set of effort choices. As in Section 14.B, we can break up the principal's (the owner's) problem into several parts: (a) What are the effort levels e that it is possible to induce'! (b) What is the optimal contract for inducing each specific etTort level e E E'! (c) Which etTort level e E E is optimal'! In a multiple-action setting, each of these three parts becomes somewhat more complicated. For example, with just two actions, part (a) was trivial: eL could be induced with a fixed wage contract, and ell could always be induced by giving incentives that were sufficiently high at outcomes that were more likely to arise when ell is chosen. With more than two actions, however, this may not be so. For example, consider the three-action case in which E = {e L, eM' ell} and the conditional density functions are those depicted in Figure 14.AA. I. As is suggested by the figure, it may be impossible to design incentives such that eM is chosen because for any w(n) the agent may prefer either eL or ell to eM' (Exercise 14.B.4 provides an example along these lines.) Part (b) also becomes more involved. The optimal contract for implementing effort choice e solves Min w(lt)
f
(14.AA.I)
\I'(n)f(nle)dn
s.t. (i)
f
v(w(n)) f(n I e) dn - g(e)
(ii) e solves
~.~~
f
~ ii
v(w(n»I(n\e) dn - g(e).
Ifwe have K possible actions in set E, the incentive constraints in problem (14.AA.I) [constraints (ii)] consist of (K - I) constraints that must be satisfied. In this case, with a change of variables in which we maximize over the level of utility that the manager gets conditional on n, say v(7I), we have a problem with K linear constraints and a convex objective function [see Grossman and Hart (1983) and footnote 7 for more on this]. However, if E is a continuous set of possible actions, say E = [0, e] c IR, then we have an infinity of incentive constraints. One trick sometimes used in this case to
APPENDIX A: MULTIPLE EFFORT LEVELS IN THE HIDDEN ACTION MODEL
In this appendix, we discuss additional issues that arise when the effort choice in the hidden action (moral hazard) model discussed in Section 14.B is more complex than the simple two-effort-choice specification e E {e L, ell} analyzed there. Here, we return
d
504
CHAPTER
1.:
PRINCIPAl.AGENT
THE
PROBLEM
----------------------------------------------------------------------------------simplify problem (14.AA.l) is to replace constraint (ii) with a ./irst-order condition (this is sometimes called the ./irst·order approach). For example, if e is a one. dimensional measure of effort, then the manager's first-order condition is
f
~:P:P:E:N~O~':X~B~'~S~O~"~U~T~'~O~N__~O_'
,--
Using Lemma 14.C.1 we can restate problem (14.C.8) as Max ..... //.1'/1
v(w(n» !e(n I e) dn - g'(e) = 0,
OBLEM WITH HIDDEN INFORMATION ___ 505 __T_H_E__P_R_'_H_C_'_P_A_l_._A_G_E_N__T __P_A________________________________
~
).[n(el/) - wl/]
(14.AA.2)
1
• I.
+ I'
[!.(n1e)] f(n Ie)
The condition that ratio Ue(n I <,)!f(n I e)] be increasing in n is the differential version of the monotone likelihood ratio property (see Exercise 14.AA.l). In general, however, a solution to the problem resulting from this substitution is not necessarily a solution to the actual problem (14.AA.I). The reason is that the agent may satisfy first-order condition (14.AA.2) even when effort level e is not his optimal effort choice. First, effort level e could be a minimum rather than a maximum; therefore, we at least want the agent to also be satisfying a local second-order condition. But even this will not be sufficient. In general, we need to be sure that the agent's objective function is concave in e. Note that this is not a simple matter because the concavity of his objective function in e will depend both on the shape of f(n I e) and on the shape of the incentive contract w(n) that is offered. The known conditions which insure that this condition is met are very restrictive. See Grossman and Hart (1983) and Rogerson (1985b) for details. Exercise 14.AA.2 provides a very simple example.
~
O.
'1'.'/.('/;;::
g(e ll , 01/) ~
W, -
gte"~
- i.)[n(e,J -
g(e l .,
0,.1
IV, -
g(e" 01/)
(iv)
IV, -
gte"~ 0,) ~
lVI/ -
g(el/' OLl.
+ 4>11
+ 'l -
>1/
= O.
(14.IlB.2)
+ >, = o.
(14.BB.3)
- >1.
(14.1313.4)
(I 4. 1lB.5)
Step 4: Steps 1 to 3 imply that 4>, = O. Suppose not: i.e., that >,. > O. Then constraint (iv) must be binding. We shall now derive a contradiction. First, substitute for 4'" in conditions (14.BB.4) and (14.BB.5) using the fact that 4>11 = 4>,. + i.from condition (14.BB.2). Then, using the fact that (e,., ell)>> O. we can write cond,t,ons (14.1l1l.4) and (14.BB.5) as
+ 4>1.[y,.(e ll . 0,.)
-II,.(eIl' 0Il)] = 0
and
+ (I + >d[g,(e/.. 011) -
g,(e" 0,,)] =
o.
But ',. > 0 then implies that
(ii) IV" - y(e", 01/) ~ V-I(ol) g(el/' 01/) ~
Od·
Step J: Iloth 1'1. and el/ are strictly positive. To see this, note that condition (14.BB.4) cannot hold at ell = 0 because n'(O) > 0 and Ye(O. 0,) = a for i = L. H. Similarly for condition (14.B8.5) and e,.
~ v·'(ii)
lVI/ -
g(el/,
Step 2: Adding conditions (14.BB.2) and (14.BB.3) implies that i' = I. Hence. constmint (i) must bind at an optimal solution.
w,J
(iii)
w, - gte"~ 0Il) "'II -
Step I: Condition (14.BB.2) implies that >11 > O. Thus, constraint (iii) must bind (hold with equality) at an optimal solution.
(I - i.)[n'(e,.) - 1I,(e/.. 011)] WI. -
(14.B8.I)
along with the complementary slackness conditions for constraints (i), (iii). and (iv) [conditions (M.K.7)]. Let us break up the analysis of these conditions into several steps.
0
s.t. (i)
~
if 1', > 0
i.[rr'(ell) - r/,(ell. 0Il)]
+ (I
OLl
if 1'" > 0
Recall problem (14.C.8): ""/1. ell
WI/ -
(iv)
-(I - i.)
APPENDIX B: A FORMAL SOLUTION OF THE PRINCIPAL-AGENT PROBLEM WITH HIDDEN INFORMATION
i.[n(el/) - w,,]
~ V-'(ol)
(iii)
-i.
Finally, to answer part (c), we need to compute the optimal contract from part (b) for each action that part (a) reveals is implementable and then compare their relative profits for the principal. With more than two effort choices, two features of the two-effort-choice case fail to generalize. First, nonobservability can lead to an upward distortion in effort. (Exercise 14.B.4 provides an example.) Second, at the optimal contract under nonobservability we can get boch an inefficient effort choice and inefficiencies resulting from managerial risk bearing.
Max
OLl
IV,]
Letting (i', 4>", 4>d ~ 0 be the multipliers on constraints (i), (iii), and (iv), respectively, the Kuhn- Tucker conditions for this problem can be written (see Section M.K of the Mathematical Appendix)
(14.AA.3)
.
).)[n(ed -
s.t. (i) w, - g(e"
where f.(n I e) = af(n I e)IDe. If we replace constraint (ii) with (14.AA.2) and solve the resulting problem, we can derive a condition for w(n) that parallels condition (14.8.10):
v'(w(;i) =
+ (I -
0, ""',.t'/ ~ 0
rr'(ed - y,.(e" 011) >
a > n'(ell ) -
g,(e ll , 011),
which implies 1'1/ > e, since n(e) - gte, 011) is concave in e. But if el/ > e, and constraint (iii) binds (which it does from Step I), then constraint (iv) must be slack
....
506
CHAPTER
14;
THE
PRINCIPAL-AGENT
PROBLEM
-------------------------------------------------------------------------------------------because we then have
l
EX E R CIS E S
~----------
Maskin. E.. and J. Riley. (l984b). Monopoly with incomplete information. Rand Journal of Economics IS:
("'" - wel =
= <
t7t-96. Milgrom. P. (1981). Good news and bad news: Representation theorems and applications. Bell Journal of £nltlomics 12: 380-91. Myerson, R. (1979). Incentive compatibility and the bargaining problem. Econometrica 47: 61-74 . Myerson. R. (1982). Optimal coordination mechanisms in generalized principal-agent problems. Jou,.,lul of Mur"emalicu/ ECOtlOmics 10: 67-tH. N.alcbufT. B.. and J. E. Stiglitz. (1983). Prizes and incentives: Towards a general theory of compensation and competition. Bdl Jour",,1 oj E('Onomj('~'I3: 21-41 Rog.erson. W. (1985a). Repeated moral hazard. Econamf!lric'a 53: 69-76. Rogerson. W. (1985h). The first·order approach to principal·agent problems. Econometrica 53: 1357-68.
g(e", fI,,) - g(e L , fI,,)
f.'," f"'"
g.(e, fI,,) de
y,(e,
Oel de
This is our desired contradiction: Slep 5: Since 4>,. = 0, we know from (14.BB.2) that 4>11 two values into conditions (14.88.4) and (14.B8.5) we have
= J..
Substituting these
1['(e,,) - Y,(e", 0,,) = 0
(14.B8.6)
EXERCISES
(14.B8.7)
14.B.I" Consider the two-elTort-level hidden aelion model discussed in Section 14.B with the gene«11 utility function u(w, e) for Ihe agent Must the reservalion utility constraint be binding
and
[1['(e,.) - y •. (e L• 0,.>]
i.
+ l·~-;: [g.(eL' 0,,) -
y.(e L • 0,.>] =
o.
in an optim
Conditions (14.88.6) and (14.B8.7) characterize the optimal values of ell and e,., respectively. The optimal values for "',. and are then determined from constraints (i) and (iii). which we have seen hold with equality at the solution.
w"
14.B.2" Derive the forst-order condition characlerizing Ihe optimal compensation scheme for Ihe IWll-clTorl·!cvel hidden aClion model studied in Section 14.B when the principal is striclly risk avc-rsl!.
An alternative approach to solving problem (14.BB.I) that avoids this somewhat cumbersome argument involves the following "trick"; Solve problem (l4.BB.I) ignoring constraint (iv). Then show that the solution derived in this way also satisfies constraint (iv). If so. this must be a solution to the (more constrained) problem (14.B8.1). (Exercise 14.88.1 asks you to try this approach.)
REFERENCES
14.1l.3" Consider a hidden action model in which the owner is risk neutral white the manager has preferences defoned over the mean and the variance of his income wand his elTort level e as follows: Expected utility = E[\\'] -O. and Lim, _ " g'(e) = vo. Possible elTort choices are. E R +. Conditional on elTori level (', the realization of profit is normally distributed with mean e and variance a'.
(a) Reslricl attenlion to linear compensation schemes w(x) = manager's cxpecled utility given w(n). e, and a' is given by ~ + (Ie -
+ (In. Show that the
~
(b) Derive the optimal conlroct when e is observable. Allen, F. (l985). Repeated principal-agent relationships with lending and borrowing. Economic utters 17: 27-31. Baron. D .. and R. Myerson. (1982). Regulating a monopolise with unknown costs. Econometrica SO: 911-30. Bernheim. B. D .• and M. D. Whinston. (1986). Common agency. Econometrica 54: 923-42.
D.sgupla, P.. P. Hammond, and E. Maskin. (1979). The impiemenlation of sociat choice rutes: Some results on incentive compatibility. Revifw of Economic S,udies 46: 18S-216. Dye. R. (1986). Optimal monitoring policies in agencies. Rand Journal of Economics 17: 339-50. Fudenbcrg. D .. B. Holmstrom. and P. Milgrom. (1990). Short-term contracts and long-term agency relationships. Journal of Economic nJftJr)' 52: 194-206. Green. J .• and N. Stokey. (1983). A comparison of tournaments and contests. JOUrlW/ of Political £('OIIOIIIY
9t: 349-64. Grossm.m. S. J .• and O. D. Hart. (1983). An amllysis of the principal-agent problem. Ecollometrka 51: 7-45. H;.m. O. D .. and B. Holmstrom. (1987). The theory of contracts. In Advances in Economic Thear}'. Fifth World Congress, edited by T. Bewley. New York: Cambridge University Press. Holmstrom. B. (1979). Moral hazard and observability. Bell Journal of Economks 10: 74-91. Holmstrom. B. (1982). Moral hazard in teams. Bell Journal of Economics 13: 324-40. Holmstrom, B.. and P. Milgrom. (1987). Aggregation and linearity in the provision of intertemporal inCCllIives. f"Ollometricll 55: 303-28. Holmstrom. B.. and P. Milgrom. (1991). Multitask principal-agcnt analyses: Incentive contracts, asset ownership, and job design. Journal (if Law, Economics. and Organizations 7: 24-52. Maslin, E.. and J. Riley. (1984a). Optimal auctions with risk averse buyers. Econometrica 52: 1473-1518.
(e) Derive the optimal linear compensalion scheme when e is not observable. What elTects do changes in II and a' have'! 14.1l.4" Consider the following hidden action model with three possible actions E = let. e,. e,}. There are two possible profit outcomes: nil = 10 and nL = O. The prObabilities of XI/ conditional on the three elTort levels are [(x,.Ie,) = i. [(XII Ie,) =!. and [(nil I',) =~. The
agent's elTort cosl function has y(e,) = \. y(e,) = ~, g(e,) =~. Finally. v(w) = manager's reservation utility is U ::::: O.
Jw. and
the
(a) What is the optimal conlract when elTori is observable? (b) Show Ihal ifelTort is not observable, then e, is nol implementable. For what levels of U(e,) would e, be implemenlable'! [Hilll: Focus on the utility levels the manager will get for Ihc two outcomes, v, and rather than on the wage payments themselves.]
1',.
(e) What is the oplimat contracl when elTori is not observable?
fl,
and let [(x"le,) = x E (0.1). What is the optimal (d) Suppose, instead, that y(e,) = contracl if elTort is observable as x approaches I? What is the optimal contract as x approaches t if il is not observable·! As x approaches I, is the level of elTort implemented higher or lower when elforl is nol observable than when it is observable?
507
508
CHAPTER
,.:
THE
PRINCIPAL·AGENT
PROBLEM
EXERCtSES
14.C.3B Suppose that in the two-state hidden information model examined in Section 14.C,
14.B.S" Suppose that in the hidden action model explored in Section 14.B the manager can not only choose how much effort to exert but can also, after observing the realization of the firm's profits It, unobservably reduce them in a way that is of no direct benefit to him (e.g., he can voluntarily offer to pay more for his inputs). Show that in this case there is always an optimal incentive scheme that is nondecreasing in observed profits.
u(w,
e, 0)
= v(w) -
gee, 0).
(a) Characterize the optimal contract under full observability. (b) Is this contract feasible when the state 0 is not observable?
14.C.4C Characterize the solution to the two-state principal-agent model with hidden information when the manager is risk averse, but not infinitely so.
14.B.6" Amend the two-effort-level model studied in Section 14.B as follows: Suppose now that effort has distinct effects on revenues R and costs C, where It = R - C. Let I.(R I e) and /r.·(C I e) denote the density functions of Rand C conditional on e, and assume that, conditional on e, Rand C are independently distributed. Assume R E [R, R), C E [C, C), and that for all e,J.(R Ie) > 0 for all R E [~, R) and IdCle) > 0 for all C ~ [c:', C). The two effort choices are now {e., e(.}. where e. is an effort choice that devotes more
14.C.S" Confirm that the analysis in Section 14.C could not change if the owner's profits depended on the state and were not publicly observable and if. letting It;(e) denote the profits in state 0; for i = L, H. It;'(e) ~ It~(e) > 0 for all e ~ O. What happens if this condition is not satisfied?
time to revenue enhancement and less to cost reduction, and the opposite is true for C!c. In
particular, assume that F.(R Ie.) < F.(R I ecJ for all R E (~, R) and that Fe(Cle,,) > FdCle.) for all C E (c:" C). Moreover, assume that the monotone likelihood ratio property holds for each of these variables in the following form: U.(RIe.)lf.(RlecJ) is increasing in R, and (J;.(CJe.)/k(Clecl) is increasing in C. Finally, the manager prefers revenue enhancement over cost reduction: that is, g(e,,) > uk.).
14.C.6C Reconsider the labor market screening model in Exercise 13.0.1, but now suppose that there is a single employer. Characterize the solution to this firm's screening problem (assume that both types of workers have a reservation utility level of 0). Compare the task levels in this solution with those in the equilibrium of the competitive screening model (assuming an equilibrium exists) that you derived in Exercise 13.0.1.
(a) Suppose that the owner wants to implement effort choice e" and that both Rand C are observable. Derive the first-order condition for the optimal compensation scheme I\'(R. C). How does it depend on Rand C?
14.C.7" (1. Tirole) Assume that there are two types of consumers for a firm's product. OH and II,.. The proportion of type 0,. consumers is i.. A type Us utility when consuming amount x of the good and paying a total of T for it is u(x, T) = Ov(x) - T. where
(b) How would your answer to (a) change if the manager could always unobservably reduce the revenues of the firm (in a way that is of no direct benefit to him)?
1_(1 -x)' vex) = ----2---. The firm is the sole producer of this good, and its cost of production per unit is c > O.
(e) What if, in addition, costs are now unobservable by a court (so that compensation can be made contingent only on revenues)?
(a) Consider a nondiscriminating monopolist. Derive his optimal pricing policy. Show that he serves both classes of consumers if either OL or i. is "large enough."
14.B.7C Consider a two-period model that involves two repetitions of the two-effort-level hidden action model studied in Section 14.8. There is no discounting by either the firm or the manager. The manager's expected utility over the two periods is the sum of his two single-period expected utilities £[v(w) - g(e)), where v'(·) > 0 and v"(·) < O. Suppose that a contract can be signed ex ante that gives payoffs in each period as a function of performance up until then. Will period 2 wages depend on period I profits in the optimal contract?
(b) Consider a monopolist who can distinguish the two types (by some characteristic) but can only charge a simple price p; to each type 0,. Characterize his optimal prices.
(e) Suppose the monopolist cannot distinguish the types. Derive the optimal two-part tariff (a pricing policy consisting of a lump-sum charge F plus a linear price per unit purchased of p) under the assumption that the monopolist serves both types. Interpret. When will the monopolist serve both types? (d) Compute the fully optimal nonlinear tariff. How do the quantities purchased by the two types compare with the levels in (a) to (e)?
14.B.SC Amend the two-effort-choice hidden action model discussed in Section 14.B as follows: Suppose the principal can, for a cost of c, observe an extra signal y of the agent's effort. Profits n and the signal )' have a joint distribution fen, y Ie) conditional on e. The decision to investigate the value of )' can be made after observing It. A contract now specifies a wage schedule wen) in the event of no investigation, a wage schedule I1'(It, }') if an investigation occurs, and a probability pen) of investigation conditional
14.C.S" Air Shangri-la is the only airline allowed to fly between the islands of Shangri-la and Nirvana. There are two types of passengers, tourist and business. Business travelers are willing to pay more than tourists. The airline, however, cannot tell directly whether a ticket purchaser is a tourist or a business traveler. The two types do differ, though, in how much they are willing to pay to avoid having to purchase their tickets in advance. (Passengers do not like to commit themselves in advance to traveling at a particular time.) More specifically, the utility levels of each of the two types net of the price of the ticket, P, for any given amount of time W prior to the flight that the ticket is purchased are given by
on n. Characterize the optimal contract for implementing effort level ell'
14.C.I C Analyze the extension of the hidden information model discussed in Section 14.C where there are an arbitrary finite number of states (0" ... , ON) where 0;+, > 0; for all i. 14.C.2" Consider the hidden information model in Section 14.C, but now let the manager be risk neutral with utility function v(w) = IV. Show that the owner can do as well when 0 is unobservable as when it is observable. In particular, show that he can accomplish this with a contract that offers the manager a compensation scheme of the form w(It) = It - a and allows him to choose any effort level he wants. Graph this function and the manager's choices in (w, e)-space. What revelation mechanism would give this same outcome?
BIISilless: Tourisc:
v - O.p - W, v - OfP - W,
where 0 < O. < 0,.. (Note that for any given level of W, the business traveler is willing to pay more for his ticket. Also, the business traveler is willing to pay more for any given reduction in W.)
..
509
510
CHAPTER
t.:
THE
PRINCIPAL.AGENT
PROBLEM
The proporlion of travelers who are tourists is .I.. Assume that the cost of transporting a passenger is e. Assume in (a) to (d) that Air Shangri·la wants to carry both types of passengers.
-
PAR
T
F
0
U
R
General Equilibrium
(0) Draw the indifference curves of the two types in (P, W)-space. Draw the airline's
isoprofit curves. Now formulate the optimal (profit·maximizing) price discrimination problem mathematically that Air Shangri·la would want to solve. [Hinr: Impose nonnegativity of prices as a constraint since, if it charged a negative price, it would sell an infinite number of tickets at this price.] (b) Show that in the optimal solution, tourists are indifferent between buying a ticket and not going at all. (c) Show that in the optimal solution. business travelers never buy their ticket prior to the night and are just indifferent between doing this and buying when tourists buy. (d) Describe fully the optimal price discrimination scheme under the assumption that they sell to both types. How does it depend on the underlying parameters i., 0•• Or, and ,,? (0) Under what circumstances will Air Shangri·la choose to serve only business travelers?
P"rt IV is devoted to an examination of competitive market economies from a
gelleral equilibrium perspective. Our use of the term "general equilibrium" refers both
14.C.9" Consider a risk·averse individual who is an expected utility maximizer with a Bernoulli utility function over wealth u(·). The individual has initial wealth Wand faces a probability IJ of suffering a loss of size L, where W> L > O. An insurance contract may be described by a pair (c" e,), where e, is the amOunt of wealth the individual has in the event of no loss and c, is the amount the individual has if a loss is suffered. That is, in the event no loss occurs the individual pays the insurance company an amount (W - e,), whereas if a loss occurs the individual receives a payment [e, - (W - L)] from the company.
to a methodological point of view and to a substantive theory. Methodologically, the general equilibrium approach has two central features. First, it views the economy as a closed and interrelated system in which we must simultaneously determine the equilibrium values of all variables of interest. Thus, when we evaluate the effects of a perturbation in the economic environment, the equilibrium levels of the entire set of endogenous variables in the economy needs to be recomputed. This stands in contrast to the partial equilibrium approach, where the impact on endogenous variables not directly related to the problem at hand is explicitly or implicitly disregarded. A second central feature of the general equilibrium approach is that it aims at reducing the set of variables taken as exogenous to a small number of physical realities (e.g., the set of economic agents, the available technologies, the preferences and physical endowments of goods of various agents). From a substantive viewpoint, general equilibrium theory has a more specific meaning: It is a theory or the determination of equilibrium prices and quantities in a system of perfectly competitive markets. This theory is often referred to as the Walrasian theory of markets [from L. Walras (1874)], and it is the object of our study in Part IV. The Walrasian theory of markets is very ambitious. It attempts no less than to predict the complete vector of final consumptions and productions using only the fundamentals of the economy (the list of commodities, the state of technology, prererences and endowments), the institutional assumption that a price is quoted for every commodity (including those that will not be traded at equilibrium), and the behavioral assumption of price taking by consumers and firms. Strictly speaking, we introduced a particular case of the general equilibrium model in Chapter 10. There, we carried out an equilibrium and welfare analysis of perfectly competitive markets under the assumption that consumers had quasilinear preferences. In that setting, consumer demand functions do not display wealth effects (except for a single commodity, called the numeraire); as a consequence, the analysis of a single market (or small group of markets) could be pursued in a manner understandable as traditional partial equilibrium analysis. A good deal of what we do in Part IV
(0) Suppose that the individual's only source of insurance is a risk·neutral monopolist (i.e., the monopolist seeks to maximize its expected profits). Characterize the contract the monopolist will offer the individual in the case in which the individual's probability of loss, IJ. is observable.
(b) Suppose, instead, that 0 is not observable by the insurance company (the individual knows 0). The parameter 0 can take one of two values {OL,OH}' where 0" > OL > 0 and Prob (Od = i.. Characterize the optimal contract offers of the monopolist. Can one speak of one type of insured individual being "rationed" in his purchases of insurance (i.e., he would want to purchase more insurance if allowed to at fair odds)? Intuitively. why does this rationing occur? [Him: It might be helpful to draw a picture in (c" e,)·space. To do so. start by locating the individual's endowment point. that is, what he gets ifhe does not purchase any insurance.] (c) Compare your solution in (b) with your answer to Exercise I3.D.2. 14.AA.I" Show that [f,(lt Ie)ll(lt Ie)] is increasing in It for all e E [a, b] c R if and only if for any e', e" E [a. b]. with e" > e', [f(ltle")II(ltle')] is increasing in It.
"L.
l4.AA.2" Consider a hidden action model with e E [0, i] and two outcomes It,, and with "L' The probability of ltH given effort level e is I(lt" Ie). Give sufficient conditions for the first·order approach to be valid. Characterize the optimal contract when these conditions arc satisfied.
"If >
l4.B8.I" Try solving problem (14.B8.I) by first solving it while ignoring constraint (iv) and then arguing that the solution you derive to this "relaxed" problem is actually the solution to problem (14.B8.I).
511
...
512
PART
IV:
GENERAL
PART
EQUILIBRIUM
IV:
GENERAL
EQUILIBRIUM
513
--------------------------------------------------------------------------- --------------------------------------------------------------------------------can be viewed as an attempt to extend the ideas of Chapter 10 to a world in which wealth effects are significant. The primary motivation for this is the increase in realism it brings. To make practical use of equilibrium analysis for studying the performance of an entire economy, or for evaluating policy interventions that affect large numbers of markets simultaneously, wealth effects, a primary source of linkages across markets, cannot be neglected, and therefore the general equilibrium approach is essential. Although knowledge of the material discussed in Chapter 10 is not a strict prerequisite for Part IV, we nonetheless strongly recommend that you study it, especially Sections 10.B to 10.0. It constitutes an introduction to the main issues and provides a simple and analytically very useful example. We will see in the different chapters of Part IV that quite a number of the important results established in Chaptcr 10 for the quasilinear situation carryover to the case of general preferences. But many others do not. To understand why this may be so, recall from Chapters 4 and 10 that a group of consumers with quasilinear preferences (with respect to the same numeraire) admits the existence of a (normative) representative consumer. This is a powerful restriction on the behavior of aggregate demand that will not be available to us in the more general settings that we study here. It is important to note that, relative to the analysis carried out in Part III, we incur a cost for accomplishing the task that general equilibrium sets itself to do: the assumptions of price-taking behavior and universal price quoting-that is, the existence of markets for every relevant commodity (with the implication of symmetric information)-are present in nearly all the theory studied in Part IV. Thus, in many respects, we are not going as deep as we did in Part III in the microanalysis of markets. of market failure, and of the strategic interdependence of market actors. The trade-off in conceptual structure between Parts III and IV reflects, in a sense, the current state of the frontier of microeconomic research. The content of Part IV is organized into six chapters. Chapter 15 presents a preliminary discussion. Its main purpose is to illustrate the issues that concern general equilibrium theory by means of three simple examples: the tlVo-consumer Edgeworth box economy; the one-consumer, one-firm economy, and the sm
properties derived from the mass nature of markets. We study the important core equivalence theorem, examine further the idea of Walrasian equilibria as the limit of noncooperative equilibria as markets grow large (a subject already broached in Section 12.F), and present two normative characterizations of Walrasian equilibria: one in terms of envy-freeness (or anonymity) and the other in terms of a marginal productivity principle. Appendix A of Chapter 18 offers a brief introduction to the cooperative theory of games. Chapter 19 covers the modeling of uncertainty in a general equilibrium context. The ability to do this in a theoretically satisfying way has been one of the success stories of general equilibrium theory. The concepts of contingent commodities, Arrow-Debreu equilinrium, sequential trade (in a two-period setting), Radner equilibrium, arnitrage, ratiollal expectations equilibrium, and incomplete markets are all introduced and studied here. The chapter provides a natural link to the modern theory of finance. Chapter 20 considers the application of the general theory to dynamic competitive economies (but with no uncertainty) and also studies a number of issues specific to this environment. Notions such as impatience, dynamic efficiency, and myopic versus overall utilicy maximi:acion are introduced. The chapter first analyzes dynamic representative consumer economies (including the Ramsey-Solow model), then generalizes to the case of a finite number of consumers, and concludes with a brief presentation of the operlapping gelleracions model. In the process, we explore a wide range of dynamic behaviors. The chapter provides a natural link to macroeconomic theory. The modern classics of general equilibrium theory are Debreu (1959) and Arrow and Hahn (1971). These texts provide further discussion of topics treated here. For extensions, we recommend the encyclopedic coverage of Arrow and Intriligator (1981, 1982, 1986) and Hildenbrand and Sonnenschein (1991). See also the more recent textbook account of Ellickson (1993). General equilibrium analysis has a very important applied dimension that we do not touch on in this book but that accounts in good part for the importance of the theory. For a review, we recommend Shoven and Whalley (1992).
REFERENCES Arrow. K.. and F. Hahn. (l971). Gtrttra! Competilil'e
Anal}'~j.~.
San Francisco: Holden-Day.
Arrow, K .• and M. Intriligator. cds. (1981). Handbook of Marh('matical Economics. Vol. 1. Amsterdam: NOrlh·Hoiland. Arrow, K., and M. Intriligator. cds. (1982). IIclIIdhook oj Mllthemdticai Eco/Zomio, Vol. 2. Amsterdam: North·Holiand
theorems of welfare economics.
Arrow. K .. and M. IntriligalOf. cds. (1986). /lm"lhoo~ o( /I1(ltIIt'IIId1ical Ecollomics, Vol. J Amsterdanr
In Chapter 17. the emphasis is, instead, on positive (or descripcive) properties of Walrasian equilibria. We study a number of questions pertaining to the predictive power of the Walrasian theory, including the existence, local and global uniqueness, and comparative statics behavior of Walrasian equilibria. Chapters 18 to 20 explore extensions of the basic analysis presented in Chapters 16 and 17. Chapter 18 covers a number of topics whose origins lie in normative theory or the cooperative theory of games; these topics share the feature that they provide a deeper look at the foundations of price-taking equilibria by exploiting
North-Holland. Dcbrcu. G. (1959). T/H?OryoJ Va/Uf', New York: Wiley. Elhckson, B. (1993). Compel/till(' Equilihrillln: TfJeory Clnd Applications. Cambridge, UK: Cambridge University Press. I-lildenbrand, W., and H. Sonnenschein. cds. (1991). lIalldhaok (~r Mafhematical EcolJolIJio, Vol. 4. Amsterdam: North-Holland. Shoven. J .. and J. Whalley. (1992). Applying General Equilihriwn Anal)'si.~. Cambridge, UK: Cambridge University Press.
Walras, L. (1874). Elements J'economie politiqlle p"rt'. Lausanne: Corbaz.
C
General Equilibrium Theory:
HAP
T
E
R
15
Some Examples
lS.A Introduction The purpose of this chapter is to present a preliminary discussion. In it, we describe and analyze thrce simple examples of general equilibrium models. These examples introduce some of the questions, concepts, and common techniques that will occupy us for the rest of Part IV. In most economics, three basic economic activities occur: production, exchange, and consumption. In Section IS.B, we restrict our focus to exchange and consumption. We analyze the case of a pure exchange economy, in which no production is possible and the commodities that are ultimately consumed are those that individuals possess as endowments. Individuals trade these endowments among themselves in the marketplace for mutual advantage. The model we present is the simplest-possible exchange problem: two consumers trading two goods between each other. In this connection, we introduce an extremely handy graphical device, the Edgeworth box. In Section IS.C, we introduce production by studying an economy formed by one firm and one consumer. Using this simple model, we explore how the production and consumption sides of the economy fit together. In Section 15.0, we examine the production side of the economy in greater detail by discussing the allocation of resources among several firms. To analyze this issue in isolation, we study the case of a small open economy that takes the world prices of its outputs as fixed, a central model in international trade literature. Section IS.E illustrates, by means of an example, some of the potential dangers of adopting a partial equilibrium perspective when a general equilibrium approach is called for. As we noted in the introduction of Part IV, Chapter 10 contains another simple example of a general equilibrium model: that of an economy in which consumers have preferences admitting a quasilinear representation.
lS.B Pure Exchange: The Edgeworth Box A pure exchange economy (or, simply, an exchange economy) is an economy in which there are no production opportunities. The economic agents of such an economy are 515
516
C HAP T E R
,.:
G ENE R ALE QUI LI B R I U M
THE Q R V:
SO M E
SECTION
E X AMP L E S
--------------------------------------------------------------------------------------------consumers who possess initial stocks, or endowments, of commodities. Economic activity consists of trading and consumption. The simplest economy with the possibility of profitable exchange is one with two commodities and two consumers. As it turns out, this case is amenable to analysis by a graphical device known as the Edyeworth hox, which we use extensively in this section. Throughout, we assume that the two consumers act as price takers. Although this may not seem reasonable with only two traders, our aim here is to illustrate some of the features of gcneral equilibrium models in the simplest-possible way.' To begin, assume that there are two consumers, denoted by i = 1,2, and two commodities, denoted by ( = 1,2. Consumer i's consumption vector is Xi = (x Ii' X 2i); that is, consumer i's consumption of commodity ( is Xli' We assume that consumer i's consumption set is IR~ and that he has a preference relation ;::i over consumption vectors in this set. Each consumer i is initially endowed with an amount W li 2! 0 of good (. Thus, consumer i's endowment vector is Wi = (WIi' W2i)' The /Owl endowment of good I in the economy is denoted by WI = WI! + WI2; we assume that this quantity is strictly positive for both goods. An al/oelllion x E IR~ in this economy is an assignment of a nonnegative consumption vector to each consumer: x = (x" x,) = «x", x,,), (x", x,,». We say that an allocation is !easihie for the economy if for I = 1,2,
'} 1
-____ jo_ -------:-1 ,: (t)21
--:,~
}
".B:
PURE
EXCHANGE:
EDGEWORTH
BOX
I'
-
fIJ!
I J
The Budgel Line Slope: - p,/p,
B,(p)
~!
0, '----r-'
0,
w"
\.
'1 Flgur. lS.B.l (len)
function of prices: Bi(p) =:Xi E ~'.: p' Xi ~ P''''i}'
The budget sets of the two consumers can be represented in the Edgeworth box in a simple manner. To do so, we draw a line. known as the nlldget line, through the endowment pointw with slope -(PI/P,), as shown in Figure IS.B.2. Consumer l's budget set consists of all the nonnegative vectors below and to the left of this line (the shaded set). Consumer 2's budget set, on the other hand, consists of all the vectors above and to the right of this same line which give consumer 2 nonnegative consumption levels (the hatched set).' Observe that only allocations on the budget line are affordable to both consumers simultaneously at prices (PI' p,)J We can also depict the preferences ;::i of each consumer i in the Edgeworth box, as in Figure IS.B.3. Except where otherwise noted, we assume that ;::i is strictly convex, continuous, and strongly monotone (see Sections 3.B and 3.C for discussion of these conditions). Figure IS.B.4 illustrates how the consumption vector demanded by consumer I can be determined for any price vector p. Given p, the consumer demands his most preferred point in B,(p), which can be expressed using his demand function as x,(p, p'w,) (this is the same demand function studied in Chapters 2 to 4; here wealth is 11', = p·w,). In Figure IS.B.S, we see that as the price vector p varies, the budget line pivots around the endowment point w, and the demanded consumptions trace out a curve, denoted by oe" that is called the ofler curve of consumer I. Note that this curve passes through the endowment point. Because at every p the endowment vector 01, = (W'I'W 2I ) is affordable to consumer I, it follows that this consumer must find every point on his offer curve at least as good as his endowment point.
(IS.B.I)
that is, if the total consumption of each commodity is no more than the economy's aggregate endowment of it (note that in this notion of feasibility, we are implicitly assuming that there is free disposal of commodities). The feasible allocations for which equality holds in (IS.B.I) could be called nonwaste!ul. Nonwasteful feasible allocations can be depicted by means of an Edgeworth box, shown in Figure IS.B.1. In the Edgeworth box, consumer I's quantities are measured in the usual way, with the southwest corner as the origin. I n contrast, consumer 2's quantities are measured using the northeast corner as the origin. For both consumers, the vertical dimension measures quantities of good 2, and the horizontal dimension measures quantities of good I. The length of the box is 01" the economy's total endowment of good I; its height is UJ" the economy's total endowment of good 2. Any point in the box represents a (non wasteful) division of the economy's total endowment between consumers I and 2. For example, Figure IS.B.I depicts the endowment vector W = «w, .. W,,), (WI2' w,,)) of the two consumers. Also depicted is another possible nonwasteful allocation, X = «x I I ' x,,), (x". x,,)); the fact that it is nonwasteful means that (XI" x,,) = (WI - XII' 01, - Xli)' As is characteristic of general equilibrium theory, the wealth of a consumer is not given exogenously. Rather, for any prices p = (p" p,), consumer i's wealth equals the market value of his endowments of commodities, P'W i = PIWli + P,W'i' Wealth levels are therefore determined by the values of prices. Hence, given the consumer's endowment vector Wi' his budget set can be viewed solely as a
2. Note, in particular, that the budget sets of the consumers may well extend outside the box. 3. There are other feasible allocations that are simultaneously affordable; but in these allocations some resources are not consumed by either consumer, and thus they cannot be depicted in an Edgeworth box. Because orthe nonsatiation assumption to be made on preferences, we will not have to worry about such allocations.
1. Alternatively. we could assume that each consumer (perhaps better called a ('onsumer typl!)
stands, not for an individual, but for a large number of identical consumers. This would make the price-taking assumption more plausible: and with equal numbers of the two types of consumers, the analysis in this section would be otherwise unaffected.
d
517
0"
-i- ----------t----I
THE
An Edgeworth box.
Figure 1S.B.2 (right)
Rudget
sclS.
518
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY:
SOME
SECTION
EXAMPLES
Indifference Curves
for
15.8:
PURE
EXCHANGE:
THE
EDGEWORTH
BOX
Good 2
<:,
~~-------------__.O,
--~~----------------------~
~
Direction of Increasing Preference for Consumer 2
Direction of Incn::asing Preference for Consumer t
Consumer I's { Net Demand for Good 2
<:,
Consumer 2's Net } Supply of Good 2
p 0,
0,
(1)11
Ly-J
Indifference Curves for <: I
~
~--------------~~----~~Goodl
sales of commodities at the going market prices. Thus, if one consumer wishes to be a net demander of some good, the other must be a net supplier of this good in exactly the same amount; that is, demand should equal supply. This gives us the notion of equilibrium presented in Definition 15.B.1.
0,
Indifference Curves for <::,
519
--------------------------------------------------------------------------------------
--------------------------------------------------------------------------------------
DefinitIon 15.B.1: A Walrasian (or competitive) equilibrium for an Edgeworth box economy is a price vector p' and an allocation x' = (xT, x;) in the Edgeworth box such that for i = 1, 2, x: <::::; xi for all xi E B ,(p').
Flgur. 15.B.3 (top leH)
Preferences in the Edgeworth box.
A Walrasian equilibrium is depicted in Figure 15.B.7. In Figure 15.B.7(a), we represent the equilibrium price vector p' and the equilibrium allocation x* = (xT, xn. Each consumer i's demanded bundle at price vector p* is and one consumer's net demand for a good is exactly matched by the other's net supply. Figure 15.B.7(b) adds to the depiction the consumers' otTer curves and their indifference curves through w. Note that at any equilibrium, the offer curves of the two consumers intersect. In fact, any intersection of the consumers' offer curves at an allocation different from the endowment point w corresponds to an equilibrium because if x* = (xT, x!) is any such point of intersection, then is the optimal consumption bundle for each consumer i for the budget line that goes through the two points wand x'. In Figure 15.B.8, we show a Walrasian equilibrium where the equilibrium allocation lies on the boundary of the Edgeworth box. Once again, at price vector p', the two consumers' demands are compatible. Note that each consumer's demand is homogeneous of degree zero in the price vector p = (p" p,); that is, if prices double, then the consumer's wealth also doubles and his budget set remains unchanged. Thus, from Definition \5.B.I, we see that if P* = (pf, p!) is a Walrasian equilibrium price vector, then so is IlP' = (Ilpf, IIp!) for any", > O. In short, only the relative prices pUp! are determined in an equilibrium.
xr,
Flgur. 15.B.4 (top right)
Optimal consumption for consumer I at prices p.
Flgur. 15.B.5 (bottom)
xr
Consumer I's offer curve. This implies that the consumer's offer curve lies within the upper contour set of w, and that, if inditTerence curves are smooth, the offer curve must be tangent to the consumer's indifference curve at the endowment point. Figure 15.B.6 represents the demanded bundles of the two consumers at some arbitrary price vector p. Note that the demands expressed by the two consumers are not compatible. The total demand for good 2 exceeds its total supply in the economy ';'" whereas the total demand for good I is strictly less than its endowment w,. Put somewhat ditTerently, consumer I is a net demander of good 2 in the sense that he wants to consume more than his endowment of that commodity. Although consumer 2 is willing to be a net supplier of that good (he wants to consume less than his endowment), he is not willing to supply enough to satisfy consumer I's needs. Good 2 is therefore in excess demand in the situation depicted in the figure. In contrast, good I is in excess supply. At a market equilibrium where consumers take prices as given, markets should clear. That is, the consumers should be able to fulfill their desired purchases and
Example \S.B.I: Suppose that each consumer i has the Cobb-Douglas utility function
=
=
II,(X,;, Xli) xiix~i-'. In addition, endowments are WI = (I, 2) and w, (2, I). At prices p = (PI' p,), consumer I's wealth is (p, + 2p,) and therefore his demands lie
on the offer curve (recall the derivation in Example 3.0.1): OC,(p) = (Il(PI
+ 2p , ), (I PI
d
- 1l)(PI p,
,»).
+ 2P
figure 15.B.6
A price vector with excess demand for good 2 and excess supply for good I.
520
Ci1A":',"ER
15
GENERAL
EQUILIBRIUM
xi
2
~~~______r-__________~'02
THEORY:
_______
EXAMPLES
SEC 1 ION
Indifference Curves Ihrough w ~-+~~ -+'O~C~,~________~02
~~
1 5 . B:
PUR E
______-r___
E X C HAN C. E:
T H £
E 0 GE W 0 A TH
BOX
~02
__
x_~ ~:_. >. ,-
,',1,----------,-I
SOME
____ :'__-I_ __ } ,', p'
Figure 15.B.9
Multiple Walrasian equilibria.
0,
o,,-,_~.~_
x~ I (b)
(a)
~-I----------~~
The Edgeworth box, simple as it is, is remarkably powerful. There are virtually no phenomena or properties of general equilibrium exchange economies that cannot be depicted in it. Consider, for example, the issue of the uniqueness of Walrasian equilibrium. In Chapter 10, we saw that if there is a numeraire commodity relalive to which preferences admit a quasilinear representation, then (with strict convexity of preferences) the equilibrium consumption allocation and relative prices are unique. In Figure 15.B.7, we also have uniqueness (see Exercise 15.B.2 for a more explicit discussion). Yet, as the Edgeworth box in Figure 15.B.9 shows, this property does not generalize. In that figure, preferences (which are entirely nonpathological) are such that the offer curves change curvature and interlace several times. In particular, they intersect for prices such that p,/p, is equal to t, I, and 2. For the sake of completeness, we present an analytical example with the features of the figure.
Figure 15.B.7 ('op)
(a) A Walrasian equilibrium. (b) The consumer's offer curves intersect at the Walrasian equilibrium allocation.
<:, Figure 15.B.8 (bottom)
A Walrasian equilibrium allocalion on Ihe boundary of the Edgeworth box.
Example 15.B.2: Let the two consumers have utility functions
u,(x",x,,) = XII -Ax 2," and ",(X",X22) = -~XI>" + x". Note that the utility functions are quasilinear (which, in particular, facilitates the computation of demand), but with respect to different numeraires. The endowments are w, = (2, r) and (1), = (r, 2), where r is chosen to guarantee that the equilibrium prices turn out to be round numbers. Precisely, r = 21/9 > O.ln Exercise 15.B.5, you are asked to compute the offer curves of the two consumers. They are:
Observe that the demands for the first and the second good are, respectively, decreasing and increasing with p,. This is how we have drawn OC, in Figure 15.B.7(b). Similarly, OC,(p) = (a(2p, + p,)/p" (I - 0()(2p, + p,)/p,). To determine the Walrasian equilibrium prices, note that at these prices the total amount of good I consumed by the two consumers must equal 3 (= W II + (1) ,,). Thus, a(pi
+ 2p!l + 0(2pi + p!) = pi
2"10 -
3. OC,(p" p,)
pi
Solving this equation yields
!1=_a_ p!
I - a
_
= (2 + r(~) (~)"IO, (~) -1/ )>> 0 9
and
(15.B.2) OC,(p" p,) =
Observe that at any prices (pi, p!) satisfying condition (15.B.2), the market for good 2 clears as well (you should verify this). This is a general feature of an Edgeworth box economy: To determine equilibrium prices we need only determine prices at which one of the markets clears; the other market will necessarily clear at these prices. This point can be seen graphically in the Edgeworth box: Because both consumers' demanded bundles lie on the same budget line, if the amounts of commodity I demanded are compatible, then so must be those for commodity 2. (See also Exercise 15.B.1.) •
(( P,p,)
-I/O
,2 + r
(p,) p; - (p,)"IO) p, »0.
NOle that, as illustrated in Figure 15.B.9, and in contrast with Example 15.B.I, consumer I's demand for good I (and symmetrically for consumer 2) may be increasing in p ,. To compute the equilibria it is sufficient to solve the equation that equates the total demand of the second good to its total supply, or
(p )"10 = 2 + r. ~ - ~ (~p)-'/9 + 2 + r (p)
d
521
522
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY:
SOME
SECTION
EXAMPLES
PURE
EXCHANGE:
THE
EDGEWORTH
BOX
523
Ahr-----~~~------_.O,
<:,
__~~--------~----------_,o,
15.B:
OC, 0 \ -*-t----~---+----------.
<:,
'
Direction of Increasing
Preference for
<:,
Consumer 2 0, OC, (a)
0,
(b)
Flgur.'S.B." (.)
Flgur.'S.B.'0
(c)
(a) Allocation x is not Pareto optimal. (b) Allocation x is Pareto optimal. (c) Allocation x is Pareto optimal.
(b)
strictly better off. Definition IS.B.2 expresses this idea in the setting of our two-consumer, pure exchange economy.
(a) and (b): Two examples of nonexistence of Walrasian equilibrium.
Recalling the value of r, this equation has three solutions for P,/P2: 2, I, and should check this). •
t (you
Definition 15.B.2: An allocation x in the Edgeworth box is Pareto optimal (or Pareto efficient) if there is no other aliocaton x' in the Edgeworth box with x: >-. x·for i = 1, 2 and x; >-j x, for some i. ,-, ,
It may also happen that a pure exchange economy does not have w')' Walrasian equilibria. For example, Figure IS.B.IO(a) depicts a situation in which the endowment lies on the boundary of the Edgeworth box (in the northwest corner). Consumer 2 has all the endowment of good I and desires only good 1. Consumer I has all the endowment of good 2 and his indifference set containing w" {x, E R~: x, -, w,}, has an infinite slope at w, (note, however, that at w" consumer I would strictly prefer receiving more of good I). In this situation, there is no price vector p' at which the consumers' demands are compatible. If P2/P, > 0 then consumer 2 optimal demand is to keep his initial bundle w 2 , whereas the initial bundle w, is never consumer I's optimal demand (no matter how large the relative price of the first good, consumer I always wishes to buy a strictly positive amount of it). On the other hand, consumer I's demand for good 2 is infinite when P2/P, = O. Note for future reference that consumer 2's preferences in this example are not strongly monotone. Figure IS.B.IO(b) depicts a second example of nonexistence. There, consumer I's preferences are nonconvex. As a result, consumer I's offer curve is disconnected, and there is no crossing point of the two consumers' offer curves (other than the endowment point, which is not an equilibrium allocation here). In Chapter 17, we will study the conditions under which the existence of a Walrasian equilibrium is assured.
Figure \S.B.II(a) depicts an allocation x that is not Pareto optimal. Any allocation in the interior of the crosshatched region of the figure, the intersection of the sets {x', ER~: x', <:::, x,} and {x~ ER~: x~ <:::2 x 2 } within the Edgeworth box, is a feasible allocation that makes both consumers strictly better off than at x. The allocation x depicted in Figure IS.B.Il(b), on the other hand, is Pareto optimal because the intersection of the sets {x; E IR~: x; <:::i Xi} for j = I, 2 consists only of the point x. Note that if a Pareto optimal allocation x is an interior point of the Edgeworth box, then the consumers' indifference curves through x must be tangent (assuming that they are smooth). Figure IS.B.11(c) depicts a Pareto optimal allocation x that is not interior; at such a point, tangency need not hold. The set of all Pareto optimal allocations is known as the Pareto set. An example is illustrated in Figure IS.B.12. The figure also displays the contract curve, the part of the
Figure 15.6.12
The Pareto set and the contract curve.
Welfare Properties of Walrasian Equilibria A central question in economic theory concerns the welfare properties of equilibria. Here we shall focus on the notion of Pareto optimality, which we have already encountered in Chapter 10 (see, in particular, Section 10.B). An economic outcome is Pareto optimal (or Pareto efficient) if there is no alternative feasible outcome at which every individual in the economy is at least as well off and some individual is
d
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY:
SOME
.. t_CTICN
EXAMPLES
1::.
.... :
,Ht.
ONE-CONSUME.H,
ONE·PROOUCER
lCONQMT
~~_,
-----------------------------------------------------------------------------------------Pareto set where both consumers do at least as well as at their initial endowments. The reason for this term is that we might expect any bargaining between the two consumers to result in an agreement to trade to some point on the contract curve; these arc the only points at which both of them do as well as at their initial endowments and for which there is no alternative trade that can make both consumers beller 01T. We can now verify a simple but important fact: Any Walrasimr equilihrium al/ocatioll x· lIecessarily belollgs to the Pareto set. To sec this, note that by the definition of a Walrasian equilibrium the budget line separates the two at-Ieast-asgood-as sets associated with the equilibrium allocation, as seen in Figures 15.B.7(3) and 15. B.l!. The only point in common between these two sets is x' itself. Thus, at any competitive allocation x', there is no alternative feasible allocation that can benefit one consumer without hurting the other. The conclusion that Walrasian allocations yield Pareto optimal allocations is an expression of the first fundamental theorem of welfare ecollomics, a result that, as we shall see in Chapter 16, holds with great generality. Note, moreover, that since each consumer must be at least as well olTin a Walrasian equilibrium as by simply consuming his endowment, any Walrasian equilibrium lies in the contract curve portion of the Pareto set. The first fundamental welfare theorem provides, for competitive market economics, a formal expression of Adam Smith's "invisible hand." Under perfectly competitive conditions, any equilibrium allocation is a Pareto optimum, and the only possible welfare justification for intervention in the economy is the fulfillment of distributional objectives. The secolld fundamell/at theorem of welfare economics, which we also discuss extensively in Chapter 16, offers a (partial) converse result. Roughly put, it says that ullder convexity assumptions (not required for the first welfare theorem), a planner can achieve allY desired Pareto optimal al/ocation by appropriately redistributing wealth ill a lump-sum fashion and then "letting tIre market work." Thus, the second welfare theorem provides a theoretical affirmation for the use of competitive markets in pursuing distributional objectives. Definition 15.B.3 is a more formal statement of thc concept of an equilibrium with lump-sum wealth redistribution.
~-+
,,
"" x·
for all
xi E IR~
such that
p"
~~
__~~______________'"o,
~ /p' "1 ,,
".~) Wealth
TransFer, 0,
0, (a)
Figure 15.8.13
(b)
The second fundamental welfare theorem. (a) Using wealth transfers. (h) Using. transfers of endowments.
indicated in the figure. the price vector p' clears the markets for the two goods, and allocation x' results. Note that this wealth transfer may also be accomplished by directly transferring endowments. As Figure 15.B.13(b) illustrates, a transfer of good I that moves the endowment vector to w' will have the price vector p' and allocation x' as a Walrasian equilibrium. A transfer of good 2 that changes endowments to w" does sO as well. In fact, if all commodities can be easily transferred, then we could equally well move the endowment vector directly to allocation x'. From this new endowment point, the Walrasian equilibrium involves no trade" Figure 15.8.14 shows that the second welfare theorem may fail to hold when preferences are not convex. In the figure, x' = (xf, x!J is a Pareto optimal allocation that is not supportable as an equilibrium with transfers. At the budget line with the property that consumer 2 would demand x!, consumer I would prefer a point other than xf (such as x'd. Convexity, as it turns out, is a critical assumption for the second welfare theorem. A failure of the second welfare theorem of a dilTerent kind is illustrated in Figure IS.B.IO(a). There, the initial endowment allocation w is a Pareto optimal allocation, but it cannot be supported as an equilibrium with transfers (you should check this). In this case, it is the assumption that consumers' preferences are strongly monotone that is violated. For further illustrations of Edgeworth box economies see, for example, Newman (1965).
Definition 15.B.3: An allocation x' in the Edgeworth box is supportable as an equilibrium with transfers if there is a price system p' and wealth transfers T, and T2 satisfying T, + T2 = 0, such that for each consumer i we have
xi' ;::1 xi
____~~~____________~o,
xi :5: p' 'Wi + Ti ·
lS.C The One-Consumer, One-Producer Economy
Note that the transfers sum to zero in Definition 15.B.3; the planner runs a balanced budget, merely redistributing wealth between the consumers. Equipped with Definition 15.B.3, we can state more formally a version of the second welfare theorem as follows: if the preferences of the two consumers in the Edgeworth box are continuous, convex, and strongly monotone, then allY Pareto optimal al/ocatioll is supporcable as an equilibrium with transfers. This result is illustrated in Figure 15.B.13(a), where the consumer's endowments are at point w. Suppose that for distributional reasons, the socially desired allocation is the Pareto optimal allocation x'. Then if a tax authority constructs a transfer of wealth between the two consumers that shifts the budget line to the location
We now introduce the possibility of production. To do so in the simplest-possible selling, we suppose that there are two price-taking economic agents, a single 4. In practice, endowments may be difficullto transfer (e.g., human capital), and so the ability to use wealth transfers (or transfers of only a limited number of commodities) may be important. It is worth obscrving that there is one attractive feature of transferring endowments directly to the desired Pareto optimal allocation: we can be assured that x· is the unique Walrasian equilibrium allocation artcr the transfers (strictly speaking this requires a strict convexity assumption on preferences).
d
526
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY:
aOME
SEC T ION
EXAMPLES
1 5 • C:
THE
0 N E - CON S U iii E R,
0 N E - PRO DUe ERE CON 0 iii Y
-------------------------------------------------------------------------------------------Budget Line Tangent to Indifference Curves at x' 0 ~~r_--~--~------------~ 1
(-:(p, w), q(p, w)) (-z(P.").q(P.··»
Figure 15.6.14
Failure of the second welfare theorem with nonconvex preferences. (a)
consumer and a single firm, and two goods, the labor (or leisure) of the consumer and a consumption good produced by the firm.' The consumer has continuous, convex, and strongly monotone preferences 2::: defined over his consumption of leisure x I and the consumption good Xl' He has an endowment of [ units of leisure (e.g., 24 hours in a day) and no endowment of the consumption good. The firm uses labor to produce the consumption good according to the increasing and strictly concave production function f(z), where z is the firm's labor input. Thus, to produce output, the firm must hire the consumer, effectively purchasing some of the consumer's leisure from him. We assume that the firm seeks to maximize its profits taking market prices as given. Letting p be the price of its output and w be the price of labor, the firm solves Max
pf(z) - wz.
xl(p', 11") = q(p', w·)
R:
z(p', w') = [ - xl(p', 11")
PXl
(15.C.4)
Figure 15.C.1 illustrates the working of this one-consumer, one-firm economy. Figure 15.C.I(a) depicts the firm's problem. As in Chapter 5, we measure the firm's use oflabor input on the horizontal axis as a negative quantity. Its output is depicted on the vertical axis. The production set associated with the production function fez) is also shown, as are the profit-maximizing input and output levels at prices (p, w), zIp, 11') and q(p, 11'), respectively. Figure 15.C.1 (b) adapts this diagram to represent the consumer's problem. Leisure and consumption levels are measured from the origin denoted 0, at the lower-lefthand corner of the diagram, which is determined by letting the length of the segment [0,,0,] be equal to [, the total labor endowment. The figure depicts the consumer's (shaded) budget set given prices (p, w) and profits n(p, w). Note that if the consumer consumes [ units of leisure then since he sells no labor, he can purchase n(p, w)/p units of the consumption good. Thus, the budget line must cut the vertical q-axis at height nIP, w)/p. In addition, for each unit of labor he sells, the consumer earns w and can therefore afford to purchase w/p units of Xl' Hence, the budget line has slope -(w/p). Observe that the consumer's budget line is exactly the isoprofit line associated with the solution to the firm's profit-maximization problem in Figure 15.C.I(a), that is, the set of points {( -z, q): pq - wz = n(p, w)} that yield profits
(15.C.2)
U(XI,Xl)
S.t.
(15.C.3)
and
(l5.C.1)
:~o
Max
(a) The firm's problem. (b) The consumer's problem.
The budget constraint in (15.C.2) reftects the two sources of the consumer's purchasing power: If the consumer supplies an amount ([ - XI) of labor when prices are (p, w), then the total amount he can spend on the consumption good is his labor earnings 11'([ - XI) plus the profit distribution from the firm n(p, 11'). The consumer's optimal demands in problem (t 5.C.2) for prices (p, w) are denoted by (x l ( p, w), Xl(P, w». A Walrasian equilibrium in this economy involves a price vector (p', w·) at which the consumption and labor markets clear; that is, at which
Given prices (p, w), the firm's optimal labor demand is z(p, w), its output is q(p, 11'), and its profits are n(p, w). As we noted in Chapter 5, firms are owned by consumers. Thus, we assume that the consumer is the sole owner of the firm and receives the profits earned by the firm n(p, w). As with the price-taking assumption, the idea of the consumer being hired by his own firm through an anonymous labor market may appear strange in this model with only two agents. Nevertheless, bear with us; our aim is to illustrate the workings of more complicated many-consumer general equilibrium models in the simplest-possible way." Letting II(XI' Xl) be a utility function representing 2:::, the consumer's problem given prices (p, w) is (XI. ,xl)E
(b)
Figure IS.C.l
s: 11'([ - xtl + n(p, 11').
5. One-consumer economies are sometimes referred to as Robinson Crusoe economies. 6. The poinl made in footnote I can be repealed here: we could imagine that Ihe firm and the consumer stand for a large number of identical firms and consumers. We comment a bit more on this interpretation at the end of this Section.
ofn(p, w).
d
527
528
C HAP T E R
1.5:
G ENE R ALE QUI LIB R I U M
THE 0 R Y:
S0 M E
SECTION
E X AMP L E S
,..0:
THE
2.2
PRODUCTION
x,
..------------- "
Slope =
I\'
x'2
Flgur. 1S.C.2 A Walrasian
I.
equilibrium.
(a)
The prices depicted in Figure IS.C.I(b) arc not equilibrium prices; at these prices, there is an excess demand ror labor (the firm wants more labor than the consumer is willing to supply) and an excess supply or the produced good. An equilibrium price vector (p*, \\'*) that clears the markets ror the two goods is depicted in Figure IS.C.2. There is a very important ract to notice rrom Figure IS.C.2: A parricillar cotlSllmplion-ieisure combillaliol1 can arise in a compelilil'c' eC/I,i1ihrium illlnci lml.\' if it maximi:es tilt' conSllmers IItilit)' slIbject to tltt' ecollomy's tee/llwlof/iral alltl CIldowmelll cOIlstraillls. That is, the Walrasian equilibrium allocation is the same allocation that would be obtained ir a planner ran the economy in a manner that maximized the consumer's well-being. Thus, we see here an expression or the rundamental theorems or welrare economics: Any Walrasian equilibrium is Pareto optimal, and the Pareto optimal allocation is supportable as a Walrasian equilibrium.7 The indispensability or convexity ror the second welrare theorem can again be observed in Figure I S.C.3(a). There, the allocation x* maximizes the welrare or the consumer, but ror the only value or relative prices that could support x* as a utility-maximizing bundle, the firm does not maximize profits even locally (i.e., at the relative prices w/p, there are productions arbitrarily close to x* yielding higher profits). In contrast, the first welrare theorem remains applicable even in the presence ornonconvexities. As Figure IS.C.3(b) suggests, any Walrasian equilibrium maximizes the well-being or the consumer in the reasible production set. Under certain circumstances, the model studied in this section can be rigorously justiOed as representing the outcome of a more general economy by interpreting the" Orm" as a representative producer (see Section 5.E) and the "consumer" as a representative consumer
(sec Section 4.D). The rormer is always possihle, but the latter-.. -that is, the exislence of a (normative) representative consumer-requires strong conditions. Jr, however. the economy 7. In a single·consumer economy, the test for Pareto optimality reduces to the question of whether the well-being of the single consumer is being maximized (subject to feasibility constraints). Note that given the convexity of preferences and the strict convexity of the aggregate production set assumed here, there is a unique Pareto optimal consumption vector (and therefore a unique
equilibrium).
'1 (b)
FIgure 1S.C.3 (a) Failure of the second welfare theorem with a nonconvex technology. (b) The first welfare theorem applies even with a nonconvex technology. is composed of many consumers with identical concave utility functions and identical initial endowments, and if society has a strictly concave social welfare function in which these consumers arc treated symmetrically, then a (normative) representative consumer exists who has the same utility function as the consumers over levels of per capita consumption. 8 (We can
also think of the representative firm's input and output choices as being on a per capita basis). For more general conditions under which a representative consumer exists, see Section 4.D.
lS.D The 2 x 2 Production Model In this section, we discuss an example that concentrates on general equilibrium effects in production. To begin, consider an economy in which the production sector consists of J firms. Each firm j produces a consumer good q} directly rrom a vector or L primary (i.e., non produced) inputs, or factors, Zj = (z1), ...• ZLj) ~ 0." Firm j's production takes place by means or a concave, strictly increasing, and differentiable production runction fi(z,). Note that there are no intermediate goods (i.e., produced goods that are used as inputs). The economy has total endowments or the L ractor inputs, (i l , ... , i,.> »0. These endowments are initially owned by consumers and have a usc only as production inputs (i.e., consumers do not wish to consume them). To concentrate on the ractor markets or the economy, we suppose that the prices of the J produced consumption goods arc fixed at p = (PI' ... ' PJ), The leading example for this assumption is that or a small open economy whose trading decisions in the world markets ror consumption goods have little effect on the world prices of 8. To see this, note that an equal distribution of wealth (which is what occurs here in the absence of any wealth transfers given the identical endowments of the consumers) maximizes sodal welfare for any price vector and aggregate wealth level.
9. Some of these outputs may be the same good; thai is, firms j and j' may produce the same commodity.
MODEL
529
530
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY:
SOME
SECTION
EXAMPLES
:)
~
p.=
=
'r ow,
«zt,.···. zt.>.···. (zrJ.···. Z0».
such that firms receive their desired factor demands under prices (P. w') and all the factor markets clear. that is. such that
qil
_
=
for all j = I ..... J Max
and
(:, ..... :J)C!!O
(=
2 x 2
PRODUCTION
for
z,
(=
I •...• L.
(15.0.3) (15.0.4)
Conditions (15.0.3) and (15.0.4) constitute a system of L + J equations in the L + J endogenous variables (w, •...• wL ) and (q, •...• qJ). Condition (15.0.3) states that each firm must be at a profit-maximizing output level given prices p and w'. If so. firmj's optimal demand for the (th input is z1j = Jcj(w. Q7>/Jw, (this is Shepard's lemma; see Proposition 5.C.2). Condition (15.0.4) is therefore the factor marketclearing condition. Before examining the determinants of the equilibrium factor allocation in greater detail. we note that the equilibrium factor allocation (zt .. ..• zJ) in this model is exactly the factor allocation that would be chosen by a revenue-maximizing planner. thus providing us with yet another expression of the welfare-maximizing property of competitive allocations (the first welfare theorem).'2 To see this, consider the problem faced by a planning authority who is charged with coordinating factor allocations for the economy in order to maximize the gross revenues from the economy's production activities:
Pj~(Zj) - w,z j .
for all
for j = I •...• J.
Jqj
" JCj(w'.
0
zj E =j(P' w)
JCj(w'. qj)
J
We denote firmj's set of optimal input demands given prices (P. w) by z(P. w) c IR~. Ikcause consumers have no direct use for their factor endowments. the total factor supply will be (i, •...• id as long as the input prices WI are strictly positive (the only casc that will concern us here). An equilibrium for the factor markets of this economy given the fixed output prices P therefore consists of an input price vcctor 1\'* = (wr •...• I\'t) »0 and a factor allocation
(zt.···. zi>
THE
conditions hold:
these goods.'o Output is sold in world markets. Factors. on the other hand. are immobile and must be used for production within the country. The central question for our analysis concerns the equilibrium in the factor markets; that is. we wish to determine the equilibrium factor prices W = (w, •...• wd and the allocation of the economy's factor endowments among the J firms." Given output prices p = (p, •...• PJ) and input prices w = (w, •...• wL ). a profitmaximizing production plan for firm j solves Max
15.0:
I ..... L.
I
pj~(z;l
(15.0.5)
j
s.t.
LZj = i.
zn
How does the equilibrium factor allocation (zt .. ... compare with what this planner does? Recall from Section 5.E that whenever we have a collection of J price-taking firms, their profit-maximizing behavior is compatible with the behavior we would observe if the firms were to maximize their profits jointly taking the prices of outputs and factors as given. That is. the factor demands (zt •. .. , z1) solve
Because of the concavity of firms' production functions. first-order conditions are both necessary and sufficient for the characterization of optimal factor demands. Therefore. the L(J + I) variables formed by the factor allocation (=r ..... =1) E R~J and the factor prices w' = (IVt •...• wl) constitute an equilibrium if and only if they satisfy the following L(J + 1) equations (we assume an interior solution here):
Max
(15.0.6)
(:I •..•• :J)~O
for j = I •...• J and { = I ....• L
(15.0.1)
Since Lj zj = i (by the equilibrium property of market clearing). the factor demands must also solve problem (15.0.6) subject to the further constraint that Lj Zj = i. But this implies that the factor demands (zt ....• z1) in fact solve problem (15.0.5): if we must have Lj Zj = i. then the total cost w' -(LI z) is given. and so the joint profit-maximizing problem (15.0.6) reduces to the revenue-maximizing problem (15.0.5).
(zt.···. z1)
and for ( = I•...• L.
(15.0.2)
The equilibrium output levels are then qj = ~(zj) for every j. Equilibrium conditions for outputs and factor prices can alternatively be stated using the firms' cost functions cl(w. qj) for j = I•...• J. Output levels (qt • ...• qj) » 0 and factor prices 11'* »0 constitute an equilibrium if and only if the following
One benefit of the property just established is that it can be used to obtain the equilibrium factor allocation without a previous explicit computation of the equilibrium factor prices; we simply need to solve problem (l5.D.5) directly. It also provides a useful way of viewing the equilibrium factor prices. To sce this. consider again the joint profit-maximization problem (15.D.6). Wc can approach this problem in an equivalent manner by first deriving an aggregate
10. See Exen:isc 15.D.4 ror an endogenous determination (up to a scalar multiple) or the pri~cs I' ~ (PI"
.. PJ)'
II. Note that once the factor prices and allocations are determined. each consumer's demands G.ln be readily determined from his demand function given the exogenou5 prices (PI" .. • PJ) and
12. Note thai maximization or economy-wide revenue rrom production would be the goal of any planner who wanted to maximize consumer welrare: it allows ror the maximal purchases of consumption goods. at the fixed world prices.
Ihe wealth derived from factor input sales and profit distribulions. Recall that the currenl model is completed by assuming that this demand is met in the world markets.
rl
MODEL
531
5:.:..::
~HAPTEK
i!i:
GENERAL
EQUILIBRIUM
THEORY:
SOME
.:)tCTION
EXAMPLES
15.0.
tHE
", -.~
f(:",:,,)
=I
FIgure 15.0.1
2)(2
PRODUCTION
__________________
MODEL
533
~o,
0,
(a) A unit isoquant. (b) The unit cost (b)
(a)
function. (a) Figure 15.0.2
production function for dollars: f(:) =
Max
p.!.(Z.)
S.t.
+ ... + pJlJ(ZJ)
(h) (a)
An inefficient factor allocation.
(b)
The Pareto set of factor allocations.
represent the possible allocations of the factor endowments between the two firms in an Edgeworth box of size i, by The factors used by firm I are measured from the southwest corner; those used by firm 2 are measured from the northeast corner. We also represent the isoquants of the two firms in this Edgeworth box. Figure 15.D.2(a) depicts an inemcient allocation z of the inputs between the two firms: Any allocation in the interior of the hatched region generates more output of hOlh goods than docs :. Figure 15.D.2(b), on the other hand, depicts the Pareto set of factor allocations, that is, the set of factor allocations at which it is not possible, with the given total factor endowments, to produce more of one good without producing less of the other. The Pareto set (endpoints excluded) must lie all above or all below or be coincident with the diagonal of the Edgeworth box. If it ever cuts the diagonal then because of constant returns, the isoquants of the two firms must in fact be tangent all along the diagonal, and so the diagonal must be the Pareto set (see also Exercise 15.B.7). Moreover, you should convince yourself of the correctness of the following claims.
=,.
Liz)=:,
The aggregate factor demands must then solve Max.~o({(z) - w'z). For every I, the first-order condition for this problem is IV, = Df(z)/elz, . Moreover, at an equilibrium, the aggregate usage of factor ( must be exactly :/ Hence, the equilibrium factor price of factor ( must be W, = (if(:)/ill, ; that is, rhe prke of factor { mu.W be exactly equal to its aggregate margi/wl productivit}' (in rerms of revenue). Since f(·) is concave, this observation by itself generates some interesting comparative statics. For example, a change in the endowment of a single input must change the equilibrium price of the input in the opposite direction. Let us now be more specific and take J = L = 2, so that the economy under study produces two outputs from two primary factors, We also assume that the production functions I,(z", ZIt), 1'(Z12' ZIt) are homogeneous of degree one (so the technologies exhibit constant returns to scale; see Section 5.B). This model is known as the 2 x 2 production model. In applications, factor I is often thought of as labor and factor 2 as capital. For every vector of factor prices W = (w" w,), we denote by cj(w) the minimum cost of producing one unit of good j and by OJ(w) = (o,iw), o'l(w» the input combination (assumed unique) at which this minimum cost is reached. Recall again from Proposition 5.C.2 that Vcj(w) = (olj(w), O'j(w», Figure IS.D.I(a) depicts the unit isoquant of firm j,
Exercise 15.0.1: Suppose that the Pareto set of the 2 x 2 production model does not coincide with the diagonal of the Edgeworth box. (a) Show that in this case, the factor intensity (the ratio of a firm's use of factor I relative to factor 2) of one of the firms exceeds that of the other at every point along the Pareto set. (b) Show that in this case, any ray from the origin of either of the firms can intersect the Pareto set at most once. Conclude that the factor intensities of the two firms and the supporting relative factor prices change monotonically as we move along the Pareto set from one origin to the other.
{(zlj' Z,j) E R~: Jj(z'j' z,;l = I}, along with the cost-minimizing input combination (o,j(w), o,iw». In Figure IS.D.I(b), we draw a level curve of the unit cost function, {(WI' w,): ciw" w,) = c}. This curve is downward sloping because as w, increases, w, must fall in order to keep the minimized costs of producing one unit of good j unchanged. Moreover, the set {(WI' w,): Cj(w" WI) 2: c} is convex because of the concavity of the cost function Cj(w) in IV. Note that the vector VCj(w), which is normal to the level curve at W = (w" w,), is exactly (o,j(w), O'j(w», As we move along the curve toward higher w, and lower
In Figure 15.0.3, we depict the set of nonnegative output pairs (q"q,) that can be produced using the economy's available factor inputs. This set is known as the production possihilil)' SCI. Output pairs on the frontier of this set arise from factor allocations lying in the Pareto set of Figure IS.D.2(b). (Exercise IS.o.2 asks you to prove that the production possibility set is convex, as shown in Figure 15.0.3.)
w" the ratio O'j(w)/o'j(w) falls. Consider, first, the efficient factor allocations for this model. In Figure IS.D.2, we
g
J
534
CHAPTER
'5:
GENERAL
EQUILIBRIUM
THEORY:
SOME
EXAMPLES
SECTION
15.0:
THE
2 x 2
PRODUCTION
MODEL
535
q, Flgur. 15.0.3 (left)
The production possibility set. Flgur. 15.0.4 (right)
The equilibrium [actor q,
prices and factor
Figure 15.D.5
intensities in an interior equilibrium.
The equilibrium factor allocation.
factor intensity condition, there is at most a single pair of factor prices that can arise liS the equilibrium factor prices of an inlerior equilibrium. t s Once Ihe equilibrium factor prices 11'* are known, Ihe equilibrium output levels can be found graphically by delermining the unique point (z!, z;! in the Edgeworth box of factor allocations at which both firms have the factor inlensities associated wilh faclor prices "", that is,
With the purpose of examining more closely the determinants of the equilibrium factor allocation (z!. z;! and the corresponding equilibrium factor prices 11'* = ("'t. w!). we now assume that the fuctor intensities of the two firms bear a systematic relation to one another. In particular. we assume that in the production of good I. there is. relative to good 2. a greater need for the first factor. In Definition IS.D.I we make precise the meaning of" greater need ". is relatively more intensive in factor 1
Definition 15.0.1: The production of good than is the production of good 2 if
and
a22 (w)
at all factor prices w = (w,. w 2). To determine the equilibrium factor prices. suppose that we have an illterior equilibrium in which the production levels of the two goods are strictly positive (otherwise. we say that the equilibrium is specialized). Given our constant returns assumption, a necessary condition for (IV!. "';! to be the factor prices in an interior equilibrium is that it satisfies the system of equations and
C,(W,. 11',) = p,.
a,,(IV')
The construction is depicted in Figure 15.0.5. An important consequence of this discussion is that in the 2 x 2 production model, if Ihe factor intensity condition holds, then as long as the economy does not specialize in Ihe production of a single good [and Iherefore (15.0.7) holds], the equilibrium factor prices depend only on the technologies of the two firms and on the oil/put prices p. Thus. the levels of the endowments matter only to the extent that Ihey delermine whether the economy specializes. This result is known in the international trade literature as the faclOr price equalizatioll theorem. The theorem provides conditions (which include the presence of tradable consumption goods, idenlical produclion technologies in each country. and price-taking behavior) under which the prices of nontradable factors are equalized across nonspecialized countries.
a,,(w) > a'2(w)
a2 ,(w)
zI, = ~~(IV') z!,
(15.0.7)
That is, at an interior equilibrium, prices must be equal to unit cost. This gives us two equations for the two unknown factor prices "'I and "".'3 Figure 15.0.4 depicts the two unit cost functions in (15.0.7). By expression (15.D.7). a necessary condition for (w,. 'v,) to be the factor prices of an interior equilibrium is that these curves cross at (w" w,). Moreover. the factor intensity assumption implies that whenever the two curves cross. the curve for firm 2 must be Oatter (less negatively sloped) than that for firm I [recall that VCj(w) = (a,j(w), a'j(w))], From this, it follows that the two curves can cross at most once. I. Hence. under the
We now present two comparative statics exercises. We first ask: How does a change in the price of one of the outputs, say Pt. affect the equilibrium factor prices and factor allocations? Figure 15.D.6(a), which depicts the induced change in Figure IS.D.4. identifies the change in factor prices. The increase in p, shifts firm I's curve
15. Note, however. that although ('\',. \\<2) may solve (15.0.7), this is not sufficient to ensure that (Ii',. \\'1) are equilibrium factor prices. In particular, even though (WI' "'2) solve (lS.D.7). no interior equilibrium may exist. In Exercise 15.D.6, you are asked to show that under the factor intensity condition, the equilibrium will involve positive production of the two goods if and only if
13. Expression (15.D.7) is the constant returns version of (15.0.3). Note Ihat the effect of the constant returns to scale assumption is to make (15.0.3) independent of the output levels (q" . .. , q,)
a'l(w) I, QI1(W) ---- > _. > - - ,
(for interior equilibria). 14. If they crossed several times, then the curve for firm 2 must cross the curve for firm 1 at least once from above. At this crossing point. the curve for firm 2 would be Sleeper than the curve for firm I, contradicting the factor intensity condition.
wht:rc \\' = (1\'1,1\'2) is the unique solution to (15.0.7). In words, the factor intensity of the overall economy must be intermediate between the factor intensities of the two firms computed at the sole vector of factor prices at which diversification can conceivably occur.
a 21 (\\')
d
:2
a 22 (\v)
536
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY:
SOME
s(
EXAMPLES
C T ION
1 5 • 0:
THE
2)( 2
PROD U C T I 0 HMO 0 E L
"" ParelO Set ~r----+ Factor Allocation afler
Factor Allocation
PI Increases
arter:, Increases
0, 0, (a)
"
I' I·
(h)
,.
-,
'I
Ftgur.'5.0.1 The Rybcszynsk i
,I
theorem.
Figure 15.0.6 The Stolper-Samuelson theorem. (a) The change in equilibrium faclOr prices. (b) The change in the equilibrium factor allocation. Denote
IA 1 = [the set {(w,. 11',): c,(w" w,) = p,}] outward toward higher factor price levels; the point of intersection of the two curves moves out along firm 2's curve to a higher level of WI and a lower level of w,. Formally, this gives us the result presented in Proposition 15.0.1.
this
2 x 2 matrix
by
A.
""(11"),,,,(11") - a"(",·),,,,( ..·)
I [ A -,
=!Ai
The factor intensity assumption implies that > O. Therefore A -, exists and we can compute it to be a,,(w')
-a,,( ....
-a,,( .... )
)J.
a,,( .... )
llenee. the entries of A -, arc positive at the diagonal and negative off the diagonal. Since = A ~ I tip. this implies that for d" = (1,0) we have dw, > 0 and (1\""2 < 0, as we wanted. -
du'
PropositIon 15.0.1: (Stolper-Samuelson Theorem) In the 2 x 2 production model with the factor intensity assumption. if Pi increases. then the equilibrium price of the factor more intensively used in the production of good i increases, while the price of the other factor decreases (assuming interior equilibria both before and after the price change).'·
wr
We have just seen that if p, increases, then /"'! increases. Therefore, both firms must move to a less intensive use of factor I. Figure 15.D.6(b) depicts the resulting change in thc equilibrium allocation of factors. As can be seen, the factor allocation moves to a new point in the Pareto set at which the output of good I has risen and that of good 2 has fallen. For the second comparative statics exercise, suppose that the total availability of factor I increases from i , to i'" What is the effect of this on equilibrium factor prices and output levels? Because neither the output prices nor the technologies have changed, the factor input prices remain unaltered
Proor: For illustrative purposes, we provide a formal proof to go along with the graphical analysis of Figure 15.0.6 presented above. Note that it suffices to prove the result for an infinitesimal change dp = (1,0). Differentiating conditions (15.0.7), we have dp, = Vc,(w')'dw = a,,(w')dw, + a,,(\\,') dll'2' dp, = Vc,(w')'dw = a,,(w') dll',
+ a,,(w') dw"
or in matrix notation,
Proposition 15.0.2: (Rybcszynski Theorem) In the 2 x 2 production model with the factor intensity assumption, if the endowment of a factor increases, then the production of the good that uses this factor relatively more intensively increases and the production of the other good decreases (assuming interior equilibria both before and after the change of endowment).
16. See Exercise 15.0.3 for a strengthening of this conclusion. We also note that. strictly speaking, the factor ioeosity condition is not required for this result. The reason is that, as we saw in Exercise 15.0.1, the firm that uses one factor, say factor t. more intensely is the same for any
poinl in the Pareto set ofractor allocations. Suppose. for example. Ihat we are as in Figure IS.D.2(b). where firm I uses faclor t more intensively. Then, when PI rises. we can see from Figure 15.0.3, and the overall revenue·rnaxirnizing property of equilibrium discussed earlier in this section, that the output of good 1 increases and that of good 2 decreases. This implies that we move along the
For further discussion of the 2 x 2 production model see, for example, Johnson (1971 ).
Pareto set in Figure IS.D.2(b) toward firm 2's origin. Therefore, recalling Exercise IS.D.I, both firms' intensity of use of factor 1 decreases. Hence. the equilibrium factor price ratio w~ /w; must increase. Finally, since firm 2 is still breaking even and its output price has not changed, this implies that w~ increases and w! decreases.
Consider the general case of an arbitrary number or factors L and outputs J. For given output prices. the zero-profit conditions [i.e., the general analog of expression (15.0.7)]
-
537
:,.;
"
538
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY:
SOME
EXAMPLES
SEC T ION
1 5 • E:
G ENE R A L
V E R SUS
PAR T I ALE QUI LIB R I U M
,
,
constitute a (nonlinear) system of J equations in L unknowns. If L > J, then there are too many unknowns and we cannot hope that the zero-profit conditions alone will determine the factor prices. The total factor endowments will playa role. If J > L, then there are too many equations and, for typical world prices, they cannot all be satisfied simultaneously. What this means is that the economy will specialize in the production of a number of goods equal to the number of factors L. The set of goods chosen may well depend on the endowments of factors. Beyond the 2 x 2 situation (the analysis of which, as we have seen, is quite instructive), the case L = J seems too coincidental to be of interest. Nevertheless, we point out that in this case the zero-profit conditions are nonlinear and that in order to guarantee a unique solution
(and versions of the Stolper-Samuelson and the Rybcszynski theorems), we need a generalization of the factor intensity condition. These generalizations exist, but they cannot be interpreted economically in as simple a manner as can the factor intensity condition of the 2 x 2 model.
lS.E General Versus Partial Equilibrium Theory There are some problems that are inherently general equilibrium problems. It would be hard to envision convincing analyses of economic growth, demographic change, international economic relations, or monetary policy that were restricted to only a subset of commodities and did not consider economy-wide feedback effects. Partial equilibrium models of markets, or of systems of related markets, determine prices, profits, productions, and the other variables of interest adhering to the assumption that there are no feedback effects from these endogenous magnitudes to the underlying demand or cost curves that are specified in advance. Individuals' wealth is another variable that general equilibrium theory regards as endogenously determined but that is often treated as exogenous in partial equilibrium theory. If general equilibrium analysis did not change any of the predictions or conclusions of partial equilibrium analysis, it would be of limited significance when applied to problems amenable to partial equilibrium treatment. It might be of comfort because we would then know that our partial equilibrium conclusions are valid, but it would not change our view of how markets work. However, things are not that simple. The choice of methodology may be far from innocuous. We now present an example [due to Bradford (1978)] in which a naive application of partial equilibrium analysis leads us seriously astray. See Sections 3.1 and IO.G for some discussion of when partial equilibrium theory is (approximately) justified.
A Tax IlIcidellce Example Consider an economy with a large number of towns, N. Each town has a single price-taking firm that produces a consumption good by means of the strictly concave production function f(z) (once again, we could reinterpret the model as having many identical firms in each town to make the price-taking hypothesis more palatable). This consumption good, which is identical across towns, is traded in a national market. The overall economy has M units of labor, inelastically supplied by workers who derive utility only from the output of the firms. Workers are free to move from town to town and do so to seek the highest wage. We normalize the price of the consumption good to be I, and we denote the wage rate in town n's labor market by w,.
Given that workers can move freely in search of the highest wage, at an equilibrium the wage rates across towns must be equal; that is, we must have w, = ... = W N = Iii. From the symmetry of the problem, it must be that each firm hires exactly MIN units of labor in an equilibrium. As a result, the equilibrium wage rate must be Iii = [,(M IN). The equilibrium profits of an individual firm are therefore f(MIN) - [,(MIN)(MIN). Now suppose that town 1 levies a tax on the labor used by the firm located there. We investigate the "incidence" of the tax on workers and firms (or, more properly, on the firms' owners); that is, we examine the extent to which each group bears the burden of the tax. If the tax rate is t and the wage in town 1 is w" the labor demand of the firm in town I will be the amount z, such that [,(z,) = t + WI' At this point, we may be tempted to argue that, since N is large, we can approximate and take the wage rates elsewhere, IV, to be unaffected by this change in town 1. Moreover, since labor moves freely, the supply correspondence of workers in town 1 should then be Oat "" < ,v, 00 at w, > II', and [0,00] at w, = Iii. Thus, taking a partial equilibrium view, the equilibrium wage rate in the town I labor market remains equal to Iii, and the labor employed in town I falls to the level z, such that [,(z t) = t + Iii (hence, some labor will shift to the other towns). By adopting this sort of partial equilibrium view of the labor market of town I, we are therefore led to conclude that the income of workers remains the same, as does the profit of every firm not located in town 1. Only the profit of the firm in town I decreases. The qualitative conclusion is that firms (actually, firms' owners) "bear" all of the tax burden. Labor, because it is free to move and because the number of untaxed firms is large, "escapes." Alas, this conclusion constitutes an egregious mistake, and it will be overturned by a general equilibrium view of the same model. We now look at the general equilibrium across the labor markets of all the towns. We know that the equilibrium wage rate must be such that w t = ... = W N and that all M units of labor arc employed. Let w(t) be this common equilibrium wage when the tax rate in town 1 is r. By symmetry, the firms in towns 2, ... , N will each employ the same amount of labor, z(t). Let z, (t) be the equilibrium labor demand of the firm in town 1 when town I's tax rate is !. Then the equilibrium conditions are (N -
I)z(l)
+ z,(t) = M. I'(z(t»
=
(IS.E.l)
w(t).
I'(z,(t» = w(t)
(IS.E.2)
+ t.
(IS.E.3)
Consider the impact on wages of the introduction of a small tax dt. Substituting from (IS.E.I) for z,(t) in (IS.E.3), differentiating with respect to t, and evaluating at t = 0 [at which point z,(O) = z(O) = (MIN», we get -(,,(MIN)(N - l)z'(O) = 1\"(0)
+ 1.
(IS.E.4)
But from (IS.E.2), we get ("(MIN)z'(O) = ",'(0).
(IS.E.5)
Substituting from (IS.E.S) into (IS.E.4) yields 11"(0) = --'-. N
Therefore, once the general equilibrium effects are taken into account, we see that
THE 0 R Y
539
S4u
l;nAt'TER
1$:
GENERAL
EQUILIBRIUM
THEORY:
SOME
EXAMPLES
EXERCISES
the wage rate in all towns falls with the imposition of the tax in town 1. However, we see that this fall in the wage rate approaches zero as N grows large. Thus, at this point, it may still seem that our partial equilibrium approximation will have given us the correct answers for large N. But this is not so. Consider the effect of the tax on total profits. The partial equilibrium approach told us that workers escaped the tax; all the tax fell as a burden on firms. But letting rr(w) be the profit function of a representative firm, the change in aggregate profits from the imposition of this tax is 17
(N - l)rr'(",)w'(O)
+ rr'(",)(w'(O) +
I) = rr'(w) (
I+ -;::;I)
N --N
N -
=
15.B.4C Consider an Edgeworth box economy. An offer curve has the gross substitute property if an increase in the price of one commodity decreases the demand for that commodity and increases the demand for the other one. (a) Represent in an Edgeworth box the shape of an offer curve with the gross substitute property. (b) Assume that the offer curves of the two consumers have the gross substitute property. Show then that the offer curves can intersect only Once (not counting the intersection at the initial endowments).
o.
Let us denote an offer curve as lIormal if an increase in the price of one commodity leads to all increase in the demand for that commodity only if the demands of the two commodities hoth in..:reasc.
Aggregate profits stay constant! Thus, all of the burden of a small tax falls on laborers, not on the owners of firms. Although the partial equilibrium approximation is correct as far as getting prices and wages about right, it errs by just enough, and in just such a direction, that the conclusions of the tax incidence analysis based on it are completely reversed. I "
17. Recall that the profits of the firm in town I are
~("~I)
(c) Represcnt in the Edgeworth box the shape of a normal offer curve that does not satisfy the gross substitute property.
(d) Show that there are preferences giving rise to ofTer curves that are not normal. Show that the demand function for such preferences is not normal (i.e., at some prices some good is inferior).
+ I).
(c) Show in thc Edgeworth box that if the olTer Curve of one consumer is normal and thai of the other salisrics the gross substitute property, then thc offer curvcs can intersect at most once (not counting the intersection at the initial endowments).
18. We note that the justifications of partial equilibrium analysis in terms of small individual
budget shares that we informally described in Sections 3.1 and IO.G do not apply here because the ··consumption·' goods in this example (jobs in different towns) arc perfect substitutes and therefore individual budget shares are not gU3nlnteed to be small at all prices.
(f) Show that two normal offer curves can intersect several times.
15.8.5'\ Verify that the ofTcr curves of Example 15.8.2 are as claimed. Solve also for the c1aimcd values of relative prices.
15.B.6" (D. Blair) Compute the equilibria of the following Edgeworth box economy (there is more than one):
REFERENCES Bradford, D. (1978). Factor prices may be constant but factor returns are not Economic ulters, 199-203.
u,(x", x'l) u,(x", x,,)
Johnson, H. G. (l97t). The Two-SeclOr Model of General Equilibrium. Chicago: Aldine-Atherton. Newman, P. (t965). The n..ory of ExclJange. Englewood Cliffs, N.J.: Prentice-Hall.
w,) + p,(L x,;(p)
-
w,) = 0 for all prices p.
(b) Argue that if the market for good I clears at prices p. » 0, then so does the market for good 2; hence, p. is a Walrasian equilibrium price vector.
15.B.2
= (1,0),
"', = (0,
I).
15.B.S" Suppose that both consumers in an Edgeworth box have continuous and strictly convex preferences that admit a quasilinear utility representation with the first good as numeraire. Show that any two Pareto optimal allocations in the interior of the Edgeworth box then involve the same consumptions of the second good. Connect this with the discussion of Chapter 10.
\5.B.\ A Consider an Edgeworth box economy in which the two consumers have locally nons.ti.ted preferences. Let x,,(p) be consumer i's demand for good ( at prices p = (p" p,).
A
W,
15.8.7" Show that if both consumers in an Edgeworth box economy have continuous, strongly 1110notone, and strictly convex preferences, then the Pareto set has no "holes": precisely, it is a connected sct. Show that if, in addition, the preferences of both consumers are homothetic, then the Pareto set lies entirely on one side of the diagonal of the box.
EXERCISES
(a) Show that P,(L; x,;(p) -
= (x,,' + (12/37)'x;,')- "', = «12/37)'x,,' + x;,')- "',
15.1l.9" Suppose that in a pure exchange economy (i.e., an economy without production), we haye two consumers, Alphanse and Betatrix, and two goods. Perrier and Brie. Alphanse and
Consider an Edgeworth box economy in which the consumers have the Cobb-
Relatrix have the utility functions:
Douglas utility functions U,(X 11 ,X21} = x~lx11" and U2(X 1 2,X22) = X12X12".6, Consumer i's endowments are (w H , w,;)>> 0, for i = 1,2. Solve for the equilibrium price ratio and allocation.
and
", =
Min{x p " (Xh')"':
(where x p • is Alphanse's consumption of Pcrrier. and so on). Alphanse starts with an endowment of 30 units of Perrier (and none of Brie); Betatrix starts with 20 units of Brie (and none of Perrier). Neither can consume negative amounts of a good, If the two consumers
How do these change with a differential change in w,,? 15.B.3 8 Argue (graphically) that in an Edgeworth box economy with locally nonsatiated preferences, a Walrasian equilibrium is Pareto optimal.
hchave as price takers, wh~1t is the equilibrium?
•
541
542
C " APT E R
1 5:
G ENE R ALE QUI LIB R I U"
THE Q R Y:
S Q .. E
E X A .. P L E S
EXERCISES
Suppose instead that Alphanse begins with only 5 units of Perrier while Betatrix's initial endowment remains 20 units of Brie. 0 units of Perrier. What happens now?
15.0.3" Show that the Stolper-Samuelson theorem (Proposition 15.0.1) can be strengthened to assert that the increase in the price of the intensive factor is proportionally larger than the increase in the price of the good (and therefore the well-being of a consumer who owns only the intensive factor must increase).
IS.B.to C (The Transfer Paradox) In a two-consumer, two-commodity pure exchange economy with continuous, strictly convex and strongly monotone preferences, consider the comparative statics of the welfare of consumer I with changes in the initial endowments w, = (w", Wll)
and
(1)2
IS.O.4 c Consider a general equilibrium problem with two consumer-workers (i = 1,2), two constant returns firms U = 1.2) with concave technologies, two factors of production (t = 1,2), and two consumption goods U = 1.2) produced, respectively. by the two firms. Assume that the production of consumption good I is relatively more intensive in factor I. Neither consumer
= (W1l,Wll)'
(a) Suppose first that the preferences of the two consumers are quasilinear with respect to
if the endowments of consumer 1 are increased to W'I » W, while Wl remains the same, then at equilibrium the utility of consumer I may decrease. Interpret
consumes either of the factors. Consumer I owns one unit of factor I while consumer 2 owns
the same numeraire. Show that
one unit of factor 2.
this observation and relate it to the theory of a quantity-setting monopoly.
(a) Set up the equilibrium problem as one of clearing the factor and goods markets (in a
(b) Suppose now that the increase in resources of consumer I constitute a transfer from consumer 2, that is, w', = w, + z and W2 = (1)2 - z with z ~ O. Under the same assumption as in (a). show that the utility of COnsumer I cannot decrease.
closed economy context) under the assumption that prices are taken as given and productions arc profit maximizing.
(e) Consider again a transfer as in (b). but this time preferences may not be quasilinear. Suppose that the transfer z is small and that similarly the change in the equilibrium (relative) price is restricted to be small. Show that it is possible for the utility of consumer I to decrease (this is called the Irallsfer paradox). A graphical illustration in the Edgeworth box suffices to make the point. Interpret in terms of the interplay between substitution and wealth effects.
(e) Suppose now that consumer I has a taste only for the first good and that consumer 2 cares only for the second good. Argue that a multiplicity of equilibria is possible.
(b) Suppose that consumer I has a taste only for the second consumption good and that consumer 2 cares only for the first good. Argue that there is at most one equilibrium.
[llint: Parts (b) and (e) can be answered by graphical analysis in the Edgeworth box of factors of production.]
(d) Show that in this Edgeworth box example (but, be warned, not more generally) the transfer paradox can happen only if there is a multiplicity of equilibria. [Hillt: Argue graphically in the Edgeworth box. Show that if a transfer to consumer I leads to a decrease of the utility of consumer I. then there must be an equilibrium at the no-transfer situation where consumer I gets an even lower level of utility.]
15.0.5" Show that the Rybcszynski theorem (Proposition 15.0.2) can be strengthened to assert that the proportional increase in the production of the good that uses the increased factor relatively more intensively is greater than the proportional increase in the endowment of the factor. IS.0.6c Suppose you are in the 2 x 2 production model with output prices (p" p,) given (the economy could be a small open economy). The factor intensity condition is satisfied (production of consumption good I uses factor I more intensely). The total endowment vector
IS.CI" This exercise refers to the one-consumer, one-firm economy discussed in Section 15.C. (a) Prove that in an economy with one firm, one consumer, and strictly convex preferences and technology. the equilibrium level of production is unique.
is:E H'.
(b) Fix the price of output to be I. Define the excess demand function for labor as
z,(w) = x,(w, w[
+ ,,(w)) + ,,(w) -
(a) Set up the equilibrium conditions for factor prices (wt, allowing for the possibility of specialization.
[,
wn
and outputs (qt, q!j
(b) Suppose that Ii' = (IV" Ii',) are factor prices with the property that for each of the two goods the unit cost equals the price. Show that the necessary and sufficient condition for the equilibrium determined in (a) to have (qr, q!j» 0 is that i belongs to the set
where IV is the wage rate. ,,(.) is the profit function. and x,(', .), ,,(.) are. respectively. the consumer's demand function for leisure and the firm's demand function for labor. Show that the slope of the excess demand function is not necessarily of one sign throughout the range of prices but that it is necessarily negative in a neighborhood of the equilibrium.
{(Z,. Z,) E R~: a,,(lv)/a2l(w)
(e) Give an example to show that there can be multiple equilibria in a strictly convex
> z,/z, > a 12 (w)/a,,(w)},
where "lj(IV) is the optimal usage (at factor prices w) of the input unit of good j. This set is called the ciiversification cone.
economy with one firm and two individuals, each of whom is endowed with labor alone. (Assume that profits are split equally between the two consumers.) Can this happen if the firm
t in the production of one
(e) The unit-dollar isoquant of good j is the set of factor combinations that produce an amount of good j of I dollar value. Show that under the factor intensity condition the unit-dollar isoquants of the two goods can intersect at most once. Use the unit-dollar isoquants to construct graphically the diversification cone. [Hint: If they intersect twice then there are two points (one in each isoquant) proportional to each other and such that the slopes of the isoquants at these points are identicaL]
operates under constant rather than strictly decreasing returns to scale?
IS.C2 A ConSIder the one-consumer. one-producer economy discussed in Section 15.C. Compute the equilibrium prices, profits. and consumptions when the production function is f(:) = :1/'. the utility function is u(x,. x,) = In x, + In and the total endowment of labor is [= I.
x,.
(d) When the total factor endowment is not in the diversification cone, the equilibrium is specialized. Can you determine, as a function of total factor endowments, in which good the economy will specialize and what the factor prices will be? Be sure to verify the inequality conditions in (0). To answer this question you can make use of the graphical apparatus developed in (e).
IS.D.I" In text. IS.0.2A Show that in the 2 x 2 production model the production possibility set is convex (assume free disposal).
d
543
)44
CHAPTER
15:
GENERAL
EQUILIBRIUM
THEORY;
SOME
EXAMPLES
15.0.78 Suppose there are two output goods and two factors. The production functions for
-
C
HAP
the two outputs are
f,(z". z,,) = 2(zll)'"
+ (z,,)'/'
and
f,(z". z,,) = (z,,)'"
Equilibrium and Its Basic Welfare
+ 2(z,,)I/'.
The international prices for these goods are p = (I. I). Firms are price takers and maximize profits. The total factor endowments are i = (i,. i,). Consumers have no taste for the consumption of factors of production. Derive the equilibrium factor allocation z!,). (zr,. z!,)) and the equilibrium factor prices (wr. w!) as a function of (i,. i,). Verify that you get the same result whether you proceed through equations (15.0.1) and (15.0.2) or by solving (15.0.5).
T
E
R
16
Properties
«zr,.
15.0.S" The setting is as in the 2 x 2 production model. The production functions for the two outputs are of the Cobb-Douglas type: f,(:II' %,,) = (:11)'"(%,,)'''
and
f'{:". Z21) = (z,,)'''(z,,)2/'.
The international output price vector is p = (I. I) and the total factor endowments vector is .: = O. Compute the equilibrium factor alloeations and factor prices for all possible values of i. Be careful in specifying the region of total endowment vectors where the economy will specialize in the production of a single good.
(=,. =,)>>
16.A Introduction
15.0.9c (The Hecbcher-Ohlin Theorem) Suppose there are two consumption goods. two factors. and two countries A and B. Each country has technologies as in the 2 x 2 production model. The technologies for the production of each consumption good are the same in the two countries. The technology for the production of the first consumption good is relatively more intensive in factor I. The endowments of the two factors arc fA E R~ and f. E R~ for countries A and B. respectively. We assume that country A is relatively better endowed with factor I. that is. ' .. /i'A > f ../i, •. Consumers are identical within and between countries. Their preferences are representable by increasing. concave. and homogeneous utility functions that depend only on the amount consumed of the two consumption goods. Suppose that factors are not mobile and that each country is a price taker with respect to the international prices for consumption goods. Suppose then that at the international prices p = (p,. p,) we have that. first. neither of the two countries specializes and. second. the international markets for consumption goods clear. Prove that country A must be exporting good I. the good whose production is relatively more intensive in the factor that is relatively
With this chapter. we begin our systematic study of equilibrium in economics where agents act as price takers. We consider a world with L commodities in which consumers and firms interact through a market system. In this market system. a price is quoted for every commodity. and economic agents take these prices as independent of their individual actions. We concentrate in this chapter on a presentation of the basic welfare properties of equilibria. Some more advanced topics in welfare economics are discussed in Chapter 18 and in Part V. We begin. in Section 16.B. by specifying the formal model of an economy to be studied here and for the rest of Part IV. Its essential ingredients-commodities. consumers. and firms-we have already encountered in Part I. The remainder of Section 16. B introduces the main concepts that will concern us throughout the chapter. We define first the normative notion of a Pareto optimal al/ocatiol1. an allocation with the property that it is impossible to make any consumer betler off without making some other consumer worse off. Then. we present two notions of price-taking equilibrium: Walrasian (or competitive) equilibrium. and its generalization. a price equilibriulH wit II transfers. The Walrasian equilibrium concept applies to the case of a private ownership economy. in which a consumer's wealth is derived from her ownership of endowments and from claims to profit shares of firms. The more general notion of a price equilibrium with transfers allows instead for an arbitrary distribution of wealth among consumers. The remaining sections of the chapter are devoted to exploring the relationships between these equilibrium concepts and Pareto optimality. Section 16.C focuses on the statement of the (very weak) conditions implying that every price equilibrium with transfers (and. hence. every Walrasian equilibrium) results in a Pareto optimal allocation. This is the first jrmdamcl1Ialtheorem of welfarc eco/lomics. a formal expression for competitive market economies of Adam Smith's claimed "invisible hand" property of markets. In Section 16.D. we study the converse issue. We state conditions (convexity assumptions arc the crucial ones) under which every Pareto optimal allocation can be
more abundant in country A.
545
s
546
CHAPTER
1&:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
PRQPERTIES
supported as a price equilibrium with transfers. This result is known as the second fundamental theorem of welfare economics. It tells us that if its assumptions are satisfied, then through the use of appropriate lump-sum wealth transfers, a welfare authority can, in principle, implement any desired Pareto optimal allocation as a price-taking equilibrium. We also discuss the practical limitations of this result. In Section 16.E, we introduce the problem of maximizing a social welfare function and relate it to the Pareto optimality concept. We uncover a close formal relationship between these two notions of welfare optimality. Section 16. F reexamines the Pareto optimality concept and associated results by making dilTerentiability assumptions and analyzing first-order conditions. There we sec how equilibrium prices can be interpreted as the Lagrange multipliers, or shadow prices. that arise in the associated Pareto optimality problem. Section 16.G discusses several applications of the concepts and results previously developed. We first present some examples that rely on particular interpretations of the L abstract commodities; one of them concerns the case of public goods. We then consider an application of our results to a world with nonconvex production sets, which leads to a brief exposition of the theory of marginal cost pricing. Appendix A deals with some technical issues concerning the boundedness of the set of feasible allocations and the existence of Pareto optima. Classical accounts of the material at the heart of this chapter are given by Koopmans (1957). Debreu (1959), and Arrow and Hahn (1971).
16.B The Basic Model and Definitions I n this chapter, we study an economy composed of 1 > 0 consumers and J > 0 firms in which there are L commodities. These L commodities can be given many possible interpretations; we discuss some examples in Section 16.G. Each consumer i = I, ... ,I is characterized by a consumption set Xi c: RL and a preference relation )::, defined on Xi' We assume that these preferences are rational (i.e., complete and transitive). Chapters 1 to 3 provide an extensive discussion of these concepts. Each firm j = I, ... , J is characterized by a technology, or production set, Y; c RL. We assume that every Y; is nonempty and closed. See Chapter 5 for a discussion of production sets and their properties. The initial resources of commodities in the economy-that is, the economy's elldowments-are given to us by a vector iiJ = (w" ... ,iiJd E RL. Thus, the basic data on preferences, technologies, and resources for this economy are summarized by ({(Xi' )::i)}!=" {lj}f= I' w). The Edgeworth box pure exchange economy discussed in Section 15.B, for example, corresponds to the case in which L = 2, 1 = 2, Xl = X, = RI;., J = I, and Y, = -R~ (the disposal technology). More generally, we say that an economy is a p"re exchallge ecollomy if its only technological possibility is that of free disposal, that is, if Y; = -R';. for allj = 1, ... ,J. Definition 16.B.1: An allocation (x, y) = (x" ... ,Xl' y, • ... 'YJ) is a specification of a consumption vector Xi E Xi for each consumer i = 1.... ,I and a production vector Yi E ~. for each firm i = 1, ... ,J. An allocation (x, y) is feasible if
SEC T ION
I i Xli =
WI
+ I i Yli for
every commodity
11. B:
t.
THE
BAS I C
MOD E LAN D
That is, if
LXi = W + LYi'
(16.B.1)
We denote the set of feasible allocations by A = {(x, y)
E
X,
X .•• X
X,
X
Y,
X • "
x ~:
LXi
= W+
i
LYJ} c: RL(/+J). J
The notion of a socially desirable outcome that we focus on is that of a Pareto optimal allocation. DefinitIon 16.B.2: A feasible allocation (x, y) is Pareto optimal (or Pareto efficient) if there is no other allocation (x', V') E A that Pareto dominates it, that is. if there is no feasible allocation (x', y') such that xi )::iXi for all i and xi >-iXi for some i. An allocation is Pareto optimal if there is no waste: It is impossible to make any consumer strictly better off without making some other consumer worse off. Note that the Pareto optimality concept does not concern itself with distributional issues. For example. in a pure exchange economy, an allocation that gives all of society's endowments to one consumer who has strongly monotone preferences is necessarily Pareto optimal. In Appendix A, we provide conditions on the primitives of the economy implying that the set offeasible allocations A is nonempty, closed, and bounded and that Pareto optimal allocations exist.
Private Ownership Economies Throughout Part IV, we study the properties of competitive private ownersltip economies. In such economies, every good is traded in a market at publicly known prices that consumers and firms take as unaffected by their own actions. Consumers trade in the marketplace to maximize their well-being, and firms produce and trade to maximize profits. The wealth of consumers is derived from individual endowments of commodities and from ownership claims (shares) to the profits of the firms, which are therefore thought of as being owned by consumers. I Formally, consumer i has an initial endowment vector of commodities Wi E R / and a claim to a share Oij E [0, I] of the profits of firm j (where iiJ = I i Wi and Li Oij = I for every firm j). Thus, the basic preference, technological, resource, and ownership data of a private ownership economy are summarized by ({(Xi' )::i)l!=I'
{Y;}f=,. {(wi,O", ... ,Ou)}!=,)· The notion of a price-taking equilibrium for a competitive private ownership economy is that of a Walrasiall equilibrium. Definition 16.6.3: Given a private ownerShip economy specified by ({(Xi, )::i)}!-I' {li)i'." {(Wi' 0" ... . , OjJl}I_,), an allocation (x', yO) and a price vector p = (p" ... , ptl constitute a Walrasian (or competitive) equilibrium if: (i) For every j,
Yi
lj; that is, for all Yi E lj.
maximizes profits in
P'Yi ~ p'Yi
Rec
DE FIN I T ION S
547
548
CHAPTER
11:
EQUILIBRIUM
(ii) For every i.
AND
xf is maximal for
ITS
~i
BASIC
WELFARE
s
PROPERTIES
in the budget set
~ C T ION
1 •. C:
THE
FIR S T
FUN 0 A MEN TAL
THE 0 REM
0 F
W ELF ARE
16.C The First Fundamental Theorem of Welfare Economics
2 {XieXi:P'Xi 5,P'wi + L OiiP•Yil·
i
(iii)
The first fundamental theorem of welfare economics states conditions under which any price equilibrium with transfers. and in particular any Walrasian equilibrium. is a Pareto optimum. For competitive market economies. it provides a formal and very general confirmation of Adam Smith's asserted "invisible hand" property of the market. A single, very weak assumption, the local nonsa/ialian of preferences (see Section 3.13). is all that is required for the result. Notably. we need not appeal to any convexity assumption whatsoever. Recall the definition of locally nonsatiated preferences from Section 3.13 (Definition 3.B.3).
Lxi = W + L yj. . i
Condition (i) of Definition 16.13.3 says that at a Walrasian equilibrium, firms are maximizing their profits given the equilibrium prices p. The logic of profit maximization is examined extensively in Chapter 5. Condition (ii) says that con· sumers are maximizing their well-being given. first. the equilibrium prices and, second. the wealth derived from their holdings of commodities and from their shares of profits. See Chapter 3 for extensive discussion of preference maximization. Finally. condition (iii) says that markets must clear at an equilibrium; that is. all consumers and firms must be able to achieve their desired trades at the going market prices.
Definition 16.C.1: The preference relation ~i on the consumption set Xi is locally nonsaliated if for every Xi E Xi and every c > O. there is an xi e Xi such that II x; - x,1I 5, r. and xi >-, Xi'
Price Eqllilihria witl! Trallsfers The aim of this chapter is to relate the idea of Pareto optimality to supportability by means of price-taking behavior. To this end. it is useful to introduce a notion of equilibrium that allows for a more general determination of consumers' wealth levels than that in a private ownership economy. By way of motivation. we can imagine a situation where a social planner is able to carry out (lump-sum) redistributions of wealth, and where society'S aggregate wealth can therefore be redistributed among consumers in any desired manner.
Intuitively. the local nonsatiation condition will be satisfied if there are some desirable commodities. Note also a significant implication of the condition: if ~, is continuous and locally nonsatiated. then any closed consumption set X, must be unbounded. Otherwise, there would by necessity exist a global (hence. local) satiation point (see Exercise 16.CI). Proposition 16.C.1: (First Fundamental Theorem of Welfare Economics) If preferences are locally nonsatiated, and if (x·, y., p) is a price equilibrium with transfers, then the allocation (x·, Y·) is Pareto optimal. In particular, any Walrasian equilibrium allocation is Pareto optimal.
Definition 16.0.4: Given an economy specified by ({(Xi. ~i)}!~" (lj}f-" iii) an allocation (x·. Y·) and a price vector p = (p" ... ,ptl constitute a price equilibrium with transfers if there is an assignment of wealth levels (w" ... , WI) with Li Wi = p'w + Li poyi such that (i) For every j.
Yi
maximizes profits in lj; that is,
(ii) For every i,
xi
is maximal for
for all ViE lj.
P'Yi 5,P'y! ~i
in the budget set
{xieXi:p'x i (iii)
Proof: Suppose that (x·. y., p) is a price equilibrium with transfers and that the associated wealth levels are (11'" •••• 11',). Recall that L, w, = p'w + LiP·yj. The preference maximization part of the definition of a price equilibrium with transfers [i.e., part (ii) of Definition 16.B.4] implies that
5,
then P'x, >
Wi}'
i
The concept of a price equilibrium with transfers requires only that there be some wealth distribution such that allocation (x', y.) and price vector peRL constitute an equilibrium. It captures the idea of price-taking market behavior without any supposition about the determination of consumers' wealth levels. Note that a Walrasian equilibrium is a special case of an equilibrium with transfers. It amounts to the case in which. for every i, consumer ts wealth level is determined by the initial endowment vector w, and by the profit shares (Oil' ... , 0,,) without any further wealth transfers, that is, where 11', = P'w, + LjOijP·yj for all i = 1•...• 1.
IfXi~iX[
then P'x, ~
Wi'
(16.C2)
That is, anything that is at least as good as xr is at best just affordable. This property is easily verified (you are asked to do so in Exercise 16.C2). Now consider an allocation (x. y) that Pareto dominates (x'. y.). That is. x, ~,xr for all i and x, >-, xr for some i. By (16.C2), we must have P' x, ~ w, for all i, and by (16.CI) P' x, > w, for some i. Hence,
L, P'x, > L,
w, = p'w
+ L p.y;, j
2. The terminology" Xi is maximal for ;::i in set B" means that x, is a prderence-maximizing choice for consumer; in the set B; that is, ."1 E B and i ;2:/ for all E B.
x x;
(16.CI)
That is, anything that is strictly preferred by consumer i to xr must be unaffordable to her. The significance of the local nonsatiation condition for the purpose at hand is that with it (16.CI) implies an additional property:
Lxi = W + L yii
Wi'
x;
Morcover, because yj is profit maximizing for firm j at price vector p,
=
E CON 0 M I C S
549
550
CHAPTER
11:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
SECTION
PROPERTIES
16.0:
THE
SECOND
FUNDAMENTAL
THEOREM
OF
WELFARE
16.D The Second Fundamental Theorem of Welfare Economics The second fundamental welfare theorem gives conditions under which a Pareto optimum allocation can be supported as a price equilibrium with transfers. It is a converse of the first welfare theorem in the sense that it tells us that, under its assumptions, we can achieve any desired Pareto optimal allocation as a market-based equilibrium using an appropriate lump-sum wealth distribution scheme. The second welf~re theorem is more delicate than the first, and its validity requires additional assumptions. To see this, reconsider some of the examples discussed in Ch-i xt, then p' Xi> w:') is replaced by the weaker requirement that anything preferred to xt cannot cost less than w· (i.e., "if Xi>-'X,~, then p'x, ~ w:'). '
Figure 16.C.1 A price equilibrium with transrers that is not a Pareto optimum.
we have p'w
+ Lj P'yj
~ p'w
+ Lj P·Yj· Thus,
L P'X
i
> p'w +
L P·Yj·
(16.C.3)
But then (x, y) cannot be feasible. Indeed, Li Xi = W + Lj Yj implies Li P' Xi = P'W + P' J'j' which contradicts (16.C.3). We conclude that the equilibrium allocation (x*, y*) must be Pareto optimal. _
L
The central idea in the proof of Proposition 16.C.1 can be put as follows: At any feasible allocation (x, Y), the total cost of the consumption bundles (x I' ... ,x,), evaluated at prices P, must be equal to the social wealth at those prices, P' W + L} P' Yj' Moreover, because preferences are locally nonsatiated, if (x, Y) Pareto dominates (x*, yO) then the total cost of consumption bundles (Xl' ... ' x,) at prices p, and therefore the social wealth at those prices, must exceed the total cost of the equilibrium consumption allocation p'(L x1j = p'w + Lj p' yj. But by the profitmaximization of Definition 16.B.4, there are no technologically feasible production levels that attain a value of social wealth at prices P in excess of P' W + Lj P' yj. The importance of the nonsatiation assumption for the result can be seen in Figure 16.C.I, which depicts an Edgeworth box where local nonsatiation fails for consumer I (note that consumer I's indifference "curve" is thick) and where the allocation x>, a price equilibrium for the price vector P = (PI' pz) (you should verify this), is not Pareto optimal. Consumer I is indifferent about a move to allocation x, and consumer 2, having strongly monotone preferences, is strictly better otT. (See Exercise 16.C.3 for a first welfare theorem compatible with satiation.) Two points about Proposition 16.C.1 should be noted. First, although the result may appear to follow from very weak hypotheses, our theoretical structure already incorporates two strong assumptions: ulliversal price quolillg of commodities (market completeness) and price taking by economic agents. In Part III, we studied a number of circumstances (externalities, market power, and asymmetric information) in which these conditions are not satisfied and market equilibria fail to be Pareto optimal. Second, the first welfare theorem is entirely silent about the desirability of the equilibrium allocation from a distributional standpoint. In Section 16.0, we study the second fundamental theorem of welfare economics. That result, a partial converse to the first welfare theorem, gives us conditions under which any desired distributional aims can be achieved through the use of competitive (price-taking) markets.
Definition 16.~.1~ Given an economy specified by ({(Xi' ;::i)}!-1' P~-lf-1' W) an allocalion (~ • V ). and a price vector p = (P" ... ,PL) # 0 constitute a price quasieqUlhbnum with transfers if there is an assignment of wealth levels (W,' ... , w,) with LiWi = p·w + LiP'Vi" such that
m
ECONOMICS
551
552
CHAPTER
11:
EQUILIBRIUM
(i) For every
i.
AND
ITS
WELFARE
PROPERTIES
SECTION
yt maximizes profits in If; that is. P'Yj ~ p·yt for all Yj Elf·
(ii) For every i. if x;>-;xi then p'x; ~ (iii)
BASIC
is,
always make. we must have p' x; =
Xi
2= p'
=~
Y=
THEOREM
OF
WELFARE
x~,
ECONOMICS
553
that
v. = {~x, E RL
X,
E V" ... ,
x, v,} E
Wi
~ lj = {~YjER":)"
E
Y" .. . ,J'JE YJ}'
Thus, V is the set of aggregate consumption bundles that could be split into I individual consumptions, each preferred by its corresponding consumer to x,*. The set l' is simply the aggregate production scI. Note that the set Y + {(v}, which geometrically is the aggregate production set with its origin shifted to (V, is the set of aggregate bundles producible with the given technology and endowments and usable, in principle, for consumption. St('l'l: Ev('ry s('t v; iscOlw('.t. Suppose that x, >-, x: and x; >-, x:' Take 0 $ (% !> I. We want to prove that <Xx; + (I - Cl)X;~, Because preferences are complete, we can assume without loss of generality that Therefore, by convexity of preferences, we have ax, + (I - a),,; ~,x;, which by transitivity yields the desired conclusion: ax, + (I - (X)x; >-, x,* [recall part (iii) of Proposition I.B.I).
x:.
Jor ever)' i. This means that we could just as well not
mention the w,'s explicitly and replace part (ii) of Definition 16.D.1 by
xr then p'
lUNOAMt.hTAl
and
j
Note also that when consumers' preferences are locally nonsatiated, part (ii) of Definition 16.D.1 implies p'x! ~ w, for every i.l In addition, from part (iii), we get L,P'x; = P'w + Lj r' y; = L, II',. Therefore, Imder Ihe assumption of focally /J(J/Isaliat.d preferences, which we
If ."(/ >i
,;,fCONO
We begin by defining, for every i, the set V. of consumptions prererred to = {x, E X,: x, >-, x~} c RL. Then define V
Part (ii) of Definition 16.0.1 is implied by the preference maximization condition of the definition of a price equilibrium with transfers [part (ii) of Definition 16.B.4]: If xi is prererence maximizing in the set {Xi E X,: P' X, ~ w,j, then no x, >-, xi with p' Xi < Wi can exist. Hence, any price equilibrium with transfers is a price quasiequilibrium with transfers. However, as we discuss later in this section, the converse is not true.
(ii')
THL
V.
Wi'
'LA = w + L yr ;
16.0:
x;.
That is, allocation (x·, y.) and price vector p constitute a price quasiequilibrium with transfers if and only if conditions (i), (ii'), and (iii) hold' Moreover, with locally nonsatiated is expenditure minimizing on the set preferences, condition (ii') is equivalent to saying that {x, E X: x, ;::,xn (see Exercise 16.D.I). Thus, our discussion later in Ihis section of the conditions under which a price quasiequilibrium with transfers is a price equilibrium with transfers can be interpreted in the locally non satiated case as providing conditions under which expenditure minimization on the set {x, e X,: x,;::, xi I implies preference maximization on the set {x,eX,: P'x,~p'xn = {x,eX,: P'x,!> w,l . •
x,;::, x;.
SI~p 2: Tire sels Valid Y + {w} are cO/wex. This is just a general, and easyto-prove, mathematical fact: The sum of any two (and therefore any number of) convex sets is convex.
xr
Slep 3: V f"\ (Y + {w}) = 0. This is a consequence of the Pareto optimality of (x·, y.). If there were a vector both in Vand in Y + {w}, then this would mean that with the given endowments and technologies it would be possible to produce an aggregate vector that could be used to give every consumer i a consumption bundle that is preferred to
x:.
Proposition 16.0.1 states a version of the second fundamental welfare theorem.
St~p
Tlrer~ is P = (PI' ... ' 1',) '" 0 alld a lIumber r Stlclr Ilral p' z ;>: r for every p'Z S r for every Z E Y + {w}. This follows directly from the separating hyperplane theorem (see Section M. G. the Mathematical Appendix). It is illustrated in Figure 16.0.1.
Proposition 16.0.1: (Second Fundamental Theorem of Welfare Economics) Consider an economy specified by ({(X;, ~;)li-" {Y;}t-" iii). and suppose that every Y; is convex and every preference relation ~; is convex [i.e., the set {X; E X;: X; ~j Xj} is convex for every x; E X;] and locally nonsatiated. Then, for every Pareto optimal allocation (x*, y*). there is a price vector p = (p" ... ,pd '" 0 such that (x·, y*. p) is a price quasiequilibrium with transfers.
4:
=E V IIl1d
Figure 16.0.1
The separation argument in the proof of the second welfare
Proof: In its essence, the proof is just an application of the separating hyperplane theorem for convex sets (see Section M.G. of the Mathematical Appendix). To facilitate comprehension, we organize the proof into a number of small steps.
theorem.
3. To see this, observe that if preferences are locally nonsalialed and p' x7 < Wi) then close to Xi with Xi >-ix7 and p' Xi < Wj. contradicting condition (ii) of Definition 16.D.1. 4. A similar observation applies, incidentally, 10 the definition of price equilibrium with transfers
x7 there is an
(Definition 16.B.4). If preferences are locally nonsatiated, we get an equivalent definition by not referring explicitly to the w:s and replacing part (ii) of the definition by (ii"): If x, >-, x~ then P'x j > P'x7- Thus. in this locally nonsatiated case, condition (ii") says that x~ is preference maximizing on {XI E Xi: p·.'t l S p' x7}.
x,
=
554
CHAPTER
16:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
PROPERTIES
SECTION
16.0:
THE
SECOND
FUNDAMENTAL THEOREM
OF
WELFARE
Budget
Seep 5: If Xi ?::" Xi for every j then p' (L, x,) ;:: r. Suppose that x, :::, xi for every
A price quasi-equilibrium that
x,
is not a price equilibrium.
Step 6: p' (Li xi) = P' (w + Lj yj) = r. Because of step 5, we have P' O::i xi) ~ r. On the other hand, Li xi = Lj yj + W E Y + (wi, and therefore P'(Li xi) s: r. Thus, p' (Li = r. Since Li xi = w + Lj yj, we also have P' (w + Lj yj) = r.
}<:,
xn
\
\
XI \
Step 7: For every j, we have P' Yj s: P' yj for all Yj E lj. For any firm j and )'j E ~, we have Yj + L ... J E Y. Therefore,
y:
P'(w+YJ+
L
y:)s:r=p,(w+Yj+
It t'j
Hence, P'Y)
L y:). ."i
0,
s: p'yj. Proposition 16.0.2: Assume that X, is convex and ?::'i is continuous. Suppose also that the consumption vector xi E Xi' the price vector p, and the wealth level Wi are such that Xi >- i xi implies p' Xi;:: Wi' Then, if there is a consumption vector x; E Xi such that p·x; < Wi [a cheaper consumption for (p, Wi))' it follows that 6 X/'>iX: implies p·x,.> Wi.
Step Ii: For every i, if Xi >-i"!' chell P'x, ~ p·x{. Consider any Xi>-IX{. Beca use of steps 5 and 6, we ha ve P{Xi +
'~i xi) ~
I'
= p.(X!
+
'~I xi).
Proof: The idea of the proof is indicated in Figure 16.0.3 (where we take p'xi = Wi only because this is the leading case; the fact plays no role in the proof). Suppose that, contrary to the assertion of the proposition, there is an Xi >- I xi with p' Xi = Wi' By the cheaper consumption assumption, there exists an x; E XI such that P' x; < lVi. Then for all IX E [0, I), we have rzx i + (I - rz)x; E Xi and P·(tXXl + (1 - IX)X;) < WI" But if rt is close enough to I, the continuity of ?::'i implies that lXX, + (I - IX)X; >- i xi, which constitutes a contradiction because we have then found a consumption bundle that is preferred to xi and costs less than lVi. • Note that in the example of Figure 16.0.2, we have WI = 0 in the price quasicquilibrium supporting allocation x·, and so there is no cheaper consumption for (p, w.)." As a consequence of Proposition 16.0.2, we have Proposition 16.0.3.
Step 9: The lI'ealth levels IVI = p'xi for i = I, ... , I support (x·, y., p) as a price 411asic(luilihrium lVith transfers. Conditions (i) and (ii) of Definition 16.0.1 follow from steps 7 and 8; condition (iii) follows from the feasibility of the Pareto optimal allocation (x·, y.) . • In Exercise 16.0.2, you are asked to show that the local nonsatiation condition is required in Proposition 16.0.1. When will a price quasiequilibrium with transfers be a price equilibrium with transfers? The example in Figure 15.B.IO(a), reproduced in Figure 16.0.2, indicates that there is indeed a problem. Figure 16.0.2 depicts the quasiequilibrium associated with the Pareto optimal allocation labeled The unique price vector (normalizing PI = I) that supports x· as a quasiequilibrium allocation is p = (1,0); the associated wealth levels are WI = p·xt = (I,O)'(O,x!,) = 0 and W2 = P·x!. However, although the consumption bundle xt satisfies part (ii) of Definition 16.0.1 (indeed, p' x I 2: 0 = WI for any XI ~ 0), it is not consumer I's preference-maximizing bundle in her budget set {(x",x21)ER~:(1,0)'(x",x21)S:0} = {(XII'X21)ER~:
x·.
Xli
=
555
Figure 16.0.2 (lett)
Line
i. By local nonsatiation, for each consumer j there is a consumption bundle Xi arbitrarily close to Xi such that >-, Xi' and therefore Xi E V;. Hence, Li Xi E V, and so p'(L :li) 2: r, which, taking the limit as Xi -+ Xi' gives P'(L, Xi) ~ r. s
ECONOMICS
Proposition 16.0.3: Suppose that for every i, Xi is convex, 0 E Xi' and :::i is continuous. Then any price quasiequilibrium with transfers that has (w" ... , wd »0 is a price equilibrium with transfers.
O}.
An important feature of the example just discussed, however, is that consumer I's wealth level at the quasiequilibrium is zero. As we shall see, this is key to the failure of the quasiequilibrium to be an equilibrium. Our next result provides a sufficient condition under which the condition MXi>-IXi implies P'X i ~ lV i" is equivalent to the preference maximization condition" Xi >-i xi implies p' Xi > Wi'"
6. If, as in all our applications. ~j is locally nonsatiated and Wj = p·xi. then Proposition 16.0.2 olTers 5ufHcienl conditions for the eq uivalence of the statements" x7 minimizes expenditure relative to p in the s~l {''(j E Xj: X,~, x;}" and" xi is maximal for 2:, in the budget set lXiE X,: P·X . ~ p·xj}." 7. A similar argument can be used to show that if X, is convex and the Walrasian demand function xj(p, \\'j} is well defined, then there is a cheaper consumption for (P. w,) if and only if there is an x; arbitrarily close to xl(p. Wi) with p' x; < lVi . In the Appendix A, of Chapter 3 the latter concept was called the locall.v cheaper consumption condition. 8. Note also that Proposition 16.0.2 generalizes the result in Proposition 3.E.t(ii), which assumed local nonsatiation. Wi = p'x7 > O. and Xj = R';.
5. Geometrically. what we have done here is show that the set 1:1 {Xi E Xj: Xi;::j x7} is contained in the closure of V (see Section M.F of the Mathernalieal Appendix for this concepl), which, in turn. is contained in the half-space (v E ilL: p' v ~ r).
m
Ftgure 16.0.3 (right)
Suppose there exists a "cheaper consumption" (an x; E Xi such that p'x; < WI)' Then if the preferred set does intersect the budget set (p' XI :S WI for some xi>jx:), it follows that the preferred set does intersect the interior of the budget set (p·x; < \Vi for some
Xi >-IXn.
CHAPTER
11:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
PROPERTIES
Consider the implications of Proposition 16.0.3 for a pure exchange economy in which W »0 and every consumer has X, = R~ and continuous, locally nonsatiated preferences. In such an economy, by free disposal and profit maximization, we must have p ~ 0 and P ~ 0 at any price quasiequilibrium 9 Thus, under these assumptions, any price quasiequilibrium with transfers in which x~ »0 for all i is a price equilibrium with transfers (since then = p' x,' > 0 for all iJ. But there is more. Suppose that, in addition, preferences are strongly monotone. Then we must have p » 0 in any price quasiequilibrium with transfers. To see this, note that p ~ 0, p ~ 0, and w » 0 imply that L, w, = p' 'v > 0 and therefore that w, > 0 for some i. But by Proposition 16.0.2, this consumer must then be maximizing her preferences in her budget set {x, E R~: P' x, :0;; w,}, which, by strong monotonicity of preferences, cannot occur if prices arc not strictly positive. Once we know that we must have p »0, we can conclude that0 and Proposition 16.D.2 applies. On the other hand, if x~ = 0, then 1\', = 0 and the result follows from the fact that x~ = 0 is the only vector in the set {x, E R'; : p' .', :s; 0:. (Exercise 16.D.3 asks you to extend the arguments presented in this paragraph to the case of an economy with production.)
w,
The second welfare theorem (combined with Propositions 16.0.2 and 16.0.3) identifies conditions under which any Pareto optimal allocation can be implemented through competitive markets and oITers a strong conceptual anirmation of the use of competitive markets, even for dealing with distributional concerns. Yet, it is important to discuss some of the practical limitations on the usc of this theoretical result. The first observation to make is that a planning authority wishing to implement a particular Pareto optimal allocation must be able to insure that the supporting prices (PI"'" pd will be taken as given by consumers and firms. If the market structure is such that price-taking behavior would not automatically hold (say, because economic agents are not all of negligible size), then the planning authority must somehow enforce these prices, either by monitoring all transactions or, perhaps, by credibly offering to buy or sell any amount of any good t at price p,. A second observation is that the information of a planning authority that wants to use the second welfare theorem must be very good indeed. To begin with, it must have sufficiently good information to identify the Pareto optimal allocation to be implemented and to compute the right supporting price vector. For this purpose, the authority must know, at least, the statistical joint distribution of preferences, endowments, and other relevant characteristics of the agents that actually exist in the economy. However, to implement the correct transfer levels for each consumer, the planning authority must know more: it must have the ability to tell who is who by observing each individual's private characteristics (e.g., preferences and endowments) perfectly. Such information is extremely unlikely to be available in practice; as a result, most common transfer schemes fail to be lump-sum schemes. For example, if the planning authority wants to transfer wealth from those who have a great deal of a highly valuable labor skill to those who do not, the only way it may have to tell which consumers are which may be by observing their aellla/
SEC T , 0 N
16 •D,
THE
SEC 0 N D
FUN DAM E N TAL
D F
W ELF A A E
earnings. But if transfers are based on observed earnings, they will cease to be lump-sum in nature. Individuals will recognize that by altering their earnings, they will change their tax burden. Finally, even if the planning authority observes all the required information, it must actually have the power to enforce the necessary wealth transfers through some tax-and-transfer mechanism that individuals cannot evade. Because of these informational and enforceability limitations, it is in practice unlikely that extensive lump-sum taxation will be possible.'o We shall sec in Section 18.0 that if these types of transfers arc not possible, then the second welfare theorem collapses in the sense that, for a typical economy, only a limited range of Pareto optima arc supportable by means of prices supplemented by the usual sort of taxation systems. For the typical economy, redistribution schemes arc distortiOlwr)'; that is, they trade off distributional aims against Pareto optimality. The analysis of this trade-oIT is the subject of second-best welfare economics, some clements of which arc presented in Chapter 22. (Chapter 23 discusses in much more depth what is and what is not implementable by a planning authority who faces informational and enforceability constraints.) In summary, the second welfare theorem is a very useful tl,eorl'lica/ reference point. But it is far from a direct prescription for policy practice. On the contrary, by pointing out what is necessary to achieve any desired Pareto optimal allocation, it serves a cautionary purpose. It is clear from our discussion that convexity plays a central role in the second welfare theorem. But it is a role that deserves a very important qualification. The interpretation of the second welfare theorem is at its strongest when the number of economic agents is large. This is so because the price-taking assumption is then enforced by the market itself (otherwise, it is almost inescapable that there must be some sort of centralized mechanism guaranteeing the fixity of prices). It turns out, however, that if consumers are numerous (in the limit, a continuum), and if the nonconvexities of production sets are bounded in a certain sense, then the assumptions of convex preferences and production sets are noC required for the second welfare theorem.
To see the idea behind this, it is useful to consider the one-consumer economy depicted in Figure 16.D.4 where, because of nonconvexities, the (trivially Pareto optimal) allocation x, = 'v cannot be price supported. Suppose, however, that we replicate the economy so that we have CIVO consumers and the total endowments are doubled to 2w. Again, the allocation x, = x, = 0) cannot be price supported, but now chis s),mmfcric al/ocacion is no longer a Pareco IIp'i,,,,,,,,. In Figure 16.D.5, we can see that the asymmetric allocation
x; = 'v +(1, -I)
and
xi =w + (-1,1)
Pareto dominates x, = x, = w.lt certainly does not follow from this that with just two replicas any allocation that cannot be price supported is not a Pareto optimum. (In Figure 16.0.6, we represent the Edgeworth box associated with Figure 16.0.5. We have drawn it so that the allocation x' is actually Pareto optimal and yet cannot be price supported). However, what 10. Note that the extent of the required lump-sum taxation depends not only on the final wealth levels (w •....• w,) but also on the initial situation. For example. if we are in private ownership Ojjp·yj and so depends on the economy. lhen the net transfer to consumer; is Wi - P'Wlconsumer's initial endowments. The Walrasian equilibria correspond to the no·taxation situations; and Ihe farther from the Walrasian wealth levels that we try to go, the larger the required transfer.
rJ
9. Indeed, if we had P, < 0, then unboundedly large profits could be generated through disposal of the (th good.
THE 0 A E M
E CON 0 M , C S
557
558
CHAPTER
16:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
SECTION
PROPERTIES
H.E:
PARETO
OPTIMALITY
AND
SOCIAL
WELFARE
OPTIMA
559
",
Ftgure 16.0.4 (top left)
'~.',
Ii)
= '"
+ (I, -I)
+I
~---'-----"
A one-consumer economy (without produclion) where the initial endowment is not supportable by prices.
,
Allocation (x;, x;) Pareto dominates allocation (v,
J,i,-I
,",,{~
I
Figure 16.0.6 (bottom)
Allocation x' = (x;, x2) is Pareto optimal but not price supportable.
0, ~,_-Y-O-~ (I}-I
In this section, we discuss the relationship between the Pareto optimality concept and the maximization of a social welfare function (see Section 4.D and Chapter 22 for more on this concept). Given a family ui (·) of (continuous) utility functions representing the preferences <:i of the J consumers. we can capture the attainable vectors of utility levels for an economy specified by ({(Xi' <:;:;)}f." { lj} f." Iii) by means of the utility possibility set:
U = {(u" ...• u,) E IR': there is a feasible allocation !Ii S;
(x, y)
The utility possibility set.
u
",
",
UP = {(II" ... , II,) E U: there is no (u'" ... , !Ii) E U such that ui ~ Ui for all i and ui >
IIi
for some i}.
Proposition 16.E.1 is then intuitive.
I
Proposition 16.E.1: A feasible allocation (x. y) = (x" ... , x,, Y, • ...• YJ) is a Pareto optimum if and only if (U,(x,). ...• u,(x,)) E UP.
»
can be shown is that if the number of replicas is large enough, then any feasible allocation that fails (significantly) to be price supportable can be Pareto dominated, and therefore any Pareto optimum must be (almost) price supportable. (See Exercise 16.0.4 for more on this.)"
16,E Pareto Optimality and Social Welfare Optima
~
Figure 16.E.I depicts this set for a two-consumer economy. (Note that we show U c R' as a closed set; Appendix A discusses sufficient conditions to guarantee that the set is indeed closed.) By the definition of Pareto optimality. the utility values of a Pareto optimal allocation must belong to the boundary of the utility possibility set. I 2 More precisely. we define the Pareto /rolllier UP. also shown in Figure 16.E.I. by
Ftgure 16.0.5 (top rtght)
0,
U
Ftgure 16.E,1 (left)
II
•
I
Proof: If (u, (x ,)•...• II, (XI f UP. then there is (U'I' .... u;) E U such that uj ;;:: U,{Xi) for all i and ui > u,(x i ) for some i. But (u; •. . "ui) E U only if there is a feasible allocation (x', y') such that ui(xi) ;;:: ui for all i. It follows then that (x'. y) Pareto dominates (x. y). Conversely. if (x. y) is not a Pareto optimum. then it is Pareto dominated by some feasible (x', y'). which means that Ui{Xi) ~ u,(x i ) for all i and ui(xi) > "i(Xi) for some i. Hence. (u,(x,), ...• u,(x,» f UP . • We also note that if every Xi and every lj is convex. and if the utility functions IIi ( .) arc concave, then the utility possibility set U is convex (see Exercise 16.E.2).1l One such utility possibility set is represented in Figure 16,E.2. Suppose now that society's distributional principles can be summarized in a social welfare /ullctioll W(u" ... , u,) assigning social utility values to the various possible vectors of utilities for the J consumers. We concentrate here on a particularly simple class of social welfare functions: those that take the linear form W(u, •...• u,) =
such that
L i.;Ui
ui(x,) for i = I•... , J}. 12. However, nol every poinl in the boundary must be Pareto optimal. Go back, for example, to Figure 16.C.1: The ulility values associated with x· belong to the boundary of the utility possibility set because it is impossible to make both consumers better off. Yet. x· is not a Pareto optimum. 13. II can be shown that under a mild technical strengthening of the strict convexity assumption on preferences (essentially the same condition used to guarantee differentiability of the Walrasian demand function in Appendix A of Chapter 3), there are in the family of utility functions ",(.) that represent ;::j some utility functions that are not only quasiconcave but also concave.
II. Two faclS eSiablished in Chapter 17 lend plausibility to this claim. First, in Section 17.1, WI! show that convexity is not required for the (approximate) existence or a Walrasian equilibrium In a l'lrgc economy. Second. in Section I7.C, we argue that the second welfare theorem can be rcphnlscd as an assertion of the existence or a Walrasian eqUilibrium ror economies in which endowments are distributed in a particular manner, and it can therefore be seen as implied by the conditions guaranteeing the general existence of Walrasian equilibria.
I
Figure 16.E.2 (right)
A convex utility possibilily sel.
560
CHAPTER
16:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
SEC T ION
PROPERTIES
",
1 I. F!
FIR S T - 0 R DE R
CON 0 I T ION S
FOR
P A A E TOO P TIM A LIT y
561
x,
Figure 16.E.4 Ftgur. 16.E.3
Maximizing the utili! y or a representative consumer.
Maximizing a linear social welfare function.
for some constants i. = p., ..... i.,).'4 Lelling 1/ = (u, •...• U,). we can also write W(I/) = i. '1/. Because social welfare should be nondecreasing in the consumer's utility levels. we assume that i. :2: O. Armed with a linear social welfare function, we can select points in the utility possibility set U that maximize our measure of social welfare by solving Max w.u
i.·I/.
(16.E.I)
Figure 16.E.3 depicts the solution to problem (16. E.I). As the figure suggests. we ha ve the result presented in Proposition 16.E.2. Proposition 16.E.2: If u· = (u1, . ..• u7) is a solution to the social welfare maximization problem (16.E.1) with i.» O. then u· E UP; that is, u· is the utility vector of a Pareto optimal allocation. Moreover. if the utility possibility set U is convex. then for any ii = (ii, • ...• ii,) E UP, there is a vector of welfare weights i. = (i." ... , i.,):2: O. i. #< 0, such that i.·ii:2: ).·u for all UE U. that is. such that ii is a solution to the social welfare maximization problem (16.E.1). Proof: The first part is immediate: if u· were not Pareto optimal, then there would exist a /I E U with u:2: u' and u #< u*; and so because ).» 0, we would have i.'u>i.'fl*, For the second part, note that if ii E UP. then ii is in the boundary of U. By the supporting hyperplane theorem (see Section M.G of the Mathematical Appendix), there exists a i. #< 0 such that i.. ii :2: i.' u for all u E U. Moreover, since the set U has been constructed so that U - R'+ c U, we must have ). :2: 0 (indeed, if )" < O. then by choosing a u E U with u, < 0 large enough in absolute value, we would have i.·/1 > i.' ,i). • Proposition 16.E.2 tells us that for economies with convex utility possibility sets, there is a close relation between Pareto optima and linear social welfare optima: Every linear social welfare optimum with weights i. » 0 is Pareto optimal, and every Pareto optimal allocation (and hence. every Walrasian equilibrium) is a social welfare optimum for some welfare weights V." ... , i.,) :2: 0.15 14. See Chapter 22 for a discussion of more general types of social welfare functions. 15. The necessity of allowing for some i' l to equal zero in the second part of this statement parallels the similar feature encountered in the characterization of efficient production vectors in
Proposition 5.F.2.
As usual. in the absence of convexity of the set U. we cannot be assured that a Pareto optimum can be supported as a maximum of a linear social welfare function. The point ,i in Figure 16.E.1 provides an example where it cannot. fly using the social welfare weights associated with a particular Pareto optimal allocation (perhaps a Walrasian equilibrium), we can view the latter as the welrare optimum in a certain single-consumer. single-firm economy. To sec this, let (x·, y.) be a P.ucto optimal allocation and suppose that i. = 0.1 .... ' i.,)>> 0 is a vector of welfare weights supporting U at (ul(xn .... ,II,(xi)). Define then a utility runction u).(.X) on aggregate consumption vectors in
X=
L,
X;
C
R " by u,(.<) =
L A;U;(X;)
Max
(16.E.2)
I
(Jr\ ..... Jr/)
s.t.
XI E
Xi ror all i and
Li X; = i.
The utility function u,(·) is the (direct) utility function of a normative representative consumer in the sense discussed in Section 4.D (see, in particular. Exercise 4.D.4). Letting Y = LJ lj be the aggregate production set. the pair (LI x7 'LJ y;) is then a solution to the problem Max u,(.<) S.t.
x=
cii +
J.
x E X, YE
Y.
This solution is depicted in Figure 16.E.4. It is important to emphasize. however, that the particular utility function chosen for the representative consumer depends on the weights (i., •...• i., ) and therefore on the Pareto optimal allocation under consideration.
16.F First-Order Conditions for Pareto Optimality This section intends to be pedagogical. The emphasis is not on minimal assumptions, and its pace is deliberately slow. Making differentiability assumptions. we show how prices and the optimality properties of price-taking behavior emerge naturally from an examination of the first-order conditions associated with Pareto optimality problems. Along the way we redo, and in some aspects also generalize, the analysis in Sections 16.C and 16.0 on the two fundamental welfare theorems.'" 16. Early contributions along the line of the discussion in this section are Allais (1953), Lange
(1942). and Samuelson (1947).
562
CHAPTER
16:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
SECTtON
PROPERTIES
",
(2)
L, Xli :0; w, + Lj YIj
I = I ..... L
(3)
F;(Y,j •..• , YLj) :0; 0
j= I ... .• J.
OPTtMALITY
Utility Possibility Set Figure 16.F.1
consumer I's utility subject to the required utility constraint for consumer 2. By varying 2's required utility level, we trace out the set of Pareto optimal points. Under our assumptions (recall that Ui ~ 0 for i ~ 2), all the constraints of problem (16.F.I) will be binding at a solution. Denote by (0 2, ... ,Ii/) ~ 0, (Ph ... ,PL) ~ 0, and (y" ... , YJ) ~ 0 the multipliers associated with the constraints (I), (2), and (3) of problem (16.F.O. respectively, and define Ii, = I. [n Exercise 16.F.2, you are asked to verify that the first-order (Kuhn-Tucker) necessary conditions for problem (16.F.l) can be written as follows (all the derivatives, here and elsewhere in this section, are evaluated at the solution):'·
(16.F.I) i
PARETO
Parameterizing the frontier of the utility possibility set when 1 = 2 by the required utility level of consumer 2.
= 2..... 1
u,(x li •...• x,,;) ~ ii,
FOR
/
(x.Y)=(x, •...• xl.Y, •...• YJ)E!;I~' x R 'J
u,(x", ...• XL')
CONDITIONS
/
that solve the following problem:
s.t. (I)
FtRST-ORDER
",
To begin with, we assume that the consumption set of every consumer is !;I~ and that preferences are represented by utility functions u,(x,) that are twice continuously differentiable and satisfy Vu,(x,)>> 0 at all x, (hence, preferences are strongly monotone). We also normalize so that u,(O) = O. The production set of firm j takes the form Y; = {y E RL: Fj(Y) :0; OJ, where /-~(r) = 0 defines firm j's transformation frontier. We assume that Fj:!;IL --+ R is twice continuously differentiable, that Fi(O):o; 0, and that V F;(y) = (c1fif.l'J)/rly'j •...• of/Yj)/iJYLj)>> 0 for all YjE !;IL. The meaning of the last condition is that if fit)~ = O. so that Yj is in the transformation frontier of Y;, then any attempt to produce more of some output or use less of some input makes the value of Fj (') positive and pushes us out of Y; (in other words, YJ is production ellicient. in the sense discussed in Section 5.F. in the production set y;).17 Note that. for the moment, no convexity assumptions have been made on preferences or production sets. The problem of identifying the Pareto optimal allocations for this economy can be reduced to the selection of allocations
Max
".F:
Xli:
iJu 0, --- 1', { i
OXfi
Problem (16.F.I) states the Pareto optimality problem as one of trying to maximize the well-being of consumer I subject to meeting certain required utility levels for the other consumers in the economy [constraints (I)] and the resource and technological limitations on what is feasible [constraints (2) and (3). respectively]. By solving problem (16.F.I) for varying required levels of utility for these other consumers ('1, •...• iii)' we can identify all the Pareto optimal allocations for this economy. Indeed. you should pause to convince yourself of this by solving Exercise 16.F.1.
oF.
:0;
0
= 0
Pt - Yj ---L = 0 oYtj
.
If Xli > 0
for all i,
t,
(16.F.2)
for all j,
t.
(16.F.3)
As is well known from Kuhn-Tucker theory (see Section M.K of the Mathematical Appendix), the value of the multiplier Pt at an optimal solution is exactly equal to the increase in consumer J's utility derived from a relaxation of the corresponding constraint, that is. from a marginal increase in the available social endowment wt of good t. Thus. the multiplier Pt can be interpreted as the marginal value or "shadow price" (in terms of consumer I's utility) of good t. The multiplier Ii" on the other hand, equals the marginal change in consumer J's utility if we decrease the utility requirement Ui that must be met for consumer i"f. I. Condition (16.F.2) therefore says that, at an optimal interior allocation, the increase in the utility of any consumer i from receiving an additional unit of good t. weighted (if i "f. I) by the amount that relaxing consumer i's utility constraint is worth in terms of raising consumer I's utility, should be equal to the marginal value PI of good t. Similarly, the multiplier Yj can be interpreted as the marginal benefit from relaxing the jth production constraint or. equivalently, the marginal cost from tightening it.
Exercise 16.F_I: Show that any allocation that is a solution to problem (16.F.I) is Pareto optimal and that any Pareto optimal allocation for this economy must be a solution to problem (16.F.I) for some choice of utility levels (ii, •...• iii)' [Him: Use the fact that preferences are strongly monotone.] Because utility functions are normalized to take nonnegative values. from now on we consider only required utility levels that satisfy U, ~ 0 for all i. The point of Exercise 16.F.I can be seen by examining the utility possibility set U in Figure 16.F.1. If we fix a required nonnegative utility level for consumer 2. we can locate a point on the frontier of the utility possibility set U by maximizing
18. Recall that for expositional ease we are not imposing any boundary constraints on the vectors h We note also that the assumption of strictly positive gradients of the functions uc( .} and fj( .} implies that the constraint qualification for the necessity of the Kuhn-Tucker conditions is satisfied. (See Section M.K of the Mathematical Appendix for the specifics of first-order conditions for optimization problems under constraints.)
17. For expositional convenience. we have taken every FJ(') to be defined on the entire R'·. A consequence of this (and the assumption that VF;(y/)» 0 for all YJ} is that every commodity is both an input and an output of the production process. Because this is unrealistic, we emphasize that no more than expositional ease is involved here.
i
563
564
CHAPTER
11:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
Hence. yPJFj/vYt;) is the marginal cost of increasing Ytl and thereby effectively tightening the constraint on the net outputs of the other goods. Condition (16.F.3) says. then. that at an optimum this marginal cost is equated. for every j. to the marginal benefit Jlt of good t. If we suppose that we have an interior solution (i.e .• x,» 0 for all i). then conditions (16.F.2) and (16.F.3) imply that three types of ratio conditions must hold (see Exercise 16.F.3):
(Ju;/Dx" iJu;/iJx n
for all i. i'. f. I'.
(16.FA)
t.1'.
(16.F.5)
for all i.j.t.I'.
(16.F.6)
for allj./.
cJFj"/iJY{"j"
iJll;/(:X~, =J!I/~Ytj
cJU;/(lx{",
vFI/vYf"1
Condition (16.F.4) says that in any Pareto optimal allocation. all consumers marginal rates of substitution between every pair of goods must be equalized [sec Figures 15.B.II(b) and 15.B.12 for an illustration in the two-good. two-consumer case]; condition (16.F.5) says that all firms' marginal rates of transformation between every pair of goods must be equalized [see Figure IS.D.2(b) for an illustration in the two-good. two-firm case]; and condition (16.0.6) says that every consumer's marginal rate of substitution must equal every firm's marginal rate of transformation for all pairs of goods [see Figure 15.C.2 for an illustration in the case of the one-consumer. one-firm model with two goods]. Conditions (l6.F.4) to (16.F.6) correspond to three types of efficiency embodied in a Pareto optimal allocation (see Exercise 16.F.4).
(16.F.7) •.•
XLi) ;;,
(2) L,xl/:!;;x,
u,
Fltri
PARETO
(16.F.9)
s.t. }', 5 fCY, •...• y,).
To explore the relationship of the first-order conditions (16.F.2) and (16.F.3) to the first and second welfare theorems. we make the further. and substantive. assumption that every 11,(') is a quasiconcave function (hence. preferences are convex) and that every fj(') is a Convex function (hence. production sets are convex). The virtue of this assumption is that with it we do not have to worry about second-order conditions; in all thc maximization problems to be considered. the first-order nceessary conditions arc automatically sufficient. In this differentiable. convex framework. conditions (16.F.2) and (16.FJ) can be uscd to establish a version of the two welfare theorems. To see this. note first that (x*, .1'*, p) is a price equilibrium with transfers (with associated wealth levels 1\', = p' for i = I •...• I) if and only if the first-order conditions for the budgetconstrained utility maximization problems
xr
and the profit maximization problems Max
(=
s.t. Fj(Yj)
#~"
I, ...• L.
,"Xli
(16.F.8) (n ..... .,))
{=
2•...• L
j = I •... • J.
The first-order conditions for this problem lead to condition (16.F.5).
P'Yj :!;;
0
are satisficd. Denoting by 7, and PI the respective multipliers for the constraints of these problems. the first-order conditions [evaluated at (x'. Y')] can be written as follows:
i = 2, ...• 1
(ii) Elficienr productioll across tec/tllologies. The aggregate production vector should be elficienr in the sense discussed in Section 5.F. That is, it should be impossible to reassign production plans across individual production sets so as to produce, in the aggregate, more of a particular output (or use less of it as an input) without producing less of another. Focusing. in particular. on the first good. this means that given required total productions (Yl' ... ,yc! of the other goods. we want to solve
(2) Fj(y):!;; 0
u«O, + y, ..... '0,. + h)
Max
_~,Pt
The first-order conditions for this problem lead to condition (I6.F.4).
s.t. (I) Lj Y'j;;' }',
CONDITIONS
(iii) Optimal agf/regllle production levels. We also must have picked aggregate production levels that generate a desirable assortment of commodities available for consumption. Keeping the utility requirements (", •...• ti,) fixed. let u(.i, •...• xc! and J(Y, .. ' .. h) denote. respectively. the value functions for problems (I6.F.7) and (16.F.8). Then we want to solve
(i) Oprimal al/oeation oj available goods aeros., consumers. Given some aggregate amounts (.i" ...• xc> of goods available for consumption purposes, we want to distribute them to maximize consumer I's well-being while meeting the utility requirements (u ...• ",) for " consumers 2•... , I. That is. we want to solve
s.t. (I) u,(x Ii"
fIRST·ORDER
16.f:
The first-order conditions of this problem lead to condition (16.F.6).
vu,./cJx". ou,.fJx{",.
~!"I/VYt j = J!j"~~Y{j: cJFj/<'.vl"j
SECTION
PROPERTIES
Pt -
I',
Pj
{
:!;;
=
0.
0 If Xtl > 0
<'f'. --"
DYfi
= 0
for all i,
t.
(16.F.IO)
for all i.
t.
(16.F.II)
Letting = p,. ,), = II"". and )'j = Pj' we see that there is an exact correspondence between conditions (16.F.2)-(16.FJ) and (16.F.IO)-(16.F.II). Since both sets of conditions are necessary and sufficient for their respective problems. this implies that the allocation (x'. Y') is Pareto optimal if and only if it is a price equilibrium with transfers with respect to some price vector P = (Pl •...• PL)' Note. moreover. that the equilibrium price Pt exactly equals i't. the marginal value of good t in the Pareto optimality problem. Suppose that. in addition. every u,(·) is concave. Then it is also instructive to examine the marginal conditions for the maximization of a linear social welfare
OPflMALITY
5b~
566
CHAPTER
16:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
function (sec Section 16.E). Consider the problem Max
L i.,u,(x u•···• XL') (2)
(16.F.12)
L, Xli S; w, + Lj YIj Fj{J',j, ...• hj)
S;
0
1'. G:
SO .. E
ill!erpretarions of tile Commodity Space
x.)'
s.l. (I)
SEC T ION
PROPERTIES
{=
I •...• L
j = I •... • J.
wh~re i., > 0 for all i. Lelling (1/1, •...• I/IL) and ('11 •...• ~J) denote the multipliers on constraints (I) and (2). respectively. the necessary and sufficient first-order conditions for this problem can be wrillen as follows:
t.
(16.F.13)
for all j. (.
(16.F.14)
for all i.
Not~ that by lelling <>, = ;''/;;'''/'' = I/I,/i." and Yj = ~j/i.,. we have an exact correspondence between (16.F.2)-(16.F.3) and (16.F.13)-(16.F.14). Therefore. any solution to (16.F.I3) and (16.F.14) is a solution to (16.F.2) and (16.F.3) and. hence. a Par~to optimum'· Conversely. any Pareto optimum that for some multipliers satislics (16.F.2) and (16.F.3) is also a solution of (16.F.13) and (16.F.14). and consequently of problem (16.F.12). for an appropriate choice of;. = (i., •...• i.,). It is also enlightening to compare (16.F.13) and (16.F.14) with the first-order conditions (16.F.1O) and (16.F.II) for the optimization problems associated with a price equilibrium with transfers. We get an exact correspondence between them by lelling p, = 1/1,. IX, = I/i.,. and {lj = ~j' Once again. the price p{ represents the marginal social value of good t. In addition. note that IX" which is the marginal utility of wealth for consumer i at prices p and wealth level Wi = p' x1. equals the reciprocal of ;.,. Hence. we can draw the conclusion presented in Proposition 16.F.1.
Proposition 1S.F.1: Under the assumptions made about the economy [in particular, the concavity of every Ui(') and the convexity of every 0(' every Pareto optimal allocation (and, hence, every price equilibrium with transfers) maximizes a weighted sum of utilities subject to the resource and technological constraints. Moreover, the weight i'i of the utility of the ith consumer equals the reciprocal of consumer i's marginal utility of wealth evaluated at the supporting prices and imputed wealth.
»).
16.G Some Applications In this section. we present some applications of the ideas covered in the previous sections of the chapter. We first discuss three examples that introduce particular interpretations of the commodity space. We then present an extension of the second welfare theorem that relies on a concept of marginal cost prici'lg. 19. R.".II that by Ihe concavity-convexity assumptions, (16.F.2)-(16.F.3) and (16.F.I3)(16.F.14) are necessary and sufficient conditions for their respective problems.
Up to now we have treated our commodities as abstractly defined objects. This has not been for formalism's sake. but to facilitate a wide applicability of the theory. It is easy to think of the case in which commodities are distinct. physically tradeable real objects. But there arc many other interesting possibilities. The theory presented in the previous sections has proven to be remarkably flexible and subtle in the interpretations that can be given to the commodities. consumption sets. preferences. and production sets. Two important examples. ('ollllllodiries COlI!illgent on the state of the world and dared (,/lllll/wdilies. are discussed extensively in Chapters 19 and 20. respectively. For the sake of completeness. we devote a few words to contingent commodities in this section. We then briefly discuss two other examples: occupational choice and I'lIhlic goads.
Example 16.G.1: COlllillgenr COllllllodities. An interesting use of artificial commodities appears in the area of general equilibrium under uncertainty. A full formal description is presented in Chapter 19. but the basic idea can be conveyed here. The usefulness of a commodity may depend on uncertain. external circumstances. For example, medical care is much more important if one is ill than if one is healthy. To ensure an efficient allocation of resources. we have to make sure not only that the right commodities are delivered to the right people but also that they are allocated under the right circumslances. that is. according to the realization of the uncertain external states. To model this type of resource allocation problem. we can use the concept of a cOlltillgell! commodity. A commodity such as medical care can be subdivided into many different "artificial commodities." each of which has the interpretation "medical care is provided under circumstance s." For example. suppose that there are I consumers in the economy. each of whom may turn out ex post to be either "sick" or "healthy." A consumer's need for medical care depends. of course. on her state of health. From an economy-wide perspective. there are then 21 different states of nature. each corresponding to a different configuration of ill health across the population. We can Iherefore imagine 2' different commodities called "medical care." one for each of these configurations. A consumer buying "medical care in state s" receives care when state s occurs.20 One of the strengths of general equilibrium theory is its ability to deal easily with an arbitrary number of commodities. There are very few results that depend on the number of commodities. and none of them is of general interest. Therefore. even though it seems difficult to conceive of a very large number of markets for a very large number of contingent commodities. all the welfare propositions that we have developed turn out to be easily applicable to this uncertainty selling (to be sure. we arc taking a theoretical. rather than a practical perspective here). In Chapter 19. we discuss these points in more detail. • 20. Thus. to purchase medical care when she is sick. the consumer actually buys a large number of different "conlingent medical care commodities" (in fact, 2'-1 of them).
A P P LtC A T ION S
567
568
CHAPTER
11:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
PROPERTIES
SECTION
r;
CI
~ 0, t" ~ 0, t" ~ 0,
SOME
APPLICATIONS
569
Suppose there arc I consumers and two commodities; a "private" good, say labor, and a public good (the theory presented extends without essential modification to any number of private or public goods). A private good is like the goods discussed up to now: A unit of the good can be consumed only once, so if consumer i consumes it then it is unavailable for consumption by others. We let Xli denote consumer j's consumption of the private good. But in the case of a (pure) public good, its consumption by consumer j does not prevent its availability to other consumers. Thus, if x, is the total amount of public good produced, then X, can be made available at no extra cost to every consumer. We assume that every consumer i has the consumption set R~ and continuous, convex, locally nonsatiated preferences ;:::, defined on pairs (x Ii' x,). The model is completed by having some amount w. of the private good as the initial total endowment (there is no endowment of the public good) and a firm that transforms amounts Z E IR+ of the private good into the public good by means of an increasing, concave production function f(z). An allocation «x •• , ... , x tl , x,), (q, z));:>: 0 is feasible if
Example 16.G.2: Occupational Choice Suppose that every individual could, in principle, work either as a classics scholar or as an economics professor. But not all individuals are equally good at both things. A way to capture the different comparative advantage is to assume that for every individual i, there is an Il i ;:>: 0 measuring how many "effective hours of economics professorial services" it takes to produce "an effective hour of classical scholarship." A relatively low Ili indicates comparative advantage in classical scholarship. Suppose also that every individual i has an amount of professorial hours that she can supply; we assume that I professorial hour can produce I effective hour of economics professorial services or 1/':1. i effective hours of classical scholarship by individual i. There is a single consumption good on which the individual i can spend her earnings. It is important to be able to imbed this problem in our formal structure because we certainly want to be able to analyze how, for example, competitive labor markets will perform when individuals have occupational choices as well as choices about how mllcil labor to supply. This is how it can be done (it is not the only possible way): suppose we list consumption and effective hours supplied as a three-dimensional vector (c i , t d , t'i)' where CI is individual j's consumption and t" ~ 0 and t" ~ 0 are the effective hours spent working as a classics scholar and as an economics professor, respectively. Because the latter two quantities are supplies-that is, services offered by the individual to the market-we follow the convention of measuring them as negative numbers. We can then define the consumption set of individual j as XI = (c" t", 1.1 ):
II.G:
q ~ f(:),
,
LXI'
+:
=
w.,
and
q=
x,.
«x;., ...
z'»
It is Pareto optimal if there is no other feasible allocation ,.<;" xi). (q', such that (x'1I' xi) ;:::i(-,'.i' xi) for all i and (X;i' xi) >-,(x'1I' xi) for some i. We now describe this model in an artificial but equivalent way, with the advantage, that, formally speaking, it reduces the public commodities to private ones and therefore makes the results of Sections 16.C and 16.0 applicable. The "trick" is to define a personalized commodity x2/ for every consumer i, to be interpreted as "commodity 2 as received by the ith consumer." Formally, consumer i cares only about good I and the ith personalized commodity. We therefore denote her consumption bundle by XI = (X1/' X2;)' The single firm is now viewed as producing a joint bundle of personalized commodities with a technology that produces these commodities in fixed proportions. Formally, its (convex) production set is
r; + t" + 11,/" ~ O}.
One should interpret the nonpositivity constraints as the inability to consume labor + 1.1 + 1111" is the time available for leisure activities. services. The amount Preferences are defined on XI' Because the consumption good is desirable, the local nonsatiation condition is satisfied. The assumption that preferences are continuous and convex is also natural. We can complete the model by having a concave production function f(z" z.) that transforms input combinations (z" z.) of effective hours of classics and economics scholarship, respectively, into the consumption good. We now have a complete general equilibrium system to which we can pose a number of interesting questions: If the occupational choice is directed by a competitive (i.e., price-taking) market system, will the outcome result in an efficient exploitation of comparative advantage? Conversely, can every efficient arrangement of occupations be sustained by a market system (supplemented perhaps by lump-sum transfers)? The results of Sections 16.C and 16.0 tell us that the answer to both questions is in the affirmative. _
r;
y = {( -z, q., ... , q,) E R~+·: z;:>: 0 and q.
= ... = ql = q ~
f(z)}.
With this reinterpretation, the model fits into the structure analyzed throughout the chapter. A price equilibrium with transfers for this artificial economy is known as a Lindahl equilibrium. 22 Definition 16.G.l: A Lindahl equilibrium for the public goods economy is a price equilibrium with transfers for the artificial economy with personalized commodities. That is, an allocation (xf, ... , xi, q., z·) E 1R21 X R x R and a price system (p" P2'" .. , P2') E R'+' constitute a Lindahl equilibrium if there is a set of wealth levels (w., ... , w,) satisfying :L w; = :L p,xf; + (L P2;)q· - p,z· and such that:
Example 16.G.3: Public Goods. The notion of a "public good" and the more general concept of an "externality" were discussed in Chapter II, where we also introduced the key idea of "personalized" prices.'· Consequently, we can be rather brief here. (The basic references on public goods were also given in Chapter 11.)
(i) q. ~ f (z·) and (L; P2;) q* - p,z* and q ~ f(z).
21. In fact, the current discussion is more general than the one in Chapter 11 because there we restricted ourselves to a quasilinear setting.
~
(L P2;) q -
p,z for all (q, z) with z ~ 0
22. More properly, we should say a Lindalll equilibrium with transfers.
l
-
570
C HAP T E R
1 6:
E QUill B R I U MAN D
ITS
BAS l e w Elf ARE
PRO PER TIE 5
SEC T ION
(iii)
SOli E
A P P Lie A T ION s
571
.',
(ii) For every i, xi = (xf;, x~;) is maximal for <':;; in the set {(x,;, x 2;) EX;: p,x,; + P2iX2i
1 I. G:
:5 Wi}.
Lxr; + z*
= w,
and x~;
= q* for every i.
The first and second fundamental welfare theorems tell us that every Lindahl equilibrium is Pareto optimal and every Pareto optimal allocation can be implemented using a Lindahl equilibrium (with appropriate wealth transfers and, perhaps, with the usual quasiequilibrium qualification).23 There is an important caveat, however: unlike economies where with large numbers of agents each agent becomes small relative to the size of the market, in markets for personalized goods, each consumer is necessarily large with respect to the market in her personalized good. As we mUltiply the number of consumers, we also multiply, as a matter of dclinition, the number of commodities. As a result, it is very unlikely that the critical assumption of price taking will be satisfied. Thus, the descriptive content of this equilibrium concept is low. Nevertheless, the second welfare theorem may still be of some interest. In particular, it tells us that if the planning authority has a means to enforce the prices, then we have a mechanism involving voluntary purchases of the public good that achieves the desired Pareto optimal allocation. Even for this purpose, however, further diflieultics arise that arc inherent to the public goods setting: First, to calculate the personalized supporting prices, statistical information (e.g. information on the distribution of preferences across the economy) will not do; the fact that prices arc personalized means that personal, private information is required. This information may be difficult to get because individuals will often not have incentives to reveal this information truthfully (see Chapters II and 23 for more on this issue). &cond, for a personalized market voluntary mechanism to work, individuals must expect to receive precisely the amount of public good they purchase. This requires that the public good be excludable; that is, there must be some procedure to deny total or partial use of the public good to anyone who does not pay for it. In many cases, such exclusion is difficult, if not impossible (consider, for example, national defense). _
Figure 16.G.1
The firm incurs a loss at Ihe prices that locally support the Pareto oplimal allocalion.
Firm's Loss Measured in } Units 01 Good 2
if price taking can somehow be relied on (perhaps because a planning authority can enforce prices), it may still be impossible to support a given Pareto optimal allocation. Examples arc provided by Figures IS.C.3(a) and 16.G.l.ln Figure 16.G.1, at the only relative prices that could support the production y* locally, the firm sustains a loss and would rather avoid it by shutting down. In Figure \S.C.3(a), on the other hand, not even local profit maximization can be guaranteed (see the discussion in Section I S.C). Although nonconvexities may prevent us from supporting the Pareto optimal production allocation as a profit-maximizing choice, under the differentiability assumptions of Section 16.F we can usc the first-order necessary conditions derived there to formulate a weaker result that parallels the second welfare theorem (see Exercise 16.G.\). Proposition IS.G.l: Suppose that the basic assumptions of Section 16.F hold" and that, in addition, all consumers have convex preferences (so utility functions are quasiconcave). If (x*, yO) is Pareto optimal, then there exists a price vector p=(p" .... PL) and wealth levels (w" ... ,w,) with L;W;=P·iiJ + LjP·yj such that:
NOllco/1vex Production Technologies and Marginal Cost Pricing
(i) For any firm j, we have
The second welfare theorem runs into difficulties in the presence of nonconvex production sets (in this section, we do not question the assumption of convexity on the consumption side). In the first place, large nonconvexities caused by the presence of fixed costs or extensive increasing returns lead to a world of a small number of large firms (in the limit, production efficiency may require a single firm, a so-called "natural monopoly"), making the assumption of price taking less plausible. Yet, even
p
= lj VF;(yj)
for some Yj> O.
(16.G.l)
(ii) For any i, xi is maximal for <':;; in the budget set {X;E
(iii)
L.; xi
Xi:
p·X; ~ Wi}.
= w + Lj yj.
The type of equilibrium represented by conditions (i) to (iii) of Proposition \6.G.1 is called a /IIaryinal cost price equilibriU/ll wit Ii transfers. The motivation for this terminology comes from the one-output, one-input case. 2S
23. Suppose that the production function I( . ) is differentiable and the indifference curves are smooth. and consider a Lindahl equilibrium that is interior. Then preference maximization implies that p,;ip, = - MRS\ .. where MRS\, is consumer j's marginal rate olsubslitution of good 2 for good I. On Ihe other hand, profit maximization entails L/ P,Jp, - -MRT", where MRT" is the
firm's marginal rate 01 transformation of good 2 for good I (the marginal cost of oUlput in terms
24. That is, (he assumptions leading up to conditions (16.F.2) and (I6.F.3). 25. We point out that for the general case, the term marginal cost price equilibrium is, strictly speaking, inappropriate. Exercise 16.G.3 explains why. However, the terminology is by now standard, and we retain it.
of input). Hence, in any Lindahl equilibrium. we must have Li M RS~ I = M R T11 • which is exactly the Samuelson optimality condition for a public good (see Section I1.G ror its derivation in the ca:..e of quasilinear prderences).
1
572
C H AFT E R
1 6:
E QUI LIB R I U MAN 0
ITS
BAS l e w ELF ARE
PRO PER TiE S
APPtn
:(
A:
APPENDIX
A:
FEAStBLE
ALLOCATIONS
"ROPEATIf:~
TECHNtCAL
(;
PROPERTIES
'-
OF
,'I:..
THE
t... r
SET
f E A ~ I 8 l E A L L 0 CAT ION S
OF
The sct of feasible allocations is
A= C
{(XI"'"
!;I/"
X
x,, YI"'"
YJ)
E
XI
X ..• X
X,
X
YI
X ..•
x Y,: LI Xi = W + Lj Yj}
!;I").
Our economic problem would not be very interesting if there were no feasible allocations for the economy or if we could give every consumer an unboundedly large consumption vector. We might therefore simply assume that A is nonempty, bounded, and, for good measure, closed (i.e., nonempty and compact). In Chapter 17, where this technical point becomes important for the study of the existence of Walrasian equilibria, we assume exactly this. Nonetheless, it is useful to give, once and for all, a set of sufficient conditions for these very basic properties to hold.
Flgur. 16.0.2 A marginal cost price equilibrium need not
be Parelo optimal.
Exercise 16.G.2: Suppose there arc only two goods: an input and an output. Show that in this case, condition (16.G.I) simply says that the price of the input must equal the price of the output multiplied by the marginal productivity of the input or, equivalently, that the price of the output equals its marginal cost.
Proposition 16.AA.1: Suppose that (i) Every Xj: (i.1) is closed; (i.2) is bounded below (i.e .. there is , > 0 such that Xlj > -r for every ( and i; in words. no consumer can supply to the market an arbitrarily large amount of any good). (ii) Every ~. is closed. Moreover, the aggregate production set Y = Lj ~:27 (ii.1) is convex: (ii.2) admits the possibility of inaction (i.e .. 0 E Y): (ii.3) satisfies the no-free-Iunch property (i.e., V ~ 0 and V E Y implies V = 0): (ii.4) is irreversible (VE Yand -VE Y implies V = 0).
As we have noted, condition (16.G.I) does not imply that the (yr,· .. , yJ) are profit-maximizing production plans for pricc-taking firms. The condition says only that small changes in production plans have no first-order effect on profit. But small changes may still have positive second-order effects (as in Figure IS.C.3, where at a marginal cost price equilibrium the firm actually chooses the production that minimi;es profits among the efficient productions) and, at any rate, large changes may increase profits (as in Figure 16.G.I). Thus, to achieve allocation (x·, y.) may require that a regulatory agency prevent the managers of nonconvex firms from attempting to maximize profits at the given prices p.2. Sec Quinzii (1992) for extensive background and discussion on the material presented in this section.
Then the set of feasible allocations A is closed and bounded [i.e .. there is r > 0 such that Ixiii < rand IVIiI < , for all i. i. I and any (x. V) EA). If. moreover. -!;I~ c Y and we can choose k j E X j for every i in such a manner that L j k j ~ W. then A is nonempty. Proof: The proof of this proposition is rather technical, and we shall not give it. Nonetheless, we shall say a few words regarding the logic of the result. The nonemptiness part is clear enough because we have X, E X, for every i and L -Xi - WE - R~ c Y. Thus, an allocation with individual consumptions (XI"'" x,) and aggregate production vector LXi - W is feasible. Similarly, the closed ness of A is a direct consequence of the closed ness assumptions on the consumption and production sets (see Exercise 16.AA.I). What remains is to show that A is bounded. To gain some understanding, suppose that J = I and Xi = R~ for every i (as long as every Xi is bounded below, the argument for general consumption sets is similar). In Figure 16.AA.I, we represent the set of feasible aggregate consumption bundles (Y + {wl) n !;I~, that is, the set of nonnegative vectors obtained when the origin of Y is shined to w2': O. It is intuitive from the figure that this set can be unbounded above only if Y contains nonnegative, nonzero vectors and so violates the no-free·lunch condition.
It should be noted that the converse result to Proposition 16.G.I, which would assert that every marginal cost price equilibrium is Pareto optimal, is not true. In Figure 16.G.2, for example. we show a one-consumer economy with a nonconvex production set. In the figure, x· is a marginal cost price equilibrium with transfers for the price system p = (I, I). Yet, allocation x· yields the consumer a higher utility. Informally, this occurs because marginal cost pricing neglects second·order conditions and it may therefore happen that, as at allocation x·, the second·order conditions for the social utility maximization problem are not fulfilled. As a result, satisfaction of the first·order marginal optimality conditions (which in the case of Figure 16.G.2 amounts simply to the tangency of the indifference curve and the production surface) does not ensure that the allocaton is Pareto optimal. (Sec Exercise 16.GA for more on this topic.)
26. In the context of Figure 16.G.I. the regulator could reach the desired outcome by merely prohibiting the firm from shutting down and otherwise letting it maximize profits at the "supporting" prices (assuming that the firm will act as a price taker; otherwise, the regulator may also need to enforce those prices).
27. Sec Section 5.B for a discussion of conditions (ji.l) to (;;.4).
L
!>I '-'
574
CHAPTER
16:
EQUILIBRIUM
AND
ITS
BASIC
WELFARE
EX ERe,s E S
PROPERTIES
x,
Koopmans. T. (1957). Three Essays on the Stale of Economic Science. New York: McGraw-Hili. Lange, O. (1942). The foundation or welfare economics. Econometrica 10: 21S-228. Quinzii, M. (1992). Increasing Returns and Efficiency. New York: Oltrord University Press. Samuelson. P. (1947). FOUildarions of Economic AnalysiS. Cambridge. Mass.: Harvard University Press.
/(Y+ {w}) n R~
EXERCISES Figure 16.AA.l The set or reasible
16.C.t" Show that ir a consumption set Xi c RL is nonempty. closed. and bounded and the prererence relation <:::, On X, is continuous. then <:::, cannot be locally nonsatiated. [Hint: Show that the continuous utility runction representing <:::, must have a maximum of X,; see Section M.F or the Mathematical Appendix on this.]
aggregate consumption vectors is compact.
YOI' should verify. however. that this intuition also depends on three facts: 0 E Y. Y is closed. and Y is convex (see Exercise 16.AA.2). Note now that if the set of feasible aggregate consumptions is bounded. then. a fortiori. so are the sets of feasible individual consumptions and the set of feasible productions (because J = I. we get this set by subtracting in from each feasible aggregate consumption vector). Hence. A is bounded.
{x,
16.C.3" In this exercise you are asked to establish the first welfare theorem under a set or assumptions compatible wilh satiation. Suppose that every X, is nonempty and convex and thai every <:::, is slrictly convex (i.e .• if xi <:::, x, and xi "" x, then axi + (I - a)x, >-, '
(b) Any pric'C equilibrium with Iransrers is a Pareto optimum.
16.C.4' Suppose that for each individual there is a "pleasure function" depending on her own consumplion only. given by u,(x i ). Every individual's utility depends positively on her own and everyone else's "pleasure" according to the utility function U,(x, •... , x,) = U,(u,(x,). u,(x,) •...• ",(XI»'
Show that ir X = (x ,•.... x,) is Pareto optimal relative to the Uk )'s. then (x, •...• XI) is also a Pareto optimum relative to the u,s. Does this mean that a community or altruists can use competitive markels to attain Pareto oplima? Does your argument depend on the concavity of the Ui·S. or the U/s?
Proposition 16.AA.2 gives an important implication of the compactness of the sct of feasible allocations for the form of the utility possibility set. Proposition 16.AA.2: Suppose that the set of feasible allocations A is nonempty. closed. and bounded and that utility functions ui (.) are continuous. Then the utility possibility set U is closed and bounded above.
16.0.1' Prove that if prererences are locally nonsatiated then the condition: "if x,
A}
C
{Xi €
Xi;::i
16.0.2" Exhibit a one-firm. one-consumer economy in which the production set is convex. the prererence relation is continuous and convex. and there is nevertheless a Pareto optimal allocation that can be supported neither as a price equilibrium with transrers nor as a price quasiequilibrium with transrers. Which condition of Proposition 16.0.1 fails?
R /..
Thus. U' is the image of the compact set A under a continuous function and is therefore itself a compact set (see Section M.F of the Mathematical Appendix). From this. the closed ness and the boundedness above of U = U' - IR~ follow directly. _
16.0.3" Suppose that we have an economy with continuous and strongly monotone preferences (consumption sets are X, = R~). Suppose also that a strictly positive production is possible; that is. there are YI e lj such that Lj Yj + OJ » O. Prove that any price quasiequilibrium with transrers must be a price equilibrium with transfers. [Hint: Show first that w, > 0 for some i and then argue that p » 0.]
REFERENCES
16.0.4C Consider a two·good exchange economy with r identical consumers. The consumption set is R!. the individual endowments are co e R! •• and the prererences are continuous and strongly mono lone but not necessarily convex. Argue that the symmetric allocation in which
Albis. M. (1953). Traile d'ecOIlOmie pure. Paris: Publications du CNRS. Arrow, K., and F. Hahn. (1971). Genrral Competitive Analysis. San Francisco: Holdcn·Day. Dcbrcu. G. (1959). Tllt'orr IIf Value, New York: Wiley.
L
>-, xr then
xr" is equivalent to the condition: "x~ is expenditure minimizing for the price vector Xj: xr }." p in the set
p' Xi ~ P'
Proof: Note that U = U' - IR~ where E
xr
p'X; ~ Wi'"
The case with several production sets is more delicate. and it is here that the irreversibility assumption comes to the rescue. Very informally. we can dcrive. as in the preceding paragraph. the boundedness of feasible aggregate productions and feasible individual consumptions. Now. the only way that unboundedness would be possible at the individual production level while remaining bounded in the aggregate is if. so to speak. the unboundedness in one individual production plan was to be canceled by the unboundedness of another. However, this would imply that the collection of all technologies in the economy (i.e.• the aggregate production set) allows the reversal of some technologies (see Exercise 16.AA.3 for more details). Incidentally, it can also be shown that irreversibility, with the other assumptions. yields the closed ness of Y, so we do not actually need to assume this separately. _
U' = {(u,(x,), ... , U/(X/)): (x. Y)
xr
16.C.2' Suppose that the prererence relation <:::, is locally nonsatiated and that is maximal ror <:::, in set E Xi: P' x,:s w.}. Prove that the rollowing property holds: "If x, <:::, then
+-
575
576
C HAP T E R
1 6:
E QUI L I 8 R I U MAN D
ITS
8 A SIC
W ELF ARE
PRO PER TIE S
E X ERe I S E S
every consumer gets her initial allocation is either a Walrasian equilibrium (for some price vector p) or, if it is not, then for r large enough it is not a Pareto optimum. [Hint: the differentiable case is simpler.]
There is a single consumer with the utility function u(e, f) = e'j" - ",
and endowment
16.E.l" Given a utility possibility set U, denote by U' c U the subset actually achieved by feasible allocations: U'
= (u,(x,), ... , u,(x,)): L
x,
= Lj YJ + w for some YJ E
(WL'
w,.). To ease the calculations, take
Wt
= wr = I and a = \.
(a) Find the optimal allocation of the endowments to their productive uses.
(b) Recognizing that the production of education entails increasing returns to scale, the government of this economy decides to control the education industry and finance its operation with a lump·sum tax on the Consumer. The consumption of education is competitive in the sense that the consumer can choose any amount of education and food desired at the going prices. The food industry remains competitive in both its production and consumption aspects. Assuming that the education industry minimizes cost, find the marginal cost of education at the optimum. Show that if this price of education were announced, together with a lump·sum tax to finance the deficit incurred when the education sector produces the optimal amount at this price for its product, then the consumer's choice of education will be at the optimal level.
lj}.
(Relative to U', the set U allows for free disposal of utility). (a) Give a two-consumer, two-commodity exchange example showing that it is possible for a point of U' to belong to the boundary of U and nOI be a Pareto optimum. (b) Suppose that every lj satisfies free disposal and 0 E lj. Also, assume that for every i, Xi = R'. and ~, is continuous and strongly monotone. Show that any boundary point of U that belongs to U' is then a Pareto optimum. [Hinl: Let u,(O) = 0 for all i and show first that U' = U ('\ R~. Next argue that if u E U is a Pareto optimum and 0 :s; u' :s; u, u' #0 u, then we must be able to reach u' with a surplus of goods relative to u.]
(0) What is the level of the lump-sum tax necessary to decentralize this optimum in the manner described in (b)?
(c) Consider an exchange economy with consumption sets equal to R~, continuous, locally nonsati"ted preferences, and a strictly positive total endowment vcctor. Show that if II = (Ill •... ,II,) is the utility vector corresponding to a price quasicquilibrium with transfers then u cannot be in the interior of U; that is, there is no feasible allocation yielding higher utility to ever)' consumer. [llillt: Show that 1\'; > 0 for some i and then apply Proposition 16.D.2.]
Now suppose there inc two consumers and that their preferences are identical to those above. One owns all of the land and the other owns all of the labor. In this society, arbitrary lump-sum taxes arc not possible. It is the law that any deficit incurred by a public enterprise must he covered by a tax on the value of land. (d) In appropriate notation, write the transfer from the landowner as a function of the government's planned production of education.
(0) Find a marginal cost price equilibrium for this economy where transfers have to be compatible with the transfer function specified in (d). Is it Pareto optimal?
16.E.2K Show that the utility possibility set U of an economy with convex production and consumption sets and with concave utility functions is convex.
16.AA.1 A Show that if every Xi and every lj is closed, then the set A of feasible allocations is closed.
16.F.l" In text. 16.F.2A Derive the first-order conditions (16.F.2) and (I6.FJ) of the maximization problem (16.F.I).
16.AA.2K Show that( Y + (w}) ('\ R~ is compact if the following four assumptions are satisfied: (i) Y is closed, (ii) l' is convex, (iii) 0 E Y, and (iv) if v E Y ('\ R~ then v = O. Exhibit graphically four examples showing that each of the four assumptions is indispensable.
16.F.3 A Derive conditions (16.F.4), (16.F.5), and (I6.F.6) from the first-order conditions (16.F.2) and (16.F.3).
16.AA.3" Suppose that Y = Y, + Y, c R'; satisfies the assumptions given in Exercise 16.AA.2 and that 0 E 1'" 0 E l',. Argue that if the irreversibility assumption holds for Y then (.I', E Y,: y, + y, + w ~ 0 for some y, E Y,} is bounded.
16.F.4 A Derive the first-order conditions (16.F.4), (16.F.5), and (16.F.6) from problems (16.F.7), (16.F.8), and (16.F.9), respectively. 16.G.1A Prove Proposition 16.G.1 using the first-order conditions (16.F.2) and (16.F.3). I6.G.2A In text. 16.G.3" Exhibit graphically a one-consumer, one-firm economy with two inputs and one output where at the (unique) marginal cost price equilibrium, cost is 1101 minimized. [/lillt: Choose the production function to violate quasi concavity.] 16.G.4" Show that under the general conditions of Section 16.G if there is a single consumer (perhaps a normative representative consumer) with convex preferences, then there exists at least one marginal cost price equilibrium that is an optimum. 16.G.S" In a certain economy there are two commodities. education (e) and food (f), produced by using labor (L) and land (D according to the production functions e = (Min (L, T})'
and
j=(LD'/'
.....
577
C
The Positive Theory
HAP
T
E
R
SEC T ION
17
'7 • B:
E QUI LIB" I U M:
D E FIN I T ION SAN DBA SIC
the three sections is the role of two sufficient conditions: the weak axiom of revealed preferellce ill Ille aggregale (a way of saying that wealth effects do not cancel in the aggregate the positive influence of the substitution effects), and the property of gruss slIhslilllliulI (a way of saying that there are not strong complementarities among the goods in the economy). In Section 17.1, we return to the role of convexity in guaranteeing the existence of Walrasian equilibrium. We qualify this role by showing that nonconvexities that are "small" relative to the aggregate economy (e.g., the indivisibility represented by a car) are not an obstacle to the (near) existence of equilibria, even if they are "large" from the standpoint of an individual agent. This chapter is of interest from both methodological and substantive points of view. From a substantive standpoint, it deals with an important theory: that of Walrasian equilibrium. Methodologically, the qucstions that we ask (e.g., does an equilibrium exis!"! Are the equilibria typically isolated? Is the equilibrium unique? Is it stable? What arc the effects of shocks') and the techniques that we use are questions and techniques that arc of relevance to any theory of equilibrium.
of Equilibrium
17.A Introduction III this chapter. we study the theoretical predictive power of the Walrasian equilibrium model. Thus. in contrast with Chapter 16. our outlook here is positive rather than
17.B Equilibrium: Definitions and Basic Equations
Ilormative.
The concept of a private ownership economy was described in Section 16.B. In such an economy. there are I consumers and J firms. Every consumer i is specified by a consumption set Xi c R/.• a preference relation <::;i on Xi' an initial endowment vector / Wi E R .• and an ownership share 0ij ~ 0 of each firm j = I, ... , J (where LI Oij = 1). Each firm j is characterized by a production set >j c RL. An allocation for such an economy is a collection or consumption and production vectors:
We begin in Section 17.B by laying the foundations for our amllysis. We recall t he basic model of a p,.ivlIIe uWllership ecullumy and the definition of a W"lrusi,," ('(/II;Ii/>,.illlll presented in Section 16.B. We then introduce the notion of an ayyregllle ncess
(x. y) = (x, •...•
XI.
y, •...• YJ)
E
X, x ...
X XI X
Yt x ... x }j.
The object of investigation in this chapter is the notion of Walrasian equilibrium. which we take as a positive prediction for the outcome of a system of markets in which consumers and firms are price takers and the wealth of consumers derives from their initial endowments and profit shares. The formal notion of a Walrasian equilibrium was already introduced in Definition 16.B.3. Definition 17.B.l repeats it. Definition 17.B.1: Given a private ownership economy specified by ({(X;,
<::;;)}f.,. {\f}t.,.
{(Wi' 0" ..... 0iJ)}I.,).
an allocation (x*, yO) and a price vector p = (p, . .... pd constitute a Walrasian (or competitive. or market. or price-taking) equilibrium if (i) For every
i. y/ E lj maximizes
(ii) For every i.
xi
profits in lj; that is.
p'¥':::; P'¥" for all YjE lj. E X; is maximal for <::;; in the budget set {X;EX;: p'x;:::; P'w;
+ L O;jp·ytJ·t j
l. The terminology "xi is m
maximizing choice in the set B: that is.
578
..
Xi E
!:i
in set B" means that X, is consumer j's prererenceB and Xi ~i xi for all xi E B.
E QUA T ION S
579
580
CHAPTER
17:
THE
POSITIVE
THEORY
OF
SECTION
EQUILIBRIUM
17.8:
EOUILIBRIUM:
DEFINITIONS
AND
BASIC
EOUATIONS
581
excess demand function of the economy is
For purposes of formal analysis, it is extremely helpful to be able to express equilibria as the solutions of a system of equations. We devote the remainder of this section to the study of how this may be done. In what follows, we aim at being very concrete and impose strong assumptions to simplify the analysis.
ZIP)
= LZ,(P).
The domain of this function is a set of nonnegative price vectors that includes all strictly positive price vectors
Exchange Economies and Excess Demand FUllctiolls Using the economy's excess demand function :( 1'). condition (17.13.1) can now be expressed more succinctly as follows:
We begin our derivation of equilibrium equations by studying the case of a pure exchange economy. Recall that a pure exchange economy is one in which the only possible production activities are those of free disposal. Formally, we let J = I and 1', = - R~. We take X; = R'. and we assume at the outset that each consumer's preferences are cOlltillUOIIS, strictly convex, and locally Ilollsatiated (shortly we shall strengthen local nonsatiation to strong monotonicity). We also assume that L; W; » O. For a pure exchange economy satisfying the above assumptions, the three conditions of Definition 17.8.1 can be equivalently restated as: (x*, yO) = (xr, ... yj) and I' E R'- constitute a Walrasian equilibrium if and only if
"I'
yr O.p·yr = O. and I' O. xr = Xi(p,p'W i ) for all i [where Xi!') is consumer i's Walrasian demand function]. (iii') Li xr - Li wi = yr· ~
IR'; is an equilibrium price vector if and only if zIp) ~ 0."
(17.13.1')
Note that if" is a Walrasian equilibrium price vector in a pure exchange economy with locally nonsatiated preferences. then " ~ O. :( 1') ~ 0, and p' :(f') = L; p' :,( 1') = L;(/,·x;(I'.P·"'i) -P''''i) = 0 (the last equality follows once again from local nonsatiation). Therefore. for every I, we not only have z,(p) ~ o. but also :,(1') = 0 if 1', > O. Thus. we sec that at an equilibrium a good 1 can be in excess supply (i.e .. haw:, (,,) < 0). but only if it is free (i.e .. only if 1', = 0).2
,xr.
(i') (ii')
E
~
To simplify matters even more. we go one step further by assuming that consumer preferences arc stfOlIl/l.r /l/OIlOtOlle. Thus. for the rest of this section (and. in fact. for all sections of this chapter except Section 17.1 and Appendix B). we let Xi = 1<'; for all,. ano aSSlIllle lhat (Ill "ttd~n.'lIn' rdllliolJs ~; are l'ollli"IUJUS~ strictly ccmr('.\", and str(IIlY/Y mOlwloll('.
Condition (i') is the only one that is not immediate. In Exercise 17.8.1. you are asked to show that it is equivalent to condition (i) of Definition 17.8.1. Conditions (i') to (iii') yield the simple result of Proposition 17.8.1.
With strongly monotone preferences. any Walrasian equilibrium must involve a strin!.r p(lsitire price vector p» 0; otherwise consumers would demand an unbounciedly large quantity of all the free goods. As a result, we conclude that with strong monotonicity of preferences, {/ price vector I' = (1', •.•.• p,J is a WlI/rasillll ('(//Ii/i/lfilllll priC{' "ector if lIlId 011/)' if it "c1clIrs lI/I markets"; that is, if 1I11d 011/.1' if it
Proposition 17.B.1: In a pure exchange economy in which consumer preferences are continuous, strictly convex, and locally nonsatiated, P ~ 0 is a Walrasian equilibrium price vector if and only if:
.'wlres tht! system
(17.B.1)
(~r
L equations in L unknowns: :,(1') = 0
for every 1 = I....• L.
(17.B.2)
or. in more compact notation. :(1') = O. Throughout this chapter, we study the properties of Walrasian equilibria largely by examining the properties of the system of equilibrium equations (17.B.2). We should point out. however. that this is not the only system of equations that we could use to characterize Walrasian equilibria. In Appendix A, for example, we discuss an important alternative system that exploits the welfare properties of Walrasi:!n equilibria identified in Chapter 16. Proposition 17.8.2 enumerates the properties of the aggregate excess demand function. in pure exchange economics with strongly monotone preferences. that are essential to the developments of this chapter.
Proof: That (17.8.1) must hold in any Walrasian equilibrium of such an economy follows immediately from conditions (i') to (iii'). In the other direction, suppose that (17.8.1) holds. If we let )'T=L,(Xi(P.P·W,)-Wi) and Xr=Xi(p,P·w i ). then (x! . ... ,xi, yT) and I' satisfy conditions (i') to (iii'). In particular, note that p'yT = P'L, (.xi(p, P'w;) - w;) = Li(p'X;(p,p'w i ) - P'w;) = 0, where the last equality follows because with local nonsatiation we have P'x;(p,p'w;) = P'W i for all i. • The vector x i ( 1', p' Wi) - W; E IR I• lists consumer i's net, or excess, demand for each good over and above the amount that he possesses in his endowment vector Wi' Condition (17.B.I) suggests that it may be useful to have a formal representation of this excess demand vector, and of its sum over the I consumers, as a function of prices. This is given in Definition 17.8.2.
Proposition 17.B.2: Suppose that, for every consumer i, X; = IR~ and <::;; is continuous, strictly convex, and strongly monotone. Suppose also that Li w; »0. Then the aggregate excess demand function zIP). defined for all price vectors P» 0,
Definition 17.B.2: The excess demand function of consumer i is 2. As a simple example. good r mighl be a "bad." Then. we would expect Iha, good rs price would be zero, consumer demand for the good would be zero, and the excess supply zt< p) = WI > 0 would be eliminated lIsing the disposal technology.
z;(p) = x;(p, P'w;) - Wi'
where x;(p, P'w;) is consumer i's Walrasian demand function. The (aggregate)
-
.-
582
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
SECTION
z(·) is continuous. z(·) is homogeneous of degree zero. P'z(p) = 0 for all p (Walras' law). There is an 5 > 0 such that Z/(P) > -5 for every commodity If pn _ p, where p # 0 and PI = 0 for some t, then
Max {z,(pn) .... , zdpn)} _
DEFINITIONS
AND
BASIC
t and all
defined for some p» [because we may have 11j(p) = 00 for some j). Nevertheless. an equilibrium price vector is still characterized by i(p) = 0. When production sets are not strictly convex, matters become more complicated because the correspondences Yj(p) may no longer be single-valued. Indeed, a production situation of considerable theoretical and practical importance-and one lhat we certainly do not want to rule out by assumption-is the case of constant returns to scale. With constant returns, however, production sets are neither strictly convex nor bounded above (except for the trivial case in which no positive amount of any good can be produced). In principle, we could still view the equilibria as the zeros of a "production inclusive excess demand correspolldellce," defined as in (17.B.3) for a subset of strictly positive prices.· Correspondences, however, do not make good equational systems (e.g., they cannot be differentiated). It is therefore usually much more convenient in such cases to capture the equilibria as the solutions of an extended system of equations involving the production and the consumption sides of the economy. We illustrate this idea in the small type discussion that follows.
p.
00.
Proof: With the exception of property (v), all these properties are direct conseq uenees of the definition and the parallel properties of demand functions.' The bound in (iv) follows from the nonnegativity of demand (i.e., the fact that X, = R~), which implies that a consumer's total net supply to the market of any good t can be no greater than his initial endowment. You are asked to prove property (v) in Exercise 17.B.2. The intuition for it is this: As some prices go to zero, a consumer whose wealth tends to a strictly positive limit [note that, because P-(L, w.l > 0, there must be at least one such consumer] and with strongly monotone preferences will demand an increasingly large amount of some of the commodities whose prices go to zero (but perhaps not of all such commodities: relative prices still matter). _
To sc.::c how an extended system of equations can be constructed, consider the case in which
production is of the linear activity type (this case is reviewed in Appendix A of Chapter 5). Say that, in addition to the dispos,,1 technologies, we have J basic activities ii" ••. , ilJ E R /·.
°
Finally, note that because of Walras' law, to verify that a price vector I' » clears all markets [i.e., has :/(1') = for all t] it suflices to check that it clears all markecs hll/ 0"<'. In particular, if 1'» and Z,(p) ='" = ZL_I(P) = 0, then because P'z(p) = LI 1',=,(,,) = and PI. > 0, we must also have ZL(P) = 0. Hence, if we denote the vector of L - I excess demands for goods I through L - I by
°
EQUILIBRIUM:
°
satisfies the properties: (i) (ii) (iii) (iv) (v)
17.B:
° °
That is. the aggregate production sct is
Y=
{!'E fR /·:.I''' ~
'jilj
for some ('1"'"
oJ)
~ o}.
Because prdcrcnccs are strongly monotone, there can be no free goods at an equilibrium (i.e., we mllst have P »0). Also, productions should be profit maximizing, and because of constant returns. these m"ximum profils must be zero. Therefore, a pair (p, 0) formed by a price vector I' E H:/i and a vector of activity levcls cr: E RJt- constitute an equilibrium if and only if they solve
i(p) = (z 1(1'), ... ,Z,._I(P»,
we see that a strictly positive price vector p is a Walrasi"n equilibrium if and only if :(/» = 0.
(17.B.4) and
Produccioll EccJllomies
for all j,
(17.B.5)
where:(') is the consumers' aggregate excess demand function of Definition 17.B.2. Note that, if so d"sired, condition (17.8.5) can be expressed as a system of equations: just replace "P'lI j ';; O. • j(P·lI j ) = 0" by "'jp·Oj + Max {O, p·Oj} = 0." Exercise 17.B.5 presents an exten-
It is possible to extend the methodology based on excess demand equations to the
gencral production case. Assume, to begin with, that production sets are closed, strictly convex, and bounded above. Then, for any price vector 1'» 0, we can let 11/1') and jj(p) be the maximum profits and the profit-maximizing production vector, respectively, for firm j. Defining
sion of the current discussion to a more general production case allowing for continuous substitution of activities. It is worth emphasizing that, at least for the case of convex technologies, there would not be much loss of conceptual generality if we assumed that the production sector of the economy was composed of a single firm endowed with a constant returns production technology. To sec this, recall from Proposition 5.8.2 that by creating for each firm j an extra, firm-specific,
(17.B.3) as the prodllccioll illclusive excess demalld junccion, we see that P is a Walrasian equilibrium price vector if and only if it solves the system of equations i(p) = 0. In Exercise 17.B.4, you are asked to show that under a weak hypothesis (that a strictly positive aggregate consumption bundle is producible using the initial endowments), the fUllction i(') satisfies properties (i) to (v) of Proposition 17.B.2. Note that if thc production sets are not bounded above, then i(p) may fail to be
factor of production, we can always assume that every >j exhibits constant returns (when we
do so, we transform each consumer·s ownership shares of the profits of the jth firm into endowments of the jth new physical resource). Because profits at an equilibrium must then be zero. we sec th"t onCe this is done there is no need to keep the identity of firms separate in order to compute the wealth of consumers. Moreover, from the point of view of production decisions, ilgain no such need arises. As we saw in Section S.E. we could as well work with the aggregate '·representative firm" Y = Lj I).
J. NOle. im:idcntally. that properties (i) to (iv) continue to hold even if preferences arc only localJy nonsatiated.
4. That is. p would be an equilibrium price vector if and only if 0 E i(p) .
..
EQUATIONS
583
584
CHAPTER
17:
THE
POSITIYE
THEOR'Y
OF
S
EQUILIBRIUM
l;.
C T I 0...
1 l . C.
E,\ 1ST ENe E
0 f-
\"I'
A l RA 5 I A N
A single constant returns Y can be interpreted as a description of a long-run state of knowledge that is freely available to every agent in the economy (i.e., to every consumer), for the purpose of setting up firms or simply for household production. In fact, we could go one step further and formally dispense with the separate consideration of firms and of the profit-maximization condition. In Exercise 17.8.6, you are asked to show that a Walrasian equilibrium can be redefined in terms of the following two-stage process: first, consumers choose a vector v, € RL subject to the budget constraint p'V, S p'W, (the equilibrium market-clearing condition is L, v, = L, w,); second, every i uses VI and the technology Y for household production of a most preferred consumption bundle.'
E a U I Lie R I U M
585
Figure 17.C.1
-s -------------------------
17.C Existence of Walrasian Equilibrium represented in Figure 17.C.I.' When p', is very small, we must have =,(1';, I) > 0; if =,(p~, I) < O. These two boundary restrictions follow by using conditions (iv) and (v) of Proposition 17.B.2 to identify the commodity with positive excess demand as the one whose relative price is very low." Because the function =,(1'" I) is continuous [condition (i) of Proposition 17.B.2]. there must be an intermediate value € [1';,1";] with =,( I) = 0 and, hence. an equilibrium price vector fllllst exist. In the general Case of more than two commodities, the proof that a solution exists is more delicate, and involves the use of some powerful mathematical tools. In Proposition 17.CI, we follow a traditional approach that invokes Kakutani's lixed-point theorem (sec Section M.I of the Mathematical Appendix). We should point out that the proof of Proposition 17.CI has to deal with the technical complication that excess demand is not defined when the prices of some commodities arc zero. The reader may actually gain a more direct insight into the nature of thc fixed-point argument from Proposition 17.C2, which contains a very simple proof for the case of excess demand functions defined for all nonzero, nonnegative prices.
When studying a positive theory, the first question to ask is: under what conditions docs the formal model possess a solution? That is, is it capable of predicting a definite outcome? This is known as the existellce problem. Conceptually, the assurance of existence of an equilibrium means that our equilibrium notion passes the logical test of consistency. It tells us that the mathematical model is well suited to the purposes it has been designed for. Although an existence theorem can hardly be the end of the story, it is, in a sense, the door that opens into the house of analysis.· The existence of a Walrasian equilibrium can be established in considerable generality. To maintain the natural How of exposition, in this section we offer a detailed examination of the existence problem for the particular case that will be our primary focus throughout the chapter: pure exchange economics modeled by means of excess demand functions. In Appendix B, we discuss the existence problem in the general case. We have seen in Section 17.B that the excess demand function z(·) of an exchange economy with L, (/), »0 and continuous, strictly convex, and strongly monotone preferences satisfies properties (i) to (v) of Proposition 17.B.2. We now argue that allY function z(·) satisfying these five conditions admits a solution, that is, a price vector P such that z(p) = O. By doing so, we establish that a Walrasian equilibrium exists under the conditions of Proposition 17.B.2. To start simply, suppose that there are only two commodities (i.e., L = 2). For this case, it is an easy matter to establish the existence of an equilibrium. First, by the homogeneity of degree zero of z(·) (condition (ii) of Proposition 17.B.2), we can normalize P2 = I and look for equilibrium price vectors of the form (P" I). Then, by Walras' law (condition (iii) of Proposition 17.B.2), an equilibrium can be obtained as a solution to the single equation Z,(P" I) = O. This one-variable problem is
1"; is very large, we have
pr
pr,
Proposition 17.C.1: Suppose that zIp) is a function defined for all strictly positive price vectors p E IR~+ and satisfying conditions (i) to (v) of Proposition 17.B.2. Then the system of equations zIp) = 0 has a solution. Hence, a Walrasian equilibrium exists in any pure exchange economy in which Li Wi» 0 and every consumer has continuous, strictly convex, and strongly monotone preferences. 7. Note that we revert to the usual mathematical convention of representing the independent variable p, on the horizontal axis. The partial equilibrium convention of putting prices on the vertical
5. This process formally reduces the production economy to an economy where only exchange takes place. But we do not mean to suggest that the induced exchange economy satisfies all of the strong assumptions that we have imposed in this chapter. 6. It should be emphasized that finding a class of conditions that guarantee the existence of a Walrasian eqUilibrium docs not say that this is the outcome that will occur whenever preferences.
in Marshall (1920) where, in COntrast with Walras (t874). prices are, in fact, dependent variables. R. tn particular, property (iv) implies that the value of intended sales is bounded. By Walr.s' law, the value of intended purchases must therefore also be bounded. Because. by property (v). intended purchases become unbounded in physical terms for some good as PI .... 0, it follows thai it must be good I whose demand becomes unbounded as PI - O. Hence, z,(pl.1) > 0 for 1'; sllOiciently small. By symmetry, as PI -x [which. by the homogeneity of degree zero of :('1, is equivalent 10 1'1 .... 0 holding PI fixed], ror p~ large enough we must have Z1(P';, I) > O. and
endowments, and technologies satisfy the assumptions of the existence theorem: the behavioral assumption of price taking and the institutional assumptions of complete markets must also hold. However, when the conditions required for existence are not satisfied by preferences. endowments, or technologies. it does suggest that the type of equilibrium under consideration may not be the right one to look for.
therefore
,l.
Z I (p;,
I) < O.
Proof of existence of an equilibrium for the case L = 2.
586
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
SECTION
6= {PER~:~P' = I} the unit simplex in RL. Because the function z(·) is homogeneous of degree zero, we can restrict ourselves, in our search for an equilibrium, to price vectors in 6. Note, howcver, that the function z(·) is well defined only for price vectors in the set E
6: p, > 0 for all f}.
We shall proceed in five steps. In the first two, we construct a certain correspondence /(.) from 6 to 6. In the third, we argue that any fixed point of /(.), that is, any 1'* with 1'* E /(1'*), has z(p*) = O. The fourth step proves that /(.) is convex valued and upper hemicontinuous (or, equivalently, that it has a closed graph). Finally, the fifth step applies Kakutani's fixed-point theorem to show that a 1'* with 1'* E f( 1'*) necessarily exists. For notational clarity, in defining the sct f(p) c 6, we denote the vectors that are clements of /(1') by the symbol q.
~
:(p).q'
for all q'E6}.
where s is the bound in excess supply given by condition (iv).'o In summary, for p' c10sc cnough to Boundary 6, thc maximal demand corresponds to some of the commoditics whose price is close to zero. Therefore, we conclude that, for large n, any q' E /(p') will put nonzero weight only on commodities whose prices approach zero. But this guarantees p'q = 0, and so q E f( ,,).
= (q E 6: q, = 0 if z,(p) < Max {z,(,,), ... , zdp)}}.
5/<,1' 5: A fixed poillt exists. Kakutani's fixed-point theorem (see Section M.I of the Mathematical Appendix) says that a convex-valued, upper hemicontinuous correspondence from a nonempty, compuct, convex set into itself has a fixed point. Since 6 is a nonempty, convex, and compact set, and since f(·) is a convex-valued upper hemicontinuous correspondence from 6 to 6, we conclude that there is a p* E 6 with E I(I'*) . •
Observe that if z(p) '" 0 for p» 0, then because of Walras' law we havc Z/(I') < 0 for some t and zr(P) > 0 for some (' '" t. Thus, for such a p, any q E f( 1') has ql = 0 for some t. Therefore, if z(p) '" 0 then f(p) c Boundary 6 = 6\Interior 6. In contrast, if z(p) = 0 then f(p) = 6.
,,*
SI<'p 2: " E
Construction of the fixed-poilll correspondence for I' E Boundary 6. If Boundary 6, we let f(p)
WALRASIAN
ZI(P") < Max:=,(p·), ... , =,.(P')}
In words: Given the current "proposal" p, the "counterproposal" assigned by the correspondence f(·) is any price vector q that, among the permissible price vectors (i.e., among the members of 6), maximizes the value of the excess demand vector :( ,,). This makes economic sense; thinking of f(·) as a rule that adjusts current prices in a direction that eliminates any excess demand, the correspondence f(·) as defined above assigns the highest prices to the commodities that are most in excess demand. In particular, we have
f(p)
OF
and therefore that, ugain, Ifi = O. To prove the above inequality, note that by condition (v) the right-hand side of the above cxpression goes to infinity with II (because" E Boundary 6, some prices go to zero as II .... 00). But the left-hand side is bounded above because if it is positive then
51£'1' /: CO/lStruction of the fixed-point correspondence for P E Interior 6. Whenevcr I' » 0, we let
f(p) = {qE6:z(p)'q
EXISTENCE
511'1' 4: The fixed-poine correspondence is cOllvex-valued and upper hemicontinuous. To establish convex-valuedness, note that, both when I' E Interior 6 and when I' E Boundary 6, f( p) equals a level set of a linear function defined on the convex set 6 [that is, a set of the form {q E 6: i.·q = k} for some scalar k and vector i. E R "], and so it is convex. (Exercise 17.C.1 asks for a more explicit verification.)" To establish upper hemicontinuity (see Section M.M of the Mathematical Appendix for definitions), consider sequences 1" .... 1', q' .... q with q' E f(p') for all II. We have to show that q E f(p). There are two cases: I' E Interior 6 and P E Boundary 6. If P E Interior 6, then p'» 0 for n sufficiently large. From q"'z(p") ~ q"z(p") for all 'I' E 6 and the continuity of z(·), we get q' z( p) ~ q" z( p) for all q'; that is, q E f( 1'). Now suppose that p E Boundary 6. Tuke any ( with PI > O. We shall argue that for II sufficiently large we have q; = 0 and therefore it must be that ql = 0; from this, 'I E I( 1') follows. Because PI > 0, there is an c > 0 such that pi > £ for II sufficiently large. If, in addition, 1" E Boundary 6 then q; = 0 by the definition of f(p'). If, instead, ,,' » 0 then the boundary conditions (iv) and (v) of Proposition 17.B.2 come into play. They imply that, for II sufficiently large, we must have
Proof: We begin by normalizing prices in a convenient way. Denote by
Interior 6 = {p
17.C:
It is instructive to examine which of properties (i) to (v) of Proposition 17.B.2 fail to hold for the excess demand functions corresponding to the Edgeworth boxes of Figures IS.B.IO(a) and (b), where, as we saw, no Walrasian equilibrium existed. In
= {q E 6: p'q = O} = {q E 6: ql = 0 if PI > OJ.
Because 1'1 = 0 for some f, we have f(p) '" 0. Note also that with this construction, no price from Boundary 6 can be a fixed point; that is, I' E Boundary 6 and p E f(p) cannot occur because P'P > 0 while P'q = 0 for all q E f(p).
9. Noto also Ihat ror 'my p E /'J., the set I( p) is always a race or the simplex /'J.; that is, it is One or the subsets or /'J. spanned by a finite subset or unit coordinates. For P E Boundary /'J., I(p) is the race or /'J. spanned by Ihe zero coordinates or p. For p E Interior /'J., I(p) is the race spanned by the coordinates corresponding to commodities with maximal excess demand. 10. In words. the last chain of inequalities says that the expenditure on commodity {is bounded because it has to be financed by. and thererore cannot be larger than, the bounded value of excess supplies.
51<'1' 3: A fixed point of f(·) is an equilibrium. Suppose that 1'* E f(p*). As we pointed out in step 2, we must have p* ~ Boundary 6. Therefore p* » O. If z(p*) '" 0, then we saw in step I that f(p*) c Boundary 6, which is incompatible with 1'* E f( pO) and 1'* »0. Hence, if p* E f(p*) we must have z(p*) = O.
1
EQUILIBRIUM
587
5&0
CHAPTER
1~
TttE
Pos,rlVE
THEORY
OF
EQUILI8RIUM
-1
I'" 0 E x T H E 0 REM
589
By Brouwer's fixed-point theorem (see Section M.l of the Mathematical Appendix) there is a 1'" E 11 such that 1'" = f(p"). We show that zip") :;; O. By Walras' law:
the case of Figure 15.B.lO(b), preferences are not convex and therefore z(') is not a function, let alone a continuous one (condition (i».l1 For Figure 15.B(IO(a), it is property (v) that fails: For any sequence of prices (pi, p;) -+ (1,0), excess demand stays bounded. Note that in the limit there is a single consumer with positive wealth, but the preferences for this consumer, while monotone, are not strongly monotone. To facilitate a clear understanding of the nature of the fixed-point argument it is helpful to consider Proposition 17.C.2, in which boundary complications are eliminated by studying continuous, homogeneous of degree zero functions zIp) satisfying Walras' law and defined for all nonnegative, nonzero price vectors. Within a framework of continuous and strictly convex preferences, this type of excess demand function is not compatible with monotone preferences but can arise with preferences that are locally nonsatiated. Recall also that the equilibrium condition when zero prices are allowed is zIp) :!> 0; see expression (17.B.I ').
0= 1"':(1") = f(p')'z(p') = [I/C/(p')]:+(p')·:(p')·
Therefore, z'(p')':(p') = O. But, as we have already pointed out, this implies :(,,') :s O. • The applicability of Proposition 17.C.1 is not limited to exchange economics. We saw, for example, in Section 17.B (and Exercise 17.B.4), that if we allow for production scts that arc closed, strictly convex, and bounded above (and if a positive aggregate consumption bundle is producible from the initial aggregate endowments), then the production inclusive excess demand function z(·) satisfies conditions (i) to (v) of Proposition 17.B.2. Hence, Proposition l7.C.1 also implies that a Walrasian cquilihrium necessarily exists in this case. We also note for later reference that Proposition l7.C.1 holds as well for a conwx-valued and upper hemicontinuous correspOIu/(,IICI! z(p) that satisfies conditions (ii) tll (v) (properly adapted) of Proposition 17.8.2. In this case, there exists a" such that 11 E :(p). (Sec Exercise 17.C.2 for more on this poin!.) Although Proposition 17.C.1 tells us that an equilibrium exists, it docs not give LIS the equilibrium price vectors or allocations in an explicit manner. The issue of how to actually find equilibria was first considered by Scarf(1973). By now, a variety of useful tcchniques are available. They are very important for applied work, whcre the ability to compute solutions is key. See Shoven and Whaley (1992) for an account of applied general equilibrium.
Proposition 17.C.2: Suppose that zIp) is a function defined for all nonzero, nonnegative price vectors p E R~ and satisfying conditions (i) to (iii) of Proposition 17 .B.2 (Le. continuity, homogeneity of degree zero and Walras' law). Then there is a price vector p" such that zIp") :!> o. Proof: Because of homogeneity of degree zero we can restrict our search for an equilibrium to the unit simplex 11 = {p E IR\: 2:1 PI = I}. Define on 11 the function z+(·) by zt(p) = Max {ZI(p),O}. Note that z+(·) is continuous and that z+(p)'z(p) = 0 implies z(p):;; O. Denote alp) = 2:1 [PI + z;(p)]. We have alp) ~ I for all p. Define a continuous function f(·) from the closed, convex set 11 into itself by
f(p) = [l/a(p)](p
E
+ z+(p)). The second welfare theorem of Section 16.0 can be seen as a particular case of the current existence result. To see this, suppose that ., = (x" ... , x,,) is a Pareto optimal allocation of a pure exchange economy satisfying the assumptions leading to Proposition 17.C.1. Then, by Proposition 17.C.t, a Walrasian equilibrium price vector" and allocation .i = (.<" ... , .<,)
Note that, corresponding to intuition, this fixed-point function tends to increase the price of commodities in excess demand. The construction of the function is illustrated in Figure 17.C.2 for the case L = 2.
exist for the economy in which endowments are W, = .'(, for all i. Since X, is afTordable ~tt prices /' for every consumer i. we must have .tj ~iXi for all i. Hence. it follows from the ()arcto optimality of x = (x" ... , x,) that .x, -,x, for all i. But since .i, is consumer i"s optimal demand given prices P lind wealth Wi = P'w; = P'X;> Xi must also be an optimal demand for consumer i for price wealth pair (p, p·x,). Hence, we see that the price vector p and the wealth levels
Figure 17.C.2
The fixed-point function for Proposition 17.C.2.
\\', = p' x, support the allocation x as a price eqUilibrium with transfers in the sense of Definition t(>.IIA."
17.0 Local Uniqueness and the Index Theorem Having established in Section 17.C (and Appendix B) conditions under which a Walrasian equilibrium is guaranteed to exist, we now begin a study of some issues related to its uniqueness or multiplicity. 11. As we shall mention shortly in this section, existence would still obtain if z(') was a convex-valued and upper hemicontinuous correspondence. In Figure 15.B.lO(b), however, excess demand fails to satisfy the convex-valuedness property.
11. The fact thaL the second welfare theorem can be viewed as a corollary of theorems asserting the existence of Walrasian equilibrium is valid much beyond the economies studied in this seclion.
1
-
590
CHAPTER
17:
THE
POSITIVE
THEORY
OF
SEC T ION
EQUILIBRIUM
1 7 • D:
l 0 CAL
U N I QUE N E S SAN D
THE
IN DE X
THE 0 REM
OC I ParelO Sct
0, Figure 17.0.1 p,
The excess demand
=I
function for an economy with multiple equilibria.
OC,
0,
For a theorist, the best of the all possible worlds is one in which the social situation being analyzed can be formalized in a manner that, on the one hand, is very parsimonious (i.e., uses as inputs only the most indisputable and sturdy traits of the reality being modeled) and, on the other, manages to predict a unique outcome. The Walrasian model of perfect competition is indeed very parsimonious. Essentially, it attempts to give a complete theoretical account of an economy by using as fundamentals only the list of commodities, the state of the technology, and the preferences and endowments of consumers. However, the other side of the coin is that the theory is not completely deterministic. We shall see in Section 17.F that the uniqueness of equilibria is assured only under special conditions. The Edgeworth box of Figure IS.B.9 and Example 15.B.2 provides a simple illustration that, under the assumptions we have made, multiplicity of equilibria is possible. Figure 17.0.1 represents the excess demand function for good I of Example IS.B.2 as a function of PI (normalizing to Pl = I). For another example of multiplicity, see Exercise 17.0.1. From the theoretical point of view, if uniqueness is not achievable, the next-best property is local uniqueness. We say that an equilibrium price vector is locally unique, or locally isolated, if we cannot find another (normalized) price vector arbitrarily close to it. If every equilibrium of an economy is locally unique, we say that the local IIniqlleness property holds for the economy. The local uniqueness property is of interest because, if it prevails, then it may not be difficult to complete the theory in any particular application. For example, history may have determined the region where equilibrium lies (it could be the region where equilibrium used to be before a small unanticipated shock to the economy), and in that region we may have a unique equilibrium. In this case, the theory retains its predictive power, albeit only locally. We say that a theory that guarantees the local uniqueness of equilibria is locally (as opposed to globally) determinate. The question is then: [s the Walrasian theory locally determinate? The example of Figure 17.0.1 suggests that it is: Every solution to the excess demand equation is locally isolated. But Figures 17.0.2 and 17.0.3 provide a counterexample. The figures depict the offer curves and the excess demand function of an exchange economy with a continuum of Walrasian equilibria. Nonetheless, we should not despair. The situation displayed in Figures 17.0.2 and 17.0.3 has an obvious pathological feel about it; it looks like a coincidence. And indeed, it was shown by Ocbreu (1970) that such an occurrence is not robust: it can happen only by accident. We now turn to a formal discussion of these issues. For the sake of concreteness we restrict ourselves, as usual, to the analysis of exchange economics formed by I consumers. Every consumer i is specified by (~;, w;), where ~; is a continuous,
slrictly convex, and strongly monotone preference relation on R~ and Wi »0. As we know, the aggregate excess demand function z(·) then satisfies conditions (i) to (v) of Proposition 17.B.2. We further assume that z(·) is continuously differentiableY Because we can only hope to determine relative prices, we normalize h = I and, as we did in Section 17.B, denote by tip) = (z,(p), ... , Z,._I(P»
the vector of exccss demands for the first L - I goods.
Flgur. 17.0.3 (right)
equilibria is possibh::
A normalized price vector
o.
ReYlllllr Ecollomies
It is useful to begin by introducing the concept stated in Definition 17.0.1.
Definition 17.0.1: An equilibrium price vector p = (Pl' ... ,PL-l) is regular if the (L - 1) x (L - 1) matrix of price effects Di(p) is nonsingular. that is, has rank L - 1. If every normalized equilibrium price vector is regular, we say that the economy is regular.
In Figure 17.0.1, every equilibrium is regular because the slope of excess demand, iJz,(I'I' 1)/iJPI' is nonzero at evcry solution. [n contrast, none of the equilibria of Figure 17.0.3 are regular because at any equilibrium price vector the slope of the excess demand function is zero. Later in this section we shall argue that, in a sense that we will make precise, "almost every" economy is regular. The significance of the technical concept of regularity derives from the fact that a regular (normalized) equilibrium price vector is isolated, and a regular economy can only have a finite number of (normalized) price equilibria. This is formally stated in Proposition 17.0.1. 13. In Appendix A to Chapter 3, we discussed conditions [or the differentiability o[ demand functions and therefore of excess demand functions.
14. Nothing in what [ollows depends on the particular normalization. It can be shown, [or example, thilt i[ zIp) = 0 and the L x L matrix Dz(p) has rank L - I, then the (L - I) x (L - I) matrix Dz(p) has rank L - I whichever good we choose to normalize. Even more, the sign of its determinant is independent o[ the normalization (see Section M.D o[ the Mathematical Appendix).
1
Acontinuum o[ equilibria is possible: Edgeworth box.
A continuum o[ 14
P = (Pi'··.' P,.- I' I) constitutes a Walrasian equilibrium if and only if it solves the systcm or L - I equations in L - I unknowns:
i(p) =
Flgur. 17.0.2 (Ieh)
excess demand.
591
592
CHAPTER
11:
THE
POSITIVE
THEORY
OF
EOUILIBRIUM
Proposition 17.0.1: Any regular (normalized) equilibrium price vector p = (p, ....• PL-" 1)
h ,
I..JI
oJ [ H E 5 SAN
u
1" H E
I N 0 EXT H E 0 REM
L
index p
1/1: :11'1 :O,Pf" II
makes sense. The next proposition (the index ,heorem) says that the value of this expression is always equal to + I.
Proof: The local uniqueness of a regular solution is a direct consequence of the inverse function theorem (see Section M.E of the Mathematical Appendix). Intuitively, this is clear enough. For any infinitesimal change in normalized prices, dp = (dp" ... , dpL_" 0) "" O,the nonsingularity of D:(p) implies that D:(p) dp "" O. Hence, we cannot remain at equilibrium. Once we know that every equilibrium is locally isolated, the finiteness of the number of equilibria is a consequence of the boundary condition (v) of Proposition 17.8.2 on the excess demand function. Because of this condition (which, recall, follows from the strong monotonicity of preferences), equilibrium is not compatible with relative prices that are arbitrarily close to zero. That is, there is an r > 0 such that if zIp) = 0 and P,. = I, then I/r < Pr < r for every I. The continuity of i(') adds to this the fact that the set of equilibrium price vectors is a closed subset of R"- '. But a set that is closed and bounded (i.e., compact) in R"-' and discrete (i.e., with all its points locally isolated) must necessarily be finite (see Section M.F of the Mathematical Appendix). _
Proposition 17.0.2: (The Index Theorem) For any regular economy. we have
L \p:.ttpl ;'O'P 1
index P =
+1.
= 11
A brief discussion of why this result is true is given at the end of this section. llere we point out some of its implications and why it is useful and significant. Note. first. that it implies that the number of equilibria of a regular economy is odd.'6 In particular, this number cannot be zero; so the existence of at least one equilibrium is a particular case of the proposition. Second, the index concept provides a classification of equilibria into two types. In a sense, Proposition 17.0.2 tclls us that thc type with positive index is more fundamental because the presence of at least one equilibrium of positive type is unavoidable. In fact, it is typically the case that any scarch for well-behaved equilibria (what this means depends on the particular application) can be confined to the positive index equilibria. Third, as we shall see in Section 17.F, the index result has implications for the uniqueness and the multiplicity of equilibria. Fourth. as we shall discuss in Section 17.E, part of the importance of the index theorem is that this is all we can hope to derive without imposing (strong) additional assumptions. We next proceed to argue that typically (or, in the usual jargon, generically) economics arc regular. Hence, generically, the solutions to the excess demand equations are locally isolated and finite in number, and the index formula holds.'7
Our next aim is suggested by reexamining Figure 17.0.1; we see that for a regular economy with two commodities, we can assert more than the finiteness of the number of equilibria. Indeed, the boundary conditions on the excess demand function z,(·) (excess demand is positive if PI is very low and negative if it is very high) necessarily imply that, for a regular economy, first, there is an odd number of equilibria and, second, the slopes of the excess demand function at the equilibrium must alternate between being negative and being positive, starting with negative. If we say that an equilibrium with an associated negative slope of excess demand has an index of + I and that one with a positive slope has an index of -I, then, no matter how many equilibria there are, the sum of the indices of the equilibria of a regular economy is always + I. With appropriate definitions, it turns out that this invariance of index property also holds in the general case with any number of commodities, where it has some important implications for comparative statics and uniqueness questions. Let us generalize the definition of the index of a regular equilibrium that we have just suggested for the case L = 2 to the case of many commodities.
(/eller;cily Arwlysis
To emphasize the wide scope of the methodology to be presented, we discuss it first in terms of a general system of equations. We then specialize our discussion to the economic problem at hand and spell out its consequences for the excess demand equations. The essence of genericity analysis rests on counting equations and unknowns. Suppose we have a system of M equations in N unknowns:
II(r, .... , ('.v) = 0,
Definition 17.0.2: Suppose that p = (p, • ...• PL _" 1) is a regular equilibrium of the economy. Then we denote
(17.0.1)
f"(r,, ... , v.v) = 0,
indexp = (_I)L-' sign ID2(p)I,
or, more compactly, I(v) = O. The normal situation should be one in which, with N unknowns and M equations, we have N - M degrees of freedom available for the
where ID2(p)1 is the determinant of the (L - 1) x (L - 1) matrix D2(p)15 If L = 2, then IOf(p)1 is merely the slope of ZI(') at p. Hence, we see that for this case the index is + I or -I according to whether the slope is negative or positive.
+1 or
~
A regular economy has a finite number of equilibria (Proposition 17.0.1). Therefore, for a regular economy, the expression
is locally isolated (or locally unique). That is. there is an c > 0 such that if p' "" P. pi = PL = 1. and lip' - pil < c, then z(p') "" O. Moreover, if the economy is regular. then the number of normalized equilibrium price vectors is finite.
15. For any number" '" 0, sign" =
... v .... A,
5 I:. C T
16. This result was first shown by Dierker (1972). 17. For advanced treatments on the topic of this section, refer to Balasko (1988) or Mas-Colell (InS).
-I according to whether" > 0 or " < O.
1.
o!:tJ
594
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
SECTION
LOCAL
UNIQUENESS
AND
THE
INDEX
f(·; q) if q' is close to q. Hence, the notion that the regularity of a system f(·; q) = 0 is typical, or generic, could be captured by demanding that for almost every q, f(·; q) = 0 be regular; in other words, that nonregular systems have probability zero of occurring (with respect to say, a nondegenerate normal distribution on RS).'9 It stands to reason that some condition will be required on the dependence of f(·; q) on q for this to hold. At the very least, f(·; q) has to actually depend on q. The important mathematical theorem to be presented next tells us that little beyond this is needed. 20
f'(u): Perlurbed S~stem
I(d
17.D:
Figure 17.0.4 The regular case is typical.
Proposition 17.0.3: (The Transversa/iCy Theorem) If the M x (N + S) matrix Of(v; q) has rank M whenever f(v; q) = 0 then for almost every q, the M x N matrix OJ(v; q) has rank M whenever f(v; q) = O.
description of the solution set. In particular, if M > N, the system should be overdetermilled and have no solution; if M = N, the system should be exactly determined with the solutions locally isolated; and if M < N, the system should be underdetermilled and the solutions not locally isolated. Clearly, all these statements are not always true (you can see this just by considering examples with linear equations). So, what does it mean to be in the "normal case"'? The implicit function theorem provides an answer: one needs the equations (which we assume are differentiable) to be independent (that is, truly distinct) at the solutions. Definition 17.0.3 captures this notion.
Heuristically, the assumption of the transversality theorem requires that there be enough variation in our universe. If Df(v; q) has rank M whenever f(v; q) = 0, then from any solution it is always possible to (differentially) alter the values of the function f in any prescribed direction by adjusting the v and q variables. The conclusion of the theorem is that, if this can always be done, then whenever we are initially at a nonregular situation an arbitrary random displacement in q breaks us away from nonregularity. In fanciful language, if our universe is nondegenerate, then so will be almost every world in it. Note one of the strengths of the theorem: the matrix Df(v; q) has M rows and N + S columns. Hence, if S is large, so that there are many perturbation parameters, then the assumption of the theorem is likely to be satisfied; after all, we only need to find M linearly independent columns. On the other hand, D../(v; q) has M rows but only N columns. It is thus harder to guarantee in advance that at a solution D.f(v; q) has M linearly independent columns. But the theorem tells us that this is so for almost every q. Observe that if M > N (more equations than unknowns), then the M x N matrix D.f(v; q) cannot have rank M. Hence, the theorem tells us that in this case, generically (i.e., for almost every q). f(v; q) = 0 has no solution.
Definition 17.0.3: The system of M equations in N unknowns f(v) = 0 is regular if rank Of(v) = M whenever f(v) = O. For a regular system, the implicit function theorem (see Section M.E of the Mathematical Appendix) yields the existence of the right number of degrees of freedom. Ir M < N, we can choose M variables corresponding to M linearly independent columns of Df(v) and we can express the values oC these M variables that solve the M equations f(v) = 0 as a Cunction oC the N - M remaining variables (see Exercise 17.0.2). Ir M = N, equilibria must be locally isolated for the same reasons as discussed earlier in this section for the system zIp) = O. And iC M > N, then rank Df(v):s; N < M Cor all v; in this case, Definition 17.0.3 simply says that, as a matter of definition, the equation system f(v) = 0 is regular if and only if the system admits no solution. It remains to be argued that the regular case is the "normal" one. Figure 17.0.4 suggests how this can be approached. In the figure, the one-i:Quation, one-unknown system f(v) = 0 is not regular [because of the tangency point of the graph of f(·) and the horizontal axis]. But clearly this phenomenon is not robust: if we slightly perturb the equation in an arbitrary manner [say that the shocked system is /,(. )], we get a regular system. On the other hand, the regularity of a system that is already regular is preserved for any small perturbation.'· This intuitive idea of a perturbation can be formalized as follows. Suppose there are some parameters q = (q" ... , qs) such that, for every q, we have a system of equations f(v; q) = 0, as above. The set of possible parameter values is RS (or an open region of RS ). We can then justifiably say that f(·; q') is a perturbation of
Let us now specialize our discussion to the case of a system of L - I excess demand equations in L - I unknowns, i(p) = O. We have seen by example that nonregular economies are possible. We wish to argue that they are not typical. To 19. More formally. we could say that in a system defined by finitely many parameters (taking values in. say, an open set) a property is generiC in the first sense if it holds for a set of parameters of full measure (i.e.• the complement of the set for which it holds has measure zero). The property is gelwrk ;11 the .~econd .~ense if it holds in an open set of full measure. A full measure set is dense but it need not be open. Hence. the second sense is stronger than the first. Yet in many applications (all of ours in fact). the property under consideration holds in an open set, and so genericity in the first sense automatically yields genericity in the second sense. In some applications there is no finite number of parameters and no notion of measure to appeal to. In those cases we could say that a property is generic in the third sense if the property holds in an open and dense set. When no measure is available, this still provides a sensible way to capture the idea that the property is typical; bUI it should be nOled that with finitely many parameters a set may be open, dense and have arbitrarily small (positive) measure. In this entire section we deal with genericity in the first sense, and we simply call it genericity. 20. for this theorem, we assume that f(v; q) is as many times differentiable in its two arguments as is necessary.
18. The perturbation should control the values and Ihe derivatives of the funclion. In technical language. it should be a C' perturbation.
1
THEOREM
595
596
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EOUILIBRIUM
SEC T ION
do so, we could resort to a wide variety of perturbation parameters influencing preferences or endowments (or, in a more general setting, technologies). A natural set of parameters are the initial endowments themselves:
L 0 CAL
U N I QUE N E S SAN 0
THE
I N 0 EXT H E 0 REM
Because of the index theorem, this picture, in which the number of equilibria changes discontinuously from 3 to 1 at some points in the space of endowments is typical of the multiple-equilibrium case. A very extensive analysis of this equilibrium set has been carried out by Balasko (1988).
We can write the dependence of the economy's excess demand function on endowments explicitly as zIp; w). We then have Proposition 17.0.4.
We conclude the discussion of genericity with two observations: First, the generic local determinateness of the theory extends to cases with externalities, taxes, or other "imperfections"
Proposition 17.0.4: For any p and w, rank D.,i(p; w) = L - 1.
leading to Ihe failure or the first welfare theorem. (See Exercise 17.0.6.) This should be clear from the generality of the malhematical techniques which. in essence. rely only on the ability 10 express the equilibria of the theory as the zeros of a natural system of equations with Ihe
Proof: It suffices to consider the endowments of a single consumer, say consumer I, and to show that the (L - 1) x L matrix D."z(p; w) has rank L - 1 [this implies that rank D.,i(p; w) = L - 1]. To show this, we can either compute D.. ,z(p; w) explicitly (Exercise 17.0.3) or simply note that any perturbation of w,' say dw" that leaves the wealth of consumer 1 at prices p unaltered will not change demand and therefore will change excess demand by exactly -dw,. Specifically, if p'dw , = 0 then, denoting dw , = (dw ll , .•• , dWL_I.,), we have D.. ,i(p; w) dWI = D..,z,(p; w) dw , = -d
same number of equations and unknowns. Second, "finiteness of the number of equilibria" is
a blunt conclusion. It is not the same if the "finite" stands for three or for a few million. lJnforlunatcly. shorl of going all the way to uniqueness condilions (as we do in Section 17.F). we have no lechnique that allows us to refine our conclusions. We want to put on record. however. that it should not be presumed that in all generality "finite" means "small." In this respect. we mention, tentatively, thai there seems to be a distinction between market equilibrium situations ror which the first welfare theorem holds (in which, indeed. examples wilh "many" elluilihria seem contrived) and situations with a variety of market failures (where examples arc easy to produce). See Exercise 17.0.7 and the discussion on "sunspots" in Section 19.F.
We are now ready to state the main result [due to Oebreu (1970)]. Proposition 17.0.5: For almost every vector of initial endowments (w" ... , w,) the economy defined by {(;::i' wiJl::l is regular."
1 7 • D:
0" ,h" ["i/ex TI,,'an'lII
E R~/+,
The index resuit (Proposilion 17.0.2) is, in its essence, a purely mathematical fact. An attempt at a rigorous proof would take us too rar afield. Nonetheless. it is instructive to give an argument for its validity. It is an argument, we note incidentally. that can be made into a rigorous proof.
Proof: Because of Proposition 17.0.4, the result follows from the transversality theorem (Proposition 17.0.3). •
Denote our given, normalized, excess demand function by ~(p). We begin by availing oursclves of some other excess demand function iO(p) with the properties that (i) there is a unique (> such that zO(ji) = 0 and (ii) signIDio(ji)1 = (_I)c-I. For example, zO(p) could be generaled from a single-consumer Cobb-Douglas economy (Exercise 17.0.8). The idea is that fO( p) is both simple and familiar to us and that, as a consequence, we can use it to learn aboul the properties of the unfamiliar t( pl. Consider the following one· parameter family (in technical language, a homotopy) of excess
See Exercises 17.0.4 to 17.0.6 for variations on the theme of Proposition 17.0.4. In Figure 17.0.5, we represent the equilibrium set E = {(w" w 2 , PI): Z(PI' 1; w) = O} of an Edgeworth box economy with total endowment w = w, + w 2 • The set E is the graph of the correspondence that assigns equilibrium prices to economies w = (w" W2)'
demand functions:
P,
zip, I) = I:(p)
___""--;-----'7 0,
Figure 17.0.5 The equilibrium scI.
+ (I
- I)io(p)
for 0 :5 I :5 I.
Thc syslcm i(p, I) = 0 has L - I equations and L unknowns: (P,'··.' PC-I,I). Typically, Iherefore, the solution set £ = ((p,I): z(p, I) = OJ has one and only one degree of freedom al any of its points (that is, it looks locally like a segment). Moreover, since this solution set cannot escape to infinite or zero prices (because of the boundary conditions on excess demand) and is closed [because or the continuity of t(p, I)], it follows that the general situation is well represented in Figure 17.0.6. In Figure 17.0.6, we depict £ as formed, so to speak, by a finite number of circle·like and segmenl-like components, with the endpoints of the segments at Ihe I = 0 and I = I boundaries. Since Ihere arc two endpoints per segment, there is an even number of such endpoints. By construclion, (> is the only endpoint at the I = 0 boundary." Therefore, there must be an odd number of endpoints at the I = 1 boundary; that is, there is an odd number of solutions to :(p) = z(P. I) = O. Suppose now that we follow a segment from end to end. Whal
21. To be quite explicit, this means that the set of endowments that yield nonregular economies
is a subset or RLI that has (LI-dimensional) Lebesgue measure zero, or, equivalently. probability zero for. say, a nondegenerate LI·dimensional normal distribution.
22. More generally, if z(p: I) is an .rbilrary homolopy. then the typical situalion is well reprcscnled by any or the Figures 17.0.I(a), (b), or (c).
597
598
CHAPTER
17:
THE
POSITIVE
Index: +
THEORY
OF
EQUILIBRIUM
+)
I= I
FIgure 17.0.6
The equilibrium set under a homotopy.
is Iho rdalion belween Ihe indices al the two ends'? A moment's reflection (keeping the il11plicit function theorem in mind) reveals that as long as we move in a given
diroction relative to I (i.e., forward or backward), the index, (_I)L -, sign IDpi(p, 1)1, does not change. and that the index changes sign precisely when we reverse direction.2) Now, a segment that begins ~tnd ends at the same boundary must reverse direction an odd number or times; heneo. tho indices at the two endpoints have opposite signs. You can verify this in Figure 17.0.6. Therefore, the sum of the indices at I = I equals the index of the lone equilibrium of i(') connected by a segment to the equilibrium p oUo(.) at the boundary I = O. It is represented by in Figure 17.0.6. The segment that connects;; to p. in £ reverses directions an even number of times (possibly none); therefore, we conclude that the index of this equilibrium at I = I equals the index of p for to(.), which, by construction, is + I. Hence, the sum of the indices al I = I is + I, as Proposition 17.D.2 asserts to be true in complete generality.
"oz,(p) L...
,
,,*
ANYTHING
aI',
_ 0 1',-
GOES:
THE
SONNENSCHEIN-MANTEL-DEBREU
for all ( and I' [or Dz(p)p
" "z,(p) L... 1', - - = -Zt(p) , apt
= 0]
(17.E.I)
(17.E.2)
for all ( and p [or p'Dz(p) = -z(p)]
These arc the excess demand counterparts of expressions (2.E.I) and (2.E.4) for demand functions. They follow, respectively, from the homogeneity of degree zero and the Walras' law properties of excess demand. More interestingly, from z(p) = L; (x;(p, P'w;) - w;) we also get
17.E Anything Goes: The Sonnenschein-Mantel-Debreu Theorem
(17.E.3)
We have seen that under a number of general assumptions (of which the most substantial concerns convexity), an equilibrium exists and the number of equilibria is typically finite. Those are important properties, but we would like to know if we could say more, especially for predictive or comparative-statics purposes (see Section 17.G). We may well suspect by now (especially if the message of Chapter 4 on the difficulties of demand aggregation has been well understood) that the answer is likely to be negative; that is, that, in general, we will not be able to impose further reslrictions on excess demand than those in Proposition 17.8.2, and therefore that no further general restrictions on the nature of Walrasian equilibria than those already studied can be hoped for. Special assumptions will have to be made to derive stronger implications (such as uniqueness; see Section 17.F). I n this section, we confirm this and bring home the negative message in a particularly strong manner. The theme, culminating in Propositions 17.E.3 and 17.E.4, is: AII)'llrillg satisjying tire jew properties tlrat we have already shown must Iwld,
call
17.E:
The analysis that follows develops the logic of this conclusion through a series of intermediate results that have independent interest. Some readers may wish, in a first reading of this section, to skip these results and examine directly the statements of Propositions 17.E.3 and 17.E.4 and the accompanying discussion of their interpretations. To be specific, we concentrate the analysis, as usual, on exchange economies formalizcd by means of excess demand equations. Focusing on exchange economies makes sense because, as we know from Chapter 5, aggregation effects are unproblematic in production. The source of the aggregation problem rests squarely with the wealth effects of the consumption side. We begin by posing a relatively simple but nonetheless quite important question: To what extent can we derive restrictions on the behavior of excess demand at a given price p. In particular, we ask for possible restrictions on the L x L matrix of price effects DZ(p).24 Suppose that z(p) is a differentiable aggregate excess demand function. In Exercise 17.E.I, you are asked to show that
Five + Equilibria Pall = I
~----..
+ 1=0
SECTION
where, as usual, S;(p, P 'w;) is the substitution matrix (see Exercise I7.E.2). Expression (I7.E.3) is very instructive. It tells us that if it were not for the wealth effects, Dz( 1') would inherit the negative semidefiniteness (n.s.d.) property of the substitution matrices. How much havoc can the wealth effects cause? Notice that the matrix
i!!~, p'w,l z,,(p) ow,
D., ',(I', p·w.)Z,(p)T
=
...
iJxI,(p, P'w,) ZU
iJw,
.
[ iJxL,(p, P'w,) iJw,
()
z" P
OXL,(p, P'w,)
ow;
.() le, p
is of rank I (any two columns, or rows, are proportional). Therefore, we could informally surmise that the wealth effect of consumer i can hurt in at most one
aClually OCCllr.
24. Note that z(p) can take any value. You need only specify a consumer with an endowment vector w such that w + z(p)>> 0 and then choose a utility function that has OJ + z(p) as the
23. To see Ihis, think of the case where L = 2. Applying the implicit function Iheorem to i,(p,. t) = O. verify Ihen that a reversal of direction occurs precisely where OZ,(p"I)/iJp, = O.
demanded point
1
THEOREM
599
600
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
SECTION
direction of price change." Thus, we should expect that if I < L then there are some negative semidefiniteness restrictions left on Dz( pl. That this is the case is formalized in Proposition 17.E.1.
:L z,(p) =
0, at most I of the I
+
THE
= L (lip, )e'(p,a')
,
SONNENSCHEIN-MANTEL-DE8REU
THEOREM
601
= L, e' a' = A.
and so we would have accomplished our objective. Can we find these L consumers? The answer is "yes." Begin by choosing a collection of
I vectors,
endowments (w\ ... . , w,) yielding strictly positive consumptions when excess demands arc =,(p) = _p,(a')T; that is, x, = w,- p,(a')T» 0 for every i. Observe then that, for cvery i = I. ... , L, thc candidate individual excess demand satisfies Wal",s' law
can be linearly independent. Since I < L, it follows that we can find a nonzero vector dp E RL such that p·dp = 0 and z,(p)·dp = 0 for all i.ln words: dp is a nonproportional price change that is compensated (i.e., there is no change in real wealth) for every consumer. But then from (17.EJ) we obtain
L dp'S,(p, P'w,) dp 5: O.
GOES:
Dz(p) = - L D.. x,(p, P'w,)Z,(p)T
(P,ZI(P), ... ,z,(p)} c IRL
dp'Dz(p) dp =
ANYTHING
and
Proposition 17.E.1: Suppose that 1< L. Then for any equilibrium price vector p there is some direction of price change dp ¥ 0 such that p'dp = 0 (hence, dp is not proportional to p) and dp'Dz(p) dp 5: O. Proof: Because z( p) =
17.E:
P'z,(p)
= -p,p'a' = 0
(because Ap = 0),
and. also. that the candidate wealth effect vector satisfies the necessary condition of Proposition 2.E.3 p·D..,x,(p, P'w,)
•
=(llp,)p'e' = I.
Figure 17.E.1 should thcn be persuasive enough in convincing us that wc can assign prdacnccs to i = I, ...• L in such a way that the chosen consumption at p is Xi. the wealth efTect vector at p is proportional to f' (and therefore must equal (1Ip,Je')." and the indifference map has a kink at x,. The figure illustrates the complete construction for the case /. = 2" In Exercise 17.E.3, you are asked to write an explicit utility function . •
Parallel reasoning should make us expect that if I ;e: L (i.e., if there are at least as many consumers as commodities), then there may not be any restriction left on Dz(p) beyond (17.E.I) and (17.E.2). After all, thc direction of an individual wealth effect vector at a given price is quitc arbitrary (and can be chosen independently of the substitution effects of the corresponding individual); and with I ;e: L wealth effect vectors to be specified, there is considerable room to maneuver. Proposition 17.E.2 confirms this suspicion.
\"li
Figure 17.E.1
Decomposition of excess demand and
price effects at a price vector p (for L = 2).
Proposition 17.E.2: Given a price vector p, let z E RL be an arbitrary vector and A an arbitrary L x L matrix satisfying p'z = 0, Ap = 0 and p'A = -z. Then there is a collection of L consumers generating an aggregate excess demand function z(·) such that zIp) = z and Dz(p) = A. Proof: To keep the argument simple, we restrict ourselves to a search for consumers that at their demanded vectors have a null substitution matrix, SlIp, P'w,) = 0, that is, whose indifference sets exhibit a vertex at the chosen poin!." We can always formally rewrite the given L x L matrix A as A = Le'a',
,
where e' is the (th unit column vector (i.e., all the entries of e' are 0 except the (th entry, which equals I) and a' is the Ith row of A [i.e., at = (an, .. . , a,d)' Suppose now that we could specify L consumers, i = I, ... , L, with the property that, for every i, consumer i has, at the price vector p, an excess demand vector z,(p) = _p,(a')T, a wealth effect vector Dw,x,(p, P'w,) = (1/p,)e', and a substitution matrix S,(p, P'w,) = 0 (where 1 L l L 0 , ••. ,a and e , •.. ,e are as defined above). Then we would have both zIp)
= LZ,(P) = -LP,(a')T = _ATp = , ,
-p'A
27. Indeed. if Dxj(p, P'w j ) = (t/e i , then 1 = p'Dxj(p, P'w i ) = "iP'e l = rlfpj. Hence. (t, = lip;. 28. At no extra cost, we could actually accomplish a bit more. We could also require the substitution matrices of the consumers i = I•...• L to be any arbitrary collection of L x L matrices S, satisfying the properties: Sj is symmetric. negative semidefinite, p,Sj = 0, and SiP = O. The spccil1cation of consumers generating excess demand z( p) and excess demand effects D:(p) at p would proceed in a manner similar (0 the proof just given except that the argument would now be applied to A - L S,. By using matrices S, of maximal rank (i.e.• of rank L - I), we could insure that the resulting L consumers display smooth indifference sets at their chosen consumptions.
=z
25. For example. it cannot hurt in any direction of price change that is orthogonal to the weallh effects vector D..,xj(p. P"W j ) or to the excess demand vector Zj(p). A more precise argument is given
in Proposition 17.E.1. 26. The term "vertex" refers to what is usually called a "kink" in the case L
= 2.
i
602
CHAPTER
17:
THE
POSITIVE
THEORY
OF
SECTION
EQUILIBRIUM
17.E:
ANYTHING
GOES:
THE
SONNENSCHEIN-MANTEL-OEBREU
Up to now, we have studied the possibility of restrictions on the behavior of excess demand at a single price vector. Although the results of Propositions 17.E.! and !7.E.2 are already quite useful, we can go further. The essence of the negative point being made is, unfortunately, much more general. Consider an arbitrary function z(p), and let us for the moment sidestep boundary issues by having zIp) be defined on a domain where relative prices are bounded away from zero; that is, for a small constant £ > 0, we consider only price vectors p with pt!Pt. ;:: £ for every ( and t'. We could then ask: "Can z(·) coincide with the excess demand function of an economy for every p in its domai~?" Of course, in its domain, z(·) must fulfill three obvious necessary conditions: it must be continuous, it must be homogeneous of degrec zero, and it must satisfy Walras' law. But for any z{·) satisfying these three conditions, it turns out that the answer is, again, .. yes.""
{p
E R~:
ptfPt· ;::
£
Ftgure 17.E.2 preferences (in the case L = 2) for the offer curve of an excess
demand function :J') such that ZI( p) = 0 h", no solution with 1/& < PI/P, < 1:.
for every ( and t'} U!.
and with values in RL. Assume that, in addition, z(·) is homogeneous of degree zero and satisfies Walras' law. Then there is an economy of L consumers whose aggregate excess demand function cOincides with zIp) in the domain p'.'O
= tZI(P)
+r
the initial endowment point. We then see in the figure that no matter how complicated the
Strictly speaking, Proposition 17.E.3 docs not yet settle our original question, "0111 we assert ill1ything more about the equilibria of an economy than what we have derived in Sections 17.e and 17.D?" The problem is that Proposition 17.E.3 characterizes the behavior or excess demand away rrom the boundary, whereas it is the power or the boundary conditions that yields some or the restrictions we have already established: existence, (generic) finiteness, oddness. the index rormula.)( To argue that we cannot hope ror more restrictions than these on the equilibrium set, we need to guarantee that ir a candidate equilibrium set satisfies them, then the construction or the "explaining" economy will not add new equilibria. The result presented in Proposition 17.E.4, whose proor we omit, provides therefore the final answer to our question."
[accordingly, zHp) = -(PI/P,)Z:(p)] [accordingly, z~(p)
p.
olT(:r curve! m~ly otherwise be, we can always fit an indifference map so that for any p E ~ we generale precisely the demands flJ j + :i(p). •
and :;(p) = !ZI(P) - r
+ :I( p) is the intersection point of the offer curve with the budget line perpendicular to
The olTer curve is continuous and. because Zi(p) = 0 has no solution in p,. it does not touch
Proof: AI the end of this section, we olTer (in small·type) a brief discussion of the general proof of this result. Here, we limit ourselves to the comparatively simple ease where L = 2. Suppose then Ihat L = 2 and that an & > 0 and a function z(·) satisfying the assumption of the proposition are given to us. The continuity and homogeneity of degree zero of z(·) imply the existence of a number r> 0 such that IZI(p)1 < r for every peP,. We now specify two functions Zl(.) and z'(·) with domain P, and values in A', which are also continuous and homogeneous of degree zero, and satisfy Walras' law. In particular, we let z:(p)
603
Construction of
Proposition 17.E.3: Suppose that z(·) is a continuous function defined on
p. =
THEOREM
= -(PI/P,)Z;(p)].
Note that zIp) = Zl(p) + z'(p) for every peP,. We shall show that for i = 1,2 the function Zl(.) coincides in the domain P, with the excess demand function of a consumer. To this elTect, we usc the following properties of Zl(.): continuity, homogeneity of degree zero, satisfaction of Walras law, and the fact that there is no peP, such that Zl(p) = O. In Exercise 17.E.4, you are asked to show by example that this last requirement is needed. Choose a WI » 0 such that WI + Zl(p) » 0 for every pep'. In Figure 17.E.2, we represent the offer curve OCI associated with Zl(.) in the domain p,. In the figure, for every pep',
PropOSition 17.E.4: For any N ;:: 1, suppose that we assign to each n = 1, ... ,N a price vector pn, normalized to IIpnll = 1, and an L x L matrix An of rank L - 1, satisfying Anpn = 0 and pn'A n = O. Suppose thaI. in addition, the index formula Ln (_1)L -1 sign IAnl = + 1 holds.)) If L = 2, assume also that positive and negative index equilibria alternate. Then there is an economy with L consumers such that the aggregate excess demand z( .) has the properties:
29. The question was posed by Sonnenschein (1973). He conjectured that the answer was that. indeed, On the domain where PI ~ t for all It the three properties were not only necessary but also suil1cicnl; that is. we could always find such an economy. He also proved that this is so for the
(i) zIp) = 0 for Ilpll = 1 if and only if p = pn for some n. (ii) Dz(pn) = An for every n.
two·eommodilY case. The problem was then solved by Mantel (1974) for any number of commodities. Mantel made use of 2L consumers. Shortly afterwards, Debreu (1974) gave a different and very simple proof requiring the indispensable minimum of L consumers. This was topped by Mantel (1976), who refined his earlier proof to show that L homothetic consumers (with no
31. Note, for example, that although a candidate function z(·) defined on
p, may
not have any
solution. we can still successfully generate it from an economy. What happens. of course, is that the equilibria of the economy (which must exist) are all outside of p.:.
restrictions in their initial endowments) would do.
30. NOle. in particular, that this result implies that for any / ~ L, there is an economy of I consumers Ihat generates z(·) on p,. We need only add to the L eonsumers identified by the
32. For this and more general results, see Mas-Colell (1977). 33. Here. A" is the L - I x L - 1 matrix obtained by deleting one row and corresponding column from A.
proposition J - L consumers who have no endowments (or. alternatively. whose most preferred consumption bundle at all price vectors in p.: is their endowment vector).
1
604
CHAPTER
17:
THE
POSITIVE
THEORY
OF
-
EQUILIBRIUM
Proposition 17.E.4 tells us that for any finite collection of price vectors {pl •... , pN} and matrices of price effects {DZ(pl), ... , DZ(pN)}, we can find an economy with L consumers for which these price vectors are equilibrium price vectors and {DZ(pl), ... , DZ(pN)}. are the corresponding price effects at these equilibria. The result implies that to derive further restrictions on Walrasian equilibria we will need to make additional (and, as we shall see, strong) assumptions. This is the subject of the next three sections. An excellent survey for further reading on the topic of this section is Shafer and Sonnenschein (1982).
SECTION
11.E:
ANYTHING
uOES:
THE
SONNEHSCHEIN-MANTEL-OEBAEU
p Revealed Preferred to p' Weak Axiom Satisfied
Figure 17 .E.4
We should point out that the initial endowments of the consumers obtained by means of Propositions 17.E.2, 17.E.3 or 17.E.4 are not a priori limited in any way. If there are constraints on permissible initial endowments. the nonnegativity conditions on consumption come into play and there may, in fact, be other restrictions on the function z(·). For example, you are asked in Exercise 17.E.5 to verify that the excess demand vectors z( p) and z( p') represented in Figure 17.E.3 cannot be decomposed into individual excess demand functions generated by rational preferences if the amount of any commodity that any consumer may possess as an initial endowment is prescribed to be at most I and if consumptions must be nonnegative.
Revealed preference for excess demand.
z'(p) #< z'(p')
and
P'z'(p')!> 0
(17.E.4)
(see Figure 17.E.4). We say that p is indirectly re""altd preferred 10 p' if there is a finite chain p' ..... p' such that p' = p, p" = p', and p' is directly revealed preferred to p" I for all II !> N - I. The SA then says:
=,
For "1'"'.1' p "lid p'. if p is (direc·tly or indirectly) revealed preferred preJ<·rred to p.
f()
P'. then p' call not be
(t/irC'<"Ily) r,'pealed
From now on, we let prices be normalized. A convenient normalization here is
IIpil' = P'p = L,. (Pt)' = I.
, ,,, , --..... ---(-I, -I):
=,
THEOREM
We say that an excess demand function z'(·) is proportionally one-IO-ont if p #< p' implies that z'(p) is not proportional to z'(p'); in particular, we have z'(p) #< z'(p'}. For proportionally one-to-one excess demand functions (and normalized prices), we can restate the "directly revealed preferred" definition (17.E.4) as
Figure 17 .E.3
Excess demand choices that cannot be decomposed due to boundary constraints.
p#< p'
and
p,z'(p')!> O.
(17.E.4')
Suppose that .,(.) is an arbitrary real-valued function of p such that .,(p) > 0 for all p € P,. The hasic observation of the proof is then the following: if z'(·) is a proportionally one-to-on" l'xC('ss clt'mand /utlcliml that 5alisjies the SA, then lhe same propert;~s are trut of the function ,,(. )z'(·). Indeed, for any p and p' the revealed preference inequalities (17.E.4') hold for z'(·) if 1II1<[ 011/.1' if they hold for .,(. }z'(·), and if z'(-) is proportionally one-to-one. then so is .,(. )z'(·). This observation suggests a way to proceed. We could look for L proportionally one-to-one excess demand functions z'(·) satisfying the SA and such that, at every p € P" the vectors {z'(p) •...• ZL(p)} constitute a basis capable of spanning z(p) by means of a strictly positive linear combination, that is, such that for every p € P, we can write z(p) = L, o,(p)z'(p) for some numbers ',( p) > O. This is precisely what we will now do. For every normalized p € P,. denote T, = {z € RL: p'Z = O} and for every i = I, ... , L, let :'( p) € Tp be the point that minimizes the Euclidean distance liz - e'li (or, equivalently, maximizes the concave "utility function" -liz - ill) for z € Tp , where e' is the ith unit vector (the column vector whose ith entry is I with zeros elsewhere). Geometrically, z'( p) is the perpendicular projection of e' on the budget hyperplane Tp; that is, z'(p) = e' - p, P. where p, is the ith component of the vector p (recall that i :s; L). Then z'(·) is proportionally one-to·one (see Exercise 17.E.6) and satisfies the SA (since it is derived from utility maximization; see also Exercise 17.E.7). Now let r > 0 be a large-enough number for us to have z(p) + rp » 0 for every normalized p € P, [such an r exists by the continuity of z( . ) and the fact that the set of normalized price vectors in P, is compact and includes only strictly positive price vectors]. For every i = I •...• L and every normalized p € P" define o,(p) = z,(p) + rp, > O. where z,(p) is the ith component
Proof of Proposition '7.E.3 continued: Although a complete proof of the proposition for the case of any number of commodities would take us too far afield. the essentials of the proof by Debreu (1974) are actually not too difficult to convey. We shall attempt to do so. We note that. when carefully examined, the proof can be seen as a generalization of the argument for the L = 2 case presented earlier. In Section 3.1. we saw that the strong axiom of revealed preference (SA) for demand functions is equivalent to the existence of rationalizing preferences. The same is true for excess demand functions: If an excess demand function z'(·) satisfies the SA (we will give a precise definition in a moment), then z'(-) can be generated from rational preferences." It is thus reasonable to redefine our problem as: Given a function z(') that. on the domain P" is continuous. homogeneous of degree zero, and satisfies Walras' law (for short, we refer to these functions as excess demand functions). can we find L excess demand functions z'(·). each satisfying the SA. such that L, z'(p) = z(p} for every p E P,? Before proceeding, let us define the SA for an excess demand function z'(·). The definition is just a natural adaptation of the definition for demand functions. We say that p is directly reveald preferred to p' if
34. We refer to the proof of Proposition 3.1.1 for the justification of this claim.
.L
605
,
606
CHAPTER
17:
THE
POSITIVE
THEORY
OF
SECTION
EQUILIBRIUM
p,
.~
/ --<'
!
p
:
~
-;(p)
OF
Suppose that the production side of the economy is given to us by an arbitrary technology Y c IRL of the constant returns, convex type (i.e., Y is a convex cone). What conditions involving only the demand side of the economy guarantee the uniqucness of equilibrium allocations?36 From the analysis presented in Section 4.0, we already know one answer: if a welfare authority makes sure that wealth is always distributed so as to maximize a (strictly concave) social welfare function, then the economy admits a (strictly concave) normative representative consumer, and the equilibrium necessarily corresponds to the unique Pareto optimum of this oneconsumer economy (as in Section IS.C). In our current framework, however, this is not a promising approach because wealth is derived from initial endowments and only by coincidence can we expect that the induced distribution of wealth maximizes a social welfare function. We will therefore concentrate on a weaker and, for the purpose at hand, more interesting condition: that the weak axiom of revealed preferellce holds for the aggregate excess demand of the consumers. To begin, suppose that z(p) = :L (x;(p, p'w;) - 00;) is the aggregate excess demand function of the consumers. For this economy with production, Proposition 17.F.I provides a useful restatement of the definition of Walrasian equilibrium (Dcfinition 17.B.I) in terms of ;(').
+ rp
Figure 17 .E.5
:"(P)
+ 'P,
Illustration of the
P,
construction
or
individual excess demand in the proof of Proposition 17.E.3.
of the vector ZIP). We claim. that L; a;(p)z;(p)
= zIP)
= zIp).
and this concludes the proof. Indeed.
-0
+ 'p
- 'p
Proposition 17.F.l: Given an economy specified by the constant returns technology Y and the aggregate excess demand function of the consumers z(·), a price vector p is a Walrasian equilibrium price vector if and only if
= zIp). Geometrically. what we are doing is projecting every a;(p)e ' On the hyperplane 0:. By the definition of .;(p). we have L; o;(p)e; zIp) + 'p. Therefore. when we project both sides. we get L; o;(p);;(p) ZIP). The construction is illustrated in Figure 17.E.5. • :: E ~I.: p';
UNIQUENESS
The Weak Axiom for Aggregate Excess Demand z,(p)+,p, ------------,. ZIP) , I /1
;'(p)
17.F:
=
=
=
(i) p' Y :5: 0 for every Y E Y, and (ii) z(p) is a feasible production; that is. z(p)
E
Y.
Proof: If p is a Walrasian equilibrium price vector, then (ii) follows from market clearing and (i) is a necessary condition for profit maximization with a constant returns technology. In the other direction, if (i) and (ii) hold, then consumptions xt = x;(p, p 'w;) for i = I, ... , I, production vector y' = z(p) E Y, and price vector p constitute a Walrasian equilibrium. To verify this, the only condition that is not immediate is profit maximization. However, y' = z(p) E Y is profit maximizing because p" y :5: 0 for all y E Y [since (i) holds] and p' y' = p- z(p) = 0 (from Walras' law) . •
17.F Uniqueness of Equilibria Up to this point. we have concentrated on the determination of the general properties of the Walrasian equilibrium model. We now take a different tack. We focus on a particular. important property-the uniqueness of equilibrium-and we investigate conditions. necessarily special, under which it obtains!5 The presentation is organized into four headings. The first contemplates a general setting with production and discusses conditions on the demand side of the economy that, by themselves (i.e., without the help of further restrictions on the production side), guarantee the uniqueness of equilibrium. The second discusses the gross substitution property, an important class of conditions with uniqueness implications for exchange economies. The third presents a limited result that relics on the Pareto optimality property of equilibrium. The fourth analyzes the role of the index formula as a source of uniqueness and nonuniqueness results. Throughout Section 17.F, we assume that individual preferences arc continuous, strictly convex, and strongly monotone.
We next define the weak axiom for excess demand functions. Definition 17.F.1: (The Weak Axiom for Excess Demand Functions) The excess demand function z(·) satisfies the weak axiom of revealed preference (WA) if for any pair of price vectors p and P'. we have z(p)
oF z(p') and p,z(p') :5: 0 implies p"z(p) > O.
36. The sel Y can be thought as an aggregate produclion sel. The restriction that Y be of constant returns is made merely for convenience of exposition. It allows us. for example. not to worry about the distribution of profits to consumers (since profits are zero in any equilibrium). Note also that the constant returns model includes pure exchange as a special case (where y= -R';).
35. Reviews for this topic are Kehoe (1985) and (1991). and Mas-Colell (1991).
L
EQUILIBRIA
607
608
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
SECTION
17.F:
UNIQUENESS
OF
EOUILIBRIA
609
Proposition 17.F.2: Suppose that the excess demand function z(·) is such that, for any constant returns convex technology Y, the economy formed by z(·) and Y has a unique (normalized) equilibrium price vector. Then z(·) satisfies the weak axiom. Conversely, if z(·) satisfies the weak axiom then, lor any constant returns convex technology Y, the set 01 equilibrium price vectors is convex (and so, il the set of normalized price equilibria is finite, there can be at most one normalized price equilibrium). Figure 17.F.l
A violation of the weak axiom implies
multiplicily of equilibria for some Y.
In words, the definition says that if p is revealed preferred to p', then p' cannot be revealed preferred to p [i.e., z( p) cannot be affordable under p']. It is the same definition used in Sections I.G and 2.F, but now applied to excess demand functions.'" The axiom is always satisfied by the excess demand function of a single individual, but it is a strong condition for lI!JqreYlIre excess demand (see Section 4.C for a discussion of this point). We lirst note that, given:('), the WA is a necessary condition for us to be assured of a unique equilibrium for every possible convex, constant returns technology Y that :(.) is coupled with. To see this, suppose that the WA was violated; that is, suppose that for some p and p' we have z(p) # z(p'), p'z(p') ,.:; 0, and p'·z(p),.:; o. Then we claim that both p and p' are equilibrium prices for the convex, constant returns production set given by
Y'
I
= lYE R .: P'y,.:;
0 and P"y,.:; O}.
Figure 17.F.1 depicts this production set for the case L = 2. Note that we have :( p) E Y' and p' Y ,.:; 0 for every Y E Y·. Thus, by Proposition 17.F.I, p is an
Proof: The first part has already been shown. To verify the convexity of the set of equilibrium prices, suppose that p and p' are equilibrium price vectors for the constant returns convex technology Y; that is, z(p) E Y, z(p') E Y, and, for any )'E Y, P'y,.:; 0 and p"Y ,.:; O. Let p" = I1.p + (I - alp' for 11. E [0, I). Note, first, that p" •.\" = ap' Y + (I - a)p"J''':; 0 for any)'E r. To show that p" is an equilibrium, we therefore need only establish that :(1''') E Y. Because 0 = p",z(p") = I1.p,z(p") + (I - l1.)p'·z(p"), we have that either p·z(p"),.:; 0 or P"z(p")":; O. Suppose that the first possibility holds, so that P':(p")":; 0 [a parallel argument applies if, instead, p'·z(p"),.:; 0]. Since :(1') E Y we have p"·z(p),.:; O. But with p"':(p),,:; 0 and p·z(p")":; O. a contradiction to the WA can be avoided only if :(p") = :(p). Hence z(p") E y.lO •
We arc therefore led to focus attention on conditions on preferences and endowments of the I consumers guaranteeing that the aggregate excess demand function z( p) fulfills the WA. To begin with a relatively simple case, suppose that all the endowment vectors w, arc proportional among themselves; that is, that w, = (1,(ii, where w is the vector of total endowments and (1, ~ 0 are shares with (1, = I. I n such an economy, the distribution of wealth across consumers is independent of prices. Normalizing prices to P'w = I, the wealth of consumer I is (x, and =,(p) = x,(p, 11.,) - w,. The aggregate demand behavior ofa population of consumers with fixed wealth levels was studied in Section 4.C. We repeat our qualitative conclusion from there: if individual wealth levels remain fixed, the satisfaction of the WA by aggregate demand (or excess demand), although restrictive, is not implausible.'o
:L
equilibrium price vector. The same is true for p'. Since z(p) # z(p'), we conclude that the equilibrium is not unique for the economy formed by z(·) and the production set yo. What about sufficiency? The weak axiom is not quite a sufficient condition for uniqueness, but Proposition 17.F.2 shows that it does guarantee that for any convex, constant returns Y, rile set of equilibrium price vecrors ;s convex. Although this convexity property is certainly not the same as uniqueness, it has an immediate uniqueness implication: if an economy has only a finite number of (normalized) price equilibria (a generic situation according to Section 17.0),38 the equilibrium must be unique.
A proportionality assumption on initial endowments is not very tenable in a general equilibrium context. It is important, therefore. to ask which new effects are at work (relative to those studied in Section 4.C) when the distribution of endowments docs not satisfy this hypothesis. Unfortunately, it turns out that nonproportionality of endowments can reduce the likelihood of satisfaction of the weak axiom by aggregate excess demand. To see this, consider the relatively simple situation in which preferences arc homothetic. Recall from Sections 4.C and 4.0 that, when endowments arc proportional, this case is extremely well behaved; not only is the WA satisfied, but the model even admits a representative consumer. Yet, as we proceed to discuss
37. A formal, and inessential, difference is that we now define the revealed preference relation on the budget sets (i.e., on price vectors) directly rather than on the choices (i.e., on commodity
39. Observe thai we have cSlablished thai eilher :(p") = z(p) or :(p") = z(p'). Since this is true for any" E [0, I]. and since Ihe function z(') is conlinuous, this implies Ihal z(p) = z(p') for any
vectors).
two equilibrium price vectors p and p'; that is, if the WA holds for =('), then every Walrasian equilibrium for the given endowments must have the same aggregate consumption vector and. hence. the same Llggrcgate production vector.
38. Although our discussion in Section 17.D focused on the case of exchange economies, its conclusions regarding generic local uniqueness and finiteness of the equilibrium set can be extended to the present production context.
40. On this point. consull also Ihe references given in Chapter 4, especially Hildenbrand (t994).
-
610
C HAP T E R
1 7:
THE
PO SIT lYE
THE 0 R Y
0 F
E QUI LIB R I U M
SECTION 17,F:
UNtQUENESS
OF
EQUtLtBRIA
611
-------------------------------------------------------------------------------------------------below (in small type), even with homothetic preferences, the WA can easily be violated when endowments are not proportional.·'
Example 17.F.l: This is an example of a failure of the WA compatible with homotheticity and even with the property of gross substitution, which we will discuss shortly. Consider a
In Section 2.F we olfered a dilferential version of the WA for the case of demand functions. 1n a parallel fashion we can also do so for excess demand functions. It can be shown that a sufficient dilferential condition for the WA is
for only the first two goods; that is. he has an excess demand function Z,(P) = Z,(P" p,) that does not depend on p, and P. and. further. is such that z,,(p) = z.. (p) = 0 for all p. Similarly. consumer 2 has preferences and endowments for only the last two goods." We claim that if there is a price vector p' at which the excess demand of the two consumers is nonzero [i.e., z,(p') '" 0 and z,(p') '" 0]. then the aggregate excess demand cannot satisfy the WA. To sec this, choose (p" p,) and (p,. p.) arbitrarily, except that P,ZII(P') + P,Zll(P') < 0 and P,Zll(P') + P.z.,(p') < O. For a> O. take q = (p'" Pl' ap,. afi.) and q' = (ap" ap" Pl' p~). Then if a> 0 is sufticiently large. we have q,z(q') < 0 and q'·z(q) < 0 (Exercise 17.F.2). •
four·commodity economy with two consumers. Consumer I has preferences and endowments
dp'Dz(p) dp < 0 whenever dp'z(p) = 0 (i.e., whenever the price change is compensated) and dp is not proportional /0 p (i.e .• relative prices change).
(17.F.I)
Allowing for the first inequality to be weak, expression (l7.F.I) constitutes also a necessary condition. 42
Under the homotheticity assumption, we have
See Exercise 17.F.3 for yet another example.
I Dw-',(p. P'w,) = - - x,(p, p·w,). p.Wj
I
Denoting S,
= S,(p, P'w,), x, = .<,(P. p·w,). x = L, x, and w = L, w"
Gross Subslilutioll
this implies (recall 17.E.3)
We now investigate the implications of a condition of a different nature from the WA. We shall sec that it yields a uniqueness result for situations that are reducible to formalization as exchange economics. To motivate (he concept (and justify its name), consider the demand function of a consumer in a two-commodity situation. At given prices the demand substitution matrix has negative diagonal entries and, as a consequence, positive off-diagonal entries: if the price of one good is raised, the compensated demand for the other good increases. However. if we do not net out the wealth effects (i.e., if we look at the effect of prices on uncompensated demand), then it is possible for an increase in the price of one good to decrease the demand of the two goods: in gross terms, the two goods may be complements. We say that the two goods are gross substitutes if this does not happen, that is. if an increase in the price of one good decreases the (uncompensated or gross) demand for that good and increases the (uncompensated or gross) demand for the other good. By extension. the same term is used in the L-commodity case for the property that asserts that when a price of one good increases, the demand of every other good increases (and, therefore, the demand of that good decreases). For L > 2, however, this is not by any means a necessary property of even compensated demand. In fact, the gross substitute property is very restrictive. Nonetheless, it can make sense for problems with a few very aggregated commodities or for those where commodities possess special symmetries (see Exercise 17.F.4).
1
Dz(p) = LS,(P.P·w,) - L xI(p,P'w,)z,(p,p'w,)T / j P'W i
=
L S, - L _1_ i p·Wj
i
+ L _1i
P'Wi
[x, _~'~' .i][X' _P'(~' p'W
.,]'
p'W
[x, _P'~' x][w, _P'~' w]' - _,-= xz(p)" p'W p'w P'w
(17.F.2)
For any direction of price change dp with dp' zIp) = 0, the first two terms on the right·hand side of equation (17.F.2) generate an elfect of the appropriate sign [the L x L substitution matrices S,(P. P'w,) and variance matrices
r
-[XI -:'.: x][x, -:'.: x
arc negative semidefinite]. the fourth is null, but the third is ambiguous. It is quite possible for this covariance term to have the wrong sign (positive) and even for it to overcome the other two terms." The situation to worry about is when (l/(p'w,llx,(p,p'w,) and WI arc positively associated within the popUlation of consumers; that is, when the consumers who consume (per dollar) more than the average (per dollar) consumption of some commodities tend to be those that are reiatively well endowed (per dollar) with those commodities. It makes sense that this case will cause difficulties: If the price of a good increases, the consumers who are (net) sellers of the good (who are likely to be those relatively well endowed with it) experience a positive wealth elfect. whereas the consumers who arc (net) buyers experience a negative wealth elfect. Hence, an increase in the total demand for the good will ensue if (net) sellers COnsume relatively more of the good (per dollar) than (net) buyers.
DefinItion 17.F.2: The function z(·) has the gross substitute (GS) property if whenever P' and P are such that, for some t, PI > PI and Pk = Pk for k "f. t, we have Zk(P') > Zk(P) for k oF t.
41. To reinforce this point. it is also worth mentioning that. in fact. jf we are free to choose initial endowments, then the class of homothetic preferences imposes no restrictions on aggregate
If, as is the case here, we are dealing with the aggregate excess demand of an economy. then the fact that z(·) is also homogeneous of degree zero has the consequence that with gross substitution we also have z{(p') < zAp) whenever p' and pare related as in Definition 17.F.2. To see this, let p = a.p, where a. = PI/p{, Note that PI = PI and p, > p; for k "f. t. Then the homogeneity of degree zero of z(·)
demand. Indeed. as we noted in Section 17.E, the basic conclusion of Proposition 17.E.2 can still be obtained with the further restriction that preferences be homothetic. See Mantel (1976) and the su rvey of Shafer and Sonnenschein (1982). 42. Suppose that dp'z(p) = (p' - p),z(p) = O. Definition I7.F.1 implies then that dp'dz = (p' - p)'(z(p') - zIp)) s; O. Going to the dilferentiallimitand using the chain rule, it follows that dp·Oz(p)dp s; 0 whenever dp'z(p) = O. 41 But this cannot happen if the ,'(,(P, p·w . ) are collinear among themselves or if the collinear among themselves. See Exercise l7.F.1.
WI
are
44. Thus, this example can also be seen as a case of positive association between endowments and demands.
1.
01:'"
Ci:AP1ER
17.
THE
POSITIVE
THEORY
Of
~;
EQUILIBRIUM
z,
E (. T ION
( will respond positively to an increase in P•. But if response.'6 •
The offer Curve of a
"
Wli
UNIOUENESS
bl~
= 0, there will be no
Proof: It suffices that we show that z(p) = z( p') cannot occur whenever P and p' are two price vectors that are not collinear. By homogeneity of degree zero, we can assume that p' ~ p and p, = PI for some I. Now consider altering the price vector p' to obtain the price vector p in L - I steps, lowering (or keeping unaltered) the price of every commodity k '" I one at a time. By gross substitution, the excess demand of good I cannot decrease in any step, and, because P '" p', it will actually increase in at least one step. Hence, =/(P) > ZI(P') . • One might hope 10 establish uniqueness in economics with production by applying the as property to the production inclusive excess demand :(.). However, the direct use of the GS property in a production context is limited. Imagine. for example. a situation in which inputs
slIhslirllliolt. 45
Figure 17.F.2 represents the offer curve of a gross substitute excess demand function L = 2. As the relative price of good I increases, the excess demand for good I decreases and the excess demand for good 2 increases. An important characteristic ofthe gross substitute property, which follows directly from its definition, is that it is additive across excess demand Junctions. In particular, if the individual excess demand functions satisfy it, then the aggregate function does also.
and Olltputs ure distinct goods. If the price of an input increases, the demand for every other
input may decrease. not increase as the as property would require, simply because the optimal level of output decreases. Indirectly. though. the gross substitute concept may still be quite helpful. Recall. in particular. that at th. end of Section 17.B. we argu.d that it is always possible to
reduce a production economy to an exchange economy in which, in effect. consumers
c,change factor inputs and then engage in horne production using a freely availabl. constant returns technology. The aggregate excess demand in this derived exchange economy for factor inputs combines elements of both consumplion and production and may well satisfy the as property."
Example 17.1'.2: Consider a utility function of the form u,(x,) = L' UI/(x,,), [f -[xIiUi,(XIi)/ui,(xl/)] < I for all ( and Xli' then the resulting excess demand function z,(p) has the gross substitute property for any initial endowments (Exercise 17.F.5). This condition is satisfied by U,(X,) = (L, ~IiXt,)lJp for 0 < p < I (Exercise l7.F.5). The limits of these preferences as p -+ I and p .... 0 are preferences representable, respectively, by linear functions and by Cobb-Douglas utility functions (recall Exercise 3.C.6). As far as the gross substitution property is concerned, Cobb-Douglas preferences constitute a borderline case. Indeed, the excess demand function for good ( is then ZIi(P) = rt.1i(P·W,)/PI - wli . [f W" > 0, the excess demand for good
What is the relationship between gross substitution and Ihe w.ak axiom? Clearly, the \V A docs not imply the as property (the latter can be violated ev.n in quasilinear, one·consumer economies). The converse relationship is not as obvious. but it is nevertheless true that the GS property does not imply the WA. [n fact, Example 17.F.I. which viola led the \VA. could perfectly well satisfy as" There is. however. one connection that is important. The gross substitute property implies that If z(p) = 0 and ;(p') ¢' 0, then p,z(p') > O.
(17.F.3)
We shall not prove condition (17.FJ) here. For the case in which L = 2, you are asked for a proof in Exercise 17.F.7. To understand (17.F.3). note that if p is the price vector of an
45. It is worth mentioning that functions satisfying the GS property arise naturally in many economic contexts. For example. if A is an (L - I) x (L - I) input~output matrix and (.' E IR'.-'.
then (. - (I - A). satisfies the (w.ak) GS property as a function of a E R'.-' (see Appendix A of Chapter 5 for the interpretation of th.s. concepts). More generally, the equation system g(a) - a associated with the fixed-point problem [i.e., find a such that y(') = aJ of an ;naeas;nrl function r/: R~ - R~ [i. •.• g(.) ~ y(.') whenever a ~ a'J satisfies it (perhaps. again. in its weak version).
46. See also Grandmont (1992) for an interesting result wh.r. a Cobb-Douglas positive representalive consumer. and therefore GS excess demand, is derived from a requirement that at any given price, the choice behavior is widely dispersed (in a certain precise sense) across consumers. Grandmonl's is an example of a model in which the individual excess demand functions may not
Note that in these cases there is no homogeneity of degree zero or Walras' law~-conditions specific to general equilibrium applications-to complement the GS property. This is significant because exploration of the implications of the OS property without homogeneity of degree zero or Walras'
satisfy the gross substitute property but th. aggr.gat. function does. 47. See Mas-Colell (1991) and Exercis. 17.F.6 for further elaborations on this point. 48. Therefore, in view of Proposition 17.F.I. we know that in a constant returns economy the fulfillment of the GS property by the excess demand of the consumers does not imply the uniqueness
law.
of equilibrium.
as property. See
EQUILIBRIA
Proposition 17.F.3: An aggregate excess demand function z(') that satisfies the gross substitute property has at most one exchange equilibrium; that is, z(p) = 0 has at most one (normalized) solution.
gross substitute excess demand function.
tells us that 0 = zl(ii) - Z/(P) = zl(ii) - ZI(P') + Z/(P') - Z/(P), However, gross substitution implies that z/(ii) - z,(p') > 0 (change sequentially each price P; for k '" I to Pl' applying the GS property at each step), and so ZI(P') - Z/(P) < O. The differential version of gross substitution is clear enough: At every p, it must be that iJz.(p)/c1PI > 0 for k '" I; that is, the L x L matrix Dz(p) has positive off-diagonal entries. In addition, when z(·) is an aggregate excess demand function, homogeneity of degree zero implies that Dz(p)p = 0, and so c:z,(p)/c1p, < 0 for all I = I, ... , L: the diagonal entries of Dz(p) are all negative. If in these definitions the inequalities arc weak, one speaks of \\'(,lIk IJ'OSS
these conditions add substantially to the power of the
OF
III tile special case oj excllange economies if the gross substitute property holds for aggregate excess demand then equilibrium is unique.
Figure 17.F.2
'(pI
;::
Exercise 17.F.16 for an
1.
-
614
CHAPTER
17:
THE
POSITIVE
THEORY
OF
SECTION
EQUILIBRIUM
:,
'7.F:
UNIQUENESS
OF
price vector p for each consumer i, we have x, ;t,w, for all i. However, by the assumption of the proposition and the first welfare theorem, (w" ...• wc) is a Pareto optimal allocation and so we must have x, -,wc for all i. But then we can conclude that x, = w, for all i, because otherwise, by the strict convexity of preferences. the allocation Ox, + !w" ... , !x, + !w,) would be Pareto superior to (w" . ..• w,) . •
Il1dex Al1alysis and Ul1iqueness ( ... and NOl1ul1iqueness) =1 - -__ :(p)
Figure 17.F.3
The index theorem (Proposition 17.D.2) provides a device to test for uniqueness in any given model. The idea is that if merely from the general maintained assumptions of the model we can attach a definite sign to the determinant of the Jacobian matrix of the equilibrium equations at any solution point, then the equilibrium must be unique. After all, the index theorem implies that sign uniformity across equilibria is impossible if there is multiplicity. As a matter of fact, we could have proceeded by means of this index methodology for many of our previous uniqueness results. Take, for example, an exchange economy. In both the WA and the GS cases, whenever z(p) = 0, the matrix Dz(p) is necessarily negative semidefinite [see the small-type discussion of expression (17.F.I) and Proposition 17.F.4]. Moreover, if an equilibrium is regular (i.e., if rank Dz(p) = L - I), the negative scmidefiniteness of Dz(p) can be shown to imply that the index of the equilibrium is necessarily + I (see Exercise 17.F.II). Hence, we can conclude that in both the WA and GS cases, any regular economy must have a unique (normalized) equilibrium price vector. Although the index methodology provides a good research tool. it is often the case that, as here, uniqueness conditions lend themselves to direct proofs. It is a notable fact that some of the more subtle uses of index analysis are not to establish uniqueness but rather to establish nonuniqueness [the first usage of this type was made by Varian (1977)]. This is illustrated in Example 17.F.3.
The revealed preference property of gross substitution.
(e,change) equilibrium and p' is not, then, sinoe :(p); 0, we have p',z(p); 0, and therefore any nonequilibrium p' is revealed preferred to p. Henoe, the requirement in (17.F.3) that ".:( p') > 0 amounts to a restricted version of the WA asserting that no equilibrium price veclor p C'In be revealed preferred to a nonequilibrium prioe vector p'. Geometrically, it says that the range of the excess demand function, {:(p'): p' »O} c RL (Le., the offer curve), lies entirely above the hyperplane through the origin with normal vector p (see Figure 17.F.3). In par;dlcl to Proposition 17.F.2, condition (17.F.3) implies the convexity of the equilibrium price set of Ihe exchange economy, that is, of (pe R'++: z(p); O} c R" (in Exercise 17.F.B, you are asked 10 show this). Interestingly, condition (17.F.3) is satisfied not only in the WA and the GS cases but also in the no-trade case, to be reviewed shortly. In Ihe differentiable case, there is a parallel way to explore the connection between the WA and gross substitution. Let z(p) = O. The sufficient differential condition (17.F.I) for the WA tells us that dp'D:(p)dp < 0 for any dp not proportional to p. Suppose now that instead of the WA, we require that Dz(p) has the gross substitute sign pattern. Because z(p) = 0, we have p' D:(p) = 0 and Dz(p)p = 0 [recall (17.E.I) and (17.E.2)]. Using these two properties it can then be shown that again we obtain dp' Dz(p) dp < 0 for any dp not proportional to p (sec Section M.D of the Mathematical Appendix). Henoe, we can conclude that aC all exc/ulIIgc "'{lIjlibrjllll/ prke veccor, the GS property yields every local restriction implied by the WA. This is summarized in Proposition 12.F.4.
Example 17,F_3: Suppose we have two one-consumer countries. i = 1,2. Countries are symmetrically positioned relative to the home (H) and the foreign (F) good. To be specific, let each country have one unit of the home good as an endowment and none of the foreign good, and utility functions u,(xu;, xF/) = Xu; - X:, for -I < p < O. Merely from symmetry considerations, it follows that there is a symmetric equilibrium p = (1.1). But we may be interested in knowing whether there are asymmetric equilibria. One way to proceed is as follows: compute the index of the symmetric equilibrium; a sufficicnt (but not necessary) condition for the existence of an asymmetric equilibrium is that this index be negative (i.e., _1).49 If we carry out the computation for the present example (you are asked to do so in Exercise 17.F.I3), we see that the index is negative if at prices p = (I, I) the wealth effects in each country are so biased toward the home good that an increase in the price of the good of country I. say, actually increases the demand for this good in country I by more than it decreases the demand from country 2. •
Proposition 17.F.4: If z(') is an aggregate excess demand function, zIp) = 0, and Oz(p) has the gross substitute sign pattern, then we also have dp-Oz(p) dp < 0 whenever dp ,",0 is not proportional to p.
Uniijlleness as an Implication of Pareto Optimality We now present a result that is not of great significance in itself but that is nonetheless interesting because it highlights a uniqueness implication of Pareto optimality. For simplicity, we restrict ourselves again to an exchange economy (see Exercise 17.F.9 for a generalization allowing for production). Proposition 17.F.S: Suppose that the initial endowment allocation (w" .. . , wc) constitutes a Walrasian equilibrium allocation for an exchange economy with strictly convex and strongly monotone consumer preferences (Le., no-trade is an equilibrium). Then this is the unique equilibrium allocation.
49. In this, as typically in any example, the excess demand function fails to be differentiable at
prices at which demand just "hits" the boundary. Typically (we could say "generically"), these prioes will not be equilibrium prioes and the validity of the index theorem is not alTected by these
Proof: Let an allocation x = (XI' ... , xc) and price vector p constitute a Walrasian cquilibrium when consumers' endowments are (w" . .. , w,). Since w, is affordable at
nondilTerentiabilities.
L
EQUILIBRIA
615
616
CHAPTER
'7:
THE
POSITIVE
THEORY
OF
~ [ CliO N
EOUILIBRIUM
17.G Comparative Statics Analysis
Proof: Let Ihe firsl consumer have endowmenls wilh Ihe prescribed amounlS of Ihe firsl L - I commodilies, and give 10 Ihis consumer arbitrary preferences, with the single reslriclion Ihal 0"" t ,(p; ,;,,) be nonsingular (it suffices for Ihis thallhe demand function of ~onsumer I salis?es a slricl normalilY condilion; again see Exercise 17.G.I). Since 0 ..,1(;;; WI) = 0~,1,(;;; w,), expression (17.G.I) lells us Ihal we are looking for an addilional collection of L consumers such Ihal Ihe resulling (L + I)-consumer economy has z(p; ';',) = 0 and
z(p; q) = (ZI(P; q), ... , Z,._,(P; q)).
Here, q E IRN is a vector of N parameters influencing preferences or endowments (or both). Throughout, we normalize P,. = I. Suppose the value of the parameters is given initially by the vector ii and that p is an equilibrium price vector for ii; that is, z(p; ii) = O. We wish to analyze the effect of a shock in the exogenous parameters q on the endogenous variable p solving the system. A first difficulty for doing so is the possibility of multiplicity of equilibrium: the system of L - I equations in L - I unknowns i('; q) = 0 may have more than one solution for the relevant values of q, and thus we may need to decide which equilibrium to single out after a shock. If the change in the values of the parameters from ii is small, then a familiar approach to this problem is available. It consists of focusing on the local effects on p, that is, on the solutions that remain near p. Assuming the differentiability of i(p; q), we may determine those eITects by applying the implicit function theorem (see Section M.E of the Mathematical Appendix). Indeed, if the system z(·; ii) = 0 is regular at the solution p, that is, if the (L - I) x (L - I) matrix Dpz(p; ij) has rank L - l,so then for a neighborhood of (p; ij) we can express the equilibrium price vector as a function p(q) = (p,(q), ... , PL-,(q» whose (L - I) x N derivative matrix at ii is
ii)] - , D.i( p; ii).
(I7.G.2) NOlc Ihal Ihe (L - I) x (L - I) matrix defined in (17.G,2) is nonsingular. Thus, we have reduced our problem 10 Ihe following: can we find L consumers whose aggregale excess demand al p is -:,(;;; ,;',) and whose aggregate (L - I) x (L - I) matrix of price effects is A = - D,o,: ,(P;';', )8-' - O,Z,(;;; ';',)1 II follows from Proposition 17.E.2 Ihal the answer 10 Ihis queslion is "yes" (nole Ihallhe reslriclions Ihal Proposilion 17.E.2 imposes on Ihe L x L matrix A place no restriclion on Ihe malrix obtained by deleling one row and one column of A) . • Proposilion 17.G.1 shows that any firsl-order effecl is possible. As in Section 17.E (recall Figurc 17.E.3), il is also the case here thaI if there are prior reslriclions on initial endowments and if consumplion musl be nonnegalive, Ihen Ihere are again comparative statics restrictions of a global character. [See Brown and Matzkin (1993) for a recent investigation of this point]
(17.G.I)
There are a number of comparative static effects that, ideally. we would like to have and that seem economically intuitive: For example, that if the endowment of one good increases, then its equilibrium price decreases. Nevertheless, strong conditions are required for them to hold. By now this should not surprise us: We already know that wealth effects and/or the lack of sufficient substitutability can undermine intuitive comparative static effects. The latest instance we have seen of this occurring has been precisely Proposition 17.G.1. The analysis of uniqueness in Section 17.E may lead us to suspect that good comparative statics effects can hold if aggregate excess demand satisfies either weak-axiom-like conditions (recall Definition 17.F.I) or gross substitution properties (sec Definition 17.F.2). This is in fact so. We consider first the implications of a weak-axiom-like restriction on aggregate excess demand.
What can we say about the first-order eITects Dp(ij)? Expression (17.G.I) and Proposition 17.E.2 [which told us that the matrix of price effects Dpz(p; ij) is unrestricted when I 2: L] strongly suggest that, without further assumptions, the "anything goes" principle applies to the comparative statics of equilibrium in the same manner that in Section 17.E it applied to the closely related issue of the effects of price changes on excess demand. We now elaborate on this point in the context of a specific example. = (w,', ... , WL_I.') Let the list of parameters under consideration be the vector of initial endowments of the first consumer for the first L - I commodities. All of the remaining endowments are kept fixed. As before we assume that i('; dJ,) = 0 is regular at the solution p. It can be shown (see Exercise 17.G.I) that if the demand function of the first consumer satisfies a strict normality condition, then rank Dp(6J,) = L - I, where p(.) is the locally defined solution function with p(J,,) = p. Proposition 17.G.I tells us that if there are enough consumers then this is all that we can say.
w,
50. In a slighl abuse of nOlalion. we leI D,z(i;; ii) stand for Ihe matrix oblained from D,z(p; Ii) by deleling Ihe lasl row and column.
COM PAR A T I v EST A TIC S
Proposition 17.G.1: Given any price vector p, endowments for the first consumer of the first L - 1 commodities J" = (w", ... , wL -",), and a (L - I) x (L - 1) nonsingular matrix 8, there is an exchange economy formed by L + 1 consumers in which the first consumer has the prescribed endowments of the first L - 1 commodities, i(ft; J,,) = 0, i(', J,,) = 0 is regular at p and Dp(J,,) = 8.
Comparative statics is the analytical methodology that concerns itself with the study of how the equilibria ofa system are affected by changes (often described as "shocks") in various environmental parameters. In this section, we examine the comparative static properties of Walrasian equilibria. To be concrete, we consider an exchange economy formalized by a system of aggregate excess demand equations for the first L - I commodities:
Dp@ = - [Dpi( p;
11. 0:
Proposition 17.G.2: Suppose that i(p; ij) = 0, where i(') is differentiable. If Dqi(p; ij) is negative definite," then (Dqi(p; ij) dq)'(Dp(ij) dq) 2: 0 for any dq,
i
1
(17.G.3)
51. This condition is independent or which particular commodity has been labeled as L (see
Section M.D of Ihe Malhemalical Appendix).
.
A HAL Y SIS
617
618
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EOUILIBRIUM
SECTION
COMPARATIVE
STATICS
first L - I goods (relative to the price of the Lth good) decrease. 53 In particular. suppose again that consumer I's initial endowment of some good decreases. By labelling commodities appropriately. we can let this good be commodity LUnder the assumption of normal demand for consumer I. a decrease in WLl. at the fixed price vector P. will decrease the excess demand for the first L - I goods. Therefore. the prices of the first L - I goods decrease and so we now reach the conclusion that we could not obtain by means of Proposition 17.G.2: if the endowments of a single good decrease then its price (relative to the price of any other good) increases. This suggests. incidentally. that the assumptions of Proposition 17.G.3 are strictly stronger than those of Proposition 17.G.2. Indeed. as we saw in Proposition 17.F.4. if z(p; ij) = 0 and the L x L matrix D,z(p; ij) satisfies the gross substitute property. then dp'D,z(p; ij) dp < 0 whenever dp '" 0 is not proportional to p. In particular. by letting drc = 0 we have that the matrix D,i(p; q) is negative definite.
Proof: The inverse of a negative definite matrix is negative definite. Therefore [D,i(p; q)] -, is negative definite (see Section M.D of the Mathematical Appendix). Hence. by (17.G.I) we have (D,z(p; ij) dq) . (Dp(ij) dq)
17.G:
= - D.i(p; ij) dq· [D,z(p; ijl] -, D.i(p; q) dq
which is precisely (17.GJ). _ The weak axiom implies the negative semidefiniteness of D,i(P; q) whenever 0 [see expression (17.F.I) and the remark following it]. Therefore. the assumption of Proposition 17.G.2 amounts to a small strengthening of this implication. Its conclusion says that for any infinitesimal shock dq in q. the induced shock to excess demand at prices fixed at P. D.i(p; q) dq. and the induced shock in equilibrium prices. D.p(ij) dq. move "in the same direction" (more precisely. as vectors in R / -' they form an acute angle). For example. a shock that at fixed prices alTects only the aggregate excess demand of the first good. 52 say by decreasing it. will necessarily decrease the equilibrium price of this good. Note that this docs 110/ say that if (I)" increases then the equilibrium priee of good I decreases. Under an assumption of normal demand. this change in w" does indeed decrease the excess demand for good I at ;; but it also alTects the excess demand for all other goods (sec Exercise 17.G.2). We next consider in Proposition 17.G.3 the implications of gross substitution (or. more precisely. of gross substitution holding locally at (p; q)).
:( p; ti) =
Expression (17.G.I) allows us to explicitly compute the effects of an infinitesimal shock. In fact. it also offers a practical computational method to estimate the local effects of small (but pcrh"ps not infinitesimal) shocks. Suppose that the value of the vector of parameters after the shock is ci and, for IE [0, I]. consider a continuous function i(' ,I) that, as I ranges from I ; 0 to I = I. distorts :(.; ti) into i('; iiI. An example of such a function, called a homotopy. is :('./) = (I - I):(';q)
+ If(·;ii).
Denote the solution set by E = {(I, pI: f(p, t) = O}. Then we may attempt to determine p(q) by following a segment in the solution set that starts at (0, p)." If ij is close to q. and the initial situation (1 is regular, then we are in the simple case of Figure 17.G.I(a): there is a unique segment that connects (0. p) to some (I. p)." Naturally. we then put P(ii) - p. If q is not close to ij but nevertheless i(' ,I) is a regular excess demand function for every I [this will be the case if. for example, z(· ./) satisfies. for every I. any of the uniqueness conditions covered in Section 17.F], then this procedure will still succeed in going from I = 0 to I = I and. therefore, in determining an equilibrium for ii.'· Unfortunately. if the shock is large, we can easily find ourselves in situations such as Figures 17.G.I(b) and 17.G.I(c), where at some t' the economy i(·. n is not regular and at (I'. Po') there is no natural
Proposition 17.G.3: Suppose that i(;;; ij) = 0, where i(·; .) is differentiable. If the L x L matrix Dpz(;;; ij) has negative diagonal entries and positive off-diagonal entries, then [Dpi(;;; ij»)" 1 has all its entries negative. Proof: Because of the homogeneity of degree zero of excess demand (recall Exercise 17.E.I). we have D,z(p;q)p=O. and so D,z(p;q)p«O. where p=(p, ..... p.-,). Denote by I the (L - I) x (L - I) identity matrix and take an r > 0 large enough for the matrix A = (l/r)D,z(p; q) + I to have all its entries positive. Then D,i(p; q) = -r[l- A]. and therefore D,l(p; q)p« 0 yields (I - A)p» 0; that is. the positive matrix A. viewed formally as an input-output matrix. is productive (see Appendix A of Chapter 5; the fact that the diagonal entries of A are not zero is inessential). Hence, as we showed in the proof of Proposition 5.AA.I. the matrix [I - A)"' exists and has all its entries positive. From [D,i(p; q)J"' = -(I/r)[1 - A)"' we have our conclusion. _
53. This conclusion holds for nonlocal shocks as well. To see this let Dz(p; q) have the gross subs,itu'e sign pattern throughout its domain and suppose that l(p; ij)« l(p: ij) for all p. For IE [0.1]. define zIp; I) = Ii(p; ii) + (I - I)i(p; ij). Denote by P(/) the solution to l(p; I) = O. Note 'hat D.i( P(/); I) dl = 1(P(/); ii) - 2(P(/); ij)« 0 for all I and therefore. by Proposition 17.G.3. Dp(/) dl « 0 for all I. But then. for any ( = I, .... L - I. we have
PI (q-) - PI (-) q = f.'[ilPI(t)]d dI I < 0 . 0
[t follows from Proposition 17.GJ and expression (17.G.I) that, given gross subSlitution. if D,z(p; q) dq «0. that is. if the excess demand for all of the first L - I goods decreases as a consequence of the shock (and therefore the excess demand for the Llh good increases), then Dp(q) dq «0. That is. the equilibrium prices of the
In Exercise 17.G.3 you can find a more direct approach to the global theory. See also Milgrom and Shannon (1994) for much more on the latter approach. 54. In practice. "following" a segment involves the application of appropriate numerical techniques; see Garcia-Zangwill (1981), Kehoe (1991). and references therein. 55. Moreover, if the shock is sufficiently small. the p so obtained is independent of the particular homotopy used. 56. However. if there are multiple equilibria at ii. then which equilibrium we find may now depend on the homotopy.
52. What this means is 'hat the excess demand of good 2 to L - I is not changed. By Walras' law. the excess demand of good L must change.
L
ANALYSIS
619
620
CHAPTER
17:
THE
POSITIVE
THEORY
o o (b)
(a)
Figure 17.G.l
OF
EOUILIBRIUM
however, that those are just two examples. Indeed, one of the difficulties in this area is the plethora of plausible disequilibrium models. Although there is a single way to be in equilibrium, there are many different ways to be in disequilibrium.
Price Ttitollllemenl
,.
We consider an exchange economy formalized by means of an excess demand function
o
z(·). Suppose that we have an initial p that is not an equilibrium price vector, so
(e)
that z(p) '" O. For example, the economy may have undergone a shock and p may be the pres hock equilibrium price vector. Then the demand-and-supply principle suggests that prices will adjust upward for goods in excess demand and downward for those in excess supply. This is what was proposed by Walras; in a ditTerential equation version put forward by Samuelson (1947), it takes the specific form
Comparative statics in the large: the general case.
conlinuation of the palh as I increases." To obtain an equilibrium i> for ij there is then no real alternative but to appeal to general algorithms for the solution of the system of equations :(.: ij) = O. It is a sobering thought that which solution we come up with at ij may be dictated more by our numerical technology than by our initial position (p; ij). This is most unsatisfactory, 'lnd it is a manifestation of a serious shortcoming-the lack of a theory of equilibrium
dp,
- = c,z,(p) de
for every (,
(17.H.I)
where dp,/til is the rate of change of the price for the fth good and c, > 0 is a constant atTecting the speed of adjustment. Simple as (17.H.I) is, its interpretation is fraught with difficulties. Which economic agent is in charge of prices'! For that malter, why must the "law of one price" hold out of equilibrium (i.e., why must identical goods have identical prices out of equilibrium)"! What sort of time does "I" represent? It cannot possibly be realtime hecause, as the model stands, a disequilibrium p is not compatible with feasibility (i.e., not all consumption plans can be simultaneously realized). Perhaps the most sensible answer to all these questions is that (17.H.I) is best thought of not as modeling the actual evolution of a demand-and-supply driven economy, but rather as a tentative trial-and-error process taking place in fictional time and run by an abstract market agent bent on finding the equilibrium level of prices (or, more modestly, bent on restoring equilibrium after a disturbance).'· The hope is that, in spite of its idealized nature, the analysis of (l7.H.I) will provide further insights into the properties of equilibria. Even perhaps some help in distinguishing good from poorly behaved equilibria. The analysis is at its most suggestive in the two-commodity case. For this case, Figure 17.H.l represents the excess demand of the first good as a function of the relative price pdp,. The actual dynamic trajectory of relative prices depends both on the initial levels of absolute prices and on the differential price changes prescribed by (17.H.I).60 But note that, whatever the initial levels of absolute prices, p,(I)/p,(t) increases at e if and only if z,(p,(t)!p,(I), I) > O. In Figure I7.H.I we see the following two features of the adjustment equations (l7.H.I).
selection.
17.H Tatonnement Stability We have, so far, carried out an extensive analysis of equilibrium equations. A characteristic feature that distinguishes economics from other scientific fields is that, for us, the equations of equilibrium constitute the center of our discipline. Other sciences, such as physics or even ecology, put comparatively more emphasis on the determination of dynamic laws of change. In contrast, up to now, we have hardly mentioned dynamics. The reason, informally speaking, is that economists are good (or so we hope) at recognizing a state of equilibrium but are poor at predicting precisely how an economy in disequilibrium will evolve. Certainly there are intuitive dynamic principles: if demand is larger than supply then the price will increase, if price is larger than marginal cost then production will expand, if industry profits are positive and there are no barriers to entry, then new firms will enter, and so on. The difficulty is in translating these informal principles into precise dynamic laws. 58 The most famous attempt at this translation was made by Walras (1874), and the modern version of his ideas have come to be known as the theory of ealonnemenl slabililY. In this section, we review two tatonnement-style models, one of pure price adjustment and the other of pure quantity adjustment. We should emphasize,
(a) Call an equilibrium (p" p,) locally slable if, whenever the initial price vector is sufficiently close to it, the dynamic trajectory causes relative prices to converge to the equilibrium relative prices p,/p, (the equilibrium is locally tolally unstable if any
57. Note that by reversing the direction of change of I we can continue to move along the segments in these two figures (this is actually quite a general facl). If p is the only solution at I = 0, as in 17.G.I(b), then the segment necessarily ends with a (I, p). Thus, in some sense we have succeeded in finding an equilibrium for ii that is associated with our initial p. But the association is very weak: it may depend on the particular homotopy and it requires the parameter-reversal procedure. If, as in Figure I7.G.I(c~ ;Hs not the only equilibrium at I = 0, then the procedure may simply not work: the segment that starts at (0, p) goes back to I = O. 58. Refer to Hahn (1982) for a general review.
59. This is, in essence, the idea of Walras (li'onnement means "groping" in French), who took inspiration from the functioning of the auctioneer·directed markets of the Paris stock exchange.
The idea was made completely explicit by Barone (t908) and by Lange (1938), who went so far as to propose the tatonnement procedure as an actual computing device for a centrally planned economy.
L
622
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
SECTION
=,
17.H:
TATONNEMENT
STABILITY
623
L=3 I: Locally Stable. Index +1 2: Locally TOlally Unslable. Index +1 3: Saddle. Index -I p,/p,
Flgur.
=, (P,/Pl. I)
17.H.l
Tatonnement
Irajeclories for L = 2.
disturbance leads the relative prices to diverge from pdp,). Then a (regular) equilibrium PI/P, is locally stable or locally totally unstable according to tlae sign of tlae slope of excess demand at Ihe equilibrium, that is, according to the index of the equilibrium (recall Definition 17.0.2). If excess demand slopes downward at p,/p, (as in Figure 17.H.l), then a slight displacement of PI/P, above pdp, will generate excess supply for good I (and excess demand for good 2), and therefore the relative price will move back toward the equilibrium level p,/p,. The effect is the reverse if excess demand slopes upward at pdp,. (b) There is sySlem stability, that is, for any initial position (PI(O), p,(O)), tlae
Flgur. 17.H.2 An example of tatonnement
trajeclories for L = 3.
price of a good goes to zero the excess demand for the good becomes positive (thus, in particular, the trajectories point inward near the boundary). However, properties (a) and (b) are both violated: There are (regular) equilibria that are neither locally stable nor locally totally unstable (they are "saddle points," such as the equilibrium labeled 3 in the figure), and from some initial positions prices may not converge to any equilibrium."' In a more positive spirit, we now argue that for the cases where we have succeeded in proving the uniqueness of Walrasian equilibrium, we are also able to establish the convergence of any price trajectory to this equilibrium (this property is called global stability).62 The next proposition covers, in particular, the weak axiom, the gross substitute, and the no-trade cases studied in Section 17.F.63 These three cases have in common that they satisfy the weak axiom when we restrict ourselves to comparisons between equilibrium and nonequilibrium prices [see the discussion of condition (17.F.3) in Section 17.F]. That is, for the unique (normalized) equilibrium price vector p* arising in these cases we have: "If z(p*) = 0 then pO. z(p) > 0 for any p not proportional to p*."
corresponding trajectory of relalive prices PI(t)/P,(t) converges to some equilibrium arbitrarily closely as t -+ ct:). For regular, two-commodity, economies, properties (a) and (b) give a complete picture of the dynamics. It is very satisfactory picture that accounts for the persistency of tatonnement stability analysis: a theory yielding properties (a) and (b) must be saying something with economic content. Unfortunately, as soon as L> 2 neither the local conclusions (a) nor the global conclusions (b) of the two-commodity case generalize. This should not surprise us, since the price dynamics in (17.H.l) are entirely driven by the excess demand function, and we know (Propositions 17.E.2 and 17.E.3) that the latter is not restricted in any way (beyond the boundary conditions). Consider an example for L = 3 and c, = c, = c, = I. In Figure I7.H.2 we represent the normalized set of prices S = {p » 0: (p,)' + (p,)' + (P3)' = I}. This normalization has the virtue that, for any excess demand function z(p), the dynamic flow p(/) generated by the differential equation dptfdt = z(p), ( = 1,2,3, remains in S [i.e., if P(O) E S then p(t) E S for all t]. This is a consequence of Walras' law:
Proposition 17.H.l: Suppose that z(p*) = 0 and p*,z(p) > 0 for every p not proportional to p*. Then the relative prices of any solution trajectory of the differential equation (17.H.l) converge to the relative prices of pO.
d(p,(t)' + p,(t)' + pit)') dt = 2p,(t)z,(p(t» + 2p,(t)z,(p(t)) + 2p,(t)Z3(P(t» = O.
Proof: Consider the (Euclidean) distance function f(p) = Lt (l/c{)(Pt - pi)'. For any trajectory p(t) let us then focus on the distance f(p(t)} at points I along the trajectory. We have
Thus, the dynamics of p can be represented by trajectories in S, the direction vector of the trajectory at any P(/) being the direction of the excess demand vector z( p(t». We conclude, therefore, that the only restrictions on the trajectories imposed by the general theory are those derived from the boundary behavior of excess demand. In Figure 17.H.2 we represent a possible field of trajectories. In the figure, when the
61. We should warn againsl deriving any comfort when prices converge 10 a limil cycle. Recall that this price tatonnement is not happening in real time. The dynamic analysis has a hope of telling us something significant only ir it converges. 62. Warning: uniqueness by itself does nol imply slabilily-excepl ror L = 2. You should Iry
60. NOle Ihal allhough Ihe change in Pf all prescribed by (I7.H.I) depends only on Ihe relative prices p,/p, for { = I, 2, Ihe change in Ihe price ralio p,/p, al I depends bolh on Ihe curren I price
10
ratio and on the current absolute levels of PI and Pl"
1
draw a counlerexample in Ihe slyle of Figure 17.H.2. 63. For a proof specific 10 Ihe gross Subslilule case. see Exercise 17.H.1.
624
C HAP T E R
1 7:
THE
P 0 5 I T I VET H E 0 R Y
0 F
E 0 U IL
I8
~ E. C T I
R I lJ ;.,
<J
h
• 1 • H:
TAT 0 NNE MEN T
S T It. B I l i T Y
6"::~
succeeds in restoring equilibrium afler a small dislurbance. Thus we see the contrast: for lalonnemenl stability. we impose few informalional reslriclions on the adjustment process [to determine Ihe change in p we only need 10 know f(p); in particular. no knowledge of Ihe derivalives of 1(') is required]. bUI convergence is guaranleed only in special circumslances. For Ihe NeWlon method. local convergence always oblains. bul to delermine Ihe direclions of price change al any p we need 10 know alllhe excess demands f(p) and all the price effecls D:(p). See Smale (1976) and Saari and Simon (1978) for classic conlributions to Ihis Iype of
- p" z(p(r» !> 0,
Newton price dynamics.
where the last inequality is strict if and only if p(r) is not proportional to p'. We conclude that the price vector p(r) monotonically approaches the price vector p' [in fact, since the same argument applies to (1.p', p(r) must be monotonically approaching any (1.p']. This does not mean that p(r) reaches a vicinity of p'. Typically it will not: the rate of approach of p(r) to p' will go to zero before p(r) gets near p'. But the rate of approach can go to zero only if p(r) becomes nearly proportional to p' as r ..... 00, in which case the relative prices do converge.· 4 •
Qlllllltity Tatollllemelll In Ihe analysis so far. prices could be out of equilibrium but quantities. that is to say Ihe amounls demanded and supplied. are always at their equilibrium (i.e.• utility and profit·maximizing) values. We now briefly consider a model in which quantities rather Ihan prices may be in disequilibrium.· 7 This is besl done in a production context. To be very concrele. suppose that there is a single production set y.6. At any moment of lime. we assume that there is given a single. fixed production vector Y E Y. Prices. however. are always in equilibrium in the sense that the general equilibrium syslem of the economy. conditional on y. generales some equilibrium price system p<.r) (that is to say. we proceed in the short run as if the short-run production sel were {y} - R';.). This describes the short·run equilibrium of the economy. What is an appropriale dynamics for this economy? It makes sense to think Ihat. whatever it is, Ihe change in production at time I, dy(r)/dl E R L , moves production in a direction thaI increases profits when Ihe price vecror ar lime I, p(y(I». is laken
We can gain further insighl inlo Ihe dynamics of lalonnemenl by carrying oul ;. local analysis. It will be more convenienl now if we fix PL = I and. consequenlly. we limil (17.11.1) 10 Ihe firsl L - I coordinales. Accordingly we denole f(p) = (z,(p) •...• 'L-'(P)). Suppose Ihal t(p·) = O. A slandard result of differenlial equalion Iheory lells us Ihal if Ihe (L - I) x (L - I) malrix Dt(p') is nonsingular (i.e.• if Ihe equilibrium is regular). Ihen Ihe behavior of Ihe trajeclories in a neighborhood of p. is conlrolled by Ihe linearizalion of Ihe syslem al p•• Ihal is. by CDt(p·). where C is Ihe (L - I) x (L - I) diagonal malrix whose Ilh diagonal entry is Ihe conslanl Ct. One says Ihal p' is locally stable if Ihere is t > 0 such Ihal P(I) .... p. whenever II P(O) - p'lI < £ (i.e., for small perturbalions Ihe equilibrium will lend 10 reslore ilself). It Ihen lurns OUI thaI p' is locally slable if and only if all Ihe eigenvalues of eDi(p') have negalive real parts. In addition, p' is locally slable irrespeclive of the speeds of adjuslment"' (i.e .• for all positive diagonal malrices C) if Df(p') is negalive definile (see Seclion M.D of Ihe Malhemalical Appendix).·· One way 10 undersland why Ihe previous local slabilily resull for Ihe latonnemenl dynamics requires strong condilions on Df(p') is 10 nole Ihal we are in facl imposing Ihe condilion Ihal the price of a commodity reacls only to Ihe excess demand or supply for the same commodily. An ideal market agenl may wanl 10 adjusl Ihese prices wilh an eye also 10 Ihe effecls of Ihe adjustmenl on Ihe excess demand for Ihe olher commodilies. One concrele possibilily is Ihe following: if excess demand allime I is f(p(I)). Ihen Ihe markel agenl adjusts prices by some amounl dp/dr = (dp,/dl •...• dpL_,/dl) so as 10 cause a proportional decrease in the magnitude of all excess demands and supplies. ThaI is. Df(p(I))(dp/dl) = -;.f(p(t)) for some i. > 0, or. if the relevant inverse exists,
as {liven:
Definillon 17.H.1: We say that the differentiable trajectory y(t) E Y is admissible if p(y(t))·(dy(t)/dt) ~ 0 for every t. with equality only if y(t) is profit maximizing for p(y(t)) (in which case we could say that we are at a long·run equilibrium). A difference belween the price and the quantity tatonnement approaches that adds appeal to the second is Ihal feasibility is now insured at any I and that, as a result. we can interpret the dynamics as happening in real time. 69 •7o Will an admissible Irajectory necessarily take us to long. run equilibrium? We cannot really explore this matter here in any detail. As usual, the answer is "only
~·I:·
I
~.
~
Ii
'!!!. =
-;.[Di(pJr'i(p)
(17.H.2)
67. We could also look al Ihe general case where both could be in disequilibrium; sec. for example. Mas·Colell (1986). 68. There is no difficulty in considering several. Also. Y can be interpreted as an individual or
dt
This adjuslmenl equalion is known as Newlon's method and is a slandard lechnique of numerical analysis. If Df(p') is nonsingular. so Ihal [Dt(p·)r' exisls. then (l7.H.2) always
as an aggregate production set.
69. Nonelheless. it is importanl 10 realize thaI. even then. Ihis is not a fully dynamic model: The optimization problems of the consumers remain static and free of expectational feedbacks and firms follow naive. short· run rules of adjustment (in a more positive spirit one might call this adap,ive. rather than naive, behavior). For an extensive analysis of market adjustment procedures in real
64. Conlinuous rcal·valued functions Ihallake decreasing values along any dynamic trajectory and the value zero only at stationary points are known as Lyapunov functions.
time. see Fisher (1983). 70. The quantity dynamics of Definition 17.H.1 are reminiscenl of Marshall (1920) and arc
65. How could we prelend to know much about speeds of adjustments? 66. NOle Ihat Ihis fils nicely wilh Proposilion 17.H.1 because Ihe revealed.preference·like property poslulaled Ihere implies the negalive (.. mijdefiniteness of Dt(p) al Ihe equilibrium price vector p•.
ortcn referred to as Marshallian dynamics, especially in a partial equilibrium context. ]n contrast. the price dynamics are frequently called Walrasian dynamics.
l
.....
626
CHAPTER
17:
THE
POSITIVE
THEORY
Of
EQUILIBRIUM
SEC T ION
Proof: Consider u(y(l) + w) where u(·) and ware respectively the utility function and the endowments of the consumer. The unique equilibrium production vector is the single production vector y that. maximizes u(y + w) on Y; recall the oneconsumer, one-firm example of Section IS.C. The argument is much simpler if we assume that u(·) is differentiable. We claim that utility must then be increasing along any admissible trajectory. Indeed,
+ w)
V ( () uyr
+w
) dy(l) '-d-I-
dyer) = !I(I)p(y(r»'d/
NON CON V E X I TIE S
627
As we have mentioned repeatedly, especially in Chapters 10 and 12, a central justification of the price-taking hypothesis is the assumption that every economic agent constitutes an insignificant part of the whole economy. Literally speaking, however, this cannot be satisfied in the model of this chapter because, formally, we allow for no more than a finite number I of consumers. (This is particularly true of our examples, where we typically have I = 2.) A straightforward reinterpretation is possible, however. We illustrate it for the case of a pure exchange economy. Suppose we consider economies whose consumers have characteristics (preferences and endowments) that fall into I given types, with r consumers of each type (a generalization to unequal numbers per type is possible; see Exercise 17.1.1). That is, the set of consumers is formed by r replicas of a basic reference set of consumers. Furthermore, an allocation denoted by (Xl' ... ' x,) is understood now to specify that each consumer of type j consumes XI (so the totality of consumers of type j consume rx i ). We observe then that the analysis and results presented up to this point are not modified by this reinterpretation; they simply do not depend in any way on the parameter r. In this way, we can conclude informally that the theory so far covers cases with an arbitrarily large number, even an infinity, of consumers; in particular, we see that any equilibrium of our earlier model is an equilibrium of the r-replica economy (for any integer r ~ I). There is, however, an important qualification. The ability to interpret the model and results in a manner that is fully independent of the number of consumers depends crucially on the convexity assumption on preferences. Without this assumption, it is not justified to neglect allocations that assign different consumption bundles to different consumers of the same type. Consider, for example, the Edgeworth box of Figure 17.1.1. If there is only one consumer of each type, then no equilibrium exists; but if we have two of each type, then there is an equilibrium. To see this, give W2 to the convex consumers, let one of the two nonconvex consumers receive the bundle XI' and let the other receive the different bundle Thus, in the nonconvex case, the
Proposition 17.H.2: If there is a single strictly convex consumer, then any admissible trajectory converges to the (unique) equilibrium.
--d-I-- =
L A R GEE CON 0 M I E SAN D
17.I Large Economies and Nonconvexities
under special circumstances." A limited, but important example (it covers the shortrun/long-run model of Section IO.F) is described in Proposition 17.H.2.
du(y(l)
1 7 • I:
> 0,
with equality only at equilibrium. Here we have used the fact that at a short-run (interior) equilibrium, the price vector p(y(I» weighted by the marginal utility of wealth !I(t) must be equal to the vector of marginal utilities of the consumer. Now, since utility is increasing, we must necessarily reach the production vector y at which utility is maximized in the feasible production set (i.e., the equilibrium). This is illustrated in Figure 17.H.3. (We are sidestepping minor technicalities: to proceed completely rigorously, we should argue that the dynamics cannot be so sluggish that we never reach the equilibrium. To do so we would need, strictly speaking, to strengthen slightly the concept of an admissible trajectory). •
x;.
Note that the single consumer of Proposition 17.H.2 could be a (positive) representative consumer standing for a population of consumers. Figure 17.1.1
Good 2 1,(I)p(r(l))
= Vu(r(t) + w)
Figure 17.H.3
Equilibrium with
An example of
nonconvex preferences in economies of changing size.
quantity tatonnemcnt.
Indifference Curves of the (Representative) } Consumer
o,L---------\---t-----"-+-__
y" R~+ Good I i
1
THE
POSIT .... £
THEORY
OF
EQUILIBRIUM
SEC T ION
I r
I = r
(the sum has r terms)
+ ... + Z,,: =" E z,(p), ... , z"
A Ii
t:.
u
E C 0 HOM i E SA'" 0
,,0,.. l.. V
J"oj
~ 1. ), I TIE S
IJL:;;1
In the previous reasoning, the convexification or aggregate excess demand, with its existence implication, depends on our ability to prescribe very carefully which of several indifferent consumptions each consumer has to choose. Only in this way can we make sure that the "ggreg"tc consumption will be precisely right. Whatever we may think about the possible processes that may lead consumers to select among indifferent optimal choices in the right proportions, there can be little doubt that it would be better if we did not have to worry about this; that is, ir, given any price, practically every consumer had a single optimal choice. It is therefore or interest to point out that, while not a necessity, this is a most plausible occurrence if the number or consumers is large. Indeed, if the dist,ihution of individual prefere/lces ;s di.'ipas('d across the populCltion (so that, in particular. no two consumers arc exactly identical B), r!Jeli ('f('1f if til<' individual eX{'es."i demands are Irue (·orrespondences. the limit average rna}' well /><, 1I (continuous) lillletiOl'. This is because, at any p, only a vanishingly small proportion of consumers may display a nonconvexity at p. We commented on this point in Appendix A to Chapter 4 and wc illustrate it further in Figure 17.1.2. Consider the Edgeworth box or Figure
:,
:,
{Zil
L
then, as r grows large, the economy must possess an allocation and price vector that constitute a "near" equilibrium."
behavior of the economy can depend on the number of replicas: when we replicate an economy, new equilibria may emerge. 71 The discussion of the previous example suggests an interesting observation: replication may actually help in the analysis of economies with nonconvexities, in the sense that an increase in the size of the economy (in terms of the number of replicas) may help insure existence of an equilibrium. Indeed, we devote the rest of this section to develop the argument that, if the economy is large enough, then the existence of an equilibrium is assured, or nearly so, even if preferences are not convex." To see this, suppose that we have an exchange economy with I types. Consider a consumer of type i. If preferences are not convex (perhaps goods are indivisible), then the excess demand of this type is a correspondence z,(p). For p »0, z,(p) is a compact set that may not be convex (as is the case for consumer I at p = p in Figure 17.1.1). Measuring in average, per-replica, terms, the excess demand correspondence of type i when there are r replicas is
z,,(p) = - (z,(p) + ... + z,(p))
1 1 . '.
E z,(p)}.
~ , ,, t
I
p,
If we examine Figure 17.1.1 again, we see that, as r -+ 00, the set Z,,(p) + {w,} fills the entire segment between the demand points x, and x;. In particular, for any tx E [0,1] and integer r we can find an integer a, E [0, r] such that la./r - txl !> I/r (note that the r numbers {I/r, . .. , rlr} are evenly spaced in the interval [0,1]). By putting a, consumers at x, and r - a, consumers at x;, we get a per-replica consumption of
,
!
, I
p,
p,
Average ~~::::;;:::;-- Excess Demand Individual ; Function Excess Demand Correspondences
la)
(b)
Ftgur. 17.1.2 Thc aggregate demand rrom dispersed individual demand is a continuous function. Ca) Aggrcgate exccss demand for the Edgeworth box or Figure 17.1.1 with a continuum of consumers or each type (i.e., ,= 00). (b) Individual excess demands arc dispersed. 17.1.1. hut with a continuum of consumers or each type. Sincc in this economy all the consumers of type I are identical, all of thcm exhibit a "consumption switch" (a nonconvexity) at preciscly the same p. Thc excess demand correspondence ror the first good in this economy is represented in Figure 17.1.2(a) as , .. .,(')." But ir tastes of type 1 consumers exhibit variation, even if slight, then we would expect that no significant group or consumers simultaneously switches at any p and therefore that, as in Figure 17.1.2(b), average demand will be well defined at any p and will change only gradually with p.
which, by taking r large enough, comes as close to <xx, + (I - <x)x; as we wish. It turns out that this convexifying property is completely general. For any p » 0 and whatever the number of commodities, as r -+ 00 the per-replica excess demand z,,(p) of type i converges as a set of the convex hull of z,(p). In the limit, the average perreplica excess demand correspondence of type i becomes z'w(p) = convex hull z,(p). In the limit, therefore, the excess demand correspondence is convex valued and the existence of an equilibrium can be established as in Section 17.C. 73 In this sense,
74. Roughly speaking, by a "near" equilibrium we mean an allocation and price vector that is close to satisfying the conditions or an equilibrium. A precise technical definition of this concept can he given. but we shall not do so here. 75. We note, as an incidental matter, that in the limit with an infinity of agents this requirement is incompatible with the existence of only a finite number or types. To deal with this case, the rormal setting would need to be extended. 76. Prccisely. z .. , (.) is the correspondence whose graph is the limit graph, as r goes to <X!. of the correspondences :,(.) defined by ,,(pI = (lj')(z(p) + ... + zip)). where the sum has , terms and =(.) is the excess demand correspondence of the two-consumer economy in the Edgeworth box or figure 17.1.1.
71. That is, ir we let E(,) denote the equilibrium price set or the, replica economy, we have E(I) c E(,), but the converse need not be true. Note, moreover,that for arbitrary," > " > I,there need not be any inclusion relationship between E(,") and E(,') (except ir ," = m,' ror some integer m > I, in which case E(,') c E(,")). 72. See Starr (t969) ror a classic contribution to this topic. 73. See the comment after the proof of Proposition t 7.C.2 regarding demand correspondences, and also Exercise t 7.C.1.
l
630
CHAPTER
11:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
APPENDIX
We comment briefly on economies with production. Suppose that the consumption side of the economy is generated, as before, as the r-replica of a basic reference set of (possibly nonconvex) consumers. There are also J production sets l). Each lJ is closed, contains the origin, and satisfies free disposal (these are all standard assumptions). In addition, we assume that there is an upper bound (a capacity bound perhaps) on every lJ; that is, there is a number s such that )'/J ~ s for all ( and 11 e lJ. The production sets may be nonconvex. It is then possible to argue that the economy will possess a near equilibrium if r is large relative to the bound s (i.e., if the size of the consumption side of the economy is large relative to the maximal size of a single firm). On the average, the production side of the economy is also being convexified, so to speak (see the small-type discussion of Section S.E for a related point)." Note that the bounded ness property of the production sets is important. Suppose, for example, that every firm has the technology represented in Figure IS.C.3. Then no matter how many consumers there are, the potential profits of every firm are infinite (as long as p, > 0). Thus, there is no reasonable sense in which a near equilibrium exists. For the averaging·out effect to work the nonconvexity in production has to be of bounded size (see Exercise 17.1.2).
A:
CHARACTERIZING
EQUILIBRIUM
THROUGH
WELFARE
EQUATIONS
",
o,~--------------~~------~~~--
I
i
I
P(,)'(w, - "(s)) = g,(s) p,(s) . p,(s)
proportional to SEA (see Exercise I7.AA.I). In words, SEA stands for the values of utility distribution parameters and the determined allocation distributes "welfare" in accordance with the "shares" s = (s" ... ,sIlo Figure I7,AA.1 illustrates the construction. An arbitrary SEA will typically not correspond to an equilibrium. How can we recognize those sEA that do? To answer this question we can resort to the second welfare theorem. From Propositions 16.D.1 (and the discussion in small type following Proposition 16.D.3), we know that, under our assumptions, associated with x(,,) there is a price vector pes) E RL that supports the allocation in the sense that, for every i, x; >-; x,(s) implies pes)' x; > p(s) , x;(s). Therefore, (x(s'), p(s'» constitutes a Walrasian equilibrium if and only if S· E A solves the system of equations
APPENDIX A: CHARACTERIZING EQUILIBRIUM THROUGH WElFARE EQUATIONS
We have seen, beginning in Section 17.B, that if our economy satisfies sufficiently nicc properties (e.g., strict convexity of preferences) then we can resort, for the purposes of the analysis, to formalizing our theory by means of highly reduced systems of equilibrium equations. In the text of this chapter, we have focused on excess demand equations, But this is not the only possibility. In this appendix, we briefly illustrate a second approach that builds on the welfare properties of equilibria. We again concentrate on a pure exchange economy in which each consumer i = I, ... ,J has the consumption set R~ and continuous, strongly monotone, and strictly convex preferences. We also assume that WI ~ 0 for all i and 1:, w, » o. We know from Chapter 16 that a Walrasian equilibrium of this economy is a Pareto optimum (Proposition 16.C.l). Therefore, to identify an equilibrium, we can as well restrict ourselves to Pareto optimal allocations. To this effect, suppose we fix continuous utility functions u,(·) for the J consumers with u,(O) = O. Then to every vector s = (s I' ... , Sl) in the simplex A = {s' E R~ : 1:, s; = I} we can associate a unique Pareto optimal allocation xes) E R~I such that (UI(XI(S», ... , UI(XI(S))) is
g;(S')
= p(s')·[w; -
x,(s')]
=0
for every i = I, ... , J.
(I7.AA.I)
The Edgeworth box example of Figure 17.AA.2 explains the point that we are currently making. This Pareto-based equation system was first put forward by Negishi (1960), and was the approach taken by Arrow and Hahn (1971) in their proof of existence of equilibrium. It can be quite useful when the number of consumers (say, the number of countries in an international trade model) is small relative to the number of commodities. In contrast, if the number of consumers is large relative to the number of commodities, then an approach via excess demand functions will be superior. A limitation of the Negishi approach is that it is very dependent on the fact that an equilibrium must be a Pareto optimum. The excess demand approach is more easily adaptable to situations where this is not so (for example, because of tax distortions; see Exercise 17.C.3).7.
77. Observe that the average is with respect to , (the size of the economy in terms of the number of consumers). not with respect to J. Ir. as , increases. J is made to vary and is kept in some approximate fix.ed proportion with r, then rrom the qualitative point or view it does not matter how we measure size (this is a possible way to interpret, in the current context, the discussion or Section 5.E). But for the validity of the convexifying etrecl there is no need 10 vary J with r. The number J may be kept fixed and, thus, J could well be small relative to r (in which case the "averaged" economy is praclically one of pure exchange) or it could be large; il could even be Ihat J = co. The last case corresponds to a model with free entry, where the equilibrium-or the near equilibrium-
determines. endogenously, Ihe set of active firms. Typically, with free entry the sel of the aClive firms increases as the number of consumers, measured by r, grows (this point has also been discussed
78. The syslems of equations (17.B.2) and (17.AA.I) can be formally conlrasted as follows. In both of Ihem. at any poinl of Ihe domain of the equal ions, consumers and firms salisfy the utility maximizalion conditions for some prices and distribution of weallh. In (l7.B.2) this dislribution of weallh is always Ihe one induced by the inilial endowments, bUI feasibility (i.e., the equalilY of demand and supply) is insured only at the solulion. In (l7.AA.I) it is the other way around: feasibililY is always satisfied, but Ihe agreement of the wealth distribution with that induced by Ihe
in Section IO.F in a partial equilibrium context; there is not much more to add here).
initial endowments is insured only at the solution.
1
(ten) Construction of the welfare-theoretic equation system: first step.
Figure 17.AA.l
Figur. 17.AA.2 (rig hi)
Construction of the welfare-theoretic equation syslem: second step.
631
632
A P PEN 0 I X
<-OUllIBR1LM
G ENE R A l A P PRO A C H
TO
f H t:
..... I a
:
f. ,.. C f.
0,.
'#I A l R A 5 I A N
E a U III B R I U M
bjJ
state conditions implying that a quasiequilibrium is automatically an equilibrium. We devote the next few paragraphs to elaborating on this point. We begin with Definition 17.BB.2
APPENDIX B: A GENERAL APPROACH TO THE EXISTENCE OF WALRASIAN EQUILIBRIUM
The purpose of this appendix is to offer a treatment of the existence question at the level of generality of the model of Chapter 16. The results presented correspond roughly to those of Arrow and Debreu (1954) and McKenzie (1959). As when dealing with the second welfare theorem in Section l6.D, and for exactly the same technical reasons, it is useful to concentrate on establishing the existence of a Walrasian quasiequilibrium. This is a weaker notion than Walrasian equilibrium in that consumers are required to maximize preferences only relative to consumptions that cost strictly less than the available amount of wealth.
Definition 17.BB.2: The Walrasian quasiequilibrium (x*, y', p) satisfies the cheaper consumption condition for consumer i if there is X;E X, such that P'x; < p'w; + LjO;jP·",*. We then have Proposition 17.BB.1. Proposition 17.BB.1: Suppose that consumption sets are convex and preferences are continuous. Then any consumer who at the Walrasian quasiequilibrium (x*, Y*, p) satisfies the cheaper consumption condition must be preference maximizing in his budget sel. Hence, if the cheaper consumption condition is satisfied for a" i, (x*. y*, p) is also a Walrasian equilibrium.
Definition 17.BB.1: An allocation (xT, ... , xi, yf . ... ,yj) and a price system P '" 0 constitute a Walrasian quasiequilibrium if (i) For every i, P'Yj:$; p'Yj* for a" YjE)j. (ii') For every i, p·x;:$; P'Wi + LjOijP'Yj*, and if Xi~iXi then P'x; ~ P'w;
B:
Proof: Suppose that i satisfies the cheaper consumption condition; that is, there is with p'X, < p'W, + Lj O'jP·)'j. If x~ fails to be preference maximizing, then there is x; such that x; ~,x~ and x; is in the budget set of consumer i. Denote x~ = (I - (1/II))x; + (I/II)X,. Then XI'EX, and P'x, < p'W, + L)O,)P·yt for all II, and xi ..... x; as II ..... 00. By continuity of preferences, for large enough II we will have x~~. xi. But then consumer i violates condition (ii') of the definition of quasi· eqllilihrillm. _ Xi E Xi
+ L O;jP'yr j
(iii) L;x; = L;w; + Lj Yj*. Definition l7.BB.l is identical to Definition 17.B.I of a Walrasian equilibrium except that the preference maximization condition (ii) of Definition 17.B.I has been replaced by the weaker condition (ii'). With local nonsatiation, condition (ii') is equivalent to the requirement that xt minimizes expenditure for the price vector p in the set {x, E X,: x,
Suppose that, for every j, 0 E Y;, and for every i, X, is convex and w, ~ i, for some .ii E X,. Suppose, in addition, that the weak condition "L' w, + LJ fi » L, x, for some U , .... , ,iJ) E Y, x ' . 'x lJ" is satisfied. Then the quasiequilibrium (x', y', p) satisfies the cheaper consumption condition for consumer i in, fore.ample, either of the following situations (Exercise 17.1lB.1): (a) p P·(L,·X,}. Hence, there is at least one consumer with wealth larger than P··x i • Hut this consumer must be maximizing preferences (by Proposition 17.BB.l), which, by the strong monotonicity property, can only occur if every price is positive (i.e., if no good is free). Although convenient, neither condition (a) nor (b) can be regarded as extremely weak. It would be unfortunate if the validity of the theory were restricted to them. But this is not so: much weaker conditions are available. In particular, McKenzie (1959) has developed a theory of il1decomposable economies that guarantees that at a quasiequilibrium the cheaper consumption condition is satisfied for every consumer (and therefore the quasiequitibrium is an equilibrium). The basic idea, informally described, is that an economy is indecomposable if, no matter how we partition the economy into two groups, each of the groups has something for which the other group is willing to exchange something of its own. (See Exercise 17.BB.2.)
yn
FIgure 17.BB.1
Discontinuity of preference maximization.
!:
634
CHAPTER
17:
THE
POSITIVE
THEORY
OF
APPENDIX
EQUILIBRIUM
We now turn to the existence of a Walrasian quasiequilibrium. The aim is to establish the general existence result in Proposition 17.BB.2.
B:
GENERAL
APPROACH
TO
THE
EXISTENCE
OF
WALRASIAN
EQUILIBRIUM
__--~--~~--------------------~o,
Proposition 17.BB.2: Suppose that for an economy with I > 0 consumers and J > 0 firms we have (i) For every i, (i.1) Xi c RL is closed and convex; (i.2) <;:i is a rational, continuous, locally nonsatiated, and convex preference relation defined on Xi; (i.3) Wi ~ ~i for some ~i E Xi' (ii) Every lj c RL is closed, convex, includes the origin, and satisfies the free-disposal property. (iii) The set of feasible allocations
A = {(x, y)
E RLI
x RLJ: Xi E Xi for all i, Y,' E lj for all
i,
Ftgu,. 17.88.2 Equilibrium does not exist: the preferences of the first COnsumer are satiated.
and
L Xi ~ L Wi + L, y,}
Proof of Proposition 17.B8.2 The approach we follow takes advantage of the fact that the reader may already have been exposed in Chapter 8 to the notion of the Nash equilibrium of a normal form game and. more particularly. to the existence results for Nash equilibrium using best-response correspondences contained in Appendix A to Chapter 8. A gametheoretic approach to the existence of Walrasian equilibrium was taken in the classic paper of Arrow and Debreu (1954). Here we follow Gale and Mas-Colell (1975).
is compact. Then a Walrasian quasiequilibrium exists. We comment briefly on the assumptions. As we have repeatedly illustrated (in Chapters 10, 15. and 16), the convexity assumptions on individual preferences and technologies cannot be dispensed with. 79 The Edgeworth box example of Figure 17.BB.2 shows that the local nonsatiation condition is also required. 80 In contrast. the assumption of rational preferences is entirely dispensable (see the comments at the end of this appendix). The free disposal condition - R~ c lj is also only a matter of convenience. sl Allowing for negative prices. we could simply drop it from our list of conditions (see Exercise 17.BB.3). The assumption (i.3) says that Wi may not belong to the consumption set but that it is possible to reach the consumption set by just eliminating some amounts of commodities from Wi'" Finally. in Appendix A to Chapter 16 we have already investigated the conditions under which the set of feasible allocations is compact.
Definition 17.BB.3: An allocation (x', y') and a price system P disposal quasiequilibrium if
if Xi
then (x·, p) cannot be a Walrasian quasiequilibrium because consuming nothing costs zero and is
preferred by the first consumer to any other consumption. 8!. Because lj is convex and closed, - R~ C lj implies lj -
R~ c lj (Exercise 5.8.5). X, for every i. With this assumption.
Proposition 17.BB.2 yields the existence of a true equilibrium. not just of a quasiequilibrium. Wj ~ .i, can be interpreted (keeping in mind the possibility of free disposal) simply as saying that consumer; could survive Economically, however, the latter assumption is considerably stronger:
without entering the markets of the economy, while the market a strictly positive amount of every good.
Wj
»X, says thac the consumer can supply to
">-; Xi
then p'x; ~ P'w;
+ L 0i/P' Y,'. /
(iii') Li xi ~ L; W;
have p »0. By profit maximization (using the free-disposal technology) and the possibility of inaction, we have p'(x1 + x! - W, - w,) ~ O. Since p'x! $ p'W" this yields p'x1 ~ p'W, > O. But
Wi»
0 constitute a free-
(i) for every i, P'Y, ~ P'Y,' for all Y,E If· (ii') For every i, P'x!' ~p'w;+ LjOijP'Y,', and
79. Recall, however, the important qualification of Section 17.1, and see also the discussion at the end of this appendix. 80. In Figure 17.8B.2, the second consumer has conventional strongly monotone preferences; but for the first consumer both commodities are bads and. thereCore. he is satiated at the origin. Also (I), »0 and w 2 » O. Suppose that x· = (xT. xf) and price vector p#-O constitute a Walrasian quasiequilibrium. Because the preferences of the second consumer 3fC strongly monotone, we must
82. A stronger condition would require that
~
I
I
1
+ L/ Y,'
and p' (Li xi - Li Wi - L/ lj') = O.
Thus. all we have done is replace in Definition 17.BB.I of a quasiequilibrium the exact feasibility condition 'Ti xi = L Wi + Lj yt" by (iii') above. That is. we allow the excess supply of some goods provided that they are free. In Exercise 17.BB.4 you arc asked to show that if one production set. say 1I. satisfies the free-disposal property and if (xf' .... x7. yf ....• yr. p) is a free-disposal quasiequilibrium. then there is l'* ~ yf such that (xf" ... x7. l,·. y! ..... yJ. p) is a Walrasian quasiequilibrium. Therefore. to establish Proposition 17.B8.2, it is enough for us to show that a free-diposal quasiequilibrium exists. We proceed to formalize the free-disposal quasiequilibrium notion as a kind of noncooperative equilibrium for a certain game among I + J + I players. The I and J players are the consumers and the firms. respectively. and their strategies are demand-supply vectors. The extra player is a fictitious market agent (a "grand coordinator") having as his strategy the prices of the L different goods. Since the set A of feasible allocations is bounded, there is r > 0 such that whenever (xl •. ··.x,.y, •...• yJ)E A we have Ix,d < rand IY'jl < r for all i,j, and t. Because we need to have compactness of strategy sets to establish existence, we begin by
635
636
CHAPTER
17:
THE
POSITIVE
THEORY
OF
APPENDIX
EOUILIBRIUM
B:
GENERAL
APPR"" .. "
Tu
IHE
E.XISTENCl
(d,
WAlRASIAN
EQUILloHlljJol
-------------------------------------------------------------------------------------Denote by x,(x, Y, p) c X, the set of consumption bundles xi so defined. Firm j: Chooses productions yj € ~ that are profit maximizing for P on ~. (Firm j's payoff function is simply its profit.) Denote by )~(x, Y, p) c ~ the set of production plans yj so defined. Market Agent: Chooses prices q €!:J. so as to solve
replacing every X, and every lj by a truncated version:
X, = {x, € X,: IXl;!
:5 r for all t}.
~ = {YJ€ lj: IYol :5 r for all
tl·
t.
Note that A c X, x ... X X, x y, x ... x Because (.£" ... ,.£" ... ,0, ... ,0) € A, it follows that .£, € X, for every i, and 0 € ~ for every j. In particular, all the strategy sets are nonempty. Lemma 17.BB.1 shows that in our search for a free-disposal quasiequilibrium we can limit ourselves to the truncated economy.
Max
Lemma 17.BB.1: If all X; and If are convex and (x", y", p) is a free-disposal quasiequilibrium in the truncated economy, that is, if (x", y", p) satisfies Definition 17.BB.3 of free-disposal quasiequilibrium with the consumption and production sets replaced by their truncated versions, then (x", y", p) is also a free-disposal quasiequilibrium for the original untruncated economy.
ilEA
Lemma 17.BB.2: Suppose that (x·, Y", p) is such that xi € x;(x", Y·, p) for all i, v," € Yj(x·, y", p) for all j, and p € p(x·, Y", pl. Then (x·, y., p) is a free-disposal quasiequilibrium for the truncated economy. Proof of Lemma 17.BB.2: We note first that P' yj ;::: 0 for every j (because 0 € 5'). By the definition of x,(·) and jij('), conditions (i) and (ii') of Definition 17.B8.3 are then automatically satisfied. Hence, the only property that remains to be established is (iii'), that is,
Lxi - LllI, - Lyt:5 0
~ O/jP' yj}'
We have P'x~:5 w/(p,Y") = P'w, The strategy sets are:
P'(LX~ i
E
~/.:
PI ;::: 0 for all t and LI Pr
= I}.
Given a strategy profile (x, Y, 1') = (X,' ... ' X" Y,' ... ' YJ' p), the payoff functions and best-responses of the different agents are: Consumer i:
Chooses consumption vectors xi € X, such that (I) p'xi :5 w,(p, y) and (2) xi ?:;,x7 for all xi' E X, satisfying P' XI < w/(p, y). (Consumer i's payoff function can be thought of as giving a payoff 1 if he chooses a consumption vector satisfying this condition, and 0 otherwise.)
(17.BB.l)
J
Only the behavior of the market agent needs comment. Given the total excess demand vector, the market agent chooses prices so as to maximize the value of this vector. Hence, he puts the whole weight of prices (which, recall, have been normalized to lie in the unit simplex) into the commodities with maximal excess demand. As we have already observed when doing the same thing in the proof of Proposition 17.C.1, this is in accord with economic logic: if the objective is to eliminate the excess demand of some commodities, try raising their prices as much as possible. Lemma 17.BB.2 says that an equilibrium of this noncooperative game yields a frec-disposal quasiequilibrium for the truncated economy.
We are now ready to set up a simultaneous-move noncooperative game. To do so we need to specify the players' strategy sets and payoff functions. To simplify notation we assign to every consumer i, price vector P and production profile Y = (YI'···' YJ), a limited liability amount of wealth
For consumer i: X, For firm j: ~ For the market agent: !:J. = {p
i
i
Denote by p(x, y, p) the set of price vectors q so defined.
Proof of Lemma 17.BB.l: Consider a consumer i (the reasoning is similar for a firm). Because (x·, y.) € A, we have Ix1,1 < r for all t; that is, the consumption bundle of consumer i is interior to the truncation bound. Suppose now that x~ fails to satisfy condition (ii') of Definition 17.BB.3 in the nontruncated economy, that is, that there is an x, € X, such that x, >-,x~, and P' x, < P'W, + LJ O/jP' yt- Denote x7 = (I - (I/n»xr + (l/n)x,. For all n we have P' xi < P'W, + Lj O'JP' yt and, by the convexity of preferences, xi ?:;,x7- Also, we can choose an n large enough to have Ixi,l < r for all t. By local nonsatiation there must then be an xi € X, such that xi >-,xi and p'xi < P'W, + Lj O'JP' yt- But then xi € X, and xi >-,x7 ?:;,X~, and so in the truncated economy xr fails to satisfy condition (ii') of Definition 17.BB.3. Thus, (x·, y., p) must not be a free-disposal quasiequilibrium in the truncated economy. This contradiction establishes the result. _
w,(p, y) = P'w, + Max{o,
(LX,-LW,-LYJ)·q.
I !
I
L
and
+ L J8/jp·yt
for all i and therefore
t):5 o.
LW' - L Y j
j
This implies L, xi - L, w, - LJ yt :5 0 because otherwise the value of the solution to problem (17.BB.l) would be positive and so P (which as we have just seen has "'(L, xi - L, w, - LJ ytl:5 0) could not be a maximizing solution vector, that is, a member of p(x", y", pl. It follows that (x·, Y·) € A and so, X?, < r for all i and t. From this we get that the budget equations are satisfied with equality (i.e., ". xi = p'W, + Lj OIjP' yt for all i) because otherwise local nonsatiation yields that for some consumer i there is a preferred consumption strictly interior to consumer i's budget set in the truncated economy, implying x~ ¢ X,(X·, y., pl. We therefore conclude that we also have P'(L, xr - LI W, - LJ ytl = O. This completes the proof. _
638
CHAPTER
17:
THE
POSITIVE
THEORY
OF
APPENDIX
EQUILIBRIUM
N OW, as we discussed in Appendix A to Chapter 8 (see the proof of Proposition
8.0.3 presented there), under appropriate conditions on the best-response correspondences, this noncooperative game has an equilibrium. Lemma 17.88.3: Suppose that the correspondences x;!"), ~(.), and p(.) are nonempty, convex valued, and upper hemicontinuous. Then there is (x', y', p) such that x7 E x,(x', yO, p) for all i, Yj' E ~(x', yO, p) for all j, and p E p(x', yO, pl. Proof of Lemma 17.BB.3: We are simply looking for a fixed point of the correspondence '1'(.) from X, x ... X X, X YI X •• '. x >J x A to itself defined by 'I'(x.y,p) = x,(x,y,p)
x··· x x,(x,y,p) x y,(x, y, p) x··· x yAx,y,p) x p(x, y, p).
The correspondence '1'(.) is nonempty, convex valued, and upper hemicontinuous. The existence of a fixed point follows directly from Kakutani's fixed point theorem (sec Section M.I of the Mathematical Appendix). _ Lemmas 17.BB.4 to 17.BB.6 verify that the best-response correspondence of this noncooperative game is nonempty, convex valued, and upper hemicontinuous· ' Lemma 17.88.4: For all strategy profiles (x, y, p), the sets xi(x, y, p), pIx, y, p) are nonempty.
~(x,
y, p), and
Proof of Lemma 17.BB.4: For Yj(x, y, p) and pIx, y, p) the claim is clear enough since we arc maximizing a continuous (in fact, linear) function on, respectively, the noncmpty, compact sets ~ and A. For Xi(X, y, p), recall that the continuity of ;:::i implies the existence of a continuous utility representation "i(') for ;:::, .• 4 Let be a maximizer of the continuous function "i(X,} on the nonempty compact budget set (Xi E Xi: p' Xi::;; Wi(p, yO)}. Then x; E Xi (X, y, pl. The budget set is nonempty because Xi E Xi and ,Xi::;; Wi' With P
x;
Lemma 17.88.5: For all strategy profiles the sets xi(x, y, p), Yj(x, y, p), and pIx, y, p) are convex. Proof of Lemma 17.BB.5: We establish the claim for Xi(X, y, pl. You are asked to complete the proof in Exercise 17.BB.6. Suppose that Xi' x; E Xi (x, y, p) and consider Xi. = aXi + (I - a)x;, for any a E [0,1]. Note first that P·Xi. ::;; W,(P, y). In addition, by the convexity of preferences we cannot have Xi :>iXi. and x; :>, x,. (Exercise 17.BB.5). So suppose that Xi. ;:::iXi' Consider now any xi e X, with P'x, < wi(p, y). Then since x, E Xi(X, y, p) we have Xi;::: i xi, and so Xi. ;:::ixi. We conclude that Xi. e Xi(X, y, pl. A similar conclusion follows if Xi. ;:::i x;. Hence, xi(x, y, p) is a convex set. _ Lemma 17.88.6: The correspondences
xi ('), ~(.), and p(.) are upper hemicontinuous.
Proof of Lemma 17.BB.6: Again, we limit ourselves to Xi(·)' Exercise 17.BB.7 asks you to complete the proof for M') and p('). 83. For the firms and the market game this result is covered by Proposition 8.DJ, but for the consumers we need a special argument (as defined, the payoff runctions of the consumers arc not continuous).
84. This was proved in Proposition lC.1 for monotone preferences on Rt As we pointed out there. however. the conclusion actually depends only on the continuity of the preference relation.
B:
GENERAL
APPRQACH
TQ
THE
EXISTENCE
OF
WALRASIAN
x;
Let p" - t p, y" - t y, x" - t x, and x;" -+ as n - t 00, and suppose that x;" E x,(x", y", pO). We need to show that x; e XI(P, x, y). From p"'x;" ::;; W,(P", yA) we get p'x; ::;; W,(pA, yA). Consider now any xi e with x;' :>iX;. Then, by the continuity of preferences, xi :>, x;" for n large enough. Hence, p". xi :?: wi(p", yO). Going to the limit we get p' x7
Xl
x;
closed-graph property that we have replaced preference maximization by the weaker objective of expenditure minimization in the definition of the objectives of the consumer. _
The combination of Lemmas I7.BB.4 to 17.BB.6 establishes that the given best-response correspondences satisfy the properties required in lemma 17.BB.3 for the existence of a fixed-point, which completes the proof of Proposition 17.BB.2. _ The assumptions on preferences and technologies can be weakened in an important respect. Our existence argument requires only that the best·response correspondence .i,(x, y, p) and rj(x, y, p) be convex valued and upper hemicontinuous. Beyond this, the proof imposes no restrictions whatsoever on the dependence of consumers' and firms' choices on the ~state" variables (x. y. pl. Thus we could allow consumers' tastes, or firms' technologies, to depend on prices (money illusion?), on the choices of other consumers or firms (a form of externalities), or even on own consumption (e.g., tastes could depend on a current reference point-a source
I).·"··
of incompleteness or non transitivity of preferences already illustrated in Chapter The following is an example of the sort of generality that can be accommodated: Suppose that consumer preferences are given to us by means of utility functions -,('; x, y, p) defined on X, but dependent, in principle, on the state of the economy. If for every (x, y, p) the conditions of Proposition 17.BB.2 are satisfied, and the parametric dependence on (x, y, p) is continuous, then a Walrasian quasiequilibrium still exists. The proof does not need any change. We can make a similar point with respect to the possibility that firms' technologies depend on external effects, with, then, an added theoretical payoff. It allows us to see that equilibrium exists if the technology of the firm is convex: il does nol mailer iflhe ~aggregale"lechnology oflhe economy is convex. See Exercise 17.B8.8 for more on this. The existence proof we have given in this appendix is an example of a "large space" proof. The fixed'point argument (in our case phrased as a Nash equilibrium existence argument) has been developed in a disaggregated domain where all the equilibrating variables have been listed separately. The advantage of proceeding this way is that the argument remains very flexible and allows us to incorporate the weakest possible conditions without extra effort (as the last paragraph has illustrated). The disadvantage, of course, is that the fixed point may be 85. Suppose, for example, that the utility function of a consumer is given to us in the form
that is, the evaluation of possible consumptions depends on the current consumption. Without loss of generality we can normalize u,(x,: x,) = 0 for every x,. Define the induced weak and strict preference relations '=/ and >-1 on Xi by, respectively, .. xj ~,x/ if ",(xi; x,) ~ 0'" and .. xi >-iXi if uj(x/; x,) > 0." Then the relations ~, and >-, contain all the relevant information for equilibrium analysis. Note, however, that it is perfectly possible for It, not to be complete and for neither It; nor r; to be transitive. See Shafer (1974) and Gale and Mas-Colell (1975) for more u;(·: x;);
on this.
86. Another example of dependence on the overall consumption vector of the economy arises if, for example. we are considering equilibrium at a given point in time. Then current consumptions in the economy (e.g., purchases of physical or financial assets) will typically affect future prices; these, in turn, will innuence current preferences via expectations.
EQUILIBRtUM
639
640
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EXERCISl!:i
EOUILIBRIUM
hard to compute and cumbersome to analyze. Usually, as we have seen in Section 17.C and in Appendix A of this chapter, it is possible to work with more aggregated, reduced systems. In fact, the general point duly made, it is worthwhile to observe that this is so even under the assumptions of Proposition 17.BB.2.17 We elaborate briefly on this. We can prove Proposition 17.BB.2 by selling up a two-player game instead of an I + J + I one'· The first player is an aggregate consumer-firm that has L, X, - IL' w,} - LI ~ as its strategy set; the second is, as before, a market agent having 6. as its strategy set. Given p E 6., the first agent responds with the set of vectors z expressible as z = 1:. x, - L, w, - LI YI' where Jj is profit maximizing in ~ for every j, and x, E X, is such that (I) p' x, :5: P'W, + LI 0'1 p' Yj and (2) x, :::,x; whenever p'x; < p'W, + LI 0'1 p' YI' As before, the market agent responds with the set of q E 6. that maximize Z'q on 6.. Once this two-person game has been set up, the proof proceeds as for Proposition 17.BB.2. You should check this in Exercise 17.BB.9. If for any p E 6. the preference-maximizing choices of consumers, x,( p), and the profitmaximizing choices of firms, )j(p), were single valued, we could go one step further and consider a game with a single player (the market agent). Given p, we would then let the best response of the market agent be the set of price VectOrsqE 6. that maximizes [1:, x,{p) - L, w, - LI y)(p»)-q on Ii. In essence, this is what we did in the proof of Proposition I7.CI.
REFERENCES Arrow. K., and G. Debrcu. (1954). Existence of equilibrium for a competitive economy. f:('(mmnerricu 22: 265-90. Arrow, K.. and F. Hahn. (1971). General Competitive Analys;s. San Francisco: Holden-Day. Arrow, K, and M. Intriligalor, eds. (1982). Handbook of Mathematical Ewnomics. vol. II. Amsterdam:
North-Holland. Barone. E. (1908). II ministro della produzione nello stato colletti vista. Giornali degli economisti. [Reprinted as: The ministry or production in the collectivist state. in Collectivist Economic Planning. edited by F. H. Hayek. London: Routledge, 1935.] Balasko, Y. (1988). Foundations of the Tlteory of General Equifibrium. Orlando: Academic Press. Becker, G. (1962). Irrational behavior and economic (heory. Journal of Political E"onomy 70: 1-13. Brown. D., and R. Matzkin. (1993). Walrasian comparative statics. Mimeograph. Northwestern University. Chipman, J. (1970). External economies of scale and competitive equilibrium. Quarterly Journal of Economics 84: 347-85.
Debreu, G. (1959~ Tlteoryof Value. New York: Wiley. Debreu, G. (1970). Eeonomies with a finite sel of equilibria. Econometrica 38: 387-92. Debreu, G. (1974). Excess demand functions. Journal of Mathematical Economics I: 15-21. Dierker, E. (1972). Two remarks on the number of equilibria of an economy. Econom.trica 40: 951-53. Fisher. F. (1983). Disequilibrium Foundations of Equilibrium Economics. Cambridge. U.K.: Cambridge University Press. Gale. 0 .. and A. Mas~Colell. (1975). An equilibrium existence theorem for a general model without ordered preferences. Journal of Mathematical Economics 2: 9-15. [For some corrections see Journal of Mathematica/Economics 6: 297-98, 1979.] Garcia, C. B.. and W. I. Zangwill. (1981). Pathways to Solutions, Fixed Points and Equilibria. Englewood
Cliffs. N.J.: Prentice-Hall. Grandmont, J. M. (1992). Transformations of Ihe commodity space. behavioral heterogeneity. and the
Hildenbrand. W. (1994). Market Demand: Theory and Empirical Evidenu. Princeton. NJ.: Princeton University Press. Hildenbrand. W.o and H. Sonnenschein, cds. (1991). Handbook of Mathematical Economics. vol. IV. Amsterdam: North-Holland. Kehoe. T. (1985). Multiplicity of equilibrium and comparative statics, Quarterly Journal of Economics 100: 119-48. Kehoc. T. (1991). Compulation and mulliplicily of equilibria. Chap. 38 in Handbook of Mathematical Economics. vol. IV, edited by W. Hildenbrand, and H. Sonnenschein. Amsterdam: North-Holland. Lange. O. (1938). On the economic theory of socialism. In On the Economic Theory of Socialism. edited by It Lippincott. Minneapolis: University of Minnesota Press. McKenzie, L. (1959). On the cltistence of general equilibrium ror a competitive market. E('(mometrica 27: 54-71. Mantd, R. (1974). On the characterization or aggregate excess demand. Journal of Economic Theory 7: 34M-53. Mantel. R. (1976). itomothctic preferences and community CltCCSS demand functions. Joumat (if Eronomk n",o,)" 12: 197-201. Marshall. A. (1920). Princip/t'.'i oj £t'mlomics. 8th ed. London: Macmillan. Mils-Colell, A. (1977). On the equilibrium price set of an exchange economy. JOllrnal of Math('",otical £conomks 4: 117-26. Ma~-Colcll, A. (1985). nil! Theor}' of Gt'nt'fal £wnomic Eqllilihrj,~m: A Dij]er('ntiahlt' Approach. Camhridge. U.K.: Cambridge University Press. Ma~-(,ulcll. A. (1986). Notes. on price and quantity latonnement. In Aloc/d.'i (~r Economic Dytlamin, edited by fL Sonnenschein. Lecture Notes in Economics and Mathematical Systems No. 264. Berlin: Springcr- Vcrl;:tg. ~b~-Coldl. A. ( 1991 ). On the uniqueness of equilihrium once again. Chap. 12 in Eqllilihrillm Tht·ory and A,'plicatirm.'i. edit .......t by W. Barnell, B. Cornel, C. O'Asprcmont, J. Gabszewicz and A. Mas-Colell. Cambridge, U.K_: Cambridge Univtrsily Press. Milgrom. P.. and C. Shannon. (1994). Monotone comparative statics. Econometrica 62: 157-180. Ncgishi, T. (1960)_ Welfare economics and existence or an equilibrium for a competitive economy. Mc.'twe(·onomiC"a 12: 92-97_ Rader. T. (1972). Thenr), lif General Economic Equilihrium. New York: Academic Press. Saari. D., and C. Simon. (1978). Effective price mechanisms. Econometrica 46: 1097-125. Samuelson. P. (1947). Foundations of Economic Anat}'sis. Cambridge. Mass.: Harvard University Press. Scarf. H. (1973). The ComplllOlioll of Economic Equilihria (in collaboralion with T. Hansen). New Haven: Yalc University Press. Sh;lrcr, W. (1974). The non-transitive consumer. Econometrica 42: 913-19_ Sharer, W .• and H. Sonnenschein. (1982). Market demand and ucess demand functions. Chap. 14 in 110m/hook of Mathematical E(·onomir.~ vol. II. edited by K. Arrow and M. Intriligator. Amsterdam:
North·Holland. Shoven. 1.. and J. Whalley. (1992). Applring Gtn"al £quilibrium. New York: Cambridge University Press. Sonnenschein, H. (1973)_ Do Walras' identity and continuity characterize the class of community excess demand runctions'! Journal of Economic Theory 6: 345-54. Smale. S. (1976). A convergent process of price adjustment and global Newton methods. Journal of MllIlrematical Economics 3: 107-20. Starr. R. (1969). Quasi-equilibria in markets with non--convex preferences. Econometrica 37: 25-38. V.uian, H. (1977). Non-Walrasian equilibria. Econometrica 45: 573-90. Walras, L. (1874). Elements d'Economie Politique Pure. Lausanne: Corbaz. [Translated as: EI('mems of Pure E(·(tn(tmic_~. Homewood, III.: Irwin, 1954.]
aggregation problem. Journal of Economic Theory 57: 1-35.
Hahn, F. (1982). Stability. Chap. 16 in Handbook of Mathematical Economics, vol. II. edited by K. Arrow. and M. Intriligator. Amsterdam: North-Holland. 87. But it is not so for the generalizations described in the previous paragraph. 88. This was the approach taken in Debreu (1959).
EXERCISES
17.B.IA Show that for a pure exchange economy with J = 1 and Y, = -R~, "J'j:5: 0, p'rt = 0, and p ~ 0" ir and only ir "yj E YI and p' yj ~ p' Yt for all Yt E YI ."
041
642
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
EXERCISES
643
------------------------------------------------------------------------------------------------17.0.2" Prove property (v) of Proposition 17.0.2. The proof of Proposition 17.B.2 in the text contains a hint. Recall also the following technical fact: any bounded sequence in RL has a convergent subsequence. 17.B.3" Suppose that z(·) is an aggregate excess demand function satisfying conditions (i) to (v) of Proposition 17.8.2. Let p. - P with some, but not all, of the components of P being zero.
17,C.3" Consider an exchange economy in which every consumer I has continuous, strongly monotone, strictly convex preferences, and w, » O. The peculiarity of the equilibrium problem to be considered is that the consumer will now pay a type of tax on his gross consumption; moreover, this tax can differ across commodities and consumers. We will also assume that total tax receipts are rebated equally across consumers and in a lump-sum fashion. Specifically, for every i there is a vector of given tax rates Ii = (t 11_ .•. tu) ~ 0 and for every price vector P » 0 the budget set of consumer i is 1
(a) Show that as n becomes large, the maximal excess demand is always obtained for some commodity whose price goes to zero. (b) Argue (if possible by example) that a commodity whose price goes to zero may actually remain in excess supply for all n. [Hint: Relative prices mailer.] 17.B.4" Suppose that there are J firms whose production sets Y ... , >J C RL are closed, " strictly convex, and bounded above. Suppose also that a strictly positive consumption bundle is producible using the initial endowments and the economy's aggregate production set Y = Li Ij (i.e., there is an x » 0 such that x e {LI w,} + Y). Show that the production inclusive aggregate excess demand function i(p) in (l7.B.3) satisfies properties (i) to (v) of Proposition 17.B.2. 17.B.S' Suppose that there are J firms. Each firm produces a single output under conditions of constant returns. The unit cost function of firm j is cJ(p), which we assume to be dilTerentiable. The consumption side of the economy is expressed by an aggregate excess demand function z(p). Write down an equation system similar to (l7.B.4)-(I7.B.5) for the equilibria of this economy. 17.8.6 C [Rader (1972)] Suppose that there is a single production set Yand that Y is a closed, convex cone satisfying free disposal. Consider the following exchange equilibrium problem. Given prices P = (P,' ... , pd, every consumer i chooses a vector V, e RL so as to maximize ?:;:, on the set {x,e Xi: P'V,:S; P'w" and x, ~ V, + y for some ye Y}. The price vector p and the choices v· = (vr, ... , vr> are in equilibrium ifL' v~ = L, w,. Show that, under the standard assumptions on preferences and consumption sets, the price vector and the individual consumptions constitute a Walrasian equilibrium for the economy with production. Interpret. 17.C.I' Verify that the correspondence f(·) introduced in the proof of Proposition l7.C.1 is convex· valued.
17.C.2C Show that a convex-valued correspondence z( -) defined on R~ + and satisfying the conditions (i) to (v) listed below (parallel to the corresponding conditions in Proposition 17.C.I) admits a solution; that is, there is a p with Oe z(p). (i) (ii) (iii) (iv) (v)
z(·) is upper-hemicontinuous.
{z;, ... , zi.} -
00.
[Hilll: If you try to replicate exactly the proof of Proposition l7.C.l you will run into difficulties with the upper-hemicontinuity condition. A possible three-step approach goes as follows: (I) Show that fo<£ > 0 small enough the solutions must be contained in 11, {p e 11: PI ~ dar all (2) argue then that for r > 0 large enough, one has z(p) c [-r, r]L for every p E 11,; finally, (3) carry out a fixed-point argument in the domain 11, x [-r, rlL. For an easier result, you could limit yourself to prove the convex-valued parallel to Proposition l7.e2. The suggested domain for the fixed-point argument is then 11 x [-r, rl'.
n;
R~: ~ (I + IIi)PIXIi :s; Wi}'
An equilibrium wilh laxes is then a price vector p »0 and an allocation (xr, ...• xl) with L. Wi such that every i maximizes preferences in Bi(p, P'W, + (II/XL" I"PIX"»'
LX~ =
(a) Illustrate the notion of an equilibrium with taxes in an Edgeworth box. Verify that an equilibrium with taxes need not be a Pareto optimum. (b) Apply Proposition 17.C.l to show that an equilibrium with taxes exists. (c) As formulated here, the taxes are on gross consumptions. If they were imposed instead on net consumptions, that is, on amounts purchased or sold, then (assuming the same rate for buying or selling) the budget set would be B,(p,
r,)
=
{x;eR~:p.(X'-W;) + ~l/lpli(x" -w,,)I:s; r,},
where the r, are the lump-sum rebates. In what way does this budget set differ from that described previously for the case of taxes on gross consumptions? Represent graphically. Notice the kinks. (d) Write down a budget set for the situation similar to (c) except that the tax rates for amounts bought or sold may be different. (c) (More advanced) How would you approach the existence issue for the modification described in (c)? 17.C.4 A Consider a pure exchange economy. The only novelty is that a progressive tax system is instituted according to the following rule: individual wealth is no longer p'W,; instead, anyone with wealth above the mean of the population must contribute half of the excess over the mean into a fund, and those below the mean receive a contribution from the fund in proportion to their deficiency below the mean. (a) For a two-consumer society with endowments w, = (1,2) and after-tax wealths of the two consumers as a function of prices.
Wz
= (2, I), write the
(b) If the consumer preferences are continuous, strictly convex, and strongly monotone, will the excess demand functions satisfy the conditions required for existence in Proposition I7.CI given that wealth is being redistributed in this way?
z(·) is homogeneous of degree zero.
For every p and z e z(p) we have p' z = 0 (Walras' law). There is 5 E R such that Zl> -5 for any z e z(p) and p. If p. _ p ",. 0, z· e z(p·) and PI = 0 for some t, then Max
Bi(p, Wi) = {Xi E
=
17.C.S" Consider a population of / consumers. Every consumer i has consumption set R~ and continuous, strictly convex preferences ?:;:i' Suppose, in addition, that every i has a household technology>; c RL satisfying 0 E t;. We can then define the induced preferences ?:;:~ on by Xi ?:;:~ xi if and only ifror any Yi e t; with xi + Yi ~ 0 there is y, e lj with x, + y, ~ 0 and Xi + Yi ?:;:,x; + yi (i.e., whatever can be done from xi, something at least as good can be obtained from Xi)'
R'.
(a) Show that induced preferences are rational, that is, complete and transitive. (b) Show that if t; is convex then induced preferences
?:;:~
are convex.
644
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
(e) Suppose that goods are of two kinds: marketed goods and nonmarketed household goods. Initial preferences ;::;, care only about household goods, and initial endowments W, have nonzero entries only for marketed goods. Use the concept of induoed preferences to set up the equilibrium problem as one that is formally a problem of pure exchange among marketed goods. Discuss.
-
E X ERe I 5 E 5
17.0.S' Show by explicit computation that the index of the equilibrium of a one-consumer Cobb- Douglas pure exchange economy is + I. 17.E.I' Derive expressions (17.E.I) and (I7.E.2). 17.E.2' Derive expression (17.E.3).
17.C.6B Let L = 2. Consider conditions (il, (iii), and (iv) of Proposition 17.B.2 Exhibit four examples such that in each of the examples only one condition fails and yet the system of equations z(p) = 0 has no solution. Why is condition (ii) not included in the list?
17.E.3" Provide explicit utility functions rationalizing at a given price vector p the individual excess demands Z,( p) and matrices of price effects Dz,(p) constructed in the proof of Proposition 17.E.2.
17.0.IB Consider an exchange economy with two commodities and two consumers. Both consumers have homothetic preferences of the constant elasticity variety. Moreover, the elasticity of substitution is the same for both consumers and is small (i.e., goods are close to perfect complements). Specifically,
17.E.4" Consider the two-commodity case. Give an example of a function z(p) defined on = {(PI' P2)>> 0: E < (p,lp2) < (lIE)}, and with values in R2, that is continuous, is homogeneous of degree zero, satisfies Walras' law, and cannot be generated from a rational preference relation. Represent graphically the offer curve associated with this function. NOle that it goes through the initial endowment point and compare with the construction used in Figure 17.E.2.
u,(x", x,,) = (2x~,
+ X~,)"P
and
u,(x", x,,) = (Xf2
+ 2x~2)''''
r,
and p = -4. The endowments are w, = (1,0) and 002 = (0,1). Compute the excess demand function of this economy and verify that there are multiple equilibria.
17.E.SA Show that the choices represented in Figure 17.E.3 cannot be generated from consumers wilh endowment vectors bounded above by (1,1) and nonnegative consumption.
17.0.2' Apply the implicit function theorem to show that if J(v) = 0 is a system of M equations in N unknowns and if at jj we haye J(V) = 0 and rank DJ(jj) = M, then in a neighborhood of jj the solution set of J(') = 0 can be parameterized by means of N - M parameters.
I7.E.6 A Show that the excess demand function Z,(P) = e' - P,P, defined for IIpll = I. is proportionally one-to-one in the sense used in the general proof of Proposition 17.E.3 (at the end of Section 17.E).
17.0.3' Carry out explicitly the computations for Proposition 17.0.4.
17.E.7" Show directly that Ihe excess demand function z,(p) = e' - P,P used in the general proof of Proposition 17.E.3 salisfies the strong axiom of revealed preference.
17.0.4c Consider a two-commodity, two-consumer exchange economy satisfying the appropriate differentiability conditions on utility and demand functions. There is a total endowment vector 6i» O. Show that for almost eyery W,« 6i the economy defined by the initial endowments w, and W2 = 6i - w, has a finite number of equilibria. This differs from the situation in Proposition 17.0.2 in that total endowments are kept fixed. [Hint: You should use the properties of the Slutsky matrix.] 17.0.5" Consider a two-commodity, two-consumer exchange economy satisfying the appropriate differentiability conditions on utility and demand functions. Set the equilibrium problem as an equation system in the consumption variables x, e R~ and X2 e R~, the price variables p e R~, and the reciprocals of the marginal utilities of wealth )., e Rand ).2 e R (neglect the possibility of boundary equilibria). The parameters of the system are the initial endowments (00" (2) e R". Prove without further aggregation that (after deleting one equation and one unknown) the system satisfies the full rank condition of the transyersality theorem. 17.0.68 The setup is identical to Exercise 17.D.S except that an externality is allowed: The (differentiable) utility function of consumer I may depend on the consumption of COnsumer 2; that is, it has the form u,(x"x2) where X, is consumer i's consumption bundle [but we still have U2(X 2 )]. Equilibrium is defined as usual, with the proviso that consumer I takes consumer 2's consumption as given. Show that, generically on initial endowments (00" ( 2 ) E R", the number of equilibria is finite. 17.0.7 8 Suppose the agents of an overall exchange economy are distributed across N islands with no communication among them. Each island economy has three equilibria. (a) Argue that the number of equilibria in the overall economy is 3N • (b) Suppose now that the islands' economies are identical and that there is a possibility of communication across the islands: free and costless transportation of commodities. Show that then the number of equilibria is 3.
17.F.lc Show that expression (I7.F.2) gives rise to a negative semidefinite matrix of price effects. D:( pl. if initial endowments are proportional among themselves or if consumptions are proportional among themselves. 17.F.2' Complete the requested verification of Example 17.F.1. 17.F.3" There are four goods and two consumers. The endowments of the consumers are "'I = (W,I.W21'0.0) and 002 = (W , 2. Wu. 0,0). Consumer I spends all his wealth on good 3 while consumer 2 does the same on good 4. Specify some values of w, and W 2 for which the corresponding excess demand of this economy does not satisfy the weak axiom of revealed preference.
17.F.4A Suppose that there are L goods but that for every consumer there is a good such that at any price the consumer spends all his wealth on that good (perhaps goods are distinguished by their location). Show that the aggregate excess demand will satisfy the (weak) gross substitute property. 17.F.Sc Complete the missing steps of Example 17.F.2. 17.F.6 c Consider a two-consumption-good, two-factor model with constant returns and no joint production. In fact. suppose that the production functions for the two consumption goods are Cobb-Douglas. Consumers have holdings of factors and have preferences only for the two consumption goods. The economy is a closed economy (at equilibrium. consumption must equal production). Suppose that the two goods are normal and gross substitutes in the demand JUII(·tion of the consumers. Define an induced exchange economy for factors of production by assuming that at any vector of factor prices the two goods are priced at average cost and the final demand for them is met. Show that the resulting aggregate excess demand for factors of production has the gross substitute property and, consequently. that there is a unique equilibrium for the overall economy.
64:1
646
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EX E R CIS E S
EOUILIBRIUM
------------------------------------------------------------------------------------------------example, be the system of excess demands corresponding to a subgroup of markets with the prices of commodities outside the group kept fixed.
17.F.7A Prove expression (17.F.3) for L = 2. 17.F.SA Show that expression (J 7.F.3) implies that the set of solutions to z(p) = 0 is convex. 17.F.9 S Consider an economy with a single constant returns production set Y. Preferences are continuous, strictly convex, and strongly monotone. Suppose that the feasible consumptions (Xl' ... ' x,) are associated with a Walrasian equilibrium. Assume, moreover, that no trade is required to allain these consumptions if Yis freely available to all consumers; that is XI - W, E Y for all i. Show then that those are the only possible equilibrium consumptions. 17.F.10A Show that expression (17.F.3) implies that Dz(p) is negative semidefinite at an equilibrium p.
=
=
17.F.J1 B Show that if z(p) 0, rank Dz(p) L - I, and Dz(p) is negative semidefinite, then, for any (, the (L - I) x (L - I) matrix obtained from Dz(p) by deleting the (th row and column has a determinant of sign (_I)L-'. [Hint: From Section M.D ofthe Mathematical Appendix you know that rank Dz(p) = L - I implies that the (L - I) x (L - I) matrix under study is nonsingular. Consider then Dz(p) - d.] 17.F.12" Show that if z(p) = 0 and Dz(p) has the gross substitute sign pallern, then the (L - I) x (L - I) matrix obtained from Dz(p) by deleting the Ith row and column has a
lIegalive dominanl diagonal (see Section M.D of the Mathematical Appendix for this concept)
and is therefore negative definite. 17.F.13 A Provide the missing computation for Example 17.F.3. B
17.F.14 Consider a firm that produces good lout of goods 1= 2, ... , L by means of a production function /(v" .. .• vd. Assume that /(.) is concave, increasing. and twice continuously differentiable. We say that I and I' are complements at the input combination v = (v" . .. , vd if 0' /(v)/ov, ov,. > O. (a) Verify that for the Cobb-Douglas production function f(v" ... , VL) = vi' x ... x ~, + ... + ~L !> I, any two inputs are complements at any •.
.r',
(b) Suppose that f(') is of the constant returns type. Show that at any. and for any there is an (' that is a complement to t at v.
t
(c) Suppose now that f(·) is strictly concave and that any two inputs are complements at any •. Let ",(p" ... , pel be the input demand functions. Show thaI, for any t, a.tlap, > 0, < 0, and < 0 for t' # I.
av,/ap,
a.tlaPr
(d) Discuss the implications of (a) to (c) for uniqueness theorems that rely on the gross substitute property.
(a) We say that g(.) satisfies the strong gross subslilule properly (SGS) if for some" > 0 every coordinate of the function "g(p) + p is strictly increasing in p and (IXU(P) + p) E [0. rJ" for every p E [0, r]H. Show that if g(p) has the SGS property then it also has the GS property. (b) Show by example that the GS property does not imply the SGS property. Establish, however, that if g(') is continuously differentiable and the GS property is satisfied then the SGS property holds.
From now on we assume that g(') satisfies the SGS property. (e) Show that there is an equilibrium, that is, a p with g(p) = O. Illustrate graphically for the case N = I. [Hinl: Quote the Tarski fixed point theorem from Section M.I of the Mathematical Appendix, or, if you prefer, assume continuity and apply Brouwer's fixed point theorem.] (d) Give an example for N = 2 where the equilibrium is not unique.
=
=
(e) Suppose that g(p) g(p') O. Show that there must be an equilibrium p' such that p. :2: P and p' :2: p'. Similarly, there is an equilibrium p- such that p- :S P and p- :S p'. [Hint: Apply the argument in (e) to the domain [Max {p" p;}. r] x ... x [Max {PH' pj,}, r].]
(f) Argue (you can assume continuity here) that the equilibrium set satisfies a strong and very special property, namely, that it has a maximal and a minimal equilibrium. That is. there are pm.. and pm;. such that g(pm .. ) = g(pm;") = 0 and pm;" S p!> p-' whenever g(p) = O. (g) Assume now that g(.) is also differentiable. Suppose that we know that at equilibrium, that is, whenever g(p) = 0, the matrix Dg(p) has a negative dominanl diagonal; that is, Dg(p)v« 0 for a .» O. Argue (perhaps non rigorously) that the equilibrium must then be
unique. (h) Suppose that g(') is the usual excess demand system for the first N goods of an economy with N + I goods in which the last price has been fixed to equal I and the overall (N + I)-good excess demand system satisfies the gross substitute property. Apply (g) to show that the equilibrium is unique. 17.F.17A [Becker (1962), Grandmont (1992)] Suppose that L = 2 and you have a continuum of consumers. All consumers have the same initial endowments; they arc not rational, however. Given a budget set, they choose at random from consumption bundles on the budget line using a uniform distribution among the nonnegative consumptions. Let z(p) be the average excess demand (= expected value of a single consumer's choice). Show that z(·) can be generated from preference maximization of a Cobb-Douglas utility function (thus the economy admits a positive representative consumer in the sense of Section 4.0).
17.F.lS" Consider a one-consumer economy with prOduction and strictly convex preferences. There is a system of ad valorem taxes I = (I" ... , Id creating a wedge between consumer and producer prices; that is, PI = (I + I,)q, where P, and q, are, respectively, the consumer and producer price for good t. Tax receipts arc turned back in lump-sum fashion. Write the definition of (distorted) equilibrium. Show that the equilibrium is unique if the production sector is of the Leontief type (a single primary factor, no joint production, constant returns) and all goods are normal in consumption. can you argue by example the nondispensability of the last normality condition? If this is simpler, you can limit your discussion to the case of two commodities (one input and one output).
17.G.l" Suppose that in an exchange economy (and with the normalization PL = I) we are given equilibrium prices p(w,) as a differentiable function defined as an open domain of the endowments of the first L - I goods of the first consumer, W, = (w", ... , w L -,.,). All the remaining endowments are kept fixed. Suppose that the demand function of the first consumer is strictly normal in the sense that Dw,x,(p, w,) » 0 through the relevant domain of (p, w,). Show then that for any';', and;; = P(';',), we have rank D.. i,(;;;';',) = L - 1 and rank Dp(';',) = L - I, where il(p; w,) is the excess demand function ~f the first consumer for the first L - 1 goods.
17.F.16C Suppose that g(p) = (g,(p), ...• gH(P» is defined in the domain [0, r]H and that g(O, . .. ,0) » (0, ... ,0), g(r, ... , r) « (0. ...• 0). Note that we do not assume Walras' law, homogeneity of degree zero, or, for that maller, continuity. The function g(') could, for
it;;;
17.G.2" The setting is as in Exercise 17.G.1 or as in Proposition 17.G.2. Suppose that cD,) = O. Show that there are economies with D,£(;;; cD,) an (L - I) x (L - I) negative definite matrix but where op,(,;")/aw,, > O. [Hint: Use Proposition 17.G.I and the arguments employed in its proof.]
647
648
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
EXERC'SES
649
------------------------------------------------------------------------------------------------I7.G.3C The setting is a. in Exercise 17.F.16. except that now we have two functions g(p) ERN and (J(p) eRN. Each of these functions satisfies the conditions of Exercise 17.F.16 (in particular the SDS property). In addition. we assume that b(-) is an upward shift of g('); that i•• (J(p) ~ g(p) for every pE [O.r]N. Prove that if (pml •• p .... ) and (~ml •• P"''') are the minimal and maximal equilibrium price vectors (see Exercise 17.F.16) for g(') and b('), respectively. then pmln ~ pmln and pm.. ~ pm... [You can assume that g(') and !I(') are continuous; if this makes things simpler. assume also that both functions have a unique solution.] Represent graphically for the case N = I. 17.H.l c Suppose that the system of excess demand functions z(p) satisfies the gross substitute property. Consider the tatonnement price dynamics dp, = Z/(P) dt
For any price vector p let o/t(p)
for every
t.
(0)
= Max {z,(p)/p, •...• zL(p)/pd.
(I) Argue that if pCt) is a solution for the above tatonnement dynamics (i.e .•
dp,(t)/dt - ZI(pCt» for every t and t) and z(pCO» ¢ 0 then o/t(pCt» should be decreasing through time. [Him: If ZI(pCt»/PI(t) = o/t(pCr)) then PI(t)/PI'(t) cannot decrease at t for any t'. Hence. Z/(pCt» cannot increase. whereas P, surely increases.]
(b) Argue that p(t) converges to an equilibrium price as t dynamics (0) Walras' law implies that LI pJ(t) = constant.]
00 •
[Hint: Recall that for the
17.H.2" There is an output good and a numeraire. The price of the output good is p. The data of our problem are given by two functions: The consumption side of the economy provides an excess demand function z(p) for the output good. and the production side an increasing inverse output supply function pCz). Both functions are differentiable. In addition. their graphs cross at (1.1). which is the equilibrium we will concentrate on in this exercise. Given this selling we can define two one-variable dynamics: (i) In Walras price dynamics we assume that at p the price increases or decreases according to the sign of the difference between excess demand and (direct) supply at p. (ii) In Marshall quantity dynamics we assume that at z production increases or decreases according to the sign of the difference between the demand price (i.e .• the inverse excess demand) and the supply price (i.e.• p(z» at z. (a) Write the above formally and interpret economically. (b) Suppose that the technology is nearly of the constant returns type. Show then that around the equilibrium (1.1) the system is always Walrasian stable but that Marshallian stability depends on the slope of the excess demand function (in what way?). (c) Write general price and quantity dynamics where prices move II la Walras and quantities la Marshall. Draw a (P. z) phase diagram and argue that in the typical case dynamic trajectories will spiral around the equilibrium.
a
(d) Go back to the technology specification of (b). Show that the system in (c) is locally stable if and only if the equilibrium is Marshallian stable.
17.1.1A Argue that the replica procedure described at the beginning of Section 11.1 does effectively include the case where the numbers of consumers of different types are not the same (a"ume, for simplicity. that the proportions of the different types are rational numbers). [Hint: Redefine the size of the original economy.]
v'.
17.1.2A Consider for a one-input. one-output problem the production function q = where " is thc amount of input. Show that the corresponding production set Y is additive but that the smallest cone containing it. yo. is not closed. Discuss in what sense the nonconvexity in Y is large. Argue that, whatever the number of consumers, there is no useful sense in which an equilibrium (nearly) exists. 17.1.3" There arc three commodities: the first is a high-quality good. the second is a low·quality good, and the third is labor. The first and second goods can be produced from labor according to the production functions f,(v) = Min {v. I} and fiv) = Min {v', I} for 0 < /1 < I. The economy has one unit of labor in the aggregate. Labor has no utility value. There are two equally sizcd classes of agents, with a very large number of each. "Rich" and "poor" have identical endowments, but the rich own all the shares in the firms of the economy. The rich spend all their wealth on the high-quality good; the poor must buy either one quality or the other-they cannot buy both. The utility function of the poor is U(XI' x,) = x, + lx" defined for (x,. x,) not both positive. (a) Which standard hypothesis of the general model docs this economy fail to satisfy? (h) Show that there can be no equilibria other than one in which both qualities of product arc produced. (e) Show that an equilibrium exists. 17.AA.IA Consider an exchange economy in which the preferences of consumers are monotone, strictly convex, and represented by the utility functions (u,(·) •... , u,(·». Show that for any (-" ......,) »0 there can be at most one Pareto optimal allocation x = (x" ... , x,) such that (III (x;) • ... , u,(x,)) is proportional to (5, •. . .• 5,). 17.AA.2" Consider the welfare-theoretic approach to the equilibrium equations described in Appendix A (the Negishi approach). The existence of a solution to the system of equations q(s) = 0 defined there follows from a fixed-point argument similar to the one carried out in Proposition 17.C.2. Assume that you are in an exchange economy with continuous, strictly convex and strongly monotone preferences, and that w, »0 for every i. Assume also that yes) turns out to be a function rather than a correspondence (a sufficient condition for this is that preferences be representable by differentiable utility functions and that at every Pareto optimal allocation at least one consumer gets a strictly positive consumption of every good). (a) Show that yes) is continuous. (b) Show that yes) satisfies a sort of Walras' law:
"L' y,(s) = 0, for every 5."
(e) Show that if s, = 0 then g,(s) > O. [Hint: If 5, = 0 then u,(x,(s» = 0 and so pCs)' x,es) = 0.] (d) Complete the existence proof. (Note that g(s) is also defined for 5 with zero components. This makes mailers simpler.)
(e) Consider the simplest price and quantity dynamics in the limit case where there are constant returns and excess demand is also a constant function. Draw the phase diagram. Suppose now that the quantity dynamics is modified by making the quantity responses depend not only on price and cost but also on the "expectation of sales. that is. on the excess demand. Will this have a stabilizing or a destabilizing effect?
17.AA.3" Suppose that. in an exchange economy, consumption sets are R~ and preferences are representable by concave. increasing utility functions u,(·). Let f1 = p. E R~: L, i., = I} be a simplex of utility weights. Suggest an equation system for Walrasian equilibrium that proceeds by associating with every i. a linear social welfare function.
17.H.3A For L = 3 draw an example similar to Figure 17.H.2 but in which there is a single equilibrium that. moreover. is locally totally unstable. Could you make it a saddle?
17_BB.1A Give a graphical example (for L = 2) of a Walrasian quasiequilibrium with strictly positive prices that is not an equilibrium for an economy in which:
fl
650
CHAPTER
17:
THE
POSITIVE
THEORY
OF
EQUILIBRIUM
-------------------------------------------------------------------------------------------------(i) (ii) (iii) (iv)
For For For For
every j, Ij = -R~. every i, XI is nonempty, closed, convex and satisfies XI + R~ C XI' every i, preferences arc continuous, convex, and strongly monotone. every i, WI e XI'
Why does this example not contradict any result given in the text (see the small-type discussion after the proof of Proposition 17.BB.I)? 17.BB.2" Consider an economy in which every consumer desires only a subset of goods and has holdings of only some goods. For the commodities desired, however, the preferences of the consumer are strongly monotone (they' are also continuous) on the corresponding nonnegative orthant. Suppose in addition that Li WI » 0 and that the economy satisfies the following jlldemmposabililY condition: It is not possible to divide consumers into two (nonempty) groups so that the consumers of one of the groups do not desire any of the commodities owned by the consumers of the other group.
Show then that any Walrasian quasiequilibrium is an equilibrium. 17.BB.3c Consider an Edgeworth box where preferences are continuous, strictly convex and locally nonsatiated (but not necessarily monotone). Suppose also that frcc disposal of commodities is not possible. Argue that, nonetheless, the offer curves must cross and, therefore, Ihat an eq uilibrium exists. Show that at equilibrium the two prices cannot be negative. In fact, at least one price must be positive (this is harder to show). 17.BB.4A Prove that if (x', y', p) is a free-disposal quasiequilibrium and Y, satisfies free disposal, then we can get a true quasiequilibrium by changing only the production of firm I. 17.BB.5 A Provide the missing step in the proof of Lemma 17.BB.5 (that is, show that the convexity of preferences implies that XI >-1 X .. and xj >-,x .. cannot both occur for Xi.
=
IXXI
+ (I
-
IX)X;).
17.BB_6 A Complete the proof of Lemma 17.BB.S by verifying the convexity of YJ(x, y, p) and of p(", y, pl. 17.BB.7A Complete the proof of Lemma 17.BB.6 by verifying the upper hemicontinuity of the correspondences jiJ(') and ;;(.). 17.BB.S" [Existence with production externalities; see Chipman (1970) for more on this topic.] There are L goods. Good L is labor and it is the single factor of production. Consumers have consumption set R~, continuous, strongly monotone, and strictly convex preferences, and endowments only of labor. Good t = I, ... , L - I is produced in sector t, which is composed of )1 identical firms. The production function of a firm in sector ( is it("t) = IXt"~' for 0 < PI :s; I. The peculiarity of the model is that the productivity coefficient IXt will not be a constant but will depend on the aggregate use of labor in sector t. Precisely, IXt
=
Yt(t "n)",
Yt> 0 and Pt
~ O.
(a) Define the notion of Walrasian equilibrium. Assume in doing so that individual firms neglect the effect on IXt of their use of labor. To save on notation, suppose also that profit shares are equal across consumers. (b) Prove the existence of a Walrasian equilibrium for the current model (make the standard additional assumptions that you find necessary). [Him: The general proof of Appendix B needs very few adaptations.]
EX ERe I S E S
651
-------------------------------------------------------------------------------------------(c) Derive and represent the aggregate production set of each sector. Which conditions on the parameters (Jt, Yt, Pt guarantee that the aggregate production set of sector t exhibits increasing. constant, or decreasing returns to scale? (d) Note that the existence conditions of (b) may be satisfied while the aggregate production set is not convex. What would happen if the externality of sector ( were internalized by putting all the firms of the sector under joint management? (e) Suppose that L = 2, (JI = 1 and individual preferences are quasilinear in labor; that is, they admit a utility function ",(x,,) + Xli' Discuss, both analytically and graphically, the bias of the equilibrium level of production relative to the social optimum.
17.88.9" Carry out the existence argument for the two-player-game approach described at the end of Appendix B.
C
Some Foundations for
HAP
T
E
R
18
SECTION
18.A Introduction
AND
Definition 18.B.1: A coalition Sc:I improves upon, or blocks, the feasible allocation x· = (xf, ... , xn e R\I if for every i E S we can find a consumption Xi ~ 0 with the properties: (i)
Xi">-i Xi
for every i e S.
(ii) Los Xi e Y + {LoS W;}. Definition \S.B.\ says that a coalition S can improve upon a feasible allocation
x' if there is some way that, by using only their endowments L i d WI and the publicly available technology Y, the coalition can produce an aggregate commodity bundle that can then be distributed to the members of S so as to make each of them better off. I. The constant returns assumption is important. With general production sels the difficulty is Ihat we cannot avoid being explicit aboul ownership shares. However, these have been defined to be profll shares, which makes our conceptual apparatus dependent on the very notion of prices whose emergence we are currently trying to explain. Thus we stick here to the case of constant returns. This is not a serious restriction: recall from Section S.B (Proposition S.B.2) that it is always possible to reduce generailechnologies to the constant returns case by reinterpreting the ownership shares as endowments of an additional "managerial" input.
18.B Core and Equilibria The theory to be reviewed in this section was proposed by Edgeworth (IS81). His aim was to explain how the presence of many interacting competitors would lead to 652
CORE
the emergence of a system of prices taken as given by economic agents, and consequently to a Walrasian equilibrium outcome. Edgeworth's work had no immediate impact. The modern versions of his theory follow the rediscovery of his solution concept (known now as the core) in the theory of cooperative games. Appendix A contains a brief introduction to the theory of cooperative games; this section, however, is self-contained. For further, and very accessible, reading on the material of this section, we refer to Hildenbrand and Kirman (19SS). The theory of the core is distinguished by its parsimony. Its conceptual apparatus does not appeal to any specific trading mechanism nor does it assume any particular institutional setup. Informally, the notion of competition that the theory explores is one in which traders are well informed of the characteristics (endowments and preferences) of other traders, and in which the members of any group of traders can bind themselves to any mutually advantageous agreement. The simplest example is a buyer and a seller exchanging a good for money, but we can also have more complex arrangements involving many individuals and goods. Formally, we consider an economy with I consumers. Every consumer i has consumption set R\, and endowment vector WI ~ 0, and a continuous, strictly convex, strongly monotone preference relation ;::" There is also a publicly available constant returns convex technology Y c: RL.' For example, we could have Y = - R\, that is, a pure exchange economy. All of these assumptions are maintained for the rest of the section. As usual, we say that an allocation x = (x" ... , XI) E R\I is feasible if:L XI = Y + LIW; for some ye Y. With a slight abuse of notation, we let the symbol I stand for both the number of consumers and the set of consumers. Any nonempty subset of consumers Sc:I is then called a coalition. Central to the concept of the core is the identification of circumstances under which a coalition of consumers can reach an agreement that makes every member of the coalition better off. Definition IS.B.! provides a formal statement of these circumstances.
Competitive Equilibria
Up to this point of Part IV, the existence of markets in which prices arc quoted and taken as given by economic agents has been assumed. In this chapter, we discuss four topics that, in essence, have two features in common: The first is that they all try to single out and characterize the Walrasian allocations from considerations more basic than those stated in its definition. The second is that they all emphasize the role of a large number of traders in accomplishing this task. In Section IS.B we introduce the concept of the core, which can be viewed as embodying a notion of unrestricted competition. We then present the important core equivalence theorem. Section IS.C examines a more restricted concept of competition: that taking place through well·specified trading mechanisms. The analysis of this section amounts to a reexamination in the general equilibrium context of the models of noncooperative competition that were presented in Section 12.F. The motivation of the remaining two sections is more normative. In Section IS.D we show how informational limitations on the part of a policy authority (constrained to use policy tools relying on self-selection, or envy freeness) may make the Walrasian allocations the only implementable Pareto optimal allocations. In Section IS.E the objective is to characterize the Walrasian allocations, among the Pareto optimal ones, in terms of their distributional properties. In particular, we ask to what extent it can be asserted that at the Walrasian allocation everyone gets her "marginal contribution" to the collective economic well-being of society. A number of the ideas of this chapter (especially those related to the core, but also some in Section IS.E) have come to economics from the cooperative theory of games. This therefore seems a good place to present a brief introduction to this theory; we do it in Appendix A.
II.B:
EOUILIBRIA
653
--------------------------------------------------------------------------------------------
1
654
CHAPTER
18:
SOME
FOUNDATIONS
FOR
COMPETITIVE
SECTION
EQUILIBRIA
II.B:
CORE
AND
EQUILIBRIA
655
------------------------------------------------- ------------------------------------------------x"
~-+--------------~------------~~
Flgur. ".8.2 ~:---+-2:;,
Figure 18.B.l
~2
Definition 18.B.2: We say that the feasible allocation x' = (x! ..... xi) E R~I has the core property if there is no coalition 01 consumers ScI that can improve upon x*. The core is the set 01 allocations that have the core property. We can see in the Edgeworth box of Figure 18.B.1 that for the case of two consumers the core coincides with the contracl curve. With two consumers there are only three possible coalitions: {I, 2}, {I}, and {2}. Any allocation that is not a Pareto optimum will be blocked by coalition {I, 2}.2 Any allocation in the Pareto set that is not in the contract curve will be blocked by either {I} or {2}. With more than two consumers there are other potential blocking coalitions, but the fact that the coalition of the whole is always one of them means that all allocations in lite core are ParelO optimal. We also observe in Figure 18.B.I that the Walrasian equilibrium allocations, which belong to the contract curve, have the core property. Proposition 18.B.1 tells us that this is true with complete generality. The proposition amounts to an extension of the first welfare theorem. Indeed, in the current terminology, the first welfare theorem simply says that a Walrasian equilibrium cannot be blocked by the coalition of the whole.' The following result, Proposition 18.B.I, shows that it also cannot be blocked by any other coalition. Proposition 18.B.1: Any Walrasian equilibrium allocation has the core property. Proof: We simply duplicate the proof of the first welfare theorem (Proposition J6.C.I). We present it for the exchange case. See Exercise 18.B.1 for the case of a general constant returns technology. Let x' = (x!, ... , be a Walrasian allocation with corresponding equilibrium
xn
2. With continuity and strong monotonicity of preferences. if a feasible allocation is Pareto dominated. then it is Pareto dominated by • reasible allocation that strictly improves the utility of ('t'('ry consumer. To accomplish this we simply transfer a very sman amount of any good from the consumer that is made better off to every other consumer. If the amount transrerred is sufficiently small then. by the continuity or prererences, the transrerring consumer is still beller olT. while. by strong monotonicity. every other consumer is made strictly better otT. 3. Keep in mind the point made in rootnote 2.
O,~--------------------------~~r_~
The core equals the contract curve in the two-consumer casco
price vector p ~ O. Consider an arbitrary coalition ScI and suppose that the consumptions (x.} •• s are such that >-. for every i E S. Then p'X, > p·w. for every i E S and therefore P·(L •• s x.) > P·(L •• s w;). But then L •• s Xi :s: L.d w, cannot hold and so condition (ii) of Definition 18.B.1 is not satisfied (recall that we are in the pure exchange case). Hence coalition S cannot block the allocation x' . •
x. x:
The converse of Proposition 18.B.1 is, of course, not true. In the two-consumer economy or Figure 18.B.1 every allocation in the contract curve is in the core, but only one is a Walrasian allocation. The core equivalence theorem, of which we will soon give a version, argues that the converse does hold (approximately) if consumers are numerous. Quite remarkably, it turns out that as we increase the size of the economy the non-Walrasian allocations gradually drop from the core until, in the limit, only the Walrasian allocations are left. The basic intuition for this result can perhaps be grasped by examining the Edgeworth box in Figure 18.B.2. Take an alloca tion such as x where consumer I receives a very desirable consumption within the contract curve. Consumer 2 cannot do anything about this: She could not end up better by going alone. But suppose now that the preferences and endowments in the figure represent not individual consumers but types of consumers and that the economy is actually composed of four consumers, two of each type. Consider again the allocation x, interpreted now as a symmetric allocation, that is, with each consumer of type 1 receiving x I and each consumer of type 2 receiving x 2 • Then matters are quite different because a new possibility arises: The two members of type 2 can form a coalition with one member of type I. In Figure 18.B.2, we see that the allocation x can indeed be blocked by giving x; to the one consumer of type 1 in the coalition and x; to the two consumers of type 2 [note that - 2(x; - W2) = (x; - W I )].4 4. Observe that all this has the flavor or Bertrand competition, as reviewed in Section 12C. Indeed. we can look at what happens with this three-member coalition as the rollowing: One or the consumers or type 1 bids away the transactions or the COnsumers or type 2 with the other consumer or type I. Although this is a topic we shall not get into, we remark that, in fact, there are strong parallels between Bertrand price competition and core competition. Note, in particular. that core competition is as shortsighted as Bertrand competition. By undercutting the other consumer or her type. the consumer or type I is only initiating a process or blocking and counterblocking (mutual underbidding in the Bertrand selling) that eventually leads to a result (perhaps th. Walrasian allocation) where she will be worse off than at the initial position.
An allocation in the contract curve that can be blocked with two replicas.
656
CHAPTER
18:
SOME
FOUNDATIONS
FOR
COMPETITIVE
EQUILIBRIA
--------------------------------------------------------------------------------------------The ability to do this depends. of course. on the way we have drawn the indifference curves. Nonetheless. as we will see. we are always able to form a blocking coalition of this sort if we have sufficiently many consumers of each type. The version of the core equivalence theorem that we will present is in essence the original of Edgeworth. as generalized by Debreu and Scarf (1963). It builds on the intuition we have just discussed. To begin, let the set H = {I, ... , H} stand for a set of types of consumers, with each type 10 having preferences ;::, and endowments w,. For every integer n > 0, we then define the N-repliCll economy as an economy composed of N consumers of each type. for a total number of consumers IN = N H. We refer to the allocations in which consumers of the same type get the same consumption bundles as eqlwl-treatment "l1ocClliolls. Proposition 18.B.2 shows that any allocation in the core must be an equal-treatment allocation. (We hasten to add that this is true for the current replica structure, where there arc equal numbers of consumers of each type. It does not hold in general; see Exercise IS.B.2.) Proposition 18.B.2: Denoting by hn the nth individual of type h, suppose that the allocation x· = (XT1' ... • xrnl' .. • xtN" .. . xj." • ... xl.tn • ... . X;"N) E R~HN I
belongs to the core of the N-replica economy. Then x' has the equal-treatment property. that is, all consumers of the same type get the same consumption bundle: for all 1 :!> m. n:!> Nand 1 :!> h:!> H. Proof: Suppose that the feasible allocation x = (X" •...• X"N) E R/;."N does not have the equal-treatment property because. say. x, .. 1- x .. for some III 1- n. We show that x does not have the core property. In particular, we claim that x can be improved upon by any coalition of H members formed by choosing from every type a worst-treated individual among the consumers of that type. Suppose without loss of generality that, for every h. consumer h I is one such worse-off individual. that is. x" for all hand n. Define now the average consumption for each type: .", = (liN) L. x,•. By the strict convexity of preferences we have (recall that consumers of type I are not treated identically)
x,. ;::.
.x, ;::, x"
for all h
and
X,>-, X " ,
(IS.B.I)
We claim that the coalition S={II •...• hl, ...• HI}. formed by H members, can attain by itself the consumptions (x""" .x,,) E R~H. Therefore. by (IS.B.l). the original nonequal-treatment allocation can be blocked by S.' To check the feasibility of (.x I' •.•• XII) E R/;." for S, note that, because of the feasibility of x = (x", ...• x/IN) E 1R/+"N. there is Y E Y such that L. Ln x,. = y + N(Lh w,), and therefore
5. Recall that preferences are strongly monotone and continuous, so that if S can achieve an allocation that does strictly better than x· for some of its members. and at least as well as x· for all of them, then it can also achieve an allocation that does strictly better for all of its members.
S l ~ ; I 0 H
1
a . 8:
C 0 A E
AND
E
u U I LIB
RI A
657
-------------------------------------------------------------------------------------------But by the constant returns assumption on Y. (I/N)y E Yand so we conclude that (x l' ...• XII) E R';" is feasible for coalition S. • Proposition IS.8.2 allows us to regard the core allocations as vectors of fixed size LH. irrespective of the replica that we are concerned with. As a matter of terminology, we call a vector (x" ...• XII) E R~" a Iype allocation and, for any replica N. interpret it as the equal-treatment allocation to consumers where each consumer of type h gets x,. A type allocation (x l ' . . . , XII) E IR~" is feasible if L, x. = Y + L. W, for some )' E Y. Note that for any replica N the corresponding equal-treatment allocation is feasible because
and Ny E Y by the constant returns assumption on Y. By Proposition IS.8.2 the core allocations of a replica economy can be viewed as feasible type allocations. Define by eN C R~H the set of feasible type allocations for which the equal-treatment allocations induced in the N-replica have the core property. Note that eN docs depend on N. Nonetheless, we always have eN. I C eN because a type allocation blocked in the N-replica will be blocked also in the (N + I )-replica by a coalition having exactly the same composition as the one that blocked in the N-replica. Thus. as a subset of RLII the core can only get smaller when N - 00. At the same time, we know from Proposition IS.B.I that the core cannot vanish because the Walrasian equilibrium allocations belong to eN for all N. More precisely, the set of Walrasian type allocations is independent of N (see Exercise IS.8.3) and contained in all eN' The core equivalence theorem (which. in the current replica context, is the formal term for the combination of Propositions 18.B.I, IS.8.2 and the forthcoming Proposition IS.B.3) asserts that the Walrasian equilibrium allocations are the only surviving allocations in the core when N - 00.
Ri
Proposition 18.B.3: If the feasible type allocation x· = (x~, ...• xii) E H has the core property for all N = 1,2, ... , that is, x· E eN for all N. then x· is a Walrasian equilibrium allocation. Proof: To make the proof as intuitive as possible we restrict ourselves to a special case: a pure exchange economy in which. for every h. ;::, admits a continuously differentiable utility representation u,(') [with Vu,(x,)>> 0 for all x,]. In addition. the initial endowments vector w, is preferred to any consumption x, that is not strictly positive. This guarantees that any core allocation is interior. We emphasize that these simplifying assumptions are not required for the validity of the result. Suppose that x = (X, •... ' XII)E RLII is a feasible type allocation that is not a Walrasian equilibrium allocation. Our aim is to show that if N is large enough then x can be blocked. We may as well assume that X is Pareto optimal (otherwise the coalition of the whole blocks and we are done) and that x, » 0 (otherwise a consumer of type /0 alone could block). Because of Pareto optimality we can apply the second welfare theorem (Proposition 16.D.I) and conclude that X is a price equilibrium with transfers with respect to some p = (P,' ...• pLl.lf X is not Walrasian then there must be some 10, say /0 = I. with P'(X, - w,) > O. Informally. type 1 receives a positive net transfer from the rest of the economy and is thus relatively favored (interpretatively. think
658
CHAPTER
1':
SOME
FOUNDATIONS
FOR
COMPETITIVE
EQUILIBRIA
SECT,ON
---------------------------------------------------------------------------------------~-
of type I as the most favored). We shall show that, as long as N is large enough, it would pay for the members of all the other types in the economy to form a coalition with N - I consumers of type I (i.e., to throw out one consumer of type I). More precisely, if a member of type I is eliminated then to attain feasibility the rest of the economy must absorb her net trade x, That, of course, presents no difficulty for the positive entries (those commodities for which the rest of the economy is a net contributor to this consumer of type I), but it is not so simple for the negative ones (the commodities where the rest of the economy is the net beneficiary). The most straightforward methodology is to simply distribute the gains and losses equally. In summary, our coalition is formed by (N - I) + N(H - I) members and, for every type I,. every member of type h gets
+ N(H
AND
We saw in Proposition 18.B.I that the half of the core equivalence theorem that asserts that Walrasian allocations have the core property generalizes the first welfare theorem. In its essence. the half asserting that. provided the economy is large, core allocations are Walrasian constitutes a version of the second welfare theorem. To understand this it may be useful to go back to the general (non replica) setup and formulate the property of a core allocation being Walrasian in terms of the existence of a price support for a certain set. For simplicity. we restrict ourselves to the pure exchange case. Given a core allocation x = (xt, ... , 4) e R~' then. in analogy with the construction used in the proof of the second welfare theorem (Proposition 16.0.1) we can define the setsy
I
_ I)
CORE
and, therefore, will also be individually favorable (recall Section 3.1 for similar arguments)" •
W,.
x; = x, + (N
1I.B:
V. =
_ I) (x, - w,).
{XI: XI >-IXrj U {WI} c: RL
Note that (N - I)x',
+ Nx~ + ... + Nx;' =
+ Nx, + ... + NXII + (x, x, + x, - W, I)w, + Nw, + ... + Nw, .
(N - I)x,
- w,)
We have L, WI E V. But there is more: Ihe CO" property for x· implies Ihal LI WI belongs Ihe houlldary of V. To see this, note that ifL' w, is in the interior of Vthen there is: e V such that:« L, w,; that is, there is x' = (xi •.. ·, X,) with x; E V. for every i and LI x; LI w,. Hence. x' is feasible, x' ". (w, •. .. ,w,), and, for every i, either x; >-1 xr or x; = WI' It follows that the set of consumers S = Ii: x;". W,} is nonempty. that x; >-,xr for every i E S, and that
= Nw, + ... + NWII = (N -
10
Hence, the proposed consumptions are feasible for the proposed coalition. Note also that the consumptions are nonnegative if N is large enough. For every h. every consumer of type h in the coalition moves from .x, to x;. Is this an improvement or a loss"! The answer is that if N is large enough then it is an unambiguous gain. To sec this. observe that p'(x, - w,) > 0 implies Vu,(x.)·(x, - w,) > 0 for every h because p and Vu,(x.) are proportional. As we can then see in Figure 18.B.3 (or. analytically, from Taylor's formula; see Exercise 18.B.4) there is Ii> 0 with the property that, for every h, u.(x. + IX(X, > u,(x,) whenever 0 < IX < Ii. Hence, for any N with (I/[(N - I) + N(H - I)]) < Ii the coalition will actually be blocking. Intuitively, we have done the following. The coalition needs to absorb x, - W,. Evaluated at the marginal shadow prices of the economy, this is a favorable "project" for the coalition since p'(x, - w,) > O. If the coalition is numerous then we can make sure that every member will have to absorb only a very small piece of the project. Hence the individual portions of the project will all be ~at the margin"
x, + (x, - "',)
I~«I~-I~=I~-I~=I~
to p'x;
change of a consumer
of type h in the blocking coalilion.
x; ~ .', +
I
(r -1)+ r(H -I)
".
(x, - w,)
ifS
hil
,_s
i.S
xr.
Figure 18.B.3
for p, > 0
I.'
Thus S is a blocking coalition. The next claim is that if P = (P""" pd ". 0 supports Vat LI w" that is, p': :2: P'(LI Wi) for all Z E V, then P must be a Walrasian price vector for x· = (xt, ... , xf). To verify this, note first that, for every i. we have xi >i xf for some xj arbitrarily close to Therefore, xi + L. .. w. E Vand so p-(x; + L •• I co,):2: P'(WI + L ... w.). Going to the limit (i.e., letting x; ~ xn this yields p'xr ~ P'W, for all i. Because LI xr s L.WI, we must therefore have p' xr = P'W, for all i.ln addition. whenever x;>-.xr we have P'(x; + L •• IW.):2: P'(w, + L •• I w.) and so p' x; :2: p·w,. If we exploit the continuity and strong monotonicity of preferences as we did in Section 16.0 (or in Appendix B of Chapter 17). we can strengthen the last conclusion
The consumption
VU.(x,) ~ ~,P
=:«
i~S
w,»
EQUIL,BR'A
659
--------------------------------------------------------------------------------------------
>
p·W;.
The key difference from the case of the second welfare theorem (studied in Section 16.0) is that V c: R" does nol need 10 be convex and that therefore a nonzero peRI;. supporting V at L, w, may not exist. The reason for the lack of convexity is that the individual sets V, c: RL need not be convex: V. is the union of the preferred set at xr, which is convex, and the initial endowment vector which will typically be outside this preferred set and therefore disconnected from it. However, if Ihe (possibly nonconvex) sets V. c: RL being added are Ilumerous. then Ihe sum L. V. c: RL is "almost" convex. Thus, the existence of (almost) supporting prices for core allocations can be seen as yet another instance of the convexifying effects of aggregation. We end by mentioning an elegant approach to core theory pioneered by Aumann (1964) and Vind (1964). It consists of looking at a model where there is an actual continuum of consumers and where we replace all the summations by integrals. The beauty of the approach is that all the approximate results then hold exactly. The core equivalence theorem, for example,
w,.
6. See Anderson (1978) for a different line of proof 'hat makes minimal assumptions on the economy.
·· ... PTER
15
SOME
FOUNDATIONS
f")R
COMPETI1IVl;,
EQUIL,BRIA
--------------------------------------------------------------------------------------------takes the form: An allocation belongs to the core if and only if it is a Walrasian equilibrium allocation.
Definition lS.C.l: The profile of actions a* = (aT, ... , ail EA, x ... equilibrium if. for every i.
<::: u;(g(a;; pea;. a!;»
Given a = (a., ... , a,) EA. x ...
X
,"-
tOUttOATluNS
OF
W"'LRA8IAN
A" define
B,(a) = (x, E R~: x, - w,::;; g(a;; p(a;, a_,ll for some a; E A,}
B(p(a), p(a)'w,) = (x, E R~: p(a)'x, ::;; p(a)'w,}
The idea of competition that underlies the theory of the core is very unstructured; there are no trading institutions and, in principle, any conceivable profitable opportunity can be taken advantage of. It is because of this, for example, that core allocations are guaranteed to be Pareto optimal. In many economic applications, however, the structure of competition is given. Trade takes place through some type of market mechanism making explicit use of prices. The set of instruments and the information available to competitors are then limited. Yet, we also expect that price taking will emerge if individual competitors are small relative to the size of the market. We have already investigated this topic in Section 12.F. We reexamine it here because there are a few general equilibrium qualifications worth taking into account. There are many models of price-mediated competition arising in applications. We will describe three of them, but before doing so, we present an abstract treatment emphasizing the main issues. Suppose there are J economic agents (abstract competitors, perhaps firms). There are also a set PeRL of possible price vectors and a set A of "market actions." Every i has a set A, c A and an endowment vector w, E RL. For every a, E A, and PEP, a trading rule defined on A x P and with values in RL , assigns a net trade vector g(a,; p) to agent i, satisfying p'g(a,; p) = O. Given an array a = (a., .. . , a/ ) of actions, there is then a market clearing process that generates a price vector p(a) E P. We also assume that every i has a utility function u,(g(a,; p) + w,), thus indirectly defined on A, x P. The previous setup suggests treating the problem by the methodology of noncooperative games, as presented in Chapter 8.
+ w;)
~:j'EAA
as the effective budget set of trader i at a. In words, B,(a) is the set of net trades that trader i can achieve through some choice of a" given that the remaining traders are choosing a_,. This set of achievable net trades will be close to the Walrasian budget
18.C Noncooperative Foundations of Walrasian Equilibria
u;(g(ai; p(a·))
-
s·
+ w;)
X
A, is a trading
for all a; E A;. 7
The concept of noncooperative equilibrium incorporated in Definition 18.C.1 is the same as the one we used in Chapters 8 and 12. As there, and in contrast to the analysis of the core in Section 18.B, such equilibria need not be Pareto optimal. The question we now pose ourselves is: Under what conditions is it the case that, if individual traders are small relative to the size of the economy, the system of markets approximates a price-taking environment in which, effectively, every trader optimizes given a competitive budget set (and in which, therefore, the equilibria will be nearly Pareto optimal).
7. As it has become customary. we follow the notation (a j • a!./)
= (aT •...• a~_I' a
j •
a:+ \ •...• aT).
if the following two types of conditions both hold: (I) Insensitivity of prices to own actions. For the boundary of B,(a) to be (almost) contained in a hyperplane, we need the price-clearing function p(a" a _,) to be very insensitive to a,." Often this will be guaranteed if the economy is large and, consequently, every competitor is of small relative size. Suppose, for example, that pea) has the form p«l/r) L, a,) where r is a size parameter (perhaps the number of consumers). It is actually quite common that the problem be given as. or can be transformed into, one in which actions affect price only through some average. At any rate, if this is the case. then the essential fact is clear: As long as p(.) depends
continuously 011 the at'eraye action (1/r) L, a, the depelldence of prices on individual actiolls (assume the A, are bounded) lI'iII become negligible as the economy becomes IlIry<,; that is, as r -+ (fJ. Continuity of p(.) is therefore a key property. (2) IlItiivitilllll spallllillg. Even if pea) is practically independent of the actions of an individual. we could still have a failure of individual spanning. That is, the boundary of the set B,(a), while flat, may be "too short," as in Figure IS.C.l(a), or even lower-dimensional, as in Figure 18.C.I(b) where it reduces to the initial endowment vector (no trade at all is possible). Individual spanning will have to be checked in every case. Verifying it will typically involve showing, first, that g(a" p) is sensitive enough to a, and, second, that Ai is large enough. We now briefly, and informally, discuss three examples illustrating these ideas."
Example IS.C.I: General Equilibrium, Single-Good Cournot Competition. This is in essence the same model studied in Section 12.C, except that we now admit a completely general form of "inverse demand function," that is, of the correspondence that assigns market-clearing prices to aggregate production decisions. This is meant to reflect the possibility of wealth effects (a hallmark of the general equilibrium approach). To be concrete, suppose that we have two goods: the first is a consumption good and the second is "money" (which is also the unit of account with price equal to I). There are r identical consumers, each endowed with a unit of money. For a price pER of the consumption good the demand of a consumer for this good is x(p) E R There are also J firms producing the consumption good out of money. Firms set quantities. Marginal cost, up to a unit of capacity, is zero. To minimize complexity, 8. Recall that p{a;. a _,) 'g(a;; p(a;, a_,ll = 0 for all a;. Therefore. if p{a;. a_i) is (almost) independent of a; then Bi(a) is (almost) contained in the hyperplane perpendicular to p{a). 9. For more on general equilibrium Cournotian models in the style of Examples IS.C.t and IS.C.2. see Gabszewicl and Vial (1972) and Novshek and Sonnenschein (1978). For a survey of .he general area, sec Mas-Colcll (1982).
EQUILIBRIA
601
662
CHAPTER
18:
SOME
FOUNDATIONS
FOR
COMPETITIVE
EQUILIBRIA
~~~~--~----------------------------------------------------------------------
----
SECTION
I •• C:
NONCOOPERATIVE
FOUNDATIONS
OF
WALRAIIAN
EQUILIBRIA
Price
x"
p(q/,)
---~
---------, W, I
I
B,(a)
I
Bj(a)
:
I I
I
I
(a)
!
(b)
Leisure
Flgur.'B.C.'
Flgur. IB.C.2
assume that the owners of the firms are a separate group of agents interested only in the consumption of money. Therefore, for any total production q = Lj q} and size parameter r, the market price of the consumption good must solve the general equilibrium system rx(p) = q, or x(p) = qlr. We suppose that for any qlr the market selects a solution p(qlr) to this equation. '0 In the quasilinear, partial equilibrium context of Chapter 12, x(·) is a decreasing continuous function and therefore its inverse p(.) exists and is conrinuous (and decreasing). It then follows that, when r is large, P(L} qjlr) is quite insensitive to the decision of any particular firm. Hence, firms are almost price takers and, as a consequence, the Cournot equilibria are almost Walrasian. Vet in the current general equilibrium context, p(.) may be unavoidably discontinuous. This is illustrated in Figure 18.C.2, where we represent the offer curve of a consumer. In the figure, there is no way to select money demands, and therefore prices, continuously over the offer curve as consumption per capita qlr ranges from 0 to J Ir.11 The location of potential Cournot equilibria will depend on how the market selects p(') in the domain [b', b"] of consumptions per capita, but the possibility of Cournot equilibria bounded away from the Walrasian equilibrium irrespective of the size of the economy is quite real. A particular price selection p(.) has been chosen in Figure IS.C.3. Note, first, that at the Walrasian equilibria of this model we must have every firm producing at capacity (and so the Walrasian equilibrium price is pO). Yet, provided r> (J).lyb'), 10. To view this example as a particular case of the abstract trading model described above. you should think or the J firms as the players. Firm j is "endowed" with a unit or the first good and ils stralegy variable is qJ e [0, I] - AJ. Finally, Ihe Irading rule is g(qj; p, I) = (- qj. pq;). . 11. The example is contrived in that "money" is a Giffen good. If consumers were not Identical, this feature would not be required.
A price equilibrium seleclion.
o
every firm producing rb'p < I (for a consumption per capita of b') constitutes a Cournot equilibrium: because p(.) is very elastic in the domain [0, b'] it will not pay any firm to contract production; and if any firm expands production, no matter how slightly, a precipitous and unprofitable drop in prices ensues. 11 [See Roberts (1980) for more on this point.] •
Oller Curve
JI, ;' Consumplion
Flgur. la.C.3
p' !;:-------7:--..;;;:::===;t~b' b" JI, Production per Consumer
An economy with no continuous price selection.
Example IS,C.2: Cournot Competition among Complements. We modify the previous example in only two respects: (I) There are two consumption goods; (2) firms are producers of either the first or the second of these. The respective number of firms is J, and J,. To be very simple we assume that the consumer has a quasilinear utility with money as numeraire. If the concave, strictly increasing utility function for the two consumption goods is ",(x" x,) then, for any total productions (q" q,), market clearing prices are
(
) _ '>'o/I( I / ) - (a",(q,lr, q,lr) a",(q,/r, q,lr»)» 0
pq"q,-
q,r,q,r-
ax,
'ax,
.
The Walrasian equilibrium productions are (JI' J,). Suppose, now, to take an extreme situation, that the two consumption goods are complements in the sense that the consumption of one is absolutely necessary for the enjoyment of the other, or "'(0, x,) = ",(x" 0) = 0 for any XI and x,. Then we claim that lack of activity (i.e., the null production of the two goods) is an equilibrium. The reason is clear enough; If q, = 0 then any positive supply ql > 0 of good one can only be absorbed by the market at p, = V, "'(q ,/r, 0) = O. Thus, no firm has an incentive to produce any amount of good I (and similarly for good 2). Economically, the difficulty is that the cooperation of at least two firms is needed to activate a market. Technically, we have a failure of continuity of clearing prices at (0,0) since p(e, 0) = 0 for all e > 0 but the limit of pte, e) as e goes to zero remains bounded away from zero.'3 • Example IS.C.3: Tradillg Poses. This example belongs to a family proposed by Shapley and Shubik (1977). It is not particularly realistic but it has at least three 12. When every firm produces ,b'/J,the profils or one firm are ,b'/)). > I/y. Bul I/y is an upper bound ror Ihe profils or any firm Ihal deviates rrom Ihe suggeSled produclion by producing more. Hence an output level or rb'll ror every firm constitutes an equilibrium. 13. The complemenlarily makes it impossible ror .p to be continuously dillerenliable al Ihe origin. Thererore, p(' ) rails 10 be continuous. This is Ihe crucial aspecl ror Ihe example. NOle Ihal discontinuity at the origin is a natural occurrence: it will arise. ror example. whenever the indifference map or.p(·) is homothetic (bul not linear). See Harl (1980) ror more on Ihis issue.
663
ob~
... HAPTER
lei:
SOME
..
uv .... OAYlu ...
t-Gft
COMPETITIVE
fQUILIBRIA
S E to T I 0
~
\
a . LI.
T 11 E l l Mil &
l O R E 0 1ST R I 8 UTI 0 N
ootl
-------------------------------------------------- -------------------------------------------------------------------------.>
Figure lB.C.4
o
An effective budget set for the trading post Example IS.C.3.
x"
virtues: it constitutes a complete general equilibrium model. all of the participants interact strategically (in the two previous examples. consumers adjust passively). and it is analytically simple to manipulate. There are L goods and J consumers. Consumer i has endowment w, € R~. The Lth commodity. to be called "money." is treated asymmetrically. For each of the first L - I goods there is a trading post exchanging money for the good. At each trading post ( :5 L - I. each consumer i can place nonnegative bids a" = (ai" a;,) € IR~. The interpretation is that an amount ai i of good I is placed at the offer side of the trading post to be exchanged for money. Similarly. an amount a; i of money is placed in the demand side to be exchanged for good I. Accordingly. the bids are also constrained by aii :5 Wt; and LIS L-' (Iii :5 WLi' Given the bids of consumer i in the trading posts (:5 L - I and prices (p, •...• PL-,.I) the mechanism is completed by the trading rule: gt(a 1h
···,
ai i
aL-l,i; PI"'" Pl.-I' I) = -
-
, QIi
PI
for all t < L - I. The trade for the money good is derived from the budget constraint of the consumer. Given a vector a = (a" •...• aL-I."" .• a" •. ..• aL-I.,) of bids for all consumers. the clearing prices in terms of money are determined as the ratio of the amount of money offered to the amount of good offered:
ria;i
PI(a)=-~.
Lian
t
= I •...• L-I.
(IS.C.I)
Note that PI(a) is well defined and continuous except when there are no offers at the trading post I [i.e .• except when aii = 0 for all i]'" A typical effective budget set for agent i is convex and. provided that L .. i (Ii. "# 0 and L ... "# 0 for alii:::; L - I. it has an upper boundary containing no straight segments (you are asked to formally verify this in Exercise IS.C.I). This reRects the fact that as a consumer increases her bid in one side of a market the terms of trade turn against her. Figure IS.C.4 gives an illustration for the case L = 2.
a;.
t4. For the special. but important. case in which there is a single trading post (i.e .• L = 2). we can go a bit farther. When L, ail > 0 and L. a,. = O. the relative price of money is still well defined: it is zero. The essential difficulty in defining relative prices arises when L, aii = 0 and L.a,. = O.
It follows from expression (IS.C.I) that approximate price taking will prevail in any trading post that is thick in the sense that the aggregate positions taken on the two sides of the market are large relative to the size of the initial endowments of any consumer. A necessary condition for thickness is that there be many consumers. But this is not sufficient: it is possible even in a large economy to have equilibrium where some market is thin and. as a consequence. a trading equilibrium may be far from a Walrasian equilibrium. In fact. any trading equilibrium for a model where a trading post I is closed (i.e .• the trading post does not exist) will remain an equilibrium if the trading post is open but stays inactive. That is. if we put a" = (a;'" ail) = 0 for all i. Economically. this is related to Example 18.C.2: it takes at least two agents (here a buyer and a seller) to activate a market. Mathematically. the difficulty is again the impossibility of assigning prices continuously when at; = 0 for all i. Up to now. in this and previous examples. all of the instances of trading equilibria not approaching a Walrasian outcome when individual competitors are small have been related to failures of continuity of market equilibrium prices. But the current example also lends itself to illustration of the individual spanning problem. Indeed. even if markets are thick and therefore prices. from the individual point of view. are almost fixed. it remains true that the trading post structure imposes the restriction that goods can only he exchanged for money on hand (in macroeconomics this restriction is called the cash-in-advance. or the Clower. constraint). Money obtained by selling goods cannot be applied to buy goods. Therefore. for a given individual the Walrasian budget set will be (almost) attainable only if the initial endowments of money are sufficient. that is. only if at the solution of the individual optimization problem the constraint Lf " L-' WL' is not binding. But there is no general reason why this should be so. Suppose. to take an extreme case. that WL' = O. Then consumer i simply cannot buy goods at all. _
ar, : :;
18.D The Limits to Redistribution In Section 16.0 we saw that. under appropriate convexity conditions and provided that wealth can be transferred in a lump-sum manner. Pareto optimal allocations can be supported by means of prices. However, as we also pointed out there. a necessary condition for lump-sum payments to be possible is the ability of the policy authority to tell who is who-that is. to be able to precisely identify the characteristics (preferences and endowments) of every consumer in the economy. In this section, we shall explore the implications of assuming that this cannot be done to any extent; that is. we shall postulate that individual characteristics are private and become public only if revealed by economic agents through their choices. We will then see that under very general conditions the second welfare theorem fails dramatically: the only Pareto optimal allocations that can be supported involve no transfers. that is. they are precisely the Walrasian allocations. Thus. if no personal information of any sort is available to the policy authority. then there may be a real conAict between equity and efficiency: if transfers have to be implemented we must give up Pareto optimality. The nature of this trade-off is further explored in Sections 22.B and 22.C. We place ourselves in an exchange economy with J consumers. Each consumer i has the consumption set R~, the endowment vector w, ~ O. and the continuous. monotone. and strictly quasiconcave utility function u,(·).
666
CHAPTER
11:
SOME
FOUNDATIONS
FOR
COMPETITIVE
EQUILIBRIA
SECTION
-------------------------------------------------------------------------------------Good 2
Good 2
11.0:
- '"')' w,
X~?:IX;.
o
(a)
(b)
LIMITS
TO
REDIITRI8UTION
Good I B
Figure 18.0.1
+ (x:
- w,))
xi ?:l x;
o,L-----------------------+-~--
x"
We begin by stating a restriction on feasible allocations designed to capture the possibility that the allocation is the result of a process in which every consumer maximizes utility subject to market opportunities that are identical across consumers. Definition 18.0.1: The feasible allocation x' = (xf, ... , x7) E RLI if self-selective (or anonymous, or envy-free in net trades) if there is a set of nel trades B c: RL , to be called a generalized budget set or a tax system, such that, for ellery i, z; = xi - Wi soilles the problem Max
Ui(Zi
661
1=3
' •.,' = (w, + (x; Good I
THE
--------------------------------------------------------------------------------------
Another representation of the example or Figure \8.0.2.
+ Wi)
li+wi~O.
15. The concept or a nonenvy allocation was introduced by Foley (1967) and that or a nonenvy nct·trade allocation by Schmeidler and Vind (1972). See Thomson and Varian (1985) ror a survey or these notions (with an emphasis on the ethical aspects).
A Pareto optimal, self-selective allocation that is not Walrasian.
Figure 1'.0.3
S.t. liEB,
Figures 18.D.I(a) and IS.O.I(b) present two examples of self-selective allocations. " In the figures the preferences and endowments of the different consumers are depicted in the same orthant. Note that if x' = (xt, ... ,xl) is self-selective then it is enough to take B = {xf, ... , xn. Thus, we could read Definition \S.O.l as saying that no consumer i envies the net trade of any other individual; among all the net trades present in the economy, the consumer is happy enough with the one assigned to her. Expositionally, we have preferred to keep separate the reality of B because we have in mind the limit situation in which there is, on the one hand, a multitude of consumers whose actions are individually imperceptible and, on the other, a policy authority that has perfect statistical information (i.e., knows perfectly the distribution of individual characteristics), but no information at all on who is who. In such a world, a viable policy instrument is to choose a set B and let each consumer select her most preferred point in it. Because this is what an income tax schedule amounts to, we also call B a tax system. We remark that, as a policy tool, the notion of a generalized budget presumes the ability to prevent individuals from choosing several times. Hence it models the income tax, adequately, but not a commodity tax. We now pose a question of the second welfare theorem type: Which Pareto optimal allocations can be supported by means of a common budget? That is, which feasible allocations are simultaneously Pareto optimal and self-selective? This
Flgur. 1'.0.2
question is both broader and more limited than the one associated with the second welfare theorem. It is broader because we allow supportability by general (nonlinear) budget sets and not only by linear hyperplanes. It is narrower because it demands that all consumers face the same budget for their net trades. The first observation is that if x' = (x!, . .. , xl) is a Walrasian allocation with equilibrium price vector pERI.., then the allocation is Pareto optimal because of the first welfare theorem, and self-selective because we can take B = {z: p·z = o}. The allocation marked x' in the Edgeworth box of Figure \S.0.2 provides an example of a self-selective allocation in the Pareto set that is not Walrasian. Figure IS.0.3 shows transparently why x' is self-selective. In this figure we bring the origin of the consumption sets of the two consumers to their initial endowment vector so that the preferences of the two consumers are expressed over net trades. Then with z{ = x{ - Wj, i = 1,2, we see that zt ;::1 zf and zr ;:2 z!. The previous Edgeworth box example suggests that there may be ample room for anonymous redistribution. The example, however, is special in that there are only two consumers or, more generally, in that all the consumers fall into two preferenceendowment types. We will now investigate the situation in which there is a multitude of consumers who moreover fall into a rich variety of types. Intuitively, this should make the compatibility of Pareto optimality and transfers more difficult because it is likely that the opportunities for envy will then be many (Pareto optimality will force the net trades to vary across consumers), and therefore the freedom in constructing the generalized budget will be limited. In Figure IS.O.4 we represent an allocation for an economy with two commodities and a continuum of types (and therefore with a continuum of consumers).16 16. Approximate versions of the following results exist for economies with a finite but large number or types.
CHI
TiR
18:
;:OUHOATION~
;,;,uME
rOR
.... OMPETITIVE
"l
[QUILI8i-IA
LIMIT~
TO
REDISTRIBUTION
-------------------------------------------------------------------------------------------Proposition 18.0.1: Suppose we have an exchange economy with a continuum of consumer types. Assume:
'E [0, I] f
"
<::0
""
'\,.:'\:--,--------- ->,
Figure 18.0.4
\~
Pareto optimal allocation with a
B~"
continuum of traders
that is self-selective
d
but not Walrasian.
Consumer types are indexed by t E [0, I] am/tlIeir preferences ;::, depend contilJuously For simplicity we take their endowments to be the same." An implication of this continuity assumption is that the set of characteristics (prderences-endowment pairs) of the consumers present in the economy cannot be split into two disconnected classes. The consumptions of the different types are distributed along the curvilinear segment cd. This allocation has the following properties:
{l1J I.
(i) It is not Walrasian. If it were then all the consumptions would lie in a straight line; in fact, different types end up exchanging the two goods at different ratios. (ii) It is self-selective. We see in Figure IS.D.4 that the consumption chosen by each consumer maximizes her utility in the generalized budget set B. Note that the frontier of any admissible budget set has to include the segment cd comprised by the consumptions actually chosen by some consumer. (iii) It is Pareto optimal. Indeed, the price vector p = (1,1) will 'make the allocation into a price equilibrium with transfers, hence a Pareto optimum. Observe that fact (iii) depends crucially on the indifference curves of every consumer exhibiting a kink at the assigned bundle. If we tried to smooth these kinks out then, because cd is curved and the preferences change continuously, the result would be the existence of two consumers with different marginal rates of substitution at their chosen point, which is a violation of Pareto optimality (there would be room for profitable exchange among these two consumers), Only if B were rectilinear could we retain Pareto optimality, but then the allocation would be a Walrasian equilibrium. Thus, it appears that if the indifference curves are smooth at the consumption points then a Pareto optimal, self-selective allocation can be something other than a Walrasian equilibrium only if the characteristics of the consumers present in the economy can be split into disconnected classes. With this motivation we can state Proposition IS.D.1. 18 17. 18.
Whal is important is that they change continuously with I. For results of this type, see. for example, Varian (1976) or Champsaur and Laroque
(1981).
(i) The preferences of all consumers are representable by differentiable utility functions. (ii) The set of characteristics of consumers present in the economy I. cannot be split into two disconnected classes. Formally, if (u(')' w), (u'(·), w·) are two preference-endowment pairs present in the economy then there is a continuous function (u('; I), w(l)) of IE [0, 1] such that (u(·; 0), w(O))
= (u(·), w),
(u(·; 1), w(l))
= (u,(-), w),
and (u(·; I), w(l)) is present in the economy for every I. Then any allocation x· = {x!}"' that is Pareto optimal, self-selective, and interior (i.e., xi » 0 for all i) must be a Walrasian equilibrium allocation. Here J is an infinite set of names. Proof: The proof is far from rigorous. It is also limited to L = 2. Let p = (p,. p,) be the price vector supporting x· as a Pareto optimal allocation. Because of dilTerentiability of the utility functions and interiority of the allocation, the relative prices p,/p, arc uniquely determined. We want to show that p'(": - w;) = 0 for all i. The first observation is Ihat at x· the equal-treatment property holds: if (u;(·), w;) = (Uk (.), w,:) then x~ = ."(r Indeed, neither i envies k nor k envies i. Hence and x: must lie in the same indilTerence curve of the common preference relation of i and k. By the strict convexity of preferences. the price vector p can support only one point in this indilTerence
xr
curve. Hence
."(~
and
x: must be equal.
If the set of net trades present in the economy consists of a single point, then this point has to be the vector 0 (otherwise the aggregate of the net trades could not be zero) and the result follows. Suppose, therefore, that there are at least two different net-trade vectors present in the economy, Zo and ',. In Figure IS.D.S, we represent them as well as the net trades z(t) of all consumers captured by the continuous parametrization given by assumption (ii). where 1= 0, I correspond, respectively, to consumers underlying Zo and ',. A key fact is that ,(I) is continuous as a function of I. This is intuitive. We have already seen that the equal·treatment property holds: Identical individuals are treated identically. Technicalities aside, the logic of the continuity of Z(I) is the same: If envy is to be prevented, then similar individuals must be treated similarly. Thus as we go from I = 0 to I = I, the net trades are moving continuously from '0 to z,. Hence the frontier of any generalized budget set 8 must actually connect Zo to ',. If so. then either this frontier is a straight segment with normal p between these two points [in which case p' (=0 - z,) = 0], or somewhere between them there is one point (,(I') in Figure IS.D.5) where the "slope" of the frontier of 8, and therefore the MRS of the consumer choosing this point, is dilTerent from p,/p, [in which case p would not be a supporting price vector). We conclude that the portion M of the frontier of 8 containing all the net trades present in the economy is a nontrivial straight segment with normal p (hence, a convex set). Since the net trades add up to zero, we must have 0 EO M. Thus, p'Z = p.(, - 0) = 0 for every Z EO M. In particular. P'(x!' - w;) = 0 for every i. See Figure IS.D.6. • J9. The expression "characteristics or consumers present in the economy" means technically "contained in the support or the distribution or characteristics induced by the population of consumers
669
670
C HAP T E R
1':
S0 M E
F 0 U N D Ii T ION S
FOR
COM PET I T I VEE QUI LIB R I A
SEC T • 0 N
Ftgur. 18.0.5. (teft)
(p" p,): But
Pareto Prices
p; p,
~~,, \"
o·
u.(x.) = t/I.(x,., ... , XL_' .• )
B
that is Walrasian.
+ Xu·
t/I. to be strictly concave. The initial endowment vector of type h is w. ~ O. An economy is defined by a profile (I" . .. ,III) of consumers of the different types, for a grand total of 1= :[.1•. For any economy (I" ... , III) we define the maximal amount of "social utility" that can be generated, as in Section 10.0. 20 It is defined on R~-' x R. We take
u(l" ... ,I,,) = Max
I,u,(x,)
+ ... + I"ul/(x,,)
S.t. (i) I,x, + ... + II/xI/ (ii)
X(h ~
0 for all (
~
(18.E.I)
~
I,w.
+ ... + II/wl/,
L - I and h.
The function v(l., ... ,I,,) is homogeneous of degree one in its arguments: l>(rl" ... , rl,,) = rv(l., ... , I,,) for all r. In particular, v(l./I, ... ,II//I)
Ftgure 18.0.6 (rtght)
Pareto optimal, self·selective allocation
In this section, we investigate the extent to which Walrasian equilibria can be characterized by the idea that individuals get exactly what they contribute to the economic welfare of society (at the margin). We will see that, once again, the assumption of a large number of consumers is crucial to this characterization. For an extensive analytical treatment of this topic we refer to the seminal contribution of Ostroy (19BO). To remain as simple as possible, we restrict ourselves to the case of quasilinear exchange economies. The Lth good is the numeraire. Suppose that our economy has H types. The concave, differentiable, strictly increasing utility function of type h is
I
= I v(I., ... , I,,).
That is, the per-capita social utility only depends on the type-composition and not on the size of the economy. Because of this we can extend the analysis to a situation with a continuum of consumers by defining v(~ ... .. '~I/) for any nonnegative vector
Q
~I"'" ~H) = Max
then the allocation is not Pareto optimal (assuming self-selectivity).
\
18,E Equilibrium and the Marginal Productivity Principle
E
U • LIB A I U II
II NOT HEll II A GIN II L
P A0 0 U C T IVIT Y
~ = (~I"'" ~II) E R~ of masses of the different types. Precisely,
If the net trade frontier is not flat,
'~~:<,
~,.~
1 •• E:
A
(lB.E.2) + ... + ~HUII(XII) (i) ~IXI + ... + ~I/XII ~ ~.WI + ... + ~IIWII'
~IUI(X.)
s.t.
(ii)
xI> ~
0 for all (
~
L - I and h.
If we have a sequence of finite economies (Ii, ... ,I';,) such that I' = :[.1. -+ 00 and (1/1')/. -+~. for every h, then we can properly regard (Il., ... , ~II) as the continuum limit of the sequence of increasingly large finite economies. Exercise IB.E.I: Show that the function v('): of degree one.
R~ -+
IR is concave and homogeneous
The function v(·) is a sort of production function whose output is social utility and whose inputs are the individual consumers themselves. Further, in the limit, every individual of type h becomes an input of infinitesimal size. For the time being, we concentrate our discussion on the continuum limit. We assume also that v(') is differentiable. 2 • Definition lB.E.l: Given a continuum population ~ = (~1" .. '~H) E R': a feasible allocation 22 (xf, ... , xk) is a marginal product, or no·surplus, allocation if
•
Uh(X h )
av(~)
=-a~h
for all h.
(18.E.3)
In words: at a no-surplus allocation everyone is getting exactly what she contributes at the margin. Proposition lB.E.l: For any continuum population ii = (ii, .. . ,iiH) »0 a feasible allocation (xf, ... , xk) » 0 is a marginal product allocation if and only if it is a Walrasian equilibrium allocation. Proof: If x· = (xt, . .. , x~) is a marginal product allocation then, using Euler's formula (see Section M.B of the Mathematical Appendix), we have
( _) v~
,,_ av(ii)
= 7:~'
a~,
,,_
= 7: ~,u,
(.)
x, .
Hence, x· solves problem (IB.E.2) for ~ = ii. Suppose now that x· = (xt, ... , x~) is a feasible allocation that gives rise to social utility vIii); that is, it constitutes a solution to problem (IB.E.2) for Il = ii. Denote by Pt, ( = I, ... ,L, the values of the multipliers of the first-order conditions associated with the constraints:[, ii,(xt. - (010) ~ 0, ( = I, ... , L, in the optimization problem (IB.E.2); see Section M.K of the Mathematical Appendix. By the quasilinear form of u,(') we have PL
for all (
:s L
=I
and
PI
= Vlt/I,(xt"
... , xl-I .• )
(IB.E.4)
- I and all I ~ h ~ H.
21. This could be derived from more primitive assumptions. 20. Because utility functions are concave the maximum utility can be reached while treating consumers of the same type equally.
P A INC I P L E
671
--------------------------------------------------------
22. We assume thai consumers of the samelype are treated equally. Feasibilily means therefore that
r,,, Jl"xt
:s;
LII P.W ...
672
CHAPTER
,.:
SOME
FOUNDATIONS
FOR
COMPETITIVE
EQUILIBRIA
--------------------------------------------------------------------------------------------It follows from (IS.E.4) that the vector of multipliers p = (P .. ... ,pd is the vector of Walrasian equilibrium prices of this quasilinear economy (recall the analysis of Section 10.0). In addition, by the envelope theorem (see Section M.L of the Mathematical Appendix), applied to problem (IS.E.2). we have (Exercise IS.E.2): ov(il) - = u. (*) x. 0/1.
+ p. (w. -
*) x •.
(IS.E.5)
Therefore. we conclude that x* is Walrasian if and only if x* solves problem (18.E.2) for /1 = /' and (18.E.3) is satisfied. that is. if and only if x* is a marginal product allocation. _ Expression (18.E.5) IS mtUltlve. The lert-hand side measures how much the maximum sum of utilities increases if we add onc cxtra individual of type II. The right-hand side tells us that there are two effects. On the one hand. the extra consumer of type It receives from the rest of the economy the consumption bundle and so she directly adds her utility u.(x:) to the social utility sum. On the other. while she contributes her endowment vector w •. Hence the net change for the receiving How much is this worth to the rest of the economy'! rest of the economy is w. The vector of social shadow prices is precisely p = (p, • ...• pd. and so the total change for the rest of the economy comes to p·(w. - x:>, Note that the Walrasian allocations are thus characterized by this second effect being null: the utility of the consumer equals her entire marginal contribution to social utility. In Exercise IS.E.4 you are asked to verify that the smoothness assumption on utility functions is essential to the validity of Proposition IS.E.!. Let us now consider a finite economy (I", ..• Ill) » O. We can define the marginal contribution of an individual of type h as
x:.
x:.
x:.
A.v(I ...... Ill) = v(I, •...• I ••...• Ill) - v(l, •...• I. - I •...• I,,).
Typically. there does not exist a feasible allocation (xT •...• x7,) with u.(x:l = A.v(/", .. , Ill) for all It. To see this. note that by the concavity of v(.) we have A.v ~ ov/O/1. [both expressions evaluated at (/, •... ,III)]. Except for degenerate cases. this inequality will be strict. Moreover. L. I.(ov/o/1.) = v(/, •...• Ill) by Euler's formula (see Section M.B of the Mathematical Appendix). and thus we conclude that L1I.(&.v) > v(I, •...• Ill); that is. it is impossible to give to each consumer the full extent of her marginal contribution while maintaining feasibility. [n contrast with the continuum case, individuals are not now of negligible size: their whole contribution is not entirely at the margin. [n particular. you should note that in a finite economy the Walrasian allocation is typically not a marginal product allocation. [t follows from expression (IS.E.5) that an allocation (xT •...• x7t) that solves problem (18.E.2) for (/1 , •...• /11l) = (I ...... 1/1) is a Walrasian equilibrium allocation if and only if
illl u.(x:) = - - (/, •...• I,,).
a/I.
But we have just argued that normally &.V(/" ...• Ill) > ov(l" . .. , IIl)/il/I•. [n words: At the Walrasian equilibrium consumers are compensated according to prices determined by the marginal unit of their endowments. But they lose the extra social surplus provided by the inframarginal units. This is yet another indication that the concept of Walrasian equilibrium stands on firmer ground in large economics.
- - - - ..... _ - - -
APPENDIX
A:
COOPERATIVE
GAME
THEORY
O/,.}
-----------------------------------------------------------------------------We have just seen that in the context of economies with finitely many consumers it is not possible to feasibly distribute the gains of trade while adhering literally to the marginal productivity principle. The cooperative theory of games provides a possibility for a sort of reconciliation between feasibility and the marginal productivity principle. It is known as the Shapley value. In Appendix A. devoted to cooperative game theory. we offer a detailed presentation of this solution concept. For an economy with profile (I, •... • Ill) the Shapley value is a certain utility vector (Sit, •...• Shll) E R"that satisfies L,l,Sh, = v(/, •...• I H ). For every type h.the utility Sh, can be viewed as an al'eraye oj marginal utilities d,v(/; •. ..• I~). The average is taken over profiles (I',' ...• /~) S (I" ...• (11 ), where the probability weight given to (I; •...• /~) equals 1//, interpreted as the probability assigned to sample size I; + ... + I~, times the probability of gelling thc profile (I; •...• 1;,) when independently sampling I; + ... + I~ consumers out of the original population with I consumers and profile (I, •...• I H ). See Appendix A for more on this formula. An allocation that yields the Shapley value (let us call it a Shapley alloration) is not related in any particular way to the Walrasian equilibrium allocation (or for that mailer to the core). Except by chance. they will be different allocations. Yet. remarkably. we also have a convergence of these concepts in economies with many consumers: the Walrasian and the Shapley allocations are then close to each other. This result is known as the value equivalence llieorelli. A rigorous proof of this theorem is too advanced to be given here [see Aumann (1975) and his references]. but the basic intuition is relatively straightforward. There are two key facts. First. if the entries of (I'" ...• /~) are large. then subtracting a consumer of type Ii amounts to very lillIe. and so
d,v(I', •...• I~) '" ov(/;, ...• I~)/o/". Second, if the entries of (I, •. ..• I H ) are large then, by the law of large numbers. most profiles (I; •.... I~) constitute a good sample of (I, •... • Ill) and are therefore almost proportional to (I, ... .• I,,). Using the homogeneity of degree one of v(') (hence the homogeneity of degree zero of "1'/<'11,). the combination of the previous two facts implies
d,t'(I'" ... ,
I~) '" cv(I', •... , I~) '" ilv(I" ... ,IH ) 0/',
0/',
for most (/'" ...• I;,). Therefore. Sh, '" ov(/" ...• IH)/O/I" which is the utility payoff of type h at the Walrasian equilibrium allocation of the economy (I, •...• 1,,).
APPENDIX A: COOPERATIVE GAME
THEORY
In this appendix. we offer a brief introduction to the cooperative theory of games. For more extensive recent accounts see Moulin (19SS). Myerson (1991). or Osborne and Rubinstein (1994).23 In Chapter 7, we presented the normal and the extensive forms of a game. The starting point for the cooperative theory is a classical third description: that of a 23. The text by Owen (1982). although not so recent. is nevertheless strong in its coverage of cooperative theory. Another useful reference is Shubik (1984). which is encyclopedic in spirit and contains a wealth of information.
674
CHAPTER
18:
SOME
FOUNDATIONS
FOR
COMPETITIVE
EQUILIBRIA
" P PEW D I X
" :
COO PER" Tt V E
G"" E
THE 0 R Y
675
------------------------------------------------------------------------------------------Definition 1B.AA.2: A game in characteristic form (/, V) is a set of players I and a rule V(·) that associates to every coalition ScI a utility possibility set V(S) c: RS •
",
Ftgur. 18.AA.l (te«)
Utility outcome" E R' and its projections. Flgur. 18.AA.2 (rtght)
A utility possibility set rorS={1.2}.
The elements of V(S) are to be interpreted as the payoffs the players in S can achieve by themselves if they jointly commit to a certain course of action. It is important to observe the expression "can achieve" is not free of subtlety. This is because the course of action undertaken by the members of I\S will typically affect the payoffs of the members of S. In applications, therefore, one should be explicit as to how V(S) is constructed. Example IS.AA.l: Economies. Consider an economy with I consumers having continuous, increasing, concave utility functions u,: R~ -+ R and endowments w, ~ O. There is also a publicly available convex, constant returns technology Y c RL. We can then define a game in characteristic form by letting V(S) = {(U;(Xi));'S:
L X; = L w, + y, y E id
yallli' ill c/wracraistic forlll. The characteristic form is meant to be a summary of the payolfs available to each group of players in a context where binding commitments among the players of the group are feasible. Although, in principle, it should be possible to derive the characteristic form from the normal or the extensive forms, the viewpoint of cooperative game theory is that it is often analytically desirable to avoid detail and to go as directly as possible to a summary description of the strategic position of the different groups of players. 24 After defining the characteristic form, we will discuss two of the main solution concepts of cooperative game theory: the core and the Shapley value. The set of players is denoted I = {I, ... ,I}. We abuse notation slightly by using the same symbol to denote the set and its cardinality. Nonempty subsets S, Tel are called coalitiolts. An olltcome is a list of utilities u = (u" ... , u,) E IR'. Given u = (u" ... , u,), the relevant coordinates to a coalition S are US = (u;);.s' Mathematically, US is the restriction (or projection) of u E IR' to the coordinates corresponding to S. We can therefore view US as a member of the Euclidean space It' spanned by these coordinates. Figure 18.AA.1 shows how outcomes for three players are evaluated by all six proper subsets: S = {I}, {2}, P}, p, 2}, p, 3}, and {2,3}.
ieS
Y} -
~.
That is, V(S) is the set of payoffs that the consumers in coalition S can achieve by trading among themselves and using the technology Y. Every set V(S) is convex (recall Exercise t 6.E.2). Figure t 8.AAJ depicts these sets for the case I = 3. • Example tS.AA.2: Majority Voting. Consider a three-player situation in which any two out of the three players can form a majority and select among a set of social alternatives A. If a E A is selected, the payoffs are u;(a) ~ 0, i = t, 2, 3. In addition, any player i has the right to unilaterally withdraw from the group and get a payoff of zero. Then we can define a game in characteristic form (I, V) as
Flgur. 18.AA.3 (leH)
A family of utility possibility sets.
V(I) = ((u,(a), uz(a), uJ(a)): a E A} - R/•. V({i, hI) = {(u;(a), u.(a)): a E A} V({i}) =
Rt~"l for all distinct pairs
{i, h}.
-lRt~.
Flgur. 1'.AA.4 (right)
Figure 18.AA.4 shows this characteristic form for a case with three alternatives,
The utility possibilily sets for the majority voting in Example
A = {a" az, aJ}. In the figure we suppose that the three decisions yield, respectively,
\8.AA.2.
",
Definition 1B.AA.1: A nonempty, closed set US c IRs is a utility possibility set for Sci if it is comprehensive: US E US
and
u's S US implies u,s E Us.
See Figure 18.AA.2 for an illustration.2>
24. Nevertheless. we note that there is a school of thought within game theory of the opinion that the condensation of information that a characteristic form represents may not do justice to the strategic complexities inherent in the making of binding commitments. Despite the cogency of this position. the analytical power of games in characteristic forms for the study of normative issues in economics has been amply demonstrated. This is more than enough reason to welcome the parsimony it brings to the analysis. 25. No,e that. as we did in Section t6.E. we build rree disposability or utility into the definition or a utility possibility set.
V(i2,
3D
",
",
",
676
C HAP T E R
1
a:
S 0 M E
F 0 UNO A T ION S
FOR
COM PET I T I VEE QUI LIB R I A
A P PEN 0 I X
--------------------------------------------------------------------------------------------
A:
COG - F. RAT lYE
G A JIll! E
T H f. 0 R Y
677
",
the utility vectors (2, 1,0), (I, 0, 2), (0, 2, I). Note that V(·) need not be convex when, as here, decisions are discrete and there are no possibilities for either randomization or any form of side transfers. _ Definition 18.AA.3: A game in characteristic form (I. V) is superadditive if for any coalitions S, Tel that are disjoint (I.e., such that that S,.., T = 0), we have if
US E
VIS) and u T E VeT), then (us, u T )
E
VIS u T).
Superadditivity means that coalitions Sand T are able to do at least as well acting together as they could do acting separately. It is an assumption we will commonly make (it is satisfied by Examples IB.AA.1 and IB.AA.2). If one of the possibilities open to the union of two disjoint coalitions is to agree to act as if they were still separated coalitions, then superadditivity should hold.
Figure 18.AA.S
It has been a constant theme of this book that often the analysis becomes much simpler when individual utility functions are quasilinear, that is, when there is a commodity ("numeraire") that can be used to effect unit-per-unit transfers of utility across economic agents. The same is true in the theory of cooperative games. Its history is, in fact, replete with instances of concepts first formulated for the transferable utility case that have later been extended to the general setting without an essential loss of intuition and analytical power. For a situation described by a game in characteristic form, what the quasilinearity, or transferable utility, hypothesis amounts to is the assumption that the sets VIS) are half-spaces (as they were, for example, in Section 10.0); that is, they are sets whose boundaries are hyperplanes in RS • Moreover, by choosing the units of utility, we can take the hyperplanes defining VIS) to have normals (I, ... , I) E RS.26 Thus, the sets V(S) will now have the form V(S)
(O,liN),O)
2
Figure 18.AA.6
Usi ng a simplex to represent a three3~-----L
____- L_ _
(O,O,'~N))
player TU-game with utitities normalized so that V({i}) = O.
~
(liN),O,O)
preferences arc quasilinear). Then the characteristic function is
= {us E IRs: .L uf :5; V(S)} •• s
v(/) = 3,
for some v(S) E R. In other words, we can view coalition S as choosing a JOlDt action so as to maximize the total utility, denoted v(S), which then can be allocated to the members of S in any desired manner through transfers of the numeraire. Figure IB.AA.5 depicts the sets V(S) for the case [ = 3. The number v(S) is called the wortlt of coalition S. Since the numbers v(S), S c [, constitute a complete description of (I, V) we provide Definition IB.AA.4.
v({I, 2}) = v({ I, 3}) = v({2, 3}) = 3
and
vIIi}) = 0, i = 1,2,3 .
-
All the developments in this appendix will be invariant to changes in the origins of individual utilities; thus we can fix these arbitrarily. The usual convention is to put v([i}) = 0 for every i. In Figure IB.AA.6 we represent a diagrammatic device for the case of three-player games that is particularly useful. Instead of working in three dimensions, we look at the two-dimensional simplex that exhibits all possible divisions of v(l) subject to the condition that U; ~ 0 for all i [which means, in the normalization just discussed, that U i ~ v( {i})]. The other selS in the diagram represent, for every two-person coalition {i, It}, the utility combinations in the simplex satisfying U; + u.:5; vIIi, It}). We now turn to a presentation of two well-known solution concepts for cooperative games: the core and the Shapley value.
Definition 1B.AA.4: A transferable utility game in characteristic form, (or TV-game), is defined by (I. v), where [ is a set of players and v(·) is a function, called the characteristic function, that assigns to every nonempty coalition S c [ a number vIS) called the worth of S. Example 18.AA_3: TV Majority Voting. Suppose that in Example IB.AA.2 (with the values of Figure IB.AA.4) we add the possibility that utility be freely transferable across players (there may be a numeraire commodity with respect to which
The Core
26. This choice or units or utility is legitimale because all the solutions to be considered are invariant to normalizations of units. See Chapter 21 for more on this point.
Ulility possibility sels for a transrerable ulility game.
",
I
1
The first solution concept we review is the core: the set of feasible utility outcomes with Ihe property that no coalition could on its own improve the payoffs of all its members. An empty core is indicative of competitive instability in the situation being modeled. If the core is nonempty and small, then we could say that coalitional
678
CHAPTER
11:
SOME
FOUNDATIONS
FOR
COMPETITIVE
APPENDIX
EQUILIBRIA
V«(I,3})
Flgur. 18.AA.7
(al A TU·game with a nonempty core.
(bl A TU game with an empty core.
(al
competition by itself brings about a sharply determined outcome. If it is nonempty and large, then coalitional competition alone does not narrow down the possible outcomes very much.
COOPERATIVE
there is identical with the concept considered here for games in characteristic fonn.21 We conclude, therefore that if a Walrasian equilibrium exists then the core is nonempty. _
2
1\
Core, Nonempty
A:
Example IS.AA.6: Single·Input, Increasing Returns Production Function. Consider a one.input, one·output world in which there is a publicly available technology f(z) which is continuous and satisfies flO) = O. There are I players. Each player i cares only about the consumption of the output good and owns an amount w, of input. Assuming that utility is transferable, we can define a TU characteristic function by v(S) = Wi)' The core of this game will be nonempty whenever the technology exhibits nondecreasing returns to scale, that is, whenever average product f(z)/z is nondecreasing. [In particular, if f(·) is convex, that is, if we have a nondecreasing marginal product, then f(z)/z is nondecreasing.] To verify this, suppose that we distribute the product proportionally:
Hr., ..
Definition 18.AA.S: Given a game in characteristic form (I, V), the utility outcome u E R' is blocked, or improved upon, by a coalition ScI if there exists u's E VIS) such that uf < u;s for all i E S. If the game is a TU game (I, v) then the outcome u = (u" .•. , U/) is blocked by S if and only if :L.s Ui < v(S).
for every Ii E I. Then, for any ScI we have
Definition 18.AA.6: A utility outcome u = (u t ' ... , u l ) that is feasible for the grand coalition [Le., u E V(llJ is in the core of the characteristic form game (I, V) if there is no coalition S that blocks u. In TU games the core is the set of utility vectors u = (u" ... , U/) satisfying the linear inequalities
I Ui ~ v(S) for all ScI ,.s
and
I
Ui
S v(1).
1.1
Figure IS.AA.7(a) depicts a three·player game with nonempty core. In contrast, in Figure IS.AA.7(b) the core is empty. See Exercise IS.AA.l for a set of necessary and sufficient conditions in the TU case for the nonemptiness of the core. Exercise IS.AA.2: Show that any TU game with a nonempty core must satisfy: For any two coalitions S, Tc I such that S,.., T = 0 and S u T = I, we have v(S)
where the inequality follows because average product is nondecreasing. We conclude that this proportional distribution of output belongs to the core. In Exercise IS.AA.3 you are asked to show that if average product. is constant then the proportional allocation is the only allocation in the core. It is also intuitively clear that the more pronounced is the degree of increasing returns, the more difficult it will be for proper subgroups to do better by their own means (they will have relatively low average product) and therefore the more we could depart from the proportional allocation while remaining in the core. Hence, for this sort of one-dimensional distribution problem, the larger the degree of increasing returns, the larger the core will be. 28 _
+ veT) :s v(l).
Example lS.AA.4: Majority Voting, Once Again. For the majority voting games described in Examples IS.AA.2 and IS.AA.3, the core is empty. In the latter, which is a TU game, this is clear enough: if u, + U 2 + U) = 3 then u, + u. < 3 for some i, h. Hence the coalition {i, h} will block. For the former (nontransferable utility) game, note that the outcomes (2, 1,0), (1,0,2), (0,2, I) are blocked, respectively, by the coalitions {2,3} using a), (1,2) using a" and {I,3} using a2' These examples constitute instances of the so called Condorcet paradox (which we have already encountered in Section I.B and will see again in S.~tion 21.C). They are illustrative of an inherent instability of majority voting. Example lS.AA.5: Economies, Again. The economic example in Example IS.AA.l was extensively studied in Section IS.B. Note that the concept of the core discussed
GAME
THEORY
679
--------------------------------------------------------------------------------------------
The Shapley Value The core tries to capture how the possible outcomes of a game may be shaped by coalitional competitive forces. It is the simplest solution concept in what could be called the descriptive side of cooperative game theory. We shall now investigate a solution concept, the value, whose motivation is normative. It attempts to describe
27. The connection of the solution concept proposed by Edgeworth (1881) with the modern game-theoretic notion of the core was made in Shubik (1959). 28. On the other hand, if I( .) exhibits strictly decreasing returns then it follows from Exercise IS.AA.2 that the core is empty [indeed. (~S) + v(T) > v(1) for any partition of I into two coalitions S. TJ.
680
';HAPTER
11:
SOME
FOUNDATIONS
FOR
COMPETITIVE
A P P E '" v I X
EQUILIBRIA
A:
COO PER A T I V EGA MET H E 0 R Y
players i, h E S, utility differences are preserved in a manner similar to the two-player case: ShieS, v) - S/,,(S\{h}. v) = Sh.(S, v) - Sh.(S\{i}, v)
",
for all ScI, i, h E S,
L ShieS, v) = v(S)
(IS.AA.2)
for all ScI,
iES
1~{2H-I----4'=-.J.._-~.----
Figure 1B.AA.8
o
u,
{(u,. u,):", + u, =
Egalitarian division for two-player games.
'111. 2})}
Expressions (IS.AA.2) determine the numbers ShieS, v), i E S, uniquely. This is clear for Sh,({i}, v). From here we can then proceed inductively. Suppose that we have dcfined S/,,(S, v) for all ScI, S 'I' I, i E S. We show that there is one and only one way to define Sh,(f, v), i E I. To this effect, note that (tS.AA.2) allows us to express every Sh,(I, v) as a function of Sh,(f, v) and of already determined numbers:
Shi(f, v) = Sh,(f, v) + Sh,(f\{ I}, v) - Sh,(1\{i}, v) Then to determine Sh,(f, ") use :[;,' Sh i(1. v)
a reasonable, or "fair," way to divide the gains from cooperation, taking as a given the strategic realities captured by the characteristic form.29 We study only the TV case, for which the theory is particularly simple and well established. The central concept is then a certain solution called the Shapley value. 30 Suppose that individual utilities are measured in dollars and that, so to speak, society has decided that dollars of utility of different participants are of comparable social worth. The criterion offairness to which value theory adheres is egalitarianism: the aim is to distribute the gains from trade equally. To see what the egalitarian principle could mean in the current TV context let us begin with a two-player game (I, v) = ({ I, 2}, v). Then the gains (or losses, if superadditivity fails) from cooperation are
v(l) - v({I}) - v({2}). Therefore, the obvious egalitarian solution, which we denote (Sh,(I, v), Sh 2 (1, v», is (see Figure IS.AA.S)
Sh;(I, v) = v({i}) + !
i = 1,2.
(IS.AA.I)
How should we define the egalitarian solution (Sh,(I, v), ... , Sh,(I, v» for an arbitrary TV-game (I, v)1 We have already solved the problem for two-player games. It is suggestive to rewrite expression (IS.AA.I) as
Sh,(I, v) - Sh,({I}, v) = Sh 2 (1, v) - Sh 2 ({2}, v), Sh,(I, v)
+ Sh 2(1, v) = v(I),
where we put Shi({i}, v) = v({i}). In words, this says that utility differences are preserved: What player 1 gets out of the presence of player 2 is the same as player 2 gets out of the presence of player 1. This points to a recursive definition: Given ScI, denote by (S, v) the TV-game obtained by restricting v(.) to the subsets of S [this is called a subgame of (I, v». Then we could say that a family of numbers {ShieS, v)}s c /,i.S constitutes an egalitarian solution if, for every subgame (S, v) and
S/,,(l. v) = 1[(1(1) I
L
= v(f).
Sh,(l\{I}, v)+
i* I
for all i'l' t.
Specifically,
L
Sh,(l\{i}, V)].
1* I
Definition 18.AA.7: The Shapley value of a game (I. v). denoted Sh(l. v) = (Sh,(I. v) ....• Sh,(l. v)).
is the single outcome consistent with expression (lB.AA.2). We can compute S/,,(l, v) in a direct and interesting manner as follows. For any ScI and if S let m(S, i) = v(S v {ill - v(S) be the marginal conrributioll of ito coalition S." For any ordering x of the players in 1 (technically, x is a one-to-one function from I to l) denote by Sex, i) C 1 the set of players that come before i in the ordering x [technically, S(x, i) = {h: x(h) < xU)}]. Note that, for any given ordering 7[, if we consider the marginal contributions of every player i to the set of the predecessors of i in Ihe ordering x, then the sum of these marginal contributions must exactly exhaust v(l); that is, LiE' m(S(x, i) = v(/). It then turns oul that S/,,(l, v) is the average marginal contribution of i to the set of her predecessors. where the average is taken over all orderings (held to be equally likely). Since the tOlal number of orderings is I! Ihis gives
n,
t S/',(I, v) = m(S(x, i), i). I! •
L
(IS.AA.3)
where the sum is taken over all orderings x of the players in I. Example IS.AA.7: Glove Market. Consider the three-player game defined by
v({ I, 2, 3}) = I,
= t, v({I, 2}) = 0, = v({2}) = v({3}) = O.
v({I, 3}) = v({2, 3}) v({!})
If the utility of a matched pair of gloves is I, while an unmatched pair is worth nothing. then this game could arise from a situation in which players I and 2 own
29. Thus. the redistributional fairness considerations that we will discuss in Sections 22.8 and
22.C, based on notions of absolute justice. are alien to the value. 30. It is named after L. Shapley. who proposed it in his Ph.D. dissertation at Princeton (in 1953).
31. Whenever we compute marginal contributions we follow (he convention ",(S.
i) = ,,((ill whenever S = 0.
L~0)
= O. Therefore.
681
682
CHAPTER
11:
SOME
FOUNDATIONS
FOR
COMPETITIVE
EQUILIBRIA
APPENDIX
A:
COOPERATIVE
--------------------------------------------------------------------------------------one right-hand glove each, while player 3 owns a left-hand glove. Let us compute Sh,(l, v). There are six possible orderings of the players: (1,2,3}, {I, 3, 2}, (2, I, 3}, (2,3, I}, (3, I, 2}, and {3,2, l}. The marginal contribution of player 3 to its predecessors in each of these orderings is, respectively: I, I, I, I, 0, and O. The average of these numbers is t; hence Sh,(I, v) = i. Similarly, we get Sh l (1, v) = Sh 2 (1, v) = t. Note that these numbers satisfy (IS.AA.2). For example, ShJ(I, v) - Sh,(1\{l}, v) =
1-! ~ Ii -
0 = Sh l (1, v) - Sh.(I\{3}, v).
-
We can give a more explicit formula than (IS.AA.3) for 5h,(I, 0). The probability that in a random ordering a given coalition Tc I, i e T, arises as the union of I and its predecessors equals the probability that j is in the Tth place," which is simply 1//, multiplied by the probability that T\{i} arises when we randomly select #- T - I members from the population I\{i}. which is (1- #- T)!(#- T- I)!/(I- I)!. Hence, we can rewrite (IS.AA.3) as 5h,(I, v)
= L
[(1- #- T)!(#- T - 1)!//!](v(T) - v(T\{i})).
In Exercise IS.AAA you are asked to verify that if we were to define the Shapley value by (IS.AAA), or (tS.AA.3), then equations (IS.AA.2) would be satisfied; this means that, indeed, (IS.AA.3) or (IS.AAA) provide correct formulas for the Shapley value. We now put on record, rather informally, some of the basic properties of the Shapley value.
LI Sh l (1, v) =
Example IS.AA.7 continued: In the glove market example a core utility outcome is (0,0, I). Moreover, this is the only outcome in the core. Indeed, if (II" " 2, II,) with L, u, = I has, say, u, > 0, then the coalition (2,3} can block by means of (0, U2 + lU" II, + lu,). In effect, at the core the two owners of right-hand gloves undercut each other until they charge a price of zero. In contrast, the Shapley value, while heavily skewed towards player 3, nonetheless leaves something to the other two players
(IS.AAA)
Tc: I,I.T
(a) Efficiency.
core. In a sense, we already know this because the Shapley value is defined for all games and there are games for which the core is empty. But the phenomenon can also occur if the core is nonempty. To see this, let us reexamine the glove market of Example IS.AA.7.
v(I); that is, no utility is wasted.
(b) Symmetry. If the games (I, v) and (1, v') are identical, except that the roles of players i and h are permuted," then Sh,(1, v) = Sh.(1, v'). In words: The Shapley values do not depend on how we label players; only their position in the game, as summarized by the characteristic function, matters. (c) Linearity. Note from (IS.AA.3) or (IS.AAA) that the Shapley values depend linearly on the data, that is, on the coefficients v(S) defining the game. (d) Dummy axiom. Suppose that a player j contributes nothing to the game; that is, v(S v (i}) - v(S) = 0 for all ScI. Then Sh,(1, v) = O. This important property follows directly from (IS.AA.3): The marginal contribution of player j to any coalition is null; hence its average is also null. These four properties fully characterize the Shapley value. Although the proof of this fact is not difficult, we shall not give it here. See Exercises IS.AA.5 and IS.AA.6 for the discussion of some examples. Given a game, the Shapley value assigns to it a single outcome. In contrast, the core solution assigns a set. We point out that the Shapley value need not belong to the 32. The symbol # T denotes the number of players in a coalition T. 33. Precisely, v(S) = v'(S) whenever i e Sand he S, v(S) = v'(S) whenever i ~ Sand h; S, v(S)=r'«S\(i})v{h}) whenever ieS and h;S, and v(S)= v'«S\(h}) v (i}) whenever heS and i r/ S.
DefinItion 18.AA.8: A game (I, v) is convex if for every i the marginal contribution of i is larger to larger coalitions. Precisely, if SeT and i E I\T, then vIS v (i}) - vIS) ;s; v(T v {il) - v(T).
Example 18.AA.8: Complementary Illputs. Let f(z" ... , ZN) be a production function displaying increasing marginal productivities with respect to all inputs' that is 2 ' , iJ f(z)/az. az, ~ 0 for all z and h, k. Suppose that every player i E I is endowed with a vecto~ of inputs w, E R"•. Then we can define a TV-game by v(S)
= fI the condition iJl f(z)/oz. az, ~ 0 for all z and h, k, is neither necessary nor sufficient for the convexity of f(·). In fact, the convexity of f(·) is far from sufficient for the convexity of the game (see Exercise IS.AA.S). _ We can then show the result in Proposition IS.AA.l. PropositIon 18.AA.1: If a game (I, v) is convex then its Shapley value utility outcome Sh(l, v) = (Sh,(I, v), . . " Shdl, v» belongs to the core (in particular, the core is nonempty). Proof: It is enough to show Ihat if j eSc T then Sh,(S, v) :S Sh,(T, v). Indeed, for any S c J this implies that v(S) = L,.5 Sh,(S, v) :S L, .. Sh,(I, v) and therefore the coalition S cannot block. To prove the claimed property it suffices to consider j e Sand T = S v (hI. Given an ordenng n of S denote by mIn, j) the marginal contribution of i to its predecessors in 5 and according to the ordering n by m'(n, i) the average marginal contribution of j to its predecessors in T when the average is taken over the # T orderings of T differing from the given ordering n of S only by the placement of ". Then I Sh,(S, v) = '--S-,
L mIn, i)
# .•
and
I SII,( T, v) = - -
L m'(n, j).
#S! •
Note that for every ordering n of S we must have m'(n, i)
GAME
THEORY
683
684
CHAPTER
11:
lOME
fOUNDATIONS
fOR
COMPETITIVE
EQUILIBRIA EXERCISES
the marginal contribution of / to its predecessors in T is still m(It, I); if we place h before i then, by the convexity condition, this marginal contribution is at least m(It, I). We conclude that Sh,rT, v) ~ Sh,eS, v), as we wanted to show. •
18.B.2" Exhibil an example of a nonequal-treatment core allocation in a three-consumer exchange economy with continuous. strictly convex, strongly monotone preferences. Can an example be given with only two consumers? A IS.B.3 Give a direct proof (i.e., not using properties of the core) that a Walrasian allocation of an economy wilh continuous, strictly convex preferences has the equal-treatment property.
REFERENCES
18.B.4' Use Taylor's formula to complete Ihe proof of Proposition IS.B.3.
Anderson. R. (1978). An elementary core equivalence theorem. Econometrica 46: 83-87. Aumann. R. (1964). Markets with a continuum of traders. Econometrica 32: 39-50. Aumann. R. (1975). Values of markets with a continuum of traders. Econometrica 43: 611-46. Champsaur. P., and G. laroque. (1981). Fair allocations in large economies. Journal of Economic Theory
IX-B.S" Consider an economy composed of 21 + I consumerS. or these, I each own One right shoe and I + I each own a left shoe. Shoes are indivisible. Everyone has the same utility function. which is Min {R. L}. where Rand L are. respectively. the quantities of right and lerl shoes consumed.
25: 269-82. Debreu. G .• and H. Scarf. (1963). A limit theorem on the Core of an economy. inlr'nOlionat Economic Review ~: 235-46. Edgeworth. F. Y. (1881). Mathematical Psychics. London: Kegan Paul. Foley. D. (1967). Resource allocation and the public sector. Yale Economic Essays 7: 45-98.
Gabszewicz. J. J.• and J. P. Vial. (1972). Oligopoly·. la Cournot· in a general equilibrium analysis. Journal of EClmomic Theory 4: 381-400. Hart, O. (1980). Perfect competicion and optimal product differentiation. Journal of Economic Theory 11: 165-99. Hildenbrand. W. and A. Kirman. (1988). Equilibrium Anal),sis. New York: Norlh·Holland. Mas·Colell, A. (1982). The Cournotian roundations or Walrasian equilibrium: an exposition or recent theory. Chap. 7 in Advanct's in Economic Theory, edited by W. Hildenbrand. New York: Cambridge University Press. Moulin, H. (1988). Ax/o,.., of Cooperative Game Theory. New York: Cambridge University Press. Myerson, R. (1991). Game Theory: Analysis of Conflict. Cambridge, Mass.: Harvard UniversilY Press. Novshek, W.. and H. Sonnenschein. (1978). Cournot and Walras equilibrium. Journal of Economic Theory
19: 223-66. Roberts. K. (1980). The limit points of monopolislic competition. Journal of Economic Theory 22: 256-278. Osborne, M. and A. Rubinstein. (1994). A Course in Game Theory. Cambridge, Mass.: MIT Press. Ostray, J. (1980). The no-surplus condition as a characterization of perfectly competitive equilibrium. Journal of Economic Theory 22: 65-9\. Owen. G. (1982). Game Theory, 2nd ed. New York: Academic Press. Schmeidler, D. and K. Vind. (1972). Fair net trades. Econometrica 40: 637-47. Shapley. L.. and M. Shubik. (1977). Trade using a commodity as a means of payment. Journal of Political Economy 85: 937-68. Shubik, M. (1959). Edgeworth's market games. In Contributions to the Theory of Games. IV. edited by R. D. Luce. and A. W. Tucker. Princeton, NJ.: Princeton University Press. Shubik, M. (1984). Game Theory in the Social Sciences. Cambridge, Mass.: MIT Press. Thomson, W., and H. Varian. (1985). Theories of justice based on symmetry. Chap. 4 in Social Goals and Social Organizations, edited by L. Hurwicz, D. Schmeidler, and H. Sonnenschein. New Vorlc: Oxford University Press. Varian. H. (1976). Two problems in the theory of fairness. Journal of Public Economics S: 249-60. Vind, K. (1964). Edgeworth allocations in an exchange economy with many traders./nlt'rnalional Economic Review 5: 165-77.
(a) Show that any allocalion of shoes that is matched (i.e .• every individual consumes the same nllmhcr of shoes of each kind) is a Parelo optimum, and conversely. (b) Which Parelo oplima are in the COre of this economy? (This time, in Ihe definition of the COre allow for weak dominance in blocking.) (e) Let P. and Pi. be the respective prices of the two kinds of shocs. Find the Walrasian equilibria of this economy.
(d) Comment on Ihe relationship between the core and the Walrasian equilibria in this economy. IS.C.I'" ESlablish the properties of effective budget sets claimed in the discussion of Example I X.C.3. You can restrict yourself to the case L = 2. IS.O.I" Consider an Edgeworth box wilh continuous, strictly COnvex and monotone preferences. Show that every feasible allocalion where both consumers are at least as well off as al their initial endowments is self.selective. IS.E.I" In texl. \S.E.2' Use the envelope theorem (see Section M.L of the Mathemalical Appendix) to derive expression (IS.E.5). IS.E.3" By considering an example with L-shaped preferences for two non.numeraire goods (hence, the utility function cannol be differentiable), argue that it is possible that at a Walrasian allocalion with a continuum of traders every trader gets less than her marginal conlribution. 18.AA.I" A collection of coalitions S" ... , SH C I is a generalized partition if we can assign a weIght b, E [0. I] to every I S n S N such that, for every player i E I. we have LI" i.S.1 b, = 1. ExhIbit examples of generalized partitions, with the corresponding weights. We say thai a TU-game (I, v) is balanced if for every generalized partition we have L. I\, ..(S.) S ..(I). where b, are the corresponding partition weights. Show that the game has a nonempty Core if and only if it is balanced. [Hint: Appeal to the duality theorem of linear programming (see Section M.M of the Mathematical Appendix).] IS.AA.2' In texl. A
EXERCISES
I8.AA.3 Show that the proportional allocation of Example IS.AA.6 is the only allocation in the core if average product is constant.
18.B.IA Show that Walrasian allocations are in the core for the model with a constant returns technology described in Section 18.B.
18.AA.4C Show that if the Shapley value is defined by formula (lS.AA.4)-or, equivalently. by (18.AA.3 )-then the preservation of differences expression (IS.AA.2) is satisfied.
685
686
CHAPTER
18:
SOME
FOUNDATIONS
FOR
COMPETITIVE
EOUILIBRIA
18.AA.5" We say that a game (I, v) is a unanimity game if there is a nonempty Sci such that v(T) = v(S) if SeT and v(T) = 0 otherwise. Show then that under the efficiency, symmetry, and dummy axioms we arc led to distribute v(S) equally across the members of S. 18.AA.6" Show that any TU-game (I, v) can be expressed as a linear combination of unanimity games. Then use the Exercise IS.AA.5 and the linearity axiom to show that there is a unique solution satisfying the efficiency, symmetry, dummy, and linearity axioms. Connect your discussion wilh the Shapley value.
CHAPTER
General Equilibrium Under
19
Uncertainty
18.AA.7c Show that the production game described in Example IS.AA.S is convex. 18.AA.8" In the context of the production example of Example IS.AA.S, give an example of a two-input production function that is convex (as a function) but for which, nonetheless, the core is empty (thus. the induced game cannot be convex). 18.AA.9" Consider the game with four players defined by v({i}) = 0, v({12}) = v({34}) = 0, = v({ 14}) = v({23}) = v({24}) = I, v({ijk}) = I for all three-player coalitions {ijk}, and ,,({I234}) = 2.
L{{ 13})
(a) Show that this is the game that you would get from the utility production technology Min {=" =,l, where z, and z, are the amounts of two factors, if the factor endowments of the four consumers are w, = w, - (1,0) and wJ = W. = (0,1). ,
(b) Show that the core of this game contains all points of the form (a. a, I - a, I E [0, I].
~)
for
(e) Show that if v({ 134}) is increased to 2, holding all other coalition values constant. there is then only one point in the core. Compare the welfare of player I at this point to what she would get at all the points in the core before the increase in v({I34}). (d) Compute the Shapley value of the game [before the modification in (e)) without using the brUle-force enumeration technique. [Hint: Use symmetry considerations and other axiomatically based simplifications to go part of the way to the answer.] (e) How does the Shapley value change under the modification of part (e)? Discuss the difference between the changes in the Shapley value and in the core. 18.AA.IO" Consider a firm constituted by two divisions. The firm must provide overhead in Ihe form of space. (x" x,), to each of them. The cost of aggregate amounts of space is given by C(x, + x,) = (x, + x,)', 0 < y < I. (a) Suppose that, whatever the usage of space (x" x,), the total cost must be exactly allocated between the two divisions. Propose a cost allocation system based on the Shapley value to accomplish this. (b) Compute the marginal cost imposed on each of the two divisions [according to the cost allocalion system identified in (a)) whenever a division increases its usage of space. (c) Suppose now that the profits accruing to the two divisions arc Il,X, and Il,X" respectively (we assume that 11, > 0 and 11, > 0), and that each division uses space to the point where marginal profits equal own marginal costs [as determined in (b)). Will this lead to an efficienl (that is, profit-maximizing) choice of overhead? (d) Is there any distribution rule ""(x,, x,), ""(x,, x,), with ""(x,, x,) + ""(x,, x,) = C(x, + .<,) for all (x" .<,), that leads to efficient decentralized choice [in the sense of (e)) for all 11" ,,? [Hillt: Consider the externality imposed by each division on the othef.]
19.A Introduction In this chapler. we apply Ihe general equilibrium framework developed in Chaplers 15 10 I H to economic silualions involving Ihe exchange and allocation of resources under conditions of uncertainty. In a sense, this chapter offers the equilibrium counterpart of the decision theory presented in Chapter 6 (and which we recommend you review at this poinl). We begin, in Seclion 19.B, by formalizing uncertainty by means of states of the world and then introducing the key idea of a contingent commodity: a commodity whose delivery is conditional on the realized state of the world. In Section 19.C we usc these tools to define the concept of an Arrow-Debreu equilibrium. This is simply a Walrasian equilibrium in which contingent commodities are traded. It follows from the general theory of Chapter 16 that an Arrow-Debreu equilibrium results in a Pareto optimal allocation of risk. In Section 19.0, we provide an important reinterpretation of the concept of Arrow-Debreu equilibrium. We show that, under the assumptions of self1u/jilling, or raliollal, expectations, Arrow-Debreu equilibria can be implemented by combining trade in a certain restricted set of contingent commodities with spot trade that occurs aft"r the resolution of uncertainty. This results in a significant reduction in the number of ex ante (i.e., before uncertainty) markets that must operate. In Section J9.E, we generalize our analysis. Instead of trading contingent commodities prior to the resolution of uncertainty, agents now trade assets; and instead of an Arrow-Debreu equilibrium we have the notion of a Radner equilibrium. We also discuss here the important notion of arbitrage among assets. The material of this section lies at the foundations of a very rich body of finance theory [good introductions are Duffie (\992) and Huang and Litzenberger (1988)]. In Section 19.F, we briefly illustrate some of the welfare difficulties raised by the possibility of incomplete markets, that is, by the possibility of there being too few asset markets to guarantee a fully Pareto optimal allocation of risk. Section 19.G is devoted to the issue of the objectives of the firm under conditions of uncertainty. In particular, it gives sufficient conditions for shareholders to agree unanimously on the objective of market value maximization. 687
688
CHAPTER
11:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
Section 19.H takes a close look at the informational requirements of the theory developed in this chapter. We see that the theory applies well to situations of symmetric information across consumers (reviewed in Section 19.H); but its applicability is more problematic in situations of asymmetric information. This provides a further argument for the techniques developed in Chapters \3 and 14 for the study of asymmetric information problems. For additional material and references on the topic of this chapter, see the textbooks of Huang and Litzenberger (1988) and Duffie (1992) already mentioned, or, at a more advanced level, Radner (1982) and Magill and Shafer (1991).
19.B A Market Economy with Contingent Commodities: Description As in our previous chapters, we contemplate an environment with L physical commodities, I consumers, and J firms. The new element is that technologies, endowments, and preferences are now uncertaill. Throughout this chapter, we represent uncertainty by assuming that technologies, endowments, and preferences depend on the state of the world. The concept of state of the world was already introduced in Section 6.E. A state of the world is to be understood as a complete description of a possible outcome of uncertainty, the description being sufficiently fine for any two distinct states of the world to be mutually exclusive. We assume that an exhaustive set S of states of the world is given to us. For simplicity we take S to be a finite set with (abusing notation slightly) S elements. A typical element is denoted s = I, ... , S. We state in Definition 19.B.I the key concepts of a (state-)contingent commodity and a (state- )contingent commodity vector. Using these concepts we shall then be able to express the dependence of technologies, endowments, and preferences on the realized states of the world. Definition 19.B.1: For every physical commodity t = 1, ... ,L and state s = 1, ... ,S, a unit of (state-) contingent commodity (s is a title to receive a unit of the physical good {if and only if s occurs. Accordingly, a (state-) contingent commodity vector is specified by
x = (x", . .. , XL'"
•. ,X,s• .••• XLS) E
IR LS •
and is understood as an entitlement to receive the commodity vector (X,.•...• x Ls ) if state s occurs.' We can also view a contingent commodity vector as a collection of L random variables, the tth random variable being (X{I" .. ,x(s). With the help of the concept of contingent commodity vectors, we can now describe how the characteristics of economic agents depend on the state of the world. To begin, we let the endowments of consumer i = I, ... ,I be a contingent commodity vector:
SECTION
19.8.
A
MARKET
ECONOMY
WITH
CONTINGENT
COMMODITIES:
The meaning of this is that if state s occurs then consumer i has endowment vector E RL. The preferences of consumer i may also depend on the state of the world (e.g., the consumer's enjoyment of wine may well depend on the state of his health). We represent this dependence formally by defining the consumer's preferences over contingent commodity vectors. That is, we let the preferences of consumer i be specified by a rational preference relation ~; defined on a consumption set X; c IR LS • (w,,;, ... ,wu ')
Example \9.B.I: As in Section 6.E, the consumer evaluates contingent commodity vectors by first assigning to state s a probability n" (which could have an objective or a subjective character), then evaluating the physical commodity vectors at state s according to a Bernoulli state-dependent utility function u,;(x,,;, ... , xu;). and finally computing the expected utility.2 That is, the preferences of consumer i over two contingent commodity vectors x" x; E X, C RLS satisfy Xi
~iX; if and only if
L, "';U'i(X"i'···' xu;) ~ L n,;u'i(X"," ...• XL,i)'
-
It should be emphasized that the preferences ~i are in the nature of ex ante
preferences: the random variables describing possible consumptions are evaluated before the resolution of uncertainty. Similarly. the technological possibilities of firm j are represented by a production set }j c IRIs. The interpretation is that a (state-)contillgent production plan Yj E RI•s is a member of >j if for every s the input-output vector (YI'j' ..• , Yuj) of physical commodities is feasible for firm j when state s occurs. Example 19.8.2: Suppose there are two states, s, and S2' representing good and bad weather. There are two physical commodities: seeds (f = I) and crops (t = 2). In this case. the elements of >j are four-dimensional vectors. Assume that seeds must be planted before the resolution of the uncertainty about the weather and that a unit of seeds produces a unit of crops if and only if the weather is good. Then Yj = (Yllj.Y2Ij'Y'2j.Y22j) = (-1.1, -1,0)
is a feasible plan. Note that since the weather is unknown when the seeds are planted, the plan (-1. 1,0,0) is not feasible: the seeds, if planted, are planted in both states. Thus. in this manner we can imbed into the structure of >j constraints on production related to the timing of the resolution of uncertainty.' _ To complete the description of an economy in a manner parallel to Chapters 16 and 17 it only remains to specify ownership shares for every consumer i and firm j. I n principle, these shares could also be state-contingent. It will be simpler, however, to let 0ji ~ 0 be the share of firm j owned by consumer i whatever the state. Of course 2: j OJ; = I for every i. 2. The discussion in Section 6.E was for L = 1. It extends straightforwardly to the current case of L ~ I.
1. As usual, a negative entry is understood as an obligation to deliver.
DESCRIPTION
689
~-----------------------------------------=
3. A similar point could be made on the consumption side. If, for a particular commodity I. any vector Xi E Xi is such that all entries X ,Jj , S = I •...• S, 3fC equal, then we can interpret this as asserting that the consumption of I takes place before the resolution of uncertainty.
SEC T ION
1 •• C:
A R ROW - DEB R E U
E QUI LIB R I U M
691
-------------------------------------------------------- -------------------------------------------------------690
CHAPTER
11:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
Example 19.B3: Consider the tree in Figure 19.B.2. We have fl'o =
«( 1.2, 3, 4, S,6}),
fI', = «( I, 2}, (3}, (4, 5, 6}), fI', = «(I}, (2}, (3}. (4}. {5}, {6}} . •
Flgur. 19.B.1 ,~O
1=
Two periods. Perfeci information at I = 1.
I
Illformalion and the Resolution of Uncertainty In the setting just described, time plays no explicit formal role. In reality, however, states of the world unfold over time. Figure 19.B.I captures the simplest example. In the figure, we have a period 0 in which there is no information whatsoever on the true state of the world and a period I in which this information has been completely revealed. We have already seen (Example 19.B.2) how, by conveniently defining consumption and production sets, we can accomodate within our setup the temporal structure of Figure 19.B.I: a commodity that has as part of its physical description its availability at I = 0 should never appear in differing amounts across states. The same methodology can be used to incorporate into the formalism a much more general temporal structure. Suppose we have T + I dates I = 0, I, ... , T and, as before, S states, but assume that the states emerge gradually through a Irte, as in Figure 19.B.2. These trees are
The partitions could in principle be different across individuals. However, except in the last section of this chapter (Section 19.H), we shall assume that the information structure is the same for all consumers. A pair (I. E) where I is a date and E e Yo is called a dale-event. Date..,vents are associated with the nodes of the tree. Each date-event except the first has a unique predecessor, and each date-evenl not at the end of the tree has one or more successors. With this temporal modeling it is now necessary to be explicit about the time at which a physical commodity is available. Suppose there is a number H of basic physical commodities (bread. leisure. etc.). We will use the double index hI to indicate the time at which a commodity Ia is produced. appears as endowment. or is available for consumption. Then x ... stands for an amount of the physical commodity h available at time I along the path of state s. Fortunately. this multiperiod model can be formally reduced to the timeless structure introduced above. To see this. we define a new set of L = H(T + I) physical commodities. each of them being one of these double-indexed (i.e .• hi) commodities. We then say that a vector Z e R's is lIIea.,urahie with respect to the family of information partitions (fI'..... , fl'T) if. for every /I/S and /1/.,'. we have that z". = =.... whenever s.s· belong to the same clement of the partition .'1;. That is. whenever sand s' cannot be distinguished at time I. the amounts assigned to the two states cannot be different. Finally. we impose on endowments w, e RI.<. consumption sets Xi c R'-' and production sets )j c RI.< the restriction that all their clements be measurable with respect to the family of information partitions. With this, we have reduced the multiperiod structure to our original formulation.
19.C Arrow-Debreu Equilibrium
Flgur. 19.B.2
I ~
0
I ~
I
I ~
2
similar to those described in Chapter 7. Here final nodes stand for the possible states realized by time I = T. that is, for complete histories of the uncertain environment. When the path through the tree coincides for two states, sand s', up to time I, this means that in all periods up to and including period I, sand s' cannot be distinguished. Subsets of S are called even/so A collection of events fI' is an information s/ruelur. if it is a partition, that is, iffor every state s there is E e fI' with SeE and for any two E,E' e fI', E ~ E'. we have E r. E' = 0. The interpretation is that if sand s' belong to the same event in fI' then sand s· cannot be distinguished in the information structure fI'. To capture formally a situation with sequential revelation of information we look at a family of information structures: (fI'o, ... ,Yo, .. . , fl'T)' The process of information revelation makes the Yo increasingly fine: once one has information sufficient to distinguish between two states. the information is not forgotten.
An inrormation tree: gradual release of information.
We have seen in Section 19.B how an economy where uncertainty matters can be described by means of a set of states of the world S, a consumption set XI c R LS, an endowment vector Wi E RLS , and a preference relation ;::i on X, for every consumer i. together with a production set lj c RLS and profit shares (0)1" .• , OJ!) for every firm j. We now go a step further and make a strong assumption. Namely, we postulate the existence of a market for every contingent commodity Is. These markets open hefore the resolution of uncertainty. at date 0 we could say. The price of the commodity is denoted Pt.. What is being purchased (or sold) in the market for the contingent commodity Is is commitments to receive (or to deliver) amounts of the physical good I if. and when, stale of the world s occurs. Observe that although deliveries arc contingent, the payments are not. Notice also that for this market to be well defined it is indispensable that all economic agents be able to recognize the occurrence of S. That is, information should be symmetric across economic agents. This informational issue will be discussed further in Section 19.H. Formally. the market economy just described is nothing but a particular case of the economies we have studied in previous chapters. We can. therefore, apply to our market economy the concept of Walrasian equilibrium and, with it, all the theory
692
CHAPTER
11:
GENERAL
EGUILIBRIUM
UNDER
UNCERTAINTY
developed so far. When dealing with contingent commodities it is customary to call the Walrasian equilibrium an Arrow-Debreu equilibrium. 4
-
SECllON
II.C:
ARROW-DEBREU
EOUILIBRIUM
693
~--+-~------------~~02
Definition 19.C.1: An allocation
(xT, ... , xr, yT, ... , y,) E X, x ... X XI X Y, X ••• x ~ c RLS(I +J) and a system of prices for the contingent commodities P = (Pl1' ... ,PLS) E R LS constitute an Arrow-Debreu equI1ibrium if:
"II
(i) For every j, Yi satisfies p'yi ~ P' YJ for all YjE)j. (ii) For every i, x7 is maximal for ~; in the budget set {x; E X;: P'x;:s; P'w;
(iii) LI x7 = LJ Yi
""
l-::f-=--.-,,_
+ '4 O;jP'yj}.
same probability
+ L; Wi'
The welfare and positive theorems of Chapters 16 and 17 apply without modification to the Arrow-Debreu equilibrium. Recall from Chapter 6, especially Sections 6.C and 6.E, that, in the present context, the convexity assumption takes on an interpretation in terms of risk aversion. For example, in the expected utility setting of Example 19.8.1, the preference relation ~, is convex if the Bernoulli utilities u,,(x,,) are concave (see Exercise 19.C.1). The Pareto optimality implication of Arrow-Debreu equilibrium says, effectively, that the possibility of trading in contingent commodities leads, at equilibrium, to an efficient allocation of risk. It is important to realize that at any production plan the profit of a firm, p' YJ' is a nonrandom amount of dollars. Productions and deliveries of goods do, of course, depend on the state of the world, but the firm is active in all the contingent markets and manages, so to speak, to insure completely. This has important implications for the justification of profit maximization as the objective of the firm. We will discuss this point further in Section 19.G. Example 19.C.l: Consider an exchange economy with 1 = 2, L = I, and S = 2. This lends itself to an Edgeworth box representation because there are precisely two contingent commodities. In Figures 19.C.1(a) and 19.C.I(b) we have WI = (I, 0), W2 = (0, I), and utility functions of the form X II u,(x II) + X 2l U;(X,,), where (XI" x,;) are the subjective probabilities of consumer i for the two states. Since WI + ClJl = (I, I) there is no aggregate uncertainty, and the state of the world determines only which consumer receives the endowment of the consumption good. Recall from Section 6.E (especially the discussion preceding Example 6.E.I) that for this model [in which the U,(') do not depend on s], the marginal rate of substitution of consumer i at any point where the consumption is the same in the two states equals the probability ratio X II/Xl/' In Figure I 9.C.1 (a) the subjective probabilities are the same for the two consumers (i.e., X II = X 11) and therefore the Pareto set coincides with the diagonal of the box (the box is a square and so the diagonal coincides with the 45-degree line, where the marginal rates of substitution for the two consumers are equal: X II /X21 = XIl/X,,), Hence, at equilibrium, the two consumers insure completely; that is, consumer i's equilibrium consumption does not vary across the two states. In Figure 19.C.I(b)
~=~ (a)
1[21
w
w
1[22
(b)
assessments. (b) No aggregale risk:
different probability assessments.
the consumer's subjective probabilities are different. In particular, XII < X ,2 (i.e., the second consumer gives more probability to state I). In this case, each consumer's equilibrium consumption is higher in the state he thinks comparatively more likely (relative to the beliefs of the other consumer). _ Example 19.C.2: The basic framework is as in Example 19.G.1. The difference is that now there is aggregate risk: W, + ClJl = (2, I). The utilities are state inde. pendent and the probability assessments are the same for the two traders: (x" x 2 ). The corresponding Edgeworth box is represented in Figure 19.C.2. We see that at any point of the Pareto set the common marginal rate of substitution is smaller than the ratio of probabilities (see Exercise 19.C.2). Hence at an equilibrium we must have PI/P2 < 1t , /X" or p,/x, < P2/X" If, say, 1t, = X, = t, then p, < P2: The price of one contingent unit of consumption is larger for the state for which the consumption good is scarcer. This constitutes the simplest version of a powerful theme of finance theory: that contingent instruments (in our case, a unit of contingent consumption) arc comparatively more valuable if their returns (in our case, the amount of consumption they give in the different states) are negatively correlated with the "market return" (in our case, the random variable representing the aggregate initial endowment). _
Figure '9.C.2
There is aggregate risk: P';"r negatively
correlated with total
---
endowment of
commodity t.
Ol~~~~,-,----/~----~----~~k_~_~_~
~ 4. See Chapter 7 or Debreu (\959) ror a succinct development or these ideas.
Figure '9.C.'
(a) No aggregale risk:
1[2
_2,/
2
...
~J:_:::::. pz
--
694
CHAPTER
II:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
19.D Sequential Trade The Arrow-Debreu framework provides a remarkable illustration of the power of general equilibrium theory. Yet, it is hardly realistic. Indeed, at an Arrow- Debreu equilibrium all trade takes place simultaneously and before the uncertainty is resolved. Trade, so to speak, is a one-shot affair. In reality, however, trade takes place 10 a large extent sequentially over time, and frequently as a consequence of information disclosures. The aim of this section is to introduce a first model of sequential trade and show that Arrow-Debreu equilibria can be reinterpreted by means of trading processes that actually unfold through time. To be as simple as possible we consider only exchange economies (see Section 19.G for some discussion of production). In addition, we take X, = R';:' for every i. To begin with. we assume that there are two dates, t = 0 and t = I, that there is no information whatsoever at t = 0, and that the uncertainty has resolved completely at r = 1. Thus, the date-event tree is as in Figure 19.B.1. Again for simplicity, we assume that there is no consumption at t = O. (We refer to Exercise 19.D.3 for the more general situation.) Suppose that markets for the LS possible contingent commodities are set up at r = O. and that (x! • ...• e RLSI is an Arrow-Debreu equilibrium allocation with prices (PIl •...• PLS)eR LS . Recall that these markets are for delivery of goods at t = 1 (they are commonly called forward markels). When period r = 1 arrives. a state of the world s is revealed. contracts are executed. and every consumer i receives x;; = (xT .... ..• xt,) e RL. Imagine now that, after this but before the actual consumption of markets for the L physical goods were to open at 1= 1 (these are called spot markets). Would there be any incentive to trade in these markets? The answer is "no." To see why. suppose that there were potential gains from trade among the consumers. That is. that there were x" = (X,'...... xu,) for i = 1•...• 1. such that 2:i Xsi $; ti Wjj and (x!" ... t X,;, ..• , X~,) ~i (Xril' •• , x:" ... , xtJ for all i, with at least one preference strict. It then follows from the definition of Pareto optimality that the Arrow-Debreu equilibrium allocation (x! •...• xrl e RLSI is not Pareto optimal. contradicting the conclusion of the first welfare theorem.' In summary. at r = 0 the consumers can trade directly to an overall Pareto optimal allocation; hence there is no reason for further trade to take place. In other words. ex ante Pareto optimality implies ex post Pareto optimality and thus no ex post trade. Matters are different if not all the LS contingent commodity markets are available at r = O. Then the initial trade to a Pareto optimal allocation may not be feasible and it is quite possible that ex post (i.e.• after the revelation of the state s) the resulting consumption allocation is not Pareto optimal. There would then be an incentive to reopen the markets and retrade.
xn
x;;.
5. Alternatively. consider the Arrow-Debreu equilibrium prices for the L contingent commodities corresponding to state s: P, = (Ph •...• p",,). Then P•• viewed as a system of spot prices at s, induces. for the initial endowment vector (x:I' ...• x:,), a null excess demand for all traders (and Iherefore clears markets). Indeed. if U,(x" ....• x.d is a utililY function for ;::, and (x~i" '. x;,) E RLS maximizes U~X1i"'" XSj) subject to L. P,'(Xd - w,,,) so, then, for any particular s, Xi~ maximizes Ui(xri•...• X,i ••..• x;) subject to P,'(X'i - W,I} S PJ'(x~ - w,.), that is, subject to PJ'(x" - x:j ) S O.
& E C T ION
" . 0:
a E QUE N T I A L
A most interesting possibility, first observed by Arrow (1953). is that. even if not all the contingent commodities are available at t = 0, it may still be the case under some conditions that the retrading possibilities at t = 1 guarantee that Pareto optimality is reached. nevertheless. That is. the possibility of ex post trade can make up for an absence of some ex ante markets. In what follows. we shall verify that this is the case whenever at least one physical commodity can be traded contingently at I = 0 if. in addition. spot markets occur at I = 1 and the spot equilibrium prices are correctly anticipated at I = O. The intuition for this result is reasonably straightforward: if spot trade can occur within each state. then the only task remaining at I = 0 is to transfer the consumer's overall purchasing power efficiently across states. This can be accomplished using contingent trade in a single commodity. By such a procedure we are able to reduce the number of required forward markets for LS to S. Let us be more specific. At I = 0 consumers have expectations regarding the spot prices prevailing at I = 1 for each possible state s e S. Denote the price vector expected to prevail in state s spot market by p, e RL. and the overall expectation vector6 by P = (Pl.···. Ps) e RLS. Suppose that. in addition. at date 1=0 there is trade in the S contingent commodities denoted by II to I S; that is. there is contingent trade only in the physical good with the label 1. We denote the vector of prices for these contingent commodities traded at t = 0 by q = (ql' ...• qs) e RS. Faced with prices q e RS at t = 0 and expected spot prices (PI>' ..• Ps) e RLS at I = I. every consumer i formulates a consumption. or trading. plan (z 1/ ••••• ZSi) e RS for contingent commodities at I = O. as well as a set of spot market consumption plans (x 1/ •••• , x,,) E RLS for the different states that may occur at t = I. Of course, these plans must satisfy a budget constraint. Let U,(') be a utility function for ;:::,. Then the problem of consumer i can be expressed formally as Max
(19.D.1)
(xl •• ··· • .xs.)ER~.s
(:u •. ·· ••s.)ERS
s.t.
(i)
L, q'Z,i
~
O.
(ii) p, x" ~ p,W"
+ p"Z"
for every s.
Restriction (i) is the budget constraint corresponding to trade at I = O. The family of restrictions (ii) are the budget constraints for the different spot markets. Note that the value of wealth at a state s is composed of two parts: the market value of the initial endowments. p, ·W". and the market value of the amounts z,' of good 1 bought or sold forward at t = O. Observe that we are not imposing any restriction on the sign or the magnitude of z". If z" < - W"i then one says that at t = 0 consumer i is selling good 1 shorl. This is because he is selling at t = 0, contingent on state s occurring. more than he has at I = I if s occurs. Hence, if s occurs he will actually have to buy in the spot market the extra amount of the first good required for the fulfillment of his commitments. The possibility of selling short is. however. indirectly 6. In principle, expectations could differ across consumers. but under the assumption of correct expeclations (soon 10 be introduced) they wiU nol.
TRADE
695
696
CHAPTER
11:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
SEC TID N
limited by the fact that consumption, and therefore ex post wealth, must be nonnegative for every s.' To define an appropriate notion of sequential trade we shall impose a key condition: Consumers' expectations must be self-fuljilled, or rational; that is, we require that consumers' expectations of the prices that will clear the spot markets for the different states s do actually clear them once date t = I has arrived and a state s is revealed. Definition 19.0.1: A collection formed by a price vector q = (q, •.. . ,qs) E contingent first good commodities at t = O. a spot price vector
W for
for every s. and. for every consumer i. consumption plans zi = (zT;•...• z$;) E RS at t = 0 and xi = (xT; . .... x$;) E IRLS at t = 1 constitutes a Radner equilibrium [see Radner (1982)] if:
1 9 • 0:
SEQ U E N T I • L
Proposition 19.0.1: We have: (i) If the allocation x· E R lSI and the contingent commodities price vector (P" ...• PsI E R~S+ constitute an Arrow-Oebreu equilibrium. then there are prices q E R~ + for contingent first good commodities and consumption plans for these commodities z· = (zT ..... zrl E RSI such that the consumptions plans x'. z·. the prices q. and the spot prices (p, • ...• PsI constitute a Radner equilibrium. (ii) Conversely. if the consumption plans x· E nlSI. z· E R SI and prices q E A~ +. (p, •...• PsI E n~s+ constitute a Radner equilibrium. then there are multipliers (II, •. ..• lIS) E n~ + such that the allocation x' and the contingent commodities price vector (II,P, •. .•• jlsPs) E R~s+ constitute an Arrow-Oebreu equilibrium. (The multiplier is interpreted as the value. at t = O. of a dollar at t = 1 and state s.)
I'.
Proof: (i) It is natural to let q, = P" for every s. With this we claim that. for every consumer i. the budget set of the Arrow-Oebreu problem. BtO = {(Xli' ... ' xS/) E R,!:
L, p,'(x"
- w,,) SO}.
is identical to the budget set of the Radner problem, (i) For every i. the consumption plans
zi. xi solve problem (19.0.1).
Br =
{(x 1/ ••••• x.,,)
E
IR'! : there are (z li • . . . • zs.> such that L. q,z" S 0 and P,(X,,- w,,) S P"Z" for every s}.
see this. suppose that X, = (Xli •...• X5') E Bto. For every s. denote (I/PI')p,·(x" - w,,). Then L,q,z" = L,PI'Z" = L,P,'(X" - w,,) s 0 and P.,·(x" - W,.> = P"Z" for every s. Hence, X, E Bf. Conversely. suppose that x,=(x" •...• xs,)EBr; that is. for some (z" •...• zs,) we have L,q,z" SO and p,(x" - w,,) S P"Z" for every s. Summing over s gives L,P,·(X" - w,,) S LJPtsZ:wi = Lsqszsi S O. Hence, Xi e Bto. We conclude that our Arrow-Oebreu equilibrium allocation is also a Radner equilibrium allocation supported by q = (Pi' •...• PIS) E RS• the spot prices (P, ....• PsI, and the contingent trades (zr, ..... zt,) E RS defined by =.;. = (I/p,,)p,·(x:' - w,,). Note that the contingent markets clear since. for every s. To
At a Radner equilibrium, trade takes place through time and, in contrast to the Arrow-Oebreu setting. economic agents face a sequence of budget sets, one at each date-state (more generally, at every date-event). We can see from an examination of problem (19.0.1) that all the budget constraints are homogeneous of degree zero with respect to prices. This means that the budget sets remain unaltered if the price of one physical commodity in each date-state (that is, one price for every budget set) is arbitrarily normalized to equal I. It is natural to choose the first commodity and to put p" = I for every s, so that a unit of the s contingent commodity then pays off I dollar in state S.8 Note that this still leaves one degree of freedom, that corresponding to the forward trades at date 0 (so we could put q, = I. or perhaps L, q, = I). In Proposition 19.0.1, which is the key result of this section, we show that for this model the set of Arrow-Oebreu equilibrium allocations (induced by the arrangement of one-shot trade in LS contingent commodities) and the set of Radner equilibrium allocations (induced by contingent trade in only one commodity, sequentially followed by spot trade) are identical. 7. Observe also that we have taken the wealth at I = 0 to be zero (that is. there are no initial endowments or the contingent commodities). This is simply a convention. Suppose. ror example. that we regard
Wlli~
the amount of good t available at t
:a=
1 in state s. as the amount of the s
contingent commodity that i owns at I = 0 (to avoid double counting. the initial endowment or commodity I in the spot market' at I = I should ,imultaneously be put to zero). The budget constraints are then: (i) L,q,(:~ - "'I") ~ 0 and (ii) p,·x,. ~ LI~I P"",, + Ph:;' ror every s. But letting
Z:i = tli +
W bl •
we see that these are exactly the constraints of (l9.D.1).
8. It rollows rrom the possibility or making this normalization that. without loss or generality. we could as well suppose that our contingent commodity pays directly in dollars (see Exercise 19.D.1 ror more on this).
TRA0 E
697
---------------------~~~~~~~~-=
Z" =
L.=!
=
(l/p,,)P,·[L,(X:' -
w,,» SO.
(ii) Choose I', so that jl,P" = q,. Then we can rewrite the Radner budget set of every consumer i as
Bf
=
{(x " •...• xs.> E ALS : there are (z" •...• zs,) such that L,q,z" S 0 and jl,p,(x" - W,.> s q,z" for every s}.
But from this we can proceed as we did in part (i) and rewrite the constraints. and therefore the budget set. in the Arrow-Oebreu form:
Bf = BtO =
{(XI/ •...• Xs,)E
ALS: L,jl,P,·(x" - w,,) SO}.
x:
Hence, the consumption plan is also preference maximizing in the budget set Bt"· Since this is true for every consumer i. we conclude that the price vector LS (II. P,.···. lISPS) E A clears the markets for the LS contingent commodities. _ Example 19.0.1: Consider a two-good. two-state. two-consumer pure exchange economy. Suppose that the two states are equally likely and that every consumer has the same. state-independent. Bernoulli utility function u(x,,). The consumers differ only in their initial endowments. The aggregate endowment vectors in the two states
698
CHAPTER
11:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
SECTION
Good 2
Spot Prices in the Two States
p 0, Initial Endowment in State 2
Good I Good 2
are the same; however, endowments are distributed so that consumer I gets everything in state I and consumer 2 gets everything in state 2. (See Figure 19.0.1.) By the symmetry of the problem, at an Arrow-Debreu equilibrium each consumer gets, in each state, half of the total endowment of each good. In Figure 19.0.1, we indicate how these consumptions will be reached by means of contingent trade in the first commodity and spot markets. The spot prices will be the same in the two states. The first consumer will sell an amount ex of the first good contingent on the occurrence of the first state and will in exchange buy an amount {J of the same good contingent on the second state. (You are asked to provide the details in Exercise
19.0.2.) • It is important to emphasize that, although the concept of Radner equilibrium cuts down the number of contingent commodities required to attain optimality (from LS to S), this reduction is not obtained free of charge. With the smaller number of forward contracts, the correct anticipation of future spot prices becomes crucial. Up to this point we have discussed the sequential implementation of an Arrow-Debreu equilibrium when there are two dates,9that is, for the date-eventtree of Figure 19.B.1. Except for notational complications, the same ideas carryover to a tree such as that in Figure 19.B.2 where there are T + I periods and information is released gradually. (See the small· type discussion at the end of Section \9.B for basic concepts and notation.) We would then have spot markets at every admissible date-event pair IE (i.e., those IE where E e.9';, the information partition at r). With H the set of basic physical commodities, we denote the spot prices by P,. e RH. At every rE we could also have trade for the contingent delivery of physical good I at each of the sucoessor date-events to tE. Denote by q,.(t + I, E') the price at tE of one unit of good I delivered at r + I if event E' is revealed (of course, we require E' e .9';., and E' c: E). The problem of the consumer consists of forming utility-maximizing plans by choosing, at every admissible rE, a vector of consumption of goods X,E' e R~ and, for every sucoessor (r + I, E'), a contingent trade Z,E,(I + I, E') of good I deliverable at (I + I, E'). Overall, the budget constraint to be satisfied at tE is
+
ASSET
One can then proceed to define a corresponding concept of Radner equilibrium and to show that the Arrow-Debreu equilibrium allocations for the model with H(T + I)S contingent commodity markets'· at I = 0 are the same as the Radner equilibrium allocations obtained from a model with sequential trade in which, at each date-event, consumers trade only current goods and contingent claims for delivery of good I at successor nodes. Exercises 19.D.3 and 19. D.4 discuss this topic further.
Good I
P,.· x,"
1I.E:
I
q,.(r
+ I, E')z,.,(1 + I, E')
!> P,E'W'E;
+ PUEZ,- ..... ,(t, E)
1£'.,Y',.,:E'cEI
where E - is the event at the date
I -
I predecessor to event E at
I.
9. To be as simple as possible, we have also assumed that there is no consumption at 1=0.
Figure 19.0,1 Reaching the Arrow-Debreu equilibrium by means of contingent trade in the first good only.
19.E Asset Markets Thc S contingent commodities studied in the previous section serve the purpose of transfcrring wealth across the states of the world that will be revealed in the future. Thcy arc. however, only theoretical constructs that rarely have exact counterparts in reality. Nevertheless, in reality there are assels, or securities, that to some extent perform the wealth-transferring role that we have assigned to the contingent commodities. It is therefore important 10 develop a theoretical structure that allows us to study the functioning of these asset markets. We accomplish the task in this section by extending the formal notion of a contingent commodity and then generalizing the theory of Radner equilibrium to the extended environment.'1 We begin again with the simplest situation, in which we have two dates, I = 0 and t = I, and all the information is revealed at t = I. Further, for notational simplicity we assume that consumption takes place only at t = 1. We view an asset, or, more precisely, a unit of an asset, as a title to receive either physical goods or dollars at t = I in amounts that may depend on which state occurs. '2 The payoffs of an asset are known as its returns. If the returns are in physical goods, the asset is called real (a durable piece of machinery or a futures contract for the delivery of copper would be examples). If they are in paper money, they are called financial (a government bond, for example). Mixed cases are also possible, Here we deal only with the real case and, moreover, to save on notation we assume that the returns of assets are only in amounts of physical good 1.13 It is then convenient to normalize the spot price of that good to be I in every state, so that, in effect, we are using it as numeraire. Definition 19.E,1: A unit of an asset, or security, is a title to receive an amount f. of good 1 at date t = 1 if state s occurs. An asset is therefore characterized by its feturn vectof f = (f" . .. ,fS) € R S.
10. A contingent commodity is a promise to deliver a unit of physical commodity h at date t if state" occurs. Recall from Section 19.B that the consumption sets have to be defined imbedding in them the inrormation measurability restrictions. thai is, making sure that at date t no consumption is dependent on inrormation not yet available. II. See Radner (1982) and Kreps (1979) [complemented by Marimon (1987)] for treatmenls in the spirit of this section. 12. As usual "title to receive" means "duty to deliver" ir the amount is negative. Although negative returns present no particular difficulty. we will avoid them. 11 This assumption also has an important simplifying reature: At any given slate the returns of all assets are in units of the same physical good. Therefore, the relative spot prices of the various physical goods in any given state do not affect the relative returns or the different assets in that state.
MARKETS
699
700
CHAPTER
11:
QENERAL
EQUILIBRIUM
UNDER
SECTION
II.E:
AISET
MARKETS
70t
-------------------------~~~~~~~~~
UNCERTAINTY
Example 19.E.l: Examples of assets include the following: (i) r = (1, ... , I). This asset promises the future noncontingent delivery of one unit of good 1. Its real-world counterparts are the markets for commodity futures. In the special case where there is a single consumption good (i.e., L = I). we call this asset the safe (or riskless) asset. It is important to realize that with more than one physical good a futures contract is not riskless: its return in terms of purchasing power depends on the spot prices of all the goods l 4
Dellnltlon 19.E.2: A collection formed by a price vector q = (q, • ... ,qK) E RK for assets traded at t = O. a spot price vector P. = (p, •• ...• PL.) E RL for every s. and. for every consumer i. portfolio plans zi = (zr;..... z.tl E RK at t = 0 and consumption plans xi = (Xri . ...• x~;) E R LS at t = 1 constit~tes a Radne, equilibrium if: (i) For every i. the consumption plans Max
zi. xi solve the problem
U,{X'i' .... XSi)
(Xli • ... JI's,)ER~S
(ii) r = (0•...• 0, I. 0•...• 0). This asset pays one unit of good I if and only if a certain state occurs. These were the assets considered in Section 19.0. In the current theoretical setting they are often called Arrow securities. (iii) r = (1,2. 1.2•...• 1.2). This asset pays one unit unconditionally and. in addition. another unit in even-labeled states. Example 19.E.2: Options. This is an example of a so-called derivative asset. that is. of an asset whose returns are somehow derived from the returns of another asset. Suppose there is a primary asset with return vector r E RS• Then a (European) call option on the primary asset at the strike price c E R is itself an asset. A unit of this asset gives the option to buy. after the state is revealed (but beforc the rcturns are paid). a unit of the primary asset at price c (the price c is in units of the "numeraire." that is. of good I). What is the return vector rIc) of the option? In a given state s. the option will be exercised if and only if r, > c (we neglect the case r, = c). Hence
(Z" •...• z~,)ER ..
s.t.
(a)
L. q.·Z.i ~ 0
(b) p, X'i !> P,W'i + L. P,.z./,.
for every s.
(ii) LiZki ~ 0 and LiX:i ~ L;W.; for every k and s. In the budget set of Definition 19.E.2, the wealth of consumer i at state s is the sum of the spot value of his initial endowment and the spot value of the return of his portfolio. Note that. without loss of generality. we can put Ph = I for all s. From now on we will do so. It is convenient at this point to introduce the concept the retllrn matrix R. This is an S x K matrix whose kth column is the return vector of the kth asset. Hence. its generic sk entry is r... the return of asset k in state s. With this notation. the budget constraint of consumer i becomes
B,j p, q,
B).1 "R~
, fo,"om' portfolio ,,' R'
~ ho~
q",
rIc) = (Max {O. rl - c} •... , Max {O, rs - c}).
Pt'(XII:- WII») [rtt ...... ruj_
For a primary asset with returns r = (4,3,2. I) specific examples are r(3.5) =
(.5. 0 • 0
(
p,'(xs, -
0).
r(2.5) = (1.5.
0.5. 0
r( 1.5) = (2.5.
1.5. 0.5, 0)._
0).
We proceed to extend the analysis of Section 19.0 by assuming that there is a given set of assets. known as an asset structure, and that these assets can be freely traded at date t = O. We postpone to the next section a discussion of the important issue of the origin of the particular set of assets. Each asset k is characterized by a vector of returns r. E RS• The number of assets is K. As before. we assume that there are no initial endowments of assets and that short sales are possible. The price vector for the assets traded at t = 0 is denoted q = (ql •... , qK)' A vector of trades in these assets. denoted by z = (z I' ••• , ZK) E R K, is called a portfolio. The next step is to generalize the definition of a Radner equilibrium to the current environment. In Definition 19.E.2. U;(·) is a utility function for the preferences ~; of consumer i over consumption plans xs;) R~s.
(XII .... '
~
. WS,)
'.
<0
~d
I
z, - Rz,
rs t , · · · , rSK
We now present a very important implication, rich in ramifications. of the assumption that unlimited short sales are possible. Namely, we will establish that knowledge of the return matrix R suffices to place significant restrictions on the asset price vector q = (q" . .. ,q,,) that could conceivably arise at equilibrium. Proposition 19.E.1: Assume that every return vector is nonnegative and nonzero' that is, '. ~ 0 and 'k '# 0 for all k.'5 Then. for every (column) vector q E RK of ~sset prices arising in a Radner equilibrium. we can find multipliers Jl = (Jl, • ...• Jls) ~ D. such that q. = L.Jl.'•• for all k (in matrix notation. qT = I"R). In words. Proposition 19.E.1 says that we can assign values (Jlt ..... Jls) to units of wealth in the different states so that the price of a unit of asset k is simply equal to the sum. in value terms. of the returns across states.'6 Because this is an important
E
14. Strictly speaking. for the term "riskless" to be meaningful we need. in addition to L that utility functions be uniform across states.
= I.
15. This assumption can be weakened substantially_ 16. As we shall see shortly. the value 1'. can also be interpreted as the implicit price the stale-contingent commodity that pays one unit of good 1 if state s occurs and nothing otherwise.
or
-------------------
702
c HAP T E R
1 I:
G ENE R ALE QUI LIB R I U..
SEC T ION
UNO E RUN C E R T .. I N T Y
1 •• E:
.. S • E T
--------------------------------------------------------------------------------We now argue Ihallhe row veclor qT must be proportional 10 the row veclor p" R e R". The entries of p' and of R are all nonnegalive and no row of R is null. Therefore p" R ~ OT and II'·R '" OT. If qT is nol proportional 10 II"R Ihen we can find ie R~ such Ihal q'i-O and p" Ri > 0 [see Figure 19.E.I(b)). BUlletting • = Ri. we would Ihen have .e V and p' •• '" 0, which we have jusl seen cannol happen. Hence qT musl be proportional to p" R; Ihal is. q T = all'. R for some real number a > O. Letting II = all', we have the conclusion of Ihc lemma. _
result. we shall give two proofs of it. The first. which we give in small type. is based on convexity theory and uses only one implication of equilibrium: the fact that q must be arbitrage free (we will give a definition of this concept shortly). The second proof uses the first-order conditions of the utility maximization problem and provides further insight into the nature of the multipliers. Proof I of Proposilion 19.E.l: Callihe syslem q E RK of assel prices arbilrage free if Ihere is no portfolio z = (z, •...• ZK) such thai q-z $ O. Rz ~ O. and Rz '" O. In words. there is no portfolio that is budgetarily feasible and that yields a nonnegative return in ewry state and a strictly positive return in some state. Nole that whether an asset price vector is arbitrage free or nol depends only on the returns of the assets and not on preferences. If, as usual, we assume thai preferences are strongly monotone, then an equilibrium asset price vector q E R" must be arbitrage free: if it were not, it would be possible to increase utility merely by adding to any currenl portfolio a portfolio yielding an arbilrage opportunity. Because there are no restrictions on short sales this addition is always feasible. In Lemma 19.E.1 we establish a resull which in view of the observation just made is formally stronger than the stalement of Proposition 19.E.1.
As we have already argued, if short sales of assets are possible and preferences are slrongly monotone (e.g., if preferences admit an expected utility representalion wilh sirictly posilive subjective probabililies for the stales), Ihen equibrium assel prices musl be arbitrage free and. therefore, Proposition 19.E.I follows from Lemma 19.E.1. _
Proof 2 of Proposition 19.E.l: For this proof we assume that preferences are represented by utility functions of the expected utility form ~(x H' •••• xs,) = L.71.,u.,(x.,) and that the Bernoulli utilities u.,(·) are concave. strictly increasing. and differentiable. We denote by v.,(P .. w.,) the indireci utility function [derived from u.,( . )] of consumer i in state s. Suppose that in the Radner equilibrium with asset prices q = (q" ...• qK) the equilibrium spot prices are P = (Pl •...• Ps) E RLS. Because unlimited short sales are possible, the optimal portfolio choice z i E RK of any consumer i is necessarily interior and. denoting w:' = P. + L. it must satisfy the following first-order conditions for some '" > 0:
Lemma 19.E.l: If the asset price vector q E RK is arbitrage free, then there is a vector of multipliers /' = (/""'" lis) ~ 0 satisfying qT = wR. Proof of Lemma 19.E.I: Note to begin with Ihal since we deal only with assets having nonnegative, nonzero returns, an arbitrage-free price vector q must have q. > 0 for every k. Also, without loss of generality. we assume thai no row of the relurn malrix R has all of its entries equal to zero," Given an arbitrage-free assel price veclor q E RK. consider Ihe convex set V= (UE
R':. = Rz for some ZE R" wilh
q'Z =
·w.,
r.. z:,.
for every k = I •...• K.
Ol·
That is, the vector of expected marginal utilities of the K assets must be proportional to the vector of asset prices. 18 With this we have attained our result. since by taking
The arbitrage freeness of q implies Ihal V () (R~ \(Oll = 0. Since both V and R~ \(Ol are convex sets and the origin belongs 10 V. we can apply the separating hyperplane theorem (see Section M.G of the Mathematical Appendix) to oblain a nonzero vector II' = (11', •...• lis) such that 11". $ 0 for any UE V and 1"'. ~ 0 lor any .E Note Ihal il must be Ihalll' ~ O. Moreover,because.E V implies -UE V. it follows Ihatll'·. = ofor any uE V. Figure 19.E.I(a) depicts this construction for Ihe Iwo-slale case.
w..
Figure 19.E.l
Vo
V = : 1': I' = R:. q': = O. :
E R<}
(a) Construction of the no-arbitrage weights. (b) Existence of an inadmissible i if qT is not proportional
R'
we have qj = L. /1.,r••. Hence. we could determine /1 = (/1" ...• /1.) by choosing any consumer i and letting /1. = /1". the marginal utility of wealth at state s of consumer i weighted by " •.1",. The multiplier '" is the Lagrange multiplier of the budget constraint at t = 0 and can therefore be viewed as the marginal utility of wealth at t = O. Hence, for any consumer i. /1., equals the ratio of the (expected) utility at t = 0 of one extra unit of wealth at t = 1 and state s. and the utility of one extra unit of wealth at t = O. See Exercise 19.E.I for more on this point. Note also that different consumers may lead to different p, = (/1" •...• /1,.) and therefore to different /1·s. The uniqueness of /1 is assured only when rank R = S. _
10Il'·R. (0)
(b)
17. U there is such a row, set Il, arbitrarily for the state s corresponding to that row, drop s from the list of stales. and proceed with the remaining slates.
18. Recall that we always take Ph = 1. Therefore, r... is the extra amount of wealth in state s
derived from an extra unit of asset k. The proportionality factor a, is the Lagrange multiplier of the problem
S.t.
I:. q.z" ,;; O.
.... R K E T S
703
704
CHAPTER
11:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
--------------------------------------------------------------------------------------------Example 19.E.3: Suppose that there is available an asset with noncontingent returns; for example, rl = (I, ... , I). Normalize the price of this asset to be I, that is, ql = I. Then if I' = (1'1"'" I's) is the vector of multipliers given by Proposition 19.E.I, we must have I' ~ 0 and L,I', = wrl = ql = I. For any other asset k we then obtain the intuitive conclusion that q. = L,I',r,. ~ Min, r.. and, similarly, q. :S: Max, r... I n Section 19.D, we proved that for the set of assets consisting of the S contingent markets in a single physical commodity we have an equivalence result between Arrow-Debreu and Radner equilibrium allocations (Proposition 19.D.I). We now generalize this result. In particular, we show that this equivalence holds for any family of S or more assets, provided that at least S of them have returns that are linearly independent (i.e., provided the effective number of assets is at least S). We begin with Definition 19.E.2. Definition 19.E.3: An asset structure with an S x K return matrix R is complete if rank R = S. that is. if there is some subset of S assets with linearly independent returns. Example 19.E.4: In the case of S contingent commodities discussed in Section 19.D, and also in Example 19.E.1 (ii), the return matrix R is the S x S identity matrix. This is the canonical example of complete markets. But there are many other ways for a matrix to be nonsingular. Thus, with three states and three assets, we could have the return matrix
-
SECTION
Example 19.E.5: Spallnillg through Options. Suppose that S = 4 and there is a primary asset with returns r = (4.3.2, I). We have seen in Example 19.E.2 that, for every strike price e, the option defined by e constitutes an asset with return vector rIc) = (Max {O, r l - c}, ... , Max {O. r. - e}). Using options we can create a complete asset structure supported entirely on the primary asset r. For example, the return vectors r(3.5), r(2.5), r( 1.5), and r and linearly independent (the matrix R has all its entries below the diagonal equal to zero). Thus, the asset structure consisting of the primary asset plus three options with strike prices 3.5, 2.5. and 1.5 is complete. More generally, whenever the primary asset is such that r, r,' for all 5 s', it is possible to generate a complete asset structure by means of options (see Exercise 19.E.2). If r, = r,. for some distinct 5 and 5'. then we cannot do so: If the primary asset does not distinguish between two states. no derived asset can do so either. -
*
(i) If the consumption plans x· = (xT •...• xT) E RLS1 and the price vector (p, •...• Ps)ER~S+
constitute an Arrow-Oebreu equilibrium. then there are asset prices q E IR~ + and portfolio plans z· = (zT •. .•• zT) E RKI such that the consumption plans portfolio plans asset prices q. and spot prices (p, • ...• PsI constitute a Radner equilibrium. (ii) Conversely. if the consumption plans x· E RLS1 • portfolio plans z· E IRKI. and prices q E R~ +. (p, • .... PsI E R~S+ constitute a Radner equilibrium. then there are multipliers (1', .... , lis) E R~ + such that the consumption plans x· and the contingent commodities price vector (I',P, •...• l'sPs) E R LS
x·.
z·.
constitute an Arrow-Oebreu equilibrium. (The multiplier 1'. is Interpreted as the value. at t = O. of a dollar at t = 1 and state 5; recall that p,. = 1.) Proor: It is entirely similar to the proor or Proposition 19.0.1. (i) Define q, = L. p,.r.. ror every k. Denote by A the S x S diagonal matrix whose s diagonal entry is p". Then q T = e' AR. where e e RS is a column vector with all its entries equal 10 I. For every i the (column) vector or wealth transrers across states (at the Arrow- Debreu equilibrium) is '"I = (PI "(xj/ -
"m,
WII)'· ..•
PS'(xti - WSj»T,
m,
We have = 0 for every i and L, = O. By completeness, rank AR = Sand, thererore, we can find vectors z~ E RK such that m, = ARz~ ror i = I.... , I - I. Letting = -(Ii
+ ... + :1-1)
we also have m t = -(ml + ... + mt_,) = ARzt. Thererore, ror each i. the portrolio zr allows consumer i to reach the Arrow-Debreu consumptions in the different states at the spot prices (p, •... , Ps)· To veriry budget reasibility note that q'zr = "ARzr = "m, = O. In Exercise 19. E.3 you are asked to complete the proor by showing that the consumption and portfolio plans x~ and z~ are not just budget feasible but also utility maximizing in the budget set. (ii) Assume, without loss or generality. that p" = 1 ror all s. By Propositon 19.E.1 we have q' = II" R ror some arbitrage weights I' = (1', •... , I's). We show that x· is an Arrow-Oebreu equilibrium with respect to (/" PI'" . • /'sPs). To this eliect. suppose that X, E RLS satisfies the Arrow-Oebreu single budget constraint, that is. L.I'.P' (x., - w,,) ~ O. Then by the complete· ness assumption there is z, e RK such that (PI '(XII - w II ) •...• Ps'(xs, - ws,llT = Rz, and, therefore, q'z, = jJ."Rz, ~ O. Hence x, also satisfies the budget constraints of the Radner is Arrow-Oebreu equilibrium. Observe next that the Radner equilibrium consumption budget reasible since (p, '(xT, - w ll ), ...• PS'(xI, - WS,llT ~ Rzr and qT = jJ." R yields
xr
*
To repeat, the importance of the concept of completeness derives from the fact that with it we can generalize Proposition 19.D.1. With a complete asset structure, economic agents are, in effect, unrestricted in their wealth transfers across states (except, of course, by their budget constraints). Therefore, at the equilibrium, their portfolio choices induce the same second-period consumptions as in the ArrowDebreu equilibrium. and so full Pareto optimality is reached. This is the content of Proposition 19.E.2.
ASSET
Proposition 19.E.2: Suppose that the asset structure Is complete. Then:
:1
which has rank equal to 3. the number of states. -
".E:
L/~$PJ'(x! - Wli) ~ IJ"Rzi = q'Z{
Thererore,
x~
sO.
is utility maximizing in the Arrow-Oebreu budget constraint. _
It is important to realize that in discussing Radner equilibria what matters is not so much the particular asset structure but the linear space,
Range R
= {v E IRs: v = Rz for some z E RK} C
IRs.
the set of wealth vectors that can be spanned by means of the existing assets. It is quite possible for two different asset structures to give rise to the same linear space. Our next result, Proposition 19.E.3. tells us that. whenever this is so, the set of Radner equilibrium allocations for the two asset structures is the same.
MARkETI
705
706
c
HAP T E R
I I:
G ENE R ALE G U , L , B R , U M
UNO E RUN C E R T A , N T Y
-----------------------------------------------------------------------------K
Proposition 19.E.3: Suppose that the asset price vector q e R , the spot prices P = (P", ., ,Ps) e R LS, the consumption plans x· = (xf, ... , x1) e R~s" and the portfolio plans (zf, ... ,z1) e RK' constitute a Radner equilibrium for an asset structure with S x K return matrix R. Let R' be the S x K' return matrix of a second asset structure. If Range R' = Range R, then x· is still the consumption allocation of a Radner equilibrium in the economy with the second asset structure. Proof: By Proposition 19.E.I, the asset prices satisfy the arbitrage condition q T = W R, for some JJ. E R~. Denote q' = [w R']T. We claim that if Range R = Range R' then B,(p, q', R') = B,(p, q, R)
We show that if
XI
e B,{p, q, R) then
X,
for every i.
(19.E.I)
e B,(p, q', R'). To see this, let
(p, '(Xli - WII)' ••• ' Ps'(xs, - WS;)T :S Rz, and q' z, :S O. Since Range R = Range R', we can find z; e Range R' such that Rz, = R'z;. But then q"zi = WR'z; = WRzl = q'zl:S 0, and therefore we can conclude that X, E B,(p, q', R'). The converse statement [if X, e B,(p, q', R') then x, E B,(p, q, R)] is proved in exactly the same way. It follows from (19.E.I) that, for every consumer i, is preference maximal in the budget set B,(p, q', R'). To argue that the asset prices q', the spot prices p = (p" ... , Ps), and the consumption allocation x· are part of a Radner equilibrium in the economy with an asset structure having return matrix R', it suffices to find portfolios (z;, ... , zi) e RK' such that, first, LI zi = 0 and, second, for every consumer i, the vector of across-states wealth transfers
xr
ml = (p,'(x!, - w lI ), •.. , Ps'(x;, -
SEC T , 0 N
I I •E:
ASS E T
the state multipliers JJ. = (JJ." •.. , JJ.s) of Proposition 19.E.\; indeed, for this we just have to solve a linear system of S independent equations in S unknowns. These mUltipliers can be interpreted as the (arbitrage) prices of the Arrow securities [Example 19.E.I(ii)]. Once we have these multipliers, we can obtain the price of any other asset k with return vector r. as q. = L. JJ..r••. Example 19.E.6: Pricing an Option. Suppose that, with S = 2, there is an asset with uncontingent returns, say r, = (I, I) and a second asset r, = (3 + IX, I - IX), with IX> O. The asset prices are q, = I and q,. We now consider an option on the second asset that has strike price c E (I, 3). Then r (.) - (3 2 C
-
+
. 0) _ 3 + IX
rx - ,~
- c (I - 1X)(3 + IX - ~ '2 2 + 2«
-
c) '1.
Therefore, the arbitrage price of the option (the only price compatible with equilibrium in the asset market) must be 3+IX-C
q,(c) = -_._- [q, - (I - IX)].
2 + 21X
(19.E.2)
An equivalent way to get the same formula is to note that, since the 2 x 2 return matrix R is nonsingular, the multipliers JJ. = (JJ." JJ.,) of Proposition 19.E.I can be determined uniquely from (I,q,) = WR. They are JJ., = (q, - (I -1X»/(2 + 21X) and JJ., = I - JJ.,. But, again from Proposition 19.E.l, we have q,(c) = Wr,(c) = JJ.,(3
+ IX -
c),
which is precisely expression (19.E.2). WSI»T
satisfies In, = R'zi. This is simple to accomplish. By strong monotonicity of preferences we have m, = Rzf, for every i. Hence, m, e Range R and therefore m, e Range R' for every i. Choose then z'" ... , zi-, such that m, = R'zi for every i = I, ... , I - I. Finally, let zi = -z', _ ... - zi-,. Then L, zi = 0 and also
m, = -(m, + ... + m,_,) = -R'(z', + ... + zi-,) = R'zi .• One says that an asset is redundant if its deletion does not affect the linear space Range R of spannable wealth transfers, that is, if its return vector is a linear combination of the return vectors of the remaining assets. It follows from Proposition 19.E.3 that the set of consumption allocations obtainable as part of a Radner equilibrium is not changed by the addition or deletion of a redundant asset. Another important fact is that a redundant asset can be priced merely by knowing the matrix of returns and the prices of the other assets. Exercise 19.E.4: (Pricing by Arbitrage). Suppose that rJ = lX,r, + lX,r,. Show that at equilibrium we must have qJ = lX,q, + lX,q,. Recall that unlimited short sales are possible. (Assume also that the return vectors are nonnegative and nonzero.) An implication of Exercise 19.E.4 is that, if the asset structure is complete, then we can deduce the prices of all assets from knowing the prices of a subset formed by S of them with linearly independent returns. A related way to see this is to note that from the prices of S assets with linearly independent returns we can uniquely deduce
Note that if the prices of the two assets r, and r, are themselves arbitrage free, then we must have 3 + IX ~ q, ~ I - IX (recall Example 19.E.3). Therefore, we learn from formula (I9.E.2) that q,(c) is nonnegative, decreasing in c, and increasing in q,. We can also show that if the asset price q, stays constant but the dispersion parameter represented by IX increases, then the option becomes more valuable. Suppose, in effect, that IX' > IX and r;(c), r,(c) are the corresponding returns of the option. Then rJ = r;(c) - r,(c) is itself an asset with nonnegative returns. We can also price it by arbitrage from r, and r, to get a qJ ~ 0 (typically, qJ > 0). But then, again by arbitrage, qi(c) = qJ + q,(c) ~ q,(c). • Everything generalizes to the case of T + 1 periods and gradual release of information. With several periods, an asset can take many forms. For example, we could have Shorl-Ierm assels available for trade in a given period and having positive returns only in the next period. Or ."'.e could have long-Ierm assets available at t = 0, tradeable at every period, and having pos,lIve returns only at the final period I = T. And, of course, there could be mixed cases of assets tradeable in a subset of periods and providing returns also in a subset of periods (not necessarily the same). Again, the equivalence result of Proposition 19.0.1 generalizes if the asset structure is complele. Suppose, for example, that our asset structure is composed only of collections of short-term assets available and tradeable at any admissible date-event pair tE and paying
cont,ngent amounts of physical good I at the immediately succeeding date-event pairs. Denote by S(lE) the number of successors at tEo If the number of assets available at tE is K(IE) we can view the return matrix at tE as an S(tE) x K(IE) matrix R(lE). The completeness condition is then the requirement that rank R(lE) = S(lE) for all admissible date-events pairs tEo In
MAR K E T S
707
708
C HAP T E R
,.:
G ENE R ALE QUI LIB R I U M
UNO E RUN C E R T A I N T Y
SECTION
In this section we explore the implications of having fewer than S assets, that is, of having an asset structure that is necessarily incomplete. We pursue this point in the two-period framework of the previous sections. 22
Example 19.E.7: Suppose that T = 2 and the date-event tree unfolds as in Figure 19.E.2. In
Figure 19.E.2
Construction of
Arrow- Debrcu prices from the equilibrium
o
values of the asset
q..
= i·trll l + i·fJ1ol. i.,. + i'e = I
the figure we have seven admissible date-events corresponding to the four terminal nodes. or slates, to the initial node. denoted a, and to two intermediate nodes, denoted band ('. In parlicular, there are no more than two branches from any node. In this case we claim that, typically, two long-lived assets should suffice to guarantee that the Radner and the ArrowDebreu equilibrium consumption allocations are the same.'· Suppose, to take a simple instance that L = I and that our two assets have return vectors r, = (I, I. I, I) and r, = (0, ;,0, I), payable at the terminal nodes. Consumption takes place only at the terminal nodes, but the assets may be traded both at the node a corresponding to I = 0 and at the nodes band c correponding to I = I. We can normalize the price of the first asset (as well as the price of final consumption) to be I at every node. 20 Denote by q" q., and q, the prices of the second asset at the respective nodes. By arbitrage (Proposition 19.E.I), applied at I = I, there are (1',,1',) :!: 0 such that 1', + 1', = I, I', = q., and (I'l' 1'.) :!: 0 such that I'l + I'. = I and I'. = q,. Again by arbitrage (applied this time at I = 0), we must have (A., i.,) :!: 0 such that i.• + i., = I and q. = A.q. + i.,q, = i.• + ;',1' •. This suggests considering the following Arrow- Debreu prices;
I',
p = (i·"I'I' i·,,111' ;·(11l'
i·
t
I14).
In Exercise 19.E.5 you are asked to show that, under the weak condition q. '" q" the set of final consumptions achievable through sequential trade with asset prices (q., q., q,) is indeed the same as the set of final consumptions achievable with the four Arrow-Debreu contingent commodities at prices p.2I •
19. We say "typically" because the existence of two assets is a necessary condition for
completeness at every node but, strictly speaking, not a sufficient condition. 20. Note thai, as should be the case, we do not normalize more than one price per budget constraint.
21. Incidentally, the assets prices provide a specific instance of what is called the marlilll/a/e properly oj assel prices: at any node the price of an asset is the conditional expectation of the final returns, where the expectation is taken with respect to some probabilities, in our case the ArrowDebreu prices.
INCOMPLErE
19,F Incomplete Markets
Section 19.D, the matrices R(IE) were identity matrices, and so the asset structure there was complete. But, to repeat, the results of Section 19.D generalize to the complete, nondiagonal case. A very interesting, and new, phenomenon is that if assets are long lived, and therefore repeatedly tradeable as information is gradually disclosed, it may be possible to implement the Arrow-Debreu equilibrium with much fewer than S assets. This is illustrated in Example 19.E.7.
I
".F:
prices of IwO sequenlially traded assets.
Markets may fail to exist for a number of reasons. One class of reasons refers to the informational asymmetries to be covered in Section 19.H: contracts for the delivery of goods can only be made contingent on states whose occurrence can be verified to the satisfaction of all contracting parties. Another class of reasons stems from transaction costs: the availability of a market is, after all, in the nature of a public good. Yet another variety of reasons comes from enforceability constraints: a promise to deliver one unit of good is worthless if delivery cannot be enforced.21 This said, we shall not delve Curther into a theory of asset determination. We will rest content for the moment with taking the incomplete situation as a reasonable description of reality. We begin by observing that when K < Sa Radner equilibrium need not be Pareto optimal. This is not surprising: if the possibilities of transferring wealth across states are limited, then there may well be a welfare loss due to the inability to diversify risks to the extent that would be convenient. Just consider the extreme case where there are no assets at all. Example 19.F.I provides another interesting illustration of this type of failure. Example 19,F_I: Sunspots. Suppose that preferences admit an expected utility representation and that the set of states S is such that, first, the probability estimates for the dillerent states are the same across consumers (i.e., 71" = 71". = 71, for all i, i', and s) and, second, that the states do not affect the fundamentals of the economy; that is, the Bernoulli utility functions and the endowments of every consumer i are uniform across states [i.e., u,.<·) = u,(·) and w" = w, for all s). Such a set of states is called a sunspol set. The question we shall address [first posed by Cass and Shell (1983)] is whether in these circumstances the Radner equilibrium allocations can assign varying consumptions across states. An equilibrium where this happens is called a
SUlISpot equilibrium.24 Under the assumption that consumers are strictly risk averse, so that the utility functions u,(·) are strictly concave, any Pareto optimal allocation (x" ... ,x,) E IRLSI mllsl be uniform across states (or state independent); that is, for every i we must have 22. For a general and advanced Ireatmenl, see Magill and Shafer (1991). 23. An example of a silualion where enforceabilily would be helped is when we are dealing with the shares of a firm (the total endowments of this asset is, therefore, positive) and no short s
24. The lerm "sunspoIS" is old, but the current meaning is recent. In the XIX century, research on Ihe "sunspol problem" Iried 10 determine the influence on the fundamentals (e.g., on agriculture) Ihal could make an unobservable signal (sunspots) have an effeci on prices. The modern problem is to determine how an observable signal with no influence on fundamentals can nonetheless have via expectations, an effect on prices. '
MARKETS
709
710
C HAP T E R
1 9:
G E HER ALE
a
U IL
I8
R
IU M
UH DER
UHC E RT A
IH T Y SECTION
= ... = xs,.
To see this, suppose that, for every i and s, we replace the consumption bundle of consumer i in state s, x" E IR~, by the expected consumption bundle of this consumer: X, = L, It,X" E R~. The new allocation is state independent, and it is also feasible because XII
=
X 2,
= ... = x"
L .x, = L L 1I,X" I
=
I
.li
I
I
U,*(:,)=U,*(z""",ZK)=U(Lzr II: +W
By the concavity of 11,(') it follows that no consumer is worse off:
~ 1I,u,(.x,) = u,(.x,) = u.( ~ 1I,X,,) ;:0: ~ 11,11,(:<,,)
I
We have seen that Radner equilibrium allocations need not be Pareto optimal, and so, in principle, there may exist reallocations of consumption that make all consumers at least as well off, and at least one consumer strictly better off. It is important to recognize, however, that this need not imply that a welfare authority who is "as constrained in interstate transfers as the market is" can achieve a Pareto optimum. An allocation that cannot be Pareto improved by such an authority is called a constrained Pareto optimum. A more significant and reasonable welfare question to ask is, therefore, whether Radner equilibrium allocations are constrained Pareto optimal. We now address this matter. 2S
25. This is a typical instance or a second-best welfare issue. We have already encountered problems of this kind in Chapters \3 and 14, and we shall do so again in Chapter 22.
I"
otl
den.otel~he .utility ind.uced by the portfolio optima Ity IS then qUite natural.
for every i.
Because of the Pareto optimality of (x" ... , x,), the above weak inequalities must in fact be equalities; that is, u,(x,) = L, 1I,II,(X,,) for every i. But, if so, then the strict concavity of u,{·) yields x', = .x, for every s. In summary: the Pareto optimal allocation (x" ... , x,) E RLSI is state independent. From the state independence of Pareto optimal allocations and the first welfare theorem we reach the important conclusion that if a system of complete markets over the states S can be organized, then the equilibria are sunspot free, that is, consumption is uniform across states. In effect, traders wish to insure completely and have instruments to do so. It turns out, however. that if there is not a complete set of insurance opportunities, then the above conclusion does not hold true. Sunspot-free, Pareto optimal equilibria always exist Uust make the market "not pay attention" to the sunspot; see Exercise 19.F.I). But it is now possible for the consumption allocation of some Radner equilibria to depend on the state, and consequently to fail the Pareto optimality test. In such an equilibrium consumers expect different prices in different states, and their expectations end up being self-fulfilling. The simplest, and most trivial, example is when there are no assets whatsoever (K = 0). Then a system of spot prices (p" .. . ,Ps) E RI.S is a Radner equilibrium if and only if every p, is a Walrasian equilibrium price vector for the spot economy defined by {(I1,(·),w,)}i:\. Jr, as is perfectly possible, this economy admits several distinct Walrasian equilibria, then by selecting different equilibrium price vectors for different states, we obtain a sunspot equilibrium, and hence a Pareto inefficient Radner equilibrium. _
INCOMPLETE
To proceed with the analysis we need a precise description of the constrained feastble set and of the corresponding notion of constrained Pareto optimality This ;: :o~t Simply .done in th~ contex~ where there is a single commOdity per stat~, that , - I. The Important ImphcatJon of this assumption is that then the amount of ~otnsumptldonb good that any consumer i gets in the different states is entirely e ermine y the portfolio z Ind ed "<' j" e , X si = L.,II Z"j',1r + w,;" Hence. we can let
L 1t'(L x,,) ~ L 1I'(L w,) = L w,.
I S S
1'.F:
z,.
Jj, • • • ,
L "
ztjrSIr.
+ W Si )
The definition of constrained Pareto
) IRK1' . DefinItion 19.F.1: The asset allocation (z if it is feasibl r " < ' . , •... ,z~ E IS constrained Pareto optimal z' .) eRKlje., L.i Zi $ 0) and If there IS no other feaSible asset allocation ( , •...• Z, E such that
Ui(z; •...• zi);:o: Ui(z" .. .. z,) with at least one inequality strict.
for every i,
In this L = I context the utility maximization problem of consumer i becomes Max
:,ER~
U ~(z", ... , ZK') s.t. q'Z,
~
O.
Suppose that z· E IRK fo . - I I . . . , . r I - , ... , ,IS a family of solullons to these individual '" . problems, for the asset pnce vector q E IRK Then q E RK . R d · if and onl if L . 2. .. IS a a ner equlhbnum price .. . y Note that thIS has become now a perfectly conventional equlhbr.lum pro.blem Wlt~ K commodities [see Exercise 19.F.2 for a discussion of the pr~pertlehs of U. (.»). ~o It we can apply the first welfare theorem (Proposition 16 C I) .. an reac the conclUSIon of Proposition 19.F.1.
,z, $?
pro~~~~~O:e:~~d1~e~~::'O;~e~h:~~h=~~:~ :qe~:~:~u~~sO;!~~t~:'::;~~Ption 9. e
00 d
;~
tthl sense that there is no possible redistribution of assets in the ~itr~tpP::~ a eaves every consumer as well ff d I off21.2. 0 an at east one consumer strictly better
. . ~~e situation c~nsidered in Proposition 19.F.l is very particular in that once the ~:Itt~;m~~~~t. ~r~oho of a consumer i~ determined, his overall consumption is fully h . It only one consumptIon good. there are no possibilities for trade oncc t ~ state Occurs. In particular, second-period relative prices do not matter i sImp y ccause there are no such prices. Things change if there is more than on~ 26. Recall that. given ZI. the consumptions in ever sl d . consumption good in every state is formally fixed to be~. ate are etenmned. Also, the price or
:r
27. We reemphasize that the term constrained is a f O ' W individuals and states only by trade in the iven set a pnate. t8lth can be .tr~nsfer~ed across suppose that there are no assets. Then the :elfare autho;t~I~~~:osee I~OW' restnctive thiS can be,
. 2~. In OUf current discussion. all consumption takes la . po ICY tnstrume~t what~oe.ver. Simplification that does not affect the vali . ~ . ce In the second penod. ThiS IS a redistribute consumption thai takes lace. dl:~ or the P~OpOSIl1on. If the welfare authority Can also will still be constrained Pareto oPti~al. In e first period, then the Radner equilibrium allocations
MARKETS
711
712
CHAPTER
19:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
consumption good in the second period, or if there are more than two periods. Consider the two-period case with L > I: Then we cannot summarize the individual decision problem by means of an indirect utility of the asset portfolio. The relative prices expected in the second period 2 ' also matter. This substantially complicates the formulation of a notion of constrained Pareto optimality. Be that as it may, there appears not to be a useful generalization of the "constrained Pareto optimal" concept in which we could assert the constrained Pareto optimality of Radner equilibrium allocations. Example 19.F.2, due to Hart (1975), makes the point.ln it we have an economy with several Radner equilibria where two of them are Pareto ordered. That is, we have a Radner equilibrium that is Pareto dominated by another Radner equilibrium. To the extent that it seems natural to allow a welfare authority, at the very least, to select equilibria, it follows that the first equilibrium is not constrained Pareto optimal.)O Example 19.F.2: Pareto Ordered Equilibria. Let 1= 2, L = 2, and S = 2. There are no assets (K = 0). The two consumers have, as endowments. one unit of every good in every state. The utility functions are of the form 7lIlUi(Xlli' X21i) + 7l2iUi(X12i, x 2,,). Note that although the probability assessments arc different for the two consumers (these probabilities will be specified in a moment), the spot economies are identical in the two states. Suppose that this spot economy has several distinct equilibria (e.g., it could be the exchange economy in Figure 15.B.9). Let p', p" E R2 be the Walrasian prices for two of these equilibria and let Vi(P) be the spot market utility associated with Ui(', .) and the spot price vector P E R2. Suppose that v,(p') > v,(p"). By Pareto optimality in the spot market, V2(P') < V2(P·). We now define two Radner equilibria. The first has equilibrium prices (p" P2) = (p', p") E R4 and the second has (p" P2) = (p., p') E R4. Because there is no possibility of transferring wealth across states, these are indeed Radner equilibrium prices and, moreover, they are so for any probability estimates 7li' However, the expected utility of these Radner equilibria for the different consumers depends on the 7l i . We can sec now that if consumer I believes that the first state is more likely than the second, that is, he has It" > }, then he will prefer the first equilibrium to the second. Indeed, 7t1l > 1 and v,(p') > v,(p") imply 7lIlV,(P') + 7l"v,(p") > 7tIlV,(P") + 7l 2l V,(p'). Similarly, if the second consumer believes that the second state is more likely than the first, that is, he has 7122 > 1, then he will also prefer the first equilibrium to the second: 7l'2 >! and v,(p') < v2(p") imply 1t'2V2(P') + 7lnv2(P") > 7l"V2(P") + 7tnv,(p'). Thus, the Radner equilibrium with prices (p', pH) Pareto dominates the one with prices (p", p') . • The consensus emerging in the literature seems to be that failures of restricted Pareto optimality (for natural meanings of this concept) are not only possible but even typical [Geanakoplos and Polemarchakis (1986)]. In Exercise 19.F.3 you are asked to develop a related optimality paradox: it is possible for the set of assets to expand and for everybody to be worse off at the new equilibrium! We shan not pursue the constrained optimality analysis
sec T ION
1 8 • G:
FIR M
B E H A V lOR
UNO E RUN C e R T A , N T Y
in any greater depth. At some point the analysis runs into the difficulty that it is hard to proceed sensibly without tackling the difficult problem of the determination of the asset structure.
We could also analyze the positive issues studied in Chapter 17 within an incomplete market sellmg. For eXIStence, there is a new set of complexities related to the fact that unbounded short sales are possible. In some contexts this may lead to existence failures (see Exercise 19.F.4).'1 New subtleties also arise for the issue of the determinacy of equilibria (i.e., the number and local uniqueness of equilibria). As we have seen in Section 17.0, with a complete asset structure we have generic finiteness. But with incomplete markets the nature
of the assets (e.g., whether real or financial) mailers, as may the size of S.
19.G Firm Behavior in General Equilibrium Models under Uncertainty In the previous sections we have concentrated on the study of exchange economies. For once, this has not been just for simplicity. The consideration of production and firms is genuinely more difficult in a context of possibly incomplete markets. The rca son relates to the issue of the objectives of the firm.12 As before, we consider a setting with two periods, t = 0 and t = I, and S possible statcs at I = I. There are L physical commodities traded in the spot markets of period t = I and K assets traded at t = O. There is no consumption at I = O. The returns of the assets are in physical amounts of the good I (which we call the numeraire). The S x K return matrix is denoted R. We introduce into our model a firm that produces a random amount ofnumeraire at date t = I (perhaps by means of inputs used at time t = O. but we do not formalize this part explicitly). We let (a ... , as) denote the state-contingent levels of produc" tion of the firm. There are also shares ()i ~ 0, with Li (), = I, giving the proportion of the firm that belongs to consumer i. We take, for the rest of this section (except in the small-type paragraphs at the end) the natural point of view that the firm is an asset with return vector a = (a" ... , as) whose shares are tradeable in the financial markets at t = 0.)) Suppose now that the firm can actually choose, within a range, its (random) productIOn plan. Say, therefore, that there is a set A c RS of possible choices of return 31. Unbounded short sales are at the origin of a discontinuity in the dependence on asset returns of the space of attainable wealth transfers across states. No matter how close asset returns (in dollar terms) may be t~ displaying a linear dependence. consumers can plan to attain. by using trades of very large magnatude, any wealth transfer in the subspace spanned by the asset returns. But when :et~rns bcco~e exactly linearly dependent, this attainable subspace suddenly drops in dimension. As 1O~lcated. thl~ can lead to an existence failure in some COntexts. The model we have analyzed in t~IS chapter IS not. however. one of those. If. as here, in every state all assets have returns in a Single good. whic~. moreover. is the same across assets, then the discontinuity does not arise.
32. The ciass,c paper on Ihis topic is Diamond (1967). For a more recent survey see Merton 29. Or the relative prices of goods between the second and third period. if we are considering more than two dates. 30. That is. the first equilibrium is not Pareto optimal relative to any set of constrained feasible allocations that includes all Radner equilibrium allocations.
(1982).
33. A minor difference with the setting so far is that the firm does really produce the vector (a l • . . . • "s), and. therer~re the total endowment of this asset is not zero. In fact, by putting L, 0, = 1 we have normahzed thiS total endowment to be I.
713
'I,I
714
CHAPTER
19:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
SECTION
-------------------------------------------------------------------------------------------Bo; =
II.G:
FIRM
BEHAVIOR
UNDER
{(xu,···, x s;) E R~: there is a portfolio Z,E RK such that p,(x" - rob) S L Phr.. z" for every 50 and q-z, S O,v(a. q)},
,
(19.G.1)
Figure 19.G.l
An example of possible production choices of the firO).
vectors (a" ... , as) E A of the firm. See Figure 19.0.1 for the case where S = 2. We assume that the return vector a E A is chosen before the financial markets of period I = 0 open. Thus, the decision is made by the initial shareholders (since shares may be sold in period I = 0, the shareholders at the end of period I = 0 may be a different set). Which production plan should these initial owners choose? It turns out that the answer is very simple if A can be spanned by the existing assets and is very dillicult if it cannot. Definition 19,G.1: A set A c R S of random variables is spanned by a given asset structure if every a E A is in the range of the return matrix R of the asset structure, that is, if every a E A can be expressed as a linear combination of the available asset returns.
If we assume, first, that A is spanned by R and, second, that we are dealing with a small project (i.e., all the possible productions a E A are small relative to the size of t he economy; e.g., a,flIL; ro,;!1 is small for all s), then we are (almost) justified in taking the equilibrium spot prices P = (PI'" . ,Ps) E RLS and asset prices q = (ql' ... ' qd E R" as constants independent of the particular production plan chosen by the firm." For the asset priccs q E R" the markel value v(a, q) of any production plan a E A can be computed by arbitrage: if a = L, ex,r, then v(a, q) = L, 'J.,q,. In Exercise 19.0.1 you are asked to show that if the firm is added as a new asset to the given list of assets, and each production plan a E A is priced at its arbitrage value v(a), then any budget-feasible consumption plan of any consumer can actually be reached without purchasing any shares of the firm (the fact can be deduced from Proposition 19.E.3). Thus, for fixed asset prices q E R" and spot prices p = (p I' .•. , Ps) E R LS, the budget constraint of consumer i is"
It follows from the form of this budget constraint that at constant prices every consumer-owner (i.e., any i with 0, > 0) faced with the choice between two production plans a, a' E A, will prefer the one with higher market value, Indeed, if v(a, q) ~ v(a', q) then B•. ; c Bo" Thus, the objective of market value maximization will be the ullanimous desire of the firm's initial owners.'" If A is not spannable by the given asset structure we run into at least two serious difficulties. The first difficulty has to do with price quoting and is common to any commodity innovation problem. Without spanning, the value of a production plan a E A cannot be computed from current asset prices simply by arbitrage, The value IS not, so to speak, implicitly quoted in the economy. Therefore, it would need to be anticipated by the agents of the economy from their understanding of the workings of the overall economy-no mean task. The second difficulty, more specific to the financial context, has to do with price laklllY. Due to the possibility of unlimited short sales there is a discontinuity in the
plausibility of the price-taking assumption. With spanning we can argue, as we did that if the project is small then the effect of production decisions on asset prices, on spot prices at I = I, is also small. But if a new asset a E A, no matter how small, IS not generated by the current asset structure, then its availability increases the span of avatlable wealth transfers by one whole extra dimension. The impact is therefore substantial, and may well have a dramatic effect on prices. 31 There is then no reason for owne~s' preferences over different production plans to be dictated merely by the tocrease to wealth at the prices prior to the introduction of the firm (see Exercise 19.0.2). These two difficulties, to repeat, are serious. There is no easy way out.
0;
A variation of the above model entirely eliminates the asset role of the firm at t = O. Let us assume that the firm's shares cannot be traded at I = 0.>8 If owners at t = 0 choose " E A, this simply means that their endowments at I = I are modified by the random variable Ilia that, recall, pays in good I (i.e., the new endowment of consumer i becomes (w" + (O,a.. 0, ... , 0» E RL for every state sl. If a E A can be spanned, then we are as in the previous model. It does not matter whether shares of the firm can be sold or not at r = O. In either case consumers can take positions in the asset markets that will guarantee that the resulting final consumptions at t = I are the same (Exercise 19.G.3). If a E A cannot be spanned, matters are different. The good news is that, because no new tradeable asset is created at r = 0, the price-taking discontinuity problem disappears. The bad news IS that there IS now another difficulty: Because there is no market for the shares at I = 0, the value of the asset cannot be computed as a deterministic amount at I = O. It is rather a
34. Both assumptions are important for this conclusion. Suppose for a moment that there are
zero 10lal endowments of the asset. Then, since the asset is redundant, Proposition 19.E.3 (see also Exercise 19.E.4) implies that at the Radner equilibrium the new asset is absorbed without any change in prices. What we are now assuming is that this remains approximately true if the total endowment
of the asset is small (i.e., if the project is small). 35. Note that the value of the initial endowments at t = 0 is the value or the shares of the firm 0,1"(<1_ 'I).
36. As long as the project is small and, consequently, the prices are almost constant.
37. Recall that short sales are possible. One way out of the dilemma is to put a bound on short sales of roughly the size of the possible production vectors. The cost of this assumption is that we have then to give up (he theory or arbitrage pricing.
38. Or perhaps we are now at the end of period 1=0 and the financial markets have already closed.
UNCERTAINTY
715
716
CHAPTER
It:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
rallClom variable at I = 1 and therefore the risk attitudes of the consumer-owners will be essential to the determination of the preferred a e A. In particular, unanimity of the consumer-owners should not be expected (see Exercise 19.G.4).
19.H Imperfect Information Up to now we have concentrated the analysis on a model where spot trading for goods occurs under conditions of perfect information about the state of the world. I n this section, we relax this feature by considering the possibility that this information is not perfect. In doing so we shall see that there is a key difference bet ween the case of symmetric information (all .traders have the same information), which is largely reducible to the previous theory, and the case of asymmetric information, where a host of new and difficult conceptual problems arise. To focus on essentials, we deal with trade in a single period. You can think of it as the period I = I of the previous treatments. In this period, one of several states s = I, ... , S can arise. Once a state occurs, we consider the simplest case in which there is a single spot market. In this market a first commodity (good, service, ... ) is traded against a second good, to be thought of as money (thus, L = 2). The price of the second good is normalized to I. We reserve the symbol p E IR for the price of the nonmonetary commodity. There are I consumers. Given probabilities !t = (!til' ... , !ts,) over the states, a random consumption vector XI = (XII"'" xs l ) E R 2S is evaluated by consumer i according to an extended von Neumann-Morgenstern utility function:
where 11,1(' ) is consumer i's Bernoulli utility function in state s. Consumer i also has an initial, state-dependent, endowment vector w, = (w lI, ... , WSi) E R2S, and a signal fllllction 0",(') assigning a real number O"I(S) E R to every state S E S. The state s occurs at the beginning of the period. We assume that, once this has happened, consumer i receives the initial endowment w" and the signal O";{s) E IR. The interpretation is that consumer i can distinguish two states s, s' E S if and only if O",(s) '" 0";{S').39 Consistent with this interpretation, we require that the endowments be measurable with respect to the signal function, that is, W,i = W,'I whenever O",(s) = O",(s') (thus, we can write w,' as w.",,,). In this manner the endowments of goods of consumer i do not reveal to him information about the state of the world that is not already revealed by the signal. After every consumer gets his signal, the spot market operates. Finally, at the end of the period, the state is revealed and consumption takes place. 40
39. Equivalently, as we did (in small type) in Section 19.B, we could use informalion par/iliolls instead of explicit signal functions. The information partition .~ associated with the signal 11k) is composed of the events (., E s: u~.,) = c} oblained by lelling C E R run over all possible values. 40. In a more general. multistage. situation, some information is revealed at the end of the period and the economy moves on to the next period.
SEcrlON
It.H:
IMPERFECT
INFORMATION
717
------------------------------------------------Symmetric Information
We say that information is symmetric if any two states, s, s' E S, are distinguishable by one consumer i if and only if they are distinguishable by every other consumer k; that is, O",(s) '" O",(s') if and only if O",(s) '" O",(s'). Thus, with symmetric information we can as well assume that all consumers share the same signal function. We therefore write 0",(') = 0"(') for all i. We can think of 0"(') as a public signal. With symmetric information the determination of the spot prices proceeds in a manner entirely parallel to what we have seen so far. Suppose that state s occurs. Then every consumer i receives the signal O"(s) and initial endowments w.w." From the signal and the prior probabilities !t, = (!t", ... , !ts;), which we take to be strictly positive, he computes his posterior probabilities over the different states s' as 11",1 rr(s) =
!t," 1:",,:,,(.1'"""I1(s)l 'Tt s 'O(
for any s' with O"(s') = O"(s), and 11,·,10"(5) = 0 otherwise. The utility of a consumption bundle x, E 1R2 conditional on the signal O"(s) is then ",(X, I 0"(5)) =
L (!t,,;! O"(s»u",(x;). "
Therefore we have, conditional on s, a perfectly well-specified spot economy. Under the usual assumption of price-taking behavior, an equilibrium price will be generated. We write this price as p(0"(5» E R,42 The concept of an information signal function lends itself to interesting compara· tive statics exercises. DefinItion 19.H.1: The signallunction 0"': S ~ R is at least as informative as 0": S ~ R il O"(S) '" O"(s') implies a'(s) '" a'(s') lor any pair s, s'. It is more informative iI, in addition, O"'(s) '" O"'(s') lor some pair s, s' with O"(s) = 0"(S').43 Two arbitrary signal functions 0"('),0"'(') may well not be comparable by the "at least as informative·' concept. If they are, it is natural to ask if the more informative signal leads to a welfare improvement. We pose the "improvement" question in an ex ante sense (see Exercise 19.H.1 for an interim and an ex post sense); that is, we want to compare the expected utility of the different consumers under 0"(') and under 0"'(') when the expectation is evaluated before s occurs. 41. The endowmenls could, of course, be the result of the execution of forward trades entered upon in the past. The measurability requirement, namely. the restriction that endowments depend on the state only through the signal, then captures the restriction that a forward contract can only be made contingent on information available at the time of the execution of the contract (strictly speaking, it should be conditional on information available at that time to the contract enforcing authority). 42. Because updated probabilities and ulility functions depend only on the values of the signal, we have imposed the natural requirement that the clearing price depends also only on the signal; that is. we write p(u(s)) rather than p($). Indeed, how could the (unmodeled) markel mechanism manage to distinguish states that no consumer can distinguish? 43. In terms of the corresponding information partitions. (he signal q'(') is more informative than a(') if the information partition a'(') refines the information partition of a(·).
or
718
CHAPTER
11:
GENERAL
EQUILIBRIUM
UNDER
SECTION
UNCERTAINTY
Consider first the decision problem of a single consumer i. Suppose. for simplicity. that the spot price pER and the consumer wealth w, E R are given and are independent of s (a more general case is presented in Exercise 19.H.2). For any signal 2S E R as follows: Subject function 17(') the consumer forms a consumption plan to the restriction that x:r' = x:.'i' whenever a(s') = a(s). the consumer chooses, for every possible state s, a consumption X:,I" in his budget set that maximizes expected utility conditional on the signal a(s). The ex ante utility of the information signal function 17(') is therefore L, x"u,,(x:r')·
Ls TCSIU$i(X~}·I). Proof: The first observation is that for any 17('), xjl" solves the problem
Max
L
1[si U.si(X s i)
S.t. Xi E
Bf(·j =
{Xi E
R 2S
: pX lsi
x,,
+ XlIi :s;;
= X,"
Wj
for every s, and
whenever a(s)
= a(s')}.
You are asked to verify this formally in Exercise 19.H.3. The second observation is that if 17'(') is at least as informative as 17('), then B711 c B(I·'. Again you should verify this in Exercise 19.H.3. It follows that moving from 17(') to 17'(') we only expand the constraint set of the consumer's problem. In doing so, the maximum value cannot decrease. The claim of Proposition 19,H.1 is intuitive: in an isolated single-person decision problem. a more informative signal does not upset the feasibility of any decision plan (hence. it brings about an expansion of the feasible set) because the decision maker always has the option not to act on the extra information provided by 17'(') over 17('). Unfortunately. this line of argument does not apply to a system with interacting decision makers. At the equilibrium the budget set of a single consumer may well be affected by the signal. even if the consumer does not use it. It suffices that the other consumers use it and that. as a result. the new information finds its way into the spot prices. Example 19.H.I shows that, because of this. it is even possible for an increase of information to make everybody (ex ante) worse off. Example \9.H.1: Suppose there are two consumers. two commodities. and two equally probable states s = I. 2. In both states the two consumers' endowments of the two physical goods are w, = (1.0) for consumer I and W2 = (0. I) for consumer 2. I n total. therefore. there is one unit of every commodity in every state. The two consumers have the same von Neumann-Morgenstern expected utility function. Their state dependent Bernoulli utility function is
u,,(x,,/. X2,/)
=
p,';;:::' + (\ - P,)jX;;.
where fll = I and fl2 = O. Thus. in state I the second good is worthless. while in state 2 the first good is worthless. Suppose first that there is no information (i.e., there is no signal function distinguishing the two states). Then there is a single spot market where every
IMPERFECT
consumer chooses amounts (XII' Xl') of the two commodities so as to maximize the expected utility function
tjx,; + !Jx,:.
xr'"
PropOSition 19.H.1: In the single-consumer problem. if the signal function 17'(') is at least as informative as the signal function 17('), then the ex ante utility derived from 17'('), :L xs;us;(x:t'), is at least as large as the ex ante utility derived from 17(')'
11,H:
By symmetry (but nonetheless compute the first order conditions) we see that at equilibrium every consumer will get half of each commodity and have an expected utility of 1/.j2. Hence. in this no-information equilibrium every consumer has managed to insure against the possibility that the good he owns turns out to be worthless. Suppose that instead we have a perfectly informative signal function revealing the state prior to the opening oC the spot market. Then the spot market equilibrium will be different under the two states, What happens now is that in each state one good is known to be worthless and. therefore. there is no possibility of trade in the spot market: Each consumer consumes his endowment, receiving a utility of 1 in one state and oC 0 in the other. Ex ante this means that under perfect information every consumer has an expected utility of t < 1/.j2. Thus. we see that the availability of a more informative signal function makes everybody worse off, The reason is that the availability oC information destroys insurance opportunities [a possibility first pointed out by HirshleiCer (1973»).44 _ Asymmetric lriformation Suppose now that inCormation is not symmetric; that is, the signal functions ai(') are private and not necessarily the same across consumers. How to proceed then? A first thought is to proceed exactly as beCore. When s occurs every consumer observes ab) and uses his signal function ai(') to update probabilities and utility functions. This defines a spot economy to which we can associate in the usual way a spot market clearing price written as p(a,(s)•... , a,(s», Note that the price p(a" ... , as) depends on all the individual signals: One says that the price aggregates the information of the market participants. In particular, the price/unction p(s) = p(a,(s), ... , a,(s» need not be measurable with respect to the individual signal functions a!-'); that is, it may be that two states s. s' E S are not distinguishable by consumer i (i.e., a/(s) = a,(s'», but are distinguished by the market [i.e .• p(a,(s)•...• a,(s» oF p(a,(s'), ...• a,(s'))]. This raises an important difficulty that we discuss by means of Example 19.H.2. 44. Suppose ,ha' our period is period 1 and that previous to it there is a period 0 in which forward trade could conceivably take place. Under the no-information scenario there can be no contingent trade at t = 0 since the two states cannot be distinguished at t :a::: l. The model considered, therefore, is as complete as it can be (hence, th. equilibrium is Par.to optimal relative to th. no-information structure). This is not so for the model with perfect information. There is then no
informational impedim.nl to th. creation of a complet. s.t of conting.nt mark.ts at I = O. With them the possibility of insurance would be r.stored. In g.n.ral, if markets arc compl.te [relative '0 the information signal function a(')] th.n th. equilibrium is a Pareto optimum [relative to IT(.») and, therefor., if information improves (and the corresponding additional markets are created) some traders may gain and some may lose (i .•.• th.r. may be distribution .ffects) but overall the new vector of ex ante expected utilities is at the frontier of an .xpanded utility possibility set. We can conclude. therefore, that if markets arc always complete relative to the information signal (i.e., a forward market contingent on every signal takes place at I = 0), then not everyone can end up worse off if the information signal function improves.
INFORMATION
719
720
CHAPTER
11:
GENERAL
EGUILIBRIUM
UNDER
UNCERTAINTY
SEC T' 0 N
Example 19.".2: Tbe economy bas two goods and two consumers. The two consumers have identical utility functions u,{x lI , X2') = PIn XII + Xl!' The parameter Pis the same for tbe two consumers and it is uncertain. It can take the values P= I and P= 2 with equal probability. (Hence, we can think that tbere are two equally probable states: One yields P= I and the other P= 2.) The two consumers bave deterministic endowments of one unit of the first good (because of quasilinearity with respect to the second good we do not need to specify endowments of the latter). The first consumer has an informative signal CT I = P that allows him to distinguish the two possible values of p. The second consumer is not informed: his signal function has CT2(P) = k for some constant k. After nature bas determined the value of P and the information CT,(f1), CT2(fJ) has been transmitted to tbe two consumers, a spot market takes place (as usual, the price of the numeraire commodity is fixed to be I). Since the first consumer knows f1, his demand, given tbe price pER of the first good, is x l1 (p; P) = PIp. The second, uninformed, consumer will equalize his expected marginal utility to the price. Hence, his demand function, wbich does not depend on p, is
«I)
x, 2( p) = I [I I P 2
+~2 2J = 2p ~.
Solving the market equilibrium equation x l1 (p; Ii) brium price function p(fJ) =
+ X'2(P)
(19.H.I) = 2 we get the equili-
;\(3 + 2fJ).
Note now tbat p( I) >F P(2). Tbis means that the price reveals the informed consumer's information. If so, then it is logical to suppose that the uninformed consumer will try to use the observed market price to infer the unobserved value of fl. There is really no good reason to prevent him from doing so. But, once he does, his demand will no longer be given by (19.H.I), and the price function p(P) specified above will no longer clear markets for every possible value of p. Tbis is the difficulty we wanted to illustrate. It suggests that what is needed is an equilibrium notion embodying a consistency requirement between the information revealed by prices and the information used by consumers. _ We have argued in Exainple 19.H.2 that it makes sense to require that the information revealed by prices be taken into account by the consumers in making their consumption plans in tbe different spot markets. Suppose, therefore, that PIs) = p(CT,(S), .•• , CT,(S» is an arbitrary price function. Interpret it as a specification of the prices expected to bold by the consumers at the different states. We could now view this price function as a public signal function and let any consumer use it in combination witb his private signal. Tbat is, wben state 5 occurs, consumer i knows that the event Ep(.,.•,(., = {s'; p(s') = p(s) and CT,(S') = CTi(S)} has occurred and updates his probability estimates of any state 5' E £p(",,,(,, to
1 •• H:
IMP E R F E C TIN FOR MAT ION
If, for the updated utility functions, the price p(s) clears the spot market for every s, then we say that the price function p(.) is a rational expectations equilibrium price s function.· ", This is expressed formally in Definition 19.H.2.
Definition 19.H.2: The price function p(.) is a rational expectations equilibrium price function if, for every s, pIs) clears the spot market when every consumer i knows that s E Ep (.) .•,(.) and, therefore, evaluates commodity bundles x, E R2 according to the updated utility function
Ls· (n •., I p(s), CT;(S»U•.;(x;). We saw in Example 19.H.2 a situation in which all privately observed information is revealed by the spot market price. This suggests the following approach to the determination of a rational expectations equilibrium price function. Imagine (this is merely a hypothetical experiment) that all the individual signal functions are known to all consumers and that for every state the vector of signal values (CT,(S), ... , CT,(S» is made public and is, therefore, usable by all consumers to update probabilities and utilities. The market-clearing price function fils) = (i(CT,(S), .•. , CT,(S» thus generated is called the pooled information equilibrium price function. If the values of (i(') distinguish all possible values of (CT" ... , CT,), that is, if (i(s) >F (i(s') whenever CTi(S) # CTi(S') for some s, 5', and i, then we say the price function (i(') is fully revealing. In other words, the price function is fully revealing if it distinguishes the occurrence of any two states that can be distinguished by some consumer. We argue now that if the pooled information equilibrium price function (i(') is fully revealing. then it must be a rational expectations equilibrium price function. For any s, (i(s) is determined under the assumption that every i knows that s E {s': CT.(S') = CT.(S) for all k}. Because the pooled information equilibrium price function (i(') is fully revealing, it follows that Is': CT.(S') = CT.(S) for all k} = Is': (i(s') = (i(s)}. Hence, for any s, fils) is a market-clearing price when every i knows that s E E..... ,.• ,(.,. We conclude that (i(') is a rational expectations equilibrium price function. In other words: If a pooled information equilibrium price function is fully revealing, then the pooled information used by consumers need not be obtained by violating any privacy constraint but can simply be derived from the public price signals. Example 19.".2 continued: If both consumers are fully informed, we have the demand functions x l1 (p) = xdp) = PIp. Thus, in this case, the pooled information equilibrium price function is (i(fJ) = p. This pooled information equilibrium price function is fully revealing and therefore a rational expectations equilibrium. _ 45. For the concept or rational expeclations equilibrium (including additional references). see Green (1973), Grossman (1977) and (1981), Lucas (1972), and Allen (1986). 46. In this section we concentrate on issues relating to inrormation transmission more than on maHers concerning spanning or completeness. But note that, as we pointed out in the case of symmetric inrormation (in rootnote 44), Ihere is no conceptual difficulty in imagining that previous to the period r = I under consideration there has been contingent trade for the delivery at r = I or amounts or physical good conditional on the values or the public signals at I = I (we call the overall situation complete if such conlingent markets exist for every possible value or the public signals). Obser~e Ihat because the SpOI price constitutes a public signal, a possible instrument of contingent trade IS an asset with returns conditional on the realized value of the spot market price at I = I; options are instances of such assets.
721
722
c HAP T E R
1 I:
G
ENE R ALE
Q
U, LI
8
R I U M U N D E RUN
CE
RTA I NT Y
[n Example 19.H.3 the pooled information equilibrium price function is not fully revealing and fails to be a rational expectations equilibrium. [n fact, in the example no rational expectations equilibrium price function exists.
Example 19.H.3: [Kreps (1977)] There are two commodities and two consumers with utility functions U'(.
As we have seen, the concept of a fully revealing equilibrium provides a useful tool for the Sludy of markets with asymmetric information. In applications it is more common tha, we encounler a slightly weaker and more natural version of the full revelation idea. In effect, in order for the pooled information equilibrium price function M..) to be a rational expectations equilibrium price function, we do not need that, for every s, p(.) reveals precisely the vector of signals (<1, (s), ... , <1,(s»; it is enough that it reveals a sufficient statistic for this vector [or a stalistic that is sufficient for every consumer i in conjunction with the private signal <1,(.»). More generally, all we need is that for every possible state s the expressed demand of every consumer i at the price p(s) is the same whether the consumer knows the pooled information signed functions (<1 I ( . ), ... , <1 /( .», and receives the signal vector (<1 I (s), ... , <1,(s», or instead knows only the signal function p(.) [or only p(.) and <1,{.»).
SECTtON
II.H:
IMPERFECT
Suppose that information is pooled. Then: (i) If <1, = 4, 5, or 6 for either i = I or i = 2, we know that (J = 2 with probability I and therefore (i(<1 <1,) = fJ/2 = I. (ii) If <1, = -1,0," or I for either i = I or i = 2, we know that (J = I with probability I and therefore P(<1" <1,) = (1/2 = I. (iii) In the remaining cases, <1, = 2 or 3 for both i = I and i = 2, the updated probabilities on the two values of (J remain J: No useful information is transmitted. Hence the elearing price is P(<1" a,l = i (Exercise 19.H.4). The price function p(.) defined by (i) to (iii) is not fully revealing: Given the value of (i(.) we cannot deduce from it the specific values of <1, and a,." Vet, the price funC1ion (1(.) is suflieient to distinguish among cases (i) to (iii), and therefore the knowledge of the single function (i(.) can replace for every consumer the knowledge of the vector of functions (<1 1(·), <1l(·». We can say, therefore, that (i(.) is a sufficient statistic for the signals, and conclude that p(.) is a rational expectations price function. _ In Example 19.H.5 price function is not a sufficient statistic but it becomes so when combined with the private signal of any consumer. Example 19.H.5: The basic economy and the signals are as in Example 19.H.4, but with three diITerences. First, there are I consumers. Second, the noise terms " arc now payoff relevant: in particular,u,(x,) = ({I + f.,) In Xli + Xli. Third, half the consumers have their noise uniformly distributed ill the interval [ -i,n whereas the other half is perfectly informed about (J, that is, 1:1 = O. The pooled information equilibrium price function is P((J,." .. . ,") = (J + (l/IXL,',). NOle that this price function reveals (J [if (J = I then (i(.) < 1.5 with probability I; if fJ = 2 then p(.) > 1.5 with probability I] but not the individual values of ',. However, a consumer· i that knows (J and <1, = (J' + " also knows t" and therefore, at any given price, expresses a demand that coincides with the pooled information demand. We conclude that the pooled information equilibrium is a rational expectation equilibrium. It is important to observe that, in contrast with Example 19.H.4, the equilibrium price function alone does not now provide a sufficient statistic. At the rational expectations equilibrium the individual utility maximization problems of half the consumers make essential use of the private signal functions. _
Example t9.H.4: The basic economy is as in Example 19.H.2 Now, however, the signal of each consumer i = 1,2 is <1, = (J' + ". The '1' for i = I and 2, are noise variables independently distributed and taking the values " = -2, -1,0, t, 2 with equal probability."
Example 19.H.S allows us to address another issue. Suppose that to our general model we add the fealure that acquiring the information signal function <1,(.) costs some small amount of money fl > O. Suppose also that I is large, so that, plausibly, the pooled information price function p(.) is not very sensitive to any single consumer i failing to acquire his signal function <1,C). Then we have the following paradox [see Grossman and Stiglitz (1976)]: If the price function (i(.) is fully revealing (or is a sufficient statistic by itself), why will any consumer i pay fl for the signal function 11,<·)7 Anyone consumer would rather not do so and attempt to free ride on the information transmitted by the price system. But if everybody proceeds in this manner, then the price function cannot be fully revealing (there is nothing to reveal)! Example 19.H.5 suggests one way out of this paradox: it can be verified that in the example there is a sufficiently small fl > 0, not dependent on the number I, such that at any fixed price p, a COnsumer i with nontrivial." even if he already knows (J, has an incentive to pay fl for the improvement of information provided by his private signal (Exercise 19.H.5). _
47. All of Ihis could be expressed in terms of underlying states s = (P, ," .,l. We would need 2 x 5 x 5 = 50 of Ihem.
we gel 10 know the value or the signals.
48. Recall that "fully revealing'· does nOI mean that we gel to know the value of (I, only that
INFORMATION
723
724
C HAP T E R
1 I:
G ENE R ALE QUI L I 8 R I U M
UNO E RUN C E R T A I N T Y
Up to now, prices may have conveyed information about an exogenously ocurring state. But in a world of asymmetric information, prices could also convey information on the consumers' endogenously chosen actions, and those actions could matter for individual utilities. For example, the final utility of a consumer may depend not only of the number of units consumed and on exogenous states but also on some other statistic depending on other consumers' actions. If we regard this statistic as a "state" then it is as if states were determined endogenously. To illustrate this point we consider another example-the market for used cars (also referred to as the "lemons market"). With this, we connect with the analysis of adverse selection in Section 13.B and bring this section to a close. Example 19.H.6: The Lemons Market. Suppose that consumers fall into two types: potential buyers and potential sellers (of, say, used cars). There are many consumers and twice as many pOlenliol sellers as potential buyers. Potential sellers have one unit of the good and potential buyers buy either one unit or none. The peculiarity of this market is that commodities are of two kinds: good and bad. Half the potential sellers have a good product and half have a bad one. The quality, known to the polential sellers, is unrecognizable to the buyers at the moment of trade. A good commodity is worth I to the potential seller and 4 to the buyer. A bad commodily is worth nothing to every consumer. We could call the stale of the market the fraction ~ E [0, I] of the commodities supplied that are of good quality. If the state of the market is ~, then a buyer paying p gets expeeted ulilily 42 - p. The problem is that the state of the market depends on the price (thUS, as before, the price provides information about the utility derived from consuming one unit of the good). Indeed, for any p > 0, every unit of bad commodity will be supplied to the market. But for p < I no unit of the good commodity will be supplied, whereas for p > I every unit of the good commodity will be supplied. Therefore, we must have ~ = 0 for p < I, ~ E [0, for p = I, and 2 =! for p> I. If a pair (~, p) satisfies these inequalities then we say that the pair (~, p) is admissible. Note that the inequalities can be equivalently expressed as pSI for 2=O,p= I for2e(O,!)andp~ I forlX-!. A potential buyer will deduce IX = 0 if he observes p < I, and C< = if p > I. Potential buyers mayor may not express a demand depending on this inference. It is natural to say that if at the admissible pair (<<, p) the total demand is not larger than the total supply then we are at a rational expectations equilibrium. In fact, in our case any admissible pair (<<, p) turns out to be a rational expectations eqUilibrium. Note, however, that for some (C<, p) the supply is larger than the demand (e.g. at IX t, p 3 no demand is expressed)." _
n
t
=
=
EXERCISES
725
REFERENCES Allen, B. (1986). General equilibrium with rational expectations. Chap. I in Contribur/OftJ to Mathematical EconomiCS, ediled by W. Hildenbrand, and A. Mas.cole11. Amsterdam. North-Holland. Arrow, K. (1953). Le role des valcurs boursieres pour la repartition la meilleurc des risques. Ecoflomttrie,
Paris: Centre National de la Rechercbe Scientifique. [Transilled
or
IS:
Arrow, K. (1964). The role of
securities in the optimal allocation risk·bearing. R~;tw of Economic Studies 31: 91-96.] Casso D.. and K. Shell. (1983). Do sunspols malter? Journal of Polilical Economy 91: 193-227. Dcbreu. G. (1959). Th.ory of Valu •. New York: Wiley. Diamond. P. (1967). The role of a stock market in a general equilibrium model with technological uncertainly. American E('onomic Review 57: 759-76. Dunie. D. (1992). D,'namic' Assrl Pricing Theory. Princeton, N.J.: Princeton University Press. Geanakoplos, 1., and H. Polcmarchakis. (1986). Exiscence, regularity and constrained 5uboptimality of competitive allocations when the asset market is incomplete. In Essays in Honor of K. Arrow,
vol. III, ediled by W. Heller, and D. Slarrell. Cambridge, U.K.: Cambridge University Press. Green, 1. (1973). Information, efficiency and equilibrium. Harvard Discussion Paper 284. Grossman. S. (1977). The existence or future mark.els, noisy rational expectalions and informational externalities. R('t,j('w of Economic Studies 44: 431-49. Grossman. S. (1981). An introduction to the theory or rational exp«tations under asymmetric information. Rt'l'it'M' I~r EnJnomic Sludi('s 48: 541-S9. Grossman. S .• and J. E. Stiglitz. (1976). Information and competitive price systems. American Ecunomic Rt'l"it'W 66: 246-53. Hart. O. (1975). On the optimality of equilibrium when the market structure is incomplete. Journal of l::nltlomk Theor), J I: 418-41 Hirshleifcr. 1. (1973). Where are we in the theory of information? Amt*rkun Economic Rn1iew, Papers and Pron!eJings 63: 31-40. Huang. C. F .• and R. Litzenberger. (1988). Foundations of Financial Economics. Amsterdam: North~
Holland. Kreps, D. (1977). A note on 'fulfilled expeelalion,' equilibria. Journal of Econmnic Th"",y 14: 32-43. Krep~ D. (1979). Three essays on capital markelS. Institute for Mathematical Studies in The Social Sciences, Technical Report. v. 298, Stanford Universily. Reprinted as Kreps, D (1987): Three essays on capilal markets. Re(l;sta Espanola d~ Economia• .-: 111-146. Lucas. R. (1972). Expectations and the neutrality of money. Journal of Economic Theory'-: 103-24. Magill. M.. and W. Shafer. (1991). Incomplete markels. Chap. 30 in Handbook of Malhernalical fconomiL's, vol. IV. ediled by W. Hildenbrand, and H. Sonnenschein. Amsterdam: North-Holland. Marimon. R. (1987). Kreps' "Three essays on capital markets" almost len years later. Revilla Espanola d. Economia 4(1): 147-71.
Merton. R. (1982). On the microeconomic theory of in"cstment under uncertainty. Chap. 13 in Handbook of Mathematical Economics, "01. II, edited by K. Arrow, and M. D. Intriligator. Amsterdam:
49. For simplicily we have chosen an example where it is always the case that the forthcoming demand is not larger than the forthcoming supply. It is because of this thai all admissible pairs (a, p) may appear as equilibria. More generally, some of Ihese pairs may be eliminated because forlhcoming demand is larger than supply. It is also worth observing Ihal it would not be legitimale to impose as an equilibrium condition Ihat demand be larger Ihan or equal to supply. Suppose, for example,lhal a = land p = 1.5. Then lotal supply is 2 while 10lal demand is I. The usual argumenl (underlying Ihe latonnement dynamics) for downward pressure on demand is Ihat some frustraled seller would altempt to sell to some buyer al a price 1.5 - t. Bul, in Ihe current conlext, how is the buyer 10 know that the commodity being offered is not of bad qualily? NOle Ihat Ihings would look differenl if it were the buyer wbo approached a random seller (Ihis is Ihe reason wby Ibe requirement that tbe demand be no larger tban Ihe supply is a natural eqUilibrium condition.) In contrast, with symmetric information it is of no consequence who approaches whom when a buyer and a seller meet. The lesson to be learned from this discussion is that with asymmetric information Ihe parlicular disequilibrium Slory mailers a lot. To push Ihe analysis forward it is Iherefore appropriale 10 refer back 10 Cbapler 13 where, in a more limiled, parlial equilibrium selling, we have already studied asymmetric information problems witb Ihe help of a melhodology well suited to the consideration of this type of microstructure.
North·Holiand. Radncr. R. (1982). Equilibrium under uncertainty. Chap. 20 in Handbook of Malhematical Economics,
vol. II. edited by K. Arrow, and M. D. Intriligator. Amsterdam: North-Holland.
EXERCISES A
19.C.I There are S slates. A consumer has, in every slatc 5, a Bernoulli utility funclion /I,(x,), where X,E R~. Suppose that, for every s, u,(·) is concave. Show that the cxpeeled ulilily funclion U(x" ... , xs) = L, n,u.(x,) defined on R'+' is concave. A
19.C.2 For Ihe model described in Example 19.C.2, show Ihat the marginal rales of subslilulion along the Parelo set are as drawn in Figure t9.C.2; that is, at any poinl of the Parelo sel Ihe marginal rale of substitution is smaller Ihan Ibe ratio of probabilities.
-
726
CHAPTER
11:
GENERAL
EQUILIBRIUM
UNDER
UNCERTAINTY
19.C.3 A Consider an Arrow-Debreu equilibrium of the economy described in Section 19.C. Suppose that L = 1 and that preferences of every consumer i admit an expected utility representation with continuous, strictly concave, and strictly increasing Bernoulli utility functions (identical across states). For every state. denote by p.. ",., and x" the equilibrium price for the s-contingent commodity, the subjective probability of consumer i for state " and the consumption of consumer i in state " respectively. Denote by p = L. p, the price of uncontingent delivery of one unit of consumption. Show that L. (n.,p - p,)x,. ~ 0 for every i. [Hinl: Usc a revealed preference argument.) Interpret.
t 9.C.4" There are a single consumption g~od, two states, and two consumers. Note that this allows the use of Edgeworth boxes. Utility functions are of the expected utility form. Bernoulli utility functions are identical across states. That is, U,(X", Xli) = "oo",(x oo ) + ",,",(XlI)
and U'('
+ ",,",(x,,),
where X.i is the amount of s-contingent good consumed by consumer i and 1[si is the subjective probability of consumer i for state •. We assume that every ",(.) is strictly concave and differentiable. The total initial endowments of the two contingent commodities are w = (w" w,) »0. We assume that every consumer gets half of the random variable W, that is, (woo' WlI) = 1w and (wl2' w n )
= }w.
(a) Suppose that consumer I is risk neutral, consumer 2 is not, and both consumers have the same subjective probabilities. Show that at an interior Arrow-Debreu equilibrium
consumer 2 insures completely. (b) Suppose now that consumer I is risk neutral, consumer 2 is not, and the subjective probabilities of the two consumers are not the same. Show then that at an interior (Arrow-Debreu) equilibrium consumer 2 will not insure completely. Which is the direction of the bias in terms of the differences in subjective probabilities? Argue also that consumer I (the risk-neutral agent) will not gain from trade.
19.C.5 A Consider an economy such as the one introduced in Section 19.C but with only one commodity in each state. There is a number I of risk-averse consumers. Preferences admit an expected utility representation. Suppose that the Bernoulli utility functions of a consumer for the good are identical across states and that subjective probabilities are the same across individuals. Individual endowments vary from state to state. However, we assume that the total endowment is nonstochastic, that is, uniform across states (if, say, I is large and the realizations of individual endowments are identically and independently distributed, then the total endowment per capita is almost nonstochastic). Set up the Arrow-Debreu trading problem. Show that the allocation in which every individual's consumption in every state is the average across states of his endowments is an equilibrium allocation.
•
EXERCISES
19.0.3 A Formulate a model similar to the two-period model of Section 19.0 with the difference that consumption also takes place at period I = O. Show that the result of Proposition 19.0.1 remains valid. 19.0.4" Consider a three-period economy, 1=0, 1,2, in which at 1=0 the economy splits into two branches and at I = I every branch splits again into two. There are H physical commodities and consumption can take place at the three dates. (a) Describe the Arrow-Debreu equilibrium problem for this economy. (b) Describe the Radner equilibrium problem. Suppose that at I = 0 and I = I there are contingent markets for the delivery of one unit of the first physical good at the following date. (c) Argue that the conclusion of Proposition 19.0.1 remains valid.
19.E.I" Consider an asset trading model such as that considered in Section 19.E. The only difference is that consumption is also possible at date I = O. Assume for simplicity that the Bernoulli utility functions on consumption are state independent and additively separable across time; that is, u,(X Oit XII) = UOj(XOi) + ulA.XIE)' where XOI. XII e RL. (a) Argue along the lines of the second proof of Proposition 19.E.1 that the conclusion of Proposilion 19.E.1 remains valid.
(b) Suppose now that there is a single physical good in every period. Express the multipliers JI, in terms of marginal utilities of consumption.
19.E.2 A Show that if a primary asset with return vector r E It' separales states, that is, if r, ¢ '" whenever s ¢ s', then it is possible to create a complete asset structure by using only options on this primary asset. You can assume that " > 0 for every s. 19.E.3 A Complete part (i) of the proof of Proposition 19.E.2 in the manner requested in the text. 19.~:.4"
In text.
19.E,5" Complete the verification that at the prices specified in Example 19.E.7 the set of consumptions achievable through sequential trade is the same as the set of consumptions achievable through ex ante trade of the four Arrow-Debreu commodities. 19.E.6 A There are two dates: At date I there are three states; at date 0 there is trade in assets. There are two basic assets whose return vectors in current dollars are " = (64. 16,4)
and
" = (0,0, I).
The markel prices of these assets are q, = 32 and q, = I, respectively. In the following you arc asked to price by arbitrage a variety of derived assets. (a) Suppose that one unit of a derived asset is described as "One unit of this asset confers the right to buy one unit of asset I at 75% of its spot value in period I (after the state of the world occurs)." Write the return vector of this asset and price it. (b) The siluation is the same as in (a) except that the asset is modified to read "One ullit of this asset confers the right to buy one unit of asset I at 75% of its spot value in period I (afler the state of the world occurs) provided the spot value is at least 10." (0) Suppose that the asset is as in (b) except that "at least 10" is replaced by "at least 19." Write down Ihe return vector and argue that this asset cannot be priced by arbitrage with the a vailable primary assets.
19.0.IA Consider the model of sequential trade of Section 19.0. The only difference is that we now assume that, for every., the .-contingent commodity pays I dollar (rather than one unit of physical good I) if state. occurs (and nothing otherwise). Write down the budget constraints corresponding to this model and discuss which price normalizations are possible.
(d) How would the analysis in (0) differ if we had in addition a riskless asset with a price equal to I? (You do not need to compute the price explicitly.)
19.0.2A Show that in Example 19.0.1 the contingent trades of the two consumers are as claimed in the discussion of the example.
(0) Suppose that now the asset is further complicated to read "One unit of this asset confers, at the choosing of the holder, either I dollar in period I or the right to buy one unit of asset
727
728
CHAPTER
11:
GENERAL
EGUILIBRIUM
UNDER
UNCERTAINTY
1 at 75% of its spot value in period I (after the state of the world occurs)." Write the return vector of this asset and price it. (f) The situation is the same as in (e) except that the asset is modified to read ·One unit of this asset confers, at the choosing of the holder, either 1 dollar in period I or the right to buy one unit of asset I at 75% of its spot value in period I (after the state of the world occurs) provided this value is at least 10." 19.F.l" Consider the sunspot model of Example 19.F.1. Argue that, under the standard conditions on preferences, there is a sunspot-frce equilibrium, whatever the asset structure. 19.F.2A Consider (for the case L = I) the utility function U t(·) on asset portfolios defined in Section 19.F. Give sufficient conditions for Ut(·) to be continuous and concave. Show also that if returns are strictly positive then U t(·) is strictly increasing. 19.F.3c The aim of this exercise is to show that it is possible in an incomplete market situation for the number of assets to increase while at the same lime everybody becomes worse off at the new Radner equilibrium. We do this in steps. (a) Construct a two-consumer economy with Iwo equally likely states and in which the distributional effects of trade are so biased that Ihe sum of the utilities at the equilibrium with complete markets is smaller than the sum of the utilities at the equilibrium with incomplete markets. (b) Now construct an economy that has four equally likely states and in which in the first two states the economy is as in (a), while in the other two states it is also as in (a) excepI that the roles of the two consumers are reversed. (c) Show then that the asset structure in which there is a single asset allowing a transfer of wealth from the first two states to the other two yields an equilibrium that is better for every consumer than the equilibrium obtained if we add two new assets, one allowing a transfer of wealth between states I and 2 and the other doing the same between states 3 and 4. 19.F.4c Exhibit an example in which with unlimited short sales a Radner equilibrium may not exist. This example requires returns denominated in more than one commodity. [Hint: Recall the no-arbitrage necessary condition.]
EXERCISES
equilibrium described in the answer 10 (b) is not reachable. Continuing to assume that each worker can purchase insurance from no more than one insurance company, determine the competitive equilibrium. Is it optimal relative to the allocations the government can achieve, supposing that the government also cannot observe disability? 19.G.IA Justify expression (l9.G.I). That is, suppose that every possible production plan of the firm can be spanned and that prices (in the asset and spot markets) are given. Now inlroduc-e Ihe firm as a new asset. Show that any consumption plan of any consumer can be reached without purchasing any shares of the firm. 19.G.2A Suppose that, in a two-consumer economy with L = I, initially there is a single asset (which, therefore, goes un traded; recall that we do not allow consumption at 1=0). Now a firm is inlroduced Ihat can produce the return vector £Q E R~. The firm is owned, with equal shares, by the two consumers. Give an example in which, no malter how small < may be (and !citing Ihe vector" E R'~ remain fixed), the introduction of the firm as an assel tradeable at I = 0 has Ihe property that at the new equilibrium one consumer is significantly belter off and Ihe olher is significanlly worse off. 19.G.3' Suppose Ihat, in an economy with L = I, a firm is introduced thai can produce the single return vector
19.F,s" Consider an economy with a single period, a single consumption good, and a single input (labor). All workers are identical ex ante. Each of them has a probability! of being able to work (in which case the production is k units of output and work causes no disutility). With probability t, the worker is unable to work (is disabled). The utility of an amount c of consumption if able or disabled is Uic), Ulc), respectively. Assume that the probability of disability is independent across workers and that there arc sufficient workers that society operates on the expected value production possibility set.
(b) Show Ihat x71 " is interim optimal in the following sense: there is no allocation X; measurable with respect 10 0(') and such that for some possible signal a(s) the expected utility of x, [conditional on o(s)) is larger than that corresponding to x7 ·'.
(a) If there is a full set of Arrow-Debreu markets that are open before the disability is known, what is the equilibrium allocation of resources? (You may need to allow for the possibility of infinitely many states.)
(e) Show that if ,,'(.) is at least as informative as 0'(.) then for every s the expecled utility generaled by conditional on a'(,), cannot be inferior to the expected utility generated by X· ' " conditional on 0'(5).
(b) Assume now that it is impossible for others to observe whether an individual really is disabled or just claims to be so and stops working. Assume that insurance markets continue to exist as a competitive industry. Assume also that the condition "U:(c.) = U4(CI) implies c. > cl " is satisfied and that each individual purchases insurance only from a single company. Show that the competitive equilibrium (you should also define what this means) is the same as the one derived in (a).
(d) Show, similarly, thai if 0'(') is at leasl as informative as a(') then for every' the expected utilily generated by Xr"'I, conditional on o(s), cannot be inferior to the expected utilily generaled by x"", conditional on a(·).
(c) Continue to assume that it is impossible for others to observe whether a claimed disability is real. But assume now that U;(c.) = U;,(c d ) implies c. < Cd' Show then that the
729
'
xr"',
19.H.2' Argue that for the validity of Proposilion 19.H.I we may allow p and W; to depend on Ihe stale. Whal is required is that, as funclions, they be measurable with respect to the original a(') and that they remain unaltered as 0(') is replaced by 0"('). 19.11.3" Complete the requested steps of the proof of Proposition 19.H.1.
-
730
c HAP T E R
":
G ENE R ALE QUI LIB R I U M
UNO E RUN C E R T A I N T Y
19.H.4 A Complet. the missing st.p in Exampl. 19.H.4.
EXERCISES
19.H.5 Carry out the r.quest.d v.rification in Exampl. 19.H.5.
equilibrium prices if they clear the two spot markets when consumer I derives informalion from p!«). p!(t» in Ihe manner just described. Suppose Ihal t ~ 0 and derive a ralional expectations equilibrium pair of prices.
19.H.6C Consider the following two-consumer. two-commodity general equilibrium model. The (Bernoulli) ulility functions of the two consumers are
(d) Show that if t = 0 then there is no rational expectations equilibrium pair of prices. Compare with Example 19.H.3.
8
U,(XI1,x1d
=
XII
+ X21>
u,(x". Xll) = (x,,)'/1
+ Xll'
Consumer I's endowment of the second good is w". H. has no endowment of the first good. Consumer 2 has no endowm.nts of the second good and his endowments of the first good depend on which of three .qually likely states occurs. The respective levels in the three states are Will. W12l. Will-
(a) Delermine the Arrow-Debreu equilibrium of this economy. You can assume that the paramelers are such' that the equilibrium is interior. (b) Suppose that the only possible markets are for the noncontingent delivery of the two goods. Set up the equilibrium problem. Is the equilibrium allocation a Pareto optimum? (c) Suppose now that before any trade takes place. and before the endowments are revealed, the two consumers are told whether or not state I has occurred. Arter the revelation of this information (and before the values of the endowments are disclosed) non·contingent trade can take place. Set up the equilibrium problem as it depends on the information available. (d) The setting is as in (e); the only difference is that contingent trade (after the revelation of the information on stat. I) is now permitted. (e) Compare .x ante (i.•.• before any announcement is known) the expected utilities attained by the two consum.rs in the equilibria of (a), (b). (c). and (d); assume that all of these equilibria are interior. When can you assert that the information available in parts (c) and (d) is socially valuable? 19.H.7 A Suppose that th.r. are two equally likely stat.s. In ev.ry state there is a spot market where a consumption good (good I) is .xchanged against the numeraire, which we denote as good 2. There are two consumers. Their utiliti.s are COll.lumer J
Consumer 2
SIOle J
Slate]
21n Xli + Xli 4Inx12+X12
4lnx ll +x21 21nx12xx12
The total endowment of the first good equals 6 in the first state and 6 + < in the second state. All the endowm.nts of Ihis good are received by Ih. second consumer. Assume also Ihal the endowments of num.raire for Ih. Iwo consumers are sufficient for us to neglecl the possibility of boundary .quilibria. Th. price of the numeraire is fixed to be one in the two states. The prices of the non·numerair. good in the IWO stat.s are denoted (p" p,). (a) Suppose Ihat when uncerlainty is resolved, bolh consumers know which state of the world has occurred. Determine the spot .quilibrium prices (p,(t), p,«» in the two states (as a function of the parameter t). (b) Assum. now Ihal wh.n a stale occurs, consumer 2 knows the state but consumer I remains uninformed (i.e.• h. musl keep Ihinking of Ihe Iwo Slales as .qually likely). Delermine, under this informal ion setup (and assuming Ihat pric.s cannot be used as signals). the spot equilibrium prices (p,(t). p,«» in Ihe two slat.s. (c) The situation is as in (b), except thaI now we allow consumer I to deduce the state of the world from prices. That is, if p, ~ p, then consumer I is actually informed, but if p, = p, he is not informed. A pair of spot prices (p!«), p!«)) constitute rational expectations
«
731
CHAPTER
Equilibrium and Time
20
SECTION
In this chapter, we present the basic elements of the extension of competitive equilibrium theory to an intertemporal setting. In the presentation, we try to maintain a balance between two possible approaches to the theory varying by the degree of emphasis on time. A first approach contemplates equilibrium in time merely as the particular case of the general theory developed in the previous chapters in which commodities are indexed by time as one of the many characteristics defining them. This is a useful point of view (the display of the underlying unity of seemingly disparate phenomena is one of the prime roles of theory), and to a point we build on it. However, exclusive reliance on this approach would, in the limit, be self-defeating. It would reduce this chapter to a footnote to the preceding ones. A second approach proceeds by stressing, rather than deemphasizing, the special structure of time. Again, we follow this line to some degree. Thus, every model discussed in this chapter accepts the open-ended infinity of time, or the fact that production takes time. Also, at the cost of some generality, we pursue our treatment under assumptions of stationarity and time separability that allow for a sharp presentation of the dynamic aspects of the theory. Sections 20.B and 20.C are concerned with the description of, respectively, the consumption and the production sides of the economy. Section 20.D is the heart of this chapter. It deals with the basic properties of equilibria (including definitions, existence, optimality, and computability) in the context of a single-consumer economy. Section 20.E (which concentrates on steady states) and Section 20.F (which is general) study the dynamics of the single-consumer model. Section 20.G considers economies with several consumers. The message of this section is that, as long as the Pareto optimality of equilibrium is guaranteed, the qualitative aspects of the positive theory of Chapter 17 extend to the more general situation and, moreover, that the properties of individual equilibria identified by the single-consumer methodology remain valid in the broader context. Section 20.H gives an extremely succinct account of overlapping-generations 732
INTERTEMPORAL
economies, a model of central importance in modern macroeconomic theory. Our interest in it is twofold: on the one hand, we want to display it as yet another instance of a useful equilibrium model; on the other hand, we want to point out that it is an example that, because of the infinity of generations, does not fit the general model of Section 20.G, and one that gives rise to some new and interesting issues having to do with the optimality and the multiplicity of equilibria. Section 20.1 gathers some remarks on nonequilibrium considerations (short-run equilibrium and tatonnement stability, learning, and so on). For pedagogical purposes, the entire chapter deals only with the deterministic version of the theory. The unfolding of time is a line, not a tree. A full synthesis of the approaches of Chapter 19 (on uncertainty) and the current one (on time) is possible. However, we view its presentation as advanced material beyond the scope of this textbook. The account of Stokey and Lucas with Prescott (1989) constitutes an excellent introduction to the general theory. A point of notation: in this chapter always means that is, lim T _" When the sum docs not run from t = 0 to t = 00 the two end-points of the sum are explicitly indicated.
I,
20.A Introduction
20.8:
UTILITY
733
----------------------------------------------------------------------------------
I:=O',
I::r
20.B Intertemporal Utility In this chapter, we assume that there are infinitely many dates t = 0, I, ... , and that the objects of choice for consumers are cOllsumption streams c = (co' ... ,c" . .. ) where c, E IR'•• C, 2': O. I To keep things simple, we will consider only consumption streams that arc hOUllded, that is, that have sup, IIc,lI < 00. Rather than proceed from the most general form of preferences over consumption streams to the more specific, we instead introduce first the very special form that we assume throughout this chapter (except for Sections 20.H and 20.1); we subsequently discuss its special properties from a general point of view. It is customary in intertemporal economies to assume that preferences over consumption streams c = (co, ... , c" ... ) can be represented by a utility function V(c) having the special form
V(c) =
L'" c5'u(c,)
(20.B.I)
,-0
where c5 < I is a discount factor and u(·), which is defined on R~, is strictly increasing and concave. This chapter will be no exception to this rule: Throughout it we assume that preferences over consumption streams take this form. However, we comment here. at some length, on six aspects of this utility function. As a matter of notation, given a consumption stream c = (co, ... , c" ... ), we let c T = (c~, cf, . .. ) denote the T-pcriod "backward shift" consumption stream, that is, the stream (c~, with c," = c, +'f for all t 2': O.
cr.... )
(I) Time impalience. The requirement that future utility is discounted (i.e., that t5 < I), implies time impatience. That is, if C = (CO,ch .•. ,c" ... ) is a nonzero consumption stream, then the (forward-) shifted consumption stream c' = (0, Co. c" ... , c,_, •... ) is strictly worse than C (see Exercise 20.B.I). It is an I. We use the terms "stream," "trajectory." "program," and "path" synonymously.
734
c HAP T E R
2 .:
E QUI L I 8 R I U MAN D
TIM E
SEC T t 0 H
2 •• 8:
I HTE RT E M P 0 AA L
----------------------------------------------------------------------------------------~
assumption that is very helpful in guaranteeing that a bounded consumption stream has a finite utility value [i.e., guarantees that the sum in (20.B.1) converges], thus allowing us to compare any two such consumption streams' and making possible the application of the machinery of the calculus. There is a strand of opinion that views this technical convenience as the real reason for the fundamental role that the assumption of time discounting plays in economics. This skeptical view on the existence of substantive reasons' is excessive. An implication of time discounting is that the distant future does not matter much for current decisions, and this feature seems more realistic than its opposite. A possible interpretation, and defense, of the discount factor a views it as a probability of survival to the next period. Then V(e) is the expected value of lifetime utility. For another interpretation, see (6) below. (2)
Srarionariry. A more general form of the utility function would be V(c)
'" u.(e.). =L
(20.B.2)
1-0
The form (20.B.I) corresponds to the special case of(20.B.2) in which u,(c,) = a'u(c, ). This special form can be characterized in terms of starionariry. Consider two consumption streams e ¢ e' such that c, = c; for r ~ T - I: that is, the two streams c and e' are one and the same up to period T - I and differ only after T - I. Observe that the problem of choosing at t = T between the current and future consumptions in e and e' is the same problem that a consumer would face at t = 0 in choosing between the consumption streams eT and e'T, the T backward shifts of e and c', respectively. Then stationarily requires that V(e)
if and only if
V(e T )
It is a good exercise to verify that (20.B.I) satisfies the stationarity property and that the property can be violated by utility functions of the form V(e) = L, a;u(e,), that is, with a time-dependent discount factor (Exercise 20.B.2). The property of stationarity should nol be confused with the statement asserting that if the consumption streams c and e' coincide in the first T - I periods and a consumer chooses one of these streams at t = 0, then she will not change her mind at T. This "property" is tautologically true: at both dates we are comparing Vee) and V(e').· The stationarity experiment compares V(e) and V(e') at I = 0, but at period T it compares the utility values of the future streams shifted to t = 0, that is, V(e T ) and V(C'T). Thus, stationarity says that in the context of the form (20.B.2), the preferences over the future are independent of the age of the decision maker. Time stationarity is not essential to the analysis of this chapter (except for Sections 20.E and 20.F on dynamics), but it saves substantially on the use of subindices.
(3) Additive separability. Two implications of the additive form of the utility function are that at any date T we have, first, that the induced ordering on consumption streams that begin at T + I is independent of the consumption stream followed from 0 to T, and, second, that the ordering on consumption streams from o to T is independent of whatever (fixed) consumption expectation we may have from T + I onward (see Exercise 20.B.3). In turn, these two separability properties imply additivity; that is, if the preference ordering over consumption streams satisfies these separability properties, then it can be represented by a utility function of the form Vee) = L, u,(c,) [this is not easy to prove, see Blackorby, Primont and Russell (1978)]. How restrictive is the assumption of additive separability? We can make two arguments in its favor: the first is technical convenience; the second is a vague sense that what happens far in the future or in the past should be irrelevant to the relative welfare appreciation of current consumption alternatives. Against it we have obvious counter-examples: Past consumption creates habits and addictions, the appreciation of a particularly wonderful dish may depend on how many times it has been consumed in the last week, and so on. There is, however, a very natural way to accommodate these phenomena within an additively separable framework. We could, for example, allow for the form V(e) = L. U,(C'_I' c,). Here the utility at period t depends not only on consumption at date I but also on consumption at date t - I (or, more generally, on consumption at several past dates). We can formulate this in a slightly different way. Define a vector Z, of "habit" variables and a household producrioll recilllology that uses an input vector C'-I at t - I to jointly produce an output vector C _ I of consumption goods at I - I and a vector z, = C._I of "habit" ' variables at I. Then, formally, u. depends only on time t variables and total utility is L. u,(z" c,). In summary: additive separability is less restrictive than it appears if we allow for household production and a suitable number (typically larger than I) of current variables. (4) Lellglh of period. The plausibility of the separability assumption, which makes the enjoyment of current consumption independent of the consumption in other periods, depends on the length of the period. Because even the most perishable consumption goods have elements of durability in them (in the form, for example, of a flow of "services" after the act of consumption), the assumption is quite strained if the length of the elementary period is very short. What determines the length of the period? To the extent that our model is geared to competitive theory, this period is institutionally determined: it should be an interval of time for which prices can be taken as constant. On a related point, note that the value of fJ also depends, implicitly, on the length of the period. The shorter the period, the closer should be to 1.
a
(5) Recursive uriliry. With the form (20.B.I) for the utility function, we have = u(c o) + aV(c ' ) for any consumption stream C = (co, C I , ••• ,e... ..). If we think of u = u(co) as current utility and of V = V(e ' ) as future utility, we see that the marginal rate of substitution of current for future utility equals and is therefore independent of the levels of current and future utility. The recursive utility model [due to Koopmans (1960)] is a useful generalization of (20.B.I) that combines two features: it allows this rate to be variable but, as in the additively separable case, it has the property that the ordering of future consumption streams is independent of the consumption stream followed in the past. Vee)
2. Hence, the completeness of the preference relation on consumption streams is guaranteed. 3. Ramsey (t928) called the assumption a 'weakness of the imagination." 4. This property is often called lime consistency. Time inconsistency is possible if tastes change through time (recall the example of Ulysses and the Sirens in Section 1.8!), but, as we have just argued, it must necessarily hold if the preference ordering over consumption streams (co . ...• c" . .. ) does not change as lime passes. In line with the entire treatment of Part IV, we maintain the assumption of unchanging tastes throughout the chapter.
a
U TtL I T Y
735
736
CHAPTER
20:
EQUILIBRIUM
AND
TIME
The recursive model goes as follows. Denote current utility by II ~ 0 and future utility by V ~ O. Then we are given a current utility function II(C,) and an aggregator function G(II, V) that combines current and future utility into overall utility. For example, in the separable additive case we have G(II, V) = II + cS V. More generally we could also have, for example, G(u, V) = II' + cSV', 0 < " S I. In this case, the indifference curves in the (II, V) plane arc not straight lines. The utility of a consumption stream c = (co, ... , c" . .. ) could then be computed recursively from (20.8.3) V(c) = G(u(c o), V(e')) = G(u(co), G(u(e,), Vee'))) = .... For (20.B.3) to make sense we must be able to argue that the influence of V(e T ) on V(c) will become negligible as T ~ 00 [so that Vee) can be approximately determined by taking a large T and letting V(CT) have an arbitrary value]. This amounts to an assumption of time impatience. In applications, it will typically not be necessary to compute V(c) explicitly. See Exercise 20.B.4 for more on recursive utility.
(6) Altruism. The expression V(c) = u(co) + b V(c') suggests a multigeneration interpretation of the single-consumer problem (20.B.I). Indeed, if generations live a single period and we think of generation 0 as enjoying her consumption according to u(co), but caring also about the utility V(c') of the next generation according to f, V(c'), then V(c) = u(co) + bV(c') is her overall utility. If every generation is similarly altruistic, then we conclude, by recursive substitution, that the objective function of generation 0 is precisely (20.B.I). The entire "dynasty" behaves as a single individual. With this we also have another justification for b < I. The inequality means then that the members of the current generation care for their children, but not quite as much as for themselves. See Barro (1989) for more on these points.
20.C Intertemporal Production and Efficiency Assume that there is an infinite sequence of dates t = 0, I, .... In each period t, there are L commodities. If it facilitates reading, you can take L = 2 and interpret the commodities as labor services and a generalized consumption-investment good (see Example 20.C.1). One of the great advantages of vector notation, however, is that in some cases-and this is one-there is no novelty involved in the general case. Thus, while you think you are understanding the simple problem, you are at the same time understanding the most general one. We shall adopt the convention that goods are nondurables. This is a convention because, in order to make a good durable, it suffices to specify a storage technology whose role is, so to speak, to transport the commodity through time. lfwe were exogenously endowed with some amount of resources (e.g., some initial capital and some amount of labor every period), we would ask what we could do with them. To give an answer, we need to specify the production technology. We already know from Chapter 5 how to do this formally by means of the concept of a production set (or a production transformation function, or a production function). With minimal loss of generality, we will restrict our technologies to be of the following form: the production possibilities at time t are entirely determined by the production decisions at the most recent past, that is, at time t - I. If we keep in mind that we can always define new intermediate goods (such as different vintages of a machine),
SEC T 10"
2 0 • C:
I N T E R T E M P 0 R ALP ROD U C T I 0"
A" 0
E F Fie lEN C Y
737
------------------------------------------------------and also that we can always define periods to be very long, we see that the restriction is minor. Thus, the technological possibilities at t will be formally specified by a production set Y c R2L whose generic entries, or production plans, are written y = (Y., y.). The indices b and a are mnemonic for "before" and "after." The interpretation is that the production plans in Y cover two periods (the "initial" and the "last" period) with y. E IRL and y. E IRL being, respectively, the production plans for the initial and the last periods. As usual, negative entries represent inputs and positive entries represent outputs. We impose some assumptions on Y that are familiar from Section 5.B: (i) Y is dosed and convex. (ii) Y n IR~/ = {O} (no free lunch). (iii) Y - IR~L C Y (free disposal). An assumption specific to the temporal setting is the requirement that inputs not be used later than outputs are produced (i.e., production takes time). This is captured by (iv) If)" = (Yo,Y.)E Y then (y.,O)E Y (possibility oJtrllllcation). In words, (iv) says that, whatever the production plans for the initial period, not producing in the last period is a possibility. A simple case is when y.. ~ 0 for every r E Y, that is, when all inputs are used in the initial period. Then (iv) is implied by the free-disposal property (iii). Example 20.C.I: Ramsey-Solow Model.' Assume that there are only two commodities: A consumption-investment good and labor. It will be convenient to describe the technology by a production function F(k, I). To any amounts of capital investment k ;:>: 0 and of labor input I ~ 0, applied in the initial period, the production function assigns the total amount F(k, I) of consumption-investment good available at the last period. Then
Y = {( -k, -I, x, 0): k
~ 0,
I
~ 0, x S;
F(k, I)} -
R~.
Note that labor is a primary factor; that is, it cannot be produced. _ Example 20.C.2: Cost-oJ-Adjustmem Model. Suppose that there are three goods: capacity, a consumption good, and labor. With the amounts k and I of invested capacity and labor at the initial period, one gets F(k, I) units of consumption good output at the last period. This output can be transformed into invested capacity at the last period at a cost of k' + y(k' - k) units of consumption good for k' units of capacity, where y(.) is a convex function satisfying y(k' - k) = 0 for k' < k and i'(k' - k) > 0 for k' > k. The term y(k' - k) represents the cost of adjusting capacity upward in a given period relative to the previous period. (Note the marginal cost of doing so increases with invested capacity of the period.) Formally, the production set Y is Y = {( - k, 0, -I, k', x, 0): k <: 0, I ;:>: 0, k' ;:>: 0, x
:$;
F(k, I) - k' - y(k' - k)} -
I\l~. _
5. See Ramsey (1928) and Solow (1956). The same model was atso inlroduced in Swan (1956).
738
CHAPTER
20:
EQUILIBRIUM
AND
TIME
--------------------------------------------------------------------------------Example 20.C.3: Two-Sec/or Model. We could make a more general distinction between an investment and a consumption good than the one embodied in Examples 20.C.1 and 20.C.2. Indeed. we could let the production set be 1
I 1
y
= {( -k, 0, -I, k', x, 0): k?; O,I?; 0, k'?; 0, x $
G(k, I, k')} - R~,
where k, k' are, respectively, the investments in the initial and the last periods. Note that the investment and the consumption good need not be perfectly substitutable [they are produced in two separate sectors, so to speak; see Uzawa (\964)]. If they are [i.e., if the transformation function .G(k, I, k') has the form F(k, I) - k'] then this example is equivalent to the Ramsey-Solow model of Example 20.C.t. If it has the form G(k.I, k') = F(k, I) - k' - y(k' - k) then we have the cost·of-adjustment model of Example 20.C.2. • Example 20.C.4: (N + I)-Sec/or Model. As in Example 20.C.3, we have a consumption good and labor, but we now interpret k and k' as N-dimensional vectors. For simplicity of exposition, in Example 20.C.3 we have taken G(k,I, k') to be defined for any k ?; 0, k' ?; O. In general, however, this could lead to the production of negative amounts of consumption good. To avoid this it is convenient to complete the specification by means of an admissible domain A of (k,I, k') combinations. Then
y=
I( -k, 0,
-I, k', x, 0): (k,I, k') E A and x $ G(k, I, k')} -
R~(H+2J.
•
Once we have specified our technology, we can define what constitutes a path of production plans. Definition 20.C.1: The list (Yo' y" ... , y" ... ) is a production path, or trajectory, or program, il y, EYe RIL lor every t. Note that along a production path (Yo,' .• , y" •• . ) there is overlap in the time indices over which the production plans y,_, and y, are defined. Indeed, both Y•. ,_' E RL and Y.. E RL represent plans, made respectively at dates / - I and /, for input use or output production at date /. Thus, we have. at every /, a net input-output vector equal to Y•. ,_, + Y., E RL (at / = 0, we put Y•. _I = 0; this convention is kept throughout the chapter)." The negative entries or'this vector stand for amounts of inputs that have to be injected from the outside at period / if the path is to be realized, that is, amounts of input required at period / for the operation of y,_1 and y, in excess of the amounts provided as outputs by the operation of y,_1 and y,. Similarly, the positive entries represent the amounts of goods left over after input use and thus available for final consumption at time /. The situation is entirely analogous to the description of the production side of an economy in Chapter 5. If we think of the technology at every / as being run by a distinct firm (or as an aggregate of distinct firms) and of p, as an infinite sequence with nonzero entries (equal to y,) only in the / and / + I places, then L, p, is the aggregate production path; and it is also precisely the sequence that assigns the net input-output vector Y•. ,_, + Yo, e RL to period t. If we had a finite horizon, the current setting would thus be a particular case of the description of production in 6. A minor point of notation: when there is any possibility of confusion or ambiguity in the reading of indices. we insert commas; for example. we write Y•. ,_I instead of Y.'_I"
SEC.TION
20.C:
'NTERTEMPORAL
PRODUCT'ON
AND
EFFICIENCY
739
-----------------------------------------------------------------------------Chapter 5. With an infinite horizon there is a difference: we now have a countable infinity of commodities and of firms instead of only a finite number. As we shall see. this is not a minor difference. It will. however. be most helpful to arrange our discussion around the exploration of the analogy with the finite horizon case by asking the same questions we posed in Section 5.F regarding the relationship between efficient production plans and price equilibria. Definition 20.C.2: The production path (Yo, ...• y" ... ) is efficient il there is no other production path (Yo, ... , y;, ... ) such that Y•. '-1
+ Yb' $
y~.'-I
+ Yb'
lor all t,
and equality does not hold lor at least one t. I n words: the path (Yo,"" y.. ... ) is efficient if there is no way that we can produce at least as much final consumption in every period using at most the same amount of inputs in every period (with at least one inequality strict). The definition is exactly parallel to Definition 5.F. I. What constitutes a price vector in the current intertemporal context? It is natural to define it as a sequence (Po, p" . .. ,p" . .. ), where p, e RL. For the moment we shall not ask where this sequence comes from. We assume that it is somehow given and that it is available to any possible production unit. The prices should be thought of as present-value prices. We shall discuss further the nature of these prices in the nex t section. Given a path (Yo, ... ,y" ... ) and a price sequence (Po, . .. ,p" ... ), the profit level associated with the production plan at / is
P'·Ybl
+ P,+l"Ys,'
We now pursue the implications of profit maximization on the production plans made period by period. Definition 20.C.3: The production path (Yo" .. , y" ... ) is myopically, or short-run, profit maximizing for the price sequence (Po' ... ,p" ... ) il lor every t we have P,'Yb, + P'+I·Y.,?; P,'Yb, + P'+I'Y~,
lor all y;e Y.
Prices (Po,' .. ,p" . .. j capable of sustaining a path (y" . ..• y ... .. ) as myopically profit-maximizing are often called Malinvaud prices for the path [because of Malinvaud (1953»).' Does the first welfare theorem hold for myopic profit maximization? That is, if (Yo .... ,Y.. ... ) is myopically profit maximizing with respect to strictly positive prices, does it follow that (Yo, ... , y ..... ) is efficient? In a finite-horizon economy this conclusion holds true because of Proposition 5.F.I, but a little thought reveals that in the infinite-horizon context it need not. The intuition for a negative answer rests on the phenomenon of capical overaccumulation. Suppose that prices increase through 7. Observe that we do not require that L. P,"(Y•. I_I + Ybl) < 00. In principle. a production path may have an infinite present value. We saw in Sections S.E and S.F. where we had a finite number of commodities and firms that individual. decentralized profit maximization and overall profit maximization amounted to the same thing. Because of the possibility of an infinite present value. the existence of a countable number of commodities and production sets makes this a more delicate matter in the current context. See Exercises 20.C.2 to 20.C.S for a discussion.
740
C HAP T E R
20:
E QUI LIB R I U MAN D
TIM E
SECTION
---------------------------------------------------------------------------------------time fast enough. Then it may very well happen that at every single period it always pays to invest everything at hand. Along such a path, consumption never takes place-hardly an efficient outcome. Example 20.C.5: With L = I, let Y = ( -k, k'): k ~ 0, k' $ k} C R2. This is just a trivial storage technology. Consider the path where Y, = (- 1,1) for all t; that is, we always carry forward one unit of good. Then Y•. _I + Y.o = - I and Y•. ,_I + Y., = 0 for alit> O. This is not efficient; just consider the path Y; = (0,0) for alit, which has Y~.'_I + = 0 for all t ~ O. But for the stationary price sequence where p, = I for all t, (Yo, ... , y" ... ) is myopically profit maximizing. _
Proposition 20.C,1: Suppose that the production path (Yo" .. , Yr" .. ) is myopically profit maximizing with respect to the price sequence (Po' ... , Pr , ••. ) » O. Suppose also that the production path and the price sequence satisfy the transversality condition Pr+,'Y., -+ O. Then the path (Yo.·'" yr' ... ) is efficient. Proof: Suppose that the path (y~ •...• y;, ... ) is such that Y•. ,_I + Y., $ Y;.'_I + Yb' for all t, with equality not holding for at least one t. Then there is e > 0 such that if we take a T sufficiently large for some strict inequality to correspond to a date previous to T, we must have T
I
.=0
T
p"(Y~.'-1 + Yb') >
I
p,'(Y•. '-1
+ Y.,) + e.
1=0
In fact, if T is very large then PT+ I • Y.T is very small (because of the transversality condition) and therefore T
I
I
T
T-l
PT' YOT
+
I (p,+ .=0
I'
y~,
+ P"Yb') >
I (p,+ .=0
I ' Y.,
+ p,' Yo,).
We must thus have either p, + 1 'y~, + p,' Y" > p,+ I 'Y., + p,' Y., for some t ::; T - I or PT' YOT > PT+ 1 'Y.T + PT' Y.T' In either case we obtain a violation of the myopic profit-maximization assumption [recall that by the possibility of truncation we have (YbT' 0) E Y]. Therefore, no such path (y~, ... , y;, ... ) can exist. Note that the essence of the argument is very simple. The key fact is that if the transversality condition holds, then for T large enough we can approximate the overall profits of the truncated path (Yo, ... , YT) by the sum of the net values of period-by-period input-output realizations (up to period T). It does not matter whether we match the inputs and the outputs per period or per firm (that is, "per production plan "). If the horizon is far enough away, either method will come down to Profits = Total Revenue - Total Cost. _
AND
EFFICIENCY
741
(ii) If the answer to (i) is yes, can we conclude that the pair (Yo, ... , y" ... ). (Po, ... , p" ... ) satisfies the transversality condition? Ti,e allswer to (ii) is "not necessarily." In Section 20.E we will see, by means of an example, that the transversality condition is definitely not a necessary property of Malinvaud prices. The answer to (i) is .. Essentially yes." We illustrate the matter by means of two examples and then conclude this section by a small-type discussion of the general situation.
Example 20,e.6: Ramsey-Solow Model Continued. In this model, we can summarize a path by the sequence (k" I" c,) of total capital usage, labor usage, and amount available for consumption. From now on we assume that k,+ 1 + c,+ I = F(k" I,) and that the sequence I, of labor inputs is exogenously given. Then it is enough to specify the capital path (k o, .•. , k" .• .). Denoting by (q" w,) the prices of the two commodities at t, we have that profits at tare q,+,F(k"I,) - q,k, - w,l, and, therefore, the necessary and sufficient conditions for short-run profit maximization at ( are
~
q,+
p"(Y'.'-1 +y.,).
By rearranging terms-a standard trick in dynamic economics-this can be rewritten as (recall the convention Y•. _I = Y~._I = 0)
PRODUCTION
(i) Is there a system of Malinvaud prices (Po, ... , p" .. .) for (Yo, ... , y" . .. ), that is, a sequence (Po,""p" ... ) with respect to which (Yo, ... ,Y" ... ) is myopically profit maximizing?
T
p,'(Y;.'-1 +yb,»PT+I'Y.T+
INTERTEMPORAL
Proposition 20.e. I tells us that a modified version of the first welfare theorem holds in the dynamic production setting. Let us now ask about the second welfare theorem: Given an efficient path (Yo, ... , y" ... ), can it be price supported? In Proposition 5.F.2 we gave a positive answer to this question which applies to the finite-horizon case. In the current infinite-horizon situation we could decompose the question into two parts:
Y',
Efficiency will obtain if, in addition to myopic profit maximization, the (present) value of the production path becomes insignificant as t ... 00. Precisely, efficiency obtains if the (present) value of the period t production plan for period t + I goes to zero, that is, if p,+ I' Y., ... 0 as t -+ 00. This is the so-called transversality condition. Note that the condition is violated in the storage illustration of Example 20.C.5.
20.C:
--------------------------------------------------------------------------------------
= V,F(k"I,)
and
I
~ = V2 F(k" I,).
q,.l
Notc that, up to a normalization (we could put qo = I), these first-order con· ditions determine supporting prices for any feasible capital path (see Exercise 20.e.6). The transversality condition says that q,+ ,F(k" I,) -+ O. If the sequence of productions F(k" I,) is bounded, then it suffices that q, -+ O. In view of Proposition 2a.e.I. we can conclude that a set of sufficient conditions for efficiency of a feasible and bounded capital path (k o, ..• , k" ... ) is that there exist a sequence of output prices (qo, ... , q" ... ) such that
~ = V,F(k" I,)
for alit
(20.C.I)
q,+1
and
q,
-+
a
(equivalently, Ijq,
-+ 00).
(20.C.2)
Because of the possibility of capital overaccumulation, (2a.e.I). which is necessary, is not alone sufficient for efficiency. On the other hand, (20.e.2) is not necessary (see Section 20.E). Cass (1972) obtained a weakened version of (20.e.2) that, with (2a.c.I),
-
742
CHAPTER
20:
EQUILIBRIUM
AND
SECTIQN
TIME
--------------------------------------------------------------------------------------
20.0:
EQUILIBRIUM:
THE
ONE-CQNSUMER
CASE
743
---------------------------------------------------------------------------------------
is both necessary and sufficient.· The condition is
"" I
L-= ,-0 q,
(20.C.2')
00.
-
Example 20.C.7: Cost of Adjustment Model continued. In the cost or adjustment model. a production plan at time t - I involves the variables k,_ .. 1,_1' k,. c,. We associate with these variables the prices q,_I' q,. s,. Profits are then
w,_"
s,(F(k,.,.I,.,) - k, -y(k,...,. k,_,))
+ q,k, -
Production Possibility Set . Y () ({Y•. ,_,) X RL)
q,_,k,., - w,.,I,.,.
(Y •. <-I fixed)
Using the first-order profit-maximization conditions with respect to k, and k,., we get the rollowing two conditions: (i) q, = s,(1 + y'(k, - k,.,»; that is. the price or capacity at r equals the investment cost in extra capacity at t. (ii) q,_, = s,(V,F(k,_,.I,_,) + y'(k, - k,_,»; that is. the price or capacity at r - I equals the return at t of one unit or extra capacity at t - I (the return has two parts: the increased production at t and the saving in the cost or capacity adjustment at t). Combining (i) and (ii).
q,_,
V,F(k,_,.I,.,)
q,
I
+ y'(k,
+ y'(k, -
- k,.,)
(20.C.3)
k,_,)
Note that ir there are no adjustment costs [i.e .• if y(.) is identically equal to zero]. then (20.C.3) is precisely (20.C.1). Observe also that. in parallel to Example 20.C.6. condition (20.C.3) determines short·run supporting prices ror any reasible capital path. _
In a general smooth model it is not difficult to explain how the supporting prices (Po • ...• P.. ' .. ) for an efficient path (yo •...• y" .. .) can be constructed. Note that. because of efficiency. every y, belongs to the boundary of Y. The smoothness property that we require is that. for every r. the production set Y has a single (normalized) outward normal q, = (q". q.,) at )" (we could. for example. normalize q, to have unit length); see Figure 20.C.1. Less geometrically. smoothness means that at y, E Y all the marginal rates of transformation (M RT) of inputs for inputs. inputs for outputs. and outputs for outputs are uniquely defined. We claim that the efficiency property implies that for every r we have that q•. "1 = {Jq" for some {J > O. Heuristically: for any two commodities their MRT as outputs at r for the production decision taken at time r - 1 must be the same as their M R T as inputs at I for the production decision taken at time r. If this were not so. it would be possible to generate a surplus of goods. The argument is standard (;ecallthe analysis of Section l6.F). Consider. for example. Figure 20.C.2. where in panel (a) we have drawn the output transformation frontier through Y•• , ' I (Le .• keeping Y,.,., fixed) and in panel (b) the input isoquant through Yo, (i.e .• keeping Y., fixed; recall the sign conventions for inputs). We see that if the slopes at these points are not the same. then it is possible to move from Y... r-I to Y;.,-l and from Ybl to y~, in such a way that
y; .• _1 + y., > Y•. ,_I + YIN,
thus contradicting efficiency.
(a)
(b)
We construct the desired price sequence (Po •...• P, •.. ·) by induction. Put Po = q", (i.e.• the relative prices at r = 0 are the M RTs between goods at the initial part of the production plan Yo EYe R'L). Suppose now that the prices (Po •...• PT) have already been determined, and that every y, up to I = T - 1 is myopically profit maximizing for these prices. Because of the first-order conditions for profit maximization at T - I. we have that PT = aq•. T ' I for some a > O. We know that q•. T ' I = {Jq .. for some {J> O. Then PT = a{Jq'T' Therefore. if we put p,.. 1 = a/iq.", we have that (PT' PT.,) = (a/iq'T' a{Jq.T) is proportional to qT = (q'T' q.T). which means that YT is profit maximizing for (PT. PT. I)' Hence we have extended our sequence to (Po • ...• PT. I) and we can keep going. Note that, as in Examples 20.C.6 and 20.C.7. the construction of the supporting short-run prices does not make full use of the efficiency. What is used is that the production path is "short-run efficient" (that is. the production path cannot be shown inefficient by changes in the production plans at a finite number of dates). The above observations can be made into a perfectly rigorous argument for the existence of Malinvaud prices in the smooth case. The proof for the nonsmooth case is more complex. It must combine an appeal to the separating hyperplane theorem (to get prices for truncated horizons) with a limit operation as the horizon goes to infinity. With a minor technical condition (call IlOnlighrness in the literature). this limit operation can be carried out.
20.D Equilibrium: The One-Consumer Case In this section. we bring the consumption and the production sides together and begin the study or equilibrium in the intertemporal setting. We shall start with the one-consumer case. As we will see in Section 20.G. the relevance or this case goes beyond the domain or applicability or the representative consumer theory or Chapter 4. An economy is specified by a short-term production technology Y c R2L, a utility Junct ion u( . ) defined on R~. a discount factor lJ < I, and, finally, a (bounded) sequence or initial endowments (W o," .• w" .. .), w, e R~. We assume that Y satisfies hypotheses (i) to (iv) or Section 20.C and that u(·) is strictly concave. differentiable. and has strictly positive marginal utilities throughout its
domain. 8. Some additional, very minor, regularity conditions on the production runction F(') are required ror the validity or this equivalence.
Prices are given to us as sequences (Po •. ..• p" . .. ) with p, e R~. As in Chapter 19 we can interpret these prices either as the prices or a complete system of rorward
Figure 2O.C.l (ten)
Smooth production scI. Flgur. 20.C.2 (rtgh')
A production path that is inefficient at T.
744
CHAPTER
20:
EQUILIBRIUM
AND
markets occurring simultaneously at t = 0 or as the correctly anticipated (present value) prices of a sequence of spot markets. We will consider only bounded price sequences. In fact, most of the time we will have IIp,lI _ 0. 9 Given a production path (Yo,"" y.. ... ), y, E Y, the induced stream of consumptions (co, ... , c,' ... ) is given by C,
+ P'+I-Y.'
Fixing T and rearranging the terms of we get
L
(It,+p,'w,)-
L, '" T p,'C, = L, '" T p'(Y•. ,_, + Y., + w,)
L
p,·C,=PT+'·Y.T
(20.0.1)
20. D:
E QUI L' • R , U M:
THE
0 N E - CON SUM E RCA S E
there is a forward market for every commodity at every date, or, in another, that assets (e.g., money) are available that are capable of transferring purchasing power Ihrough time (see Exercise 20.0.1 for more on this). Secondly, observe that the strict monotonicity of u(·) implies that if we have reached utility maximization then, a fortiori, total wealth (denoted w) must be finite; that is,
Lit, + LP,'w, < 00.
w=
for every t.
T , ~"
Moreover, at the equilibrium consumptions the budget constraint of (20.0.4) must hold with equality. An important consequence of the last observation is that at equilibrium the transversality condition is satisfied. Formally, we have Proposition 20.D.1. Proposition 20.0.1: Suppose that the (bounded) production path (yt,···, vi, . .. ) and the (bounded) price sequence (Po, ... ,p" ... ) constitute a Walrasian equilibrium. Then the transversality condition P,+,' 0 holds.
Y:, -
Proof: Denote
c: = Y:.,_, + y:' + w,. By expression (20.0.1) we have L (It,+p,'w,)- L p,·C,=PT+'·Y.T· 1$
T
IS T
Since cach of the sums in the left-hand side converges to w < conclude that PH' • Y:T -+ O. •
00
as T
-+ 00.
we
IS T
IS T
Expression (20.0.1) is an important identity. It tells us that the transversality condition is equivalent to the overall value of consumption not being strictly inferior to wealth (i.e., there is no escape of purchasing power at infinity). The definition of a Walrasian equilibrium is now as in the previous chapters. One only has to make sure that a few infinite sums make sense. Definition 20.0.1: The (bounded) production path (y~, ... , y~" .. ), y~ E Y, and the (bounded) price sequence p = (Po' ... ,p" ... ) constitute a Walrasian (or competitive) equilibrium if:
Y:.,_, Y6, + w, ~ 0
(i) c~ = + (ii) For every t,
1t,
for all t.
(20.0.2)
= P,'Yb, + p,+,'V:, ~ P,'Yb + p,+,·Y.
(20.0.3)
for all Y = (Vb' Y.) E Y. (iii) The consumption sequence (ct, ... ,cr . .. ) ~ 0 solves the problem Max
L, o'u(c,) s.t.
(20.0.4)
L,P,'c, ~ L,1t, + L,p,·w,.
Condition (i) is the feasibility requirement. Condition (ii) is the short-run, or myopic, profit-maximization condition already considered in Section 20.C (Definition 20.C.3). The form of the budget constraint in part (iii) deserves comment. Note first that there is a single budget constraint. As in Chapter 19, this amounts to an assumption of completeness, which means, in one interpretation, that at time t = 0 9. Keep in mind that prices are to be thought
or as
measured in current-value terms.
745
---------------------------------------------------------------------------------------
+ Y'u + 00"
= Y.... _I
If c, ~ 0 for every t, then we say that the production path (Yo, ... ,y" . .. ) is feasible: Given the initial endowment stream the production path is capable of sustaining nonnegative consumptions at every period. To keep the exposition manageable from now on we restrict all our production paths and consumption streams to be bounded. Delicate points come up in the general case, which are better avoided in a first approach. Alternatively, we could simply assume that our technology is such that any feasible production path is bounded. Given a production path (Yo, . .. ,Y... .. ) and a price sequence (Po, ... ,P.. · .. ), the induced stream of profits (Ito, ... , It" ... ) is given by
n, = P,·Y'n
s [C
TIME
-----------------------------------------------------------------------------------
Another implication of Definition 20.0.1 by
w
<
00
is the possibility of replacing condition (ii) of
(ii') The production path (Y6,"" Y:,. ,,) maximizes total profits, in the sense that for any other path (Yo,' , , ,y" . .. ) and any T we have ,=T
L
.=0
(p,' Y.,
+ p,+ "
Y.,) ~
L (p,' Y:, + p,+,' Y:,) < 00. f
Clearly, (ii') implies (ii), and (ii) with w < 00 implies (ii') (see Exercise 20.0.2). Thus, at equilibrium prices, the test of myopic and of overall profit maximization coincide. Could a similar statement be made for an appropriate concept of myopic utility maximization? We now investigate this question. Definition 20.0.2: We say that the consumption stream (co' ... , c,' ... ) is myopically, or short-run, utility maximizing In the budget set determined by (Po' ... ,p" ... ) and w < 00 if utility cannot be increased by a new consumption stream that merely transfers purchasing power between some two consecutive periods. The key fact is presented in Exercise 20.D.3. Exercise 20.D.3: Show that a consumption stream (eo,' .. ,c" . .. ) »0 is short-run utility maximizing for p = (Po.' ..• P, • ...) and w < 00 if and only if it satisfies L, p,' e, = wand the collection of first-order conditions: For every t there is )., > 0 such that ;., p, = Vu(c,)
and
;.,p,+, = oVu(e,+ I)'
(20.0.5)
It follows from (20.0.5) that ;., p, = Vu(c,) and. ).,_, p, = oVule,). Therefore, i"-J = ,\)., and so J. o = 8').,. Hence letting i. = )'0' we see that (20.0.5) is actually
746
C H "P T E R
20:
E 0 U I LIB R I U 101
AND
SECTION
TIM E
--------------------------------------------------------------------------------equivalent to For some ).,
J.p,
= /J'Vu(c,)
for all t.
(20.0.6)
Once we realize that myopic utility maximization in a budget set amounts to
(20.0.6), we can verify that overall utility maximization follows. This is done in Proposition 20.0.2. Proposition 20.0.2: If the consumption stream (CO" .•• c,' ... ) satisfies Lt Pt' ct = W < OCJ and condition (20.0.6), then it is utility maximizing in the budget set determined by (Po' ... 'P t ' ... ) and w. Proof: We firsl nole that we cannot improve upon (co, . .. , c" ... J by transferring purchasing power only through a finite number of dates. Indeed, (20.D.6) implies that the first·order sumcient conditions for any such constrained utility maximization problem are satisfied. Suppose now that (co, . ", c;, ... ) is a consumption stream satisfying the budget constraint and yielding higher total utility. Then for a sufficiently large T, consider the stream (f~ ... " ,.~, ... J with ,.; = c; for I :S Tand c; = c, for I> T. Because /J < I, there is f. > 0 such that if T is large enough then there is an improvement of utility of more than 2e in going from (1'0' ... , "" ••• J to (co, ... , c;, ... J. Since w < 00, the amount I:t> riP,' (c, - c;)1 can be made arbitrarily small. Hence, for large T the stream (c;;' . .. , c;, . ..) is almost budget feasible. It follows that it can be made budget feasible by a small sacrifice of consumption in the first period resulting in a utility loss not larger than e. Overall, it still results in an improvement. But this yields a contradiction because only the consumption in a finite number of periods has been altered in the process. _
20.0:
EOUILtBRIUM:
THE
ONE-CONSUMER
(ii) Suppose instead that Wo = I and w, = 0 for t > O. There is, however, a linear production technology transforming every unit of input at t into IX > 0 units of output at I + I. Because of the boundary behavior of the utility function, consumption will be positive in every period, and therefore the technology will be in operation at every period. The linearity of the technologies then has the important implication that the equilibrium price sequence is completely determined by the technology. Putting Po = I, we must have P, = I/a'. Wealth is IV = PoWo = I, and therefore the equilibrium consumptions must be c: = [.5'(1 - t5)]/p, = (exb)'(1 - b). Note that, as long as I ~ a < I/b, both the price and the consumption sequences are bounded. Observe also the interesting fact that for this example we have been able to compute the equilibrium without explicitly solving for the sequence of capital investments. (iii) We are as in (ii) except that we now have a general technology F(k) transforming every unit k" of investment at t into F(k,) units of output at r + I. This output can then be used indistinctly for consumption or investment purposes at I + I. That is, c,., = F(k,) - k,. I' The logarithmic form of the utility function allows for a shortcut to the computation of equilibrium prices. Indeed, say that (Po," . 'PI' ... ) are equilibrium prices and (c~, ... ,c~, ... ), (k~, ... ,k:, . .. ) equilibrium paths of consumption and capital investment. Then we know that at any T a constant fraction 0 of remaining wealth is invested. That is,
PT.,k~., = b( L
p,c:)
= t5PT+,F(k:J.
'2: T+ 1
Example 20.0.1: In this example we illustrate the use of conditions (20.0.6) for the computation of equilibrium prices. Suppose that we are in a one-commodity world with utility function L, /J' In c,. Given a price sequence (Po,' .. ,p" ... ) and wealth IV, the first-order conditions for utility maximization (20.0.6) are
b'
).p, = c,
for all I,
and
L p,c, = IV. ,
Hence, W = L, p,C, = (1/).) L, b' = (1/).)[1/(1 -.5)] and so p,c, = .5'/i. = .5'(1 - b)1V for all r. Note that this implies a conslant rale of savings because PTcdeL, " T p,c,) = I - 6, for all T (Exercise 20.0.4).'0 We now discuss three possible production scenarios. (i) The economy is of the exchange type; that is, there is no possibility of production and we are given an initial endowment sequence (wo. ... , WI' ... ) » O. Then the equilibrium must involve = w, for every I, and therefore, normalizing to L, p,W, = I, the equilibrium prices should be
c:
0'(1 - b) P,=----
w,
for every t.
10. Logarithmic utility functions facilitate computation and are very important in applications. However, they are not continuous at the boundary (In 1:, _ - 00 as c, - 0) and therefore violate one of our maintained assumptions. This does not affect the current analysis but should be kept in mind.
CASE
747
----------------------------------------------------------------------------------
Therefore, we must have k:., = of(k:J for every I. With ko = Wo = I given, this allows us to iteratively compute the sequence of equilibrium capital investments. The sequence of prices is then obtained from the profitmaximization conditions P,. ,F'(k:J- p, = o.• Since a Walrasian equilibrium is myopically profit maximizing and satisfies the transversality condition (Proposition 20.0.1), we know from Proposition 20.C.1 that it is production efficient (assuming p, » 0 for all I). Can we strengthen this to the claim that the full first welfare theorem holds? We will now verify that we can. In the current one-consumer problem, Pareto optimality simply means that the equilibrium solves the utility-maximization problem under the technological and endowment constraints: Max
L b'u(c,), S.t. (", =
Yu,l-l
(20.0.7)
+ Ybt + w,;,:: 0
y,
and
Proposition 20.0.3: Any Walrasian equilibrium path planning problem (20.D.7).
(V~,
E
Y for all I.
... , V?, . .. ) solves the
Proof: Oenote by B the budget set determined by the Walrasian equilibrium price sequence (Po,· .. , p" ... ) and wealth IV = L,1!, + L, p,·w" where 1[,
= p,'Y:'
+ P,+l'Y:,,+l
748
CHAPTER
EQUILIIIRIUM
20:
AND
TIME
SEC T ION
--------------------------------------------------------------------------------------------for all I. That is.
have
B = {(co•...• c; •.. .): c; ~ 0 for all t and
L, p,'c; !> w}.
By the definition of Walrasian equilibrium. the utility of the stream (c~ •...• c: •... ) defined by = Y:.,_1 + + £0, is maximal in this budget set. It suffices. therefore, to show that any feasible path (y~•. ..• y~, ... ), that is, any path for which y~ E Y and c~ = Y;.,_1 + Y;' + co, ~ 0 for all I, must yield a consumption stream in B. To see this note that, for any T.
c:
Y:,
L
p,'c;=
1ST
L (P,'Yb,+P,+I'Y;,)+Pr'Ybr+ L rST-l
c; »
E QUI L III R I U M:
THE
0 N E • CON. U MER
CAS E
749
0 for all t and. moreover, for it to be legitimate to determine the sign of
L, b'(u(C;) -
u(cn) = ~T(U(CT) - u(cm + b T' '(u(c~+ ,) - u(4+ ,))
by signing the first·order term b T Vu(cf)'(cr - cf) + bT+' Vu(cf+ ,)'(CT" - cf, ,) = PT'(Y;T - rtT) + PT' ,'(Y~r - Y:T) = PTe Y;T
+ PT+ I" Y~T - PTe Y:T - PT+ I" Y:T > O.
But this conclusion contradicts the assumption that
p,·W,.
(y~,
... ,
Y: .... )solves (20.0.7). •
1sT
By the possibility of truncation of production plans, we have (Ybr, 0) E Y. Therefore, by short-run profit maximization, p,' Y,r :$; 1[r and p,' + P, + I • :$; 1[, for all t:$; T- I. Hence,
y"
L
L
p,'c;:$;
1ST
which implies
20. D:
-----------------------------------------------------------------------------
L, p,'c;
1ST
1[,+
L
p,'w,:$;w
y:,
for all T,
1ST
:$; w. •
Let us now ask for the converse of Proposition 20.0.3 (i.e .• for the second welfare theorem question; see chapter 16): Is any solution (Yo, ... , y" ... ) to the planning problem (20.0.7) a Walrasian equilibrium? In essence. the answer is "yes," but the precise theorems are somewhat technical because. to obtain a well-behaved price system (i.e .• a price system as we understand it: a sequence of nonzero prices). one needs some regularity condition on the path. We give an example of one such result." Proposition 20.0.4: Suppose that the (bounded) path (Y6,"" vi, ... ) solves the planning problem (20.0.7) and that it yields strictly positive consumption (in the sense that, for some t > O. en = Yi•. '-l + Yib' + COft > t for all t and t). Then the path is a Walrasian equilibrium with respect to some price sequence (Po.···.P" ... ). Proof: We provide only a sketch of the proof. A possible candidate for an equilibrium price system is suggested by expression (20.0.6): p, = b'VU(Cn
for all t,
y:.,_,
where c: = + Y:' + w,. Because (c~, ...• c: •... ) is bounded above and bounded away from the boundary (uniformly in t) we have L, IIp,1I < 00, which implies the transversality condition. In turn. by expression (20.0.1) this yields L, Poc: = L, (n, + p,'w,) = w < 00. Therefore. by Proposition 20.0.2, the utility·maximization condition holds. I! remains to establish that short·run profit maximization also holds. To that effect suppose that this is not so. that is. that for some T there is y' E Y with PT·Y~
+ PT+t 'y~ > PT'Y:T + PT+ I'Y:T =
nT'
Let (y; •. ..• y;, ... ) be the path with Yr = y' and y; = Y: for any t >F T. Let (co •... , c; •. .. ) be the associated consumption stream. Because of the convexity of Y and the strict positivity for us to property of (c~ •. ..• c: •. .. ) we can assume that YT = Y' is sufficiently close to
yr·
II. A general treatment would involve, as in Sections IS.C or 16.D,lhe application ora suitable
version (here infinite·dimensional) of the separating hyperplane theorem. The next result gets around this by exploiting the differentiability of It(·). I! is thus parallel to the discussion in Section 16.F.
The close connection between the solutions of the equilibrium and the planning problem (20.0.7) has three important implications for. respectively. the existence. uniqueness. and computation of equilibria. The first implication is that it reduces the question of the exiSlence of an equilibrium to the possibility of solving a single optimization problem. albeit an infinite-dimensional one. Proposition 20.0.5: Suppose that there is a uniform bound on the consumption streams generated by all the feasible paths. Then the planning problem (20.0.7) attains a maximum; that is, there is a feasible path that yields utility at least as large as the utility corresponding to any other feasible paths. The proof. which is purely technical and which we skip. involves simply establishing that, in a suitable infinite-dimensional sense. the objective function of problem (20.0.7) is continuous and the constraint set is compact. The second implication is that it allows us to assert the uniqueness of equilibrium. Proposition 20.0.6: The planning problem (20.0.7) has at most one consumption stream solution. Proof: The proof consists of the usual argument showing that the maximum of a strictly concave function in a convex set is unique. Suppose that (yo •...• y, •. .. ) and (yo •...• y; •... ) are feasible paths with L, .5'u(c,) = L, .5'u(c;) = Y. where (co •...• c, •. .. ) and (co •. ..• c; •.. .) are the consumption streams associated with the two production paths. Consider y~ = lY, + !y;. Then the path (y~, ...• y~ •. .. ) is feasible and at every t the consumption level is c~ = !c, + !c;. Hence. L, .5'u(c~) ~ Y. with the inequality strict if c, l' c; for some t. Thus. if c, '# c; for some t. the paths (yo, ... , y, •.. .). (y~ •...• y; •... ) could not both solve (20.0.7). • The third implication is that Proposition 20.0.3 provides a workable approach to the computation of the equilibrium. We devote the rest of this section to elaborating on this point.
The Computation of Equilibrium and Euler Equations It will be convenient to pursue the discussion of computational issues in the slightly restricted setting of Example 20.C.4, the (N + I)-sector model. To recall. we have N capital goods, labor, and a consumption good. We fix the endowments of labor to a constant level through time. A function G(k. k'). gives the total amount of consumption good obtainable at any t if the investment in capital goods at t - I is
-
750
c HAP T E R
2 0:
E QUI LIB R I U"
AND
T
I .. E
SEC T ION
20.
D: E QUI LIB R I
U .. :
THE
0
NE· CON. U .. E RCA. E
751
-------------------------------------------------------------------------------------------given by the vector k E RH. the investment at t is required to be k' E R~. and the labor usage at t - I and t is fixed at the level exogenously given by the initial endowments. We denote by A c RH X RH the region of pairs (k. k') E R2H compatible For notational with nonnegative consumption [i.e .• A = Ilk. k') E R2H: G(k. k') ~ convenience. we write u(G(k. k'» as u(k. k'). We assume that A is convex and that 11(', .) is strictly concave. Also. at t = 0 there is some already installed capital invcstment ko and this is the only initial endowment of capital in the economy. In this economy the planning problem (20.0.7) becomes'2
On
(20.0.8) s.t. (k,_" k,) E A for every t. and ko =
ko ·
From now on we assume that (20.0.8) has a (bounded) solution. Because of the strict concavity of u(· • . ) this solution is unique. For every / ~ I the vector of variables k, E RH enters the objective function of (20.0.8) only through the two-term sum o'u(k,_,. k,) + 0'+ 'u(k,. k, + d. Therefore, diffcrentiating with respect to these N variables. we obtain the following necessary conditions for an interior path (k o•...• k, •. .. ) to be a solution of the problem (20.0.8): 13 for every n :> Nand /
~
I.
In vector notation. for every t
~
I.
(20.0.9)
Conditions (20.0.9) are known as the Euler equations of the problem (20.0.8). Example 20.0.2: Consider the Ramsey-Solow technology of Example 20.C.1 (with I, = I for all t). Then. u(k. k') = u(F(k) - k') and A = Ilk. k'): k' $; F(k)}. Therefore, the Euler equations take the form -II'(F(k,_d- k,)
+ ou'(F(k,) -
k,+,)F'(k,) = O.
for all/2'o I
or
~=F'(k,)
for all t
~
I.
ou (c,+ ,)
In words: the marginal utilities of consuming at t or of investing and postponing consumption one period are the same. _ Example 20.0.3: Consider the cost-of-adjustment technology of Example 20.C.2 (except that as in Example 20.0.2 we fix " = I for all t and drop labor as an explicitly considered commodity) and suppose we have an overall firm that tries to maximize the infinite discounted sum of profits by means of a suitable investment policy in capacity. Output can be sold at a constant unitary price that. with a constant rate 12. By convention we put u(k_ 1• ko) == O. 13. The expression "interior path" means that (kit k,. I) is in the interior of A for all t. For the interpretation of the expression to come, recall also that k. and k~ stand, respectively. for the nth and Ihe (N + 1I)lh argumenl or u(k, k').
••
of interest. gives a present value price of 0'. Thus the problem becomes that of maximizing L, o'[F(k'_I) - k, - y(k, - k,_,)]. The Euler equations are then -I-y'(k,- k,_,)
+ o[F'(k,) +y'(k,+,
-
k,n =0
for all t ~ 1.
In words: the marginal cost of a unit of investment in capacity at t equals the discounted value of the marginal product of capacity at t plus the marginal saving in the cost of capacity expansion at t + I. Note that. iterating from t = I. we get I
+ y'(k,
- k o) =
L li'(F'(k,) -
,,,I
I).
In words: At the optimum. the cost of investing in an extra unit of capacity at t = I equals the discounted sum of the marginal products of a maintained increase of a unit of capacity. '4 See Exercise 20.0.5 for more detail.'· _ Suppose that a path (k o•..•• k" ... ) satisfies the Euler necessary equations (20.0.9). From their own definition, and the concavity of u(· •. ). it follows that the Euler equations are also sufficient to guarantee that the trajectory cannot be improved upon by a trajectory involving changes in a single k,. In fact, the same is true if the changes are limited to any finite number of periods (see Exercise 20.0.6). Thus. we can say that the Euler equations are necessary and sufficient for short-run optimization. The question is then: Do the Euler equations (or. equivalently. short-run optimization) imply long· run optimization? We shall see that. under a regularity property on the path (related. in a manner we shall not make explicit. to the transversality condition 16). they do. We say that the path (k o•. ..• k, •.. .) is strictly interior if it stays strictly away from the boundary of the admissible region A. [More precisely. the path is strictly interior if there is £ > 0 such that for every t there is an £ neighborhood of (k,. k,+ I) entirely contained in A.] Proposition 20.0.7: Suppose that the path (*0' ... ' k" ... ) is bounded, is strictly interior, and satisfies the Euler equations (20.0.9). Then it solves the optimization problem (20.0.8). Proof: The basic argument is familiar. Ir 0'0'" .. k, •. .. ) does not solve (20.0.8). then there is a feasible trajectory (/<0' ...• k;, . ..) that gives a higher ulility. To simplify the reasoning suppose Ihat this trajectory is bounded. Then. by the concavity of the objective function, the boundedness of (i'o." .• k, •. ..) and its strict interiority. we can assume that. Cor every r. k; is so close to k, that (k;. k,+ I) E A. We can now take T large enough for L'
~=~.- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - -
752
C HAP TEll
20:
E QUI L I • II I U M
A HD
TIM E
---------------------------------------------------------------------------------------
S E C'T ION
k; = k; for I:!> T and k; = k, for I > T. The new trajectory is admissible [note that (k T•kT+ I) E A); it coincides with 0,••. ... k; .... ) up to T and with (I, ...... k, .... ) after T. Moreover. if T is large enough. it still gives higher utility than 0, •..... k" .. .). But this is impossible because. as we have already indicated. the Euler equations imply short·run optimization. that is. they are the first-order conditions for the optimization problem where we are constrained to adjust only the variables corresponding to a finite number of periods (see Exercise 20.D.6). _
E QUI L III III U M:
THE
0 N E • C 0 H SUM E A
equations with initial condition (ko- k,) and then for fixed ko search for an initial condition k, generating a bounded infinite path. Example 20.D.4: Consider a Ramsey-Solow model with linear technology F(k) = 2k and utility function L, (1/2)'ln c,. Then u(k,_,. k,) = In (2k'_1 - k,) and the period-t Euler equation is (see Exercise 20.D.7) k'+1 = 3k, - 2k'_I'
This difTerence equation has the solution k, = ko + (k, - ko)(2' - I). If k, < ko. then k, eventually becomes negative. If k, > ko• then k, is unbounded. The only value of k, generating a bounded k, is k, = ko' Therefore. t/I(k o) = ko for any ko' It is instructive to see what happens if we try k, ~ ko. Then, the path induced by the ditTerence equation is feasible and. in fact, we have a constant level of consumption C, = 2k,_, - k, = 2ko - k,. Thus, for k, > k o• we have here an example of a path that is compatible with the Euler equations but that is not optimal. because at k, = ko we get a higher level of constant consumption." _
It may be helpful at this stage to introduce the concept of the value function V(k) and the policy fUllctioll t/I(k). Given an initial condition ko = k. the maximum value attained by (20.D.8) is denoted V(k). and if (k o• k l ••••• k, •... ) is the (unique) trajectory solving (20.D.8) with ko = k. then we put t/I(k) = k,. That is. t/I(k) e IRN is the vector of optimal levels of investment. hence of capital. at t = I when the levels of capital at t = 0 are given by k. What accounts for the importance of the policy function is the observation that if the path (k o•...• k" ...) solves (20.D.8) for ko = ko then. for any T, the path (k T • ••• ,k T +, •. ..) solves (20.D.8) for ko = k,.. Thus, if (k o, ...• k" . .. ) solves (20.D.8) we must have
k,+, = t/I(k,) for every t.
2 0 • D:
CA. E
753
---------------------------------------------------------
The dynamic programming approach exploits the recursivity of the optimum problem (20.D.8). namely. the fact that
V(k) =
(20.D.IO)
Max
u(k. k')
+ 0 V(k'),
(20.D.11)
,'""Uh(i,l')EA
and we see that the optimal path can be computed from knowledge of ko and the policy function t/I(.). But how do we determine t/I(')1 We now describe two approaches to the computation of t/I(.). The first exploits the Euler equations; the second rests on the method of dynamic programming. The Euler equations (20.D.9) suggest an iterative procedure for the computation of t/I(k). Fix ko = k and consider the equations corresponding to k,. With ko given, we have N equations in the 2N unknowns k, e IRN and k2 eRN. There are therefore N degrees of freedom. Suppose that we try to fix k, arbitrarily [equivalently. we try to fix - V2u(ko. k,), the marginal costs of investment at t = I] and then use the N Euler equations at t = 1 to solve for the remaining k2 unknowns [equivalently, we adjust the commitments for investment at t = 2 so that the discounted marginal payofT of investment at t = I. bV,u(k" k 2 ). equals the preestablished marginal cost of investment at t = I, i.e. - V2u(ko, k,)]. Suppose that such a solution k2 is found [by the strict concavity of u(·), if there is one solution then it has to be unique]. We can then repeat the process. The N Euler equations for period 2 are now exactly determined: Both kl and k2 are given. but we still have the N variables k) corresponding to t = 3 with which we can try to satisfy the N equations of period 2. Suppose that we reiterate in this fashion. There are three possibilities. The first is that the process breaks down somewhere, that is, that given k, _, and k, there is no solution k,+, [or, more precisely, no solution with (k,. k,+ ,) e A]; the second is that we generate a sequence that is unbounded (or nonstrictly interior); the third is that we generate a bounded (and strictly interior) sequence (k o, k" ... ,k" ... ). In the third case, by Proposition 20.D.7 we have obtained an optimum, and since by Proposition 20.D.6 the optimum is unique, we can conclude that given ko, tile third possibility (the trajectory startillg at ko and kl is strictly interior alld bounded) can occur for at most one value of k l . If it occurs, tltis value of kl is precisely t/I(k o). Thus, the computational method is: Solve the ditTerence equation induced by the Euler
and obtains t/I(k) as the vector k' that solves (20.D.II). This, of course, only transforms the problem into one of computing the value function V(·). However, it turns out that. first, under some general conditions [e.g.• if V(·) is bounded] the value function is the only function that solves (20.D.11) when viewed as a functional equation, that is, V(·) is the only function for which (20.D.lI) is true for every k. and, second, that there are some well-known and quite effective algorithms for solving equations such as (20.D.II) for the unknown function V(·). (Sec Section M.M. of the Mathematical Appendix.) We end this section by pointing out two implications of the definition of the value function (sec Exercise 20.D.8): (i) The value fUllction V(k) is concave. (ii) For every perturbacion parameter z e IRN with (k
V(k + z)
~
+ z, t/I(k» e A we have
u(k + z, t/I(k» + oV(t/I(k)).
(20.D.12)
Suppose that N = I and (k, t/I(k)) is interior to A. For later reference we point out that from (i), (ii), and V(k) = u(k. t/I(k)) + 0 V(t/I(k» we obtain
V'(k) and, if V(·) is twice-differentiable,
= V,u(k. t/I(k»
VH(k) ~ V:,u(k, t/I(k)). (See Figure 20.0.1 and Exercise 20.D.9. 18 ) 17. Hence. when k, > k., the Euler equations lead to capital overaccumulation. We note. without further elaboration, that given a path satisfying the Euler equal ions we could use the equations themselves to determine a myopically supporting price sequence. However, jf k I > ko Ihis sequence will violate the transvcrsality condition. 18. The expression V;~f(') denotes the ij second partial derivative of the real-value function f(·).
1
754
C HAP T E R
2 0:
E QUI LIB R I U MAN 0
5 EC T,0 N
TIM E
----------------------------------------------------------------------~
V(k +:)
Figure 20.0.1
Along an optimal P"h the value function is majorized by the
utilities of singleperiod adjustments.
20.E Stationary Paths, Interest Rates, and Golden Rules In this section, we concentrate on the study of steady states. This study constitutes a first step towards the analysis of the dynamics of equilibrium paths. We refer to Bliss (1975), Gale (1973), or Weizsiicker (1971) for further analysis of steady-state theory. We begin with a production set Y c I1\llL satisfying the properties considered in Section 20.C. Recall that a production path is a sequence (yo, ... , y" ... ) with y, E Y for every r. DefinItion 20.E.1: A production path (yo, ... , y" ... ) is stationary, or a steady state, if there is a production plan y = (Yb' Y.) E Y such that y, = I' for all t> O. Abusing terminology slightly, we refer to the "stationary path (y, ... , y, .. .)" as simply the "stationary path y." The first important observation is that stationary paths rhat are also efficient are supportable by proportional prices. '9 This is shown in Proposition 20.E.!. ProposItion 20.E.1: Suppose that Ii E Y defines a stationary and efficient path. Then, there is a price vector Po E I1\lL and an ex > 0 such that the path is myopically profit maximizing for the price sequence (Po' expo' ... , ex'PD' ... ). Proof: A complete proof is too delicate an affair, but the basic intuition may be grasped from the case in which production sets have smooth boundaries. For this case we can, in fact, show that every (myopically) supporting price sequence must be proportional. By the efficiency of the path (y, . .. , )" ... ), the vector y must lie at the boundary of Y. Let q = (qo, q,) be the unique (up to normalization) vector perpendicular to Y at y. Also, by the small type discussion at the end of Section 20.C, there exists a price sequence (Po, ... , P..... ) that myopically supports this efficient path. Because }' E Y is short-run profit maximizing at every r we must have (P .. P.. ,) = ).,(qo, q ,) for some ;., > O. Therefore, p, = ).,qo and p,+, = ;.,q, for all r. In particular, p, = ;., _, q, and p, = ;., qo. Combining, we obtain p, = (}.,/;., _ ,) p, and
+,
+,
+,
t 9. To prevent possible misunderstanding, we warn that establishing the inefficiency of a given stationary path will typically require the consideration of nonstationary paths.
I·
'I
S TAT' 0 N A A Y
PAT H S,
'N TEA EST
A ATE S ,
AND
+,
= ()., + ,/;., ) p,. From this we get ).,/)., _, = )., + ,/)., for all t 2: 1. Hence, denoting this quotient by ex, we have p,+, = exp, = ex 2 p,_, = ... = ex'+'po' The factor ex has a simple interpretation. Indeed, r = (I - ex)/ex [so that p, = (I + r)p,+ ,] can be viewed as a rate oj interest implicit in the price sequence (see Exercise 20.E.I). Proposition 20.E.I is a sort of second welfare theorem result for stationary paths. We could also pose the parallel first welfare theorem question. Namely, suppose that (Y, ... , }', ... ) is a stationary path myopically supported by a proportional price sequence with rate of interest r. If r> 0, then P, = (1/(1 + r»'po -+ 0 and therefore the transversality condition p, y. -+ 0 is satisfied. We conclude from Proposition 20.C.1 that the path is efficient. If r ~ 0, the transversality condition is not satisfied (p, docs not go to zero), but this does not automatically imply inefficiency because the transversality condition is sufficient but not necessary for efficiency. Suppose that r < 0 and, to make things simple, let us be in the smooth case again. Consider the stationary candidate paths defined by the constant production plan y, = (Y. + r.e,),. - u), where e = (1, ...• 1) E RL. This candidate path uses fewer inputs (or produces more outputs) at t = 0 and generates exactly the same net input-output vector at every other t. Therefore, if for some £ > 0, the candidate path is in fact a feasible path; that is. if y, E Y, then the stationary path y is not efficient (it overaccumulates). But if Y has a smooth boundary at y, the feasibility of y, for some r. > 0 can be tested by checking whether y, - y = £(e, - e) lies below the hyperplane determined by the supporting prices (Po, [1/(1 + r)]po). Evaluating. we have £(1 - 1/(1 + r»po·e < 0, because r < O. Conclusion: For'£ small enough, the stationary path)' is dominated by the stationary path y,. We record these facts for later reference in Proposition 20.E.2. p,
__~=---u(k +:, ",(k)) + 6V(I{t(k))
2 0 . E:
Q 0 L DEN
AULES
755
-----------------------------------------------------------------------
Proposition 20.E.2: Suppose that the stationary path (y, ... , y . .. ), YE Y, is myopically supported by proportional prices with rate of interest r, then the path is efficient if r > 0 and inefficient if r < O. We have not yet dealt with the case r = 0, which as we shall see, is very important. 20 We will later verify in a more specific setup that efficiency obtains in this case. Let us now bring in the consumption side of the economy and consider stationary equilibrium pat Its. Assuming differentiability and interiority, a stationary path (y, ... , y, ... ) that is also an equilibrium can be supported only (up to a normalization) by the price sequence p, = IJ' Vu(c). where c = Yb + y.; recall Proposition 20.D.4 and expression (20.D.6). That is. a slationar.v equilibrium is supported by a price sequence clIlbodyillf} (I proportiollalil,l' Jacror equal 10 rite discount Jactor .5, or, equivalently, with rate of interest r = (I - .5)/.5. Definition 20.E.2: A stationary production path that is myopically supported by proportional prices p, = rx'Po with (X = .5 is called a modified golden rule path. A stationary production path myopically supported by constant prices p, = Po is called a golden rule parh. 20. No'e that 0 is 'he ra'e of growth implicit in the pa,h (Y, ... , y, .. .). In a more general treatment we could allow for a constant returns technology and for the production path to be proportional (but not necessarily stationary). Then Proposition 20.E.2 remains valid with 0 replaced by the corresponding rate of growth.
756
c
H .. PTE R
20,
E
Q
U , L' B R , U II
.. NOT' II E
SEC T ION
2 0 • E:
S TAT ION A R Y
P It. T H S.
I N T ERE S T
RAT E S.
AND
G 0 L DEN
RULES
757
----------------------------------------------------------------------------------- ---------------------------------------------------------------------Depending on the technology and on the discount factor b, there may be a single or there may be several modified golden rule paths (see the small-type discussion at the end of this section). But in any case we have just seen that a statiollary equilibrium path is lIecessarily a modified goldell rule path. Thus, we have the important implication that the calldidates for statiollary eqllilibrium paths (y, . .. ,y, ... ) are
Qu'put
completely determined by the technology and the discount faclOr and are independent <J{ the lIIility fllnctioll u(·).
To pursue the analysis it will be much more convenient to reduce the level of ahstraction. Consider an extremely simple case, the Ramsey-Solow model technology of Example 20.C.1. We study trajectories with I, = I for all t (imagine that there is available one unit of labor at every point in time). We can then identify a production path with the sequence of capital investments (k o, ... , k" ... ). Given (k o, ... , k" .. .), denote r, = V,F(k" 1) - I. Thus, r, is the lIel (i.e., after rcplacing capital) marginal productivity of capital. Suppose that k, > 0 and that the sequence of output prices (qo, ... ,q" .. .) and wages (w o, ... , H'" ... ) myopically price supports the given path. Then, by the first-order condition for profit maximization, we have q,+,(1 + r,) - q, = O. Hence r, is the output rate of interest at time t implicit in the output price sequence (qo, ... ,q" ... ). Let us now focus on the stationary paths of this example. Any k ;,: 0 fixed through time constitutes a sleady stale. With any such steady state we can associate a constant surplus level c(k) = F(k, 1) - k and a rate of interest r(k) = V,F(k, I) - I, also constant through time. 1 ' Therefore, the supporting price-wage sequence is with H'o
V1F(k,l)
q. = V,F(k, 1>"
Denote by w(k) the real wage wo/qo so determined. It is instructive to analyze how the steady-state levels of consumption e(k), the rate of interest r(k), and the real wage w(k) depend on k. Let k be the level of capital at which the steady-state consumption level is maximized [i.e., k solves Max F(k, 1) - k). Note that k is characterized by r(k) = V, F(k, 1) - 1 = O. Thus k is precisely the goldell rule steady state. The construction is illustrated in Figure 20.E.I, where we also represent the modified golden rule k. [characterized by r(kd ) = V,F(k., 1) - 1 = (I - 0)/0). Observe that if k < k then r(k) > O. As we saw in Proposition 20.E.2, r(k) > 0 implies that the steady state k is etl1cicnt (thus, in particular, the modified golden rule is etl1cient: it gives less consumption than the golden rule but it also uses less capital). Similarly, if k > k then r(k) < 0 and we have inefficiency of the steady state k. What about k?" We now argue that the golden rule steady state k is efficient. A graphic proof will be quickest. Suppose we try to dominate the constant path k by starting with ko < k, so that consumption at I = 0 is raised. Since the surplus at t = 1 must be at least
21. Thus. c(k) is the amount of good constantly available through time and usable as a flow
for consumption purposes. 22. Recall that the associated price sequence is constant and that the transversality condition is therefore violated.
Figure 2O.E.l
I"pu,
-k I
k.
t
The production technology of the Ramsey-Solow model and the golden rule.
I
Golden Mod,fied _ Rule Golden Rule J k,
='0 -
(<(kl- [F(ko.ll- kol) = F(k o, 1)- «kl
1-r---_ -: ~
k: Golden Rule k.: Modified Golden Rule
Surplus
-
-------~--
Figure 20.E.2
Ramsey-Solow model: the golden rule is efficient.
-k
Inpu' F(k, 1)- k
-k,
e(k), the best we can do for k, is
k, = F(k o, I) - e(k) = F(k o, I) - ko
+ ko
- e(k)
< ko.
because F(k o, I) - ko < elk). This new best possible value of k, is represented in Figure 20.E.2. In the figure we also see that as the process of determination of k, is iterated to obtain kl' k3 and so on we will, at some point get a k, < O. Hence. the path is not feasible, and we conclude that a constant k cannot be dominatcd from the point of view of efficiency: the attempt to use less capital at some stage will inexorably lead to capital depletion in finite time. From the form of the production function, three "neoclassical" properties follow immediately (you are asked to prove them in Exercise 20.E.4): (i) As k increases, the level elk) increases monotonically up to the golden rule level and then decreases monotonically. (ii) The rate of interest r(k) decreases monotonically with the level of capital k. (iii) The real wage w(k) increases monotonically with the level of capital. (For the validity of this property you should also assume that production function F(k, I) is homogeneous of degree one.) From the study of the steady states of the Ramsey-Solow model we have learnt at least six new things: First, the rate of interest is equal to the net marginal productivity of capital; second, the golden rule (i.e., zero rate of interest) path is characterized by a surplus-maximizing property among steady states; third, the golden rule is efficient; fourth, fifth, and sixth, we have the three neoclassical properties.
758
C HAP T E R
2 0:
E QUI LIB R I U MAN D
TIM E
SEC T ION
• 0 • F,
DYNA MIC S
759
-------------------------------------------------------------------------- --------------------------------------------------------------------------supported by proportional prices q, = (1 + r(k))q.. , where, to insure profit maximization, q,/q,., must be equal to the slope of the level curve through (k, k) (you should verify this in Exercise 20.E.6). Therefore, the efficient steady states, those with r(k) :2: 0, correspond to the subset of the diagonal that goes from the origin to the golden rule, where r(k) = O. In the special case of the Ramsey-Solow model we have G(k, k') = F(k, I) - k' and therefore the level curves of G(k, k') admit a quasilinear representation with respect to k' (i.e., they can be generated from each other by parallel displacement along the k' axis). In Exercise 20.E.7 you are asked to show that this guarantees the satisfaction of the neoclassical properties. In general, however, it is clear from Figure 20.E.3 that we may, for example, have two different k. k < k such that, at the diagonal, the corresponding level curves have the same slope and therefore ,(k) = ,(k) (contradicting the second neoclassical property). In particular, while the golden rule is unique [if the function G(k, k') is strictly concave], there may be several modified golden rules [this is the case if, say, the discount factor J is equal to the interest rate r(k)].
How general is all of this? That is, can we make similar claims for the general model with any number of goods? The answer, in short, is that the three neoclassical properties mayor may not hold in a world with several capital goods, but the other three, duly interpreted, remain valid with great generality. Attempting to give proofs of all this would take us into too advanced material [see Bliss (1975) or Brock and Burmeister (1976)], but perhaps we can provide some intuition. Suppose we consider the general (N + I )-sector technology of Example 20.C.4. That is, G(k, k') is the amount of consumption good available at any period if kERN is the vector of levels of alpital used in the previous period and the investment required in the period is k' E R' (we also let I, = I for all t). At a steady-state path we have k' = k. Denote by C(k) = G(k, k) the level of consumption associated with the steady state k. If G(', .) is a concave function then so is C(·). In particular, VC(k) = 0 characterizes the steady state with maximal level or consumption. Consider a steady steady k. By Proposition 20.E.I, this steady state can be myopically supported by a proportional price sequence s, E R, q, E RH. Here s, is the price of the consumption good in period t, and q, is the vector of prices of investment in period t. Because of proportionality there is an r(k) such that s, = (I + r(k))s .. I' q, = (I + r(k»q,., for all t. Because of profit maximization, V, G(k, k) =1- q, _, s,
and
V,G(k, k)
=-
I s,
q,
for all t
20.F Dynamics
(20.E.I)
(you are asked to verify this in Exercise 20.E.5). Therefore, • VG(k) = V,G(k, k)
+ V,G(k, k)
I r(k) = - (q,_, - q,) = _. q" S,
5,
that is. at any time an extra dollar invested in a permanent increase of any clIpital good yields r(k) clollars ill extra value of (permallent) consumption. In this precise sense the rate of interest measures the marginal productivity of capita\. We see again that VC(k) = 0 (the necessary and sufficient condition for maximum steady-state consumption) is equivalent to r(k) = O. Hence, the golden rule property holds: a steady-state level k yields maximal consumption across steady states if and only if it has associated with it a zero rate of interest. We add that we could also prove that the golden rule path is efficient. As we have already indicated, the neoclassical properties do not carryover to the general selting. A taste of the possible difficulties can be given even if N = I, that is, for the two-sector model of Example 20.C.3. In Figure 20.E.3 we represent the level curves of G(k, k'). The steady states correspond to the diagonal, where k = k'. Every steady state k can be myopically
45'
k'
//// G(k. k')
= constant
Ftgure 20.E.3 An example with several modified golden rules.
In this section, we olTer a few observations on the vast topic of the dynamic properties of equilibria. The basic framework is as in the previous section: a one-consumer economy with stationary technology and utility. The arbitrarily given initial conditions" will typically not be compatible with a stationary equilibrium situation (e.g., the steady-state level of capital may be higher than the initial availability of capital). Therefore, the typical equilibrium path will be nonstationary. How compliCllted can the equilibrium dynamics be? Can we, for example, expect convergence to a modified golden rule? This would be nice, as it would tell us that our models carry definite long-run predictions. We can gain much insight into these matters by considering a variation of the two-sector model of Example 20.C.3. We assume that the technology produces consumption goods (possibly of more than one kind) out of labor and a capital good. There is, as initial endowment, one unit of labor in each period, and we let u(k, k') stand for the maximum utility that can be attained in any given period if in the previous period k E R units of capital were installed and the current investment is required to be k' (and, in both periods, a unit of labor is used). There is a positive initial endowment of capital only at t = O. Also, we take u(', .) to be strictly concave and differentiable. We know from Proposition 20.0.3 and 20.0.4 that the equilibrium paths can be determined by means of the following planning problem: Max
IJ'u(k,_" k,)
(20.F.t)
s.t. k, ;:: 0 and kG = k is given. Suppose that V(k) and !/I(k) are value and policy functions, respectively, for the problem (20.F.I). These concepts were introduced in Section 20.D. As we explained there, the equilibrium dynamics are entirely determined by iterating the policy function [sec expression (20.0.10)]. That is, given kG' the equilibrium trajectory is (k", k" k" ... ) = (kG' !/I(k o ), ",(",(k o)), .. .). 23. That is, the initial endowment sequence (w o•.
• W f '"
.).
760
c HAP T E R
2 0,
E QUI L I 8 R I U MAN 0
TIM E
SEC T ION
2 Q • F:
0 Y N A .. I C S
761
------------------------------------------------------------------------------ -----------------------------------------------------------------------------k'
k'
Flgur. 20.F.' (leH)
A single, stable steady state. Flgur. 20.F.2 (right)
Several steady states, no cycles.
ko k'
k'
Flgur. 2O.F.3 (leH)
A single steady state and a cycle of period 2. Flgur. 20.F.4 (right)
A cycle of period 3: chaos.
Note that a steady-state path (k, ... , k, ... ) is an equilibrium path (for ko = k), and therefore a modified golden rule steady state path for discount factor (j (see Definition 20.E.2 and the discussion surrounding it), if and only if k = I/I(k). Figures 20.F.I through 20.F.4 represent four mathematical possibilities for this equilibrium dynamics. In Figure 20.F.I, we have the simplest possible situation: a monotonically increasing policy function with a single steady state k. The steady state is then necessarily globally stable; that is, k, -+ k for any ko. In Figure 20.F.2, the policy function is again monotonically increasing, but now there are several steady states. They have different stability properties, but it is still true that from any initial point we converge to some steady state. In Figure 20.F.3, the steady state is unique, but now the policy function is not increasing and cycles are possible. Finally, in Figure 20.F.4 we have a policy function that generates a cycle of period 3. It is known that a one-dimensional dynamical system exhibiting a nontrivial cycle of period 3 is necessarily chaotic [see Grandmont (1986) for an exposition of the mathematical theory). We cannot go here into an explanation of the term "chaotic" in this context. It suffices to say that the equilibrium trajectory may wander in a complicated way and that its location in the distant future is very sensitive to initial conditions. The theoretical possibility of chaotic equilibrium trajectories is troublesome from the economic point of view. How is it to be expected that an auctioneer will succeed in computing them; or even worse, how would a consumer exactly anticipate such a sequence?
Unfortunately, the "anything goes" principle that haunted us in Chapter 17 in the form of the Sonnenschein-Mantel-Debreu theorem (Section 17.E) reemerges here in the guise of the Boldrin-Montruocio theorem [see Boldrin and Montruocio (1986)]: Any candidate policy function I/I(k) can be generated from some concave u(k, k') and" > O. We will not state or demonstrate this theorem precisely, but the main idea of its proof is quite accessible. We devote the next few paragraphs to it. Suppose for a moment that for a given u(', .) our candidate 1/1(') is such that oJ!(k) solves, for every k, the following "complete impatience" problem: Max
u(k, k').
(20.F.2)
t':
This would be the problem of a decision maker who did not care about the future. While this is not quite the problem that we want to solve, it approximates it if we take" > 0 to be very low. Then the decision maker cares very little about the future and therefore its optimal action k' will, by continuity, be very close to I/I(k). Hence, in an approximate sense, we are done if we can find a u(', .) such that oJ!(k) solves (20.F.2) for every k. In order for a oJ!(k) > 0 to solve (20.F.2), u(k, .) cannot be everywhere decreasing in its second argument (the optimal decision would then be k' = 0). [n the simplest version of the Ramsey-Solow model (Example 20.C.I), the returns of k', the investment in the current period, accrue only in the next period, and therefore the utility function u(k, k') is decreasing in k'. But in the current, more general, two-sector model there is no reason that forces this conclusion. Suppose, for example, that there are two consumption goods. The first is the usual consumption-i~vestment good, while the second is a pure consumption good not perfectly substitutable with the first. Say that with an amount k of investment at time t - lone gets, jointly, k units of the consumption-investment good at time t and k units of the second consumption good at time t - I. Accordingly, with k' units of the consumption-investment good invested at lone gets, jointly, k' units of the consumption-investment good at t + 1 and k' units of the second consumption good at t. Thus, if k and k' are the amounts of investment at 1 - I and I, respectively, then the bundle of consumption goods available at 1 is (k - k', k'). Hence, the utility function u(',') has the form u(k, k') = u(k - k', k'), where a(·, .) is a utility function for bundles of the two consumption goods. Therefore, our problem is reduced to the following: Given oJ!(k) can we find a(·, .) such that oJ!(k) solves Max,. u(k - k', k') for all k in some range? The problem is represented in Figure 20T5. 24 We see from the figure that the problem has formally become one of finding a concave utility function with a prespecified Engel curve at some given prices (in our case, the two prices are equal). Such a utility function can always be obtained. It is a well-known, and most plausible fact that the concavity of a(·) imposes no restrictions on the shape that a single Engel curve may exhibit (see Exercise 20.F.I). The news is not uniformly bad, however. In principle, as we have seen, everything may be possible; yet there are interesting and useful sufficient conditions implying a
24. We also assume that of;(k) < k for all k.
762
C HAP T E R
, 0:
E QUI LIB R I U MAN D
8 ECTI Q N
TIM E
, 0 • F:
DYNA M I CS
763
---------------------------------------------------------------------------------=-~~~~--------------------------------Second Hence, if the discount factor t5 is close to 1, it is a plausible conclusion that 1!/I'(k)1 < 1 Consumption Good
for all k. In technical language: !/I(') is a contraction, and this implies global convergcnce to a unique steady state.'· In Exercise 20.F.2 you are invited to draw the policy functions and the arrOw diagrams for this case. A particular instance of a contraction is exhibited in Figure 20.F.1. ~_ _
(k - y,(k). y,(k))
Figure 20.F.S
Construction of an
arbitrary policy function in the completely impatient casco
45 First Consumption Good
well-behaved dynamic behavior. We discuss two types of conditions: a low discollllr of time and ('ross derivatives of uniform posit;l'£, sign.
Low Discoullr of Time One of the most general results of dynamic economics is the turnpike rheorem, which, informally, asserts that if rhe one-period utility function is strictly concave and tile decision maker is very patient, then rhere is a single modified golden rule sready scare
rhar, moreover, attracts the optimal trajectories from any initial position. I n the context of the two-sector model studied in this section, we can give some intuition for the turnpike theorem. Suppose that the value function V(k), which is concave, is twice-differentiable." At the end of Section 20.D, we saw that since by definition,
V(k
Cross Derivative of Uniform PositilJe Sign
+ z) 2: u(k + z, !/I(k» + W(t/I(k»
for all z and k (with equality for z = 0), we must have
V'(k) = V,u(k, t/I(k»
and
V"(k) 2: V?,u(k, !/I(k»
for all k.
Also for all k, t/I(k) solves the first-order condition
V2 u(k, t/I(k»
+ t5 V'(!/I(k)) = O.
(20.F.3)
Differentiating this first-order condition, we have (all the derivatives are evaluated at k, !/I(k) and assumed to be nonzero)
V 2 u(') t/I'(')
Because
= - Vi,"(';'+ t5V"(')
Vl 11l(') :;:; 0 and SV;'Il(') :;:; b V"(-) :;:; 0, it follows that 1\11'(')1 :;:; \Vi,u(
The lurnpike Iheorem is valid for any number of goods. The precise slatement and the proof of the Iheorem are subtle and technical [see McKenzie (1987) for a brief survey], but the main logic is simply conveyed. Consider the extreme case where there is complete patience, that is, "only the long-run matters." A difficulty is that it is not clear what this means for arbitrary paths; but at least for paths that are not too "wild," say for those that from some time become cyclical, it is natural to assume that it means that the paths are evaluated by taking the average utility over the cycle. Observe now that for any cyclical nonconstant path. the strict concavity of the utility function implies that the constant path equal to the mean level of capital over the cycle yields a higher utility. It may take some time 10 carry out a transition from the cycle to the constanl path (e.g., it may be necessary to build up capital) bul, as long as this can be done in a finite number of periods, the cost of the transition will not show up in the long run. Hence the cyclical nonconslant path cannot be optimal for a completely patienl optimizer. By continuity, all this remains valid if 0 is very close 10 I. We can conclude. therefore, that if a path tends to a nonconstant cycle then we can always implement a finite Iransition to a suitable constant "long-run average," for a relatively large long-run gain of utility and a relatively low short-run cost. In fact, this conclusion remains valid whenever a path does not stabilize in the long-run. It follows that the optimal path must be asymptotically almost constant, which can only be the case if the path reaches and remains in a neighborhood of a modified golden rule steady state (recall from Section 20.E that those are the only constant paths that can be equilibria, and therefore optimal)."
~i~u~~~ u(. )\. I
We shall concern ourselves here with the particular case of the two-sector model studied so far where V,u(k, k') > 0 and V,u(k, k') < 0 for all (k, k'). By a cross derivative of uniform positive sign we mean that V12 u(k, k') > 0, again at all points of tlte domain. In words: An increase in investment requirements at one date leads 10 a situation of increased productivity (in terms of current utility) of the capital installed the previous date. Examples are the classical Ramsey-Solow model u(F(k) - k') and the cost-of-adjustment model u(F(k) - k' -y(k' - k)) (see Exercise 20.F.3). We shall argue that under tltis cross derivative condition tIle policy function is increasing (as in Figures 20.F.1 or 20.F.2), and therefore the optimal path converges to a stationary path. To prove the claim, it is useful to express \II(k) as the k' solution to Max
u(k, k')
+ bV
(20.FA)
(II',VI
s.t. V:;:; V(k'),
By the concavity of 11(') we have (see Sections M.e and M.D of the Mathematical Appendix)
26. We nole thai y,(.) need not be monotone and Ihe convergence may be cyclical. although the cycles will dampen through lime.
25. For a (very advanced) discussion of this assumption. see Santos (1991).
i
~L
27. Also. with ~ close 10 1, the modified golden rule willlypically retain the uniqueness property of the golden rule.
764
C HAP T E R
2 0:
E
a
U I liB R I U MAN 0
TIM E
765 -------------------------------------------------------------------- ----------------------------------------------------------------------------Indifference Curve for Ihe Ulilily + W, k fixed
u(k, k')
v
SEC T ION
-",-",//
k:I-_ _ _. ....L... /~=------ !/I,(')
>k
k;'
r-A
r----~~/-"..."...------!/I(. ) / / / ~ Transilory Shock //
_.",&----V(k')
,,/
"" ",,""
!/I(k)
S EVE R A l e 0 N SUM E R S
Permanenl Shock / /
J /
I'(o/I(k))
E QUI l I • R I U M:
k'
Indifference Curve for Ihe Ulilily , u(i. k') + J V, k fixed
i:
J. . . . . . . .
2 0 • G:
/
/"
,, I, ,,I
~5'
I
f
Figure 2O.F.6
k'
where V(·) is the value function. For fixed k, problem (20.F.4) is represented in Figure 20.F.6. The marginal rate of substitution (M RS) between current investment k' and future utility V at s = (I/I(k), V(I/I(k))) is O/b)V2u(k. I/I(k» < O. Suppose now that we take k > k. Then the indifference map in Figure 20.F.6 changes. Because V'211(k. I/I(k)) > O. the MRS at s is altered in the manner displayed in the figure. that is. the indifference curve becomes flatter. But we can see then that necessarily I/I(k) > I/I(k), as we wanted to show. The cross derivative condition does not. by itself. imply the existence of a single modified golden rule. Thus, we could be in Figure 20.F.2 rather than in Figure 20.F.!. Note, however. that in many cases of interest it may be possible to show directly that the modified golden rule is unique. Thus. in both the classical Ramsey-Solow model of Example 20.C.1 and in the cost-of-adjustment model [with )1'(0) = 0] of Example 20.C.2. the modified golden rule is characterized by F'(k) = lib. Hence it is unique and. because the policy function is increasing. we conclude that every optimal path converges to it. We also point out that if the cross derivative is of uniform negative sign. then. by the same arguments. 1/1(') is decreasing. While this allows for cycles. the dynamics are still relatively simple. In particular. the non monotonic shape associated with the possibility of chaotic paths (Figure 20.F.4) cannot rise. See Deneckere and Pelikan (1986) for more on these points. Figure 20.F.6 is also helpful in illuminaling Ihe dislinclion bel ween Irallsilory and permanelll shocks. One of Ihe importanl uses of dynamic analysis in general. and of global convergence turnpike results in particular. is in the examination of how an economy at long-run rest reacts 10 a perlurbalion of the dala al lime I = I. In an exlremely crude c1assificalion, Ihese perturbalions can be of Iwo Iypes: (i) Transitory shocks affeci the environment of Ihe economy only al I = I; Ihal is, Ihey aller ko or. more generally. u(ko• . ). Ihe ulilily function at I = I. Then Figure 20.F.6 allows us to see how the equilibrium path will be displaced. The (k'. V) indifference curve of u(k o• k') + {, V changes, bul the constraint function V(k') remains unaltered. Hence. afler the (transilory) shock, Ihe new k', corresponds 10 the solulion of the optimum problem depicled in Figure
With the uniform positive sign cross derivative condition, the policy funclion is increasing.
20.F.6 bUI with Ihe new indifference map. From
t =
2 on we simply follow Ihe old policy
function.
(ii) Permanelll shocks move the economy to a new utility function u(k. k') constant over time. Then Ihe entire policy funclion changes to a new ';('). In terms of Figure 20.F.6 Ihere would be a change in both the indifference curves and the constraint. The new kf is now harder 10 determinc and to compare with the preshock k, or. for the same shock at period I. wilh k;'; bul il can oftcn be done. We pursue Ihe matter through Example 20.F.1. Example 20.F.I: Consider the separable utilily u(k. k') = y(k) + h(k'). This could be the inveslmenl problem of a firm: g(k) is Ihe maximal revenue obtainable wilh k. and -h(k') is Ihe cost of investment. Then Vl,u(k. k') = 0 at all (k. k'). Our previous analysis of Figure 20.F.6, lells us Ihal in this case o/t<") is constant; that is. from any ko the economy goes in one step to ils sleady-state value k. This is illustrated in Figure 20.F.7. Suppose now Ihere is a shock variable 0 such thai u(k, k'. 0) = g(k. 0) + h(k', 0). wilh the preshock value being 0 = O. The economy is initially at its steady Slate k. If there is a Iransitory shock to a small 0 > O. then from the analysis of Figure 20.F.6 we can see thai k', ~ Ie according to a'h(k. O)/ak' ao ~ O. (Exercise 20.F.4 asks you to verify this.) To evaluate the effecls of a permanent shock 10 a small 0> 0 (and therefore to a new !/I,('» Ihe lerm i)2 Viii, O)/i)k
Flgur. 20.F.7
An example of dynamic adjustment under transitional and permanent shocks.
ao = a'g(k, O)/ak iJO
also matters [Ihc previous equalilY follows from expression (20.F.3)]. Suppose. for example, Ihat Ihe shock is unambiguously favourable. in the sense thai a'g(k,OJ/ak ao > 0 and c'h(le. O)/ak' DO > O. Then a careful analysis of Figure 20.F.6, would allow us to conclude thai kf> k;' > k. (Exercise 20.F.5 asks you to verify this. Note that the indifference map of Figure 20.F.6 is quasilinear wilh respecI 10 V.) Figure 20.F.7 illustrales Ihis case further. _
~O,G Equilibrium: Several Consumers Up to now we have had a single consumer. or, to be more precise, a single type of consumer. The extension of the definition of equilibrium to economies with several consumers, say I. presents no particular difficulty. We simply have to rewrite Definition 20.D.1 as in Definition 20.G.!. Definition 20.G.1: The (bounded) production path (Yd, ... , Yi, . .. ), Yi E Y, the (bounded) price sequence (Po ... . ,Pt' ... ) ;:>: 0, and the consumption streams
766
c HAP T E R
2 0:
E QUI LI B R I U MAN 0
SECTION
TIM E
20.G:
EOUILIBRIUM:
SEVERAL
CONSUMERS
767
----------------------------------------------------------------------------- --------------------------------------------------------------------------(C
Oi '" ., C~'"
.)
<': 0, i = 1, ... ,I, constitute a Walrasian (or competitive) equi-
librium if:
LC~ =
(i)
Y:.,_' + Yb, + LW'i' for all t.
(20.G.1)
(u~, u!):
(ii) For every t,
u
(20.G.2)
for all Y = (Yb" y.,) E Y. (iii) For every i, the consumption stream (C Oi , .. problem Max
O'i is
·, c:;, ... ) <': 0 solves the
Lb:ui(ci ) , s.t.
where
(20.G.3)
L, P,'C'i:!> L, O'i 1t , + L, P,'w'i = Wi'
consumer i's given share of period t profits.
The first, and very important, observation to make is that, in complete analogy 28 with the finite-horizon case (see Section l6.C), the first welfare theorem holds. Proposition 20.G.1: A Walrasian equilibrium allocation is Pareto optimal. Proof: The proof is as in Proposition 16.C.1. Let the Walrasian equilibrium path under consideration be given by the production path (yo, ... , y~, ., .), the consumption streams (CO" ... , c~, . .. ), i = I, ... ,I, and the price sequence (Po, ... , p" ••• ). Suppose now that the paths (Yo> ... ,y" ... ), and (COl' ••• , C,I' •.• ) <': 0, i = I, ... , I, are feasible [i.e., they satisfy condition (i) of Definition 20.G.I] and are Pareto preferred to the Walrasian equilibrium. By the utility-maximization condition we have L, P,'C,I <': w, for all i, with at least one inequality strict. Hence,
L P,(L C,,) = L (L p"c,,) > L W,. f
j
f
I
(20.GA)
~ p,-(~ C,.) = ~ P,-(Y•. ,_, + Y., + ~ W,.) L P"Y"'-' + L P,'Y., + Lr Li p,'W" I
= L (p,-,' Y•. ,-, + p,y•. ,-,) + L L p,W" I~
1
i
t
:!>L1t, + L L p,'W" = L Wi' i
We saw in Sections 16.E and 16.F that, under the assumption of concave utility functions, a Pareto optimal allocation of an economy with a finite number of commodities can be viewed as the solution of a planning problem, As described in Figure 20.G.I, the objective function of the planner is a weighted sum of the utilities of the different consumers (the weights being the reciprocal of the marginal utilities of wealth at the equilibrium with transfers associated with the particular Pareto optimum). The arguments of Section 16.E (in particular, Proposition \6.E.2) apply as well to the current infinite-horizon case. Therefore, Proposition 20.G.I has, besides its substantive interest, a significant methodological implication. It tells us that the prices, productions, and aggregate consumptions of a given Walrasian equilibrium correspond exactly to those of a certain single-consumer economy. We give a more precise statement in Proposition 20.G.2. In it we restrict ourselves to the case of a common discount factor, namely, b, = b for all i. Proposition 20.G.2: Suppose that (Ye!, ... , Yi, ... ) is the production path and (Po' ... 'P t ' • • . ) is the price sequence of a Walrasian equilibrium of an economy with I consumers. Then there are weights ('1" ... ,'11) » 0 such that(Yti, ... , Yi, ... ) and (Po, ... , p" ... ) constitute a Walrasian equilibrium for the one-consumer economy defined by the utility Lt btU(C,), where u(c,) is the solution to Max L; '1i Ui(C ti ) S.t. Li Cli :!> c,.
I
Because of the profit maximization condition we get 29
=
Utilities of a Pareto Optimal Allocation (could be the Walrasian allocation)
Proof: We will not give a rigorous proof, but the result is intuitive from Figure 20.G. I. From there we see (technically this involves, as in Proposition \6.E.2, an application of the separating hyperplane theorem) that there are weights ('11' ... ' '1,)>> 0 such that the equilibrium consumption streams maximize Li '1i(L, b'u,(c,i» over all feasible consumption streams, or, equivalently (it is here that the assumption of a common discount factor matters), the aggregate equilibrium consumption stream, solves the two-step planning problem specified by the definition of u(e,) and the maximization of L, b'u(c,). Because we already know (Proposition 20.DA) that this is tantamount to the one-consumer equilibrium problem, we are done. _
,
But this conclusion contradicts (20.GA). 28. Note also that, in the terminology of Chapter 19, the market structure is complete: Every consumer has a single budget constraint and, therefore, only prices limit the possibilities of transferring wealth across periods. 29. Recall that, by convention, Y... _ t = O.
Proposition 20.G.2 allows us to conclude that the one-consumer theory developed in the last three sections remains highly relevant to the several-consumer case. 30 Somewhat informally, we can distinguish two types of properties of an equilibrium. 30. More generally, it remains highly relevant to any equilibrium model that guarantees the Pareto optimality
or equilibria.
Ftgur. 2O.G.1 The Walrasian equilibrium as a solution of a planning problem.
768
cH"
PTE R
2 0:
E QUI L I • R I U"
"N 0
T I .. E
-----------------------------------------------------------------------------The internal properties are those that refer only to the structure of an equilibrium viewed solely in reference to itself (e.g., convergence to a steady state); the eXlernal properties refer to how the equilibrium relates to other possible equilibrium trajectories of the economy (e.g., uniqueness or local uniqueness). The message of Proposition 20.G.2 is that, because of Pareto optimality, the internal properties of an equilibrium of an economy with several consumers are those of its associated one-consumer economy. The implications of the one-consumer theory should not, however, be pushed beyond the internal properties. The reason is that the weights defillillg the planning problem depend on the particular equilibrium considered. For example, it is perfectly possible for there to be more than one equilibrium, each a Pareto optimum but supported by different weights. What can be said about the determinacy properties of equilibrium; for example, about the fmiteness of the number of equilibria? We will not be able to give a precise treatment of this matter, in part because it is very technical and in part because it is still an active area of research where the ultimate results may not yet be at hand. The basic intuition, however, can be transmitted. We begin by pointing out another implication of Proposition 20.G.!. Formally, our infinite-horizon model involves infinitely many variables (prices, say) and infinitely many equations (Euler equations, say). This is most unpleasant, as the mathematical theory described in Section 17.0 applies only (and for good reasons, as we shall see in Section 20.H) to systems with a finite number of equations and unknowns. However, Proposition 20.G.1 allows us to view the equilibrium problem as one of finding not equilibrium prices but equilibrium weillht, ~. If we do this then the equilibrium equations in our system are I - I in number, the same as the number of unknowns. More precisely, the ith equation would associate with the vector of weights'l = ('1" ... , ~,), L~; = I, the wealth "gap" of consumer i:
L, p,(q)'c,,(q) - L, (O"n,(q) + p,(~),w,;) = 0, where p,(~), c,;('Il, and n,(~) correspond to the Pareto optimum indexed by ~. See Appendix A of Chapter 17 for a construction similar to this. At any rate, once looked at as a wealthequilibrating problem across a finite number of consumers, the central conjecture should be that, as in Chapter 17, the equilibrium set is nonempty and generically finite. That is, equilibrium exists and, except for pathological cases, there are only a finite number of weights solving the equilibrium equations (we could similarly go on to formulate an index theorem). Technical difficulties" aside, this central conjecture can be established in a wide variety of cases [see Exercise 20.G.3 and Kehoe and Levine (1985)). We end this section with two remarks. The first derives from the question: Is there a relationship, a "correspondence," between internal and external properties? At least in a first approximation the answer is "no," We have seen that in a onc-consumer economy the
equilibrium is unique, but the equilibrium path may be complicated. Similarly, in a several· consumer economy there may be several equilibria, or even a continuum, each of them nicely converging to a steady state. 32
31. These have to do with guaranteeing the differentiability of the relevant functions.
32. The simplest, trivial, example is the following. Suppose that L = 2, I = 2 and that there is no possibility of interternporal production. Individual endowments are constant through time and the utility functions are concave. Then the intertemporal Walrasian equilibria correspond exactly to the infinite, constant repetitions of the one-period Walrasian equilibria (you are asked to prove this in Exercise 20.G.4). Because there are may be several of those, we obtain Qur conclusion.
SEC T ION
2 0 • H:
0 Y E R L " P PIN G
G ENE R " T ION S
769
-----------------------------------------------~~~~~~~ The second remark brings home the point that Pareto optimality is key to an expectation of generic determinacy. Consider, as an example, a model of identical consumers but with an externality. The utility function, u(k, k', e), now has three arguments: k and k' are the capital investments in the previous and the current periods, respectively, and e is the level of currently felt externality. Given, for the moment, an exogenously fixed externality path (eo, . .. , e" . .. ), the (bounded, strictly interior) capital trajectory k, is an equilibrium if it solves the planning problem for the utility functions u(', " e,), that is, if it satisfies the Euler equations: V,u(k,_"k"e,)
+ oV,u(k"k,."e.. ,) = 0
for all t.
An overall equilibrium must take into account the technology determining the externality. Say that this is e, = k,; that is, the externality is a side product of current investment. Hence, the equilibrium conditions are
",u(k,_" k" k,) + 0 V,u(k" k.. " k,. ,)
=
0
for all t.
(20.G.5)
Suppose that starting from an equilibrium steady state (k, = ii for all I), we try, as we did in Section 20.0, ~o generate a different equilibrium by fixing ko = ii, taking k, to be slightly different from k, and then iteratively solving (20.G.5) for k, + ,. A sufficient (but not necessary) condition for this method to succeed is that Idk,.,/dk,1 < land Idk,.,/dk,_.1
dk,
=
V1,u(')
+ V1,u(·) + oVi,u(·) + Vf,u('))
o(Vi,u(')
(20.G.6)
If there are no externalities [i.e., if V1,u(') = Vf,u(') = 0] then the cO'lcavity of u(·, ,) implies that expression (20.G.6) is larger than I in absolute value (you should verify this in Exercise 20.G.5). Thus, in agreement with the discussion of Section 20.0, we are not then able to find a non·steady-state solution of the Euler equations. But if the externality effects are significant enough, inspection of expression (20.G.6) tells us immediately that dk1+ ,/dk, can perfectly well be less than 1in absolute value. The same is true for dk1+ ,/ok,_" and therefore we can conclude that robust examples with a continuum of equilibria are possible.
2Q,H Overlapping Generations In the previous sections we have studied economies that, formally, have an overlapping structure of firms but only one (or, in Section 20.G, several), infinitely long-lived, consumer. We pointed out in Section 20.B that in the presence of suitable forms of altruism it may be possible to interpret an infinitely long-lived agent as a dynasty. We will now describe a model where this cannot be done, and where, as a consequence, the consumption side of the economy consists of an infinite succession of consumers in an essential manner. To make things interesting, these consumers, to be called generations, will overlap, so that intergenerational trade is possible. The model originates in Allais (1947) and Samuelson (1958) and has become a workhorse of macroeconomics, monetary theory, and public finance. The literature on it is very extensive; see Geanakoplos (1987) or Woodford (1984) for an overview, Here we will limit ourselves to discussing a simple case with the purpose of highlighting, first, the extent to which the model can be analyzed with the Walrasian equilibrium
770
c HAP T E
R
2 0:
E QUI l I 8 R I U MAN 0
TIM E
SEC TIC. N
-------------------------------------------------------------------------------------methodology and, second, the departures from the broad lessons of the previous sections. We shall classify these departures into two categories: issues relating to optimality and issues relating to the multiplicity of equilibria. We begin by describing an economy that, except for the infinity of generations, is as simple as possible. We have an infinite succession of dates t = 0, I, ... and in every period a single consumption good. For every I there is a generation born at time I, living for two periods, and having utility function u(c." COl) where c., and c.. are, respectively, the consumption of the Ith generation when young (i.e., in period C), and its consumption when old (i.e., in period 1 + I); the indices b and a are mnemonic symbols for "before" and "after." Note that the utility functions of the dilTerent generations over consumption in their lifespan are identical. We assume that 11(', .) is quasiconcave, differentiable and strictly increasing. Every generation 1 is endowed when young with a unit of a primary factor (e.g., labor). This primary factor does not enter the utility function and can be used to produce consumption goods contemporaneously by means of some production function J(:)." Say that J(I) = I. Under the competitive price-taking assumption, total profits at C, in terms of period-I good, will be f. = I - 1'(1) and, correspondingly, labor payments will be I - t. Thus, we may as well directly assume that the initial endowments of generation 1 0 units of consumption good. Now let (Po, ... , p" ... ) be an infinite sequence of (anticipated) prices. We do not require that it be bounded. For the budget constraint of the different generations we take P,C.,
+ P.. ,Ca, $; (I
- t)p,
for t > 0
(20.H.I)
and POC.O
+ P'CaO
$;
(1- t)po
+ t(~
p,) + M.
(20.H.2)
These budget constraints deserve comment. For 1 > 0, (20.H.I) is easy to interpret. The value of the initial endowments, available at I, is (I - t)p,. Part of this amount is spent at time 1 and the rest, (I - t)p, - P,c." is saved for consumption at t + I. The saving instrument could be the title to the technology, which would thus be bought from the old by the young at 1 and then sold at 1 + I to the new young (after collecting the period t + I return). The price paid for the asset is the amount saved, that is, (I - 0) p, - P,c.,. The direct return at 1 + I is tp, + I and so, if the asset market is to be in equilibrium, the selling price at t + I should be (I - t)p, - P,C., - cp,+ ,. In summary, in agreement with the budget constraint (20.H.l) this leaves (I - ")p, - P,C., to be spent at 1 + I. The constraint (20.H.2) for 1 = 0 is more interesting. Its right-hand side is the value of the asset to generation O. Note that asset market equilibrium requires that
33. The assumption thal production is contemporary with input usage fits well with the lenglh
or I he period being long.
20. H:
0 Y E R lAP PIN G
G ENE RAT ION.
771
----------------------------------------------------------------------------this value should be at least the Jundamental value, that is, t(l:, p,).)' Indeed, the value of the asset at t = 0 equals the profit return EPO plus the price paid by the young of generation I. At any T, the price paid by the young of generation Tshould not be inferior to the direct return tPT+ I' In turn, at T - 1 it should not be inferior to the direct return plus the value at T; that is, it should be at least t(PT + PT+ I)' Iterating, we get the lower bound t(p, + ... + PT+ ,) for the price paid by generation I, which, going to the limit and adding tpo, gives tel:, p,) as a lower bound for the value to generation O. Thus, in terms of expression (20.H.2) a necessary condition for equilibrium is M 0). We did not do so in Sections 20.D or 20.G because with a finite number of consumers, bubbles are impossible at equilibrium. The equality of demand and supply implies that the (finite) value of total endowments plus total profits equals the value of total consumption, and therefore at equilibrium no individual value of consumption can be larger than the corresponding individual value of endowments and profit wealth (you should verify this in Exercise 20.H.I). We will see shortly that under some circumstances bubbles can occur at equilibrium with infinitely many consumers. It would therefore not be legitimate to eliminate them by definition. The definition of a Walrasial1 equilibrium is now the natural one presented in Definition 20.H.1. Defin1tion 20.H.1: A sequence of prices (Po • ...• P, • ... ). an M
c:._,
In a process reminiscent of the iterative procedure (presented in Section 20.D) for the determination of the policy function from the Euler equations, Figures 20.H.I and 20.H.2 describe how we could attempt to construct an equilibrium. Normalize to Po = I. Suppose that we now try to arbitrarily fix caO ' At equilibrium, c. o = I, and thus PI is determined by the fact that p,/Po must equal the marginal rate of substitution of u(', .) at (I, caO ). Also, c. 1 = I - Cao. This now determines p,. Indeed, p,. the price at period 2, should.be fixed at a value that induces a level of demand by generation I in period I of precisely c., [under the budget set given by p" p, and wealth (I - c)p,J. With this, the demand of generation I in period 2, and therefore the residual amount c" left in that period for generation 2, has also been determined. But then we may be able to fix Pl at a value that precisely induces the right amount of demand by generation 2 in period 2, that is, c. 2 . If we can pursue this construction indefinitely so as to generate an infinite sequence (p" ... , P.... .), then we have found an equilibrium. In Figure 20.H.I, where t > 0, there is a single price sequence (with Po = I) that can be continued indefinitely, and therefore a single equilibrium path.
34. Strictly speaking, we are saying that if the consumption good prices are given by (Po.···. P,.···) and the asset prices present no arbitrage opportunity, then the price of the asset should be at least as large as its fundamental value.
772
SEC T ION
CHAPTER
20:
EOUILIBRIUM
AND
TIME
~~~~~~~----------------------
Suppose first that s > O. We say then that the asset is real (it has "real" returns). At an equilibrium the wealth of generation 0, (1 - s)Po + seL p,) + M, must be finite (how could this generation be in equilibrium otherwise?). Therefore, if s > 0, it follows that L, p, < OOB An important implication of this is that the aggregate (Le., added over all generations) wealth oj society, which is precisely L, P.. isjinite. In Proposition 20.H.1 we now show that, as a consequence, the first welfare theorem applies for the model with, > O.
Offer Curve
Proposition 20.H.1: Any Walrasian equilibrium (Po,." ,p" .. .), {{c~,. c:,l},=o, with L, Pt < 00 is a Pareto optimum; that is, there are no other feasible consumptions {{Cb" ca,)};':,o such that u{cb" Ca,) :? U{Cb,. c:,) for all t :? 0, with strict inequality for some t.
"'' -:)1-,
f',', , \'" "",_-+, '{'" I ___
Proof: We repeat the standard argument. Suppose that {(c... c.,)}~o Pareto dominates {(e:.. c:,J},":.o· From feasibility, we have c:, + C:.'_I = I and c., + C•. '_I :; 1 for every I. Therefore, L, p,(c:' + C:.,_I) = L, p, and L, p,(c., + C•. '_I) :; L, p,. Because L, p, < 00, we can rearrange terms and get
Figure 20.H.l
~""l--
C. 2
L'--C-'.-,---c...L,,~l"'-'--t+-~-C-o-'n~s-umplion in
first Period of Life
Overlapping generations: construction of
the equilibrium (case t > 0).
L (Pt c., + p,+ I C.,) :; L (p,c:' + p,+ , C:,) = L, p, < 00. Because the utility function is increasing and (c:" e:') maximizes utility in the p,c:, + p, + IC:, for every t, with at budget set we conclude that P,c., + PH I least one strict inequality. Therefore, L, (p" C.I + p,+ IC.,) > L, (p,c:, + p,+ IC:')Contradiction. _
C., ;::
Consumption in Second Period of Life 45' =
0 V E R LAP PIN G
Pareto Optimality
Consumption in Second Period of Life
c.,
2 0 _ H:
~
•
No-Trade Sleady-Stale COl
C.o
Consumption in first Period of Life
G ENE RAT ION S
773
------------------------------------------------------------------------------
Proposition 20.H.1 is important but it is not the end of the story. Suppose now that the asset is purely nominal (i.e., , = 0; for example, the asset could be fiat money, or ownership of a constant returns technology). Then it is possible to have equilibria that are /lot optimal. I n fact, it is easy to see that we can sustain autarchy (i.e., no trade) as an equilibrium_ Just put M = 0 (no bubble, worthless fiat money) and choose (Po, ... , p". _.) so that, for every t, the relative prices p,/p, + I equal the marginal rate of substitution of u(', .) at (I, 0), denoted by p. This no-trade stationary state (also called tlte nonmonetary steady state) where every generation consumes (1,0) is represented in Figure 20.H.2. As it is drawn (with p < I), we can also see that the no-trade outcome is strictly Pareto dominated by the steady state (y, 1 - y) [or, more precisely, by the consumption path in which generation 0 consumes (I, 1 -)') and every other generation consumes (y, 1 - y)]. What is going on is simple: in this example the open-ended ness of the horizon makes it possible for the members of every generation I to pass an extra amount of good to the older generation at t and, at the same time, be more then compensated by the amount passed to them at 1+ 1 by the next generation_ Note that, in agreement with Proposition 20.H.I, the lack of optimality of this no-trade equilibrium entails p,/p, + I = P< I for all t; that is, prices increase through time. It is also possible in the purely nominal case for an equilibrium with M > 0 not to be Pareto optimal. Note first if {(e:,. c:,)},"';,o, (Po"", P..... J and M constitute an
Figure 20_H.2
Overlapping generations: construction of
equilibria (case t = 0)
It corresponds to the stationary consumptions (y, 1 - y) and the price sequence p, = 0:', where 0: = (I - s - y)/(l - y) < L Note that the iterates that ~gin at a ~alue c.o '" 1 - y unavoidably "leave the picture," that is, bec~me unfe~s~~le. In Flg~re 20.H.2, where s = 0, there is a continuum of equlhbna: any tnltlal conditIOn
C.o :; 1 - y can be continued indefinitely. .. . It is plausible from Figures 20.H.1 and 20.H.2 that the existence of an eqUlllbnum can be guaranteed under general conditions- This is indeed the case [see Wilson
35. You can also verify this graphicaJly by examining Figure 20.H.l.
(1981)].
1
774
CHAPTER
20:
EQUILIBRIUM
AND
TIME
equilibrium, then we have (recall that cto = I) p,+ ,c:,
= p,(1
- ct,)
= p,c:,,_, = ... = p,e:o = M
for every t.
-
c.,
/
Monetary Sleady State
1-",
Consumption in First Period of Life
0 V E R LAP PIN G
ProposItion 20.H.2: Suppose that at an equilibrium we have LtP t <
00.
Then M =
o.
Multiplicity of Equilibria
Flgur. 20.H.3 The monetary steady state is Pareto optimal.
Consumption in Second Period of Life
20. H:
Proof: The sum of wealths over generations is L, p, + M < 00. The value of total consumption is L, p, < 00. The second amount cannot be smaller than the first (otherwise some generation is not spending its entire wealth). Therefore M = O. •
Thus, M = 0 can occur only at a no-trade equilibrium. In Figure 20.H.2, there is a continuum of equilibria indexed by COl for y s; s; I. The no-trade equilibrium corresponds to = I. But for every COl < 1 with COl > Y we have a nonstationary equilibrium trajectory with trade (hence M > 0) which is also strictly Pareto dominated by the steady state (y, I - y). Nonetheless, it is still true that for any equilibrium with COl> Y we have Mfp, -+ 0; that is, in real terms the value of the asset becomes vanishingly small with time. For CO, = y, matters are quite different. We have a steady-state equilibrium (called the monetary steady state) in which the price sequence p, is constant and therefore the real value of money remains constant and positive. This monetary steady state is the analog of the golden rule of Section 20.E and, as was the case there, we have that. in spite of L, p, < 00 being violated. the monetary steady state is Pareto optimal. We will not give a rigorous proof of this. The basic argument is contained in Figure 20.H.3. There we represent the indifference curve through (y, I - y) and check that any attempt at increasing the utility of generation 0 by putting c.. < y leads to an unfeasible chain of compensations; that is, it cannot be done. The discussion just carried out of the examples in Figures 20.H.2 and 20.H.3 suggests and confirms the following claim, which we leave without proof: In the purely nominal case, of all equilibrium paths the Pareto optimal ones are those, and only those, that exhibit a bubble whose real value is bounded away from zero throughout time. It is certainly interesting that a bubble can serve the function of guaranteeing the optimality of the equilibria of an economy, but one should keep in mind that this happens only because an asset is needed to transfer wealth through time. If a real asset exists then this asset can do the job. If one does not exist then the economy, so to speak, needs to invent an asset. To close the circle, we point out that if there is a real asset then not only is a bubble not needed but, in fact, it cannot occur.
c.,
SEC T ION
G ENE RAT ION S
775
-------------------------------------------~~~~~~~~
We have already seen, in Figure 20.H.2, a model with a purely nominal asset (i.e., c = 0) and very nicely shaped preferences (the offer curve is of the gross substitute type) for which there is a continuum of equilibria. Of those, one is the Pareto optimal monetary steady state and the rest are nonoptimal equilibria where the real value of money goes to zero asymptotically. The existence of this sort of indeterminacy is clearly related to the ability to fix with some arbitrariness the real value of money (the "bubble") at t = 0, that is Mfpo. It cannot occur if bubbles are impossible, as, for example, in the model with a real asset (i.e., £ > 0) where, in addition, we know that the equilibrium is Pareto optimal. One may be led by the above observation to suspect that the failure of Pareto optimality is a precondition for the presence of a robust indeterminacy (i.e., of a continuum of equilibria not associated with any obvious coincidence in the basic data of the economy). This suspicion may be reinforced by the discussion of Section 20.G, .where we sa~ that the Pareto optimality of equilibria was key to our ability to claim the generic determinacy of equilibria in models with a finite number of consumers. Unfortunately, with overlapping generations the number of consumers is infinite in a fundamental way,'" and this complicates matters. Whereas with a rear asse~ the Pareto optimality of equilibria is guaranteed and the type of indeterminacy of Figure 20.H.2 disappears, it is nevertheless possible to construct nonpathological examples with a continuum of equilibria. The simplest example is illustrated in Figure 20.H.4. The figure describes a real-asset model with the steady state (y, I - }'). Suppose that, in a procedure we have resorted to repeatedly, we tried to construct an equilibrium with c. o slightly different from I - y. Then, normalizing to Po = I, we would need to use p, to clear the market of period 0, P2 to do the same for period I, and so on. In the leading case of Figure 20.H.I, we have seen that this eventually becomes unfeasible. A change in p, that takes care of a disequilibrium at t - I creates an even larger disequilibrium at t, which then has to be compensated by a change of a larger magnitude in p, + I In an explOSive process that finally becomes impossible. But in Figure 20.H.4, the utility function is such that, at the relative prices of the steady state, a change in the pnce of the second-period good has a larger impact on the demand for the first-period good than on the demand for the second-period good. Hence. the successive adjustments necessitated by an initial disturbance from c = 1 _ " dampen with each iteration and can be pursued indefinitely. We conclud~ that a~ equilibrium exists with the new initial condition. As a matter of terminology, the
, 36. By this v
rn·lOY consumers are sufllClently SImilar for them to be "approximated" by a finite number of rcprcsen ta t i vcs.
776
C HAP T E R
2 D:
E QUI LIB R I U MAN D
TIM E
------------------------------------------------------------------- ---------------------------------------------------------------------SECTICN
Consumption in Second Period of Life
1-,' e. o
Figure 2O.H.4
An example of a continuum of (Pareto optimal) eqUilibria in the real asset case.
Consumplion in First Period of Life
locally isolated steady state equilibrium of Figure 20.H.1 is called de/ermina/e. and the one of Figure 20.H.4 is called inde/ermina/e. 37 It is interesting to point out that the leading case of unique equilibrium (Figure 20.H.1) in a real-asset model corresponds to a gross substitute excess demand function. while Figure 20.H.4 represents the sort of pronounced complementaries that were sources of nonuniqueness in the examples of Sections Is.B (recall also the discussion of uniqueness in Section 17.F). The connection between non uniqueness and indeterminateness is actually quite close. and you are asked to explore it in Exercise 20.H.2. Here we simply mention that gross substitution is not a necessary condition for uniqueness. It can be checked. for example. that in the real asset model the steady state remains the only equilibrium if consumption in both periods is normal in the demand function of u(· • . ) and if the corresponding excess demand (z.(P •• p,). z,(P •• p,» satisfies for all P•• p,.
2D.H:
OVERLAPPING
GENERATIONS
In Chapter 17 (sec Seclion 17.D and Appendix A of Chapter 17) we argued that, with Pareto optimalily, an equilibrium problem with a finite number of consumers could be represented by means of a finile number of equations with the same number of unknowns. From this we claimed that generic determinacy was the logical conjecture to make for this case. In Section 20.G wc extended this argument to the model with a finite number of infinitely long-lived consumers. However, the current overlapping generations problem has a basic difference in formal structure: there is no natural trick allowing us to see the equilibrium as anything but the zeros of an infinite system of equations (of the excess demand type, say). Mathematically, this is significant. To give an example, intimately related to the issues we are discussing, suppose that f: R" ~ R' is a linear map that is onto (i.e., fIx) = Ax, where A is a nonsingular matrix). Then 0 is the unique solution to fIx) = O. But suppose now that f(·) maps bounded sequences into bounded sequences and that it is linear and onto. Then fIx) = 0, or, equivalently, I.(x" ... , x" ... J = 0 for all I, need not have a unique solution. A simple example is the back ward shift, that is, I. (x" ... , x" ... ) = x, + 1, where any (~, 0, ... , 0, ... ) is sent to zero. What can we say about the dynamics of an equilibrium? We saw that the "anything goes" principle applied to the one-consumer model. It would be surprising if it did not apply here; indeed. in Figures 20.11.5 and 20.1-1.6 we provide nonpathological examples with cycles.'8 Note
Cl1nsumption in p,,;od of Life
S~COIlJ
Consumption in Second Period of Life
,
451~
",
Offer Curve
OlTer Curve
"k------__c-- ___ _
," I
t
:
,
"
,
,,,
'
,
"
t'
-----+------~~
"""
(20.H.3) 1- c
Expression (20.H.3) permits a price increase in the first period of life to lead to an increase in demand in this period (a possibility ruled out by gross substitution); but. if so. it requires the increase of demand in the second period of life to be larger. Geometrically speaking. the condition is that the slope of the offer curve in the (c •• c,) plane should never be positive and less than I. Note that in Figure 20.H.4 this is violated at the steady state. Condition (20.H.3) is known as the de/erminacy condilion. If the reverse inequality holds at the steady state. then. as in Figure 20.H.4. there is a continuum of equilibria all converging to the steady state (the steady state is therefore ilJde/ermilJa/e).
37. Observe that, at least in the context of the relatively simple model we are now discussing, there is little room for cases intermediate between uniqueness or the existence of a continuum of equilibria.
I
Consumption in First Period of Life
that in Figure 20.H.6 we have a three-period cycle: chaos rears its head. In the gross substitute example of Figure 20.H.1 we have monotone convergence to the steady-state. In a sense, the gross substitule case is the analog of the approach based on the sign of the second derivatives described in Section 20.F. Note that in the overlapping generations situation the factor of discount is not a meaningful concept and, therefore, there is no analog of a dynamic theory based on palience. In Section 20.G we also mentioned, quite loosely, that there did not seem to be, for the case of a finite number of agents with Pareto optimality, a close relation between the determinacy and the dynamic properties of equilibrium. In the current setting the connection is closer, at least in the following sense: If equilibrium trajectories with cycles can Occur, then there are infinitely many equilibria.
38. In p<.IrlicuJar, no inferior goods are required for these examples.
Consumption in First Period of Life Figure 20.H.5 (Ieh)
Complementary consumptions: example of a period·2 equilibrium path. Figure 20.H.6 (right)
Complementary consumptions: example of a period-3 equilibrium path.
777
778
C HAP T E R
2 0:
E 0 U I I I B R I U MAN D
TIM E
------------------------------------------------------------------------20.1 Remarks on Nonequilibrium Dynamics: Tatonnement and Learning The dynamic analysis that has concerned us so far in this chapter is of a different nature from, and should not be confused with, the dynamics studied in Section 17.H. The dynamics here display the temporal unfolding of an equilibrium (an internal property of the equilibrium. in the terminology of Section 20.G). whereas in Section 17.H we were trying to assess the dynamic forces that. in real or in fictional time, would buffet an economy disturbed from its equilibrium (hence, we were looking at an external property). As we saw, nonequilibrium dynamic analysis raises a host of conceptual problems, yet it may offer useful insight into the plausibility of the occurrence of particular equilibria. This remains valid in the setting of intertemporal equilibrium. Abstracting from technical complexities, the analysis and the results of Section 17.H can be adapted and hold true for the infinite-horizon, finite number of consumers model of Section 20.G. On the other hand, as we have seen. the temporal framework has its own special theory. which could conceivably be illuminated by specific nonequilibrium considerations. We make three remarks in this direction. Short-Run Equilibrium and Permanent Income 39 Suppose that (Po •. ..• P, • ... ) is the equilibrium price sequence of an economy with L goods and I consumers. Consumers are as in Section 20.0. Then at the equilibrium consumptions we have (assuming interiority) (j'
Vu,(c,,) = A,p,
for aliI and every i.
(20.1.1)
This is just (20.0.6). The variable A, is the marginal utility of income. or wealth. and the vector of reciprocals ('I, •...• 'I,) = O/A, •...• I/A,) can serve as the weights for which the given equilibrium maximizes the weighted sum of utilities (see Section 20.G). It follows from (20.1.1) that the short-run demands (i.e .• the demands at t = 0) are entirely determined by Po and the marginal utilities of wealth )'i' Denote this demand by co,(P. ).,). In the spirit of tatonnement dynamics. suppose that Po is perturbed to some p~. What will happen to demand at t = C1! If the A, remain fixed. then (20.1.1) implies that short-run demand behaves as the demand for non-numeraire goods in a quasilinear utility model with concave utility functions. In particular, differentiating (20.1. I) we see that the L x L matrix of short-run price effects
DpOco,(Po. A,) = ).,[D 2 u,(co,)]-'
':.,1.
:~
is negative definite (by the concavity of u,('» and. therefore. so is the aggregate L, DpOco,(Po, i.,). In more economic terms. as long as the )., stay fixed there are no wealth effects present in the short-run demands. Substitution prevails and. consequently, the short-run equilibrium is unique and globally tatonnement stable. In reality, however. after a change in Po we should expect that i., will have changed at the new consumer optimum. But if the rate of discount is close to I (i.e .• if agents 39. See Bewley (1977) for more on this topic. The term "permanent income" is standard and so we use it rather than "permanent wealth."
ij
l
SECTION
20.1:
REMARKS
ON
NONEOUILIBAIUM
DYNAMICS
779
---------------------------------------------------------------------~~~ are patient) then the change in )., should be small: The current period is not significantly more important than any other period and. therefore. it will account for only a small fraction of total utility and expenditure. Hence. we could say that partial equilibrium analysis is justified in the short run (recall the discussion of partial equilibrium analysis in Section 10.G). In summary: If consumers are sufficiently patiellt. thell the short-run eqUilibrium is unique and globally stable (for the tatonnement dynamics). The (Short-Run) Law of Demand in Overlapping Generations Models We now look at the short-run equilibrium of the overlapping generations model of Section 20.H. This is an example of a model where wealth effects matter in the short run and, therefore, the permanent income approach does not apply. We consider the version of the model with a real asset and normal goods and ask whether the stability of the fictional-time tatonnement dynamics at a given date t helps us to distinguish among types of equilibria. Because there is a single good per period. the stability criterion for a single period is simple enough-it amounts to the law of demand at time t. That is. we say that an equilibrium (Po •...• p" ... ) is tatonnement stable at time I if an (anticipated) increase in PI' all other prices remaining fixed. results in excess supply in that period (note that only generations t - 1 and t will alter their consumption plans). We know that if the excess demand function of the generations is of the gross substitute type. then there is a unique equilibrium (which is in steady state). (See Figure 20.H. L) Moreover. the definition of gross substitution tells us that the law of demand is satisfied at any t. This gives us a first link between the notions of determinate equilibrium and tatonnement stability. This link can be pushed beyond the gross substitute case. Take a steady-state equilibrium price sequence (I, P.·· .• P'.·· .). By the homogeneity of degree zero of excess demand functions (t o (· • • ), z.(·, .)). which implies the homogeneity of degree -I of Vz.(· •. ) and V=.C •. ). we have (you should verify this in Exercise 20.1.1) V,z.(I/p. I)
+ V,z.(I. p)
=
pV2 z.(I. p) + V,Zb(l. p)
=
-V,z.(I. p)
+ V,z,(I. pl.
The negativity of the left-hand side is the tatonnement stability criterion. that is. the law of demand at a single market,.o while the negativity of the right-hand side (i.e .• the requirement that wealth effects are not so askew that a decrease in the price In one period increases the demand of the young in that period by less than it increases the demand of these same young for their consumption in the next period) is the criterion for the determinacy of the steady state [see expression (20.H.3»). Recall that determinate means that there is no other equilibrium trajectory that remains in an arbitrarily small neighborhood of the steady state. We conclude that there is an exact correspondence: a sleady-state eqllilibrillm is (shorl-run. locally) tatonnemelll .\lanle at any t if alld only if il is tielerminate!'
40. If p, is changed infinitesimally then the demand of the old changes by V,z.(p'-',p') while the demand oflhe young changes by V,z,(p', pO+'). Because V,z(', .) and V,z(', .) are homogeneous of degree -I. the total change equals (lIp') V,z.(l/p, 1) + (lip') V,z,(l. p) 41. In this "ir and only ir statement we neglect borderline cases.
780
c HAP T E"
• 0:
E QUI LIB" I U M
" NOT I M E
~~~~~~-------------------------------------------We have confined ourselves to the real asset case to avoid a complication. With a purely nominal asset the previous concept of tatonnement stability loses the power to discriminate between determinate and indeterminate steady-state equilibria, unless we restrict ourselves a priori to monetary steady states (to see this, consider the simplest gross substitute case). The learning concept to be presented in the remainder of this section does not suffer from this limitation.
--
SEC T ION
•
O.
I:
"E M A" K SON
NON E QUI LIB" I U M
0 YNA M I
CS
Consumption in Second Period of Life
"
45~
Offer Curve
Leaming We now briefly discuss a nonequilibrium dynamics that takes place in real time and that can be interpreted in terms of learning. The framework is that of the overlapping generations of Section 20.H and, to be as simple as possible, we focus on the purely nominal asset case. We describe first how the short-run equilibrium (i.e., the equilibrium at a given period t) is determined. We suppose that there is a certain fixed amount of fiat money M (denominated, say. in dollars). The excess demand of the older generation at date t ~ 1 is then M/p,. The excess demand of the younger generation at the same date depends on p, but also on the expeclation P:+, of the price at I + 1. Given P:+" the price p, is a temporary equilibrium al lime I ~ 1 if Zb(P,. P:+ ,) + (M/p,) = O. Thus. given a sequence of price expectations (pi •... • P:, ... ). we generate a sequence of temporary equilibrium prices (p" . ..• p" ... ). But. how are expected prices determined? To take them as given does not make much sense. The sequence of realizations should feed back into the sequence of expectations. The self-fulfilled, or rational, expectations approach (which we have implicitly adhered to in this chapter) imposes a correct expectations condition: P:+, = p,+, for every I!' An alternative is to require that P:+l (the price expected at I to prevail at I + I) be an extrapolation of the past (and current) realizations Po, ...• p,. In this approach we think of consumers as engaged in some sort of learning and of expectations responding adoptively to experienced outcomes!) To be specific. let us take a not very realistic, but very simple, extrapolation rule: + I = p, _ I (i.e .• the price at I + 1 expected by the young at I ~ 1 is the price that ruled in the most recent past). Equivalently (given the fixed amount of fiat money M). the young at t expect to consume at t + I, when old, the same amount consumed by the old at I - 1. The equation for the determination of p, is then
P:
(20.1.2) By Walras' law. (20.1.2) can equivalently be written as M
z.(p"p,_,)=--.
p,-,
(20.1.3)
Consumption in First Period of Life
Given an arbitrary initial condition Po. we can then compute the sequence of temporary equilibrium realizations (p, •...• p" . ..) by iteratively using (20.1.2) or (20.1.3). Note that in doing so, the planned excess demands in (20.1.2) will be realized but those in (20.1.3) may not (because p,+, may not be equal to p,_,). We represent the dynamic process in Figure 20.1.1. In the figure, c., and c!, stand for the planned consumptions of generation I at times t and t + I. respectively. Given M/p,_, we get c!, from (20.1.3), and c., from the fact that planned consumptions are in the offer curve. Finally (20.1.2) moves us to the next value M/p,. For generation I we also show the actual consumption vector (COl' c.,). From Figure 20.1.1 we can see an interesting fact: The learning dynamics exactly reverses the equilibrium dynamics (compare with Figure 20.H.2).·· For the gross substitute case shown in the figure. this means that all the trajectories tend to the monetary steady state. Hence, in the limit we have a true self-fulfilled expectations equilibrium. Consumers have learned their way into equilibrium. so to speak. For the crude learning dynamics we are considering, this need not be so for the case of a general ofTer curve (an infinite sequence with systematic prediction error is quite possible), but the property of exact reversal of equilibrium dynamics suffices to provide, once again. a test for the well-behaved ness of steady states that reinforces the intuitions developed earlier: A steady state is (locally) stable for rile learllillg dynamics if and only if it is determinate (i.e .• "locally isolated").
42. The lerm "self-fulfilled" is justified because the sequence of expectations (Pl.· .. , pr•... ) induces a sequence of realizations identical to itselr. The term "rationa'" is justified by the fact that, given (pr •.... pt... .). a member of generation t should. in principle. be able to compute the price realization p,+ I and therefore verify the correctness of pt+ I_ 43. We should emphasize, first, that all this is a nonequilibrium story and, second, that we
cannot rigorously discuss learning without explicitly introducing an uncertain environment.
44. More precisely. if (PI' ... . PI' .. .) is the sequence of realizations of the adaptive expectations dynamics. then for any T there is an equilibrium sequence (Po •... , p;, ... ) such that p; = PT-I for every' < T.
Figure 20.1.1
Learning dynamics.
781
782
CHAPTER
20:
EQUtLtBRtU M
AND
TtME
REFERENCES Allais, M. (1947). Ecorwmi••1 Inter'l. Paris: ImprimOrie National.. Barro, R. (1989). The Ricardian approach to budget deficits. Jm",,,,1 oj Economic Pm".clives 3: 37-54. Bewley, T. (1977). The permanent income hypothesis: A theoretical formulation. Journal oj Economic Theory 16: 252-92. Blackorby, c., D. Primon~ and R. Russell (1978~ Dualily, SeparabililY, and Funclional Struclure: Theory and Economic A.pplications. Amsterdam: North-Holland. Blanchard, 0., and S. Fischer. (1989). UCIUrtS on Macroeconomics. Cambridge, Mass.: MIT Press. Boldrin, M. and L. Montruocio. (1986). On the indeterminacy of capital aocumulation paths. Journal oj Economic Theory ~ 26-39. Bliss, C. (1975~ Capital Theory and Ihe DiS/ribulion of Income. Amsterdam: North-Holland. Brock., W. A., and E. Burmeister. (1976). Regular economies and conditions ror uniqueness of steady-statcs in optimal multisector economic models. International Economic Review 17: IOS-20. Cass, D. (1972). On capital overaocumulation in the aggregative, neoclassical model of economic growth: a complete characterization. Journal of Economic Theory of: 200-23. Deneckere, R. and J. Pelikan. (1986). Competitive chaos. Journal of Economic Theory 40: 13-25. Gale, D. (1973). On the theory of interest. American Mathemalical Monlhly 88: 853-68. Geanakoplos, J. (1987). Overlapping generations. Entry in The N~ Palgra~: A Dictionary of Economics, edited by J. Eatwell, M. Milgate, and P. Newman. London: Macmillan. Grandmont, 1. M. (1986). Periodic and aperiodic behavior in discrete one dimensional systems. In Contributions to Mathematical Economics. edited by W. Hildenbrand, and A. Mas-Colell.
Amsterdam: North-Holland. Kehoe, T., and D. Levine. (1985). Comparative statics and perfect foresight. Economelrica 53: 433-54. Koopmans, T. C. (1960). Stationary ordinal utility and impatience. Econometrica 28: 287-309. Malinvaud, E. (1953). Capital aocumulation and efficient allocation of resources. Econometrica 21: 223-68. McKenzie, L. (1987). Turnpike theory. Entry in The New Palgrave: A Dlclionary oj Economics, edited by J. Eatwell, M. Milgate, and P. Newman. London, Macmillan. Ramsey, F. (1928). A mathematical theory of saving. Economic Journal 38: 543-49. Samuelson, P. A. (1958). An exact consumption-loan model of interest without the social contrivance of money. Journal of Polillcal Economy 66: 467-82. Santos. M. S. (1991). Smoothness or the policy runction in discrete time economic models. Econometrica 59: 1365-82. Solow, R. M. (1956). A contribution to the theory or economic growth. Quarterly Journal of Economics 70: 65-94. Stokey, N., and R. Lucas, with E. C. Prescott. (1989). Recursive Melhodsin Economic Dynamics. Cambridge, Mass.: Harvard University Press. Swan, T. W. (1956). Economic growth and capital aocumulation. Economic Record 32: 334-61. Uzawa. H. (1964). Optimal growth in a two-sector model or capital accumulation. Review of Economic Studies 31: 1-24. Weizsiicker, C. C. von (1971). Steady State Capital Theory. New York: Springer-Verlag. Wilson, C. (1981). Equilibrium in dynamic models with an infinity of agents. Journal of Economic Theory 24: 95-1 I 1. Woodford, M. (1984). Indeterminacy of equilibrium in the overlapping generations model: a survey. Mimeograph, Columbia University.
EXERCISES 20.8.1" Adopting the definition of lime impalience given in comment (I) of Section 20.B, show that a utility function of the form (20.B.I) exhibits time impatience. 20.B.2" Verify that a utility function of the form (20.B.I) is stationary according to the definition given in comment (2) of Section 20.B. Also, exhibit a violation of stationarity with a utility function of the form V(c) = L~o cl:u(c,).
E X E R CIS E S
20.B.3" With reference to comment (3) of Section 20.B, write c = (c', CO) where c' = (co, ... , c,), c" = (C<+I'· .. ). Suppose that the utility function V(-) is additively separable. Show that if V(C', c") ;;, v(c', c'") for some C', then V(c', c");;, V(c', for all c'. Show that if V(c', C);;, V(C', C') for some e, then V(c', c") ;;, V(c", CO) for all c'. Interpret.
n
20.B.4 c Show that in a recursive utility model with aggregator function G(u, V) = u" + cl V", 0<. < I, ~ < I, and increasing, continuous one-period utility u(c,), the utility Vic) of a bounded consumption stream is well defined. [Hinl: Use (20.B.3) to compute the utility for consumption streams truncated at a finite horizon. Then show that a limit exists as T _ <:fl. Finally, argue that the limit satisfies the aggregator equation.] 20.B.5' Show that the utility function V(c) on consumption streams given by (20.8.1) is concave. Show also that the additively separable form of V(·) is a cardinal property. 20.C.1' Given the price sequence (Po, P,' ... ' p" .. •), p, E R L , define for every I and every commodity ( the rate of interest from I to I + I in terms of commodity t (this is known as the own rale of ililereSI of commodity ( at I). 20.C.2' Show that if the path (yo, ... , y" ... ), is myopically profit maximizing for (Po, P,'···' P.. ···);;, 0, then (yo,··., y" ... ) is also profit maximizing for (Po, P,' ... ' p" ... ) over any finite horizon, in the sense that, for any T, the total profits over the first T periods cannot be increased by any coordinated move involving only these periods. 20.C.3' Define an appropriate concept of weak efficiency and reprove Proposition 20.C I, requiring only that (Po, ... , p" ... ) is a nonnegative sequence with some nonzero entry. 20.C.4" Suppose that the production path (Yo, . .. ,y" . .. ) is bounded (i.e., there is a fixed. such that 11.1',11 ,;. for allt),that (Po, ... , p" •.. )>> 0, and that L~o p, < 00. We say that the path (Yo, ... , .1'" ••• ) is overall profit maximizing with respect to (Po, ... , p" . .. ) if
f
(p,Y,,+P<+I·Y•. ,);;'
1"'0
for any other production path
(y~,
...
,Y;, ... ).
that if (yo, ... , y..... ) is (Po, ... ,P .. ·· .) »0, then it is efficient.
(a) Show
t (p,y;,+p,.,.y~.,)
1=0
overall
profit
maximizing
with
respect
to
(b) Show that if (yo, ... , Y... .. ) is myopically profit maximizing with respect to (Po, ... , p". .)>> 0, then it is also overall profit maximizing.
20.C.5 c Say that a production path (Yo, ... , Y.. ... ), is T-efficient, for T < <:fl, if there is no other production path (y~, ... , Y;, ... ) that, first. dominates (yo, ... , }'" ... ) in the sense of efficiency and, second, is such that the cardinality of {I: y, '" Y;} is at most T. (a) Show that if (yo .... , y" ... ) is myopically profit maximizing with respect 10 (Po,· ., P.. ···J» 0, then (yo,···, J ..... ) is T-efficient for all T < Xl. (b) Show thaI if the technology is smooth (in the sense used in the small-type discussion at the end of Section 20.C; assume also that the outward unit normals to the production frontiers are strictly positive), then 2-efficiency implies T-efficiency for all T < <:fl. (c) (Hardcr) Show that the conclusion of (b) fails for general linear activity technologies. Exhibit an example. [Hill!: Rely on chains of intermediate goods.] 20.C.6 A Consider the Ramsey-Solow technology of Example 20.CI, as continued in Example 20.C6. The exogenous path of labor endowments is (10' ... , I..... ). Given a production path (k o,· .. ,k" ... ), we determine a sequence of consumption good prices (qo, . .. ,q" ... ) by the requirement that (q,/q,.,) = V,F(k"I,) for aliI. Show then that a sequence of wages w, can
783
784
c HAP
TE R
2 0:
E QUI LIB R I U"
AND
T I .. E
---------------------------------------------------------------------------------be found so that the path determined by (k o,' .. , k" ... ) is myopically profit maximizing for the price seq uence determined by «q .. wo), ..• , (q.. w,), ... ). 20.D.I A Consider the budget constraint of problem (20.0.3). To simplify, suppose that we are in a pure exchange situation. Write the budget constraint as a sequence of budget constraints, one for each date. To this effect, assume that money can be borrowed and lent at a zero nominal rate or interest. 20.D.2A Show that condition (jj') in Section 20.0 (it is stated just before Definition 20.0.2) implies condition (ii) of Definition 20.0.1. Show that, conversely, condition (ii), together with W = L, p,W, + L, 1t, < 00, implies condition (ii'). 20.D.3 A in text. 20.0.4A Complete the computations requested in Example 20.0.1. 20.D.Sc In the context of Example 20.0.3, compute the Euler equations for the optimal investment policy when the production function has the form F(k) = k', 0 < a < I, and the adjustment cost function is given by g(k' - k) = (k' - k)', with (i> I, for k' > k, and by g(k' - k) = 0 for k' S k. Say as much as you can about the policy. In particular, determine the steady·state trajectory of investment. 20.D.6 8 Verify the claim made in the proof of Proposition 20.0.7 that the Euler equations (20.0.9) are the first-order necessary and sufficient conditions for short-run optimization. In other words: they are necessary and sufficient for the nonexistence of an improving trajectory differing from the given one at only a finite number of dates. 20.D.7 A With reference to Example 20.0.4, show that, for the functional forms given, the Euler equations are as indicated in the example: k,., = 3k, - 2k,., for every I. Also verify that the solution to this difference equation given in the text is indeed a solution, that is, that it satisfies the equation. 20.D.8 A Verify that the value function V(k) does satisfy the properties (i) and (ii) claimed for it at the end of Section 20.0. 20.D.9 A Argue that the properties (i) and (ii) of the value function referred to in Exercise 20.0.8 yield the two consequences, concerning V'(k) and V"(k), claimed at the end of Section 20.0. 20.E.IA Discuss in what sense the term r defined after the proof of Proposition 20.E.1 can be interpreted as the rate of interest implicit in the proportional price sequence. 20.E.2" Suppose that the production set Y c RL is of the constant return type and consider production paths that are proportiollal (but not necessarily stationary), that is, paths (Yo .. ..• .1'..... ) that satisfy .1', = (I + II)Y,., for aliI and some II. (a) Argue that the conclusion of Proposition 20.E.1 remains valid for proportional paths. (b) State and prove the result parallel to Proposition 20.E.2 for proportional paths. 20.E.3" Suppose that in the Ramsey-Solow model k solves Max (F(k, I) - k) (see Figure 20.E.2). Show that if k, S k - • for aliI, then the path determined by (k o,' .. , k" ... ) is efficient. [Hilll: Compute prices and verify the transversality condition.] 20.EAA Prove the three neoclassical properties stated at the end of the regular type part of Section 20.E. 20.E.S A Carry out the requested verification of expression (20.E. I).
E X E R CIS E S
785
------------------------------------------------------------------20.E.6
A
Carry out the verification requested in the discussion of Figure 20.E.3. A 20.E.7 In the Ramsey-Solow model. two dilTerent steady states are associated with different rates of interest. This is not so in the example illustrated in Figure 20.E.3, at first sight very similar. The key difference is that in the Ramsey-Solow model the consumption and investment goods are perfect substitutes in production. Clarify this by proving, in the context of the example underlying Figure 20.E.3, that if the two goods are perfect substitutes then r(k) # r(k) whenever k # i:. [Hilll: Their being perfect substitutes means that G(k, k' + a) = G(k, k') - a for any a < F(k. k'j.] A 20.E.8 Consider the proportional production paths with rate of growth equal to II > 0 (recall Exercise 20.E.2) in the context of a Ramsey-Solow technology of constant returns. Show that among these paths the one that maximizes surplus (at I = I, or, equivalently, normalized surplus or surplus" per capita") is characterized by having the rate of interest equal to II. This path is also called the lIoldell rule sleady slale palh. A 20.E.9 Argue that, for the one-Consumer model of Section 20.0, the golden rule path cannot arise as part of a competitive equilibrium. [Hint: The key fact is that J < I.] c 20.F.l Consider two arbitrary functions y,(w) and y,(w) that are defined for w > 0, take nonnegalive values, and satisfy y,(w) + y,(w) = w for all w. Suppose also that they are twice continuously ditTerentiable. Show thaI for any a > 0 there is a utilily function for lwo commodities. u(x" x,), that is increasing and concave on the domain {(x"x,): x, + x, S a} and is such that (y,(w), y,(w)) coincides wilh lhe Engel curve functions for prices p, = I, p, = I and wealth w < a. [Hinl: Let u(x" x,) ~ (x, + x,)'/2 - £[(x, - y,(x, + x,))' + (x, - y,(x, + x,))'] and take £ to be small enough. Verify then that Vu(x" x,) is strictly positive and D'u(x" x,) is negative definite for any (x,. x,) such that 0 < x, + x, SiX. and that the Engel Curve is as required.] 20.F.2A Suppose lhat. for k E R., the policy function
20.C.2A Consider an exchange model wilh two consumers. Utilily functions are of the form (20.8. I) and bOlh consumers have the same discount factor. There are no restrictions on the num ber of commodilies L or on the lata I endowmenls at any I. Show that at a Pareto optimal allocalion lhe following holds: for every consumer, the in-period marginal ulilities of wealth of lhe consumer is lhe same across periods (and equal to the overall marginal utility of wealth of lhe consumer). Interpret and discuss what this means in terms of intertemporal and inlerindividual lransfers of wealth.
786
CHAPTER
20:
EQUILIBRIUM
AND
TIME
-----------------------------------------------------------------------------
PAR T
F
I V E
20.G.3 8 The situation is the same as that of Exercise 20.0.2 (0) Parametrize the Pareto frontier of the utility possibility set by the ratio of marginal utilities of wealth of the two consumers.
(b) Then express the equations of equilibrium Ii la Negishi (see Appendix A of Chapter 17). That is, write down one equation in one unknown (the ratio of marginal utilities of wealth) whose zeros are precisely the equilibria of the model. (c) Argue in terms of the methodology discussed in Section 17.D that generically there are only a finite number of equilibria. Be as precise as you can.
Welfare Economics and Incentives
20.G.4A Prove the claim made in footnote 32. Be explicit about the form of the equilibrium price seq uences. 20.G.S" Verify that the concavity of the utility function implies that the expression (20.0.6) is larger than one in absolute value if there is no externality (i.e .• if Vi,u(') = V:,u(·) = 0). 20.H.I" Show that in the context of Sections 20.0 or 20.0 (a finite number of consumers) no bubbles can arise at equilibrium. 20.H.28 In the framework of Section 20.H do the following (diagrammatic proofs are permissible). (s) Show that if condition (20.H.3) is satisfied then. in the real asset case. the steady state is the only equilibrium. (b) Show that if condition (20.H.3) is satisfied then. in the purely nominal asset case, the monetary steady state is the only equilibrium that is a Pareto optimum. (c) Conversely. suppose that condition (20.H.3) is violated with strict inequality at p, = P.' Show then that. for the purely nominal asset case, there is more than one Pareto optimal equilibrium.
v«,)
+ ~v«.). Investigate (d) (Harder) Suppose that the utility function is of the form which conditions on v(') and ~ imply that the excess demand function satisfies condition (20.H.3). [Him: Recall Example 17.F.2 for a special case.] 20.1.1 A Verify the computation requested in the part of Section 20.1 headed "The (short·run) law of demand in overlapping generations models."
Part V is devoted to a systematic presentation of a number of issues related to the foundations of welfare economics. a topic that we have encountered repeatedly throughout the text. The point of view is that of a policy maker engaged in the design and implementation of collective decisions. In Chapter 21. we review classical social choice theory. The central question of this theory concerns the possibility of deriving the objectives of the policy maker as an aggregation of the preferences of the agents in the economy. and of doing so in a manner that could be deemed as satisfactory according to a number of desiderata. The difficulties of accomplishing this task are dramatically illustrated in Arrow's impossihiliry rheorem. which we present and discuss in detail. On a more positive note. we also discuss the assumption of single-peaked preferences and analyze the performance of majority voting under it. In Chapter 22, we admit, to a variety of extents. the possibility of explicit value judgments as to the comparability of individuals' uti!:!y levels. Most of the chapter is dcvoted to a presentation of welfare economics in the Bergson-Samuelson tradition. Towards this end. we develop the apparatus of utility possibility sets and social welfare functions and emphasize the distinction between first-best and second-best problems. We also offer an account of axiomatic bargaining theory. an approach that emphasizes the compromise, rather than the constrained optimality. nature of social decisions. In Chapter 23. we recognize that, in actuality. a policy maker rarely knows individuals' preferences with certainty; rather, this information is typically only privately known by the individuals themselves. The presence of information that is observed privately by rational, self-interested actors generates severe incentivecompatibility. second-best constraints. We offer a detailed analysis of what can and cannot be implemented collectively when these incentive constraints are taken into consideration. The content of Chapter 23 links with the game theory covered on Part II and revisits a number of themes first broached in Part III.
787
C HAP T E R
Social Choice Theory
21
21.A Introduction In this chapter, we analyze the extent to which individual preferences can be aggregated into social preferences, or more directly into social decisions, in a "satisfactory" manner-that is, in a manner compatible with the fulfillment of a variety of desirable conditions. Throughout the chapter, we contemplate a set of possible social alternatives and a population of individuals with well-defined preferences over these alternatives. In Section 21.B, we start with the simplest case: that in which the set of alternatives has only two elements. There are then many satisfactory solutions to the aggregation problem. I n our presentation, we focus on a detailed analysis of the properties of aggregation by means of majority voting. In Section 21.C, we move to the case of many alternatives and the discussion takes a decidedly negative turn. We state and prove the celebrated Arrow's impossibility theorem. In essence, this theorem tells us that we cannot have everything: If we want our aggregation rule (which we call a social welfare functional) to be defined for any possible constellation of individual preferences, to always yield Pareto optimal decisions, and to satisfy the convenient, and key, property that social preferences over any two alternatives depend only on individual preferences over these alternatives (the pairwise independence condition), then we have a dilemma. Either we must give up the hope that social preferences could be rational in the sense introduced in Chapter 1 (i.e., that society behaves as an individual would) or we must accept dictatorship. Section 21.D describes two ways out of the conclusion of the impossibility theorem. In one we allow for partial relaxations of the degree of rationality demanded of social preferences. In the other, we settle for aggregation rules that perform satisfactorily on restricted domains of individual preferences. In particular, we introduce the important notion of single-peaked preferences and, for populations with preferences in this class, we analyze the role of a median voter in the workings of pairwise majority voting as an aggregation method. Section 21.E sets the aggregation problem more directly as one of aggregating individual preferences into social decisions. It introduces the concept of a social choice 789
790
CHAPTER
2.:
SOCIAL
CHOICE
THEORY
--
S E C T ,ON
------------------------------------------------------------------~-
jllnction, and proceeds to give a version of the impossibility result for the latter. Essentially, this result is obtained by replacing the pairwise independence condition (which is meaningless in the context of this section) by a mono tonicity condition on the social choice function. This condition provides an important link to the incentive-based theory of Chapter 23. General references and surveys for the topics of this chapter are Arrow (1963), Moulin (1988), and Sen (1970) and (1986).
21.B A Special Case: Social Preferences over Two Alternatives We begin our analysis of social choice by considering the simplest possible case: that in which there are only two alternatives over which to decide. We call these alternative x and alternative y. Alternative x, for example, could be the "status-quo," and alternative y might be a particular public project whose implementation is being contemplated. The data for our problem are the individual preferences of the members of society over the two alternatives. We assume that there is a number I < 00 of individuals, or agents. The family of individual preferences between the two alternatives can be described by a profile t
("'I>' .. , "'I) E R ,
where "'i takes the value 1,0, or -I according to whether agent i prefers alternative x to alternative y, is indifferent between them, or prefers alternative y to alternative x, respectively.' Definition 21.B.1: A social welfare functional (or social welfare aggregator) is a rule F(~" ... , "',) that assigns a social preference, that is,F(a., ... , "',) E {-1, O. 1}. to every possible profile of individual preferences (IX ••...• a,) E { -1. O. 1}'. All the social welfare functionals to be considered respect individual preferences in the weak sense of Definition 2I.B.2. Definition 21.B.2: The social welfare functional F(", ••.. .• "',) is Paretian, or has the Pareto property. if it respects unanimity of strict preference on the part of the agents. that is. if F(l • ...• 1) = 1 and F( -1 •...• -1) = - I. Example 2I.B.I: Paretian social welfare functionals between two alternatives abound. Let (fJ, •...• fJ,) E IR'+ be a vector of nonnegative numbers, not all zero. Then we
2 •• 8:
ASP Eel A l e A S E:
SOC I ALP REF ERE NeE SOY E R
TWO
could define F(IX" ... , ",,)
= sign 2., Pi(X',
where. recall, for any a E R, sign a equals I, 0, or -I according to whether a > 0,
a = 0, or a < 0, respectively. An important particular case is majority voting, where we take p, = I for every i. Then F(IX, •... , ",,) = I if and only if the number of agents that prefer alternative x to alternative y is larger than the number of agents that prefer y to x. Similarly, F(a, •...• a,) = -I if and only if those that prefer y to x are more numerous than those that prefer x to y. Finally, in case of equality of these two numbers, we have F(IX" ...• IX,) = 0, that is. social indifference. a Example 2I.B.2: DictatorslJip. We say that a social welfare functional is dictatorial if there is an agent h, called a dictator, such that, for any profile ("'I" .. ,"',), "'. = I implies F(cx" ... , "',j = I and, similarly, "'. = -I implies F("" , •.• , "',j = -I. That is. the strict preference of the dictator prevails as the social preference. A dictatorial social welfare functional is Paretian in the sense of Definition 2I.B.2. For the social welfare functionals of Example 21.B.1, we have dictatorship whenever "'. > 0 for some agent It and "'i = for i ¥ It. since then F(a" ... , a,) = "' •. a
°
The majority voting social welfare functional plays a leading benchmark role in social choice theory. In addition to being Paretian it has three important properties, which we proceed to state formally. The first (symmetry among agents) says that the social welfare functional treats all agents on the same footing. The second (neutrality between alternatives) says that, similarly, the social welfare functional does not a priori distinguish either of the two alternatives. The third (positive responsiveness) says, more strongly than the Paretian property of Definition 2I.B.2, that the social welfare functional is sensitive to individual preferences. Definition 21.B.3: The social welfare functional F("' ...... "',) is symmetric among agents (or anonymous) if the names of the agents do not matter. that is. if a permutation of preferences across agents does not alter the social preference. Precisely. let 1I:{1, ... . /} -+ {1 ....• 1} be an onto function (i.e., a function with the property that for any i there is h such that 1I(h) = i). Then for any profile (a, •...• IX,) we have F(cx •. ... ,,,,,) = F(IX'I')'" . ,a,I'I)' Definition 21.B.4: The social welfare functional F("', •.... :x,) is neutral between alternatives if F(a, •...• a,) = -F(-IX, •... , -:x,) for every profile (:x, •...• a,). that is. if the social preference is reversed when we reverse the preferences of all agents.
rormally the principle involved. Note. in particular. that this specification precludes the usc of any
Definition 21.B.S: The social welfare functional F(~, •. ..• :x,) is positively responsive if. whenever (IX, •...• 7,) ~ (ex; .... • ~;). ("', ....• IX,) ¥ ("'; .... , IX;). and F(",; • ...• cx;) ~ O. we have F(a, • ...• (1.,) = + 1. That is. if x is socially preferred or indifferent to y and some agents raise their consideration of x. then x becomes socially preferred.
"cardinal" or "intensity" information between the two alternatives because this intensity can only be calibrated (perhaps using lotteries) by appealing to some third alternative. A rortiori, the specification also precludes the comparison of feelings of pleasure or pain across individuals. In Chapter 22, we discuss in some detail matters pertaining to the issue of interpersonal comparability of utilities.
It is simple to verify that majority voting satisfies the three properties of symmetry among agents, neutrality between alternatives, and positive responsiveness (sec Exercise 21.B.I). As it turns out, these properties entirely characterize majority voting. The result given in Proposition 21.B.1 is due to May (1952).
I. In the whole of this chapter we make the restriction that only the agents' rankings between the two alternatives matter for the social decision between them. In Section 21.C we will state
A L T ERN AT, V E S
791
792
CHAPTER
.,:
SOCIAL
CHOICE
THEORY
Proposition 21.B.1: (May's Theorem) A social welfare functional F(IX" ... ,lXd is a majority voting social welfare functional if and only If it is symmetric among agents, neutral between alternatives, and positive responsive. Proof: We have already argued that majority voting satisfies the three properties. To establish sufficiency note first that the symmetry property among agents means that the social preference depends only on the total number of agents that prefer alternative x to y, the total number that are indifferent, and the total number that prefer y to x. Given (IX., ... ,IX,), denote n+(IX., ... ,IX,) = #(i:IX,= I),andn-(IX., ... ,IX,)= #(i:IX,= _1).2 Then symmetry among agents allows us to express F(IX., ... , IX,) in the form F(IX., ... , IX,) = G(n+(IX., ... , IX,), n-(IX., ... , IX,». Now suppose that (IX., ... ,IX,) is such that n+(IX ...... IX,) = n-(IX., ... ,IX,). Then 11+( -IX I, ... , -IX,) = n-(IX., ... , IX,) = n+(IX., ... , IX,) = n-( -IX ..... , -IX,), and so
--
SECTION
".C:
THE
GENERAL
CASE:
ARROW'S
IMPOSSIBILITY
there are I agents, indexed by i = I, ...• J. Every agent i has a rational preference relation ::::, defined on X. The strict preference and the indifference relation derived from ::::, are denoted by >-, and -" respectively.3 In addition, it will often be convenient to assume that no two distinct alternatives are indifferent in an individual preference relation ::::,. It is therefore important, for clarity of exposition, to have a symbol for the set of all possible rational preference relations on X and for the set of all possi ble preference relations on X having the property that no two distinct alternatives are indifferent. We denote these sets, respectively, by fit and 9. Observe that 9 c: .'11." In parallel to Section 21.B, we can define a social welfare functional as a rule that assigns social preferences to profiles of individual preferences (::::" ... , ::::,) E fit'. Definition 21.C.1 below generalizes Definition 21.B.I in two respects: it allows for any number of alternatives and it permits the aggregation problem to be limited to some given domain .<1 c: .!Jt' of individual profiles. In this section, however, we focus on the largest domains, that is, .<1 = fit' and .
F(IX., ... , IX,) = G(n+(IX., ... , IX,), n-(IX., ... , IX,»
= G(n+(-IX., ... , -IX,),n-(-IX., ... , -IX,» = F(-IX., ... , -IX,) = -F(IX., ... ,IX,).
The last equality follows from the neutrality between alternatives. Since the only number that equals its negative is zero, we conclude that if n+(IX., ... , IX,) = n-(IX., ... , IX,) then F(IX., ... , IX,) = O. Suppose next that n+(IX ...... IX,) > n-(IX ..... ,IX,). Denote H = n+(IX., ... , IX,), J = n-(IX., ... , IX,); then J < H. Say, without loss of generality, thatlXl = I for iSH and IX, s; 0 for i> H. Consider a new profile (IX;, ... , lXi) defined by lXi = IXI = I for is; J < H, IX; = 0 for J < iSH. and IXi = IX, S 0 for i > H. Then n+(IX; .... • lXi) = J and n-(IX;, ... • lXi) = n-(IX ..... , IX,) = J. Hence F(IX;, ... , lXi) = O. But by construction, the alternative x has lost strength in the new individual preference. Indeed, (IXI'''''IX,)~ (1X·...... lXi) and IXJ+. = I >O=IX;+ •. Therefore, by the positive responsiveness property, we must have F(IX., ... ,IX,) = I. In turn, if n-(IX., ... ,IX,) > n+(IX ..... ,IX,) then n+(-IX ..... ,-IX,» n-( -(X., ... , -IX,) and so F( -IX ..... , -IX,) = I. Therefore, by neutrality among alternatives: F(IX., ... ,IX,)= -F(-IX., ... , -IX,) = -I. We conclude that F(IX., ... , IX,) is indeed a majority voting social welfare functional. _ In Exercise 2I.B.2, you are asked to find examples dilTerent from majority voting that satisfy any two of the three properties of Proposition 21.8.1.
2LC The General Case: Arrow's Impossibility Theorem We now proceed to study the problem of aggregating individual preferences over any number of alternatives. We denote the set of alternatives by X, and assume that
Definition 21.C.1: A social welfare functional (or social welfare aggregator) defined on a given subset .<1 c: iJt' is a rule F:.
3. Recall from Sec. ion I.B .hal >-, is formally defined by lelling x >-, y if x <:, y holds bUI Y?:i x docs not. That is. x is preferred to y if x is at least as good as y but y is not as good as x. Also. the indifTercncc relation -j is defmed by letting x -j y if x 2:j)' and y 2:jx. From Proposition I.B.I we know that if 2:j is rational, that is, complete and transitive, then >j is irrcnexive (x >jX cannot occur) and transitive (x >j'y and y >jX implies x >iX), Similarly, -j is fI:nexive (x -jX for all x EX). transitive (x - j y and)' -jZ implies x -,x) and symmetric (x - j y implies Y - j x). 4. Formally, the preference relation ~j belongs to ~ if it is reflexive (x ~I x for every x EX). transitive (x ;::i y and )' ~j z implies x ~I x) and IOtal (if x "# y then either x ~i y or y ~i x, but not
bOlh). Such preference relations are often referred to as slricl preferences (although Slricl-lolal preferem'('''' would be less ambiguous) or even as linear orders, because these are the properties of the usual "larger than or equal to" order in the real line. 5. In particular, there are no individual utility levels and, therefore. there is no meaningful sense
in which any conceivable information on individual utility levels could be compared and matched up. We refer again to Chapter 22 (especially Section 22.D) for an analysis of the problem that 2. Recall the notation # A = cardinality of the set A = number of clemen Is in the set A.
focuses on the information used in the aggregation process.
THEOREM
793
794
C HAP T E R
2,:
• 0 CI AL
C HOI C E
THE 0 R Y
----------------------------------------------------------------------~
Definition 21.C.2: The social welfare functional F:sI -+ fJl is Paretian if. for any pair of alternatives {x. y} c X and any preference profile (1::;, •...• 1::;,) e sI. we have that x is socially preferred to y. that Is. x Fp(I::;, •...• 1::;,) y. whenever x >-1 y for every i. In Example 2l.C.l we describe an interesting class of Paretian social welfare functionals. Example 21.C.l: The Borda Count. Suppose that the number of alternatives is finite. Given a preference relation 1::;, e fJl we assign a number of points c,(x) to every alternative x e X as follows. Suppose for a moment that in the preference relation 1::;, no two alternatives are indifferent. Then we put c,(x) = n if x is the nth ranked alternative in the ordering of 1::;,. If indifference is possible in 1::;, then c,(x) is the average rank of the alternatives indifferent to x. 6 Finally. for any profile (1::;"" .• 1::;,) E fJl' we determine a social ordering by adding up points. That is. we let F(I::;, •...• I::;,)efJl be the preference relation defined by xF(I::; ...... I::;,)y if L, c,(x) ~ L, c,(y). This preference relation is complete and transitive [it is represented by the utility function -c(x) = - L,C,(X)]. Moreover. it is Paretian since if x >-, y for every i then c,(x) < c,(y) for every i. and so L, c,(x) < L, c,(y) . • We next state an important restriction on social welfare functionals first suggested by Arrow (1963). The restriction says that the social preferences between any two alternatives depend only on the individual preferences between the same two alternatives. There are three possible lines of justification for this assumption. The first is strictly normative and has considerable appeal: it argues that in settling on a social ranking between x and y. the presence or absence of alternatives other than x and y should not matter. They arc irrelevant to the issue at hand. The second is one of practicality. The assumption enormously facilitates the task of making social decisions because it helps to separate problems. The determination of the social ranking on a subset of alternatives does not need any information on individual preferences over alternatives outside this subset. The third relates to incentives and belongs to the subject matter of Chapter 23 (see also Proposition 2I.E.2). Pairwise independence is intimately connected with the issue of providing the right inducements for the truthful revelation of individual preferences. Definition 21.C.3: The social welfare functional F:sI .... tJt defined on the domain sI satiSfies the pairwise independence condition (or the independence of irrelevant alternatives condition) if the social preference between any two alternatives {x. y} c X depends only on the profile of individual preferences over the same alternatives. Formally'. for any pair of alternatives {x. y} eX. and for any pair of preference profiles (1::;, ....• 1::;,) E d and (1::;., •...• 1::;;) E d with the property that. for every i. Xl::;iY <=> xl::;;y and yl::;iX <=> yl::;;x. 6. Thus if X = {x. y. z} and x~, y -, z then e,(x) = 1. and e,(y) = e,(z) = 2.5. 7. The expressions Ihat follow are a bil cumbersome. We emphasize Iherefore thaI Ihey do nothing more than to capture formally the statement just made. An equivalent formulation would be: for any {x.y} c: X. if ~,!{x.y} = ~;!{x.y} for all i. then F(~, ..... ~r)!{x.y} = F(;:'" ... , ;:;)1 {x. y}. Here;: !{x, y} 51ands for Ihe restriction of Ihe preference ordering;: 10 Ibe set {x, y}.
SECT'ON
'1.C:
THE
GENERAL
CASE:
ARROW'S
'MPOSSIB'L'TY
THEOREM
795
--------------------------------------------------------------------------we have that xF(I::;, •...• 1::;,) y
xF(I::;; ..... 1::;;) y
yF(~, •...• ~,)
yF(~·, •...• I::;;)x.
and
x
Example 21.C.l: conrinued. Alas. the Borda count does not satisfy the pairwise independence condition. The reason is simple: the rank of an alternative depends on the placement of every other alternative. Suppose. for example. that there are two agents and three alternatives {x. y. z}. For the preferences x >-, z >-, y.
y>-,x>-,z we have that x is socially preferred to y [indeed. e(x) = 3 and ely) = 4]. But for the preferences
x >-; y >-', z. y>-,z>-,x we have that .v is socially preferred to x [indeed. now c(x) = 4 and ely) = 3]. Yet the relative ordering of x and y has not changed for either of the two agents. For another illustration, this time with three agents and four alternatives {x. y. t. w}. consider
z >-, X
>-, J' >-, lV.
z >-,x >-,y>-, w. y
>- , z >-, w >- J x.
Here. y is socially preferred to x [c(x) = 8 and ely) = 7]. But suppose now that alternatives z and w move to the bottom for all agents (which because of the Pareto property is a way of saying that the two alternatives are eliminated from the alternative set): x >-', y >-', z >-', IV.
>-2 y >-, z >-', w. y >-3X >-3 z >-', IV. x
(21.C.l)
Then x is socially preferred to y [e(x) = 4. c(y) = 5]. Thus the presence or absence of alternatives z and w matters to the social preference between x and y. Another modification would take alternative x to the bottom for agent 3:
x >-;}' >-; z >-';
lV,
>-:;y >-:;z >-:; lI', }' >-; z >-; IV >-; x. X
Now .r is socially preferred to x [which. relative to the outcome with (21.C.l). is a nice result from the point of view of agent 3]. • The previous discussion of Example 21.C.l teaches us that the pairwise independence condition is a substantial restriction. However. there is a way to proceed that will automatically guarantee that it is satisfied. It consists of determining the social prefcrence between any given two alternatives by applying an aggregation rule that uses only the information about the ordering of rhese two alternatives in
796
CHAPTER
21:
SOCIAL
CHOICE
THEORY
individual preferences. We saw in Section 21.B that, for any pair of alternatives, there are many such rules. Can we proceed in this pairwise fashion and still end up with social preferences that are rational. that is. complete and transitive? Example 21.C.2 shows that this turns out to be a real difficulty. Example 21.C.2: The Condorcet Paradox." Suppose that we were to try majority voting among any two alternatives (see Section 21.B for an analysis of majority voting). Does this determine a social welfare functional? We shall see in the next section that the answer is positive in some restricted domains d c !Jtl • But in general we run into the following problem. known as the Condorcet paradox. Let us have three alternatives {x. y. z} and three agents. The preferences of the three agents are
x >-. y >-1 z.
----
--
SECTION
21.C:
THE
GENERAL
CASE:
ARROW'S
IMPOSSIBILITY
Definition 21_C.4: Given F('), we say that a subset of agents ScI is: (i) Decisive for x over y if whenever every agent in S prefers x to y and every agent not in S prefers y to x, x is socially preferred to y. (ii) Decisive if, for any pair {x, y} eX, S is decisive for x over y. (iii) Completely decisive for x over y if whenever every agent in S prefers x to y, x is socially preferred to y. The proof will proceed by a detailed investigation of the structure of the family of decisive sets. We do this in a number of small steps. Steps 1 to 3 show that if a subset of agents is decisive for some pair of alternatives then it is decisive for all pairs. Steps 4 to 6 establish some algebraic properties of the family of decisive sets. Steps 7 and 8 use these to show that there is a smallest decisive set formed by a single agent. Steps 9 and 10 prove that this agent is a dictator.
Z>-2X>-2Y. Y >- 3Z>- 3x.
Then pairwise majority voting tells us that x must be socially preferred to Y (since x has a majority against y and. a fortiori. y does not have a majority against x). Similarly. y must be socially preferred to z (two voters prefer y to z) and z must be socially preferred to x (two voters prefer z to x). But this cyclic pattern violates the transitivity requirement on social preferences. _ The next proposition is Arrow's impossibility theorem. the central result of this chapter. It essentially tells us that the Condorcet paradox is not due to any of the strong properties of majority voting (which. we may recall from Proposition 2I.B.l. are symmetry among agents, neutrality between alternatives, and positive responsiveness). The paradox goes to the heart of the matter: with pairwise independence there is no social welfare functonal defined on !Jtl that satisfies a minimal form of symmetry among agents (no dictatorship) and a minimal form of positive responsiveness (the Pareto property). Proposition 21.C.1: (Arrow's Impossibility Theorem) Suppose that the number of alternatives is at least three and that the domain of admissible individual profiles. denoted d, is either d = 91 1 or d = iJl l • Then every social welfare functional F:.flI -+ !Jt that is Paretian and satisfies the pairwise independence condition is dictatorial in the following sense: There is an agent h such that, for any {x, y} eX and any profile (;::; I' . . . , ;::;/) ed, we have that x is socially preferred to y. that is, x Fp (;::;, •••. , ;::;d y, whenever x >-hY' Proof: We present here the classical proof of this result. For another approach to the demonstration we refer to Section 22.D. It is convenient from now on to view I not only as the number but also as the set of agents. For the entire proof we refer to a fixed social welfare functional F: d -+ !Jt satisfying the Pareto and the pairwise independence conditions. We begin with some definitions. In what follows, when we refer to pairs of alternatives we always mean distinct alternatives.
8. This example was already discussed in Section 1.8.
Step I: If for some {x. y} c X. ScI is decisive for x over y. then. for any alternative z # x. S is decisive for x over z. Simiiarly,Jor any z # y. S is decisive for z over y. We show that if S is decisive for x over y then it is decisive for x over any z # x. The reasoning for Z over y is identical (you are asked to carry it out in Exercise 21.C.l). If z = y there is nothing to prove. So we assume that z # y. Consider a profile of preferences (;::; ••...• ;::;1) e d where
for every i e S and for every i e I \S. Then. because S is decisive for x over y. we have that x is socially preferred to y, that is. xF,(;::; ••...• ;::;/)Y' In addition. since y;::;/z for every iel. and F(') satisfies the Pareto property it follov.s that y F,(;::; ...... ;::;1) z. Therefore. by the transitivity of the social preference relation. we conclude that x F,<;::; ••... , ;::;1) z. By the pairwise independence condition. it follows that x is socially preferred to z whenever every agent in S prefers x to z and every agent not in S prefers z to x. That is. S is decisive for x over z. Step 2: If for some {x. y} c X, ScI is decisive for x over y and z is a third alternative, then S is decisive for z over w andfor w over z. where w e X is any alternative distinct from z. By step I. S is decisive for x over z and for z over y. But then. applying step 1 again, this time to the pair {x. z} and the alternative w. we conclude that S is decisive for w over z. Similarly, applying step 1 to {z. y} and w. we conclude that S is decisive for z over w. Step 3: If for some {x. y} c X. ScI is decisive for x over y. then S is decisive. This is an immediate consequence of step 2 and the fact that there is some alternative z E X distinct from x or y. Indeed. take any pair {v. w}. If v = z or w = z. then step 2 implies the result directly. If v # z and w # z. we apply step 2 to conclude that S is decisive for z over w. and then step 1 [applied to the pair {z. w}] to conclude that S is decisive for v over w.
THEOREM
797
798
C HAP T E R
2 1:
SOC I ALe HOI C E
THE 0 R Y
--------------------------------------------------------------~ Step 4: If ScI and Tel are decisive, then S 1"1 T is decisive. Take any triple of distinct alternatives {x, y, %} c X and consider a profile of preferences (;::" ... , ;::;/) E d where
for every i E S\(S
1"1
x >-,% >-,y
for every i E S
y >-,x >-i%
for every i E T\(S
y
>-i% >-,x
1"1
T),
T, 1"1
T),
for every i E I\(S u T).
Then z F.(;::; ..... , ;::;/) y because S (=[S\(S 1"1 T)] u (S 1"1 T» is a decisive set. Similarly, x F.(;::; .. ... , ;::;/) z because T is a decisive set. Therefore, by the transitivity of the social preference, we have that x F.(;::;" ... , ~tl y. It follows by the pairwise independence condition that S 1"1 T is decisive for x over y, and so, by step 3, that S ("\ T is a decisive set. Slep 5: For any ScI, we have that either S or its complement,I\S c I, is decisive. Take any triple of distinct alternatives {x, y, z} c X and consider a profile of preferences (;::;" ... , ;::;/) E d where
x >-,% >-,y
for every i E S
Y>-iX >-i%
for every i E I\S.
Then there are two possibilities: either x F,(;::;" ... , ~/) y, in which case, by the pairwise independence condition, S is decisive for x over y (hence, by step 3, decisive), or y F(;::; ..... , ;::;/) x. Because, by the Paretian condition, we have x F,(~" ... , ;::;/) z, the transitivity of the social preference relation yields that y F,(;::; ..... , ;::;/) z in this case. But then, using the pairwise independence condition again, we conclude that I \S is decisive for y over z (hence, by step 3, decisive). Slep 6: If ScI is decisive and SeT, then T is also decisive. Because of the Pareto property the empty set of agents cannot be decisive (indeed, if no agent prefers x over y and every agent prefers y over x, then x is not socially preferred to y). Therefore 1\ T cannot be decisive because otherwise, by step 4, S ("\ (l \ T) = 0 would be decisive. Hence, by step 5, T is decisive. Step 7: If ScI is decisive and it includes more than one agent, then there is a slrict subset S' c S, S' ,;. S, such that S' is decisive. Take any he S. If S\{h} is decisive, then we are done. Suppose, therefore, that S\{It} is not decisive. Then, by step 5,I\(S\{h}) = (I\S) u {It} is decisive. It follows, by step 4, that {hI = S ("\ [(I\S) u {h}] is also decisive. Thus, we are again done since, by assumption, {h} is a strict subset of S. Step 8: There is an hE I such that S = {hI is decisive. This follows by iterating step 7 and taking into account, first, that the set of agents I is finite and, second, that, by the Pareto property, the set I of all agents is decisive. Step 9:
If ScI is decisive Ihen, for any {x, y} c X, S is completely decisive for
x over y.
We want to prove that, for any T c I\S, x is socially preferred to y whenever every agent in S prefers x to y, every agent in T regards x to be at least as good as
--
SECTION
21.0:
SOME
POSSIBILITY
RESULTS:
REITRICTED
y, and every other agent prefers y to x. To verify this property, take a third alternative Z E X, distinct from x and y. By the pairwise independence condition it suffices to consider a profile of preferences (;::;" ... , ;::;/) E sf where for every i E S, x >-,z >-,y x>-,y>-,z for every i E T,
y>-,Z >-,x
for every i
E
I\(S u T).
Then x F,(;::;" . .. , ;::;,) z because, by step 6, S v Tis decisive, and % F,(;::;I' . .. , ;::;,) y because S is decisive. Therefore, by the transitivity of social preference, we have that x F,(;::;" ... , ;::;,) y, as we wanted to show. Slep 10: If, for some he I, S = {hI is decisive, then h is a dictator. If {It} is decisive then, by step 9, {h} is completely decisive for any x over any y. That is, if the profile (;::;., ... , ;::;,) is such that x >-.y, then x F,(;::;., ... , ;::;,) y. But this is precisely what is meant by h E I being a dictator.
The combination of steps 8 and 10 completes the proof of Proposition 21.C.1. •
21.D Some Possibility Results: Restricted Domains The result of Arrow's impossibility theorem is somewhat disturbing, but it would be too facile to conclude from it that "democracy is impossible." What it shows is something else-that we should not expect a collectivity of individuals to behave with the kind of coherence that we may hope from an individual. It is important to observe, however, that in practice collective judgments are made and decisions are taken. What Arrow's theorem docs tell us, in essence, is that the institutional detail and procedures of the political process cannot be neglected. Suppose, for example, that the decision among three alternatives {x, y, %} is made by first choosing between x and y by majority voting, and then voting again to choose between the winner and the third alternative %. This will produce an outcome, but the outcome may depend on how the agenda is set-that is, on which alternative is taken up first and which is left for the last. [Thus, if preferences are as in the Condorcet paradox (Example 21.C.2) then the last alternative, whichever it is, will always be the survivor.] This relevance of procedures and rules to social aggregation has far-reaching implications. They have been taken up and much emphasized in modern political science; see, for example, Austen-Smith and Banks (1996) or Shepsle and Boncheck (1995). In this section, we remain modest and retain the basic framework. We explore to what extent we can escape the dictatorship conclusion if we relax some of the demands imposed by Arrow's theorem. We will investigate two weakenings. In the first, we relax the rationality requirements made on aggregate preferences. In the second, we pose the aggregation question in a restricted domain. In particular, we will consider a restriction-single-peakedlless-that has been found to be significant and useful in applications.
Less Than Full Social Rationality Suppose that we keep the Paretian and pairwise independence conditions but permit the social preferences to be less than fully rational. Two weakenings of the rationality of preferences are captured in Definition 21.0.1.
DOMAINS
799
800
CHAPTER
21:
SOCIAL
CHOICE
THEORY
SEC T ION
Definition 21.0.1: Suppose that the preference relation;:: on X is reflexive and complete. We say then that: (i) ;:: is quasitransitive if the strict preference >- induced by;:: (i.e. x >-y = x;:: Y but not y ;:: x) is transitive. (ii) ;:: is acyclic if ;:: has a maximal element in every finite subset X' c X, that is, {xEX':x;::yfor all VEX'} '" 0. A quasi transitive preference relation is acyclic, but the converse may not hold. Also, a rational preference relation is quasitransitive, but, again, the converse may not hold" Thus the weaker condition is acyclicity. Yet acyclicity is not a drastic weakening of rationality: Note, for example, that the social orderings of the Condorcet paradox (Example 21.C.2) also violate acyclicity. (For more on acylicity see Exercise 21.0.1.) We will not discuss in detail the possibilities opened to us by these weakenings of social rationality. There are some but they are not very substantial. We refer to Sen (1970) for a detailed exposition. The next two examples are illustrative. Example 21.0.1: Oligarchy. Let I be the set of agents, and let Sci be a given subset of agents to be called an oligarchy (the possibilities S = {II} or S = I are permitted). Given any profile (;::., ... , ;::1) E ",I, the social preferences are formed as follows: For any x, y E X, we say that x is socially at least as good as y [written x F(;::., ... , ;:: I) y] if there is at least one h E S that has x ;::. y. Hence, x is socially preferred to y if and only if every member of the oligarchy prefers x to y. In Exercise 21.0.2 you should verify that this social preference relation is quasitransitive but not transitive (because social indifference fails to be transitive). This is the only condition of Arrow's impossibility theorem that is violated (the Paretian condition and pairwise independence conditions are clearly satisfied). Nonetheless, this is scarcely a satisfactory solution to the social aggregation problem, as the aggregator has become very sluggish. At one extreme, if the oligarchy is a single agent then we have a dictatorship. At the other, if the oligarchy is the entire population then society is able to express strict preference only if there is complete unanimity among its members. _ Example 21.0.2: Vetoers. Suppose there are two agents and three alternatives
{x, y, z}. Then given any profile of preferences (;::1' ;::2)' we let the social preferences coincide with the preferences of agent I with one qualification: agent 2 can veto the possibility that alternative x be socially preferred to y. Specifically, if y >- 2 x then)' is socially at least as good as x. Summarizing. for any two alternatives
9. Suppose that?::: is quasitransitive. Assume for a moment that it is not acyclic. Then there is some finite set X' c X without a maximal element for ~. That is, for every.'( e X' there is some y EX' such Ihal y >- x (i.e., such Ihal y ~ x bUI nol x ~ y). Thus, for any integer M we can find a
chain
x' >- x' >-. ,,>- x". where x· E X' for every m =
I .... , M. If M is larger than the number
of alternatives in X', then there must be some repetition in this chain. Say that x·' = x'" for m > m', By quasi transitivity. x"" >- x'" = x"", which is impossible because >- is irrenexive by definition. Hence, ;:: must be acyclic. An example or an acyclic but not quasitransitive relation will be given
in Example 2I.D.2. The relation >- derived from a rational prererence relation ~ is transitive (Proposition I.B.J). An example of a quasitransitive. but not rational. prererence relation is given
in Example 2t.D.t.
2 1 • D:
S0 ME
P D S SIB I LIT Y
RES U L T S:
RES T RIC TED
{v, w} c {x, y, z} we have that v is socially at least as good as w if either V;::I w, or v = y, w = x and v >- 2 w. In Exercise 21.0.3 you should verify that the social preferences so defined are acyclic but not necessarily quasitransitive. _
Single-Peaked Preferences We proceed now to present the most important class of restricted domain conditions: single-peakedness. We will then see that, in this restricted domain, nondictatorial aggregation is possible. In fact, with a small qualification, we will see that on this domain pairwise majority voting gives rise on this domain to a social welfare functional. Definition 21.0.2: A binary relation
x
y
then
z >- y
and If Y
> z - y.
In words: There is an alternative x that represents a peak of satisfaction and, moreover, satisfaction increases as we approach this peak (so that, in particular, there cannot be any other peak of satisfaction). Example 21.0.4: Suppose that X = [a,b] c R and ~ is the "greater than or equal to" ordering of the real numbers. Then a continuous preference relation;:: on X is single peaked with respect to ~ if and only if it is strictly convex, that is, if and only if, for every WE X, we have ay + (I - a)z >- W whenever y;:: W, Z ;:: w, y '" z, and a E (0,1). (Recall Oefinition 3.B.5 and also that, as a matter of definition, preference relations generated from strictly quasiconcave utility functions are strictly convex.) This fact accounts to a large extent for the importance of single-peakedness in economic applications. The sufficiency of strict convexity is actually quite simple to verify. (You are asked to prove necessity in Exercise 21.0.4.) Indeed, suppose that x is a maximal element for;::, and that, say, x > z > y. Then x ;:: y, y ;:: y, x '" y, and Z = ax + (I - a)y for some a E (0, I). Thus, z >- y by strict convexity. In Figures 2I.D.I and 21.0.2, we depict utility functions for two preference relations on X = [0, I]. The preference relation in Figure 21.0.1 is single peaked with respect to ;::" but that in Figure 21.D.2 is nol. _ Definition 21.0.4: Given a linear order : c 9t the collection of all rational preference relations that are single peaked with respect to
D 0 M A INS
801
802
CHAPTER
21:
SOCIAL
CHOICE
THEORY
---------------------------------------------------------------Ftgure 21.D.1 (left)
Utility
Utility
Preferences are sing). peaked with resP
--
SECTION
21.D:
SOME
POSSIBILtTY
RESULTS:
RESTRICTED
DOMAINS
Utility
Figure 21.0.2 (right)
Preferences are not
single peaked with respect to ~.
"--_ _ _ _ _ _ _ _----' "(x)
o
I
restricted domain of preferences ar'". This amounts to the requirement that all individuals have single-peaked preferences with respect to the same linear order;;::. Suppose that on the domain ar~ we define social preferences by means of pairwise majority voting (as introduced in Example 21.8.1). That is, given a profile n::" ... , ::::t> e ar~ and any pair {x, y} c X, we put x i(::::" ... , ::::, )y, to be read as "x is socially at least as good as y", if the number of agents that strictly prefer x to y is larger or equal to the number of agents that strictly prefer y to x, that is, if #(ie/:x>-IY};;:: #{ie/:y>-,x}. Note that, from the definition, it follows that for any pair {x, y} we must have either x i(::::" ... , ::::,) y or y i(::::" ... , ::::,) x. Thus, pairwise majority voting induces a complete social preference relation (this holds on any possible domain of preferences). In Exercise 21.0.5 you are asked to show in a direct manner that the preferences of the Condorcet paradox (Example 2I.C.2) are not single peaked with respect to any possible linear order on the alternatives. In fact, they cannot be because, as we now show, with single-peaked preferences we are always assured that the social preferences induced by pairwise majority voting have maximal elements, that is, that there are alternatives that cannot be defeated by any other alternatives under majority voting. Let (::::,' ... , ::::,) e ar~ be a fixed profile of preferences. For every i e I we denote by XI e X the maximal alternative for ::::, (we will say that Xi is "i's peak"). Definition 21.0.5: Agent h e I is a median agent for the profile (::::1' ... , ::::,) e ar~ if #{ie/:x;;;::xh};;::2I
and
I # {'le/:xh;;::xi } ~2'
A median agent always exists. The determination of a median agent is illustrated in Figure 2I.D.3. If there are no ties in peaks and # I is odd, then Definition 21.D.5 simply says that a number (I - 1)/2 of the agents have peaks strictly smaller than x. and another number (I - I )/2 strictly larger. In this case the median agent is unique. Proposition 21.0.1: Suppose that ~ is a linear order on X and consider a profile of preferences (::::1"" , ::::,) where, for eve!y i, ::::; is single peaked with respect.to ;;::. Let h e I be a median agent. Then Xh F(:::: l' ••• , ::::,) y for every y e X. That IS, the peak xh of the median agent cannot be defeated by majority voting by any other alternative. Any alternative having this property is called a Condorcet winner. Therefore, a Condorcet winner exists whenever the preferences of all agents are singlepeaked with respect to the same linear order.
Flgur. 21.0.3
Agent 5 is the Median Agent 3 VOlers
3 VOlers
Proof: Take any y e X and suppose that x. > y (the argument is the same for y > x.). We need to show that y does not defeat x, that is, that #{ie/:x.>-,y};;:: #(ie/:y>-,x.}.
Consider the set of agents Sci that have peaks larger than or equal to x., that is, S = {i e I: XI ;;:: x.}. Then x, ~ x. > y for every i e S. Hence, by single-peakedness of :::: I with respect to ~, we get x. >-, y for every j e S. On the other hand, because agent h is a median agent we have that #S ~ 1/2 and so # {i e I: y >-,x.} ~ #(/\S) ~ 1/2~ #S:;, #{ie/:x.>-,y}._ Proposition 21.0.1 guarantees that the preference relation i(::::" ... , ::::,) is acyclic. It may, however, not be transitive. In Exercise 21.0.6 you are asked to find an example of nontransitivity. Transitivity obtains in the special case where I is odd and, for every i, the preference relation ::::, belongs to the class 9"~ c ar~ formed by the rational preference relations:::: that are single peaked with respect to ~ and have the property that no two distinct alternatives are indifferent for ::::. Note that, if I is odd and preferences are in this class, then, for any pair of alternatives, there is always a strict majority for one of them against the other. Hence, in this case, a Condorcet winner necessarily defeats any other alternative. Proposition 21.0.2: Suppose that I is odd and that ;;:: Is a linear order on X. Then pairwise majority voting generates a well-defined social welfare functional F: .'Ji'~ -+ .'11. That is, on the domain of preferences that are single-peaked with respect to ;;:: and, moreover, have the property that no two distinct alternatives are indifferent, we can conclude that the social relation F(::::1' ... , ::::,) generated by pairwise majority voting is complete and transitive. Proof: We already know that i(::::" ... , ::::,) is complete. It remains to show that it is transitive. For this purpose, suppose that x i(::::" ... , ::::,)y and y i(::::" ... , ::::,) z. Under our assumptions (recall that I is odd and that no individual indifference is allowed) this means that x defeats y and y defeats z. Consider the set X' = {x, y, z}. If preferences are restricted to this set then, relative to X', preferences still belong to the class .'Ji'~, and therefore there is an alternative in X' that is not defeated by any
Determination of a median for a single-peaked family.
803
804
CHAPTER
21:
SOCIAL
CHOICE
SECTION
THEORY
-------------------------------------------------------------~
21.0:
SOME
POSSIBILITY
RESULTS:
RESTRICTED
DOMAINS
605
---------------------------------------------------------------x = R'
Figure 21.0.4
Indifference curves for the preferences of Example 21.0.5.
other alternative in X'. This alternative can be neither y (defeated by x) nor z (defeated by y). Hence, it has to be x and we conclude that x transitivity. _
FC??:." . .. , ~,) z, as
required by
In applications, the linear order on alternatives arises typically as the natural order, as real numbers, of the values of a one-dimensional parameter. Then, as we have seen, singlepeakedness follows from the strict quasiconcavity of utility functions, a restriction quite often satisfied in economics. It is an unfortunate fact that the power of quasiconcavity is confined to one-dimensional problems. We illustrate the issues involved in more general cases by discussing two examples. Example 21.0.5: Suppose that the space of alternatives is the unit square, that is, X = [0, I]'. The generic entries of X arc denoted x = (x" x,). There are three agents I = {t, 2, 3}. The preferences of the agents are expressed by the utility functions on X:
o example is that the cone spanned by the nonnegative combinations of the gradient vectors of the three utility functions equals the entire R' (see Figure 21.0.4). Exercises 21.0.7 and 21.0.8 provide further elaboration on this issue. The reason why in two (or more) dimensions, quasiconcavity does not particularly help is that, in contrast with the one-dimensional case, there is no sensible way to assign a "median" to a set of points in the plane. This will become clear in the next, classical, Example 21.0.6 which we now describe. Example 21.0.6: Euclidean Preferences. Suppose that the set of alternatives is R'. Agents have preferenees <: represented by utility functions of the form u(y) = -lly - xII, where x is a fixed alternative in R". In words: x is the most preferred alternative for <: and other alternatives are evaluated by how close they are to x in the Euclidean distance. The indifference curves of a typical consumer in R' are pictured in Figure 21.0.5. In the current example, the set R" does double duty. On the one hand, it represents the set of alternatives. On the other, it also stands for the set of all possible preferences because every x E R' uniquely identifies the preferences that have x as a peak." Given two distinct alternatives y, Z E R', an agent will prefer y to z if and only if his peak is closer to y than to z. Thus, the region of peaks associated with preferences that prefer y to z is
u,(x"x,) = -2x, - x" u,(x"x,)=x, +2x" u,(x" x,)
=
x, - x,.
These preferences are represented in Figure 21.0.4. Every utility function is linear and therefore preferen~es are convex (also, they have a single maximal clement on X).'· But, we will now argue that for every x E X there is ayE X preferred by two of the agents to x. To see this we take an arbitrary x = (x" x,) E [0, I]' and distinguish three cases: (i) If x, = 0, then y = (t x,) is preferred by agents 2 and 3 to x. (ii) If x, = I, then y = (x" t) is preferred by agents I and 3 to x. (iii) If x, > 0 and x, < I, then y = (x, - c, x, + e) E [0, I]' with c > 0, is preferred by agents I and 2 to x. You should verify the claims made in (i), (ii), and (iii). _ The situation illustrated in Example 21.0.5 is not a peculiarity. The key property of the
10. The preferences
or this example are not strictly convex. This is immaterial. Without changing
the nature of the example we could modify them slightly so as to make the indifference curve map strictly convex.
A(y,z)
= {XE R": IIx -
yll < IIx-zlI}.
See Figure 21.0.6 for a representation. Geometrically, the boundary of A(y, z) is the hyperplane perpendicular to the segment connecting y and z and passing through its midpoint. We will consider the idealized limit situation where there is a continuum of agents with Euclidean preferences and the population is described by a density function g(x) defined on R', the set of possible peaks. Then given two distinct alternatives y, z E R", the fraction of the total population that prefers y to z, denoted m,(y, z), is simply the integral of g(') over the region A(y, z) c R'. When will there exist a Condorcet winner? Suppose there is an x· E R' with the property that any hyperplane through x· divides R' into two half-spaces each having a total mass of ! according to the density g('). This point could be called a median for the density g('); it coincides with the usual coneept of a median in the case n = I. A median in this sense is a Condorcet winner. It cannot be defeated by any other alternative because if y ¥- x· then A(x·, y) is larger than a half-space through x· and, therefore, m,(x·, y) ~ t. Conversely, if x·
11. For an example in the same spirit where the two roles arc kept separate, see Grandmont (1978) and Exercise 21.0.9.
Flgur. 21.0.5 (left)
Euclidean preferences in R'. Flgur. 21.D.6 (right)
The region of Euclidean preferences that prefer y to z.
806
CHAPTER
21:
SOCIAL
CHOICE
SECTtON
THEORY
-------------------------------------------------------------------~ A(x"
+ "I,x")
=>
m•. ix"
+ "I, X") > i Figure 21.D.7
If x" is not a median then it is not a Condorcet winner.
Figure 21.D.8
---
-------~-------(a)
----
--B(b)
is not a median then there is a direction q E R' such that the mass of the half-space {z E R': q'Z > q·x·J is larger than 1. Thus, by continuity, if < > 0 is small then the mass of the translated half-space A(x" + f.q, x") is larger than t. Hence x· +I the existence of a median imposes so many conditions (there are many half-spaces) that it becomes a knife-edge property. Figure 21.0.8 provides examples. In Figure 21.0.8(a), the density g(') is the uniform density over a rectangle [a case first studied by Tullock (1967)]. The center of the rectangle is then, indeed, a median. But the rectangle is very special. The typical case is one of nonexistence. In Figure 21.0.8(b), the density g(-) is the uniform density over a triangle. Then no median exists: Through any point of a triangle we can draw a line that divides it into two regions of unequal area." •
SOCtAL
CHOtCE
21.E Social Choice Functions
x = R'
g(z) dz > i
21.E:
since every plane through x" divides th, rectangle into two figures of equal area. (b) Uniform distribution over a triangl~ There is no median.
The task we set ourselves to accomplish in the previous sections was how to aggregate profiles of individual preference relations into a coherent (i.e. rational) social preference order. Presumably, this social preference order is then used to make decisions. In this section we focus directly on social decisions and pose the aggregation question as one of analyzing how profiles of individual preferences turn into social decisions. The main result we obtain again yields a dictatorship conclusion. The result amounts, in a sense, to a translation of the Arrow's impossibility theorem into the language of choice functions. It also offers a reinterpretation of the condition of pairwise independence, and provides a link towards the incentive-based analysis of Chapter 23. As before, we have a set of alternatives X and a finite set of agents I. The set of preference relations ~ on X is denoted iJt. We also designate by fJ' the subset of iJt consisting of the preference relations ~ E iJt with the property that no two distinct alternatives are indifferent for ~. Definition 21.E.1: Given any subset .01 c iJtl, a social choice function f: d .... X defined on .<1 assigns a chosen element f(~", .. , ~/) E X to every profile of individual preferences in d. The notion of social choice function embodies the requirement that the chosen set be single valued. We could argue that this is, after all, in the nature of what a choice is. 13 More restrictive is the fact that we do not allow for random choice. 14 If X is finite, every social welfare functional F(') on a domain d induces a natural social choice function by associating with each (~U" .• ~,) E d a most preferred element in X for the social preference relation F(~I'''''~')' For example, if, as in Proposition 21.0.2, d c fJ'~ is a domain of single-peaked preferences, I is odd, and F(') is the pairwise majority voting social welfare functional defined on .<1, then for every (~I"'" ~,) the choicef(~I"'" ~,) is the Condorcet winner in X. We now state and prove a result parallel to Arrow's impossibility theorem. Recall that for Arrow's theorem we had two conditions: the social welfare functional had to be Paretian and had to be pairwise independent. Here we require again two conditions: first, the social choice function must be, again, (weakly) Paretian; and, second, it should be monotonic. We define these concepts in Definitions 21.E.2 and 21.E.4, respectively.
13. Nevertheless, allowing for multivalued choice sets (that is, allowing there to be more than 12. See Caplin and Nalebuff (1988) for further analysis. They show thai under a restriction on the density function (called "logarithmic concavity" and satisfied. in particular. for uniform densities
over convex sets), there are always points ("generalized medians") in R' with the property that any hyperplane through the point divides RIt into two regions. each of which has mass larger than tIe. This means that these points cannot be defeated by any other alternative if the majority required is not 1 but any number larger than t - (lIe) > ~, 64% say. Of course, a 64% rule becomes less decisiv~ than a 50010 rule: There will now be many pairs of alternatives with the property that no member or the pair ddeats the other.
FUNCTIONS
807
-----------------------------------------------------------------------
one acceptable social choice) is natural in some contexts, and certain assumptions on social choice
may be more plausible in the multi valued case. 14. Note also the contrast between the definition or choice runction here and the similar concept
of choice rule in Section I.e. There we contemplated the possibilily of there being several budgets and of the choice depending On the budget at hand. Here the budget is fixed (it is always X) but the choice may depend on the profile of underlying individual preferences. Clearly we could, but will not, consider situations that encompass both cases. Another contrast with Section I.e is that here we limit ourselves to single-valued choice.
SECTION
808
CHAPTER
2':
SOCIAL
CHOICE
2'.E:
SOCIAL
CHOICE
THEORY
Definition 21.E.2: The social choice function I: d .... X defined on d c iJll is weakly Paretian if for any profile (<:::, ....• <:::/) e d the choice 1(<:::, • ...• <:::/) eX is a weak Pareto optimum. That is. If for some pair {x. y} c X we have that x >-; Y for every i. then y -# 1(<:::, • ...• <:::/)' In order to define monotonicity we need a preliminary concept. Definition 21.E.3: The alternative xeX maintains its position Irom the profile (<:::, ... , . <:::/) e iJll to the profile (<:::; ....• <:::;) e iJll if x <:::; Y implies
x <:::j y
for every i and every y e X. In other words. x maintains its position from (<:::1 •...• <:::d to (<:::', •...• <:::;) if for every i the set of alternatives inferior (or indifferent) to x expands (or remains the same) in moving from <:::, to <:::;. That is. L(x. <:::,) = {ye X: x <:::, y} c L(x. <::::> = {ye X: x <:::; y}.
Note that the condition stated in Definition 21.E.3 imposes no restriction on how other alternatives different from x may change their mutual order in going from <:::, to C~.lS
Definition 21.E.4: The social choice function I: d .... X defined on d c iJll is monotonic if for any two profiles (<:::, ....• <:::d e oW. (<:::; •... , <:::i) e d with the property that the chosen alternative x = 1(<:::" .... <:::1) maintains its position from (<:::, ..... <:::/) to (<:::;, ...• <:::i). we have that 1(<:::; •...• <:::;) = x. In words: The social choice function is monotonic if no alternative can be dropped from being chosen unless for some agent its disirability deteriorates. Are there social choice functions that are weakly Paretian and monotonic? The answer is w yes." For example. in Exercise 21.E.1 you are asked to verify that the pairwise majority voting social decision function defined on a domain of singlepeaked preferences is weakly Paretian and monotonic. But what if we have a universal domain (i.e.• d = !Jtl or d = !?I)? A not very attractive class of social choice functions having the two properties in this domain are the dictatorial social choice functions. Definition 21.E.S: An agent h e / is a dictator for the social choice function I: d .... X if. for every profile (<:::, .... , <:::/) e d. 1(<:::, • ...• <:::/) is a most preferred alternative for <:::h in X; that is. 1(<:::, •...• <:::/) e {xeX: x <:::h y for every VEX}.
A social choice function that admits a dictator is called dictatorial. In the domain !?I. a dictatorial social choice function is weakly Paretian and monotonic. (This is clear enough. but at any rate you should verify it in Exercise 2I.E.2. where you are also asked to discuss the case d = !Jtl.) Unfortunately. in the universal domain we cannot get anything better than the dictatorial social choice functions. The impossibility result of Proposition 21.E.1 establishes this.
J5. As in Section 3.B. the sets L(x. ~i) are also referred to as lower contour sets.
~SltlOn 21.E.1: Suppose that the number of ~Iter~ati~es is at lea~t thre~an~
p ~hat the domain of admissible preference ~roflle~ IS elt.her d =. iJl ~r d - !? . Then every weakly Paretian and monotonic social chOice function I. d .... X Is dictatorial. · The proof of the result will be obtained as a corollary of Arrow's impossibility P roo.f . . I I' theorem (Proposition 21.C.1). To this effect. we proceed to denve a socia we ,are functional F(') that rationalizes f(<:::, • ...• <:::/) for every profile (<:::" .... <:::/) e d. We will then show that F(') satisfies the assumptions of Arrow's theorem. hence
yielding the dictatorship conclus.i~n. We begin with a useful definition. 1 Definition 21.E.6: Given a finite subset X' c X and a profile (<:::, .... , <:::/) e iJl. , we say that the profile (<:::; ....• <:::;) takes X' to the top Irom (<:::, •... , <:::/) If. for
every i.
x >-j y x <:::; Y <=> x <:::; y
for x e X' and y f: X'. for all x, yeX'.
In words: The preference relation <:::: is obtained from <:::, by simply tak~ng every alternative in X' to the top. while preserving the internal. (weak~r str:c~) ordering among these alternatives. The ordering among alternal1ve.s not ,m X IS arbitrary. For example. if x >-, y >-,z >-,w. then the preference relatIOn >-, defi~~ by y>-iw >-iz >-ix takes {y.w} to the top from <:::,. Note, also. tha.tif.(<:::', •. :.,' ~~) takes X' to the top from (<:::, •...• <:::/)' then every x e X mamtams ItS position m going from (<:::" ...• <:::/) to (<:::', •...• <:::;). For the rest of the proof we proceed in steps: Step I: If both the profiles (<:::'1 ..... <:::i)ed and (<:::; .... , <:::i) ed take X' c X to the topfrom (<:::1 ..... <:::/)' then f(<:::'" .. ·• <:::i) =f(<:::~ ... ·• <:::i)· For every i and x e X' we have {yeX:x<:::;y} = {yeX:x<:::i y} = {yeX:x<:::,y} uX\X'.
By the weak Pareto property,!(<:::'" ...• <:::;) e X'. Thus. f(<:::'I' ...• <:::;) e X' maintains its position i.1 going from (<:::', •...• <:::j) to (<:::~ •...• <:::i)· Therefore. by the monotonicity off(·). we conclude thatf(<:::'I.· ... <:::j) = f(<:::~ ... ·• <:::i)· Step 2: Definition of F(<:::, •...• <:::/)' For every profile (<:::, ..... <:::/) e d we define a certain bi~~ry. rela.tion F(<:::, ..... <:::/) on X. Specifically. we let x F(<:::I ..... <:::/)Y' (read as ,x IS so.clally at least as good as y") if x = y or if x = f(?:;'I' ...• <:::;) when (<:::'" ...• <:::1) e d IS an.y profile that takes {x.y} c X to the top from the profile (<:::1 •...• <:::/)' By step I thiS is well defined. that is. independent of the particular profile (<:::; •...• <:::i) chosen. Step 3: For every profile (<:::1 •...• <:::/) e d. F(<:::I •...• <:::/) is a rational preference relation. Moreover. F(<:::I •...• <:::/) e!?; that is. no two distinct alternatives are socially indifferent. Because f(-j is weakly Paretian. it follows that when (<:::'1 ... ·• <:::;) takes {x. y} to the top from (<:::" .... <:::1) we must have f(<:::'I .... • <:::;) e {x. y}. Therefore. we conclude that either x F(<:::I" ..• <:::/) y or y F(<:::I •. ··• <:::/) x. but. because of step I. not both (unless x = y). In particular. F(?:;I •. · .• <:::/) is complete.
FUNCTIONS
609
810
CHAPTER
21:
SOCIAL
CHOICE
THEORY
To verify transitivity. suppose that x F(I:::;I' ..•• 1:::;,) y and y F(I:::;I" ..• 1:::;,) z. We can assume that the three alternatives {x. y. z} are distinct. Let (I:::;~ •...• 1:::;1) e sI be a profile that takes {x, y. z} to the top from (1:::;" •..• 1:::;,). Because f(·) is weakly Paretian, we have f(I:::;~ •. .• , 1:::;1) e {x. y. z}. Suppose that we had y = f(I:::;~ •. ..• 1:::;1). Consider a profile (1:::;; •...• 1:::;;) e sI that takes {x. y} to the top from (I:::;~ •... , I:::;j). Since y maintains its position from (1:::;';,. " • 1:::;/) to (1:::;'1'·· •• 1:::;;), it follows from monotonicity that f(I:::;',,· .. , 1:::;;) = y. But (1:::;'" ... ,1:::;;) also takes {x.y} to the top from (1:::;, ..... 1:::;,): the relative ordering of x and y, the two alternatives at the top, has not been altered in any individual preference in going from (1:::;" •..• 1:::;,) to (1:::;', •...• 1:::;;). Therefore we conclude that y F(I:::; ..... , 1:::;,) x. which contradicts the assumption that x F(I:::;" .... I:::;,)y. x #< y. Hence.y #
--
REF ERE" C E S
F(I:::;, •...• 1:::;,) in Xl must also be a most preferred alternative for h; that is, f(I:::;, ..... 1:::;,) 1:::;. x for every xe X. Hence agent h is a dictator. _ Finally, we mention the following corollary (Proposition 2I.E.2) to hint at the connection between Proposition 21.E.1 and the issue of incentives to truthful preference revelation, a topic that is studied extensively in Chapter 23. proposition 21.E.2: Suppose that the number of alternatives Is at least three and that f: 13" -+ X is a social choice function that is weakly Paretian and satisfies the following no-incentive-to·misrepresent condition:
f(I:::;" . .. , I:::;h-I. I:::;h' I:::;h+l"'" 1:::;,) I:::;h f(I:::;"" " I:::;h-" I:::;i.. I:::;h+l" ..• 1:::;,) for every agent h. every I:::;i, e 13'. and every profile (I:::; I....• 1:::;,) e !?'. Then f(') is dictatorial. Proof; In view of Proposition 21.E.l it suffices to show that f: !?, -+ X must be monotonic. Suppose that it is not. Then without loss of generality we can assume that. for some agent h. there are preferences 1:::;, e!? for agents i #< h, and preferences 1:::;;. 1:::;:' e & for agent h. such that. denoting
x =f(I:::;,.· ... 1:::;.-,.1:::;;.1:::;0+, •...• 1:::;,> The social welfare functional F: sI -+!? rationalizes f: sI -+ X; that is, for every profile (1:::;" ...• 1:::;,) e sI,f(I:::;" . ..• 1:::;,) is a most preferred alternative for F(I:::;" . ... 1:::;,) in X. Step 4:
This is intuitive enough since F(') has been constructed from f(·). Denote x = f(I:::;" ... , 1:::;,) and let y #< x be any other alternative. Consider a profile (1:::;'" ... , 1:::;;) e sI that takes {x. y} to the top from (1:::;" .... 1:::;,). Since x maintains position from (1:::;" ...• 1:::;,) to (1:::;; •...• 1:::;;). we have x = f(I:::;'" .••• 1:::;;). Therefore, x F(I:::;" ... , 1:::;,) y.
Step 5: The social welfare functional F: sI -+ !? is Paretian. Clear if x >-, y for every i then, by the Paretian property of f('), we must have x = f(I:::;'" . .. , 1:::;;) whenever (1:::;;, .•. ,1:::;;) takes Ix. y} to the top from (1:::;" ..• , 1:::;,). Hence x F(I:::;" ... , 1:::;,) y, and by step 3 we conclude that x F.(I:::;" ... , 1:::;,) y.
and y =f(I:::; .. · ... 1:::;.-" 1:::;;.1:::;0+" •..• 1:::;,).
we have that xl:::;; z implies xl:::;; z for every z e X, and yet y #< x. There are two possibilities: Either y >-; x or xl:::;; y. If y >-; x then the no-incentive-to-misrepresent condition is violated for the "true" preference relation 1:::;. = 1:::;; and the misrepresentation I:::;~ = 1:::; •. If xl:::;;: y then xl:::;;; y. Therefore, since no two distinct alternatives can be indifferent, x >-~" y. But if x >-;. y then the no-incentive-to-misrepresent condition is violated for the "true" preference relation 1:::;. = 1:::;; and the misrepresentation I:::;~ = 1:::;;. •
REFERENClS
Step 6: The social welfare functional F: sI -+!? satisfies the pairwise independence condition. This follows from step I. Suppose that (1:::;" ... , 1:::;,) e.. Then x F(I:::;" ... , 1:::;,) y. But (I:::;i, ... , 1:::;;) also takes {x, y} to the top from (1:::;', •... , 1:::;;). Hence. x F(I:::;·" . .•• 1:::;; )y, as we wanted to prove.
Step 7:
The social choice function f: sI -+ X is dictatorial. By Arrow's theorem (Proposition 21.C.1) there is an agent he I such that for every profile (1:::;, ..... I:::;,)esl we have xF.(I:::;" .... I:::;,)y whenever x>-.y. Therefore, f(I:::;" ...• 1:::;,) [which by step 4 is a most preferred alternative for
811
--------------------------------------------------------------------
Arrow, K. 1. (\963). Social Choice and Individual Values. 2d ed. New York: Wiley. Austen-Smith. D .. and J. S. Banks. (1996). Positit~ Polirical Theory. Ann Arbor. University of Michigan Press. Caplin. A.. and B. NalebulT. (t988). On 64%-majority voting. Econometrica 56: 787-814. Grandmont. J-M. (1978). Intermediate preferences and majority rule. EconomrrricQ.46: 317-30. May, K. (1952). A set of independent, necessary and sufficient conditions (or simple majority decision. Econometrica 20: 680-84. Moulin. H. (1988) Axioms olCoop
812
CHAPTER
21:
SOCIAL
CHOICE
THEORY
-------------------------------------------------------------------~
EXERCISES 2I.B.I' Verify that majority voting between two alternatives satisfies the properties of symmetry among agents, neutrality between alternatives, and positive responsiveness. 2I.B.2' For each of the three properties characterizing majority voting between two alter· natives according to Proposition 21.B.1 (symmetry among agents, neutrality between alter· natives, and positive responsiveness) exhibit an example of a social welfare functional F(~" ... , a,) distinct from majority voting and satisfying the other two properties. This shows that none of the three properties is redundant for the characterization result. 2I.B.3' Suppose there is a public good project that can take two levels k e {O, I}, where k = 0 can be interpreted as the status quo. The cost, in dollars, of any level of the public good is zero. There is a population I of agents having quasilinear preferences (with dollars as numeraire) over levels of the public good and money holdings. Thus, the preferences of agent i are completely described by the willingness to pay V; e R for the level k = lover the level k = O. The number V; may be negative (in this case it amounts to the minimum compensation required). Show that a majority rule decision over the two levels of the public project guarantees a Pareto optimal decision over the set of policies constituted by the two levels of the public project (with no money transfers taking place across agents) but not over the larger set of policies in which transfers across agents are also possible. Compare and contrast the majority decision rule (a "median ") with the Pareto optimum decision rule for the case in which transfers across agents are possible (a "mean "). 21.C.1' Provide the requested completion of step 1 of the proof of Proposition 21.C.1. 21.C.2 8 We can list the implicit and explicit assumptions of the Arrow impossibility theorem (Proposition 21.C.1) to be the following; (a) The number of alternatives is at least 3. (b) Universal domain: To be specific, the domain of the social welfare functional F(') is at'. (e) Social rationality: That is, F('=" ... , is a rational preference relation (i.e. complete and transitive) for every possible profile of individual preferences. (d) Pairwise independence (Definition 21.C.3). (e) Paretian condition (Definition 21.C.2). (f) No dictatorship: That is, there is no agent h that at any profile of individual preferences
'=,)
imposes his strict preference over any possible pair or alternatives (see Proposition
21.C.1 for a precise definition). For each of these six assumptions exhibit a social welfare functional F(') satisfying the other five. This shows that none of the conditions is redundant for the impossibility result. 21.C.J' Show that there are social welfare functionals F: dt' ... Jt defined on dt' (i.e., individual indifference is possible) satisfying all the conditions of Arrow's impossibility theorem (Proposition 21.C.1) and for which, however, the social preferences are not identical to the preferences of any individual. [Hint: Try a lexical dictatorship in which the nth-ranked dictator imposes his preference if and only if every higher ranked dictator is indifferent.] 21.0.1 8 Suppose that X is a finite set of alternatives. Construct a reflexive and complete on X with the property that has a maximal element on every strict preference relation is not acyclic. subset X' c X, and yet
'=
'=
'=
21.0.2' Verify that the social preferences generated by the oligarchy example (Example 2I.D.I) are quasitransitive but that social indifference may not be transitive. Interpret.
--
EXERCISES
2\.0.3' Show that the social preferences generated by the vetoers example (Example 2I.D.2) are acyclic but not necessarily quasi transitive. Show also that in spite of the veto power of agent 2 it may happen that alternative x is the only maximal alternative for the social preferences.
21.0.4' With reference to Example 21.D.4, show that a continuous preference relation X = [0, I] is single peaked only if it is striclly convex.
'= on
21.05' Give a direct proof that none of the six linear orders possible among three alternatives can make the three preferences involved in the Condorcet paradox (Example 2I.C.2) into a single-peaked family. 21.0.6 8 Give an example with an even number of agents and single-peaked preferences in which pairwise majority voting fails to generate a fully transitive social preference relation. 21.0.7 c Suppose that X is a convex subset of R' with the origin in its interior. There are three agents i = 1,2,3. Every i has a continuously differentiable utility function U;: X ... R. Assume that the cone in R' spanned by the set of gradients at the origin {Vu,(O), Vu,(O), Vu,(O)} is the entire R'. Show the following: (a) There are three alternatives x, y, z e X that constitute a Condorcet cycle (i.e., there is a strict majority for x over y, y over z, and z over x). (b) (Harder) Given any x e R', there is aye R' such that II x - y II < II x II and y is preferred by two agents to the origin 0 e R'. That is, if you think of the origin as the status-quo then for any x we can find a strict majority that prefers, over the status-quo, an alternative that moves us closer to x. [Hint: You can safely assume that the utility functions are linear.] 21.0.Sc The situation is as in Exercise 21.0.7 except that now, at the origin, the gradients of the utility functions constitute a pointed cone (i.e. the cone does not contain any half-space). Assume also that utility functions are quasiconcave. (s) Argue that at the origin there is an agent who is a directional median in the sense that any alternative having a strict majority against the origin must make this agent strictly better off. (b) Suppose now that at every x e X the cone spanned by {Vu,(x), Vu,(x), Vu,(x)} is pointed. Then according to (a) there is a directional median agent at every x e X. Show that this directional median agent can change with x and that Condorcet cycles are possible. (e) The situation is as in (b). Show that, if the directional median agent is the same at every x e X, then there can be no Condorcet cycle. 21.0.9C (Grandmont) Consider a set of alternatives X. Given three rational preference and ,=' if x y and relations ,=, ;::', ,=" on X, one says that ,=" is intermediate between x ,=' y implies x ,=" y. That is, for every alternative y the intersection of the upper contour sets for and ,=' is contained in the upper contour set for ,=".
'=
'=
'=
(0) Show that if u(x) and u'(x) are utility functions for preferences on X then, for any
positive numbers y and "', the preference relation represented by "'u(x) + yu'(x) is intermediate between the preference relations represented by u(x) and u'(x). (h) Suppose we are given N functions h,(x), ... , hH(x) defined on X. The preferences of agents are represented by utility functions ofthe form u,(x) = P,h,(x) + ... + PNhH(x), where P= (P" ... , PH) e R". •. Show that for any two alternatives x, y e X, the set B(x, y) = {P e R"..: u,(x) > u,(y)} is the intersection of R".. with a translated half-space.
813
814
CHAPTER
21:
SOCIAL
CHOICE
THEORY
--------------------------------------------------------------------------------(e) Argue that the conclusion from (b) is still correct if a parametrization of utility functions u,(x) by a fJ eRN is such that whenever fJ" is a convex combination of fJ and fJ' then the preferences represented by u,-(x) are intermediate between the preferences represented by u,(x) and u,.(x). (d) Continuing with the parametrization of (b), suppose that we take the limit situation where the population of agents is represented by a density g(fJ) over R~ +. We say that po is a median agenl for g(.) if every hyperplane in RN passing through fJ' divides RN into two regions having equal mass according to the density g('). Show that a median agent for an arbitrary g(') mayor may not exist. (e) In the framework of (d), suppose there is a median agent p., that g(p.) > 0, and that x· is the single most preferred alternative of the median agent. Show then that x· defeats any other alternative in pairwise majority voting. (f) Show that the Euclidean preferences in Example 21.0.6 can be put into the framework of this exercise by keeping the sets of alternatives and of agents conceptually separated. 21.0.10· The purpose of this exercise is to illustrate the use of single·peakedness in a policy problem. Specifically, we consider the problem of determining by majority voting a tax level for wealth redistribution. Suppose that there is an odd number I of agents. Each agent has a level of wealth w, > 0 and an increasing utility function over wealth levels. The mean wealth is W, and the median wealth is w*, (a) Interpret the distributional significance of a ditTerence between wand W·. (b) Consider a proportional tax rate I e [0, 1] identical across agents. The set of alternatives is X = [0, I), the set of possible levels of the tax rate. Tax receipts are redistributed uniformly. Thus, for a tax rate I, the after·tax wealth of agent i is (1 - I)w, + IW. Show that the preferences over X of all agents are single peaked and that the Condorcet winner I, is I, = 0 or I, = I according to whether w· > w or w· < W, respectively. Interpret. (e) Now suppose that taxation gives rise to a deadweight loss. Being very crude about it, suppose that a tax rate of Ie [0, I] decreases the pretax level of agent i's wealth to w,(I) = (1 - I)W, [thus, the average tax receipts are 1(1 - I)w and the ex post wealth level of agent i is (I - I)'w, + 1(1 - I)W). Show that preferences on wealth levels are again single peaked (but notice that the after·tax wealth level may not be a concave function of the tax rate). Show then that I, ~ t. Also, I, = 0 or I, > 0 according to whether w· > !w or w· < tw, respectively. Compare with (b) and interpret. (d) Let us modify (e) by assuming that the deadweight loss atTects individual wealth ditTerently: A tax rate of I e [0, I] decreases pretax wealth of agent i to (1 - I')W, [this is theoretically more satisfactory than the situation in (e) since we know from first principles that at I = 0 a small increase in I should have a second·order etTect on tolal welfare]. Show then that individual preferences on tax rates need no longer be single peaked. 21.0.11 8 Consider a finite set of alternatives X and a set of pleferences f!I"" single'peaked with respect to some linear order ~ on X (note that we rule out the possibility of individual indifference). The number of agents is odd. As we have seen in Proposition 21.0.2, a possible class of social welfare functionals F: f!I~ .... f!I that satisfy the Paretian and pairwise independence conditions are those where we fix a subset Sci composed of an odd number of agents (a kind of oligarchy) and let the members of this subset determine social preferences by pairwise majority voting. Show by example that this is nOI the only possible class of social welfare functionals F: 9"", .... 9' that satisfy the Paretian and pairwise independence conditions.
EXERCISE.
815
-----------------------------------------------------~~~~~ 21.D.12 A Suppose that the total cost c > 0 of a project has to be financed by levying taxes from three agents. Therefore, the set of alternatives is X = {(r .. 1,,1,) ~ 0: 1\ + I, + I, = c}. The financing scheme is to be decided by majority voting. (a) Show that no strictly positive alternative (1,,1,,1,)>> 0 can be a Condorcet winner. (b) Discuss what happens with alternatives (1"1,, I,) where I, = 0 for some i. 21.0.\38 We have a population of agents (to be simple, a continuum) with Euclidean preferences in R'. The preferences of the agents fall into a finite number J of types. Each type is indexed by the most preferred point Xi' We assume that the x/s are in "general position," in the sense that no three of the x/s line up into a straight line. We denote by l1.i e [0, 1] the fraction of the total mass of agents that are of type j. (a) Suppose that J is odd and 11., = ... = I1. J • Prove that if ye R' is a Condorcet winning alternative, then ye {x ..... , xJ}. That is, the Condorcet winning alternative must coincide with the top alternative of some type. Does this remain true if J is even? (b) (De Marzo) Suppose now that there is a dominant type, that is, a type h such that 11., > l1.i for every j ~ h. Prove that if there is a Condorcet winning alternative ye R', then y = x,. That is, only the top alternative of the dominant type can be a Condorcet winning alternative. 21.0.14 8 In this exercise we verify that we cannot weaken the definition of single-peakedness to require only that preference be weakly increasing as the peak is approached. Suppose we have five agents and five alternatives {x, y, z, v, wI. The individual preferences are
w)-,X-,Y-SZ"""SV 1
(a) Show that there is no Condorcet winner among these alternatives; that is, every alternative is defeated by majority voting by some other alternative. (b) Show that there is a linear order ~ on the alternatives such that the preference relation of the five 1gents satisfies the following property: "Preferences are weakly increasing as we approach, in the linear order ~, the most preferred alternative of the agent." (e) Verify that the alternatives could be viewed as points in [0, I] and that the preferences of each agent could be induced by the restriction to the set of alternatives of a quasiconcave utility function on [0. 1]. [Note: ui(t) is quasiconcave if (I e [0,1): u,(r) ~ y} is convex for every y.] (d) (Harder) Extend the previous arguments and constructions into an example with the following characteristics: (i) There are five agents; (ii) the space of alternatives equals the interval [0, 1]; (iii) every agent has a quasiconcave utility function on [0, 1] with a single maximal alternative; and (iv) there is no Condorcet winner in [0, 1]. 21.E.IA Consider a finite set of alternatives X and suppose that there is an odd number of agents. The domain of preferences is sf = 9'~, where ~ is a linear order on X (i.e., preferences are single peaked and individual indifferences do not arise). Show that the social choice function that assigns the Condorcet winner to every profile satisfies the weak Pareto and the monotonicity conditions.
816
CHAPTER
21,
SOCIAL
CHOICE
C
THEORY
----------------------------------------------------------------------------------21.E.2 A Suppose that the set of alternatives X has N < 00 elements and that the alternatives are given to us with labels that go from I to N, that is, X = {Xl'.·.' x N }. Consider the social choice function defined on !lI' (i.e~ we allow for individual indifference) by lening f(':;:;" ... , ':;:;,) be the alternative that has the smallest label among all the alternatives that are most preferred by the first agent. Show that this social choice function is dictatorial, weakly Paretian and monotonic. For the sake of completeness, carry out the same verification if the domain of f(·) is !?'.
Elements of Welfare Economics
HAP
T
E
R
22
and Axiomatic Bargaining
21.E.3' As requested, complete the proof of step 3 of Proposition 21.E.1. 2I.E.4A Suppose that the number of alternatives is finite and that F:.JJI -+ i? is a social welfare functional satisfying the weakly Paretian and the pairwise independence condition on some domain .<1 c #'. The induced social choice function assigns to every profile the socially most preferred alternative. Give two examples where the induced social choice function is not monotonic. One example should be for the two·alternative case and .JJI = !pI, and the other should be for a case with more than two alternatives. [Him: Choose .<1 to be very smaIL]
22.A Introduction [n this chapter, we continue our study of welfare economics. The main difference from Chapter 21 is that here the cardinal aspects of individual utility functions will be at center stage. Moreover, we will not eschew exploring the implications of assuming that utilities are interpersonally comparable. [n Section 22.B, we present the concept of the utility possibility set. We also emphasize the distinction between first-best and second-best welfare problems. [n Section 22.C, we first posit the existence of a policy maker, or social planner, endowed with coherent objectives in the form of a social welfare function. The role of policy consists, precisely, in the maximization of the social welfare function subject to the constraint represented by the utility possibility set. We then analyze a variety of practically useful examples. The section concludes with a brief discussion of the compensat ion e,;1 erion. In Section 22.D, we probe the extent to which interpersonal comparisons of utility underlie the use of social welfare functions. We do this by analyzing the implications of axioms that postulate the invariance of social preferences to changes in the origins and units of individual utility functions. This section links naturally with Chapter 21 as, again, it relies on the concept of a social welfare functional and, through a different road, it takes us back to Am,w's impossibility theorem. Sections 22.E and 22.F deal with a somewhat different topic: axiomatic bargaining theory. The aim is now to formulate and analyze reasonable criteria for dividing among several agents the gains (or losses) from a cooperative endeavor. [n Section 22.E, we study the simplest case: that in which either there is complete cooperation (with the possible outcomes of cooperation described by a utility possibility set) or the outcome is a given threat point. We present several solutions for this case, among them the classical Nash bargaining solution. [n Section 22.F, we restrict ourselves to the situation in which the utility is transferable among agents. We allow, however, for the possibility of cooperation among subgroups of agents. A classical solution is then the Shapley value, of which we give a brief account. We also provide an illustration of an interesting application of the Shapley value to a problem of allocating joint costs to individual projects. 817
818
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
ANO
AXIOMATIC
22.B Utility Possibility Sets
Definition 22.B.1: The utility possibility set (UPS) is the set E
R': u, S u,(x), ... , u,
s
u,(x) for some x E
Xl
C
RL.
The Pareto frontier of U is formed by the utility vectors u = (u" . .. , u,) E U for which there is no other u' = (u;, ... , u;) E U with U; u i for some i. To gain some insight into the characteristics of UPSs, and in particular, into the important distinction between first-best and second-best policy problems, we discuss some examples. Example 22.B.1: Exchange Economies. Suppose that we focus on the exchange and production economies with L commodities and I consumers studied in Chapter 10 and in Part IV. The set of alternatives X c RLI then stands for the set of feasible consumption allocations x = (x" ... , x,). The utility functions of the different consumers have the form ",(x) = II,(X,); that is, consumer i's utility from an allocation depends only on her own consumption. In Exercise 22.B.I you are asked to show that under standard conditions (including the concavity of the utility functions), the UPS of this economy is a convex set. In the particular case where the utility functions are quasilinear! we saw in Section 10.0 that the boundary of U is a hyperplane. The general case and the quasilinear case are illustrated in Figures 22.8.1 and 22.B.2, respectively. _
J. For general introductions to public economics, see Atkinson and Stiglitz (1980), Laffonl (1988). and Starrett (1988). At a more advanced level, see Guesnerie (1995). Phelps (1973) contains a good compilation of basic articles emphasizing conceptual foundations. 2. As usual, in this case we also neglect the nonnegativity constraints on numeraire.
22. B:
UTI LIT Y
P 0 S SIB I LIT Y
8 ETS
819
",
",
As a first step in the study of policy decision problems, this section is concerned with the description of the set of options available to a policy maker. The following section will consider the objectives of the policy maker. I The starting point of the analysis is a nonempty set of alternatives X and a collection of I agents. In contrast with Chapter 21, where we used preference relations, we will now assume that agents' tastes are given to us in the form of utility functions II,: X -+ R. One may wonder what is the exact meaning of the utility values II.eX): DO they have cardinal or ordinal significance? Are they comparable across individuals? These questions will be considered in Section 22.0. For current purposes there is no need to answer them. It is a traditional, and firm, principle of welfare economics that policy making should not be paternalistic. At a minimum, this means that alternatives that cannot be distinguished from the standpoint of agents' tastes should not be distinguished by the policy maker either. We are therefore led to the idea that only the agents' utility values for the different alternative should matter and therefore that the relevant constraint set for the policy maker is the utility possibility set [introduced by Samuelson (1947)], which we now define.
U = {(u" ... , ud
SEC T ION
BARGAINING
Figure 22.B.1 (leH)
A utility possibility set.
u
Flgur. 22.B.2 (right)
",
",
Example 22.B.I corresponds to a first-best situation. A first-best problem is one in which the constraints defining X are only those imposed by technology and resources. The policy maker cannot produce from a void and, therefore, must respect these constraints, but otherwise she can appeal to any conceivable policy instrument. If, as is often the case, there are other restrictions on the usable instruments, we say that we have a second-best problem. The restrictions can be of many sorts: legal, institutional, or, more fundamentally, informational. The last type were amply illustrated in Chapters 13 and 14 (and will be seen again in Chapter 23). We should warn, however, that the conceptual distinction between first-best and second-best problems is not sharp. In a sense, adverse selection or agency restrictions are as primitive as technologies and endowments.
Example 22.8.2: Ramsey Taxation. Consider a quasilinear economy with three goods, of which the third is the numeraire. The numeraire good can be freely transferred across consumers (more formally, one of the policy instruments available to the policy maker is the lump-sum redistribution of wealth). The first two goods are produced from the numeraire at a constant marginal cost equal to I. Consumers face market prices that are eq·.al to marginal cost plus a commodity tax whose level is fixed by the policy maker. Tax proceeds are returned to the economy in lump-sum form. Finally, the amounts consumed are those determined by the demand functions of the different consumers. We know from the second welfare theorem (Section 16.0) that any utility vector in the first-best UPS can be reached with the above instruments (it suffices to set the tax rates at a zero level and distribute wealth appropriately). But suppose that we now have an unavoidable distortion-the policy maker is constrained to raise a total amount R of tax receipts. This has then become a second-best problem. To determine the corresponding second-best UPS, note first that, since the numeraire is freely transferable across consumers, the boundary of this set is still linear, as in the first-best case (i.e., as in Figure 22.8.2). Hence, to place this boundary it suffices to find the level of prices P" P2 that maximizes V(P" Pi), the indirect utility function of a representative consumer (which, up to an increasing transformation, equals the
A utility possibility set: transferable utility.
820
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
aggregate consumer surplus; see Section 4.D and Chapter 10 for these concepts). J Denote by XI(PI' p,) and X,(PI' p,) the aggregate demand functions. Then we must solve the problem Max
--- --
BARGAININQ
v( PI' p,) (PI - I)x,(p" p,)
S.t.
+ (p,
(PI - 1) oxI(p" p,) OP1
_ dxl(p,) .),(PI - 1) - - = (1 -1.)X,(p.) dp,
and
Denoting by tl = (PI - I)lpl the tax rate on good I, we can write this condition in elasticity form as 0:
t1 = - -
for some
0:
> O.
PO S SIB I LIT Y
+ (p,
_ 1) ox,(p" P1) = O. OP1
(22.B.2)
Example 22.B.4: Few Policy Instruments. In Examples 22.B.2 and 22.B.3 we have assumed that the unrestricted transfer of numeraire across consumers is one of the instruments available to the policy maker. Because of this, in those two examples the UPS had a "full" frontier, that is, a frontier that is an (1- I)-dimensional surface. In addition, quasilinearity insured that this surface was flat (and therefore that the UPS was convex). We now explore the implications of limiting the extent to which the numeraire is transferable. We assume that we have two goods and that the utility functions of I consumers are quasilinear with respect to the first good (which is untaxed). Arbitrary transfers of numeraire are not permitted, however. The policy maker now has a single instrument: a commodity tax (or subsidy) on the second good. Again, this good can be produced at unit marginal cost. The policy maker's surplus (or deficit) is given back to the consumers according to some fixed rule (hence, no arbitrary transfers of numeraire are permitted). Say, to be specific, that this rule is that the surplus-deficit is absorbed by the first consumer. Then the (second-best) UPS is [denoting by Vi(P1) the indirect utility f.. nction of consumer i]
.( p, - - 1) dx,(p,) - - - = (1 - j.') X, (-) p, . dp,
j.
and
UTI L , T Y
Note that except in the separable case, where ox,(p" p,)lop, = 0, we have p, i' 1; that is, even if the initial distortion involves only the first market, second-best efficiency requires creating a compensatory distortion in the second market [this point was emphasized by Lipsey and Lancaster (1956)]. This is an intuitive result: suppose that we were to put P1 = 1; then the last (infinitesimal) unit demanded of the second good makes a contribution P1 - 1 = 0 to the total surplus (recall that p, will equal the marginal utility for good 2). Therefore, a small tax on good 2 is desirable because its effect is to divert some demand toward good I, where the contribution to total surplus of the last unit demanded is PI - 1 > O. •
There is i. < 0, such that
0:
22. B:
assume that PI is fixed at some level PI > I. s The policy instruments are any transfer of numeraire across agents and the level of a commodity tax on the second good. The net revenue in the two markets is given back to consumers in a lump-sum form. The solution P1 of the surplus-maximization problem is then characterized by the first-order conditions (see Exercise 22.B.3)
- I)x,(p" p,) ;e: R.
Suppose, to take the simplest case, that the utility functions of the different consumers are additively separable. This means that the two demand functions can be written as x ,(PI) and x,(p,). Then the first-order conditions satisfied by a solution (PI' p,) of the maximization problem are (carry out the calculation in Exercise 22.B.2):
tl =-£I(PI)
SEC T ION
(22.B.I)
£1(P1)
Expression (22.8.1) is known as the Ramsey taxation formula [because of Ramsey (1927)]. An implication of it is that if the demand for good 1 is uniformly less elastic than that for good 2, then the optimal tax rate for good I is higher. This makes sense: For example, if the demand for good I is totally inelastic then there is no deadweight loss from taxation of this good (see Section IO.C) and therefore we could reach the first-best optimum by taxing only this good.' • Example 22.B.3: Compensatory Distortion. The basic economy is as in Example 22.8.2, except that we do not necessarily assume that the utility functions of the consumers are additively separable. The distortion is now of a different type. We
U = {u
E
R': u :S (V,(P1)
+ (p,
- 1) Li Xi(P1), v1(p,), ... , V,(P1» for some P1 > O}.
Two points are worth observing. The first is that U does not need to be convex (you should show this in Exercise 22.B.4; recall from Proposition 3.D.3 that the indirect utility functions are quasi-convex. An example is represented in Figure 22.B.3. The second is that U is defined by means of a single parameter, P2' and therefore its Pareto frontier (which, naturally,lies in R') is one-dimensional. See Figure 22.B.4 for a case with I = 3. This feature is entirely typical. As long as the instruments available to the policymaker are fewer than I - 1 in number, the frontier of the UPS cannot be (I - 1)-dimensional. Note that when there is free transferability of numeraire across
3. Because total surplus equals consumer surplus plus the fixed amount of tax revenues R. by maximizing consumer surplus we maximize total surplus. We nole also thai the assumption that the amount R must be raised through commodity taxation is somewhat artificial in a context where lump-sum redistribution is possible. We make the assumption, in this and the next example, merely 10 be pedagogical. Alternalively, we could rule oul Ihe possibilily of lump-sum Iransfers. In this case the exercise carried out in this example (and the next) determines the first-order conditions ror Ihe problem of maximizing the sum of individual utilities (the "purely utilitarian social welfare function" in the terminology of Section 22.C) 4. We should warn that the formulas in (22.B.I) constitute only first-order conditions. As we shall see in Ihe forthcoming examples, second-best problems are frequently nonconvex and therefore the satisraction of first-order conditions does not guarantee that we have determined a true maximum.
5. More generally. we could think of the market for good I as being beyond the control of the policy maker and giving rise, perhaps because of a monopolistic structure. to a price higher than marginal cost.
1
SET S
821
822
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
BAROAININO
SECTION
~
22 •• :
UTILITY
POSSIBIL'TY
SETS
823
---------------------------------------------------------------------------------",
Figure 22.B.3 (left)
",
A nonconvex second-best utility possibility set (Example 22.B.4).
Flgur. 22.B.6
Figure 22.B.4 (rlghl)
",
",
"J
A second-best utility possibility set for a case with few instruments: low-dimensional Pareto frontier (Example 22.B.4).
the I consumers, this automatically gives us the necessary minimum of I - I instruments. _
",
A nonconvex utility possibility set for a first·best problem wilh externalities (Example 22.8.5).
Chapter 6). then the (expected, or ex ante) UPS is convex since it is just the set of convex combinations of the utility vectors in the UPS associated with deterministic policies. There is no general theoretical reason to prevent the policy making from randomizing. On the other hand, the practical admissibility of stochastic policies cannot be decided on a priori grounds either. We conclude this section with a final example [borrowed from Atkinson (1973)] that highlights the contrast between first-best and second-best problems. Example 22.B.6: Unproductive Taxation. Suppose that there are two commodities and two consumers. We call the first commodity "labor", or leisure, and the second the "consumption good." There is a total of one unit of labor which is entirely owned by the first consumer. The consumption good can be produced by the first consumer from labor at a constant marginal cost of I (there is also free disposal). The first consumer has a utility function u,(x, I> X2') and the second has "2(X22)' In Figure 22.B.7 we illustrate the construction of the first-best Pareto frontier for this model. Suppose that u l is given. Then, subject to attaining the level of utility ", for consumer I, we want to give to consumer 2 as much utility as possible. If consumer I gets (x 11' X2') then the labor supply is I - XII and the amount of consumption good available for consumer 2 is I - XI' - X2" Thus, we should first determine (X'I' X2') by minimizing x,, + X2' subject to u,(x, I> X2') ~ "" and then let U2 = u2(1 - XII - .x2d· We now study the second-best problem where consumer I cannot be forced to supply labor. The only available policy instrument for providing consumption good
Example 22.B.5: First-best N onconvexities. In Example 22.B.4 the possible nonconvexity of the UPS is due to the second-best nature of this set.lflump-sum transfers of numeraire were allowed, then the corresponding first-best UPS would be convex. Yet a first-best UPS may also be nonconvex. Two familiar sources of nonconvexities in first-best problems are indivisibilities and externalities. As for the first, suppose that there are two locations and two agents with identical locational tastes (in particular, they both prefer the same location). There are only two possible assignments of individuals to locations and therefore the UPS will be as in Figure 22.B.5. As for externalities, suppose that there is a single good and that the utility functions of two consumers are u,(x,) = x, and u,(x .. x,) = x,/x,. Then the UPS is as in Figure 22.B.6 (see Appendix A of Chapter II for more on nonconvexities due to externalities). _ Examples 22.B.4 and 22.B.5 have provided instances where the UPS is nonconvex. There is a procedure that permits one, in principle, to convexify the UPS. It consists of allowing the policy maker to randomize over her set of feasible policies. If random outcomes are evaluated by the different agents according to their expected utility (see
Figure 22.B.7
",
Construction of the first·best Pareto frontier for Example 22.8.6.
Figure 22.B.5
A nonconvex utility possibility set for a first-best locational problem (Example
22.B.5).
u
",
"
x" Labor Supply
824
C HAP T E R
2 2:
E L E MEN T 8
0 F
W ELF ARE
E CON 0 M I C 8
AND
A X 10M A TIC
• A RQ AI NINQ
---........
.,
Flgur. 22.B.8 (le"1
Construction of Ih, second-best Pareto frontier for Example 22.B.6. Flgur. 22.B.8 (rlghll
First-best and second-best utility possibilily sels for the unproductive taxalio n example (Example 22.B.6).
x"
Labor Supply
to consumer 2 is a linear tax t(1 - XI') on whatever amount of labour the first consumer decides to supply given the tax rate. The construction of the secondbest frontier is illustrated in Figure 22.B.8. For t ~ O. consumer I will choose X'I so as to maximize u,(x". (I - t)(I - XII»' Observe that this is as if she had chosen the point in her offer curve corresponding to the price vector (I. 1/(1 - t». Denote this point by xl(t) = (xl,(t). X21 (t». The utility of consumer 2 is then u 2 (t(1 - XII(t)))· The first-best and second-best UPS are displayed in Figure 22.B.9. 6 In the second-best case the figure also depicts the locus of utility pairs Q c RI obtained as t ranges from 0 to I. that is.
Q = {(UI(XI(t)). uit(1 - XI I (t)))) e
R2: O:S; I:S;
I}.
Note that Q does not coincide with the Pareto set of the second-best UPS because it exhibits a characteristic nonmonotonicity. The economic intuition underlying it is clear: if t is low. consumer 2 will get very little of the consumption good; but if I is very high. the situation is not much better. Consumer 2 will now get a large fraction of the labor supplied by consumer I. but for precisely this reason not much labor will be supplied by consumer I. • We can distill yet another lesson from Example 22.B.6. We see in Figure 22.B.9 that it is quite possible for the first-best and second-best Pareto frontiers to have some points in common; that is. there may well be second-best Pareto optima that are first-best Pareto optima. Yet Figure 22.B.9 tells us that it would be quite silly to select a point in the second-best Pareto frontier merely according to the criterion of proximity to the first-best frontier. The resulting selection may be distributionally 7 very biased. The investigation of more sensible selection criteria will be the purpose of Section 22.C.
6. Again, the second-best frontier mayor may not be convex.
7. We may add Ihal il may also be uninteresting from Ihe point of view of policy: in Figure 22.B.9 the only second-best policy that yields a first-best result is t = 0, that is, no policy al all!
--
SECTION
22.C:
SOCIAL
WELFI.RE
FUNCTIONS
AND
SOCIAL
22,C Social Welfare Functions and Social Optima In Section 22.8 we described the constraint set of the policy maker. or social planner. The next question is which particular policy is to be selected. The application of the Pareto principle eliminates any policy that leads to utility vectors not in the Pareto frontier. Yet this still leaves considerable room for choice." which. by necessity. must now involve trading off the utility of some agent against that of others. In this section we assume that the policy maker has an explicit and consistent criterion to carry off this task. Specifically. we assume that this criterion is given by a social welfare function W(u) = W(u ...... u,) that aggregates individuals' utilities into social utilities. We can imagine that W(u) reflects the distributional value judgments underlying the decisions of the policy maker." In Section 22.E (and subsequent ones) we will discuss a somewhat different approach. one that puts more emphasis on the bargaining. or arbitration. aspects of the determination of the final policy selection. In the current section, we refrain from questioning the assumption of interpersonal comparability of Ittility, which is implicit in our use of levels of individual utility as arguments in the aggregator function W(u l ••••• u,). Section 22.0. which links with the analysis of Chapter 21, is devoted to investigating this matter. Thus, for a given social welfare function W(·) and utility possibility set U c R', the policy maker's problem is Max
W(u l ••••• u,)
(22.C.1)
s.t.(ul.···.u,)eU. A vector of utilities. or the underlying policies. solving problem (22.C.1) is called a social optimum. If the problem has a second-best nature. and we want to emphasize this fact. then we may refer to a constrained social optimum. We now present and discuss some of the interesting properties that a social welfare function (SWF) may. or may not. satisfy. (i) N onpaternalism. This first property is already implicit in the concept itself of a SWF. It prescribes that in the expression of social preferences only the individual utilities matter: Two alternatives that are considered indifferent by every agent should also be socially indifferent. The planner does not have direct preferences on the final alternatives. (ii) Paretiall property. Granted the previous property. the Paretian property is an uncontroversial complement to it. It simply says that W(·) is increasing; that is, if It; <': It; for all i, then W(u') <': W(u). and if u; > U; for all i. then W(u') > W(u). We also say that W(· ) is strictly Paretian if it is strictly increasing; that is. if u; <': U; for all i and u; > u; for at least one i. then W(u') > W(u). If W(·) is strictly Paretian then a solution to (22.C.I) is necessarily a Pareto optimum. 8. Only exceptionally will the Pareto frontier consist of a single point. Recall also that, as we saw in Example 22.B.3, in second-best situations with few instruments, the requirement of Pareto optimality may not succeed in ruling out many policies.
9. This approach 10 welfare economics was firsllaken by Bergson (1938) and Samuelson (1947).
OPTIMA
825
826
CHAPTER
•• :
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
---
BARGAINING
Figure 22.C.l (left)
K /
/
A symmetric social welfare function.
/
/
/
".C:
",
",
SOCIAL
WELFARE
FUNCTIONS
AND
SOCIAL
/~.'~
The optimum of a symmetric, strictly concave social welfare function on a
Invarianl When Reflecled on Diagonal
",
Symmetric and Convex
",
(iii) Symmetry. The symmetry property asserts that in evaluating social welfare all agents are on the same footing. Formally, W(·) is symmetric if W(u) = W(u') whenever the entries of the vector u [e.g., u = (2,4,5)] constitute a permutation of the entries of the vector u' [e.g., u' = (4, 5, 2)]. In other words, the names of the agents are of no consequence, only the frequencies of the different utility values matter. The indifference curves of a symmetric W(·) are represented in Figure 22.C.1 for a two-agent case. Geometrically, each indifference curve is symmetric with respect to the diagonal. Note also that, because of this, if the indifference surfaces are smooth then the marginal rates of substitution at any u = (Ul> •.. ,u,) with identical coordinates are all equal to I. (iv) Concavity. Finally, a most important property is the concavity of W(·). We saw in Chapter 6 that, in the context of uncertainty, the (strict) concavity of a utility function implies an aversion to risk. Similarly, in the current welfare-theoretic context it can be interpreted as an aversion to inequality condition. A straightforward way to see this is to simply note that if W(·) is concave and W(u) = W(u'), then W(tu + tu') ~ W(u) [with the inequality strict if u ". u' and W(·) is strictly concave]. Another is to observe that if the UPS is convex and symmetric, then the utility vector that assigns the same utility value to every agent is a social optimum of any symmetric and concave SWF (see Figure 22.C.2 and Exercise 22.C.l).10 Thus, with convex UPSs and concave, symmetric SWFs some inequality is called for only if, as will typically be the case, the UPS is not symmetric. It is to be emphasized that in general, and especially for second-best problems, the UPS may not be convex. This means that even if W(·) is concave the identification of social optima is not an easy task. A utility vector that satisfies the first-order conditions of problem (22.C.1) may not satisfy the second-order conditions or, if it does, it still may not constitute a global maximum. We can gain further insights by discussing some important instances of social welfare functions. 10. The set U c R' is symmetric if" e U implies "' e U for any"' e RL that differs from" only by a permutation of its entries. The interpretation of the symmetry property of a UPS is that there is no bias in the ability to produce utility for different agents. In other words, from the point of view of their possible contributions to social welfare. all agents arc identical.
symmetric and con"x utility possibility set ~ egalitarian.
(a)
"I
(b)
",
(c)
OPTIMA
Figure 22.C.3
Figure 22.C.2 (right)
/'
/
/
--
SECTION
",
Example 22.C.I: UriliIarian. A SWF W(II) is pllrely utilitarian if it has the form W(II) = L; U; [or, in the nonsymmetric situation, W(u) = LI PiU,]. In this case, the indifference hypersurfaces of W(·) are hyperplanes. They are represented in Figure 22.C.3(a). Note that W(·) is strictly Paretian. In the purely utilitarian case, increases or decreases in individual utilities translate into identical changes in social utility. The use of the purely utilitarian principle goes back to the very birth of economics as a theoretical discipline. In Exercise 22.C.2 you are asked to develop an interpretation of the purely utilitarian SWF as the expected utility of a single individual "behind the veil of ignorance." Another line of defense, based also on expected utility theory, has been offered by Harsanyi (1955); see Exercise 22.C.3. Because only the total amount of utility matters, the purely utilitarian SWF is neutral towards the inequality in the distribution of utility. It is important not to read into this statement more than it says. In particular, it does not say "distribution of wealth." For example, if there is a fixed amount of wealth to be distributed among individuals and these have strictly concave utility functions for wealth, then the purely utilitarian social optimum will be unique and distribute wealth so as to equalize the marginal utility of wealth across consumers. If, say, the utility functions are identical across individuals then this will choose as the unique social optimum the vector in the Pareto frontier that assigns the same utility to every agent (see Exercise 22.C.1 for generalizations). _ Example 22.C.2: Maximin. A SWF is of maximin or Rawlsian type [because of Rawls (1971)) if it has the form W(u) = Min {u l , ... , u,} [or, in the nonsymmetric case, W(u) = Min {P,u" ... , p,u,}]. In other words, social utility equals the utility value of the worst-off individual. It follows that the social planning problem becomes one of maximizing thl utility of the worst-off individual." The (L-shaped) indifference curves of the maximin SWF are represented in Figure 22.C.3(b). II. One could refine this criterion by adopting a lexical, or serial, maximin decision rule. First maximize the utility of the worst-ofT, then choose among the solutions of this first problem by maximizing the utility of the next worst-off, and so on. With this. the objectives of the policy maker can still be expressed by a le:dm;n social welfare ordering of utility vectors, but the ordering is not conlinuous and cannot be represenled by a SWF (compare with Example 3.CI). Even so, the refinement is natural and important. For example. we are then guaranteed that the social optimum is a Pareto optimum. You 3re asked to show all this in Exercise 22.C.4. Note that the maximin SWF is Paretian but not strictly Paretian. This makes for some difficulties. In Figure 22.C.4 the
Social welfare functions. (a) Purely utilitarian. (b) Maximin or Rawlsian. (c) Generalized utilitarian.
827
828
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
",
",
/ /
45'
/
FIgure 22.C.4
// / /
--- --
/
FIgure 22.C.S (rIght)
Maximin 0plimum Ulililarian 0plimum
Q
Q~------~~--~
"I
. O' . I FII"·be,, pllmum POlo
~
"I
It is reasonably intuitive that this concave SWF will have strong egalitarian implications. In fact, the preference for equality is quite extreme. Suppose, in effect, that U E RI is an arbitrary UPS and that u E U has all its coordinates equal. Then u fails to be the Rawlsian social optimum only if u is not Pareto optimal. Hence, if there is a u = (u., ... , u / ) in the Pareto frontier of U with all its coordinates equal, then u is a maximin optimum. Note, in contrast, that for a purely utilitarian SWF we reached the social optimum at complete equality only in the case where U is convex and symmetric. In Figure 22.C.4, which continues the analysis of Example 22.6.6, we depict a situation where maximin optimization leads to the selection of a policy (a tax level) that does not yield complete equality. Nonetheless, even in this case, the purely utilitarian social optimum is significantly more unequal than the maximin optimum. _
Example 22.C.3: Generalized Utilitarian. A SWF is generalized utilitarian if it has the form W(u) = LI g(u,) [or, in the nonsymmetric case, W(u) = LI g,(u / )), where g(.) is an increasing, concave function. The generalized utilitarian SWF is strictly Paretian and could be regarded as an instance of the purely utilitarian case where the individual utility functions u,(·) have been replaced by g(u,(·». This is not, however, a conceptually useful point of view. The point is precisely that, given the individual utility functions, there is a deliberate social decision to attach decreasing social weight to successive units of individual utility. The social indifference curves for this case are represented in Figure 22.C.3(c). We can also verify in Figure 22.C.4 and 22.C.5 that the equality implications of the generalized utilitarian SWF are intermediate between those of the purely utilitarian and of the maximin SWFs. _ Example 22.C.4: COnstalll Elasticity. An instance of generalized utilitarian functions that is very useful in applications is provided by the family defined by social utility functions g(.) whose marginal utilities have constant elasticity. This is a family in which attitudes towards inequality can be adjusted by means of a single parameter p '2! O. point at the boundary of U with equal coordinates is a maximin optimum but not a Pareto optimum. In the figure we have selected as "maximin optimum'" the leximin optimum (which. by definition, is a maximin optimum itself).
Range of generali~ utilitarian optima for Example 22.B.6 and the constant elasticity SWF of Example 22.C.4 (p E [0, <Xl]).
22.C:
SOCIAL
WELFARE
FUNCTIONS
AND
SOCIAL
For the rest of the example, individual utility values are restricted to be nonnegative. Then, for any p '2! 0, we let
(t.~)
A maximin oplimum for Example 22.8.6.
/
SECTION
BARGAINING
gp(u,) = (I - p)u! -p
and
if p'# I, if p = I.
Note that, as claimed, the elasticity of g~(u,> is constant because we have u,g7(u,>/g~(u,) = -p for all values u,. Taking into account that, for p '# I, h(W) = [1/(1 - p)] WI/(I-P) is an increasing transformation of W, we can represent the generalized utilitarian social preferences in a particularly convenient manner as w,,(u) =
(L, u! -0)11'-0
for p'# I,
w,,(u) =
L, In u,
for p = I.
and
Thus, we obtain the CES functions that are well known from demand and production theories (see Exercises 3.C.6 and 5.C.IO, respectively). Note that for p = 0 we get Wo(u) = L, u" the purely utilitarian case, and as p -+ 00 we get w,,(u) -+ Min {u l , . · . , U/}, the maximin case. (See Exercise 22.C.5.) In Figure 22.C.5 we depict the range of solutions to Example 22.B.6 as we vary p. We see that as the aversion to inequality increases (that is, as p -+ 00) the optimal tax rate increases. Note, however, that even for very high p we do not approach complete equality. On the other hand, none of these second-best solutions corresponds to the point in the Pareto frontier that is also Pareto optimal for the first-best problem. The latter distributes utility so unequally that the equity considerations underlying any symmetric and concave SWF leads us to sacrifice some first-best efficiency for an equity gain. _
The Compensation Principle We could ask ourselves to what an extent we can do welfare economics without social welfare functions. If the purpose of the SWF is the determination of optimal points in a given Pareto frontier, then resorting to them seems indispensable. This is the usage of social welfare functions that we have emphasized up to now; but in practice this is not the only usage. Often, the policy problem is given to us as one of choosing among several different utility possibility sets; these may correspond, for example, to the UPS associated with different levels of a basic policy variable. 12 If we have a social welfare function W(·), then the choice among two utility possibility sets U and U' should be determined by comparing the social utility of the optimum in U with that of the optimum in U'. However, even if there is no explicit social welfare function one may attempt to say something meaningful about this problem using revealed preference-like ide \s. This is the approach underlying the compensation principle (already encountered in Sections 4.0 and to.E). Let us first take the simplest case: that in which we have two utility possibility sets such that U c U'. Then one is very tempted to conclude that U' should be preferred to U. This would certainly be the case if the points that would be chosen 12. Formally. we can reduce this problem to the previous one by considering the overall UPS formed by the union of the UPSs over which we have to choose. But this may not be the most convenient thing to do because it loses the sequential presentation of the problem (first choose among UPS. then choose the utility vector).
OPTIMA
829
,",vV
(,;HAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
BARGAININQ
---......
---
SECTION
22.0:
IN VARIANCE
PROPERTIES
OF
SOCIAL
WELFARE
FUNCTIONS
831
Figure 22.C.7 Flgur. 22.C.&
",
U' passes the weak compensation test over (U, u).
",
within each of V and V' were the optima of a social welfare function. But even if no social welfare function is available the set V' might still be considered superior to V according to the following strong compensation test: For any possible U E V there is a u' E V' such that uj ;,: u, for every i. That is, wherever we are in V it is possible to move to V' and compensate agents in a manner that insures that every agent is made (weakly) better off by the change to V'. If the compensation is actually made, so that every agent will indeed be made better off by a switch from V to V', there is no doubt that the switch should be recommended. But if compensation will not occur, matters are not so clear: By choosing V' over V based only on a potential compensation we are neglecting quite drastically any distributional implication of the policy change. In fact, it is even possible that the change leads to a purely egalitarian worsening (see Exercise 22.C.6). Recall from Section 10.0 that in the quasilinear case we always have V c V' or V' c V. This is because the boundaries of these sets are hyperplanes determined by the unit vector (hence parallel). In addition, this property also guarantees that the strong compensation criterion (which in Sections 3.0 and 10.E we called simply the compensation criterion) coincides with the choice we would make using a purely utilitarian social welfare function. In this quasilinear case, therefore, the strong compensation criterion does not neglect distributional issues to a larger extent than do purely egalitarian social welfare functions. Matters are more delicate when we compare two utility possibility sets V and V' which are such that one is not included in the other, that is, whose frontiers cross (see Figure 22.C.6). Suppose that we know that the outcome with utility possibility set V is the vector u E V, and that we are considering a move to V'.'3 If u E V', and we were to allocate utility optimally in V' according to a social welfare function, then the move to V' would be advisable. More generally, whenever u E V', the move from (V, u) to V' passes the following weak compensation test: There is au' E V' such that u; ;:: u, for every i. That is, given that we know that the outcome at V is u, we could move to V' and compensate every agent in a manner that makes every agent (weakly) better off. In Figure 22.C.6, V' passes the test with respect to (V, u) but not with respect to (V, u). Again, if the compensation is actually paid, then the weak compensation criterion 13. For e~ample. the original U could correspond to some underlying economy and u could be Ihe ulility values of a market equilibrium.
carries weight. If it is not paid, then it is subject to two serious criticisms. The first is the same as before (it disregards distributional consequences). The second is that it may lead to paradoxes. As in Figure 22.C.7, it is possible to have two utility possibility sets U and U', with respective outcomes u E U and u' E U', such that U' passes the weak compensation test over (U, u) and V passes the weak compensation test over (U', u').ln Exercise 22.C.7 you are asked to provide a more explicit example of this possibility in an economic context. Further elaborations are contained in Exercise 22.C.8.
22.D Invariance Properties of Social Welfare Functions In this section, we probe deeper into the meaning of the comparisons of individual utilities implicit in the definition of a social welfare function. The significance of the matter derives from the fact that whereas a policy maker may be able to identify individual cardinal utility functions (from revealed risk behavior, say), it may actually do so but only up to a choice of origins and units. Fixing these parameters unavoidably involves making value judgments about the social weight of the different agents. It is therefore worth examining the extent to which such judgments may be avoided. Thus, following an approach to the problem taken by d'Aspremont and Gevers (1977), Roberts (1980), and Sen (1977), we explore such questions as: What are the implications for social decisions of requiring that social preferences be independent of the units, or the origins, of individual utility functions?" To answer these types of questions, we need to contemplate the dependence of social preferences on profiles of individual utility functions. Thus, the social welfare functionals introduced in Chapter 21 provide a natural starting point for our analysis. However, we mudify their definition slightly by specifying that individual characteristics arc given to us in the form of individual utility functions u,(') rather than as individual preference relations. From now on we are given a set of alternatives X. We denote by 'fI the set of all possible utility functions on X, and by iJt the set of all possible rational (i.e., complete and transitive) preference relations on X.
I !
J
14. In addition to the previous references, you can consult Moulin (1988) for a succinct presentation of the material of this section.
A paradox: V' passes the weak com· pensation test over (V, u), and V passes the weak com· pensation test over (U', u').
832
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
Definition 22.0.1: Given a set X of alternatives. a social welfare functional F: Cfil ..... 9t is a rule that assigns a rational preference relation F(ii, • ...• iii) among the alternatives in the domain X to every possible profile of individual utility functions (u,(·) ....• UI('» defined on X. The strict preference relation derived from F(u, •. ..• u,) is denoted Fp(ii, • ...• u,) .. • As in Chapter 21. we will concern ourselves only with social welfare functionals that are Paretian. Definition 22.0.2: The social welfare functional F: Cfil ..... 9t satisfies the (weak) Pareto property. or is Paretian. if. for any profile (u, • ...• iii) e Cfil and any pair x. veX. we have that u,{x);;o: ii,{V) for all i implies x F(ii, •.. .• iii) V. and also that u,{x) > u,{V) for all i implies x Fp(ii, •. ..• ii,) V. The first issue to explore is the relationship between these social welfare functionals and the social welfare functions of Section 22.C. A social welfare function W(·) assigns a social utility value to profiles (Ul •.••• UI) e RI of individual utility values. whereas a social welfare functional assigns social preferences to profiles (u ,....• u,) of individual utility functions (or. in Section 21.C, of individual preference relations). From a social welfare function W(·) we can generate a social welfare functional simply by letting F(UI' ...• u,) be the preference relation in X induced by the utility function u(x) = W(ul(x) •...• UI(X». The converse may not be possible. however. In order to be able to "factor" a social welfare functional through a social welfare function, the following necessary condition must. at the very least. be satisfied. Suppose that the profile of utility functions changes, but that the profiles of utility values for two given alternatives remain unaltered; then the social ordering among these alternatives should not change (since the value given by the social welfare function to each alternative has not changed). That is. the social ordering among two given alternatives should depend only on the profiles of individual utility values for these alternatives. Apart from being formulated in terms of utilities, this property is analogous to the pairwise independence condition for social welfare functionals (Definition 21.C.3). We keep the same term and state the condition formally in Definition 22.0.3. Definition 22.0.3: The social welfare functional F: Cfil ..... 9t satisfies the pairwise independence condition if. whenever x. veX are two alternatives and (u, • .... UI) eCfi I • (ii; • ...• iii) e Cfil are two utility function profiles with ii,{x) = D;{x) and u,(V) = ii;(V) for all i. we have xF(ii, ..... iil)V
<=>
xF(u; ..... ul)v.
The necessary pairwise independence condition is almost sufficient: In Proposition 22.0.1 we now see that if the number of alternatives is greater than 2, and the Pareto and pairwise independence conditions are satisfied. then we can derive from the social welfare functional a social preference relation defined on profiles (u l •...• UI) e 9t1 of utility values.'6 A standard continuity condition then allows us to represent this
BARGAINING
833
---...... ----------------------------------------------------------SECTION
22.0:
IN VARIANCE
PROPERTIES
OF
SOCIAL
WELFARE
preference relation by means of a function W(u, •...• u,). thereby yielding a social welfare function. proposition 22.0.1: Suppose that there are at least three alternatives in X and that the Paretian social welfare functional F: Cfi' ..... 9t satisfies the pairwise independence condition. Then there is a rational preference relation;::; defined on HI [that is. on profiles (u, . ...• uJl e HI of individual utility values] that generates F(·). In other words. for every profile of utility functions (u, •... • iii) e Cfil and for every pair of alternatives x. veX we have x F(ii, . ...• uJl V
<=>
(u,(x) •...• iii (x)) ;::; (u,(V) •. ··• iil(v)).
Proof: The desired conclusion dictates directly how ;::; should be constructed. Consider any pair of utility profiles u = (u, •. ..• u,) e R' and u' = (u; •. ..• ui) e R'. Then we let u;::; u' if x F(u ...... u,)y for some pair x.yeX and a profile (u, •. ..• u,) e Cfi' with Ui(X) = Ui and Ui(Y) = u; for every i. We argue first that the conclusion u ;::; u'. is independent of the particular two alternatives and the profile of utility functions chosen. Independence of the utility functions chosen is an immediate consequence of the statement of the pairwise independence condition. Proving independence of the pair chosen is a bit more delicate. It sunices to show that if we have concluded that u;::; u' by means of a pair x. Y then. for any third alternative z (recall that by assumption there are third alternatives). we obtain the same conclusion using the pairs x, z or z. y." We carry out the argument for x. z (in Exercise 22.0.2 you are asked to do the same for z. y). To this effect. take a profile of utility functions (UI •...• UI) e Cfi' with Ui(X) = Ui. Ui(Y) = u;. and ui(z) = u; for every i. Because we have concluded that u;::; u' using the pair x, y. we must have x F(u, •... , il,) y. By the Pareto property. we also have Y F(u, •. .. , UI) z. Hence. by the transitivity of F(u, •. ..• UI)' we obtain x F(u, •.. . , u,) z. which is the property we wanted. It remains to prove that ;::; is complete and transitive. Completeness follows simply from the fact that the preference relation F(u, •...• UI) is complete for any (u l •...• ,i,) e Cfil. As for transitivity. let u;::; u';::; u·. where u, u'. u· e R'. Take three alternatives x. Y. z e X and a profile of utility functions (u, ..... u,) e Cfil with Ui(X) = Ui. ui(Y) = u;. and ui(z) = u, for every i. Since u;::; u' and u' ;::; UN. it must be that x F(u, •. .. , u,) y and y F(u, ... .• UI) z. Because of the transitivity of F(u, •. ..• U/). this implies x F(u , •...• u,) z. and so u ;::; UN. Hence. ;::; is transitive. _ By the Pareto condition, the social preference relation;::; obtained in Proposition 22.0.1 is monotone. You are asked to show this formally in Exercise 22.0.3.
Exercise 22.0.3: Show that if the social welfare functional F: Cfi' ..... 9t satisfies the Pareto property. then a social preference relation;::; on utility profiles for which the 17. Indeed. suppose Ihat we initially used the pair (x. y). Consider any other pair (v. w). If v = x u'. Hence, let the chain of (v. w). There y) - (x. z) -
15. Tha' is. x Fi", ..... "lb' if x F(", ..... u,)y but not y F(u, ..... "I)X. 16. In Exercise 22.D.1 you can find examples showing that the Pareto condition and Ihe restriction on the number of alternatives cannot be dispensed with for the result of Proposition
or w = y then we have just claimed that we get the same ordering between u and t' # x and w # y. If. in addition, v # y. then we reach the same ordering by replacements: (x. y) - (v. y) _ (v. w). Similarly. if w # x we can use (x. y) - (x. w) remains the case (v. w) = (y. x). Here we use a third alternative. z. and the chain (x.
22.D.1.
(y,z) - (y.x).
FUNCTIONS
834
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
conclusion of Proposition 22.0.1 holds must be monotone in the sense that if u' then u' ;:: u, and if u' » u then u' :> u.
~
u
The social preference relation;:: on R' obtained in Proposition 22.0.1 need not be continuous or representable by a utility function. Consider, for example, a lexical dictatorship (say that there are two agents and let u:> u' if U I > u; or if U I = u; and U2 > u;) and recall from Example 3.C.1 that this type of ordering is not representable by a utility function. Nonetheless, we want to focus on social welfare functions and so from now on we will simply assume that we deal only with social welfare functionals that, in addition to the assumptions of Proposition 22.0.1, yield a continuous social preference relation ;:: on R'. As in Section 3.C, such a social preference relation can then be represented by a utility function: in fact, a continuous one. This is then our social welfare function W(u" ... , u,). Note that any increasing, continuous transformation of W(·) is also an admissible social welfare function. In summary, we have seen that the existence of a social welfare function generating a given social welfare functional amounts, with some minor qualifications, to the satisfaction of the pairwise independence condition by the social welfare functional. Therefore, we will concern ourselves from now on with a social welfare functional F: Oft' -+ {It that can be generated from an increasing and continuous social welfare function W: R' -+ R, or equivalently, from a monotone and continuous rational preference relation;:: on R'. We will discover that, in this context, natural utility invariance requirements on the social welfare functional have quite drastic effects on the form that we can choose for W(·) and, therefore, on the social welfare functional itself. Definition 22.0.4: We say that the social welfare functional F: Oft' -+ {It is invariant to cammon cardinal transformations If F(ii" ... , ii,) = F(ii;, ... , iii) whenever the profiles of utility functions (ii" ... , ii,) and (ii;, ... , iii) differ only by a common change of origin and units, that Is, whenever there are numbers (J > 0 and IX such that ii,{x) = (Jii;{x) + IX for all i and x eX. If the invariance Is only with respect to common changes of origin (i.e., we require (J = 1) or of units (i.e., we require IX = 0), then we say that F(') is invariant to common changes of origin or of units, respectively. It is hard to quarrel with the requirement of in variance with respect to common cardinal transformations. Even if the policy maker has the ability to compare the utilities of different agents, the notion of an absolute unit or an absolute zero is difficult to comprehend. We begin by analyzing the implications of invariance with respect to common changes of origin. Suppose that the social welfare functional is generated from the social welfare function W(·). We claim that the invariance with respect to common changes of origin can hold only if W(u) = W(u') implies W(u + lXe) = W(u' + lXe) for all profiles of utility values u e R', u' e R' and IX e R, where e = (I, ... , I) is the unit vector. Indeed, let W(u) = W(u') and W(u + lXe) < W(u' + lXe). Consider a pair x, y e X and profile (17" ..• , 17,) e U' with u,(x) = and u,(y) = for every i. Then x F(u I' ••• , 17,) y. However, x F(u;, .•• , 17;) y does not hold when 17;0 = 17,(') + IX, can tradicting the invariance to common changes of origin. Geometrically, the assertion that W(u) = W(u') implies W(u + lXe) = W(u' + lXe) says that the indifference curves of W(·) are parallel with respect to e-they are
u,
u;
--- --
BARGAINING
SECTION
22.0:
INVARIANCE
PROPERTIES
OF
SOCIAL
WELFARE
FUNCTIONS
835
Figure 22.0.1
Indifference map of a social welfare function invariant to identical
changes of utility origins.
obtained from each other by translations along the e direction (see Figure 22.0.1). In Proposition 23.0.2 [due to Roberts (1980)], we show that this property has an important implication: up to an increasing transformation, the social welfare function can be written as a sum of a purely utilitarian social welfare function and a dispersion term. Proposition 22.0.2: Suppose that the social welfare functional F: Oft' -+ {It is generated from a continuous and increasing social welfare function. Suppose also that F(') is invariant to common changes of origins. Then the social welfare functional can be generated from a social welfare function of the form W(U" ... , Ut) = 0 - g(u, - 0, ... , u, - 0).
(22.0.1)
where 0 = (1//) LjUj. Moreover, if F(') is also independent of common changes of units, that is, fully invariant to common cardinal transformations, then g(.) is homogeneous of degree one on its domain: {s e R': LjSj = OJ. Proof: By assumption the social welfare functional F: Oft' -+ {It can be generated by a continuous and monotone preference relation;:: on R'. Moreover the invariance to identical changes of units implies that if u - u' then u + lXe - u' + lXe for any IX e R. We now construct a particular utility function W(·) for ;::. Because of continuity and monotonicity of;:: there is, for every u e R', a single number IX such that u - lXe. Let W(u) denote this number. That is, W(u) is defined by u - W(u)e. (See Figure 22.0.2 for a depiction.) Because of the monotonicity of preferences, W(·) is a legitimate utility representation for ;::.1. The fi.st part of the proof will be concluded if we show that W(u) - ii depends only on the vector of deviations (u, - 17, ... , u, - 17) = u - ae, that is, that if u - iie = u' - a'e then W(u) - 17 = W(u') - a'. But this is true because u - W(u)e and the invariance to common changes of origin imply that if u - ae = u' - a' e then u'
= u + (ii' -
ii)e - W(u)e
+ (U'
- a)e
= [W(u) + (a' -
a)]e
18. Up to here this is identical to the parallel construction in consumption theory carried out in Proposition 3.CI. We refer to the proof of the latter for details.
836
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
BARGAINING
----
--
SEC T ION
22•D:
I NVA R, ANCE
origin case.
and therefore, W(u') = W(u) + (ii' - ii) as we wanted. The construction is illustrated in Figure 22.0.2.'9 To prove the second part, suppose that F(') is also invariant to common changes of units. Because F(') is generated from W(·), this can only happen if for every u - u' and {J > 0 we have {Ju - {Ju'. But then u - W(u)e implies {Ju - {JW(u)e, and so W({Ju) = {J W(u) for any u e R' and {J > O. That is, W(·) is homogeneous of degree one, and since g(.) coincides with - W(·) on the domain where ii = 0, we conclude that g(.) is also homogeneous of degree one. _ Going further, if the policy maker is not empowered with the ability to compare the absolute levels of utility across consumers, then the social welfare functional must satisfy more demanding invariance notions. Delinltlon 22.0.5: The social welfare lunctional F: 'fI' -+ [Jt does not allow interpersonal comparisons of utility il F(ii, . ... , ii,) = F(ii; •. ... iii) whenever there are numbers p; > 0 and r1.; such that ii,{x) = p;ii/(x) + r1.; lor all i and x. If the invariance is only with respect to independent changes 01 origin (Le .. we require P; = 1 lor all i). or only with respect to independent changes of units (Le., we require that r1.; = 0 for all i), then we say that F(') is invariant to independent changes of origins or of units. respectively. We have then Proposition 22.0.3. 20 19. We can gain some intuition on the form of this utility function by noticing its similarity to the quasilinear representations in consumer theory. Here we can write any vector u e R' as u = lie + (u - lie) and indifference sets can he obtained by parallel displacements in the direction e. In consumer theory we can write any vector x E RL as x = (x,. 0, ...• 0) + (0. Xl •••• I xd and indifference sels are parallel in the direction (1.0, ...• 0). Similarly. lhe conclusion in both cases is
thatthere is a utility function that is linearly additive in the first lerm (i.e., in the direction in which indifference sets are parallel), 20. See d'Aspremont and aevers (1977) for more results of this type.
0 F
soc I A L
W ELF ARE
proposition 22.0.3: Suppose that the social welfare functional F: 'fI' -+ [Jt can be generated from an increasing. continuous social welfare lunction. If F(·) is invariant to independent changes of origins. then F(') can be generated from a social welfare function W(·) of the purely utilitarian (but possibly nonsymmetric) form. That is, there are constants b; ~ 0, not all zero. such that W(u" ...• u,) =
Flgur. 22.0.2 Construction of the social welfare fUnetion of form (22.D.\) for the invariant to identical changes of
PRO PER TIE S
L b;u;
for all i.
(22.D.2)
Moreover. if F(') is also invariant to independent changes of units [Le., if F(') does not allow lor interpersonal comparisons of utility], then F is dictatorial: There is an agent h such that. for every pair x, VEX, iih(x) >iih(y) implies x Fp(ii, •... , ii,) y. Proof: Suppose that ;::; is the continuous preference relation on R' that generates the given F(·). For a representation of the form (22.0.2) to exist, we require that the indifference sets of ;::; be parallel hyperplanes. Since we already know from Proposition 22.0.2 that those sets are all parallel in the direction e, it suffices to show that they must be hyperplanes, that is, that if we take two u, u· e R' such that u - u'. then for u" = !U + !u· we also have u· - u - u'. The invariance of F(') with respect to independent changes of origins means. in terms of ;::;. that for any r1. e H' we have u + a;::; u" + r1. if and only if u ;::; UN. Take r1. = 1(u' - u). Then u + r1. = u" and u" + r1. = u'. Hence. u;::; u· if and only if UN;::; u'. If u;:: u" then u·;::; u' and so u· - u. If u' > u then u' > u· which contradicts u - u'. We conclude that u· - u - u'. as we wanted. Once we know that indifference sets are parallel hyperplanes. the same construction as in the Proof of Proposition 22.D.2 will give us a W(·) of the form (22.0.2). In addition, the Pareto property yields b, ~ 0 for all i. Finally, suppose that F(') is also invariant to independent changes of units. Then dictatorship follows simply. Choose an agent h with b. > O. Take u, u' e R' with u, > u~. Then, by invariance to independent changes of units. we have that L, b,u, > L, b,u; if and only if b,u, + e L, .. , bl", > b.u. + eLI .. , blu; for any e > O. Therefore, since b,u, > b,u. we get. by choosing e > 0 small enough, that LI blu, > L, b,u;. Thus. agent h is a dictator (show, in Exercise 22.0.4, that in fact b, = 0 for all i >F II). _ We point out that for the dictatorship conclusion of Proposition 22.0.3, it is not necessary that F(') be generated from a social welfare function. It suffices that it be generated from a social preference relation on R'. Proposition 22.0.3 (extended in the manner indicated in the last paragraph) has as a corollary the Arrow impossibility theorem of Chapter 21 (Proposition 2I.C.I), which is, in this manner, obtained by a very different methodology. Indeed, suppose that F(') is a social welfare functional defined. as was done in Chapter 21, on profiles of preference relations (;::;" ... , ;::;,) e !Jl'. Then we can construct a social welfare functional 1"(.) oefined on profiles of utility functions (ii" .... ii,)e'fl' by letting F'(ii" ... , ii,) = F(;::;" .•.• ;::;,). where ;::;, is the preference relation induced by the utility function ii,(·). In Exercise 22.0.5 you are asked to verify, first, that F'(') inherits the Paretian and pairwise independence conditions from F('), second, that F'(') does not allow for interpersonal comparisons of utility and, third. that a dictator for F'(') is a dictator for F(·).
FUN C T , 0 N S
837
838
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
Other invariance properties of social welfare functionals have been found to be of interest. We mention two. We say that the social welfare functional F: 'fI1 .... {It is invariant to common ordinal transformations if F(ii" ... , ii,) = F(ii;, ... , iii) whenever there is an increasing function y(-) such that ii,(x) = y(ii;(x» for every x e X and all i. The interpretation of this invariance is that although the social planner has no notion of individual utility scales she can, nonetheless, recognize that one individual is better off than another (but the question "by how much?" is meaningless). An example is provided by the social welfare functional induced by the symmetric Rawlsian social welfare function W(u) = Min {u" ... , U/}, With this SWF, the ordering Over policies depends only on the ability to determine the worse-off individual (see Exercise 22.0.8 for further elaboration). We say that a social welfare function W(·) generating a given social welfare functional F: '1l' .... R is independent of irrelevant individuals if, when we split the set of agents into any two groups, the social preference among utility vectors in one of the groups is independent of the level at which we fix the utilities of the agents in the other group (we should add that, if so desired, the condition can be formulated directly in terms of the social welfare functional). This is a sensible requirement It says that the distributional judgments concerning the inhabitants of, say, California, should be independent of the individual welfare levels of the inhabitants of, say, Massachusetts. As in the formally similar situation in consumer theory (Exercise 3.G.4), a social welfare function for I > 2 agents that is continuous, increasing, and independent of irrelevant individuals has, up to an increasing transformation, the addilively separable form W(u) = L, g,(u,); that is, W(u) is generalized utilitarian, possible nonsymmetric. Moreover, under weak conditions it is also true that the only social welfare functions that, up to increasing transformations, both admit an additively separable form and are invariant to common changes of origin are the utilitarian W(u) = L, b,u" Thus, from an invariance viewpoint we can arrive at the utilitarian form for a social welfare function by two roads: one, Proposition 22.0.3, is based on invariance to independent changes of origins; the other, just mentioned, is based on independence of irrelevant individuals and invariance to common changes of origins. See Maskin (1978) for more on this. Example 22.0_1: Fix an alternative x* and define a social welfare functional F(') by associating to every profile of individual utility functions (ii., ... , ii,) the social preference relation generated by a utility function Vex) = L, g,(u,(x) - ii~x*». Then, informally, this social welfare functional is both invariant to independent changes of origins and independent of irrelevant individuals, but it is neither utilitarian nor dictatorial. Note, however, that this functional cannot be generated from a social welfare function because it is not pairwise independent the social preference among two alternatives may depend on the ueiliey of ehe lhird afternative x* . •
-- --
BARGAININQ
SECTION
22.E:
THE
AXIOMATIC
BARGAINING
bargaining games (such as those considered in Appendix A of Chapter 9) by adopting an axiomatic point of view. Thus, the approach is more related to ideas of cooperative game theory (as reviewed in Appendix A of Chapter 18).2' For current purposes, the description of a bargaining problem among 1 agents is composed of two elements: a utility possibility set U c H' and a threat, or status-quo, point u* E U. The set U represents the allocations of utility that can be settled on if there is cooperation among the different agents. The point u* is the outcome that will occur if there is a breakdown of cooperation. Note that cooperation requires the unanimous participation of all agents, in which case, to repeat, the available utility options are given by U c R/. If one agent does not participate, then the only possible outcome is the vector 11*. This setup is completely general with two agents and, because of this, the two-agent case is our central reference case in this section. With more than two agents, the assumption is a bit extreme, since we may want to allow for the possibility of partial cooperation. We take up this possibility in Section 22.F. Throughout this section we assume that U c R' is convex and closed and that it satisfies thcfree disposal property U - R I. C U (i.e. if u' ::;; u and u E Uthen u' E U). As in Delinition 22.B.1, U c R' could be generated from a set of underlying 22 alternatives X, which could well include lotteries over deterministic outcomes. For simplicity we also assume that u* is interior to U and that {u E U: u ~ u*} is bounded.
Definition 22.E.l: A bargaining solution is a rule that assigns a solution vector f(V, u*) E V to every bargaining problem (V, u*).21 We devote the rest of this section to a discussion of some of the properties one may want to impose on f(·) and to a presentation of four examples of bargaining solutions: the egalitarian, the utilitarian, the Nash and the Kalai-Smorodinsky solutions. We should emphasize, however, that a strong assumption has already been built into the formalization of our problem: we are implicitly assuming that the solution depends on the set X of feasible alternatives only through the resulting utility values. Definition 22.E.2: The bargaining solution f(·) is independent of utility origins (IUO). or invariant Co independent changes of origins, if for any Il = (Il, ... ,Il,) E H' we have for every i f,(V', u* + Il) = f,(V, u*) + Il, whenever U' = {(u.
+ Il., ...
,u,
+ <X,): u E V}.
The IUO property says that the bargaining solution does not depend on absolute scales of utility. From now on we assume that this property holds. Note that we therefore always have feU, 11*) = feu - {II'}, 0) + II'. This allows us to normalize our problems to 11* = O. From now on we do so and simply write f(U) for f(U, 0).
22_E The Axiomatic Bargaining Approach In this section, we briefly review an alternative approach to the determination of reasonable social compromises. The role of a planner endowed with her own preferences is now replaced by that of an (implicit) arbitrator who tries to distribute the gains from trade or, more generally, from cooperation in a manner that reflects "fairly" the bargaining strength of the different agents. The origin of the theory is game-theoretic. However, it sidesteps the construction of explicit noncooperative
21. For general introductions to the material of this section, see Roth (1979). Moulin (1988). and Thomson (1995). 22. In principle. the underlying set X and Ihe corresponding utility functions on X could be dilTerent for dilTerent U ell'. For the Iheory that follows all that matters is the utility sel U. 23. Thus, a bargaining solution is a choice rule in the sense of Chapter 1. If an underlying alternalive sel X is kepi fixed and. Iherefore, the form of U, generated as in Definition 22.B.I, depends only on the utility functions. we can also regard the bargaining solution as a choice of function in the sense of Definition 2t.E.1.
APPROACH
839
840
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
BARGAINING
---
u,
SECTION
22.E:
THE
AXIOMATIC
BARGAINING
APPROACH
u,
Figure 22.E.2
The property o[ independence of
Symmelric Sel Flgur. 22.E.1
Threal Poinl
It should not be forgotten, however, that a change in the threat point (which will now show up as a change in U) will affect the point settled on. Definition 22.E.3: The bargaining solution f(·) is independent of utility units (IUU). or invariant to independent changes of units. if for any P= (Pt •...• P,) e R' with Pi> 0 for all i. we have fi(U') = PJi(U)
whenever U' = {(PtUt •....
for every i
P,u,): ue U}.24
With independence of utility origins (implicitly assumed in Definition 22.E.3), independence of utility units tells us that, although the bargaining solution uses cardinal information On preferences, it does not in any way involve interpersonal comparisons of utilities. Definition 22.E.4: The bargaining solution f(') satisfies the Pareto property (P). or is Paretian. if. for every U. flU) is a (weak) Pareto optimum. that Is. there is no u e U such that u, > fAU) for every i. Definition 22.E.S: The bargaining solution f(') satiSfies the property of symmetry (S) if whenever U c R' is a symmetric set (Le .. U remains unaltered under permutations of the axes;2S see Figure 22.E.l). we have that all the entries of flU) are equal. The interpretation of the symmetry property is straightforward: if, as reflected in U, all agents are identical, then the gains from cooperation are split equally.
Definition 22.E.6: The bargaining solution f(') satisfies the property of individual rationality (IR) if flU) ~ o. In words: the cooperative solution does not give any agent less than the threat point (recall also that, after normalization, we consider only sets U with 0 e U). It is a sensible property: if some agent got less than zero, then she would do better by opting out and bringing about the breakdown of negotiation. The next property is more substantial. 24. Geomelrically. U' is obtained from U by stretching the different axes by the rescaling factors ((J, ..... {J,). 25. More precisely. if U E U then u' e U for any u' differing from u only by a permutation of ils entries.
'"
The symmetry property [or bargaining solutions.
\'
841
----------------------------------------------------------------------
Threat Point
Definition 22.E.7: The bargaining solution satisfies the property of independence of irrelevant alternatives (IIA) if, whenever U' c U and flU) e U', it follows that flU') = flU).
The I1A condition says that if flU) is the "reasonable" outcome in U and we consider a U' that is smaller than U but retains the feasibility of f(U), that is, we only eliminate from U "irrelevant alternatives," then flU) remains the reasonable outcome (sec Figure 22.E.2). This line of justification would be quite persuasive if we could replace" reasonable" by "best." Indeed, if f( U) has been obtained as the unique maximizer on U of some social welfare function W(u), then the IIA condition is clearly satisfied [if f (U) maximizes W(·) on U then it also maximizes W(·) on U' c U]. We note that while the converse is not true, it is nonetheless the case that, in practice, the interesting examples where I1A is satisfied involve the maximization of some SWF. We proceed to present four examples of bargaining solutions. To avoid repetition, we put on record that all of them satisfy the Paretian, symmetry, and individual rationality properties (as well as, by the formulation itself, the independence of utility origins). You are asked to verify this in Exercise 22.E.I. In Exercise 22.E.2 you are asked to construct examples violating some of these conditions. Example 22.E.1: Egalitarian Solution. At the egalitarian solution fe('), the gains from cooperation are split equally among the agents. That is, for every bargaining problem U c IR t , f..(U) is the vector in the frontier of U with all its coordinates equal. Figure 22.E.3 depicts the case I = 2. Note also that. as illustrated in the figure, every f..(U) maximizes the Rawlsian social welfare function Min {u" ... , ut } on U. The egalitarian solution satisfies the IIA property (verify this). Clearly, for this olution. utility units are comparable across agents. and so the lUU property is not satisfied. 20 • Example 22.E.2: Utilitariall Solution. For every U we now let J.(U) be a maximizer of L, u, on U n IR~. If U is strictly convex, then this point is uniquely defined and. therefore. on the domain of strictly convex bargaining problems the IIA property is satisfied. As witt the previous example. the solution violates the IUU condition. Figure 22.E.4 illustrates the utilitarian solution in the case I = 2. • 26. Do not forget that the utility values arc nol absolute values but rather utility differences from the threal point. It is because of this that changes of origins do not matter.
irrelevant alternatives [or bargaining
solutions.
842
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
ANO
AXIOMATIC
.ARaAININa
---....... u, /
/
45'
//
--
SECTION
22.E:
THE
AXIOMATIC
.ARGAINING
APPROACH
",
Flgur. 22.E.3 (teft) The egalitarian solution for bargaining problems.
---.-_ _f.",.(U)/
Flgur. 22.E.6 Flgur. 22.E.4 (right)
"'-
u,
Threat Point
u,
length [b, f.,(U)] = length [I.(U), aJ
U
, ,
I t
i
I
I
", ~,
-----11 1"2 = constant
Figure 22.E.5
The Nash solution for bargaining problems.
'"
Threat Point
Example 22.E.3: Nash Solution. For this solution, we take a position intermediate between the two previous examples by requiring that f.(V) be the point in V n R/+ that maximizes the product oC utilities u l x ..• x u" or, equivalently, that maximizes Li In U i (this corresponds to the case p = 1 in Example 22.C.4). In Figure 22.E.5, we provide an illustration Cor I = 2.ln this case, the Nash solution has a simple geometry: !.( V) is the boundary point oC V through which we can draw a tangent line with the property that its midpoint in the positive orthant is precisely the given boundary point f.( V); see Exercise 22.EJ. As with the egalitarian and the utilitarian examples, the Nash solution satisfies the IIA property (because it is defined by the maximization oC a strictly concave Cunction). Interestingly, and in contrast to those solutions, the condition of independence of ucility units (IUU) holds for the Nash solution. To see this, note that Li In Ui <: Li In is equivalent to LI In PiUI = LI In UI + LI In PI ~ LI In + LI In PI = Li In Pi"; Cor any constants PI > O. The Nash solution is thereCore invariant to whatever origins or units we wish to fix. It depends only on the cardinal characteristics of the utility functions of the agents over the underlying set of alternatives.
u;
For some rescaling factors ('I" '12) the Nash solution is simultaneously egalitarian and utilitarian.
In u. + In U z = constant
The utilitarian solution for bargaining problems.
u;
There is a way to view the Nash solution as a synthesis of the egalitarian and the utilitarian solutions designed to accomplish the invariance to units: Given a bargaining problem U, the Nash solution is the only utility outcome that, for some rescaling of unils of utility, coincides simultaneously with the utilitarian and the egalitarian solutions. More formally, suppose that
> 0 arc transformation rates of the given units into new units comparable across agents. If the utilitarian and the egalitarian solutions coincide when applied to the rescaled U' (= {('I,ll" ... , ~,II,): (u, •.. . , u,) E U}) then the chosen point U E U must be such that, first, it maximizes Li '1;lI j on U and. second, for some y > 0 it satisfies "IU, = ... = '1,U, = Y. that is, ~, = ,.( I/u,) for every i. Consider now any u' E U. We have L, q,u;!> L, 'I'"' and therefore L,( I/u,)u; !> L, (1/11,)11,. Since (I/u, .... , I/u,) is the gradient of the concave function Li In u, at (,i" ... , ,i,), this implies L, In u; !> L, In Il, (see Section M.C of the Mathematical Appendix). Hence ,i maximizes L, In u, on U, that is, u= f.,(U).". See Figure 22.E.6 for an illustration of the argument. In Exercise 22.E.3 you should show the converse-that the Nash solution is simultaneously utilitarian and egalitarian for appropriate choice of units.
• The Nash solution was proposed by Nash (1950), who also established the notable fact that it is the only solution that satisfies all the conditions so far. Proposition 22.E.l: The Nash solution is the only bargaining solution that is independent of utility origins and units, Paretian, symmetric, and independent of irrelevant alternatives. 2 • Proof: We have already shown in the discussion of Example 22,E.4 that the Nash solution satisfies the properties claimed. To establish the converse, suppose we have a candidate solution f(·) satisfying all the properties. By the independence of utility origins, we can assume, as we have done so far, that f(·) is defined on sets where the threat point has been normalized to the origin. Given now an arbitrary V, let Ii = f.(V) and consider the sets
U' = {II E IR':
L II,/Ii, S; I}
and
27. To rcpc,ll in more geometric terms: the hyperplane with normal ('110 •..• ",) passing through ii leaves U belo,,"' it (because of the utilitarian property). Thus, it suffices to show that the set (/I: L, In II,';; L, In Ii,} lies above the hyperplane. BUI note that this follows from the fact that, because of the egalitarian property. (~" ... , ~,) is proportional to (I/Ii, •... , 1/';,), which is the gradient of the concave function Li In at U. 28. Note that we do not assume individual rationality explicitly: tt turns out to be implied by the other conditions.
"i
843
844
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
--- ---
SECTION
BARGAINING
(left) The N~sh solution is determmed uniquely from . the independell
Figure 22.E.7
",
",
= f.(U')
",
See Figure 22.E.7 for an illustration in the case I = 2. Note that V c V' because the concave function Llln u, has gradient (1Iu ,•...• Ilu,) at U. the point where it reaches its maximum value in the convex set V. The set V" is symmetric and, therefore. by the symmetry and Paretian properties. we conclude that J(V") = (I .... ,I). By the independence of utility units. it then follows that J(V')=(u, ..... u,)=u [observe that u E V" if and only if (u,u, •...• Ii,u,) E V'l Finally. since Ii E V and V c V'. the independence of irrelevant alternatives property yields J(V) = Ii = !.(V). which is the result we wanted. _ Example 22.E.4: Kalai-Smorodinsky Solution. This will be an example of a solution that does not satisfy the independence of irrelevant alternatives property. It was proposed by Kalai and Smorodinsky (1975). Given a bargaining problem VcR'. denote by u'(V) E R the maximum utility value that agent i could attain by means of some vector in V n R'+. See Figure 22.E.8 for the case I = 2. To motivate the solution suppose that agent i has all the bargaining power (i.e .• she can make a take-or-Ieave offer to the remaining agents). Then the outcome would give u'( V) to agent i and nothing to the remaining ones. 29 Hence. we could regard the numbers u'( V) as rough measures of the contributions of the respective agents to the cooperative endeavor and argue. perhaps. that if cooperation takes place then the solution should be the Pareto optimal allocation where the utilities of the different agents are proportional to (u'(V) •...• u'(V»; in other words. where utilities are proportional to the expected utilities that would obtain if we chose with equal probability the agent making a take-or-Ieave offer. This is precisely the KalaiSmorodinsky solution 1.( V). Its construction is indicated in Figure 22.E.8. The Kalai-Smorodinsky solution satisfies the Paretian and the symmetry properties. As with the Nash solution. it does not involve interpersonal comparisons oj utilities. However. it is different from the Nash solution and. therefore (by Proposition 22.E.I), it cannot satisfy the IIA property. In Exercise 22.E.4 you are asked to verify all this. _ 29. We neglect cases where the utility vector that gives ui(U) to agent remaining ones is Pareto dominated.
j
BARGAINING
and nothing to the
APPROACH
845
= J.(U') = f.(U') Fltur.
",
Figure 22.E.B (right)
Threat Point
AXIOMATIC
Pareto, and independence of irreleV3?t alternatives properlles (Proposition 22.E.I).
Midpoinl
",
THE
45'
umts, symmetry,
Threat Point
II.E:
Threat Point
The KalaiSmorodinsky SOIUlio. for bargaining problems.
The conditions summarized so far arc by no means an exhaustive list of the properties that have been found to be of interest in the study of the bargaining problem. We conclude with a few informal comments about some others. (i) Linearity, or decomposability, properties. Suppose that given two bargaining problems, U c R' and V' c R', we consider «V + (I - «)V' c R', for «e [0, I]; we may think of this as, for example, a randomization between the two problems. Then we may want + (I - «)V') = «[(V) + (I - «)[(V'); that is, we may wish that all agents be indifferent between coming to a settlement before or after the resolution of uncertainty. This is a strong requirement, and none of the solutions we have studied satisfies it. In fact. it can be shown that, essentially (i.e., with a few other weak conditions), it is only satisfied by the modified version of the utilitarian solution that docs not impose individual rationality. The same conclusion is reached if we consider V + V' and ask that the overall settlement [(V + V') equals the sequential settlement I(V' + (f(V)}, f(V». Recall that by the 1U0 condition the last expression equals[(V') + [(V).
[(<
(ii) Monotonicity properties. A bargaining solution [(.) is monotone if [(V) S [(V') whenever V c V', that is, if whenever the utility possibility set expands (keeping the threat point fixed at the origin), it is to everyone's advantage. The monotonicity requirement is stronger than may appear at first sight because the utility possibility set may expand in ways that arc very asymmetric across agents. In fact, you can verify in Figure 22.E.9 that neither the utilitarian, nor the Nash, nor the Kalai-Smorodinsky solutions satisfy monotonicity. On the other hand, the egalitarian solution clearly docs. In Exercise 22.E.S you arc invited to verify that the egalitarian solution is essentially the only symmetric and Paretian bargaining solution that satisfies the monotonicity condition. In Exercise 22.E.6 you can also check that the Kalai-Smorodinsky solution for / - 2 is characterized, with the IUO, IUU, P, and S properties, by a certain condition of partial monotonicity. (iii) Consisten, v properties. This type of property concerns the mutual fit of the bargaining solutions when we apply them to problems with different numbers of agents. Let [,(.) be a family of bargaining solutions (one for every set of agents I). Suppose. to be specific, that we start with / = {I, 2, 3}. Take any i. say i - I. and imagine that, conditional on final cooperation, we give agent I the utility level [I (V), but that after making this commitment we reopen the negotiation between the two remaining agents. These two agents then have to find a settlement between themselves in the set V' = {(U2' u,): (f1(V), u., u,) e V} c R2. It is then natural to apply the solution in our family, that is. ['\tIl(V'). Our family is consistent if
22.E.'
Lack of mono tonicity of the utilitarian, Nash, and Kalai-Smorodinsky bargaining solutions.
846
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
!'''''(U') = (f\(U),f~(U)). In words: the renegotiation leads exactly to the outcome of the
initial negotiation. The utilitarian and the Nash solution are consistent (in general, any solution obtained by maximizing a generalized utilitarian SWF is consistent; see Exercise 22.E.7). The Kalai-Smorodinski solution is not (verify this in Exercise 22.E.8). It is an interesting, and nontrivial, fact that the consistency axiom can be used instead of IIA in the characterization of the Nash solution [see Lensberg (1987) and Thomson and Lensberg (1992)].
22.F Coalitional Bargaining: The Shapley Value The analysis of Section 22.E is limited in one important respect: it does not contemplate the possibility of situations intermediate between the full cooperation of all agents and the complete breakdown represented by the threat-point outcome. When there are more than two agents this is definitely restrictive. In this section, we make allowance for the possibility of partial cooperation and discuss how this may influence the eventual distribution of the gains from cooperation. We are given a set 1 of agents and we proceed by specifying a possible set of utility outcomes to every potential subset of cooperators ScI. When S # 1 these utility outcomes are interpreted as a description of what may occur if bargaining breaks down and the members of S end up cooperating among themselves. For the purpose of simplicity, we limit ourselves to the case where utility is comparable across agents (we fix individual units to have the same social utility value) and freely transferable among them. We can then represent the total amount of utility available to the members of ScI, if they cooperate, by a number v(S) or, equivalently, by the utility possibility set {u E IRs: LloS u, ~ vIS)}. In cooperative game theory, which we reviewed in Appendix A of Chapter 18, the rule that assigns vIS) to every ScI is known as a game in characteristic form, a subset ScI is usually referred to as a coalition, and the number v(S) is known as the worth of the coalition S. The situation in which there are no gains from partial cooperation is captured by a characteristic form where v(S) = LIES veil for every S # I. When we interpret the vector (v(t), ... , v(I) as the threat point, this situation reduces to the bargaining problem studied in Section 22.E. In the current world of transferable utility, the egalitarian, Nash, and Kalai-Smorodinsky bargaining solutions JO lead to the same proposal: The gains from cooperation should be split equally among the agents; that is, agent i E 1 should get veil
+ ~ (v(l) I
-
L V(h»).
hEr
In fact, any bargaining solution that is independent of utility origins, Paretian, and symmetric makes this recommendation (Exercise 22.F.t). The question we will try to answer in this section is the following: Assume that, in an environment of games in characteristic form, all the members of 1 decide on cooperation, and therefore on distributing v(I) among themselves. What is then the proper generalization of the equal split solution? It stands to reason that the solution will have to reflect, in some manner, the numbers v(S), ScI, since these incorporate 30. The utilitarian solution is not uniquely defined in the transferable utility case.
BARGAININQ
----........
---
SEC T ION
2 2 • F:
C 0 A LIT ION A L
• A R Q A I N I N G:
THE
• HAP LEY
the information on how valuable the contribution of a particular agent is compared to that of another.
Dellnltlon 22.F.1: Given the set of agents I, a cooperative solution f(·) is a rule that assigns to every game v(·) in characteristic form a utility allocation fly) E RI that is feasible for the entire group, that is, such that LJ,{v) ;5; v(l). Reiterating the analytical strategy of Section 22.E, we continue by stating a number of desirable properties for a solution. The first three are merely variations of properties already encountered above. Dellnltlon 22.F.2: The cooperative solution f(·) is independent of utility origins and of common changes of utility units if, whenever we have two characteristic forms v(·) and v'(·) such that vIS) = pv'(S) + 1:;.5 IX; for every ScI and some numbers IX" ..• ,lXI' and p > 0, it follows that fly) = pt(v') + (IX" ... , IXJl. From now on we assume the property of Definition 22.F.2. Because of it we can, in particular, normalize v(·) to v(i) = 0 for all i. Definition 22.F.3: The cooperative solution f(·) is Paretian if characteristic form v(·).
L; f;(v)
= v(l), for every
Definition 22.F.4: The cooperative solution f(') is symmetric if the following property holds: Suppose that two characteristic forms, v(·) and V(·) differ only by a permutation n: 1 -+ 1 of the names of the agents; that is, VIS) = vIneS)) for all ScI. Then the solution also differs only by this permutation; that Is, f,{v') = f.(,)(v) for all i. The property defined next, in Definition 22.F.5, underlines the fact that we are trying to solve not the welfare-theoretic problem of distributing total utility equitably but rather the more limited problem of distributing equitably the surplus that can be attributed to the cooperation among agents, given the realities of the particular bargaining situation captured by the characteristic form. Definition 22.F.S: The cooperative solution f(·) satisfies the dummy axiom if, for all games v(·) and all agents i such that v(S u {i}) = yeS) for all ScI, we have ,,{v) = v(i) (= 0). In words: If agent i is a dummy (i.e., does not contribute anything to any coalition), then agent i does not receive any share of the surplus. There arc a number of cooperative solutions that satisfy the above properties. The most important is the Shapley value [proposed by Shapley (\953)]. We refer to Appendix A of Chapter 18 on cooperative game theory for examples, motivation, and extended discussion of this concept. Here we limit ourselves to offering a definition. Suppose that we consider an arbitrary ordering of the agents or, formally, an arbitrary permutation n of the names {I, ... , I}. Then g, .•(i) = v({h: nth) ~ n(i)}) - (>({h: nth) < n(i)})
represents how much agent i contributes when she joins the group of her predecessors in the ordering. This is the amount the predecessors would agree to pay i if she had
YAL U E
847
648
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
ANO
AXIOMATIC
all the negotiating power, that is, if she could make a take-it-or-leave-it offer. 31 Note that L, g•.• U) = v(l) for all permutations It. The agents do not come to us ordered. They all stand on the same footing. We may account for this by giving every agent the same chance of being in any position, thereby making all positions equally likely. Equivalently, we could take the (equal weighting) average of agent i contributions over all permutations It (there are I! of these). This is precisely the Shapley value solution. Definition 22.F.6: The Shapley value solution f.<-) is defined by f.,{v) =
~ ~ g •.• (i)
lor every i.
(22.F.1)
It is simple to verify that 1.(.) satisfies all the properties listed so far (see Exercise 22.F.2). For further discussion, see Exercises 22.F.3 and 22.F.4 (and, to repeat, Appendix A of Chapter 18). In Exercise 22.F.S we describe another cooperative solution (the nucleolus).
--- .----
and we could resort to the Shapley value as a way to split C(I) among the different projects. An example may help to clarify the point. Example 22.F.1: This is a favorite example for academics. A professor of economics based in the United States is planning a grand tour of Europe that will take her to three universities, one in Britain (B), one in Spain (S), and one in Germany (G). The total air fare comes to 1600 dollars. How is this to be reimbursed by the three institutions? Suppose that after some research it turns out that C(B) = C(S) = qG) - 800, C(BS) = C(BG) = 1000, and C(SG) = 1400. The Shapley value calculation (carry it out!) then gives c. = 400 and Cs = Co = 600. This split does indeed seem to reflect well the comparative ease of managing a side trip to Britain if already going to some of the other destinations. -
REFERENCES d'Aspremonl.
The Cos/·AlIoca/ion Problem
The following discussion is in the nature of an appendix as there is no immediate conceptual connection with the previous material. The link is that we again make use of the Shapley value. l l Suppose that we have a set 1 of projects and that a policy decision to carry them forward has already been taken. For some reason (e.g., accounting or financing rules), the total cost C(/) must be allocated exactly to the different projects; that is, we must specify (c" ... , c,) such that Li Cj = C(I). What is a reasonable way to determine such cost allocation rules? This is the cost-allocation problem. We must first of all emphasize that, from the point of view of first-best optimality, the cost allocation rules should not be used to guide investment, that is, to decide which projects to carry out. Loosely speaking, we have seen in Section 16.G that the correct efficiency prices for inputs (note that we can think of projects as inputs to the production function for welfare) do not need to cover total cost exactly (see Exercises 22.F.6 and 22.F.7). Because we wish to avoid the temptation to use cost allocation rules in this way we insist that the set of projects to be implemented be exogenously given. An alternative would be to recognize that the cost-covering constraint confers to the welfare problem a second-best nature. That is, we could attempt to maximize social welfare subject to the condition that input (i.e., project) prices must be fixed at levels that exactly cover costs. This approach was taken by Boiteux (1956) in the context of the theory of the regulated firm, and the solution is closely related to Ramsey pricing (see Example 22.B.2 and Exercise 22.F.6). Any welfare·theoretic approach, however, would need to use information on individual preferences. If this cannot, or should not, be done, we are still left with an unresolved the problem. A suggested approach proceeds then as follows: suppose that we have information on the cost of every subset of projects (this is far from an innocuous demand); that is, we know qT) for every Tc I. Then, formally, C(.) is a cooperative game in characteristic form 31. In other words, an offer that, if rejected by some predecessor, would permanently preclude the possibility of agent i or any successor from joining the coalition of the predecessors. To verify the informal claim that the offer will be g•. ,(i), determine how much the last agent in the ordering will get and proceed by backward induction. 32. See Young (1994) for an introductory account to cost-allocation and related problems.
REFERENCES
BARGAININQ
c., and
L. Gevers. (1977). Equity and the informational basis or collective choice. Review
of Economic Studies 44: 199-209. Atkinson. A. (1973). How progressive should income-tax be? In Economic JWliu, Selected Readings,
edited by E. Phelps. London: Penguin Books. Atkinson, A., and J. Stiglitz. (1980). Leclures on Public Economic•. New York: McGraw·HiII. Bergson. A. (1938). A rdormulation of certain aspects of welfare economics. Quarterly Journal of Economic. 51: 310-34. Boi.eux, M. (1956). Sur la gestion des monopoles publiques ..treint. Ii I'Cquilibre budgetaire. Econo,""lrica 14: 22-40. [Translated in Journal of Economic 1Mory (1991) 3: 219-40.] Harsanyi, J. (1955). Cardinal welfare, individual ethics, and interpersonal comparabih.y of utility. Journal of Polilical Economy 61: 309-21. [Also in Phelps (1973).] Guesnerie, R. (1995). A Conlribulion 10 Ihe Pure 1Mary of Taxation, Cambridge, U.K.: Cambridge University Press. Kalai, E., and M. Smorodinsky. (1975). Other solutions to Nash's bargaining problem. Economelrica 43: 513-t8. Laffont, J.•J. (1988). Fundamentals of Public Economics. Cambridge, Mass.: MIT Press. Lipsey, R. C, and K. Lancaster. (1956). The general theory of the second best. Review of Economic Siudies 14: 11-32. t.ensberg. T. (1987). Stability and coUective rationality. Economelrica 55: 935-62. Maskin, E. (t 978). A theorem on utilitarianism. Review of Economic Siudies 42: 93-96. Moulin. H. (1988). Axioms a/Cooperative Decision Making. Cambridge. U.K.: Cambridge University Press. Nash, J. F. (1950). The bargaining problem. Econometrica 18: 155-62. Phelps, E.• ed. (1973). Economic Justice, Selecred Readill/l. London: Penguin Books. Ramsey. F. (1927). A contribution to lhe theory or taxation. Economic Journal 37: 47-61. Rawls. J. (1971). A Theory of Justice. Cambridge. Mass.: Harvard University Press. Roberts. K. (1980). Possibility theorems with interpersonally comparable welrare levels. Review of Economic Studies 47: 409-20. Roth, A. (1979). Axiomalic Models of Bargaining. New York: Springer-Verlag. Samuelson, P. (1'17). Foundalions of Economic Analysis. Cambridge, Mass.: Harvard University Press. Sen. A. (1977). On weights and measures: inrormational constraints in social welrare analysis. Econometrica 45: 1539-72. Shapley, L. (t953). A value for n-person games. In Contributions 10 /he Theory of Ga,"". /I. Annal. of Mathematics Studies. 28. edited by H. Kuhn, and A. Tucker. Princeton. N.J.: Princeton University Press. Starrelt, D. A. (1988). Foundations of Public Economics. Cambridge, U.K.: Cambridge University Press. Thomson. W.. and T. Lensberg. (1992). The Theory of Bargaining with a Variable Number of Agents. Cambridge. U.K.: Cambridge University Press.
849
850
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMIC.
AND
AXIOMATIC
Thomson, W. (1995). Cooperalive models or bargaining. In Handbook olGame Th.ory, Vol. II, edile
---
BARGAINING
II X IRe I • I •
Is it plausible to you? If you wanted to escape from it, how would you do it? What does this all say about the independenoe axiom as applied to social decisions? Suppose now that there are I agents and that in addition to the social utility function U(·) we are also given I individual preference relations ~ defined on the same set of lotteries X. We assume that they are also represented by utility functions of the expected utility form
EXERCISES
for 1-1, .... 1.
22.B.I A Give sufficient conditions for the convexity of the first·best utility possibility set in the context of the exchange economies of Example 22.8.1. 22.B.2A Derive the first·order conditions stated in Example 22.B.2. 22.B.3 A Derive the first·order conditions (22.B.2) of Example 22.B.3. 22.B.4" Show as explicitly as you can that the utility possibility set of Example 22.B.4 may not be convex. 22.C.l A Suppose that the utility possibility set U c R' is symmetric and convex. Show that the social optimum of an increasing, symmetric, strictly concave social welfare function W(·) assigns the same utility values to every agent. [Note: A set U is symmetric if u e U implies u' E U for any u' obtained from u by a permutation of its entries.] Observe that the same conelusion obtains if W(·) is allowed to be just concave, as in the utilitarian case, but U is required to be strictly convex. 22.C.2A Suppose that we contemplate a decision maker in an original position (or ex·ante, or hehind the veil oj ignorance) before the occurrenoe of a state of the world that will determine
which of I possible identities the decision maker will have. There is a finite set X, of possible final outcomes in identity i. Denote X = X, x ... X X,. (a) Appeal to the theory of state·dependent utility presented in Section 6.E to justify a utility function on X of the form
U(x" ... , x,) = ",(x,)
+ ... + u,(x,).
Interpret and discuss the implications of this utility function for the usage of a purely utilitarian social welfare function. (b) Suppose that X, = ... = X, and the preference relation on X defined by the utility function in (a) is symmetric. What does this imply for the form of the utility function? Discuss and interpret. 22.C.3" We have N final social outcomes and we consider a set of alternatives X that is the set oftotteries over these outcomes. An alternative can be represented by the list of probabilities assigned to the different final outcomes, that is, p = (p" . .. , p.) where p. :?; 0 for every nand P,
+ ... + P. = I.
We assume that we are given a social preferenoe relation ~ on X that is continuous and conforms to the independence axiom. Thus, it can be represented by a utility function of the expected utility form U(p) = u,p,
+ ... + ".P •.
From now on we assume that this social utility function U(·) defined on X is given. (0) Suppose that there are two final outcomes and that they are specified by which of two individuals will reoeive a oertain indivisible object. Suppose also that social preferenoes are symmetric in the sense that there is social indifferenoe between the lottery that gives the object to individual 1 for sure and the lottery that gives the object to individual 2 for sure. Show that all the lotteries must then be socially indifferent. Discuss and interpret this conclusion.
851
-----------------------------------------------------------------------
We say that the social utility function U(·) is Parelian if we have U(p) > U(p') whenever U,(p) > U,(p') for every i. (b) Consider a case with N - 3 and I = 2 and illustrate, in the 2-dimensional simplex of lotteries, how the indifferenoe map of the utility functions of the two agents and of the social utility function fit together when the social utility function is Paretian. (c) Exhibit a case where the Paretian condition determines uniquely the social indifferenoe map (recall that we are always assuming the independenoe axiom fot social preferences!). Argue. however. that in general the Paretian condition does not determine uniquely the social indifferenoe map. In fact. exhibit an example where any social utility function is Paretian. (d) Argue (you can restrict yourself to N - 3 and I - 2) that if the social utility function U(p) is Paretian then it can be written in the form U(p) - Il,U,(p)
+ ... + Il,U,(p)
where Il, :?; 0 for every i and Il, oF 0 for some I. What does this conclusion say for the usage of a purely utilitarian social welfare function? Interpret the Il, weights, as well as the fact that they need not be equal across individuals. 22.C.4A The 'eximin ordering, or preference relation, on R' has been mentioned in footnote 11 of this chapter when discussing the Rawlsian SWF. It is formally defined as follows. Given a vector u = (u, •. ..• u,) let u' e A' be the vector that is the nondecreaslng rearrange"",nl or u. That is, the entries of u' are in nondecreasing order and its numerical values (mUltiplicities included) are the same as for u. We then say that the vector u is at least as good as the vector Q in the leximin order if u' is at least as good as Q' in the lexicographic ordering introduced in Example 3.C.1. (a) Interpet the definition ofthe leximin as a refinement ofthe Rawlsian preference relation. (b) Show that the leximin ordering cannot be represented by a utility function. It is enough to show this for I = 2. (c) (Harder) Show that the social optimum of a leximin ordering is a Pareto optimum. You can limit yourself to the case I = 3. 22.C.s B Consider the constant elasticity family of social welfare functions (Example 22.C.4). Argue that w,,(u) .... Min {u, •...• u,} as p .... 00. 22.C.6 A Suppose that U and U' are utility possibility sets and that we associate with them Pareto optimal utility outcomes ii e U and u' e U', respectively. Show graphically that: (8) It is possibk for U' to pass the strong compensation test over U and yet for the outcome with U' to be worse than the outcome with U, as measured by the purely utilitarian SWF.
(b) If the utility possibility sets are derived from a quasilinear economy and U' passes the weak compensation test over U. then it also passes the strong compensation test and, moreover, the outcome for U' is a utilitarian improvement over the outcome for U. Is this conclusion valid if we evaluate social welfare by a nonutilitarian SWF? 22.C.7 B Construct an explicit example of two Edgeworth box economies, differing only in their distributions of the initial endowments. such that the utility possibility set of each one
852
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
EXERCISES
BARGAININQ
passes the weak compensation test over the utility possibility set of the other, when the utility outcome in the laller is chosen to correspond to one of its competitive equilibria.
22.0.3 is valid under the weakened assumption that F is generated from a social preference relation on R I.)
22.C.8 A Suppose we have two utility possibility sets U, U' with respective outcomes U E U and u' E U'. We say that (U', u') passes the Kaldor compensation test over (U, u) if U' passes the weak compensation test over (U, u) and U does not pass the weak compensation test over
12.0.68 This exercise is concerned with social welfare functions satisfying expression (22.0.1).
(U',u').
(a) For I = 2, represent graphically a situation where Kaldor comparability is possible and one where it is not. (b) Observe that Kaldor comparability is asymmetric. Oefine your terms.
22.0.18 In this exercise we verify the indispensability of the assumptions of Proposition 22.0. I.
= lx, y}. The social
(a) Suppose there are three agents and only two alternatives, that is, X welfare functional is given by if and only if
ii,(x)
~
ii,(y) for every i
Y F(ii,. ii,. ii,)x
if and only if
ii,(y)
~
u,(x) for at least one i.
and Check that the social preference relation is always complete, that the social welfare functional cannot be represented by means of a social welfare function, and that only the condition on the number of alternatives fails from Proposition 22.0.1.
X
(e) Argue that if in (22.0. I) the function g(') is homogeneous of degree one and differentiable. then it must be linear (and so we arc back to the utilitarian case). 12.0.78 Consider the constant elasticity family of social welfare functions studied in Example
F(ii,. ii,. ii,)y
(b) Now we have three agents and three alternatives, that is, X = welfare functional is given by
(c) Show that the symmetric Rawlsian social welfare function W(u) = Min lu" •.. , u/l can be wrillen in the form (22.0. I). What about nonsymmetric Rawlsian social welfare functions? [Him: Check the condition of invariance to common changes of origins.] (d) Give other examples satisfying (22.0.1), in particular, examples with g(.) ~ 0 and intermediate between the utilitarian and the Rawlsian cases. Interpret them.
(e) Show that Kaldor comparability may not be transitive.
X
(a) Show that the nonsymmetric utilitarian function W(u) = L b,u , can be wrillen in the form (22.0.1). (b) Show that if W(·) is symmetric and g(O) = 0 then g(') ~ O.
lx, y, z}.
The social
F,(ii" ... , iii) y F,(ii ... , iii) z "
for every (ii" ... , iii) E fli. Show that, again, no representation by means of a social welfare function is possible and that, of the assumptions of Proposition 22.0.1, only the Paretian property fails to be satislied. (c) Exhibit an example in which the only condition of Proposition 22.0.1 that fails to be satisfied is pairwise independence. 22.0.2A Carry out the verification requested in the second paragraph of the proof of Proposition 22.0. I. 22.0.3 A In text. 22.0.4A A social welfare functional F is lexically dictatorial if there is a list of n > 0 agents h" ... , h. such that the strict preference of h, prevails socially, the strict preference of h, prevails among the alternatives for which hi is indifferent, and so on. (a) Show that if F is lexically dictatorial then F is Paretian, is pairwise independent, and does not allow for interpersonal comparisons of utility. (b) Under what conditions can a social welfare functional that is lexically dictatorial be generated from a social welfare function? (c) Show that if a dictatorial social welfare functional is generated from a social welfare function W(u) = L b,u then b, = 0 for every i distinct from the dictator. "
22.0'sc Complete the proof of Arrow's impossibility theorem along the lines suggested in the last paragraph prior to the small-type text at the end of Section 22.0. (Assume that Proposition
22.C.4. (a) Show that the social welfare functionals derived from SWFs in this family are invariant to common changes of units. (b) Show that the only members of this family which arc also invariant to common changes of origins, and therefore admit a representation in the form (22.0.1), are the purely utilitarian (i.e., p = 0) and the Rawlsian (i.e., p = 00). 22.0.8 8 This is an exercise on the property of invariance to common ordinal transformation. (a) Show that the symmetric, Rawlsian social welfare function satisfies the property. (b) Show that the anti-Rawlsian function W(u) = Max lui, ... , u/l also satislies it. (c) Show that the property is satislied for dictatorial social welfare functionals. (d) (Harder) Suppose that I = 2 and W(u) = W(u') for two vectors u, u' E R2, with
u; < u, < u, < ul' Assume also that W(·) is increasing. Show that the induced social welfare functional cannot be invariant to identical ordinal transformations. From this, argue informally (you can do it graphically) that for I = 2 a continuous, increasing social welfare function that is also invariant to identical ordinal transformations must be either dictatorial, Rawlsian, or anti-Rawlsian. 22.E.IA Verify that the bargaining solutions in Examples 22.E.I to 22.E.4 are independent of utility origins, Paretian, symmetric, and individually rational. It is enough if you do so for I = 2. 22.E.2A State nonsymmetric versions of the four bargaining solutions studied in Section 22.E (egalitarian. utilitarian, Nash, and Kalai-Smorodinsky). Motivate them. 22.E.3 8 This is an exercise on the Nash solution. (a) Verify that for I = 2, [.(U) is the boundary point of U through which we can draw a tangent line with th. property that its midpoint in the positive orthant is precisely the given boundary point J.( U). (b) Verify that if U c RI is a bargaining problem then there are rescaling units for the individual utilities with the property that the Nash solution becomes simultaneously egalitarian and utilitarian. 22.E.4A Verify that the Kalai-Smorodinsky solution satisfies the property of independence of utility units but violates the property of independence of irrelevant alternatives. You can restrict yourself to I = 2.
853
854
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
22.E.5 B This is an exercise on Ihe monolonicily property. (0) Show Ihat Ihe egalitarian solulion is the only bargaining solution that is independent of ulilily origins, Paretian, symmetric and monotonic. [Hint: Consider first a family of symmetric ulility possibility sels with linear boundaries. Notice then that for any two sets U, U' we always have U,... U' c U and U ,... U' c U'.]
(b) (Harder) Suppose that f(') is a bargaining solution that is independent of utility origins, Paretian, and strongly monotonic [if U c U' then flU) !> flU') and, in addition, if flU) is interior to U' then f(U)« flU')]. Show that there is a curve in R' starting at the origin and strictly increasing such that, for every U, f(U) is the intersection point of the boundary of U with this curve. You can restrict yourself to the case I = 2. 22.E.6c Let I = 2. A bargaining solution fO is partially monotone if when U c U' and u'(U) = u'(U'), that is, U' expands U only in the direction of agent j ~ i, we have fi( U') :l: fi( U) for j ~ i. Argue that the Kalai-Smorodinsky solution is characterized by the following properties: independence of utility origins and units, Pareto, symmetry, and partial monotonicity. [Hint: use sets U such that U' c U and II'(U) = II'(U'), II'(U) = II'(U')]. 22.E. 7 A Consider a family of bargaining solutions f'(') such that, for every set of agents I, f'(') is independent of utility origins and is generated by maximizing the social welfare function L. g(u,) on normalized bargaining problems U c R', where g(') is increasing, strictly concave, and independent of the particular I considered. Show that the family f' is consistent. 22.E.Sc Show by example that the Kalai-Smorodinsky solution is not consistent. It is enough to consider three agents and its subgroups of two agents. 22.E.9 A This exercise is aimed at showing the independence of the assumptions of Proposition 22.E.1. To this effect, give five examples such that for each of the five assumptions of Proposition 22.E.1 there is one of the examples that violates this assumption but satisfies the
remaining four.
EXERCIIEI
22.F.1 A Show that in the transferable utility case any bargaining solution that is invariant to independent changes of origin, symmetric, and Paretian divides the gains from cooperation equally among the agents. 22.F.2A Show that the Shapley value cooperative solution presented in Section 22.F satisfies the following properties: invariance to independent changes of utility origins, in variance to common changes of utility units, Paretian, symmetry, and the dummy axiom. 22.F.3 A Suppose that for a given set of agents I we take two characteristic forms v and v' and consider their sum v + v'; that is, v + v' is the characteristic form where (v + v'XS) = v(S) + v'(S) for every ScI. (0) Verify that the Shapley value is linear in the characteristic form; that is, /.,(v f,,(v) + f,,(v') for all v, v' and i.
+ v') =
(b) Interpret the linearity property as a postulate that agents are indifferent to the timing of resolution of uncertainty when we randomize among bargaining situations. 22.F.4 c The linearity property of the previous exercise can be restated in a perhaps more intuitive form. We say that a characteristic form v(') is a IInanimity game if for some Tel we have that v(S) = v(T) if T c S, and v(S) = 0 otherwise (thus, the bargaining situations of Section 22.E correspond to T = I). (0) Show that the independence of utility origins and invariance to common changes of utility units, Pareto, symmetry, and dummy axiom properties imply that, for a unanimity game v('), any cooperative solution f(·) assigns the values [,(v) = (1/nv(1) if JET, and J.(v) = 0 otherwise.
(b) We say that the cooperative solution f(·) is weakly linear if for any v and v' differing only by a unanimity game [i.e., there is Tel and «E R such that .'(S) - v(S) + « if T c S, and v'(S) v(S) otherwise] we have that f,(v') f,(v) + «IT if JET, and ft.u) - ft.v) otherwise. Show that if, in addition to the properties listed in (a), the cooperative solution f(') is weakly linear, then it is fully linear, that is, f(v + v') f(v) + f(v') for any two characteristic forms v and v'.
=
=
=
22.E.l0A Give an example of a utilitarian bargaining solution (Example 22.E.2) that violates the property of independence of irrelevant alternatives. [Hint: It suffices to consider I = 2. Also, the violation should involve sets U that are convex but not striclly convex.] 22.E.l1 c Go back to the infinite horizon Rubinstein's bargaining model discussed in the Appendix A to Chapter 9 (specifically, Example 9.AA.2). The only modification is that the two agents are risk averse on the amount of money they get. That is, each has an increasing, concave, differentiable utility function 1I,(m,) on the nonnegative amounts of money that they receive. The factor of discount ~ < I is the same for the two agents. Also 11,(0) = O. The total amount of money is m. (a) Write down the equations for a subgame perfect Nash equilibrium (SPNE) in stationary strategies. Argue that there is a single configuration of utility payoffs that can be obtained as payoffs of a SPNE in stationary strategies. (b) Consider the utility possibility set
U = {(II,(m,), 1I,(m,» E R': m, + m, = m} - R~. Show that if ~ is close to 1 then the payoffs of a SPNE in stationary strategies are nearly equal to the Nash bargaining solution payoffs. (c) (Harder) Argue that every payoff configuration of a SPNE can be obtained as the payoff configuration of a SPNE in stationary strategies. Thus, the uniqueness result presented in Example 9.AA.2 extends to the case in which the agents have strictly concave, possibly different, utility functions for money.
855
--....... -------------------------------------------------------~~~==
BARGAININQ
(c) Show that the Shapley value is the only cooperative solution that satisfies the following prope.'ties: independence of utility origins and in variance to common changes of utility units, Parellan, symmetry, dummy axiom, and linearity. 22.F.sc In this exercise we describe another cooperative solution for a game in characteristic form: the nucleolus. For simplicity we do it for the particular case in which 1= 3, v(1) = v(2) = v(3) = 0, and 0 !> v(S) !> v(l), for any group S of two agents. Given a ulility vector II = (II" II" II,) :l: 0 and an ScI the excess of S at II is e(u, S) = v(S) - L,.s u,. We define the first maximllm excess as m,(II) = Max {_(II, S): 1 < #S < 3}. Choose a two-agent coalition S such that m,(u) = e(u, S). Then we define the second maximum excess as m,(II) = Max {e(II, S'): I < #S' < 3 and S' ~ S}. We say that an exactly feasible [i.e., II, = v(l)] utility profile II = (II" II" 1I,):l: 0 is in thr nucleolus if for any other such profile II' we have either m,(II) < m,(u') or m,(u) = m,(u') and m,(II) !> m,(u').
L,.,
(a) Show that if u = (u" u" u,) is in the nucleolus then either the three excesses for two-agent coalitions are identical or two are identical and the third is larger. (b) Show that there is one and only one utility profile in the nucleolus. [Hint: Argue first that there is a two-agent coalition S such that e(u, S) = m,(u) for every profile in the nucleolus.] From now on we refer to this profile as the nucleolus solution. (c) Argue that the nucleolus solution is symmetric.
CHAPTER 856
CHAPTER
22:
ELEMENTS
OF
WELFARE
ECONOMICS
AND
AXIOMATIC
BARGAININQ
---......
(d) Suppose that agent I is a dummy. Then ", = 0 at the nucleolus solution. (e) Suppose that tv(J):$ v(S) for any coalition S of two agents. Show then that at the nucleolus profile the three excesses for two-agent coalitions are identical.
Incentives and Mechanism
23
(f) Compute and compare the Shapley value and the nucleolus for the characteristic form: v(l) = v(2) = v(3) = 0, v({i, 2}) = v{(l, 3}) = 4, v({2, 3}) = 5, v(J) - 6. (g) Show that if the core is nonempty (see Appendix A to Chapter 18 for the definition of the core in this context) then the nucleolus utility profile belongs to the core.
Design
22.F.6" Consider a regulated firm that produces an output by means of a cost function c(q). Assuming a quasilinear economy, the consumer surplus generated by q is Seq). (a) Suppose that c(q) is strictly concave (i.e., strictly increasing returns to scale). Show that at the first·best price the firm will not cover costs. Conversely, for any q suppose that the price p(q) is determined so that the cost is covered; that is, p(q) = c(q)/q. Show that if q is then determined so as to have p(q) = S'(q), we will not reach the first-best optimum. l11ustrate graphically. (b) Suppose that the quantity produced, q, has to be determined under the constraint that with p = S'(q) we have pq :2: c(q). Solve this second-best weIrare problem. l11ustrate graphically.
(e) Interpret the units of output as "projects." For any production decision q, what is the cost allocation suggested by the Shapley value?
22.F.7 c This exercise is similar to Exercise 22.F.6, except that the firm now produces two outputs under the separable cost functions c,(q,), c,(q,). The surplus S,(q,) + S,(q,) is also separable. (a) The second-best problem [first studied by Boiteux (1956)J is now richer than in Exercise 22.F.6. Suppose that the quantities q" q, have to be determined so that with p, = S'(q,) and p, = S'(q,) we have p,q, + p,q, :2: c,(q,) + c,(q,) (equivalently, at the chosen prices demand must be served and cost covered). Derive first-order conditions for this problem. Make them as similar as possible to the Ramsey formula of Example 22.B.2. (b) (Harder) Interpret the units of outputs as projects. Suppose that these units are very small, so that a given production decision (q" q,)>> 0 represents the implementation of many projects of each of the two types. Can you guess, given (q" q,), what is an approximate value for the cost allocation suggested by the Shapley value? [Hint: For most orderings of projects, any particular project will have preceding it an almost peneet sample of all the projects.J (e) Suppose that for the productions (9" 9,), the Shapley value cost allocation assigns cost per unit of c, and c, (note that "projects" of the same type receive the same cost imputation). Suppose also that c, = oS,(q,)/oq, and c, = oS,(ii,)/oq,. Interpret. Argue that, in general, these productions will not correspond to either the first-best or the second-best optima of the problem.
23.A Introduction In Chapter 21, we studied how individual preferences might be aggregated into social preferences and ultimately into a collective decision. Howe~er, an .im~~rtant !eature of many settings in which collective decisions must be made tS that mdlVlduals actual preferences are not publicly observable. As a result, in one way or another, individuals must be relied upon to reveal this information. In this chapter, we study how this information can be elicited, and the extent to which the information revelation problem constrains the ways in which social decisions can respond to individual preferences. This topic is known as the mechanism design problem.
Mechanism design has many important applications throughout economics. The design of voting procedures, the writing of contracts among parties who will come to have private information, and the construction of procedures for deciding upon public projects or environmental standards are all examples. I The chapter is organized as follows. In Section 23.B, we introduce the mechanism design problem. We begin by illustrating the difficulties introduced by the need to elicit agents' preferences. We also define and discuss the concepts of social choice functions (already introduced in Section 21.E), ex post efficiency, mechanisms, implementation, direct revelation mechanisms, and trutliful implementation. In Section 23.C, we identify the circumstances under which a social choice function can be implemented in dominant strategy equilibria when agents' preferences are private informdtion. Our analysis begins with a formal statement and proof of the revelation principle, a result that tells us that we can restrict attention to direct revelation mechanisms that induce agents to truthfully reveal their preferences. Using this fact, we then study the constraints that the information revelation
I. Simple examples of the last two applications were encountered in Sections 14.C and J I.E, respectively. 857
858
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
----------------------------------------------------------------~
problem puts on the set of implementable social choice functions. We first present the important Gibbard-Satterthwaite theorem, which provides a very negative conclusion for cases in which individual preferences can take unrestricted forms. In the rest of the section, we go on to study the special case of quasilinear environments, discussing in detail Groves-Clarke mechanisms. In Section 23.D, we study implementation in Bayesian Nash equilibria. We begin by discussing the expected externality mechanism as an example of how the weaker Bayesian implementation concept can allow us to implement a wider range of social choice functions than is possible with dominant strategy implementation. We go on to provide a characterization of Bayesian implementable social choice functions for the case in which agents have quasilinear preferences that are linear in their type. As an application of this result, we prove the remarkable revenue equivalence rheorem for auctions. In Section 23.E, we consider the possibility that participation in a mechanism may be voluntary and study how the need to satisfy the resulting parricipation ('onsrraillls limits the set of implementable social choice functions. Here we prove the important Myerson-Satterthwaite theorem, which shows that, under very general conditions. it is impossible to achieve ex post efficiency in bilateral trade settings when agents have private information and trade is voluntary. In Section 23.F, we discuss the welfare comparison of mechanisms. defining the notions of ex ante and interim incentive efficiency, and providing several illustrations of the computation of welfare optimal Bayesian mechanisms. Appendices A and B are devoted to, first, a discussion of the issue of multiple equilibria in mechanism design and. second, the issue of mechanism design when agents know each others' types but the mechanism designer does not (so-called complete information environments). References for further reading are provided at the start of the various sections. We would be remiss. however. not to mention here two early seminal articles: Mirrlees (1971) and Hurwicz (1972).
23.B The Mechanism Design Problem In this section. we provide an introduction to the mechanism design problem that we study in detail in the rest of the chapter. To begin. consider a setting with I agents. indexed by i = I •...• I. These agents must make a collective choice from some set X of possible alternatives. Prior to the choice. however. each agent i privately observes his preferences over the alternatives in X. Formally. we model this by supposing that agent i privately observes a parameter. or signal. Ii, that determines his preferences. We will often refer to Ii, as agent fs rype. The set of possible types for agent i is denoted 0,. Each agent i is assumed to be an expected utility maximizer, whose Bernoulli utility function when he is of type 0, is u,(x. 0,). The ordinal preference relation over pairs of alternatives in X that is associated with utility function u,(x,O,) is denoted ;t,(O,). Agent i's set of possible preference relations over X is therefore given by
fJP, = {;t,: ;t, = ;t,(O,) for some 0, e 0,}. Note that because 0, is observed only by agent i. in the language of Section 8.E
--
SECTION
22.1:
THE
MECHANIIM
DEIION
we are in a setting characterized by incomplete information. As in Section 8.E, we suppose that agents' types are drawn from a commonly known prior distribution. In particular. denoting a profile of the agents' types by 9 = (0 1 , . . . ,0,), the probability density over the possible realizations of 9 e 0 1 x ... X 0, is .p(.). The probability density .p(.) as well as the sets 0 ..... ,0, and the utility functions u;(·.O,) are assumed to be common knowledge among the agents. but the specific value of each agent fs type is observed only by i. 2 Because the agents' preferences depend on the realizations of 0 = (0 1 , •••• 0,). the agents may want the collective decision to depend on O. To capture this dependence formally. we introduce in Definition 23.B.I the notion of a social choice function. a concept already discussed in Section 21.E.3
Dellnltlon 23.B.1: A social choice function is a function f: 0 1 x ... x 0 t - X that. for each possible profile of the agents' types (0 1, ••• • 0,). assigns a collective choice f(Ol.···,O,)eX·
One desirable feature for a social choice function to satisfy is the property of ex post efficiency described in Definition 23.B.2. Dellnltlon 23.B.2: The social choice function f: 0 1 x ... X 0, - X is ex post efficient (or Paretian) if for no profile 0 = (0 1, •.•• 0,) is there an x e X such that Ui(X, Oil ui(f(O), 0i) for some i. Definition 23.B.2 says that a social welfare function is ex post efficient if it selects, for every profile 0 = (0 1, •••• 0,). an alternative f(O) e X that is Pareto optimal given the agents' utility functions UI(·'OI) •...• u,(·.O,). The problem faced by the agents is that the Ois are not publicly observable, and so for the social choice f(OI, .... 0,) to be chosen when the agents' types are (0 1, •••• 0,). each agent i must be relied upon to disclose his type 9,. However. for a given social choice function f(·). an agent may not find it to be in his best interest to reveal this information truthfully. We illustrate this information revelation problem in Examples 23.B.1 through 23.B.4. which range from very abstract to more applied settings.
2. The formulation here is restrictive in one sense: in some settings of interest. agents' preferences over outcomes depend not only on their own observed signals bUI also on signals observed by others (e.g .• agent ts preferences over whether 10 hold a picnic indoors may depend on agent j's knowtedge of likely weather condilions). Through most of this chapler. we focus on the case in which an agent's 'lreferences depend only on his own signal, known as the private values case. We generalize our analysis in Section 23.F. 3. In Section 21.E an agent's type was equivalent to his ordinal preferences over X, and so a social choice function was defined there simply 35 a mapping from Uti x ... X 91, to X. Moreover, it was assumed there that for all i we have {II, = 11, the set of all possible ordinal prderence orderings on X. 4. Two points should be noted about this definition. First. it restricts attention to delerministic social choice funclions. This is largely for exposilional purposes; ahhough much of Ihe chapter considers deterministic social choice functions. in Sections 23.0 to 23.F we allow social choice functions that assign lotreries over X. Second, as in Section 21.E, we limit our attention to singte·vatued choice functions.
PROBLEM
859
860
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIOH
Example 23.B.I: An Abstract Social Choice Selling. In the most abstract case, we are given a set X and, for each agent i, a set fJl l of possible rational preference orderings on X. To consider a very simple example, suppose that X = {x, y, z} and that J = 2. Suppose also that agent I has one possible type, so that 0, = {B,}, and that agent 2 has two possible types, so that 0, = {O'"Oi}. The agents' possible preference orderings fJl l = (;::I(B I )} and fJl 2 = (;::,(8'2), ;::2(02)} are given by
;::1 (B,)
----...
--
SEC T ION
23. B:
THE
ME C HAN I I M
DEli 0 N
z
y
y
Figure 23.B.1
In the social choice runction that selccts a Walrasian equilibrium ror each prererence profile, agent 2 has an incentive to claim to be type when he is really type 8;.
y x
x
[A higher positioned alternative is strictly preferred to a lower positioned one; so, for example, x :>,(B,) y :>,(B I ) z.] Now suppose that the agents wish to implement the ex post efficient social choice function f(·) with
Ao"0:,) = y
and
f(B" 0i) = x.
8,
0,
are (;:: ,(ii,), ;::2 (Oi))], an allocation that he strictly prefers to f(O"Oll when his preferences are ;::2 (0:,). _
If so, then agent 2 must be relied upon to truthfully reveal his preferences. But it is apparent that he will not find it in his interest to do so: When O2 = 0;, agent 2 will wish to lie and claim that his type is 0'2' In abstract social choice settings, a case of central interest arises when fJl l is, for each agent i, equal to fJI, the set of all possible rational preference relations on X. In this case, an agent has many possible false claims that he can make and, intuitively, it may be very difficult for a social choice function always to induce the agents to reveal their preferences truthfully. We will see a formal illustration of this point in Section 23.C when we present the Gibbard-Satterthwaite theorem. _
Example 23.B.3: A Public Project. Consider a situation in which J agents must decide whether to undertake a public project, such as building a bridge, whose cost must be funded by the agents themselves. An outcome is a vector x = (k, I" ... , I,), where k E (0, I} is the decision whether to build the bridge (k = I if the bridge is built, and k = 0 if not), and II E R is a monetary transfer to (or from, if II < 0) agent i. The cost of the project is c;:: 0 and so the set of feasible alternatives for the J agents is
Example 23.B.2: A Pure Exchange Economy. Consider a pure exchange economy with L goods and J consumers in which agent i has consumption set R~ and endowment vector WI = (W'h"" w u )>> o(see Chapter 15). The set of alternatives is
The constraint Lltl ~ -ck reflects the fact that there is no source of outside funding for the agents (so that we must have c + LI I, :s 0 if k = I, and LI II :S 0 if k = 0). We assume that type O;'s Bernoulli utility function has the quasilinear form
X = {(k, I" ... , I,): k E
to, I}, liE R for all i, and L II ~ -ck}. I
X
=
«x" ... ,XI): xIER~
and LXf/ ~ LW(i for I
(=
I, ... L}.
I
In this setting it may be natural to suppose that fJI;, each consumer i's set of possible preference relations over alternatives in X, is a subset of fJl E , the set of individualistic (i.e., depending on XI only), monotone, and convex preference relations on X. To consider a simple example, suppose that J = 2, that consumer I has only one possible type, so that 0, = (B,} and fJI, = (;::, (B,)}, and that for consumer 2 we have;1l2 = iJI E • Imagine then that we try to implement a social choice function that, for each pair (;::, (0,), ;::, (0,)), chooses a Walrasian equilibrium allocation (note that this social choice function is ex post efficient). As Figure 23.B.I illustrates, consumer 2 will not generally find it optimal to reveal his preferences truthfully. In the figure, f(BI,Oll is the unique Walrasian equilibrium when preferences are (;::, (ii,), ;::2 (0'2)) [it is the unique intersection ofthe consumers' offer curves OC, and OC:, occurring at a point other than the endowment point]. However, by claiming that he has type Oi, which has as its offer curve OC;, consumer 2 can obtain the allocation f(OI' 0;) [the unique Walrasian equilibrium allocation when preferences
861
-E-4--------------------------~~
;::2 (0:')
x
PRO B L E II
uI(x, 01) = Olk + (ml + II), where lii l is agent i's initial endowment of the numeraire ("money") and 0, E R. We can then interpret 01 as agent i's willingness to pay for the bridge. In this context, the social choice function f(O) = (k(O), 1,(0), ... ,1,(0» is ex post efficient if, for all 0,
k(O)=t
ifLO,;:: c, I
(23.B.I)
otherwise,
and
L 1,(0) =
-ck(O).
(23.B.2)
I
Suppose that the agents wish to implement a social choice function that satisfies (23.B.I) and (23.B.2) and in which an egalitarian contribution rule is fOllowed, t~at is, in which 11(0) = -(cj I)k(O). To consider a simple example, suppose that 0, = to,} for i '# I (so that all agents other than agent I have preferences that are known) and
862
CHAPTER
23:
'NCENT'VES
AND
MECHAN'SM
DES,GN
IICTION
L,,., e,
L,,., e, L,,..e, -
e,
~I = (c - '0"L e, + &) + tit. - ~I =
(C(l-I
I) _
L
,,..
e, +&) + m•.
Example 23.B.4: Allocation of a Single Unit of an Indivisible Private Good. Consider a setting in which there is a single unit of an indivisible private good to be allocated to one of 1 agents. Monetary transfers can also be made. An outcome here may be represented by a vector x = (y" ... , Y/O t., ... , t,), where y, = I if agent i gets the good, y, = 0 if agent i does not get the good, and t, is the monetary transfer received by agent i. The set of feasible alternatives is then
= {(y"
e
e
(l
But, for & > 0 small enough, this is less than m" which is agent I's utility if he instead claims that 8. = 0, a claim that results in the bridge not being built. Thus, agent I will prefer not to tell the truth. Intuitively, under this allocation rule, when agent I causes the bridge to be built he has a positive externality on the other agents (in the aggregate). Because he fails to internalize this effect, he has an incentive to understate his benefit from the project. _
x
DIIIGN
eo
Agent l's utility in this case is
+ m. -
MICHANII ..
e.
e,
8.
TNI
Two special cases that have received a great deal of attention in the literature deserve mention. The first is the case of bilateral trade. In this case we have I ... 2; agent I is interpreted as the initial owner of the good (the "seller"), and agent 2 is the potential purchaser of the good (the "buyer"). When ~ > there are certain to be gains from trade regardless of the realizations of 8. and 82; when ~, > 2 there are certain to be no gains from trade; finally, if ~2 < and ~, < 2 then there may or may not be gains from trade, depending on the realization of 8. The second special case is the auction setting. Here, one agent, whom we shall designate as agent 0, is interpreted as the seller of the good (the" auctioneer") and is assumed to derive no value from it (more generally, the seller might have a known value 80 = different from zero). The other agents, I, ... , I. are potential buyers (the "bidders").5 To illustrate the problem with information revelation in this example. consider an auction setting with two buyers = 2). In the previous examples, we simplified the discussion of information revelation by assuming that only one agent has more than one possible type. We now suppose instead that both buyers' (privately observed) valuations 8, are drawn independently from the uniform distribution on [0, I] and that this fact is common knowledge among the agents. Consider the social choice function f(8) = (Yo(8), y.(8), Y2(0), 10 (8), t.(8), t 2(8» in which
0, = [0, (0). Suppose also that C> > c(1 - 1)/1. These inequalities imply, first, that with this social choice function agent I's type is critical for whether the it is; if 8, < C it is not), and that the bridge is built (if 8, ~ C sum of the utilities of agents 2, ... , I is strictly greater if the bridge is built under c(1 - 1)/1> OJ. this egalitarian contribution rule than ifit is not built [since Let us examine agent I's incentives for truthfully revealing his type when e, = c - L,o" + & for & > O. If agent I reveals his true preferences, the bridge will be built because
L,,., e,
21 •• :
... ,y"t., ... ,t,):y,E{O, I} and t,EIR for all i, LY'
,
e
i)
=
if 0,
82 ;
= 0 if 8. < O2
(23.8.3)
if 8, < 8 2 ;
= 0 if 8. ~ 8 2
(23.B.4)
yo(8) = 0
for all 8
(23.8.5)
t.(8) = -8.y,(8)
(23.B.6)
t 2 (8) = - 8 2y,(8)
(23.8.7)
t o(8) = -(t.(8)
= I, and LI,;5; OJ.
e,y, + (Iii, + I,),
where 1ft, is once again agent i's initial endowment of the numeraire ("money"). Here IIi E IR can be viewed as agent i's valuation of the good, and we take the set of possible valuations for agent i to be 0, = [Q" 0,] c R. In this situation, a social choice function flO) = (y.(O), ... , y,(O), 1.(8), ... ,1,(0)) is ex post efficient if it always allocates the good to the agent who has the highest valuation (or to one of them if there are several) and if it involves no waste of the numeraire; that is, if for all 0 = (0., ... ,8,) E 0. X ••• x 0/0 y,(O)(O, - Max{O., ... , O,}) = 0
~
h(O) = I Y2(8) = I
+ t 2(8)).
(23.B.8)
In this social choice function. the seller gives the good to the buyer with the highest valuation (to buyer 1 if there is a tie) and this buyer gives the seller a payment equal to his valuation (the other, low-valuation buyer makes no transfer payment to the seller). Note that f(·) is not only ex post efficient but also is very attractive for the seller: if f(·) can be implemented, the seller will capture all of the consumption benefits that are generated by the good. Suppose we try to implement this social choice function. Assume that the buyers are expected utility maximizers. We now ask: If buyer 2 always announces his true value, will buyer I find it optimal to do the same? For each value of 8., buyer l's problem is to choose the valuation to announce, say ~" so as to solve
We suppose that type e;,s Bernoulli utility function takes the quasilinear form u,(x,
PlloallM
863
------------------------------------------------------------------------
Max i,
(II. - ~.) Prob (8 2 ~ ~.)
or Max i,
for all i
and
(8. - ~.)~ •.
S. Note that, for ease of notation, we take there to be I
,
.~
+ 1 agents
in the auction setting.
864
CHAPTER
23:
INCENTIYES
AND
MECHANISM
DESIGN
The solution to this problem has buyer 1 set 01 = Bd2. We see then that if buyer 2 always tells the truth, truth telling is not optimal for buyer I. A similar point applies to buyer 2. Intuitively, for this social choice function, a buyer has an incentive to understate his valuation so as to lower the transfer he must make in the event that he has the highest announced valuation and gets the good. The cost to him of doing this is that he gets the good less often, but this is a cost worth incurring to at least some degree." Thus, we again see that there may be a problem in implementing certain social choice functions in settings in which information is privately held. (For a similar point in the bilateral trade context, see Exercise 23.B.2.) Although buyers have an incentive to lie given the social choice function described in (23.B.3) to (23.B.8), this is not true of all social choice functions in this auction setting. To see this point, suppose we try to implement the social choice function j(.) that has the same allocation rule as that above [i.e., in which the functions y,(.) for i = 0, 1,2 are the same as those described in (23.8.3) to (23.B.5)] but instead has transfer functions 1,(0) = -O,y,(O) I,(B) = -OIY,(O) 10 (0) = -(1,(0)
+ t,(B».
In this social choice function, instead of buyer i paying the seller an amount equal to his own valuation 0, if he wins the object, he now pays OJ, where j ¢' i; that is, he pays an amount equal to the second-highest valuation. Consider buyer I's incentives for truth telling now. If buyer 2 announces his valuation to be 0, $ BI , buyer I can receive a utility of (0, - 0,) ;:: 0 by truthfully announcing that his valuation is B,. For any other announcement, buyer I's resulting utility is either the same (if he announces a valuation of at least 0,) or zero (if he announces a valuation below 0,). So if 0, $ 0" announcing the truth is weakly best for buyer I. On the other hand, if buyer 2's announced valuation is (), > B" then buyer l's utility is 0 if he reveals his true valuation. However, buyer 1 can receive only a negative utility by making a false claim that gets him the good (a claim that his valuation is at least (),). We conclude that truth telling is optimal for buyer I regardless of what buyer 2 announces. Formally, in the language of the theory of games, truth telling is a weakly dominant strategy for buyer I (see Section 8.B). A similar conclusion follows for buyer 2. Thus, this social choice function is implementable even though the buyers' valuations are private information: it suffices to simply ask each buyer to report his type, and then to choose j(O).' • Examples 23.B.I to 23.B.4 suggest that when agents' types are privately observed the information revelation problem may constrain the set of social choice functions that can be successfully implemented. With these examples as motivation, we can now pose the central question that is our focus in this chapter: What social choice junctiolls can be implemented when agellls' types are privale information?
6. This /rade-otT is similar 10 that faced by a monopotist (see Section t2.B): when the monopolist raises his price, he lowers his sales but makes more on his remaining sales. 7. For other examples of implementable social choice [unctions, see Exercise 23.8.1.
-- ----
SECTION
23.8:
THE
MECHAN"M
DESIGN
To answer this question, we need in principle to begin by thinking of all the possible ways in which a social choice function might be implemented. In the above examples we have implicitly imagined a very simple scenario in which each agent i is asked to directly reveal 0, and then, given the announcements (0" ... ,0,), the alternative j(O" ... , 0,) E X is chosen. But this is not the only way a social choice function might be implemented. In particular, a given social choice function might be indirectly implemented by having the agents interact through some type of institution in which there are rules governing the actions the agents may take and hoW these actions translate into a social outcome. To illustrate this point, Examples 23.B.5 and 23.B.6 study two commonly used auction institutions. Example 23.B.5: Firsl-Price Sealed-Bid Auclion. Consider again the auction setting introduced in Example 23.B.4. In a first-price sealed-bid auction each potential buyer i is allowed to submit a sealed bid, b,;:: O. The bids are then opened and the buyer with the highest bid gets the good and pays an amount equal to his bid to the seller.· To be specific, consider again the case where there are two potential buyers (I = 2) and each 0, is independently drawn from the uniform distribution on [0, I]. We will look for an equilibrium in which each buyer's strategy M') takes the form MO,) = CI.,O, for CI., E [0, I]. Suppose that buyer 2's strategy has this form, and consider buyer I's problem. For each B, he wants to solve Max bl
(0, - b,) Prob(b,(B,)
$
b,).
:
Since buyer 2's highest possible bid is CI., (he submits a bid of CI., when 0, = I), it is evident that buyer I should never bid more than cx,. Moreover, since B, is uniformly distributed on [0, I] and b,(B,) $ b, if and only if B, $ (b,/cx,), we can write buyer I's problem as Max
(B, - b,)(bdCl.,)·
bIE{O.1I2J
The solution to this problem is b,(B,) =
{W' CI.,
if tB, $ ietB, >
CI." CI.,.
By similar reasoning, iftB, $ CI." if!B,>CI.,.
to,
Letting CI., = CI., = t, we see that the strategies b,(O,) = for i = 1,2 constitute a Bayesian Nash equilibrium for this auction. Thus, there is a Bayesian Nash equilibrium of this first-price sealed-bid auction that indirectly yields the outcomes specified by the social ~hoice function j(O) = (Yo(B), y,(B), y,(O), t o(9), 1,(0), t,(O»
8. If there are several highest bids, we suppose that the lowest numbered of these bidders gets the good. We could equally well randomize among tho highest bidders if there are more than one, but 'his would require that we expand 'he set of alternatives to A(X), the set of all lotteries over X. In fact. we do precisely this when we sludy auctions in Sections 23.0 and 23.F.
PROBLEM
865
866
C H A , TEA
23:
INC E N T lYE 8
AND
M E C HAN IBM
DE' I Q N
--------------------------~~~~----------------~ in which
y,(O) = I
• E C T ION
23. 8:
THE
M E C HAN
II M
DE' I Q N
a Bayesian game of incomplete information. That is, leUing ii,(sl •...• s" 0,) = u,(g(s, •... ,s,), 0,). the game
(23.B.9)
Yl(/I) = I
if /I, < 0l;
(23.B.l0)
[I, {S,}, {ii,(')},e, x .. · x e"t/>(')]
yo(/I) = 0
for all 0
(23.B.11)
is exactly the type of Bayesian game studied in Section R.E. Note that a mechanism could in principle be a complex dynamic procedure. in which case the elements of the strategy sets S, would consist of contingent plans of action (see Chapter 7)." For the auction setting, the first-price sealed-bid auction is the mechanism in which S, = IR+ for all i and, given the bids (b ..... , b,) E R~. the outcome function g(b" ... , b,) = ({y,(b" ... , b, {t,(b, •.... b, )}I-,) is such that
t,(O) = -!O,y/(O)
(23.B.12)
tl(O) = -tOlY'(O) totO) = -(t,(8)
(23.B.l3)
+ t,(/I».
(23.B.14)
nt-"
•
y,(b, ..... b.>
Example 23.8.6: Second-Price Sealed-Bid Auction" Once again, consider the auction setting described in Example 23.B.4. In a second-price sealed-bid auction each potential buyer i is allowed to submit a sealed bid, b, ~ O. The bids are then ~pened and the buyer with the highest bid gets the good, but now he pays the seller an amount equal to the second-highest bid.'o By reasoning that parallels that at the end of Example 23.B.4, the strategy b,(O.) .= 0, for all /I, E [0, I] is a weakly dominant strategy for each buyer i (see ExerCise 23.B.3). Thus, when I = 2 the second-price sealed-bid auction implements the social choice function flO) = (yo(O). y,(O), Yl(O). t o(8). t,(O), tl(O» in which y,(O) = I
if 0, ~ 0l;
=
0 if 8, < Ol
y,(O) = I
if 0, < 0l;
=
0 if /I, ~ /l l
YolO) = 0
for all /I
=I
if and only if i
= Min{j: b) =
Max{b, ..... bdl.
t,(b, •...• b,) = -b,y,(b l •••• , b,).
In the second-price sealed-bid auction, on the other hand. we have the same strategy sets and functions y,('), but instead t,(b ...... b,) = -Max{b):J I< i}y,(b ..... ,b,). A strategy for agent i in the game of incomplete information created by a mechanism r is a function s,: e, -+ S, giving agent i's choice from S, for each possible type in e, that he might have. Loosely put, we say that a mechanism implements social choice function f(·) if there is an equilibrium of the game induced by the mechanism that yields the same outcomes as f(·) for each possible profile of types 8 = (0" ...• 0,). This is stated formally in Definition 23.B.4. Deflnlllon 23.B.4: The mechanism r = (5, ..... 5,. g(.)) implements social choice function f (.) if there is an equilibrium strategy profile (sH·) ..... s1 (.)) of the game induced by r such that g(s~(O,) •...• 51(/1,)) = f(/I, •...• /I,) for all (/I, .... . 0.) E e, x ... x e,.
1,(0) = -/lly,(/I)
Il(O) = -O'Yl(O) 10(0) = -(1,(/1)
Note, however, that we have not specified in Definition 23.B.4 exactly what we mean by an Mequilibrium". This is because, as we have seen in Part II, there is no single equilibrium concept that is universally agreed upon as the appropriate solution concept for games. As a result, the mechanism design literature has investigated the implementation question for a variety of solution concepts. In Sections 23.C and 23.D we focus on two central solution concepts: dominant strategy equilibrium and Bayesian Nash equilibrium. ll Note also that the notion of implementation that we have adopted in Definition 23.B.4 is in one sense a weak one: in particular, the mechanism r may have more (/'an one equilibrium, but Definition 23.B.4 requires only that one of them induce outcomes in accord with f(·). Implicitly, then, Definition 23.B.4 assumes that. if multiple equilibria exist, the agents will play the equilibrium that the mechanism designer wants. Throughout the chapter we shall stick to this notion of implementation. Appendix A is devoted to a further discussion of this issue.
+ tl(/I» . •
Examples 23.B.5 and 23.B.6 illustrate that, as a general matter, we need to consider not only the possibility of directly implementing social choice functions by asking ag~nts .to ~evea~ their .types but also their indirect implementation through the design of institutions In whIch the agents interact. The formal representation of such an institution is known as a mechanism. Definition 23.B.3: A mechanism r = (5, ..... 5,. g(.)) is a collection of I strategy sets (5, ..... 5,) and an outcome function g: 5, x ... X 51 -+ X. A mechanism can be viewed as an institution with rules governing the procedure for making the collective choice. The allowed actions of each agent i are summarized by the strategy set S,' and the rule for how agents' actions get turned into a social choice is given by the outcome function g('). For~alIy, th.e mechanism r combined with possible types (e" ... , e/)' probabIlIty denSity t/>(.), and Bernoulli utility functions (u,(·), ... , ul (·» defines
I 1. Note also that we are representing the game created by a mechanism using its normal form. For all the analysis Ihat follows in Ihe text this will be sufficient. In Appendix B. however. we consider a case where the extensive form representation is used. 12. Appendix B considers several other equilibrium concepts in the special context of camp/ere information settings in which the players observe each others' types.
9. This auclion is also called a Vickrey aucr;on. afler Vickrey (1961). 10. If there is more than One highest bid. we again select the lowest-numbered of these bidders.
J
, A
0 8 LEM
867
,-----------------------------------------------------
868
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
The identification of all social choice functions that are implementable may seem like a daunting task because, in principle, it appears that we need to consider all possible mechanisms-a very large set. Fortunately, an important result known as the revelation principle (to be formally stated and proven in Sections 23.C and 23.D) tells us that we can often restrict attention to the very simple type of mechanisms that we were implicitly considering at the outset, that is, mechanisms in which each agent is asked to reveal his type, and given the announcements (~" ... ,~,), the alternative chosen is f(~" ... , ~,) E X.13 These are known as direct revelation mechanisms, and formally constitute a special case of the mechanisms of Definition 23.B.3.
--
SECTION
STRATEGY
IMPLEMENTATION
Because of the revelation principle. when we explore in Sections 23.C and 23.D the constraints that incomplete information about types puts on the set of implementable social choice functions, we will be able to restrict our analysis to identifying those social choice functions that can be truthfully implemented. Finally, we note that, in some applications, participation in the mechanism may be voluntary, and so a social choice function must not only induce truthful revelation of information but must also satisfy certain participation (or individual rationality) constraints if it is to be successfully implemented. In Sections 23.C and 23.D, however, we shall abstract from issues of participation to focus exclusively on the information revelation problem. We introduce participation constraints in Section 23.E.
23.C Dominant Strategy Implementation
Definition 23.B.6: The social choice function f (.) is truthfully implementable (or incentive compatible) if the direct revelation mechanism r = (El" ... ,El" f(·J) has an equilibrium (sf(·) .... , s1 (. J) in which s1 (Oi) = 0i for all 0i E Eli and all i = 1•... ,I; that is. if truth telling by each agent i constitutes an equilibrium of r = (El, ....• El" f(·J).
In this section, we study implementation in dominant strategies. '4 Throughout we follow the notation introduced in Section 23.B: The vector of agents' types 0= (0" . .. ,0,) is drawn from the set El = El, x ... x El, according to a probability density t/J('), and agent i's Bernoulli utility function over the alternatives in X given his type 0, is u,(x, 0,). We also adopt the notational convention of writing 0_, = (0 1, ••• ,0,_"0,+,, ... ,0,), 0 = (0" 0_/), and El_/ = El, x ... x El/_ I x Eli+' X • " x El,. A mechanism r = (S" ... ,'Slo g(.)) is a collection of I sets S" . .. ,S" each S/ containing agent i's possible actions (or plans of action), and an outcome function g: S -+ X, where S = S, x ... X S,. As discussed in Section 23.B, a mechanism r = (S, .... , Slo g(')) combined with possible types (El ..... , El,), density t/J('), and Bernoulli utility functions (u,(·), ... , u,(·» defines a Bayesian game of incomplete information (see Section 8.E). We will also often write L, = (5" ... ,S,_"5,+,, ... ,s,), 5 = (5/,5_/), and S_/ = S, X . . . X S,_, X S/+' x .. · X S,. Recall from Section 8.B that a strategy is a weakly dominant strategy for a player in a game if it gives him at least as large a payoff as any of his other possible strategies for every possible strategy that his rivals might play. In the present incomplete information environment, strategy 5/: El, ... S, is a weakly dominant strategy for agent i in mechanism r = (S" ... ,S" g(.» if, for all 0, E El/ and all possible strategies for agentsj '" i, L,(·) = [s,(·), ... ,5,_,('),5,+,('), ... ,51 (,)],1$
To offer a hint as to why we may be able to restrict attention to direct revelation mechanisms that induce truth telling, we briefly verify that the social choice functions that are implemented indirectly through the first-price and second-price sealed-bid auctions of Examples 23.B.5 and 23.B.6 can also be truthfully implemented using a direct revelation mechanism. In fact, for the second-price sealed-bid auction of Example 23.B.6 we have already seen this fact, because the social choice function implemented by the second-price auction is exactly the social choice function that we studied at the cnd of Example 23.B.4 in which truth telling is a weakly dominant strategy for both buyers. Example 23.B.7 considers the first-price sealed-bid auction. Example 23.B.7: Truthful Implementation of the Social Choice Function Implemented by the First-Price Sealed-Bid Auction. When facing the direct revelation mechanism (El" ... , El,,f(·)) with f(O) = (Yo(O), y,(O), Y2(0), totO), t,(O), t 2(0)) satisfying (23.B.9) to (23.B.14), buyer I's optimal announcement ~, when he has type 0, solves Max ;,
DOMINANT
The first-order condition for this problem gives ~, = 0,. So truth telling is buyer l's optimal strategy given that buyer 2 always tells the truth. A similar conclusion follows for buyer 2. Thus, the social choice function implemented by the first-price sealed-bid auction (in a Bayesian Nash equilibrium) can also be truthfully implemented (in a Bayesian Nash equilibrium) through a direct revelation mechanism. That is, the social choice function (23.B.9) to (23.B.l4) is incentive compatible. _
Definition 23.B.S: A direct revelation mechanism is a mechanism in which Si = Eli for all i and g(O) = f(O) for all 0 EEl, x ... x El,. Moreover, as we shall see, the revelation principle also tells us that we can further restrict our attention to direct revelation mechanisms in which truth telling is an optimal strategy for each agent. This fact motivates the notion of truthful implementation that we introduce in Definition 23.B.6 (we are again purposely vague in the definition about the eqUilibrium concept we wish to employ).
23.C:
(0, - t~,) Prob(02 ~ ~,)
E, .[U,(g(5,(O,), 5_,(0 _,»,0,)10,] ;;:: E,Au,(g(§" L,(O _/»,0/)10/]
for all §, E S/. (23.C.I)
or Max ;,
869
-----------------------------------------------------------------------
Condition (23. ':.1) holding for all
(0, - t~,)/l,.
L
,(-l and 0, is equivalent to the condition that,
14. Good sources for fu"her reading on the subject of this section are Dasgupta, Hammond and Maskin (1979) and Green and Larront (1979). IS. The expecta'ion in (23.C.1) is taken over realizations of E (,L,.
13. Some early versions of the revelation principle were derived by Gibbard (1973). Green and Larront (1977). Myerson (1979). and Dasgupta. Hammond and Maskin (1979).
e_,
1
870
c HAP T E R
2 3:
INC E N T lYE.
AND
III E C HAN I • III
0 E• IQ N
----------------------------------------------------------------------for all O,e e .. u,(g(s,(O,), L,), 0,) 2: u,(g(J.. L,), 0,)
(23.C.2)
for aIlJ,eS, and all s_,eS_,.'6This leads to Definition 23.C.1. DefinItIon 23.C.1: The strategy profile s·(·) = (st( .), ... ,s1(')) is a dominant strategy equilibrium of mechanism r = (S" ... ,SI' g(.)) if, for all i and all 01eel' u;(g(s7(01)' L;). 0;) 2: u;(g(si. s _1).0;)
--
• E C T ION
23. C:
DO III I N A III T
• T R AT E Q Y
whether a particular J(.) is truthfully implementable in the sense introduced in Definition 23.C.3.
DefinItion 23.C.3: The social choice function f(') is truthfully implementable in dominant strategies (or dominant strategy incentive compatible. or strategy-proof. or straightforward) if s1(01) = 01 for all O;e e; and i = 1, ... ,I is a dominant strategy equilibrium of the direct revelation mechanism r = (e ,. ... ,e l • f( That is. if for all i and all 0; e e;.
'».
for all sieS;and all s_;eS_;.
ul(f(O;. 0_1). 0;) 2: UI(f(OI. 0_;). ( 1)
We now specialize Definition 23.B.4 to the notion of dominant strategy equilibrium. DefinItion 23.C.2: The mechanism r = (S" ...• SI' g(.)) implements the social choice function f (.) in dominant strategies if there exists a dominant strategy equilibrium of r. s·(·) = (st(·) •... ,si(·)). such thatg(s·(O)) = flO) for allOee. The concept of dominant strategy implementation is of special interest because if we can find a mechanism r = (S" ... ,SI' g('» that implements J(.) in dominant strategies, then this mechanism implements 1(') in a very strong and robust way. This is true in several senses. First, we can feel fairly confident that a rational agent who has a (weakly) dominant strategy will indeed play it." Unlike the equilibrium strategies in Nash-related equilibrium concepts, a player need not correctly forecast his opponents' play to justify his play of a dominant strategy. Second, although we have assumed that the agents know the probability density t/>(.) over realizations of the types (0" ... ,01 ), and hence can deduce the correct conditional probability distribution over realizations of 0_ .. if r implements J(.) in dominant strategies this implementation will be robust even if agents have incorrect, and perhaps even contradictory, beliefs about this distribution. In particular, agent i's beliefs '8 regarding the distribution of 0 _, do not affect the dominance of his strategy Third, it follows that if r implements J(') in dominant strategies then it does so regardless of the probability density t/>( '). Thus, the same mechanism can be used to implement 1(') for any t/>(.). One advantage of this is that if the mechanism designer is an outsider (say, the "government"), he need not know t/>(.) to successfully implement J(. ). As we noted in Section 23.B, to identify whether a particular social choice function Fortunately, it turns out that for dominant strategy implementation it suffices to ask
16. Condilion (23.C.2) follows from (23.C.1) simply by selling ,_,(0_,) = ,_, for all O_,E 0 _,. To see that (23.C.2) implies (23.C.1). consider the case where S_, is a finite set. Then. for any".
Thus. (23.C.2) implies (23.C.I). 17. We leave aside the question of what might happen if an agent has several weakly dominant strategies. This is Ihe issue of multiple equilibria Ihat we discuss in Appendix A. Even so. we at least mention one conclusion from that discussion: The problem of multiple equilibria is relatively small when we arc dealing with dominant strategy equilibrium. 18. In fact. the implementation of 1(' ) using r is also robusl to substantial relaxations of the hypothesis that agents maximize expected utility.
(23.C.3)
lor all (J;e e; and all O_le e_;. The ability to restrict our inquiry, without loss of generality, to the question of whether 1(') is truthfully implementable is a consequence of what is known as the revelation principle Jor dominant strategies. Proposition 23.C.1: (The Revelation Principle for Dominant Strategies) Suppose that there exists a mechanism r = (S" ...• SI' g(.)) that implements the social choice function f(·) in dominant strategies. Then f(') is truthfully implementable in dominant strategies. Proof: If r = (S" ... , SJo g(.» implements 1(') in dominant strategies, then there exists a profile of strategies s·(·) = (sr(·), ... , s1 (.» such that g(s·(O» = J(O) for all and, for all i and all 0, e e"
°
u,(g(s:(O,), L,), 0,) 2: u,(g(J.. L,), 0,)
for all §, e S, and all i and all 0, e e ..
sr (.).
J(.) is implementable. we need, in principle, to consider all possible mechanisms.
1111 P L ElliE N TAT ION
L,
u,(g(sr(O,), s!,(O_,», 0,) 2: u,(g(snb,), s!,(O_,», 0,)
for all (J, e e, and all all i and all 0, e e"
°_, °
(23.C.4)
e S_,. Condition (23.C.4) implies, in particular, that for all
e
e_,. Since g(s·(O»
(23.C.S)
= J(O) for all 0, (23.C.S) means that, for
u,(J(O.. 0_,), 0,) 2: u,(J(b.. 8_,), 0,)
for all (J, e e, and all _I e e _,. But, this is precisely condition (23.C.3), the condition for 1(') to be truthfully implementable in dominant strategies. • The intuitive idea behind the revelation principle for dominant strategies can be put as follows: Suppose that the indirect mechanism r = (S" . . , ,Sh g(.» implements J(.) in dominant strategies, and that in this indirect mechanism each agent i finds playing s~ (0,) when his type is 0, better than playing any other 5, e S, for any choices e S_, by agentsj ¥ i. Now consider altering this mechanism simply by introducing a mediator who says to each agent i: "You tell me your type, and when you say your type is 0" I will play s~(O,) for you." Clearly, if sr(O,) is agent i's optimal choice for each 0, r- e; in the initial mechanism r for any strategies chosen by the other agents. then agent i will find telling the truth to be a dominant strategy in this new scheme. But this means that we have found a way to truthfully implement 1(')' The implication of the revelation principle is that to identify the set of social choice functions that are implementable in dominant strategies, we need only identify those that are truthfully implementable. In principle, for any 1('), this is just a matter of checking the inequalities (23.C.3).
s_,
871
872
--------------------------------------------------------------CHAPTER
23:
INCENTIVEI
AND
MECHANI8M
DEIIGN
SEC T ION
2 3 . C:
0 C MIN A H T
& T RAT E G Y
IMP L E MEN TAT
ION
The inequalities (23.C.3), which are necessary and sufficient for a social choice function f(·) to be truthfully implementable in dominant strategies, can be usefully thought of in terms of a certain weak preference reversal property. In particular, consider any agent i and any pair of possible types for j, and 8';. If truth telling is a dominant strategy for agent i, then for any O_lee_ 1 we must have
e;
uM(O;, 0_,), 0;)
~
uM(Oj, 0_,), 0;)
uM(O;', 0_,), OJ)
~
u,(f(Oj, 0_,), OJ).
/.(0;,8_,) must lie
in shaded set ~.(O:)
°_,)
L,(x, 0,) =
{z e X: u,(x, 0,)
0; 0,
Figure 23.C.2 depicts a change in some agent i's type from to in an exchange setting in which agent i's preferences satisfy the single-crossing properry that we discussed in Sections 13.C and 14.C. In the figure. we denote agent i's allocation in outcome f(O,. O2 ) by J,(O" O2 ), According to Proposition 23.C.2. h(O;'. 0_,) must lie in the shaded region of the figure if truth telling is to be a dominant strategy for agent i. Thus. the characterization in Proposition 23.C.2 can be seen as a multiperson extension of the truth-telling constraints that we encountered in Section 14.C (here they must hold for every possible 0 _, e 0 _,). In the remainder of this section we explore in more detail the characteristics of social choice functions that can be truthfully implemented in dominant strategies.
~ u,(z. 0,»).
Using this lower contour set we get the characterization of the set of social choice functions that can be truthfully implemented in dominant strategies that is given in Proposition 23.C.2.
The Gibbard-Satterthwaite Theorem
Propoaltlon 23.C.2: The social choice function f(') is truthfully implementable in dominant strategies if and only if for all i. all 0_; e <:L;. and all pairs of types for agent i. 0; and OJ e e;, we have ..
f(Oi. 8_;) e L;(f(o;. 0_;). 0;)
and
f(8';. 0_;) e L;(f(Oi. 0_;), Oil.
The Gibbard-Satterwaite theorem was discovered independently in the early 1970s by the two named authors [Gibbard (1973). Satterthwaite (1975)]. It is an impossibility result similar in spirit to Arrow's theorem (Proposition 21.C.1), and has shaped the course of research on incentives and implementation to a great extent. It shows that for a very general class of problems there is no hope of implementing satisfactory social choice functions in dominant strategies. In what follows. we let 9 denote the set of all rational preference relations;:: on X having the property that no two alternatives are indifferent, and we recall that Yf, = {;::,: ;::, = ;::, (0,) for some 0, e e,} is agent i's set of possible ordinal preference relations over X. We denote by fie) the image of f(·); that is,f(e) = {x e X: flO) = x for some 0 e e}. In Definitions 23.C.4 and 23.C.5 we also recall two properties of social choice functions introduced and discussed in Section 21.E.
(23.C.6)
The idea behind this preference reversal characterization of the social choice functions that can be truthfully implemented in dominant strategies is illustrated in Figures 23.C.1 and 23.C.2. In Figure 23.C.l. we represent the social choice function f(·) for each possible configuration of types (8,. ( 2 ) in a situation in which there are two agents (I = 2). two possible values of 0,. and three possible values of 2 , Consider agent I's incentives to tell the truth. If truth telling is a weakly dominant strategy for agent I. then when his type changes from /1', to 11';, he must experience a weak preference reversal between outcomes f(ll',. ( 2 ) and f(O~, ( 2 ) for each possible value of 2 , A similar point applies for agent 2.
°
°
0,
o·, o·,
0;
o·,
f(8',.O;)
f(O'.. o;)
f(O;,o;)
f(U;,O;)
f(U;, 0;)
f(U;, 0;)
0,
Figure 23.C.l
For agent 1 to find truth telling to be his dominant strategy, he must experience a weak preference reversal between outcomes f(O,. 0,) and f(Oj. 02) when his type changes from 8j to OJ, for each possible 0,.
DefinitIon 23.C.4: The social choice function f(') is dictatorial if there is an agent i such that, for aI' 0 = (0, • ... , OJ) e e.
flO) e {x EX: u;(x,
(I;) ~
u;(y,
(I;)
Figure 23.C.2
The weak preference reversal property of Proposition 23.C.2 when preferences satisfy the singlecrossing property.
and
That is, agent i's preference ranking of flO;, 0_,) and f(8';, 0_,) must weakly reverse when his type changes from OJ to OJ, with him weakly preferring alternative flO;, to f(Oj, 0_,) when his type is 0;, but weakly preferring alternative f(8';, 0_,) to flO;, 0_,) when his type is 8';. In the reverse direction, ifthis weak preference reversal property holds for allO_,e 0_, and all pairs 0;, OJ ee" then truth telling is indeed a dominant strategy for agent j (check this in Exercise 23.C.l). This weak preference reversal property can be succinctly stated using agent j's lower contour sets. Define the lower contour set of alternative x when agent j has type 0, by (see Section 3.B):
873
--------------------------------------------------~~~~~=
for all y EX}.
In words: A social choice function is dictatorial if there is an agent i such that f(·) always chooses one of i's top-ranked alternatives.
DefinitIon 23.C.S: The social choice function f(') is monotonic if, for any 0, if 0' is such that L;(f(O), 0;) c L;(f(O), 0i) for all i [i.e., if L;(f(O). 0;) is weakly included in L;(f(O), 0;) for all i], then f(O') = f(O).
874
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
----
Xli
Xli
Monotonicity requires the following: Suppose that flO) = x. and that the I agents' types change to a 0' = (0', •...• 0;) with the property that no agent finds that some alternative that was weakly worse for him than x when his type was 0, becomes strictly preferred to x when his type is 0;. Then x must still be the social choice. The rnonotonicity property is illustrated in Figure 23.C.3 for an exchange setting. In the figure. [,(0,.0.,) represents agent i's allocation in outcome f(O" 0.,). The figure depicts a change in agent i's type from 0, to a 0; having the property that L.(f(O,. 0.,).0,) c L,(f(O,. 0.,).0;). If f(·) is monotonic. then f(o;.O .,) = f(O,. 0.,). With these definitions we now state and prove the Gibbard-Satterthwaite theorem. Proposition 23.C.3: (The Gibbard-Satterthwaite Theorem) Suppose that X is finite and contains at least three elements. that yti = 9 for all i. and that 1(0) = X.'· Then the social choice function 1(') is truthfully implementable in dominant strategies if and only if it is dictatorial. Proof: It is immediate that a dictatorial f(·) is truthfully implementable (check for yourself that every agent will tell the truth). We now show that if f(·) is truthfully implementable in dominant strategies then it must be dictatorial. The argument consists of three steps.
Step 1: If yt, = 9 for all i. and f(·) is truthfully implementable in dominant seracegies. then f(·) is monotonic. Consider two profiles of types 0 and 0' such that L,(f(O). 0,) c L,(f(O). Of) for all i. We want to show that flO') = flO). Let us begin by determining f(O'" O2., .. ,0,). By Proposition 23.C.2 we know that we must have f(o",02, ... ,0,) E L,(f(O),O,). Hence, f(O'" 0" . . . ,0,) E L,(f(O), 0',). But Proposition 23.C.2 also implies that flO) E L,(f(O'" 0" ... ,0,), 0',). Since, by hypothesis, no two alternatives can be indifferent in preference relation <::, (0',), this must imply that f(O'" 0" ... ,0,) = flO). The same line of argumem can be used to show next that f(O'" (J'" 0, • ... , 0,) = flO). Indeed, proceeding iteratively, we establish that flO') = flO). Thus, f(·) is monotonic. Seep 2: efficient.
If yt, = 9 for all i, f(·) is monotonic, and f(0) = X, then f(·) is ex post
19. S'rictly speaking. fini'eness of the set X is not required for the result. But in the absence of finiteness, our assumption that agents are expected utility maximizers may not be compatible with the condi,ion that 91, = fJ' (e.g.• if X = R';. the lexicographic preference relation studied in Example 3.e.1 is a strict preference relation that is not representable by a utility function). For a proof that Proposition 23.C.3 continues to be true if we let X be an arbitrary set and !it, be the set of all continuous preferences on X. see Barbera and Peleg (1990).
Flgur. 2l.C.3 If f(·) is monolonic, then flO;. 0.,) = /(6.. 6.,).
--
SEC T ION
23.
c:
DO MIN ANT
S T RAT E G Y
IMP L E MEN TAT ION
To verify this, suppose not. Then there is a 0 E 0 and ayE X such that u;(y. 0,) > u,(f(O). 0,) for all i (recall that no two alternatives can be indifferent). Because f(0) = X. there exists a O' E 0 such that flO') = y. Now choose a vector of types 0" E 0 such that, for all i, u,(y, 0;') > u;(f(O),O~) > u,(z, Oil for all z >F- flO). y. (Remember that all preferences in 9 are possible.) Since L,(y, 0;) c L,(y, Oil for all i, monotonicity implies that flO") = y. But. since L,(f(O), 0,) c L,(f(O). IJ~) for all i. monotonicity also implies that f(O") = flO): a contradiction because y >F- flO). Hence. f(') must be ex post efficient.
Step 3: A social clooice function f(·) tloar is monotonic and ex post efficient is necessarily dictatorial. Step 3 follows directly from Proposition 21.E.1. Together, steps I to 3 establish the result. _ It should be noted that the conclusion of Proposition 23.C.3 does not follow if
X contains two clements. For example. in this case, a majority voting social choice function (sec Section 21.E) is both nondictatorial and truthfully implementable in dominant strategies (Exercise 23.C.2). Note also that when .iff, = .'1' for all i, any ex post efficient social choice function muse have f(0) = X (verify this in Exercise 23.C.3). Thus, the Gibbard-Satterthwaite theorem tells us that when 91, = ? for all i, and X contains more than two elements. the only ex post efficient social choice functions that are truthfully implementable in dominant strategies are dictatorial social choice functions. Given this negative conclusion, if we are to have any hope of implementing desirable social choice functions, we must either weaken the demands of our implementation concept by accepting implementation by means of less robust equilibrium notions (such as Bayesian Nash equilibria) or we must focus on more restricted environments. In the remainder of this section, we follow the latter course, studying the possibilities for implementing desirable social choice functions in dominant strategies when preferences take a quasilinear form. Section 23.D explores the former possibility: It studies implementation in Bayesian Nash equilibria. Proposition 23.C.3 is readily extended in two ways. First. the resuh's conclusion still follows whenever dt; contains [j' (the set of all rational preference relations having the property that no two alternatives are indifferent). and so it extends to environments in which individual indifference is possible. This is stated formally in Corollary 23.C.1. Corollary 23.C.1: Suppose that X is finite and contains at least three elements. that iJ' c: iii, for all i, and that ((0) = X. Then the sociat choice function f(') is truthfully implementable in dominant strategies if and only if it is dictatorial. Proof: It is again immediate that a dictalorial social choice function is truthfully implementable. We now show that under the stated hypotheses f(·) must be dictatorial if it is truthfully implementable. An implication of Proposition 23.C.3 is that there must be an agent h such that /(0) E {x EX: u,(.<. 0,) 2: u,(y, 0,) for all Y EX} whenever <::,(0,) e iJ' for all i (see Exercise 23.C.4). Without loss of generality, let this be agent 1. Suppose now that the result is not true. Then there is a profile of types 0' E 0 such that f(O') fIx E X: u,(.~. 0;) 2: u,(y. 0;) for all
875
876
C HAP T E R
2 3:
INC E N T I V E SAN 0
M E C HAN ISM
0 E 8 I G H
Y EX}. Let Z E (X EX: "/(X, 0;) '" "/(Y' 11;) for all yE X}. Now consider a profile of types A" E e such that (i) ;::,(0;,) E 9' for all i = I, ... , I; (ii) for all agents i '" I, ",(f(IJ'), OJ) > ",(z, 0;') > ",(x, Oil for all x;. (f(IJ'), .}; and (iii) "/(" OJ) > "1(f(IJ'), OJ) > ",(x, OJ) for all x;. (f(O'), z}. Consider the profile of types (O~, 0'" ... ,0;). By Proposition 23.C.2, we must have flO') E L,(f(Oj, 0'" ... , 0;), O~), and so it must be that I(~, 0'" .. ,' 11;) ~ f(lJ'). The same argument can be applied iteratively for all i '" Ito show that f(~, ... , 1Ii-" 0;) = f(O')· Next, note that (by Proposition 23.C.2) we must have f(O~, ... , OJ_I' 11;) E LI(f(O"), OJ). Hence, f(O") E (t, flO')}. But (by Proposition 23.C.2) we must also have f(O") E LI(j(O;, ... , 0; -,,0;),0;), and since ",(z, 0;) > ",(f(O'), 0',) this means we cannot have flO") = z. Hence, flO") = flO'). But, since ",(z, OJ) > ",(f(O'), 0;,), this contradicts agent J being a dictator whenever ;::,(0,) E 9' for all i . •
--
SEC T ION
23•
c:
0 0 .. I NAN T
S T RAT E G Y
I .. P L E .. E N TAT ION
877
of alternatives is therefore'o
x=
{(k,'" ... ,I,): k e K,I,E R for all i, and
L I, ~ OJ.
Note that this environment encompasses the cases studied in Examples 23.B.3 and 23.B.4: Example 23.el: A Public Projecl. We can fit a generalized version of the public project setting of Example 23.8.3 into the framework outlined above. To do so, let K contain the possible levels of a public project (e.g., if K = (0, I), then either the project is "not done" or "done") and denote by elk) the cost of project level k E K. Suppose that v,(k, 0,) is agent i's gross benefit from project level k and that, in the absence of any other transfers, projects will be financed through equal contribution [i.e., each agent i will pay the amount c(k)/I).l' Then, we can write agent i's nel benefit from project level k when his type is 0, as v,(k, 0,) = v,(k, 0,) - (c(k)//). The t,'s are now transfers over and above the payments e(k)/I . •
As our second extension, we can derive a related dictatorship result for social choice
functions whose image fee) is smaller than X. We first offer Definition 23.C.6. Deflnttlon 23.C.6: The social choice function f(') is dictatorial on set X c X if there exists an age"nt i such that. for all a = (a" ... ,0,) E e, flO) E (x E X: u,(x, 0,) '" u,(y, 0,) for all YEX}.
Example n.e2: AI/ocalion of a Single Unil of an Indivisible Private Good. Consider the environment described in Example 23.B.4 in which an indivisible unit of a private good is to be allocated to one of I agents. Here the "project choice" k = (y, •... , y,) represents the allocation of the private good and K = {(y" ...• Y/): y, E {O, I} for all i and :L.v, = I}. Agent i's valuation function takes the form v,(k, 0,) = (), y,. •
This weaker notion of dictatorship requires only that f(·) select one of the dictator's most preferred ahernatives in X, rather than in X. Corollary 23.C.2: Suppose that X is finite, that the number of elements in fee) is at least three, and that 9' c!il, for all i = 1, ... ,I. Then f(·) is truthfully implementable In dominant strategies if and only if it is dictatorial on the set fee).
A social choice function in this quasilinear environment takes the form f(·) = (k(-), 1 1(-), ••. ,1,(') where, for all 0 e 0. k(O) e K and L,I,(O) ~ O. Note that if the social choice function f(·) is ex post efficient then, for all 0 e 0, k(O) must satisfy
Proof: It is immediate that f(·) is truthfully implementable if it is dictatorial on the sct I(e), and so we now show that under the stated hypotheses f(·) must be dictatorial on sct f(9). If f: e .... X is truthfully implementable in dominant strategies when the sct of alternatives is X, then the social choice function /: 9 .... f(9) which has /(0) = f(8) for all 0 E 8 is truthfully implementable in dominant strategies when the sct of alternatives is f(8). By Corollary 23.C.I, /(.) must be dictatorial. Hence, f(·) is dictatorial on the set 1(9). •
,
L
/
L
v,(k(O), 0,) ~
i-1
v,(k, ()I)
for all k e K.
(23.C.7)
1-1
We begin with a result that identifies a class of social choice functions that satisfy (23.C.7) and that are truthfully implementable in dominant strategies.
The implication flowing from Corollary 23.c.2 is therefore this: When !ii, c 9' for all i, the set of social choice functions which have an image that contains at least three elements and which are truthfully implementable in dominant strategies is exactly the sct of social choice functions that can be implemented (indirectly) by restricting the sct of possible choioes to some subset X c X and assigning a single agent i to choose frcom within this set.
Proposition 23.C.4: Let k*(') be a function satisfying (23.C.7). The social choice function f(') = (k*('), t , (·), ... , til')) is truthfully implementable in dominant strategies if, for all i = 1, ... , I, t,(O) =
[.L. ,,,,,
vj(k*(O), OJ)]
+ hi(O-i)'
(23.C.8)
where h,{-) is an arbitrary function of O-i'
Quasilinear Environments: Groves-Clarke Mechanisms
Proof: If truth is not a dominant strategy for some agent i, then there exist 0" and 0_, such that
In this subsection we focus on the special, but much studied, class of environments in which agents have quasilinear preferences. In particular. an alternative is now a vector x = (k,I" ... ,1/), where k is an element of a finite sct K, to be called the "project choice," and I, e R is a transfer of a numeraire commodity ("money") to agent i. Agent i's utility function takes the quasilinear form u,(x,O,) = v,(k, 0,)
v,(k'(b" 0 _,),0,)
b,.
+ 1,(0,.0_,) > v,(k'(O" 6_,), 0,) + 1,(0" 0_,).
20. Observe that X is not a compact set. This explains what might appear as a small paradox: in this setting, there arc no dictatorial social choice runctions because any agent i. when allowed to pick his best ahernative in X, faces no bound on how much money he can extract from the other
+ (m, + I,),
where m, is agent i's endowment of the numeraire. We assume that we are dealing with a closed system in which the I agents have no outside source of financing. The sei
agents.
21. NOlhing we do depends on this choice tor the "base" method of contribution.
,
J
s
878
c HAP TEA
2 3:
INC E N T lYE SAN D
M E C HAN IS M
Substituting from (23.C.S) for I,(~" 0_ 1) and 1,(0" 0_ 1), this implies that I
L v;(k*(~"
I
L vj(k*(O), OJ),
0_ 1), OJ) >
j<1
j-I
which contradicts k*(·) satisfying (23.C.7). Thus, f(·) must be truthfully implementable in dominant strategies. _ revelation mechanism r = (0 .... . ,0 / '/(')) in which f(·) = II (.» satisfies (23.C. 7) and (23.C.8) is known as a Groves mechanism or Groves scheme [after Groves (1973)].22 In a Groves mechanism, given the announcements 0 _I of agents j '" i, agent i's transfer depends on his announced type only through his announcement's effect on the project choice k*(O). Moreover, the change in agent i's transfer that results when his announcement changes the project decision k is exactly equal to the effect of this change in k on agents j '" i. Put differently, the change in agent i's transfer reflects exactly the externality that he is imposing on the other agents. As a result, agent i is led to internalize this externality and make an announcement-namely, truth-that leads to a level of k that maximizes the I agents' joint payoff from the project, LI v,(k, 0,). A special case of a Groves mechanism was discovered independently by Clarke (1971) and is known as the Clarke, or pivotal, mechanism. This mechanism corresponds to the case in which h,(O_,) = - Lj ", Vj(k!/(O_I), OJ} where, for all 0_ 1 E 0_" k!,(O_I) satisfies A direct
(k('), II (.), ... ,
L
vj(k!,(O_I), OJ) ~
J ",
L
=[ L
vj(k*(O), OJ)J - [
j ",
L
V;(k!I(O_I),Oj)J.
h,(8_,) ~ -
L J"
v,(k!,(O_,),OJ)
IMP L E MEN TAT ION
Proof: Note first that we can always write
L j~
v;(k*(O,. 0 _,). OJ)
+ h,(O,. 0 _,).
(23.C.9)
i
What we want to show. then. is that the function h,(') must in fact be independent of 0, if f(') is truthfully implement able in dominant strategies. Suppose that it is not; that is, that f(·) is truthfully implementable in dominant strategies but that for some 0,.0,. and 0_,. we have 11,(0,.0_,) i' h,({},. 0_,). We now consider two distinct cases. (i) k*(O,. 0 _,) = k*({J,. 0_,): If f(') is truthfully implementable in dominant strategies. then by (23.C3) we have ,',(k*(O,. 0_,). 0,)
+ 1,(0,. 0_,)",
v,(k*(O,. 0_,). 0,)
+ 1,(0,.0_,)", v,(k*(O,. 0_,).0,) + 1,(0, ,0_,).
v,(k*(O,. 0_,). 0,)
+ 1,(0"
0_,)
and
°_,)
Since. k*(O" = k*(O,. 0 _,). these two inequalities imply that 1,(0,.0_,) so by (23.C9) we have 11,(0,.0_,) = h,(O,. 0_,): a contradiction. (ii)
= 1,(0,. 0 _,).
and
k*«(},. () _,) i' k*(O,. 0 _,): Suppose. without loss of generality. that h,(O,. 0_,) > h,(O,. 0_,).
Consider the type
22. We will sometimes be a little loose in our terminology and simply refer to a social choice function f(·) satisfying (ne. 7) and (23.e.8) as a Groves mechanism. 23. Note that the social choice function of the Clarke mechanism satisfies the feasibility condition that L, 1,(0) ~ 0 for all O. Indeed, examining (23.C.8), we see that a sufficient (but nOI necessary) condition for a Groves scheme to satisfy the condition that L, I,{O) ~ 0 for all 0 is that
S T A ATE Q Y
Proposition 23.C.5: Suppose that for each agent i = 1..... 1. {vi(', Oil: 0iE a;} = 'f"; that is. every possible valuation function from K to R arises for some 0iE 0 i . Then a social choice function (.) = (k*('). t,(') .... , td'» in which k*(') satisfies (23.C.7) is truthfully implementable in dominant strategies only if tk) satisfies (23.C.8) for all i = 1..... 1.
j ",
Note that agent i's transfer is 0 if his announcement does not change the project decision relative to what would be ex post efficient for agents j '" i in isolation [i.e., if k*(O) = k!,(O_I)], and is negative if it docs change the project decision [i.e., if k*(O) '" k!,(O _I)], that is, if agent i is "pivotal" to the efficient project choice. Thus, in the Clarke mechanism agent i pays a tax equal to his effect on the other agents if he is pivotal to the project decision, and he pays nothing otherwise.'3
D 0 MIN ANT
We have seen so far that social choice functions satisfying (23.C.7) and (23.C.8) are truthfully implementable in dominant strategies. Are these the only social choice functions satisfying (23.C.7) that are truthfully implementable? The result given in Proposition 23,C.5 [due to Green and Laffont (1979)] provides one set of conditions under which the answer is "yes."" In it. we let 'f" denote the set of all possible functions v: K ---+ R.
for all k E K.
j ",
2 3 • C:
It is interesting to note that in the case of allocation of a single indivisible unit of a private good, the Clarke mechanism is precisely the social choice function implemented by the second-price sealed-bid auction (see Example 23.B.6). In particular: (i) k*(O) is the allocation rule that gives the item to the agent with the highest valuation; (ii) agent i is pivotal precisely when he is the buyer with the highest valuation; and (iii) when he is pivotal his "tax" is exactly equal to the second-highest valuation (in particular, in this case Lj,., v;(k*(O), OJ} = O. and Lj,,' Vj(k'~.,(O_,}. OJ} is equal to the amount of the second-highest valuation}.
1,(0,.0 _,) =
v;(k, OJ)
That is, k!,(O_I) is the project level that would be ex post efficient if there were only the I - I agents j '" i. Agent i's transfer in the Clarke mechanism is then given by t,(O)
SEC T ION
DES I Q N
tn E Eli
such that
-L
v;(k*(O,. 0_,). OJ)
J~'
v,(k,
0:)
=
{
-L
(23.CIO)
v,(k *(0,,0_,). OJ) + r.
J .. i -00
otherwise.
24. For another. see Ihe small-Iype discussion at the end of this section and Exercise 23.e.10.
879
880
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
We wiIl argue that for a sufficiently smaIl t > O. type 01 will strictly prefer to falsely report that he is type 0, when the other agents' types are 0_,. To sec this. note first that k·(o;. 0_,) = k·(O,. 0_,) since setting k = k·(O,. 0_,) maximizes v.(k. 0;) + LI'" vl(k. 01)' Thus. truth teIling being a dominant strategy requires that v,(k·(O,. 0_,). 01)
+ 1,(01. 0_,) ~ v,(k·(O,.
°
_,).0,)
+ 1,(0" 0_,).
or. substituting. from (23.C.9) and (23.C.IO).
But by the logic of part (i). h,(D:.
t
+ h,(o;. 0_,)
t
+ h,(b,.
~
h,(O,. 0_.).
°_,) = h,(O,. °_,) because k·(o;. °_,) = k·(O,. °_,). This gives
°_,)
~ h,(O,.
°_,).
(23.C.ll)
By hypothesis we have h(O,.O_,) > h(b,.o_,). and so (23.C.ll) must be violated for smaIl enough t > O. This completes the proor. _ Thus. when all possible functions v,(·) can arise for some 8, EO,. the only social choice functions satisfying (23.C.7) that are truthfully implementable in dominant strategies are those in the Groves class. Groves mechanisms and budgel balance Up to this point. we have studied whether we can implement in dominant strategies a social choice function that always results in an efficient choice of k [one satisfying (23.C.7)]. But ex post efficiency also requires that none of the numeraire be wasted. that is. that we satisfy the budgel balance condilion: LI,(8)=0 forall8E0.
,
(23.C.12)
We now briefly explore when fully ex post efficient social choice functions [those satisfying bOlh (23.C.7) and (23.C.l2)] can be truthfully implemented in dominant strategies. Unfortunately. in many cases it is impossible to truthfully implement fully ex post efficient social choioe functions in dominant strategies. For example. the result [due to Green and Laffont (1979)] in Proposition 23.C.6. whose proof we omit, shows that if the set of possible types for each agent is sufficiently rich. then no social choice functions that are truthfully implementable in dominant strategies are ex post efficient.2> Proposition 23.C.6: Suppose that for each agent i = 1 •...• I. {vk .8j ): 8j E OJ} = 1'"; that is. every possible valuation function from K to R arises for some 8j E OJ. Then there is no social choice function ((.) = (k·(·). t,(·) •...• t,(·» that is truthfully implementable in dominant strategies and is ex post effiCient. that is. that satisfies (23.C.7) and (23.C.12). Thus. under the hypotheses of Proposition 23.C.6. the presence of private information means that the 1 agents must either accept some waste of the numeraire
---- --
SEC T ION
DO MIN ANT
& T RAT E G Y
IMP L E MEN TAT ION
[I.e.• have L, 1,(8) < 0 for some O. as in the Clarke mechanism] or give up on always having an efficient project selection [i.e .• have a project selection k(8) that does not satisfy (23.C.7) for some OJ. One special case in which a more positive result does obtain arises when there is at least one agent whose preferences are known. For notational purposes. let this agent be denoted "agent 0". and let there still be 1 agents, denoted i = I •... • 1. whose preferences are private information (so that we are now lelling there be 1 + I agents in total). The simplest case of this phenomenon. of course. occurs when agent 0 has no preferences over the project choice k. that is. when his preferences are u,(x) = "'0 + 10, We saw one example of this kind in Example 23.B.4 when we considered auction sellings (agent 0 is then the seller). Another example arises in the case of a public project when the project affects only a subset of the agents in the economy (so that agent 0 represents all of the other agents in the economy). When there is such an agent. ex post efficiency of the social choice function still requires that (23.C.7) be satisfied; but now ex post efficiency is compatible with any transfer functions c,(·) •...• 1,(') for the 1 agents with private information, as long as we set coCO) = - L, .. o 1,(0) for all 8. That is. in this (I + I)-agent selling. the Groves mechanisms identified in Proposition 23.C.4 (in which only agents i = I •... • 1 announce their types) are ex post efficient as long as we set the transfer of agent 0 to be coCO) = - L" ot,(O) for all O. In essence. the presence of an "outside" agent 0 who has no private information allows us to break the budget balance condition for those agents who do have privately observed types. We should offer. however. one immediate caveat to this seemingly positive result: Up to this point. we have not worried about whether agents will find it in their interest to participate in the mechanism. As we will see in Section 23.E. when participation is voluntary. it may be that no ex post efficient social choice function is implementable in dominant strategies even when such an outside agent exists. The differentiable case
It is common in applications to encounter cases in which K = R. the v.<-.O,) functions are assumed to be twice continuously differentiable with o'v,(k. O,)/ok' < 0 and o'v,(k. O,)/ilk 00. oF 0 at all (k.O,). and each 0, is drawn from an interval [Q,. ii,l c R with Q, oF ii,. In this case a great deal can be said about the set of social choice functions that can be truthfully implemented in dominant strategies. Exercise 23.C.9 develops this point fully. Here we will simply show how. in this environment. we can easily derive a number of our previous results. Note first that for any continuously differentiable social choice function J(') = (k(·)., ,(.)•...• c,(·)). if truth telling is a dominant strategy for agent i. then agent i's first-order condition implies that. for all
°_,.
iJv,(k(O,. 0_,). 0,) ok(O,. 0_,) ak 00,
+ a,,(O,. 0_,)
= 0
00,
(23.C.\3)
at all 0, E (Q,. ii,). Integrating (23.C.13) with respect to the variable 0" this implies that for all profiles of types (0,.0_,) we have .(0.0 .)-_ .(0.
l.
25. For another negative result, see the small-type discussion at the end of this section and Exercise 23.C.1O.
23. C:
I'
_I
t.
_11
°
_I
)_l"oV,(k(s.o_,).S)Ok(s.o_')d s. !. ok os
(23.C.\4)
Consider now any social choice function J(-) = (k·(·).!,(·) •...• !,(·)) that satisfies
881
882
CHAPTER
23:
INCENTIVES
AND
MECHANISM
---
OEIION
(23.C.7). Under our present assumptions k·(·) must satisfy. for all O.
t
ov/kO(O). 0) = O. )=, ok
(23.C.15)
',(Q,. 0_,) + L
L
vj(k·(O" 0_,), OJ) -
J ...
i",1
v)(k·(Q" 0_,), 0). I
But this is true if and only if ,,(0) satisfies (23.C.8). Thus, in this selling. Groves mechanisms are the only social choice functions satisfying (23.C.7) that are truthfully implementable in dominant strategies. 26 Consider now the question of budget balance when there is no outside agent. We will show that satisfying (23.C.15) and budget balance is impossible in this differentiable sening when 1=2 [for I > 2 see Laffont and Maskin (1980) and Exercise 23.C.10]. By (23.C.13). for all 0= (0,,0,), we have
E. ,[ui(g(si(Oi)' S~i(O_i»' Oil I0;] ~ E.,[U;(g(§i' S~i(O-i»' 0ill0;]
Definition 23.0.2: The mechanism r = (5" ... ,Sj. g(.» implements the social choice function f(') in Bayesian Nash equilibrium if there is a BayesIan Nash equilibrium of r, s·(·) = (si(·) • ... ,sr such that g(s'(O» = flO) for all E e.
00,
Thus, for all
°
ok
(.».
()',,(O)
_iJ'_v-,-,,(k_O-,:(O~)._O-,-,,)_ ilk_'(_O) o_k_·(_O) o~
il~
+ ov,(k·(O). 0,) o_'_kO_(O_)
00,
ft
OO,il~
(23.C.16)
Delinltlon 23.0.3: The social choice lunction f(') is truthfully implementable in Bayesian Nash equilibrium (or Bayesian incentive compatible) il Sf(Oi) = 0i for al\ O,E i and i = 1•... , J is a Bayesian Nash equilibrium 01 the direct revelation mechanism r = (e" .... e" f(·)). That is. illor all i = 1•... , J and all
and
il',,(O)
00,00,
_o'_v",,(k-:-·~(O-,-,)._O~,) _ok_'(_O) _ok_·(_O) + ov,(kO(O), 0,) _o'_k'_(_O) ok' 00, 00, ok 00,00,'
e
(23.C.17)
O,E
O'V,(kO(O). 0,) ok'
+ o'v,(k·(O). 0,)] ok'(O) ilk'(O) ok'
00,
00,
ei•
(23.D.l)
If we have budget balance. then ,,(0) = -,,(8) for all 0, and so we must have 0',,(8)/00,00, = -0',,(0)/00,00,. But this would imply, by adding (23.C.16) and (23.C.17), and using (23.C.15). that
[
°
As with implementation in dominant strategies (see Section 23.C). we will see that a social choice function is Bayesian implementable if and only if it is truthfully implementable in the sense given in Definition 23.0.3.
00,
= (0" 0,),
00,00,
23.D Bayesian Implementation
for al\ 5i E Si'
ilv,(k·(O). 0,) ok'(O)
00,
IMPLEMENTATION
e
and
0,,(0)
BAYESIAN
Definition 23.0.1: The strategy profile s·(·) = (si(·) •.... si(·» Is a Bayesian Nash equilibrium of mechanism r = (5" ...• 5,. g('» if, for all i and all O,E i •
ov,(k·(O). 0,) ok·(O) ok
23.0:
In this section. we study implementation in Bayesian Nash equilibrium. 17 Throughout we follow the notation introduced in Section 23.B: The vector of agents' types 0=(0, •...• 0,) is drawn from set e=e, x· .. xe, according to probability density
Moreover. using Ihe implicit function theorem and our assumptions on the v,(·) functions. we See that k·(·) is continuously differentiable and that it has nonzero partial derivatives. elk·( 0)/ elO, 10 0 for all i. We now substitute for iJv,(k·(s, 0_,). s)/ok in (23.C.14) using (23.C.15). Doing so. we derive that, for all profiles (0,,0_,),
=
SECTION
for al\
0, E e i .
The ability to restrict our inquiry. without loss of generality. to the question of whether f(·) is truthfully implementable is a consequence of the revelation principle for Bayesian Nash equilibrium.
= 0
•
which is impossible under our assumptions. 27. Good sources for fUrlher reading on 'he subject of this section are Myerson (1991) and Fudenberg and Tirole (1991). 28. As in Section 8.E, we restrict our attention to pure strategy equilibria.
26. This argument generalizes to any case in which k·(·) is continuously differentiable.
j
883
-----------------------------------------------------------------------
884
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIQN
Proposition 23.0.1: (The Revelation Principle for Bayesian Nash Equilibrium) Suppose
that there exists a mechanism r = (S, . .... S,. g(.» that Implements the social choice function f(') in Bayesian Nash equilibrium. Then f(') is truthfully implementable in Bayesian Nash equilibrium.
-- --
Proof: If r = (S" ...• s,. g(.» implements f(·) in Bayesian Nash equilibrium, then there exists a profile of strategies s*(·) = (sr(·) • ...• sj (.)) such that g(s*(O» = flO) for all O. and for all i and all 0, EEl,.
£ •. ,[",(9(5:(0,). s!.,(O_,». 0,)18,] ;:: £ •. ,[U,(g(5,. s!.,(O_,). 0,)10,]
(23.0.2)
for all 5, E S,. Condition (23.0.2) implies. in particular. that for all i and all 0, E El ..
£ •. ,[",(9(S:(O,). s!.,(O_,)). 0,)10,] ;:: £ •. ,[u,(y(st(b,). s!.,(O_,». 0,)10,]
(23.0.3)
for all 0, EEl,. Since g(s*(O» = f(8) for all 8. (23.0.3) means that. for all i and all O,EEl,. (23.0.4) for all 0, EEl,. But. this is precisely condition (23.0.1). the condition for f(-) to be truthfully implementable in Bayesian Nash equilibrium. _ The basic idea behind the revelation principle for Bayesian Nash equilibrium parallels that given for the revelation principle for dominant strategy implementation (Proposition 23.C.t): If in mechanism r = (S, •...• S,. y(.)) each agent finds that. when his type is 0,. choosing sf(8,) is his best response to the other agents' strategies. then if we introduce a mediator who says "Tell me your type. 8" and I will play sf(O,) for you." each agent will find truth telling to be an optimal strategy given that all other agents tell the truth. That is. truth telling will be a Bayesian Nash equilibrium of this direct revelation game. The implication of the revelation principle is. once again, that to identify the set of implementable social choice functions (now in Bayesian Nash equilibrium) we need only identify those that are truthfully implementable. 19 We can note immediately that the Bayesian implementation concept is a strictly weaker notion than the notion of dominant strategy implementation. Since every dominant strategy equilibrium is necessarily a Bayesian Nash equilibrium. any social choice function that is implementable in dominant strategies is a fortiori implementable in Bayesian Nash equilibrium. Intuitively put, when we compare the requirements for truthful implementation of a social choice function f(·) in dominant strategies and in Bayesian Nash equilibrium given in equations (23.C.3) and (23.0.1). respectively. we see that. with Bayesian implementation. truth telling need only give agent i his highest payoff averaging over all possible types 0_, that might arise for Ihe oIlier agenes. In contrast. the dominant strategy concept requires that truth telling be agent i's best strategy for every possible 8 _,. Thus. we can reasonably hope to be
SECTION
23.0:
BAYEStAN
IMPLEMENTATION
able to successfully implement a wider range of social choice functions in Bayesian Nash equilibrium than in dominant strategies. The drawback. of course, is that we can be less confident about this implementation relative to implementation in dominant strategies because it depends on the agents (and any outside mechanism designer) knowing the density
The Expected Externality Mechanism Let us return to the quasilinear setting studied in Section 23.C. In particular. an alternative is now a vector x = (k. I, •.... I,). where k is an element of a finite set K. and I, E IR is a transfer of a numeraire commodity ("money") to agent i. Agent i's utility function takes the quasilinear form
u,(x. 0,) = v,(k. 0,) + (m, + I,).
(23.0.5)
,ii,
where is agent i's endowment of the numeraire. lo For simplicity. we shall henceforth normalize If"~ = 0 for all i. We assume here that the I agents have no outside source of financing. and so X = {(k.I, •... • 1,): k E K. I, E IR for all i. and :L I,:S; O}. A social choice function in this environment takes the form f(·) = (k(·). 1, (.) •...• 1,(' ). Note that if f(-) is ex post efficient then, for all 0 E El.
L v,(k(O). 0,) ~ L v,(k. 0,)
for all k E K
(23.0.6)
and LI;((}) =0.
(23.D.7)
In Proposition 23.C.6 we saw that conditions exist in which no social choice 29. Note that Proposition 23.0.1 is what we implicitly relied on in Section 14.C when. in sludying the principal-agent problem with hidden infonnation, we restricted our focus to direct revelalion mechanisms that induced truth telling by the agent. Formally. Proposition 23.0.1 tells us that the equilibrium outcome arising from any contract between the principal and the agent can be replicated using a direct revelation mechanism that induces the agent to truthfully reveal his type.
30. Unlike the analysis in Section 23.C (see Exercise 23.C.11). the developments that follow depend not only on preferences over certain outcomes having a quasilinear form but also on the fact that. with this Bernoulli utility function, each agent i is risk neutral with respect to lotteries
over his monetary transfer .
.J.
885
886
CHAPTER
2':
INCENTIVES
AND
MECHANISM
DESION
SECTION
23.D:
.AYESIAN
IMPLEMENTATION
887
~-------------------------------------------------------------------
function f(·) = (k(·), t l (·), • •• , t,(·» satisfying both (23.0.6) and (23.0.7) is truth. fully implementable in dominant strategies. We will now show that it Is possible to implement such a social choice function in Bayesian Nash equilibrium whenever the agents' types are statistically independent of one another [Le., when the density
[ L
vj(k·(O"
0_,), OJ)J + h,(O_,),
Intuitively, the form of the h,(') functions in (23.D.9) can be thought of as follows: We have seen that when the agents' types are (0 1, ••• ,0,), each agent I = I, ... ,I receives a payment equal to ~,(O,) [the first term in (23.0.8»). Now, if each agent contributes an equal 1/(1 - I) share of all of the other agents' payments, the payments from a given agent i to each of the other I - I agents will total [1/(/ - I)] Lj,,' ~j(Oj)' and agent i will receive from these agents in return payments that total to ~,(O,). Agent i's net transfer will therefore be ~,(O,) - (1/(/ Lj,,' ~J..OJ)' This direct revelation mechanism is known as the expected externality mechanism [due to d'Aspremont and Gerard-Yaret (1979) and Arrow (1979)]. In summary, we have shown that when agents' Bernoulli utility functions take the form in (23.0.5), and agents' types are statistically independent, there is an ex post efficient social choice function that is implementable in Bayesian Nash equilibrium. Although this is an interesting result, it is not the end of the story, even when we restrict our attention to Bernoulli utility functions of the form (23.0.5) and a statistically independent distribution of types. The reason is that while the expected externality mechanism implements an ex post efficient social choice function, its transfer functions imply a particular distribution of utility across the various types of the agents, and we may wish to consider other mechanisms, possibly ones involving social choice functions that are not ex post efficient, that alter this distribution. One reason why this may be important is that, in many applications of interest, agents are free to opt out of the mechanism, and so any mechanism that we wish to implement must not only be incentive compatible in the sense that we have studied so far, but must also satisfy individual rationality (or participation) constraints that assure that each agent i actually wishes to participate in the mechanism. If the expected externality mechanism does not satisfy these constraints, we will need to consider other mechanisms that do. We will have more to say about this issue in Sections 23.E and 23.F, but for now suffice it to say that for this reason, as well as others, we may be interested in identifying all of the social choice functions that are Bayesian implementable in this environment. In the remainder of this section we do this for the special, but often-studied, class of cases in which agents' preferences take a form that is linear in their type, and their types are independently distributed.
I»
(23.0.8)
j'"
where, for now, we take h,(') to be an arbitrary function of 0 _I' Note that the expectational term in (23.0.8) represents the expected benefits of agents j ¥ i when agent i announces his type to be 0, and agents j ¥ I tell the truth. (As such, it is a function of only agent i's actual announcement OJ-it is not a function of the actual announcements 0_, of agents j ¥ I.) Thus, the change in agent I's transfer when he changes his announced type is exactly equal to the expected externality of this change on agents j ¥ I. We check first that any social choice function 1(,) with the form (23.D.8) is Bayesian incentive compatible. To see this, note that when agents j ¥ I announce their types truthfully, agent I finds that truth telling is his optimal strategy because (using statistical independence of 0, and 0_,) Eo_,[v,(k"(O),O;)
+ t,(O)IO,]
= E._,[
t t
vj(k·(O), OJ)] + E._,[h,(O_,)]
j-I
~ E•. , [ J-I vj(k·(b" 0_,), OJ)] + Eo_, [h,(O_,)] = E._,[v,(k·(b" 0_,), OJ)
+ t,(b" 0_,)10.]
for all 0, E 0" where the inequality follows because k*(·) satisfies (23.0.6). What remains is to show that we can choose the hk) functions (for I = I, ... , I) so that we also satisfy the budget balance condition (23.0.7). For notational ease, define ~,(O,) = Ei.,[Lj",v/k·(0,,9_,),9 j We now let
».
h,(O_,) =
-C ~ I) J~' ~j(Oj)'
(23.0.9)
Bayesian Incelltive Compatibility with Linear Utility
for i = I, ... ,I. With this choice for the h,(') functions, we have
Suppose now that each agent i's Bernoulli utility function takes the form
L t,(O) = L e,(O,) + L h,(O _I)
u,(x, 0,) = O,r,(k)
i i i
=
L ~,(O,) -
(-I 1 ) L L ~/Oj) - I 'J'"
L, ~,(O,) -
(-I 1 ) L (/ - l)~,(0,) -I ,
,
=
+ (nl, + t,).
As before, we shall normalize Iii, = 0 for all i. We also suppose that each agent i's type lies in an interval 0, = [~"lJ,] c IR with ~, ¥ 0" and that the agents' types are statistically independent. We let the distribution function of 0, be denoted,('), and we assume that it has an associated density 0 for all 0, E [~" 0,]. We begin by deriving a necessary and sufficient condition for a social choice function fe) = (k(·), t 1('), .... t,(·» to be Bayesian incentive compatible. It is convenient to define 1,(0,) = E._Jt,(O" 0_,)); this is agent i's expected transfer given that he announces his type to be (j, and that all agents j ¥ i truthfully reveal their types. Likewise, we let [\(0,) = E._ ,[1"(k(O,, 0 .1))] denote agent i's expected "benefit"
=0.
31. See Fudenberg and Tirole (1991) for a discussion of the case of correlated Iypes and for rurther rererences.
J
888
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
conditional on announcing b,. Because of the form of agents' utility functions, we can write agent i's expected utility when he is type 0, and announces his type to be 0, (assuming that all agents j .;. i tell the truth) asH
E._.[u,(f{O" 0_,), 0.)10;] = O,v,{O,)
+ 1,(8,).
(23.0.10)
+ 1,(0,),
giving agent i's expected utility from the mechanism conditional on his type being 0, when he and all other agents report their true types. Proposition 23.0.2: The social choice function f(·) = (k(·). t,(·) •...• t,(·» is Bayesian incentive compatible il and only if. for all i = 1..... I.
f'·
(i) v,(·) is nondecreasing. (ii) VitO,) = Vi(Q,)
+
(23.D.11)
lor all 0,.
v,(s) ds
23.0:
BAYESIAN
IMPLEMENTATION
Hence.
U;(O,) ~ U,(b,) + (O, - O,)ii,(O,) = O,v,(O,) + 1,(0,). Similarly. we can derive that
U,(O,) ~ U,{O,)
+ (0, - O,)v,(O,) = O,v,(O,) + 1,(0,).
So I{') is Bayesian incentive compatible. _
It is also convenient to define for each i the function
U,{O,) = O,v,(O,)
---
SECTION
(23.D.12)
!.
Proof: (i) Necessity. Bayesian incentive compatibility implies that for each have
0, > 0, we
and
Proposition 23.0.2 shows that to identify all Bayesian incentive compatible social choice functions in the linear setting. we can proceed as follows: First identify which functions k( .) lead every agent i's expected benefit function v,(') to be nondecreasing. Then. for each such function. identify the expected transfer functions I, (.)..... I, (.) that satisfy condition (23.0.12) of the proposition. Substituting for U,(·). these are precisely the expected transfer functions that satisfy. for i = I, ...• I.
1,(0,) = 1M,)
+ ~,v,(~,) - O,v,(O,) +
f" v,(s) ds ~,
for some constant l,{~,). Finally. choose any set of transfer functions (tl(O) •. ..• t,(O» such that E, ,[t,(O,. 0_,)) = 1,(0,) for all 0,. In general. there are many such functions t,(· •. ); one. for example. is simply t,(O,. = 1,(0,)." We now illustrate one implication of this characterization result for the auction setting introduced in Example 23.B.4. Some further implications of Proposition 23.0.2 are derived in Sections 23.E and 23.F.
°_,)
Auctions: tile revenue equivalence theorem Thus, - (0.)
v"
~ U,(O,) - U,(II,) ~ -(0)
b,- 0,
v"
•
(23.0.13)
Expression (23.0.13) immediately implies that v,(') must be nondecreasing (recall that we have taken b, > 0,). In addition, letting 8, -+ 0, in (23.D.13) implies that for all 0, we have
Ui(O,) = v,(O,) and so
U,(O,) =
U,{~,)
+
f"
v,(s) ds
for all 0,.
!,
(ii) Sufficiency. Consider any 0, and 0, and suppose without loss of generality that 0,. If (23.0.11) and (23.0.12) hold, then
0, >
u;(o,) - U,(O,) =
', f" f'·
_ v,(s) ds
;:: _ v,(8,) ds I.
= (0, - O,)v,(8,).
32. Observe Ihat the agent's preferences here over his expected benefit jj, and expected transfer I, satisfy the single-crossing property that played a prominent role in Sections I3.C and 14.C.
889
-----------------------------------------------------------------------
Let us consider again the auction setting introduced in Example 23.B.4: Agent 0 is the seller of an indivisible object from which he derives no value. and agents 1•... , 1 are potential buyers. 34 It will be convenient. however, to generalize the set of possible alternatives relative to those considered in Example 23.B.4 by allowing for a random assignment of the object. Thus. we now take y,(O) to be buyer i's probability of getting the object when the vector of announced types is = (0" ...• 0,). Buyer i's expected utility when the profile of types for the 1 buyers is 0= (0" ... ,0,) is then 0, y,(O) + tote). Note that buyer i is risk neutral with respect to lotteries both over transfers and over the allocation of the good. This setting corresponds in the framework studied in Proposition 23.0.2 to the case where we take k=(yl .... ,y,). K={(yl .... ,y,):y,E[O.I] for all i = 1, ... ,1 and L, y, ~ I}, and v,{k) = Yo' Thus, to apply Proposition 23.0.2 we can write VitO,) = y,(O,), where y,(8,) = E._.[y,(O" 0_,)] is the probability that i gets the object conditional on announcing his type to be 8, when agents j .;. i announce their types truthfully, and U,(O,) = 0d,(Od + 1,(0,).
°
33. However. if we wish the social choice function f(') = (k('), ,,(.)•...• 1,('» to satisfy some further properties, such as budget balance. only a subset (possibly an empty one) of the transfer functions generating the expected transfer functions (t.(8.). ...• t,(8,» may have these properties. 34. We note that our assumption that the seller in an auction setting derives no value rrom the
object is not necessary for the revenue equivalence tbeorem. (As we shall see, tbe result characterizes the expected revenues generated for the seller in different auctions. and so is valid for any utility function that the seUer might bave.) In the absence of tbis assumption, however, tbe seller in an auction will generally care about more than just the expected revenue he receives.
890
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DEIIGN
We can now establish a remarkable result, known as the revenue equivalence theorem. 3 >
--- --
(.v,(O,lO, -
t
MO,) dO,) -
dO,J - U,(Q.).
I ->,(0.> ,(O'»)J( ,., .n >j(III») dO, ... dll,J , (23.0.16)
By inspection of (23.0.16). we see that any two Bayesian incentive compatible social choice functions that generate the same functions (y,(O) •...• y,(II» and the same values of (U,(Q,) •...• U,(Q,» generate the same expected revenue for the seller. _ As an example of the application of Proposition 23.0.3, consider the equilibria of the first-price and second-price sealed-bid auctions that we identified in Examples 23.8.5 and 23.B.6 (where the buyers' valuations were independently drawn from the uniform distribution on [0. I]). for these equilibria. the conditions of the revenue equivalence theorem are satisfied: in both auctions the buyer with the highest valuation always gets the good and a buyer with a zero valuation has an expected utility of zero. Thus, the revenue equivalence theorem tells us that the seller receives exactly the same level of expected revenue in these equilibria of the two auctions (you can confirm this fact in Exercise 23.0.3). More generally. it can be shown that in any symmetric auction setting (i.e .• one where the buyers' valuations are independently drawn from identical distributions). the conditions of the revenue equivalence theorem will be met for any Bayesian Nash equilibrium of the first-price sealed-bid auction and the (dominant strategy) equilibrium of the second-price sealed-bid auction (see Exercise 23.0.4 for a consideration of symmetric equilibria in these settings). We can conclude from Proposition 23.0.3. therefore. that in any such setting the first-price and second-price sealed-bid auctions generate exactly the same revenue for the seller.
U,(Q,).
Moreover. integration by parts implies that
.v,(s) dS)>,(O,) dO, =
>j(Oj»)dll, ...
I
,.L, U,(Q,).
.v,(s) dS)>,(O,) dO,
.v,(s) dS) >,(0,) dO,J -
n
[ y,(O, •...• 0,)(0, [ f.O,' ... fO' 0, ,=,
r
[r r r(r (r (r =
E[ -1.(0,)] =
(23.0.15)
E[ -1,(0)) = E•.[ - 1,(0,))
(.v,(0;) 11, - U,(Q,) -
P A A TIC I PAT ION
Thus. the seller's expected revenue is equal to
Proof: By the revelation principle. we know that the social choice function that is (indirectly) implemented by the equilibrium of any auction procedure must be Bayesian incentive compatible. Thus, we can establish the result by showing that if two Bayesian incentive compatible social choice functions in this auction setting have the same functions (y,(O) •...• y,(O)) and the same values of (U,(Q,), ... , U,(Q,» then they generate the same expected revenue for the seller. To show this. we derive an expression for the seller's expected revenue from an arbitrary Bayesian incentive compatible mechanism. Note. first. that the seller's expected revenue is equal to :Lf., E[ -1,(0». Now.
r
23. E:
or. equivalently.
6' ... f6' y,(O, •...• 0,)(0, - I - <1>,(0,»)( [ f,,~, >,(0,) ,.
Proposition 23.0.3: (The Revenue Equivalence Theorem) Consider an auction setting with I risk-neutral buyers. in which buyer i's valuation Is drawn from an interval [Q;.6;] with Q; ;I< 6; and a strictly positive density 11>;(·) > 0, and in which buyers' Iypes are statistically independent. Suppose that a given pair of Bayesian Nash equilibria of two different auction procedures are such that for every buyer i: (i) For each possible realization of (0 ...• 0,). buyer i has an identical probability " of getting the good in the two auctions; and (ii) Buyer i has the same expected utility level in the two auctions when his valuation for the object is at its lowest possible level. Then these equilibria of the two auctions generate the same expected revenue for the seller.
=
5 ECT, 0 N
23.E Participation Constraints
.v,(O,),(O,) dO,)
In Sections 23. B to 23.0. we have studied the constraints that the presence of private information puts on the set of implementable social choice functions. Our analysis up to this point. however, has assumed implicitly that each agent i has no choice but to participate in any mechanism chosen by the mechanism designer. That is. agent i's discretion was limited to choosing his optimal actions within those allowed by the mechanism. In many applications. however. agents' participation in the mechanism is voluntary. As a result. the social choice function that is to be implemented by a mechanism must not only be incentive compatible but must also satisfy certain participation (or individual rationality) constraints if it is to be successfully implemented. In this section. we provide a brief discussion of these additional
Substituting. we see that (23.0.14)
35. Versions of the revenue equivalence theorem have been derived by many authors; see McAfee and McMillan (1987) and Milgrom (1987) for references as well as for a further discussion of the result.
.l
CON 5 T R A I N T 5
891
892
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
--------------------------------------------------------------constraints on the set of implementable social choice functions. By way of motivating our study, Example 23.E.I provides a simple illustration of how the presence of participation constraints may limit the set of social choice functions that can be successfully implemented. Example 23.E.l: Participation Constraints in Public Project Choice. Consider the following simple example of public project choice (recall our initial discussion of public project choice in Example 23.S.3). A decision must be made whether to do a given project or not, so that K = {O, I}. There are two agents, I and 2. For each agent i, 0, = {~, 8}, so that each agent either has a valuation of~, or a valuation of iI. We shall assume that 8 > 2~ > O. The cost of the project is c E (2~, 8). Suppose that we want to implement a social choice function having an ex post efficient project choice; that is, one that has k·(III' (1 2) = 1 if either 0 1 or 112 is equal to 8, and P(OI' (1 2) = 0 if II, = 112 =~. In the absence of the need to insure voluntary participation, we know from Section 23.C that we can implement some such social choice function in dominant strategies using a Groves scheme. Suppose, however, that each agent has the option of withdrawing from the mechanism at any time (perhaps by withdrawing from the group), and that, if he does, he will not enjoy the benefits of the project if it is done, but will also avoid paying any monetary transfers. Can we implement a social choice function that achieves voluntary participation and that has an ex post efficient project choice?l6 The answer is "no." To see this, note that if agent I can withdraw at any time, then to insure his participation it must be that I,(~, 8) ~ -~. That is, it must be that whenever his valuation for the project is ~, he pays no more than ~ toward the cost of the project. Now consider what agent l's transfer must be when both agents announce that they have valuation 8: If truth telling is to be a dominant strategy, then t, (0, 0) must satisfy
Ok·(O, 0)
+ 1, (0, 0) ~
Ok·(~, 0)
+ t,(~, 0),
or, substituting for k·(O, 0) and k·(~, ii),
0+ t,(O, 8) ~ 8 + t,(~, 8). Since t,(~, 0) ~ -~, this implies that 1, (0, 8) ~ -~. Thus, we conclude that agent I must not make a contribution toward the cost of the project that exceeds ~ when (II" ( 2 ) = (0,8). Moreover, by symmetry, we have exactly the same constraint for agent 2's transfer when (II" (1 2 ) = (8, 8), namely, t , (8, 8) ~ -~. Hence, t,(O, 0) + t 2 (O, 0) ~ -2~. But if this is so, then because 2~ < c, the feasibility condition t,(O, 8) + t 2(8, 8) S -c cannot be satisfied. We conclude,therefore, that it is impossible to implement a social choice function with an ex post efficient project choice when the agents can withdraw from the mechanism at any time. Note also that the presence of an "outside agent" (say "agent 0") who does not care about the project decision does not help at all here when that agent can also withdraw from the mechanism at any time. This is because, to insure this agent's participation, his transfer to(lI .. (1 2 ) must be nonnegative for every realization of
36. Note that any social choice funclion thai fails to have both agents participate is necessarily e., post inefficient because one of the agents is excluded from the benefits of Ihe project.
--
IECTION
U.E:
PARTICIPATION
(0 (1 2 ), In particular, w.,e ..must h~v: t o(8, 82 ~ 0, and so we must fail to satisfy the " feasibility condition to(II, II) + t,(II, II) + t 2(O, II) s -c. • As a general matter, we can distinguish among three stages at which participation constraints may be relevant in any particular application. First, as in Example 23.E.I, an agent i may be able to withdraw from the mechanism at the ex post slage that arises after the agents have announced their types and an outcome in X has been chosen. Formally, suppose that agent i can receive a utility of ii,(II,) by withdrawing from the mechanism when his type is 11,.37 Then, to insure agent i's participation, we must satisfy the ex post participation (or individual rationality) constraints 3 ' (23.E.I) In other circumstances, agent i may only be able to withdraw from the mechanism at the interim stage that arises after the agents have each learned their type but before they have chosen their actions in the mechanism. Letting U,(O.lfl = E•. ,[u,(f(II" 11_,), II,) IlIa denote agent i's interim expected utility from social choice function f( . ) when his type is II" agent i will participate in a mechanism that implements social choice function f(') when he is of type II, if and only if U,(O,lfl is not less than ii,(O,). Thus, interim participation (or individual rationality) constraints for agent i require that for all 0,.
(23.E.2)
In still other cases, agent i might only be able to refuse to participate at the ex ante stage that arises before the agents learn their types. Letting U,(f) = E.,[U,(O;l fl] = E[u,(f(II.. II _,),0,)] denote agent i's ex ante expected utility from a mechanism that implements social choice function f('), the ex ante participation (or individual rationality) constraint for agent i is U,(f) ~ E.,[ii,(II,)].
(23.E.3)
Participation constraints are of the ex ante variety when the agents can agree to be bound by the mechanism prior to learning their types. When, instead, agents know their types prior to the time at which they can agree to be bound by the mechanism, we face interim participation constraints. 39 Finally, if there is no way to bind the
37. We assume that agent i's utility from withdrawal depends only on his own Iype. 38. We assume throughout that it is always optimal to insure that each agent is always willing to participate. In fact, however, there is no loss of generality from assuming this: When agents can "not participate: any outcome that can arise when some subset I' of the I agents does not participate, say x', should be included in the set X. Because we can always have the mechanism select x' in the circumstanoes when this subset of agents would have refused to participate, if the set X is defined appropriately we can always replicate the outcome of any mechanism that causes non participation with a mechanism in which all agents are always willing to participate. 39. Recall that the assumption in a Bayesian game that types are drawn from a commOn prior density is often merely a modeling device for how agents form beliefs about each others' types (see Section 8.E). That is, for analytical purposes we may be representing a setting in which agents' types are already determined but are only privately observed by assuming that there has been a prior random draw of types from a commonly known distribution; but there may not actually be any such prior stage at which the agents could possibly interact.
CONITIIAINTI
893
894
CHAPTER
23:
INCENTIVES
AND
MECHANISM
--
DEIIGN
---------------------------------------------------------------------~
agents to the assigned outcomes of the mechanism against their will, then we face ex post participation constraints.'· Note that if f(·) satisfies (23.E.I), then it satisfies (23.E.2); and, in turn, if it satisfies (23.E.2), then it satisfies (23.EJ). However, the reverse is not true. Thus, the constraints imposed by voluntary participation are most severe when agents can withdraw at the ex post stage, and least severe when they can withdraw only at the ex ante stage. In summary, when agents' types are privately observed, the set of social choice functions that can be successfully implemented are those that satisfy not only the conditions identified in Sections 23.C and 23.0 for incentive compatibility (in, respectively, either a dominant strategy or Bayesian sense, depending on the equilibrium concept we employ) but also any participation constraints that are relevant in the environment under study. In the remainder of this section, we illustrate further the limitations on the set of implementable social choice functions that may be caused by participation constraints by studying the important Myerson-Salterthwaite theorem [due to Myerson and Satterthwaite (1983)].
SECTION
23.£:
PARTlelPATION
news: Whenever gains from trade are possible, but are not certain,'1 there is no ex post efficient social choice function that is both Bayesian incentive compatible and satisfies these interim participation constraints. Thus, under the conditions of the theorem, the presence of both private information and voluntary participation implies that it is impossible to achieve ex post efficiency. (For an illustration of the result for specific functional forms, see Exercise 23.E.7.)
Proposition 23.E.1: (The Myerson-Satterthwaite Theorem) Consider a bilateral trade setting in which the buyer and seller are risk neutral, the valuations 0, and O2 are independently drawn from the intervals [~" 8,] c R and [~2' 62] c R with strictly positive densities, and (Q" 0,) ("\ (~2' 62) '" 0. Then there is no Bayesian incentive compatible social choice function that is ex post efficient and gives every buyer type and every seiler type nonnegative expected gains from participation.
Proof: The argument consists of two steps:
Slep I: In any Bayesian incentive compatible and interim individually rational social choice function f(·) = [y,(.), y,{- ),t,(' ),t,(-)) in which y,(O" 0,) + y,(O" 0,) = I and t,(O" 0,) + t,(O" 0,) = 0, we musl have
f fi, i,
The Myerson-Satterthwaite Theorem
Consider again the bilateral trade setting introduced in Example 23.B.4. Agent I is the seller of an indivisible object and has a valuation for the object that lies in the interval 0 1 = [~I' 61] c R; agent 2 is the buyer and has a valuation that lies in 0, = [~" 6,] c R. The two valuations are statistically independent, and 0, has distribution function <1Ik) with an associated density "'k) satisfying ",,(0,) > 0 for iii]. We let y,(O) denote the probability that agent i receives the good all 0i E given types 9 = (0 1,9,), and so agent i's expected utility given 0 is O,y,(O) + t,(O) (we normalize mi = 0 for all i). The expected externality mechanism studied in Section 23.0 shows that in this setting we can Bayesian implement an ex post efficient social choice function (or what, in this environment, we might call a "trading rule"). A problem arises with the expected externality mechanism, however, when trade is voluntary. In this case, every type of buyer and seller must have nonnegative expected gains from trade if he is to participate. In particular, if a seller of type 01 is to participate in a mechanism that implements social choice function f('), that is, if participation in the mechanism is to be individually rational for this type of seller, it must be that UI(O.!f> ~ 0" because this seller can achieve an expected utility of 0, by not participating in the mechanism and simply consuming the good. Likewise, a buyer of type 0, can always earn zero by refusing to participate, and so we must have U,(O,If> ~ O. Unfortunately, these interim participation constraints are not satisfied in the expected externality mechanism (you are asked to verify this in Exercise 23.E.I). The Myerson-Satterthwaite Theorem tells us the following disappointing piece of
"
"
.
.
y,(O"0,)
[(
I - <1>,(0,») (0, + <1>,(0,»)] - 4>,(0,)4>,(0,) dO, dO, ~ O.
0, - - - - - 4>,(0,)
4>,(0,)
~~
To see this, note first that the same argument that leads to (23.0.15) can be applied here to give [throughout the proof we suppress the argument f in U,(0.lf) and simply write U,(O,)]:
mi'
fi, fi, y,(O,. 0,)(0, -
£[ -i,(O,)) = [
"
I - <1>,(0'»)4>,(0,)4>,(0,) dO, dO,] -
"
.
4>,(0,)
.
Also, because (23.D.12) implies that
V,(Q,)
=
VItO,) -
i, ii, f ~I
U,(~,). (23.E.5)
y,(O" 0,)4>,(0,) dO, dO"
!l
condition (23.0.15) also implies that
-
[ fi, fi, y,(O"O,) (<1>,(0,») 0, + - - 4>,(0,)4>,(0,) dO, dO, ] - V,(O,).
£[ -1,(0,)] =
" .
"
4>,(0,)
.
(23.E.6)
Then, since J',(O,. 0,) = I - y,(O,. 0,) we have
-
£[-1,(0,))=
-
[ f" f" (0 , + <1>,(0.1) - - 4>,(O,)tJ>,(O,)dO,dO, ] 4>,(0,) ,,~,
[ f fi, i' ~,
"
»)
y,(O,.o,) ( II, + - '(0 - ' 4>,(0,)4>,(0,) dO, dll, 4>,(0,)
e,) '"
]-
U,(ii,).
41. That is, whenever (~" 8,)" (q" 0 (or equivalently, 8, > ~, and 8, > ~,), so that for some realizations of 0 = (0 1.8 2 ) there are gains from trade but for others there are not.
40. For example, if the mechanism can lead an agent into bankruptcy, the provisions of bankruptcy law provide an elfective lower bound on ex post utilities.
J
CONSTRAINT.
895
896
CH"PTER
23:
INCENTIVES
.. NO
"ECH .. NIS ..
--- --
DESIGN
But dO, dO,J [ f•i,, 'fi,, (0, + <1>,(0'»)<1>,(0,)<1>,(0,) <1>,(0,)
= [
fi, [0,<1>,(0,) + <1>,(0,)] dO,J !,
= [0,<1>,(0,)]::
=0,. Thus. £[ -1,(0,)]
0, - [
=
fi, fi, y,(O" 0,)(0, + <1>,(0'»)<1>,(0,)<1>,(0,) dO, dO,J <1>,(0,) "
.
U,(O,).
h
.
~~
Now. the fact that 1,(0" 0,) + 1,(0,,0,) = 0 implies that £[ -1,(0,,0,)] + £[ -1,(0,,0,)] = So. adding (23.E.5) and (23.E.7) we see that
o.
[U,(O,) - 0,] + U,(q,) =
f"i, fi," }"(O,, 0,)[(0, -
I - <1>,(0,») _ (0,
+ ,(O'»)J,(O,),(O,) dO, dO,.
<1>,(0,)
<1>,(0,)
But individual rationality implies that U,(O,) ~ 0, and U,(q,) ~ 0, which establishes (23.E.4). Seep 2: Cundilion (23.£.4) cannol be salisfied },,(O,. 0,) = 0 whenever 0, < 0,.
if
y,(O" 0,) = I whenever 0, > 0, and
Suppose it were. Then the left-hand side of (23.E.4) could be written as
f f";'I.' .• i
,
"
"
oI
[(0, - I - <1>,(0,) - 0,)<1>,(0,) - ,(O,)J,(O,) dO, dO, <1>,(0,) =
=
f"[( f"[(
J";,,.,.',I <1>,(0,) dO,
"
I - <1>,(0,) ) 0, - - - - - 0, <1>,(0,) <1>,(0,) "
!,
I - <1>,(0,) . - ) ,(Mm{O"O,}) . O,-----Mm{O"O,} ,(O,)dO, <1>,(0,)
= -
f
~ [I
J
- <1>,(0,)]<1>,(0,) dO,
+ f~ . [(0, - 0,)<1>,(0,) + (<1>,(0,) -
i,
= -
f"
I)] dO,
'1
~l
I
[I - <1>,(0,)]<1>,(0,) dO,
+ [(0, - 0,)(<1>,(0,) - 1)],:
23.F:
OPTI .... L
... VEII .. N
whether trade will occur and at what price.·' By the revelation principle, we know that the social choice function that is indirectly implemented in a Bayesian Nash equilibrium·) of such a mechanism must be Bayesian incentive compatible. Moreover, since participation is voluntary, this social choice function f(·) must satisfy the interim individual rationality constraints that UI (8 1 1f) ~ 8 1 for all 8 1 and U,(O,1 f) ;;>: 0 for all 0,. Thus, the Myerson-Satterthwaite theorem tells us that, under its assumptions, no voluntary trading institution can have a Bayesian Nash equilibrium that leads to an ex post efficient outcome for all realizations of the buyer's and seller's valuations.
:
1
23,F Optimal Bayesian Mechanisms In Sections 23.B to 23.E we have been concerned with the identification of implementable social choice functions in environments characterized by incomplete information about agents' preferences. In this section, we shift our focus to the welfare evaluaton of implementable social choice functions. We begin by developing several welfare criteria that extend the notion of Pareto efficiency that we have used throughout the book in the context of economies with complete information to these incomplete information settings. With these welfare notions in hand, we then discuss several examples that illustrate the characterization of optimal social choice functions (and, by implication, the optimal direct revelation mechanisms that implement them). We restrict our focus throughout this section to implementation in Bayesian Nash equilibria, discussed in detail in Section 23.0. Unless otherwise noted, we also adopt the assumptions and notation of Section 23.0. Good sources for further reading on the subject of this section are Holmstrom and Myerson (1983), Myerson (1991), and Fudenberg and Tirole (1991). For economies in which agents' preferences are known with certainty, the concept of Pareto efficiency (or Pareto optimality) provides a minimal test that any welfare optimal outcome x E X should pass: There should be no other feasible outcome:i E X with the property that some agents are strictly better off with outcome :i than with outcome x, and no agent is worse off. The extension of this welfare test to social choice functions in settings of incomplete information should read something like the following: The social choice function 1(') is efficient if it is feasible and if there is no other feasible social choice function that makes some agents strictly better off, and no agents worse off.
<0,
where the inequality follows because 0, > completes the argument. •
SECTION
q,
and
q,
< 0,. This contradicts (23.E.4) and
Recalling the revelation principle for Bayesian Nash equilibrium (Proposition 23.0.1), the implication of the Myerson-Satterthwaite theorem can be put as follows: Consider any voluntary trading institution that regulates trade between the buyer and the seller. This includes, for example, any bargaining process in which the parties can make offers and counteroffers to each other, as well as any arbitration mechanism in which the parties tell a third party their types and this third party then decides
To operationalize this idea, however, we need to be more specific about two things: First, what exactly do we mean by a social choice function being ~feasible',? Second,
42. Strictly speaking, for a direct application of Proposition 23.E.I, the date of delivery and consumption of the good must be fixed (so the bargaining processes studied in Appendix A of Chapter 9 would not count). But through a suitable reinterpretation Proposition 23.E.l can be applied to settings in which trade may take place over real time, where not only delivery of the good matters but also the rime of delivery (sec Exercise 23.E.4 for details). 43. And, hence, in any perfect Bayesian or sequential equilibrium (sec Section 9.C).
.. ECH .. NIIN.
897
898
CHAPTER
23:
INCENTIVES
AND
MECHANI8M
DEIION
precisely what do we mean when we say that no other feasible social choice function .. makes some agents strictly better off, and no agent worse off"? Let us consider the first of these issues. The identification of the set of feasible social choice functions when agents' preferences are private information has been discussed extensively in Sections 23.0 and 23.E. Suppose that we define the set FBtc = {f: El .... X: f(·) is Bayesian incentive compatible}.
(23.F.I)
The elements of set Fatc in any particular application are the social choice functions that satisfy condition (23.0.1), the condition that assures that there is a Bayesian Nash equilibrium of the direct revelation mechanism r = (0 , .. , Elto f(·» in which " truth telling is each agent's equilibrium strategy. Likewise, following the discussion in Section 23.E, we can also define the set
FI /( = {f: El .... X: f(·) is individually rational}.
(23.F.2)
The set FIR contains those social choice functions that satisfy whichever of the three types of individual rationality (or participation) constraints (23.E.I)-(23.E.3) are relevant in the application being studied. If no individual rationality constraints are relevant (i.e., if agents' participation is not voluntary), then we simply have FtR = {f: El .... X}, the set of all possible social choice functions. The content of our discussion in Sections 23.0 and 23.E is therefore that the set of feasible social choice functions in environments in which agents' types are private information is precisely F* = FBIC n F,/(. Following Myerson (1991), we call this the incentive feasible set to emphasize that it is the set of feasible social choice functions when, because of incomplete information, incentive compatibility conditions must be satisfied. Now consider the second issue: What do we mean when we say that no other feasible social choice function would "make some agents strictly better off, and no agents worse off"? The critical issue here has to do with the liming of our welfare analysis. In particular, is the welfare analysis occurring before the agents (privately) learn their types, or after? The former amounts to a welfare analysis conducted at what we called in Section 23.E the ex ante stage (the point in time at which agents have not yet learned their types); the latter corresponds to what we called in Section 23.E the interim stage (the point in time after each agent has learned his type, but before the agents' types are publicly revealed). To formally define the different welfare criteria that arise in these two cases, let us once again denote by U.(0.lf) agent i's expected utility from social choice function f(·) conditional on being of type 0 Also let U,(f) = E.,[ U,(O;lf)] denote agent i's ex ante expected utility from social "choice function f(·). We can now state Definitions 23.F.1 and 23.F.2. Definition 23.F.1: Given any set of feasible social choice functions F, the social choice function f(') E F is ex ante efficient in F if there is no I (.) E F having the property that Vi(/) ~ V;(f) for all i = 1, ... , I, and Vi(/) > Vi(f) for some i. Definition 23.F.2: Given any set of feasible social choice functions F, the social choice function f(·) E F is interim efficient in F if there is no 1(·) E F having the property that Vi(Oil/) ~ Vi(Oilf) for all OiE Eli and all i= 1, . .. ,1, and Vi(Oill) > Vi(Oilf) for some i and 0iE Eli' The motivation for the ex ante efficiency test is straightforward: If agents have not yet learned their types, then when comparing two feasible social choice functions
---
IECTION
U.F:
O,TIMAL
.AYEIIAN
MECHANIIMI
899
---------------------------------------------------------------------we should evaluate each agent's well-being using his expected utility over all of his possible types. However, when our welfare analysis occurs after agents have (privately) learned their types, things are a bit trickier. Although the agents each know their types, we-as outsiders-do not know them. Thus, the appropriate notion for us to adopt in saying that one social choice function !(.) welfare dominates another social choice function f(·) is that !(.) makes every possible type of every agent at least as well off as does f('), and makes some type of some agent strictly better off. This leads to the concept of interim efficiency given in Definition 23.F.2. Proposition 23.F.1 compares these two notions of efficiency. Proposition 23.F.1: Given any set of feasible social choice functions F, if the social choice function f(') E F is ex ante efficient in F. then It is also Interim efficient in F. Proof: Suppose that f(·) is ex ante efficient in F but is not interim efficient in F. Then there exists an !(')EF such that U,(O.!!)~ U,(O.!f) for all O,EEl, and all i = I, ... , I, and U,(O.!!) > V,(O.! f) for some i and 0, EEl,. But since, for all i, U.(f) = E.,[U,(O.!f)] and V,(!) = E.,[U,(O,I!)], it follows that V/(!) ~ V/(f) for all j = I, ... , I, and V,(!) > V,(f) for some i, contradicting the hypothesis that f(·) is ex ante efficient in F . • The ex ante efficiency concept is more demanding than is interim efficiency (and so fewer social choice functions f(·) pass the ex ante efficiency test) because a social choice function !(.) can raise every agent's ex ante expected utility relative to the social choice function f(') even though !(.) may lead some type of some agent i to have a lower expected utility than he does with f(·). Putting together the elements developed above, we conclude that when agents' types are already determined at the time we are conducting our welfare analysis, the proper notion of efficiency of a social choice function in an environment with incomplete information is interim efficiency in F·, the set of Bayesian incentive. compatible and individually rational social choice functions.·· On the other hand, if our analysis is conducted prior to agents learning their types, then the proper notion of efficiency is ex ante efficiency in F*.·5 These two notions are often called simply ex ante incentive efficiency and interim incentive efficiency [the terminology is due to Holmstrom and Myerson (1983)], where the modifier "incentive" is meant to convey the point that the set F* is being used.·' These two welfare notions differ from the ex post efficiency criterion introduced in Definition 23.B.2. To see their relationship to it more clearly, Definition 23.F.3 44. These cases often correspond to situations in which our assumption that the agents' types are drawn from a known prior distribution is being used merely as a device to model agents' beliefs about each others' types, as described in Section 8.E, rather than as a description of any actual prior lime at which the agents could interact or our welfare analysis might have been done. 45. This case often arises in contracting problems when, It the time of contracting, the agents anlicipate thatlhey will later come to acquire privale information aboutthcir types. Then the natural welfare standard to use in comparing different contracts (i.e, different mechanisms) is the ex ante criterion. The principal-agent model studied in Section 14.C and Example 23.F.1 below is an example along Ihese lines. 46. However, since the relevant individual rationality constraints Vlry from one application to another, it is usually clearer to describe precisely the sct F within which efficiency is being evaluated.
900
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
develops the ex post efficiency notion in a manner that parallels Definitions 23.F.I and 23.F.2. Dellnltlon 23.F.3: Given any set of feasible social choice functions F. the social choice function f (.) e F is ex post efficient in F if there is no 1(·) e F having the property that Uj(/(O).Oj)~Uj(f(O).Oj) for all i=I ..... 1 and all Oee. and u,(i(O), OJ) > uj(f(O), OJ) for some i and 0 e e. The ex post efficiency test in Definition 23.F.3 conducts its welfare evaluation at the ex post stage at which all agents' information has been publicly revealed. Using this definition. we see that a social choice function f(·) is ex post efficient in the sense of Definition 23.B.2 if and only if it is ex post efficient in the sense of Definition 23.F.3 when we take F = {f: e ... X}. Note that the criterion of ex post efficiency in (f: e -+ X }, or more generally. of ex post efficiency in when individual rationality constraints are present. ignores issues of incentive compatibility. As a result, it is appropriate as a welfare criterion only if agents' types are in fact publicly observable. Because F* c Fill' allocations that are ex ante or interim incentive efficient need not be ex post efficient in this sense. Indeed. the Myerson-Satterthwaite theorem (Proposition 23.E.1) provides an illustration of this phenomenon for the bilateral trade setting: under its assumptions. no element of F * is ex post efficient. Examples 23.F.1 to 23.F.3 provide further illustrations. (For one way in which the notion of ex post efficiency is nevertheless still of interest in settings with privately observed types. see Exercise 23.F.1.) Note also that even in settings in which agents' types are public information, the use of ex post efficiency in F,. as our welfare criterion is appropriate only when agents' types are already determined. When our welfare analysis instead occurs prior to agents learning their types, the appropriate notion is instead the stronger criterion that f(·) be ex ante efficient in These two notions are sometimes called ex post classical efficiency and ex ante classical efficiency [again, the terminology is due to Holmstrom and Myerson (1983)] to indicate that no incentive constraints are involved in defining the feasible set of social choice functions.
F,.
F,.,
In the remainder of this section we study three examples in which we characterize welfare optimal social choice functions. In Examples 23.F.I and 23.F.2. it is supposed that one agent who receives no private information chooses a mechanism to maximize his expected utility subject to both incentive compatibility constraints and interim individual rationality constraints for the other agents. These two examples therefore amount to a characterization of one particular interim incentive efficient mechanism. In Example 23.F.3, we provide a full characterization of the sets of interim and ex ante incentive efficient social choice functions for a simple setting of bilateral trade with adverse selection. Example 23.F.l: A Principal-A gem Problem with Hidden Information. In Section 14.C we studied principal-agent problems with hidden information for the case in which the agent has two possible types. Here we consider the case where the agent may have a continuum of types. Recall from Section 14.C that in the principal-agent problem with hidden information. the principal faces the problem of designing an optimal (i.e .• payoff maximizing) contract for an agent who will come to possess private information. In doing so, the principal faces both incentive constraints and
--- ---
SEC T ION
23. F:
0 P TIM ALB AYE. I A N i l E C HAN I • II •
a reservation utility constraint for the agent. Recall also from Section 14.C that, in the limiting case in which the agent is infinitely risk averse, the agent must be guaranteed his reservation utility for each possible type he may come to have, and so this contracting problem is identical to the contracting problem that would arise if the agent already knew his type at the time of contracting. Here we shall set things up directly in these terms, assuming that the agent already possesses this information when contracting occurs. With this formulation. the principal's optimal contract can be viewed as implementing one particular interim incentive efficient social choice function. (When the agent actually does not know his type at the time of contracting and is infinitely risk averse. then this social choice function is also ex ante incentive efficient.) To introduce our notation. we suppose that the agent (individual I) may take some observable action e e R. (his "effort" or "task "level) and receives a monetary payment from the principal of t,. The agent's type is drawn from the interval [~, 6]. where ~ < 6 < O. according to the distribution function <1>(') which has an associated density function >(.) that is strictly positive on [Q,6]. We assume that this distribution satisfies the property that [0 - «I - <1>(0»/>(0))] is nondecreasing in IJ.47 The agent's Bernoulli utility function when his type is IJ is u,(e. 11.0) = I, + IJg(e). where g(.) is a differentiable function with g(O) = O. g(e) > 0 for e > O. g'(O) = O. g'(e) > 0 for e > O. and g"(') > 0; that is. Og(e) represents the agent's disutility of effort (recall that 0 < 0). with higher effort levels leading to an increasing level of disutility to the agent. Note that a larger (i.e., less negative) level of 0 lowers. at any level of e. both the agent's total level of dis utility and his marginal dis utility from any increase in e. As noted above, we suppose that the agent must be guaranteed an expected utility level of at least ii for each possible type he may have. The principal (individual 0) has no private information. His Bernoulli utility function is uo(e. (0 ) = v(e) + 10, where 10 is his net transfer and v(.) is a differentiable function satisfying v'(·) > 0 and v"(·) < O. A contract between the principal and the agent can be viewed as specifying a mechanism in the sense we have used throughout this chapter. By the revelation principle for Bayesian Nash equilibrium (Proposition 23.D.I). the equilibrium outcome induced by such a contract, formally a social choice function that maps each possible agent type into effort and transfer levels, can always be duplicated using a direct revelation mechanism that induces truth telling. Thus, the principal can confine his search for an optimal contract to the set of Bayesian incentive compatible social choice functions f(·) = (e('), 10 (' ),1 , ('» that give the agent an expected utility of at least ii for every possible value of O. In what follows. we shall (without loss of generality) restrict attention in our search for the principal's optimal contract to contracts that have to(lJ) = - t,(IJ) for all IJ (i.e., that involve no waste ofnumeraire). The principal's problem can therefore be stated as Max
E[v(e(O» - t,(IJ)]
f(·)"'(~{·I.II(·»
S.t. f(·) is Bayesian incentive compatible and individually rational. 47. For a discussion of how the analysis changes when this assumption is not satisfied, see Fudenberg and Tirol. (1991).
901
902
CH~PTEA
23:
INCENTIVES
"NO
MECH .. NISM
DEIIGN
SECTION
The present model falls into the class of models with linear utility studied in Section 23.0 [specifically, in the notation of Proposition 23.0.2, k = e, vl{k) = g{e), and VI{O) = g{e{O» here]. Letting VI{O) = II{O) + Og{e{O» denote the agent's utility if his type is 0 and he tells the truth, Proposition 23.0.2 can be used to restate the principal's problem in terms of choosing the functions e{') and VI (.) to solve E[v{e{O»
Max
+ Og{e{O» -
UI{O)]
(23.FJ)
r
s.t. (i) e{') is nondecreasing (ii) UI{O) = UI @ +
g(e(s» ds for all 0
u for all O~
(iii) UI(O);::
Constraints (i) and (ii) are the necessary and sufficient conditions for the principal's contract to be Bayesian incentive compatible, adapted from Proposition 23.0.2 [constraint (i) follows because g(.) is increasing in e], while constraint (iii) is the agent's individual rationality constraint. Note first that if constraint (ii) is satisfied, then constraint (iii) will be satisfied if and only if UI{Q) ~ u. As a result, we can replace constraint (iii) with (iii')
UI{Q)
~
U.
Next, substituting for UI(O) in the objective function from constraint (ii), and then integrating by parts in a fashion similar to the steps leading to (23.0.14), problem (23.FJ) can be restated as
[f
Max
S.t.
{v{e(o))
+ g{e{O»(0 - I ~(:O»)}4>{O)dOJ - VI @
(23.Fo4)
(i) e{') is nondecreasing (iii') UI{Q) ~ u.
It is now immediate from (23.Fo4) that in any solution we must in fact have = U. Thus, we can write the principal's problem as one of choosing e(') to solve
UI (Q)
Max n')
[f
{v{e(o»
+ g(e(O»
(0 - I
~(~~0»)}4>(0) dO] -
u
(23.F.5)
S.t. (i) e{') is nondecreasing. Suppose for the moment that we can ignore constraint (i). Then the optimal function e(') must satisfy the first·order condition·· v'(e{O»
+ g'(e(l!» ( 0 -
1- cI>{O») -- = 0 4>{I!)
for alll!.
U.F:
OPTI .... L
."'EII .. N MECHANISMS
903
----------------------------------------------------------------------
(23.F.6)
But note that, under our assumption that [I! - {(I - cl>{1!))/4>(0))] is nondecreasing in 0, the implicit function theorem applied to (23.F.6) tells us that any solution e{') to this relaxed problem must in fact be nondecreasing. Thus, (23.F.6) characterizes the solution to the principal's actual problem (see Section M.K of the Mathematical Appendix). The optimal VI {') [and, hence, ll{')] is then calculated from constraint (ii) of (23.FJ) using this optimal e{') and the fact that UI{Q) = ii.
It is interesting to compare this solution with the optimal contract for the case in which the agent's type is observable. This contract solves
Max
E[v{e{I!» - tl{I!)]
s.t. tl{l!)
+ Og(e(O)) ~ u for
all O.
Hence, the optimal task level in this complete information contract is the level e*(O) that satisfies, for all 0, v'(e*(O» + g'(e*{O))1! = O. Note that e*(O) is the level that arises in any ex post (classically) efficient social choice function. In contrast, the principal's optimal e(') when I! is private information is such that v'{e(l!» + g'(e(I!»I!{> 0 at alll!.< 0, = 0 at 0 = O. We see then that e(l!) < eO(O) for all I! < 0, and e(0) = eO{O). This is a version of the same result that we saw for the two· type case in Section 14.C. In the optimal contract, the type of agent with the lowest disutility from effort (here type 0; in Section 14.C, type 0,,) takes an ex post efficient action, while all other types have their effort levels distorted downward. The reason is also the same: doing so helps reduce the amount the agent's utility exceeds his reservation utility for types 0 > Q (his so·called information rents). To see this point heuristically, suppose that starting with some function e(') we lower e(iI) by an amount de < 0 for some type ~ E (Q, 0) and lower this type's transfer to keep his utility unchanged!' The decrease in the transfer paid to type a is ag'{e(O» de, while the direct effect on the principal is v'{e{~» de. At the same time, according to constraint (ii), this change in e{~) lowers the utility level, and hence the transfer, that must be given to all types 0 > ~ by exactly g'{e{~» de. The expected value of this reduction in the transfers paid to these types is -(I - cI>{O»g'(e{O»de. If the original e{') is an optimum, the sum of the first two changes in the principal's profits (those for type ~) weighted by the density of type ii, [v'(e(a» + ag'(e(a))] 4>(0) de, plus the reduction in payments to types 0 > a, (I - cI>(a»g'(e(a» de, must equal zero. This gives exactly (23.F.6). • Example 23.F.2: Optimal Auctions. We consider again the auction setting introduced in Example 23.Bo4. Here we determine the optimal auction for the seller of an indivisible object (agent 0) when there are I buyers, indexed by i = I, ... , I. Each buyer has a Bernoulli utility function 0, y,(I!) + t,{I!), where y,{O) is the probability that agent i gets the good when the agents' types are I! = (Ol' ... ' 0,). In addition, each buyer i's type is independently drawn according to the distribution function <1>,(') on [Q" 0;] c: Ii! with Q, -F 0, and associated density 4>,(') that is strictly positive on [Q" 0;]. We assume also that, for i = I •...• 1, 1- cI>,(O,)
0,----4>,{O,)
is nondecreasing in 0,. ~o 49. We say "heuristically" because to do this rigorously we need to perform this reduction in
48. It can be shown that under our assumptions, the optimal contract is interior, that is, has e(0) > 0 for (almost) all O.
e over an interval of types and then take limits.
50. For a discussion of the case in which this assumption is not met, see Myerson (1981).
904
CHAPTER
23:
'NCENT'VES
AND
MECHAN'SM
--
DES'GN
A social choice function in this environment is a function f(·) = (YO('), •.. ,y,(.), 10 (') ••••• 1,(')) having the properties that. for all fi e e. Yi(fi) e [0, I] for all i, :[,,,0 y,(fi) = I - yo(fi), and lo(fi) = - L,o'o 1.(fi).51 The seller wishes to choose the Bayesian incentive compatible social choice function that maximizes his expected revenue E,[lo(fi)) = - E.[L.o'o I,(fi)] but faces the interim individual rationality constraints that V,(fi,) = fi,Y.(fi,) + 1,(0,)
I ifl'O
ij·
f
[y,(fi,)O, - V,(fi;)]
(iii) V,(O,) == V,(Q;) (iv) U,{fii )
~
~ I.
0,.
0 for all j 'f- 0 and 0,.
We note first that if constraint (iii) is satisfied then constraint (iv) will be satisfied if and only if V,(Q')
Vi(Q,) 2: 0 for all i 'f- 0 and 0i'
Next. substituting into the objective function for V,(O;) using constraint (iii), and following the same steps that led to (23.0.16), the seller's problem can be written as one of choosing the y,(.) functions and the values V,(Q,), ... , V,(Q,) to maximize B' fij, [ I, y,(fi" fs, ... !, i-'
»)][ n' ]
cIl,{fi ... , 0,) ( 0i -I--
i-'
,
I
i-'
U,(Qi)
subject to constraints (i). (ii), and (iv'). It is evident that the solution must have V,W,) = 0 for all i == I •... , I. Hence. the seller's problem reduces to choosing functions y, (- )•...• y,(-) to maximize
[.± y,(fi, ..... fi,)(Oi f"li, ... fij,~".,
n
I - cIl,(fi,»)][
(23.F.8)
subject to constraints (i) and (ii). Let us ignore constraint (i) for the moment. Define •
(23.F.9)
Throughout the chapter. we have restricted attention to "private values" settings in which agents' utilities depend only on their own types. In a number of settings of economic interest, however, an agent i's utility depends not only his own type, fi" but also on the types of other agents, O_i' Thatis, agent i's Bernoulli utility function may take the form u,(x, fi) rather than u,(x, fi,), where 11 = (11,,11_,). Fortunately, all of the concepts of implementation for Bayesian Nash equilibrium that we have studied in Sections 23.0 to 23.F extend readily to this case. For example, we can say that the social choice function f e ... X is Bayesian incentive compatible if for all i
J.(fi.)==fi._I-cIl,(fi,). ••
IAYEIIAN
[Note that J,(fi,) = Max {O, Max ... , J.(O.)} is a zero probability event.] But, given our assumption that J,(') is nondecreasing in fi/o (23.F.9) implies that Yi(') is nondecreasing in 0" which in turn implies that y,(.) is nondecreasing. Thus the solution to this relaxed problem actually satisfies constraint (i), and so is a solution to the seller's overall problem (see Section M.K of the Mathematical Appendix). The optimal transfer functions can then be set as I,(fi) == V,(fi,) - fi,y,(fi i), where V,(Oi) is calculated from constraint (iii). A few things should be noted about (23.F.9). First, observe that when the various agents have differing distribution functions eII,('), the agent i who has the largest value of J,( 0,) is nOI necessarily the same as the agent who has the highest valuation for the object. Thus, the seller's optimal auction need not be ex post (classically) efficient. Second. in the case of symmetric bidders in which Q, = Q and J,(') == J(.) for i = I •...• 1. when Q> 0 is large enough so that J@ > O. the optimal auction always gives the object to the bidder with the highest valuation and also leaves each bidder with an expected utility of zero when his valuation attains its lowest possible value. We can therefore conclude, using the revenue equivalence theorem (Proposition 23.0.3). that the first-price and second-price sealed-bid auctions are both optimal in this case. Third, the optimal auction has a nice interpretation in terms of monopoly pricing. Consider, for example, the case in which I == I and Q, = O. Then conditions (23.F.9) tell us that the optimal auction gives the object to the buyer (agent I) if and only if J,(O,) == [fi, - «(1 - eII,(II,))/O. Suppose we think instead of the seller in this circumstance simply naming a price p and letting the buyer then decide whether to buy at this price. The seller's expected revenue from this scheme is p(1 - cIl.(p)), and so the first-order condition for his optimal posted price, say p*, is (I - cIl,(p*)) p* 0, and with probability 0 if J,(O,) < O. exactly as in the optimal auction derived above. Indeed, given the revenue equivalence theorem we can conclude that in this case this simple posted price scheme is an optimal mechanism for the seller. [For more on the monopoly interpretation of optimal auctions, see Exercise 23.F.5 and Bulow and Roberts (1989).] •
o.
+ r;: y,(s) ds for all i 'f- 0 and
OPT'MAL
y,(fi) = 0
(23.F.7)
(ii) For all fi: Yi(O) e [0, I] for all i # 0, L, .. o y,(fi)
23.f:
problem if and only if for all i = I, ... ,I we have and
~,
s.t. (i) Yo{.) is nondecreasing for all i #
SECT'ON
Then inspection of (23.F.8) indicates that y,(.) • ... , y,(.) is a solution to this relaxed 51. Once again we restrict attention. without loss of generality. to social choice functions involving no waste of either the numeraire or the good (there is always an optimal social choice func,ion for 'he seller with this form).
".j
MECHANII . . .
905
906
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DeSIGN
SECTION
23.F:
OPTIMAL
BATEIIAN
MECHANIIMI
907
------------------------------------------------------------------- ,------------------------------------------------------------------and all Ii, EO"
function (h,I L , YH,I H ) is Bayesian incentive compatible if (23.F.l0)
for all b, EO,. Our third and last example, which studies a simple bilateral trade setting with adverse selection (see Section I3.B for more on adverse selection), falls within this class of models. Another difference from the analysis of Examples 23.F.I and 23.F.2 is that here we shall characterize the enlire selS of both ex ante and interim incentive efficient social choice functions. Example 23.F.3: Bilaleral Trade wilh Adverse Selection [from Myerson (1991)]. Consider a bilateral trade setting in which there is a seller (agent I) and a potential buyer (agent 2) of one unit of an indivisible private good. The good may be of high quality or low quality, but only the seller observes which is the case. To model this, we let the seller have two possible types, so that 0, = {IiL,Ii H }, and we assume that Prob (Ii H ) = .2. Both the buyer's and the seller's utilities from consumption of the good depend on the seller's type. In particular, letting Y denote the probability that the buyer receives the good, and letting I denote the amount of any monetary transfer from the buyer to the seller, we suppose that'2 u,(y,IIIi L ) = I
+ 20(1
U2(y,lllid = 30y - I ,
- Y),
u,(y,IIIiH ) = I
+ 40(1
- y), (23.F.II)
u 2 (y,IIIi H ) = SOy -I.
A social choice function in this setting assigns a probability of trade and a transfer for each possible value of Ii" and so can be represented by a vector (h, IL' YH, I H). We suppose that trade is voluntary, and that as a result any feasible social choice function must satisfy interim individual rationality constraints for both the buyer and the seller. For the seller this means that, for each type he may have, his expected utility must be no less than his utility from refusing to participate and simply consuming the good. Hence, we must have
+ 20(1- yd ~ 20 IH + 40(1 - YH) ~ 40.
(23.F.IS) and (23.F.l6) Condition (23.F.15) requires that truth telling be an optimal strategy for the seller when his type is 0H; (23.F.16) is the condition for truth telling to be optimal when his type is 0L' Thus, (h, IL' YH' I H) is a feasible social choice function if and only if it satisfies the incentive compatibility constraints (23.F.15)-(23.F.l6) and the interim individual rationality constraints (23.F.I2)-(23.F.14). These constraints imply that any feasible social choice function possesses the following three properties (Exercise 23.F.9 asks you to establish these points): (i) No feasible social choice function is ex post (classically) efficient. (ii) In any feasible social choice function, YH :S: hand IH :S: I L. (iii) In any feasible social choice function, the expected gains from trade for a low-quality seller are at least as large as the expected gains from trade for a high-quality seller, that is, IL - 20YL ~ IH - 4OYH' We now proceed to characterize the interim and ex ante incentive efficient social choice functions for this bilateral trade problem. To determine the interim incentive efficient social choice functions, we need to determine the (h, IL' YH,I H) that solve, for each possible choice of ii'H 2: 0 and ii2 2: 0, the following problem (we have simplified the incentive compatibility and individual rationality constraints by eliminating constants on both sides of the inequalities, and have removed a constant from the objective function as well'): Max (YI.IE(O.l).f,.• YHE(O.l)"n)
S.t.
(23.F.I2)
IL
(23.F.I3)
For the buyer, on the other hand, interim individual rationality simply requires that he receive nonnegative expected utility from participation (recall that he does not observe Ii,). Hence, we must have
(IR 2 )
(23.F.14)
Note from (23.F.II) that if Ii, were publicly observable then, for each value of Ii" there would be gains from trade between the buyer and the seller. Because of this I (see fact, any ex post (classically) efficient social choice function has YH h
=
=
Exercise 23.F.8). Because Ii, is only privately observed, the set of feasible social choice functions is the incentive feasible set F*, the set of Bayesian incentive compatible and (interim) individually rational social choice functions. In the present context, the social choice 52. We assume that there is no waste ofeitber the good or the numeraire, so that the probability that either the buyer or the seller consumes the good is I, and any transfer from the buyer goes to the seller.
IL - 20h
(i) (ii) (iii) (iv)
(23.F.17)
IH - 40YH ~ IL - 40YL IL - 20YL 2: IH - 20YH IH - 40YH 2: ii'H .2(50YH - I H) + .8(30h - I L ) ~ ii 2 •
Problem (23.F.17) characterizes interim incentive efficient social choice functions by maximizing the interim expected utility of the type 0L seller subject to giving the type IiH seller an interim expected utility of at least iitH ~ 0, giving the buyer an interim expected utility of ii2 ~ 0 (since the buyer acquires no private information, this is equivalent to giving him an ex ante expected utility of ii, ), and satisfying the seller's incentive compatibility constraints. We now proceed to characterize the solution to problem (23.F.l7) through a series of steps. Slep I:
Any solulion 10 problem (23.F.I7) has h = I: Ihal is, in any inlerim
incenlive efficienl social choice funclion, Irade is cerlain 10 occur when Ihe good is of low qualilY.
To see this, suppose that (yt,lt, y~,I~) solves (23.F.17) but that yt < I. Consider a change to social choice function (PL, i L , PH' iH )=(yt +£, It + 30£, y~, I~) 53. In essence, we have expressed all of these in terms of the agents' gains from trade.
908
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
where £ > O. For a sufficiently small £ > 0, this new social choice function satisfies all of the constraints of problem (23.F.l7) (check this), and raises the value of the objective function-but this contradicts the optimality of (y!,/!, y~,/n
--- 1-I
SECTION
I_
I_ =
Given step I, if a solution to (23.F.17), say
(y!,/Ly~,/~)
has
y~ =
= 40YH I_ =
26y_ + 8 - '"
26y_
+8 The optimal level of (YII' III) in problem (23.F.l9) for a given pair ("II' U,) ~ O.
} Level SeIS of I_ - 20y_
Figure 23.F.2 (rlghl)
The shaded set contains those pairs (YH,I H) arising in
"IH
interim incentive
In any solution 10 (23.F.17), constraint (ii) is binding (i.e., holds with O
y_
If conslrainl (ii) binds and YL;Z: YII' Ihen conslraint (i) is necessarily
If constraint (ii) binds then III - IL = 20(YII - yd. If YL ;z: YII then this implies that III - IL ;z: 4O(YII - yd, or, equivalently, III - 40YII ;z: IL - 40Yv Hence, constraint (i) is satisfied.
Given steps I to 4, we can simplify problem (23.F.17). In particular, we see that (h, IL' YII,III) is interim incentive efficient if and only if YL = I and (I L, YII,III) solve
Max (ILE(O,
IL - 20
(23.F.IS)
I],Y/I.,,,E(O, lU
s.t. (ii') IL - 20 = III - 20YII' (iii) III - 40YII ;z: (iv) .2(50YII - III) + .8(30 - I L) ;z: ii 2 • Substituting from constraint (ii') of problem (23.F.l8) for IL in its objective function and in constraint (iv) we see that we can determine the optimal values of (YII,III) by solving III - 20YII
The shaded set in Figure 23.F.2, {(YII' III): YII E [0, I], III;Z: 40YII' and 'H ~ + S}, depicts al/ of the pairs (YII,I H) that arise in interim incentive efficient social choice functions. These are determined by performing the analysis in Figure 23.F.I for each possible pair (ii lll , ii 2 ) ~ 0 [a sample pair (iH, iH) is also depicted in Figure 23.F.2]. Note that in any interim incentive efficient social choice function we have YII ~ 4/7. As above, for each pair UH, ill) in this set, we can determine an interim incentive efficient social choice function UL' iL , YH' iH) by setting YL = I and iL = 20 + iH - 20YH' Now, out of this set of interim incentive efficient social choice functions, which are ex ante incentive efficient? (Recall that the set of ex ante incentive efficient social choice functions is a subset of the interim incentive efficient set. Note also that although we now employ an ex ante welfare criterion, the set of participation constraints defining F· continue to be the same interim participation constraints used above.) The buyer's and seller's ex ante expected utilities in an interim incentive efficient social choice function (YL,I L , YH,I H) are
S.t.
U, = .S(I L
+ 20(1 -
YL»
+ .2(IH + 40(1 -
;z: ii,", (iv') 26YII - III + S ;z: ii 2 • (ii) III - 40YII
The solution for a given pair of values of ii," ;z: 0 and ii2 ;z: 0 is depicted in Figure 23.F.1. The pairs (YII' III ) that satisfy constraints (ii) and (iv') of (23.F.19) lie in the shaded set. Also drawn are two level sets of the objective function 111 - 20YII' The optimal pair (YII,III) for these values of a, II and ii2 is the point labeled (y~,/~). The corresponding values of IL and h in this interim incentive efficient social choice function are then Y! = I and I! = 20 + I~ - 20y~.
YH»
and Since YL = I and IL = 20 + III - 20YH in any interim incentive efficient social choice function, these expected utilities can be written as functions of only (YH,I H) as follows:
(23.F.19)
(rHEIO.IJ.f/l)
Y-
26YII
ii,",
Max
MECHANIS . . .
I, then
Suppose that the social choice function (y!,I!, y~,I~) is a solution to (23.F.17) in which constraint (ii) is not binding in the solution. Consider instead the social choice function (h, i L, ill' ill) = (y!,I! + £, y~ + £,I~ + 45£) for £ > O. For small enough £ > 0, this alternative social choice function satisfies ali of the constraints of problem (23.F.17) (note that it satisfies ill < I because, by step 2, y~ < I; check the other constraints too). Moreover, it yields a larger value of the objective function of (23.F.17) than (Y!, I!, y~, I~)-a contradiction. This establishes step 3. Slep 4: satisfied.
tH
I_
I_ =
BAYEIIAN
= y~ = I). But we have
already noted above that no such social choice function is incentive feasible (i.e., is an element of F·). Slep 3: equality).
OPTIMAL
Figure 23.F.1 (len)
Step 2: Any solulion 10 problem (23.F.Il) has YII < I; Ihal is, in any inlerim incenlive efficienl social choice function, trade does nOI occur with certainly when Ihe good is of high qualilY. (y!,I!, y~,I~) is ex post (classically) efficient (i.e., it has Y!
40y_ + ",_
23. F:
UI = 24
+ IH
- 24YH'
U2 = 8
+ 26YH
- I H·
In Figure 23.F.3, for an arbitrary point
efficient social choice functions.
909
910
CHAPTER
23:
INCENTIVES
AND
MECHANISM
---
DESIGN
v, VI Flgur. 23.F.3
YH
is therefore precisely the heavily traced line segment in Figure 23.F.3. 54 Notice that in every such social choice function the interim individual rationality constraint for a high-quality seller binds: the high-quality seller receives no gains from trade. _
APPENDIX A: IMPLEMENTATION AND MULTIPLE EQUILIBRIA
The notion of implementation that we have employed throughout the chapter (e.g., in Definition 23.B.4) is weaker in one potentially important respect than what we might want: Although a mechanism r may implement the social choice function f(·) in the sense of having an equilibrium whose outcomes coincide with f(·) for all oE 0, there may be other equilibria of r whose outcomes do not coincide with f(·). In essence, we have implicitly assumed that the agents will play the equilibrium that t he mechanism designer wants if there is more than one. ss This suggests that if a mechanism designer wishes to be fully confident that the mechanism r does indeed yield the outcomes associated with f(·), he might instead want to insist upon the stronger notion of implementation given in Definition 23.AA.\ (as in Definition 23.8.4, we are deliberately vague here about the eqUilibrium concept to be employed). Definition 23.AA.1: The mechanism r = (S" . .. ,S/, g(.» strongly implements social choice function f: 0, x ... x 0/ -+ X if every equilibrium strategy 54. Note that YH :s; 4/7 < 1 in any ex ante incentive efficient social choice function. This may seem at odds with our conclusion in Section Il.B that the ex snte efficient outcome that gives firms zero expected profits has all workers accepting employment in a firm (the structure of the model in Section I3.B parallels that here). The difference is that in Section 13.B we did not impose any interim individual rationality constraints on the workers, effectively supposing that the government could compel workers to participate (pay any taxes, etc.). Sec Exercise 23.F.10. 55. One possible argument rOf this assumption is that in a direct revelation mechanism that lruthfully implements social choice function f(·), the truth.telling equilibrium may be focal (in the sense discussed in Section 8.0).
The set of ex ante incentive efficient social choice functions corresponds to thOSe interim incentive efficient social choice functions with (YH,r ) lying in the heaVily H traced line segment.
APPENDIX
A'
IMPLEMENTATION
AND MULTIPLl
lQUILI.llrA
911
--------------------------------------------------------------profile (sr(·), ... , si(·» of the game induced by r has the property that g(sr(O,), ... , si(0rl) = f(O" ... , Orl for all (0" ... ,0/1. 56
Let us consider first the implications of this stronger concept for implementation in dominant strategy equilibria. In Exercise 23.C.8 we have already seen a case where, in the direct revelation mechanism that truthfully implements a social choice function f(·) in dominant strategies, some player has more than one dominant strategy, and when he plays one of these dominant strategies the outcome in f(·) does not result (see also Exercises 23.AA.I and 23.AA.2). Thus, with dominant strategy implementation we may have mechanisms that implement a social choice function f(·) but that do not strongly implement it. Nevertheless, there are at least two reasons why the multiple equilibrium problem may not be too severe with dominant strategy implementation. First, whenever each agent's dominant strategy in a mechanism r that implements f(·) is in fact a strictly dominant strategy (rather than just a weakly dominant one), mechanism r also strongly implements f(·). This is always the case, for example, in any environment in which agents' preferences never involve indifference between any two elements of X. Second, when a player has two weakly dominant strategies, he is of necessity indifferent between them for any strategies than the other agents choose. Thus, to play the "right" equilibrium in this case, it is only necessary that each agent be willing to resolve his indifference in the way we desire. In contrast, with Nash-based equilibrium concepts such as Bayesian Nash equilibrium, if mechanism r has two equilibria, then in each equilibrium each player may have a strict preference for his eqUilibrium strategy given that the other agents are playing their respective equilibrium strategies. Having agents play the Mright" equilibrium is then not just a matter of resolving indifference but rather of generating expectations that the desired eqUilibrium is the one that will occur. Example 23.AA.I illustrates the problem. Example 23.AA.l: Multiple Equilibria in the Expected Externality Mechanism. Consider again the expected externality mechanism or Section 23.D. Suppose that we are in a setting with two agents (I = 2) in which a decision must be made regarding a public project (see Example 23.B.3). The project may be either done (k - I) or not done (k = 0). Each agent's valuation (net of funding the project) is either OL or OH (so 0, = (OL' 0H} for i = 1,2), where OH > 0 > OL and OL + OH > O. The agents' valuations are statistically independent with Prob (0, = 8Ll = ). E (0, I) for I = 1,2. In the expected externality mechanism, each agent i announces his valuation and agent;os transfer when the announced types are (8" 8,) has the form t,(8" 8_,) = E._.[O_;k*(O;,IL;)] + h,(0_,),wherek*(8 I ,O,1 = OifO I = 0, = 8L ,and k*(8 .. 8,) = I otherwise. As we saw in Section 23.D, in one Bayesian Nash equilibrium of this mechanism, truth telling is each agent's equilibrium strategy. But this truth-telling equilibrium is not the only Bayesian Nash equilibrium. In particular, there is an equilibrium in which both agents always claim that 0H is their type. To see this, consider agent j's optimal 56_ The "strong" terminology is not standard; in the literature it is not uncommon, for example. to see the strong implementation concept simply referred to as .. implementation ....
912
c H ... PTE R
23;
INC E N T I V E S
... N D
M E C H ... N ISM
-
DES I G N
------------------------------------------------------------------------------strategy if agent - j will always announce 0H' Whichever announcement agent j makes, the project is done. Thus, regardless of his actual type, agent j's direct benefit (i.e., O,k·(O" 62 » is not affected by his announcement (it is OL if he is of type OL' and 0H if he is of type 0H)' It follows that agent j's optimal strategy is to make an announcement that maximizes his expected transfer. Now, agent j's expected transfer if he announces 0" is ().OL + (1 - )')OH) + h,(O,,), whereas if he announces OL his expected transfer is (I - )')OH + h,(OH)' Hence, agent j will prefer to announce 0" regardless of his type if agent - j is doing the same. It follows that both agents always announcing 0", and the project consequently always being done, constitutes a second Bayesian Nash equilibrium of this mechanism. _
A P PEN D I X
B:
IMP L E MEN TAT tON
IN
EN V I RON MEN T'
WIT H
COM P LET E
observe 0, there is still an implementation problem: Because no outsider (such as a court) will observe 0, the agents cannot write an enforceable ex ante agreement saying that they will choose outcome f(O) when agents' preferences are O. Rather, they can only agree to participate in some mechanism in which equilibrium play yields f(O) if 0 is realized. 57 Note that a complete information setting can be viewed as a special case of the general environment considered throughout this chapter, in which the probability density
e
We shall not pursue here the characterization of social choice functions that can be strongly implemented in Bayesian Nash equilibria. A good source of further reading on this subject is Palfrey (1992). We also refer to Appendix B, where we discuss many of these issues for the special context of complete information environments. There are, however, two important points about strong implementation that we wish to stress here. First, when trying to strongly implement a social choice function f('), we cannol generally restrict attention to direct revelation mechanisms. The reason is that when we replace a mechanism r = (S" ... ,S" g('» with a direct revelation mechanism, as envisioned by the revelation principle, we may introduce new, undesirable equilibria. (See Exercises 23.C.8 and 23.AA.l for an illustration.) Second, because a social choice function f(·) can be strongly implemented only if it can be implemented in the weaker sense studied in the text of the chapter, all of the necessary conditions for implementation that we have derived arc still necessary for f(·) to be strongly implemented. Thus, for example, the conclusions of the GibbardSatterthwaite theorem (Proposition 21C.3), the revenue equivalence theorem (Proposition 23.D.3), and the Myerson-Satterthwaite theorem (Proposition 23.E.l) all continue to be valid when we seek strong implementation.
Definition 23.88.1: The mechanism r = CS" ... , SI' g(.)) implements the social choice lunction IC') in Nash equilibrium if, for each profile of the agents' preference parameters 0 = (0" . .. ,0Jl E e, there is a Nash equilibrium of the game induced by r, 5·(0) = Cst(O), . .. ,s1CO)), such that g(s*CO)) = 1(0). The mechanism r = C5" ... , 5" g(.)) strongly implements the social choice lunction (C·) in Nash equilibrium if, for each profile of the agents' preference parameters 0= (0" . .. , Otl E e, every Nash equilibrium of the game induced by r results In outcome 1(0). The first point to note about implementation in Nash equilibria is that if we are satisfied with the weaker notion of implementation that we have employed throughout the text of the chapter (as opposed to the strong implementation concept discussed in Appendix A), then any social choice function can be implemented in Nash equilibrium as long as I <.>: 3. To see this, consider the following mechanism: each agent j simultaneously announces a profile of types for each of the I agents. If at least I - 1 agents announce the same profile, say ~, then we select outcome f(~).
Throughout the chapter we have restricted attention to single-valued social choice functions. It is sometimes natural, however, to consider social choice correspondences that can specify more than one acceptable alternative for a given profile of agent types. In this case, we would say that mechanism r = (S" ... , S" g(.)) strongly implements the social choice correspondence f(') if every equilibrium s·(·) of the game induced by r has the property that g(0·(8)) e flO), that is, if, for every 0, all possible equilibrium outcomes are acceptable alternatives according tof(·)·
57. This type of setting is often natural in contracting problems, where it is frequently reasonable to suppose lhal the parlies will come to know a 101 about each olher thaI is not verifiable by any outside enforcer of their contract.
APPENDIX B: IMPLEMENTATION IN ENVIRONMENTS WITH COMPLETE INFORMATION
58. Thus, we can think of the complete information environmenl as a case in which agents receive signals lhat are perfectly correlated. There are several ways to formalize this. Perhaps the simplest is to suppose that each agent i's preference parameler is drawn from some set e,. An agent's signal (or Iype), which is now represenled by 8, - (8", ... ,8,,) e e, is a vector giving agent i 's observation of his and every other agent'. preference parameters. Thus, the set ofpouible "types" for agent i in the sense in which we have used this term throughout the chapter, is now &, - e for each; = 1, ... , I. The probability density .p(.) on the set of possible types 9, x··· X &, then satisfies lhe property that (ii" ... , ii,) > 0 if and only if ii, z··· = ii" and agent i's Bernoulli utility funclion has the form u.(x, iii) = ii.(x, 8,,).
In this appendix we provide a brief discussion of implementation in complete information environments. An excellent source for further reading on this subject is the survey by Moore (1992) [sec also Maskin (1985»). In the complete information case we suppose that each agent will observe not only his own preference parameter 0" but also the preference parameters 0 _/ of all other agents. However, while the agents will observe each others' preference parameters, we suppose that no outsider does. Thus, despite the agents' abilities to
59. The same is true for strong implementation in dominant strategies.
...
IN FOR MAT ION
913
914
C HAP T E R
23:
INC E N T I V E SAN D
M E C HAN ISM
DEI I G N
Otherwise we select outcome Xo E X (xo is arbitary). With this mechanism, for every profile IJ, there is a Nash equilibrium in which every agent announces IJ, and the resulting outcome is f(IJ), because no agent can affect the outcome by unilaterally deviating. Although this mechanism implements f(·) in Nash equilibrium, it is obviously not a very attractive mechanism [i.e., its implementation of f(·) is not very convincing] because there are so many other Nash equilibria that do not result in outcome f(lJ) when the preference profile is IJ. Indeed, with this mechanism, given the profile of preference parameters IJ, there is a Nash equilibrium resulting in x for every x E f(0) = (x E X: there is a IJ E 0 such that f(lJ) = x}. We see then that for Nash implementation with I ?; 3, the entire problem of satisfactorily implementing a given social choice function revolves around the issue of successfully dealing with the multiple equilibrium problem discussed in Appendix A. Given this observation, what social choice functions can we strongly implement in Nash equilibrium? The simple but powerful result in Proposition 23.BB.1 comes from Maskin's (1977) path-breaking paper on Nash implementation. Proposition 23.88.1: If the social choice function f(') can be strongly implemented in Nash equilibrium, then f(·) is monotonic· o Proof: Suppose that r = (S" ... , S,' g(.» strongly implements f(·). Then when the preference parameter profile is IJ, there is a Nash equilibrium resulting in outcome f(O); that is, there is a strategy profile s* = (st, . .. , s1) having the properties that g(s*) = flO) and g(§ .. s!,) E L;(f(IJ), IJ,) for all §, E S, and all i. o , Now suppose that f (.) is not monotonic. Then there exists another profile of preference parameters IJ' E 0 such that L;(f(IJ), IJ,) c L;(f(IJ), 1Jj) for all ~ but flO') #' f(IJ). But s' is also a Nash equilibrium under preference parameter profile 0', because g(§" s!;) E L;(f(O), Oil for all §, E S, and all i. Hence, r does not strongly implement J( .I-a contradiction. As an illustration of the restriction imposed by monotonicity, Proposition 23.BB.2 records one implication of this result. Proposition 23.88.2: Suppose that X is finite and contains at least three elements, that iJI, = 9' for all i, and that f(0) = X. Then the social choice function f(') is strongly implementable in Nash equilibrium if and only if it Is dictatorial. Proof: To strongly implement a dictatorial social choice function in Nash equilibrium, we need only let the dictator choose an alternative from X. In the other direction, the result follows from steps 2 and 3 of the proof of the Gibbard-Satterthwaite theorem (Proposition 23.C.3). • One lesson to be learned from Propositions 23.BB.1 and 23.BB.2 is that dealing with the multiple equilibrium problem can potentially impose very significant restrictions on the set of implementable social choice functions.
-- -
APPENDIX
B:
IMPLEMENTATION
preference parameter is 0,.
ENVIRONIlENTI
WITH
COMPLETE
Maskin (1977) also showed that, when 1> 2, monotonicity is almost, but not quite, a sufficient condition for slrong implementalion (we omit the proof; see Moore (1992) for a discussion of this and more recent results, including consideration of the case 1 - 2). Maskin's added condition, known as no veto power, requires that if 1 - I agents all rank some alternative x as their best alternative then x = [(0).
Propo.ltlon 23.88.3: If I
~
3, (.) is monotonic, and f(') satisfies no veto power. then ft·) is strongly
implementable in Nash equilibrium.
No veto power should be thought of as a very weak additional requirement; indeed, in any setting in which there is a desirable and transferable private good, it is trivially satisfied: no two agents then ever have the same top· ranked alternative (each wants to get all of the available transferable private good). Thus, in these commonly studied environments, monotonicity of [e) is a necessary and sufficient condition for [(.) to be strongly implement able in Nash equilibrium. multivalued social choice correspondence [(.) is said to be monotonic if whenever E [(0'). The necessary and sufficient conditions for Nash implementation in Propositions 23.BB.1 and 23.BB.3 carryover to the multivalued case. (In fact, for Proposition 23.B8.3, Maskin's result actually establishes that there is a mechanism that has, for each preference profile 0, a set of Nash equilibrium outcomes exactly equal to the set [(0); this type of implementation of a social choice correspondence is commonly called [ull implement arion.] It may be verified, for example, that if [(0) is equal to the Pareto set for all 0 (the set of all ex post efficient outcomes in X given preference profile 0), then [(.) is monotonic. Thus, in any setting with a transferable good, the Pareto set social choice correspondence is strongly implementable in Nash equilibrium. A
x
E
[(0) and L,(x. 0,) c L,{x, 0;) for all i then x
Implementation using Extensive Form Games: Subgame Perfect Implementation So far we have seen that the need to ~ knock out" undesirable equilibria (formalized through the notion of strong implementation) can significantly restrict the set of implementable social choice functions. This suggests the possibility that we may be more successful if we use instead a refinement of the Nash equilibrium concept. Indeed, recent work has shown that such refinements can be very powerful. Here we briefly illustrate how, by considering dynamic mechanisms and employing the equilibrium concepts ror dynamic games discussed in Section 9.B, we can expand the set of strongly implementable social choice functions. Example 23.BB.1: [Adapted from Moore and Repullo (1988)]. Consider a pure exchange economy (see Example 23.B.2) with two consumers in which each consumer has two possible individualistic preference relations: if 0; = of, then consumer i has Cobb-Douglas preferences; if IJ; = IJ~, then consumer i has Leontief preferences. These two possible preference relations for consumers I and 2 are depicted in Figure 23.BB.1. Suppose that we wish to strongly implement the social choice function
f(lJ) = 60. Sec Definition 23.C.S. 61. Recall that LAx, 8,) c X is agent i's lower contour set for outcome x E X when agent i's
'N
Xc
if O. =
Of,
{ XL
ifO. =
ot,
where XC and XL are the allocations depicted in Figure 23.BB.1. Note that consumer 1 always prefers XC to XL, and the reverse is true for consumer 2. By inspection or
INFORMATION
915
916
CHAPTER
23:
INCENTIVES
AND
MECHANISM
-
DESIGN
----r.=!1<=:::±=----t------,O,
REF ERE NeE.
Barbera. S., and B. Peleg. (1990), Strategy-proor voting schemes with continuous preferences. Social Choice and Welfare 7: 3t-38. Baron, D., and R. B. Myerson. (1982). Regulating a monopolist with unknown costs. Econometrica
so: 911-30. Bulow. J.. and J. Roberts. (1989). The simple economics of optimal auctions. Journal oj Political Economy
97: 1060-90. Clarke. E. H. (1971). Multipart pricing or public goods.
P.blic Choice 1: 19-33. Cramton. P., R. Gibbons, and P. Klemperer. (1987). Dissolving a partnership efficiently. Econometrica 55:
-~,(O~) I
~,(O~)
615-32. Dana, 1. D., Jr., and K. Spier. (1994). Designing a private industry: Government auctions with endogenous market structure. Journal of Public Economics 53: 127-47. Dasgupta, P., P. Hammond. and E. Maskin. (1979). The implementation of social choice rules: Some general results on incentive compatibility. Review of Economic Studies 46: 185-216. d'Aspremont. c., and L. A. Gerard-Varer. (1979). Incentives and incomplete information. JourMI 0/ Public
~,(O~)
i i
~,(O~)
Fig.,. 23.BB.l
Preferences and outcomes in Example 23.BB.1.
EnmomiL's 11: 25-45.
Fudenberg. D.. and J. Tirole. (1991). Game Theory. Cambridge. Mass.: MIT Press. Gibbard. A. (1973). Manipulation oC voting schemes. Economelrica 41: 587-601.
Figure 23.B8.I, we see that f(·) is not monotonic, because L,(x c, Of) c L,(x c, O~) for i = 1,2 but f(Or, O~) ¥- f(O~, O~). Hence, by Proposition 23.BB.I, f(·) cannot be strongly implemented in Nash equilibrium. Suppose, instead, that we construct the following three-stage dynamic mechanism: Stage /: Slage 2: Stage 3:
917
----------------------------------------------------------------------------------
Green. J. R.. and 1.-J. Laffonl. (1977). Characterization of satisfactory mechanisms for the revelation of preferences for public goods. EnmonU'lricu 45: 427-38. Green. J. R.• and J.-J. Laffonl. (I 979).lncentivt's in Public Decision Making. Amsterdam: North-Holland. Gresik. T. A., and M. A. Sallerthwaite. (1989). The rate at which a simple market becomes efficient as the number of traders increases: An asymptotic result for optimal trading mechanisms. Jourrwt of Gccmomic 11IC'ory 48: 304- 32. Groves, T. (1973). Incentives in teams. Econametrka 41: 617-31. Hulmstrom. B., and R. B. Myerson. (1983). Efficient and durable decision rules with incomplete information. E"onomelrica 51: 1799-1819. Hurwicz. L. (1972). On informationally decentralized systems. In Decision and Organization, edited by C. B. McGuire. and R. Radner. Amsterdam: North-Holland. Laffont. J.-J .• and E. Maskin. (1980). A differential approach to dominant strategy mechanisms. Econometrica 48: 1507-20.
Agent 1 announces either "L" or "c" If he announces "L," XL is immediately chosen. If he announces "C," we go to stage 2. Agent 2 says either "agree" or "challenge." If he says "agree," then XC is immediately chosen. If he says "challenge," then we go to stage 3. Agent I chooses between the allocations y and z depicted in Figure 23.BB.1.
°
It is straightforward to verify that, for each possible profile of preferences 0 = (0" 2 ), the unique subgame perfect Nash equilibrium of this dynamic game of perfect information results in outcome flO) (see Exercise 23.BB.I). Thus, f(·) can be strongly implemented if we consider dynamic mechanisms and take subgamc perfect Nash equilibrium as the appropriate solution concept for the games induced by these mechanisms. _
Maskin. E. (1977). Nash equilibrium and wef[are optimality. MIT Working Paper. Maskin. E. (1985). The theory oC implementalion in Nash equilibrium: A survey. In Social Goals lJnd Social Organization: Essays in Honor 0/ Elisha Pazner. edited by L Hurwicz" D. Schmeid1er. and H. Sonnenschein. Cambridge. U.K.: Cambridge University Press. Maskin, E.• and 1. Riley. (1984). Monopoly with incomplete information. Rand Journal of Economics IS:
171-96. McAfee. R. P., and J. McMillan. (1987). Auctions and bidding. Journal of Economic Literature lS: 699-738. Milgrom. P. R. (1987). Auction theory. In Advances in Economic Theory: Fifth World Congress.
In fact, Moore and Repullo (1988) [see also Moore (1990)] show that the use of dynamic mechanisms and subgame perfection expands the set of strongly implementable social choice functions dramatically compared to the use of the Nash equilibrium concept. Even stronger results are possible with other refinements; see, for example, Palfrey and Srivastava (1991) for a study of strong implementation in undominated Nash equilibrium (i.e., Nash equilibria in which no agent is playing a weakly dominated strategy)·2
edited by T. BeWley. Cambridge. U.K.: Cambridge University Press. Mirrlccs. J. (1971). An exploration in the theory or optimal income taxation. Review of Economk Studies 3H: 175-208. Moore. 1. (1992). Implementation. contracts. and renegotiation in environments with complete information. In Advan£·t>!. in Economic Theory: Sixth World Congress. vol. I. edited by J..J. Laffont. Cambridge, U.K.: Cambridge University Press. Moore. 1.. and R. Repullo. (1988). Subgame perfect implementation. Econometrica 56: 1191-1220. Myerson, R. B. (1979). Incentive compatibility and the bargaining problem. EconOnu!lrica 47: 61-73. Myerson. R. B. (1981). Optimal auction design. MatMmotics a/Operation Re..arch 6: 58-73. Myerson. R. B. (199t). Game Theory: Analysis a/Conflict. Cambridge. Mass.: Harvard University Press.
REFERENCES
Myerson. R. B.• and M. A. Satterthwaite. (1983). Efficient mechanisms Cor bilateral trading. Economic Theory 18: 265-81.
Abreu, D., and H. Matsushima. (1994). Exact implementation. JourntJl of Economic Theory 64: 1-19. Arrow, K. (1979). The properly rights doctrine and demand revelation under incomplete inronnation. In Economics and Human Welfare. edited by M. Boskin. New York: Academic Press.
Jour",,1
0/
Palfrey. T. R. (1992). Implementation in Bayesian equilibrium: The multiple equilibrium problem in mechanism design. [n Advances in Economic Theory: Sixth World Congress, vol. I, edited by J..J. Laffonl. Cambridge. U.K.: Cambridge University Press. Palfrey. T.. and S. Srivastava. (1991). Nash implementation using undominated strategies. Econometrica
62. Very positive results have also been obtained recently for implementation in iteratively
undominaled Slrategies; see. for example, Abreu and Matsushima (1994).
59: 479-501.
....
918
C HAP T E R
23:
INC E N T I V E 8
AND
ME C HAN 18 MOE. I
Q
N
-------------------------------------------------------Satterthwaite, M. A. (1975). Strategy-proofness and Arrow's conditions: Existence: and correspondence theorems for voting procedures and social welfare functions. Journal of Economic Theory 10: 187-217. Vickrey. W. (1961). Counterspeculation, auctions, and competitive sealed tenders. Journal of Finance 16: 8-37.
-
I . E" C I • E •
alternative x, if more agents prefer x, over xJ than prefer xJ over x, (it may select either x, or x, if the number of agents preferring x, over x, equals the number preferring x, over x,)] is truthfully implementable in dominant strategies. 23.C.3 A Show that when 91, = i? for all i. any ex post efficient social choice function J(.) has
[(e)=x. EXERCISES
23.B.l A Consider the setting explored in Example 23.B.I. where 91, = {:t,(O')l and 91, = {:t,(Oi). :t,(Oi)l. For each of the following social choice functions J(.). will agent 2 be willing to truthfully reveal his preferences?
[(0,. Oil = y. f(o,. 0;) = y. (b) f(O,.Oi) = z. f(O,. Oil = x. (e) [(0,. 0i) = z. [(0,. Oil = y. (a)
[(0,. Oil = y,f(O,. 0;) =
I:
e
=
23.C.sc Show that in an environment with single-peaked preferences having the property that no two alternatives are indifferent (see Section 21.0) and an odd number of agents, the (unique) social choice function that always selects a Condorcet winner (see Section 21.0) is truthfully implementable in dominant strategies. c
The property of a social choice function identified in Proposition 23.C.2 is called
independent person-by-person monotonicity (lPM) by Dasgupta, Hammond, and Maskin (1979). In this exercise, we investigate its relationship with the monotonicity property defined in
z.
23.B.2A Consider a bilateral trade setting (see Example 23.B.4) in which both the seller's (agent I) and the buyer's (agent 2) types are drawn independently from the uniform distribution on [0. I]. Suppose that we try to implement the social choice functionJ( ,)=(y,(. ).y,( .). I,O./,{'» such that y,(O,. 0,) = \ if 0, ~ 0,; =0 if 0, < 0,. y,(O,. 0,) = \ if 0, I, (0,.0,)
e,
23.C.6
(d) f(O,.021 = z,f(O,. Oil = z. (e)
ee,
23.C.4A Show that if [: X is truthfully implementable in dominant strategies when the set of possible types is for i = I•...• I. then when each agent i's set of possible types is c e i (for i = 1, ... ,1) the social choice function 9 - X satisfying 1(0) [(0) for all oE is truthfully implementable in dominant strategies.
= «0,
Definition 23.C.5. (a) Show by means of an example that if [(.) satisfies IPM, it need not be monotonic (this can be done with a very simple example).
(b) Show by means of an example that if [(.) is monotonic, it need not satisfy IPM.
> 0,; =0 if 0, SO,.
(e) Prove that if [(.) satisfies IPM, and if 91, c i? for all i. then J(.) is monotonic.
+ O,)y,(O,. 0,).
1,(0,.0,) = -!(O,
(d) Prove that if [(.) is monotonic, and 91, = i? for all i. then J(') satisfies IPM.
+ O,)y,(O,. 0,).
Suppose that the seller truthfully reveals his type for all 0, worthwhile to reveal his type? Interpret.
E
[0. I]. Will the buyer find it
23.B.3 8 Show that b,(O,) = 0, for all 0, E [0. I] is a weakly dominant strategy for each agent i in the second-price sealed-bid auction. 23.B.4 c Consider a bilateral trade setting (see Example 23.B.4) in which both the seller's and the buyer's types are drawn independently from the uniform distribution on [0. I]. (a) Consider the double auclion mechanism in which the seller (agent I) and buyer (agent 2) each submit a sealed bid. bi ~ O. If b, ~ b,. the seller keeps the good and no monetary transfer is made; while if b, > b" the buyer gets the good and pays the seller the amount !
°_,.
23.C.2 8 Show that. for any I. when X contains two elements (say. X = {x,. x,}). then any majority voting social choice function [i.e.• a social choice function that always chooses
13.C.,c A social welfare functional F(·) (see Section 21.C) satisfies the property of nonnegative responsivene55 if for all x. y EX, and for any two pairs of profiles of preferences for the I agents (:t" .... :t,) and (:t; .... ,:til such that x :toY =- x:t; y and x >-,Y" x >-; y for all ~ we have x F(:t, .....
:t,)y =-xF(:t;, ... , :t;)y
and X
F,(:t, •... , :t,) y ... x
Fp(:t', .... , :t;) y,
where x F,( . ) y means • x F( . ) y and not y F( .) x." Show that if the social choice function [(.) maximizes a social welfare functional F{') satisfying nonnegative responsiveness [in the sense that for all (0, •... ,0,) we have [(0, •...• 0,) {x e X: x F(:t,(O,), ... , :t,(O,» y for all Y E X lJ. then [(.) is truthfully implementable in dominant strategies.
=
A 23.C.S Suppose that 1=2, X = {a, b, c, d, el, = {O',' O;l, and = {O" Oil, and that the agents' possible preferences are (a-b means that alternatives a and b are indifferent):
e,
e,
a-b
a b
a-b
a b
d
d
d
d
919
920
CHAPTER
23:
INCENTIVES
AND
MECHANISM
---
DEIIGN
Consider the social choice function 1(0) =
{ab
if 0 = (0;, 0i), otherwise.
--
EXERCISES
immediately when the seller's valuation is 0 and the price agreed to when the seller has valuation 0, is (10 + 0,)/2. What is the earliest possible time at which trade can occur when the seller's valuation is 9? 23.0.2" Consider a bilateral trade selling in which each 0, (i = 1,2) is independently drawn from a uniform distribution on [0, I].
(a) Is 1(-) ex post efficient? (b) Does it satisfy the property identified in Proposition 23.C.2?
(a) Compute the transfer functions in the expected externality mechanism.
(c) Examine the direct revelation mechanism that truthfully implements 1(-). Is truth telling each agent's unique (weakly) dominant strategy? Show that if an agent chooses his untruthful (weakly) dominant strategy, then 1(') is not implemented.
(b) Verify that truth telling is a Bayesian Nash equilibrium. 23.03' Reconsider the first-price and second-price sealed-bid auctions studied in Examples 23.8.5 and 23.B.6. Verify that the revenue equivalence theorem holds for the equilibria identified there.
23.C.9c Suppose that K = R, the v,(', 0,) functions are assumed to be twice continuously differentiable, O. is drawn from an interval [Q" Ii.], o'v,(k, O,)/ok' < 0, and o'v,(k, O,)/ok 00, > O. Show that the continuously differentiable soeial choice function f(-) - (k(·),I,(·), ... , 1,('» is truthfully implementable in dominant strategies if and only if, for all i = I, ... ,I,
23.0.4 c Consider a first-price sealed-bid auction with I symmetric buyers. Each buyer's valuation is independently drawn from the interval [~, 0] according to the strictly positive density 1/>(').
k(O) is nondecreasing in 0,
and
(a) Show that the buyer's equilibrium bid function is nondecreasing in his type. 1,(0,,0 _,) = 1,(0,.0 _.) _ -
f."
(b) Argue that in any symmetric equilibrium (b·(·), . .. ,b*('» there can be no interval of types (0',0"),0' '" 0", such that b'(O) is the same for all 0 E (0', 0"). Conclude that b'(') must therefore be strictly increasing.
ov.(k(s, 0 -.), s) ok(s, 0 -,) ds.
!,
ok
Os
23.C.IO" (8. Holmstrom) Consider the quasilinear environment studied in Section 23.C. Let
(c) Argue, using the revenue equivalence theorem, that any symmetric equilibrium of such an auction must yield the seller the same expected revenue as in the (dominant strategy) equilibrium of the second-price sealed-bid auction.
k'(') denote any project decision rule that satisfies (23.C.7). Also define the function V'(O) = L:, v,(P(O), 0,).
(a) Prove that there exists an ex post efficient soeial choice function [i.e., one that satisfies condition (23.C.7) and the budget balance condition (23.C. I 2)] that is truthfully implementable in dominant strategies if and only if the function V·(·) can be written as V'(O) = L:, 1'1(0 _,) for some functions V,(·)" .. , 1'1(.) having the property that V,(.) depends only on 0_, for all i.
=
23.0.Sc For the same assumptions as in Exercise 23.0.4. consider a sealed-bid all-pay auction in which every buyer submits a bid, the highest bidder receives the good, and every buyer pays the seller the amount of his bid regardless of whelher he wins. Argue that any symmetric equilibrium of this auction also yields the seller the same expected revenue as the sealed-bid second-price auction. [Hint: Follow similar steps as in Exercise 23.0.4.]
=
(b) Use the result in part (a) to show that when 1=3, K R, 9, R. for all i, and v,(k, 0,) = O,k - (!)k' for all i an ex posteffieient social choice function exists that is truthfully implementable in dominant strategies. (This result extends to any I > 2.)
23,0,6c Suppose that I symmetric individuals wish to acquire the single remaining ticket to a concert. The ticket office opens at 9 a.m. on Monday. Each individual must decide what time to go to get on line: the first individual to get on line will get the ticket. An individual who waits I hours incurs a (monetary equivalent) disutility of {JI. Suppose also that an individual showing up after the first one can go home immediately and so incurs no waiting cost. Individual i's value of receiving the ticket is 0., and each individual's 0, is independently drawn from a uniform distribution on [0, I]. What is the expected value of the number of hours that the first individual in line will wait? [Hint: Note the analogy to a first-price sealed-bid auction and use the revenue equivalence theorem.] How does this vary when {J doubles? When I doubles?
(c) Now suppose that the v,(k, 0,) functions are such that V·(·) is an I-times continuously differentiable function. Argue that a necessary condition for an ex post efficient social choice function to exist is that, at all 0, o'V'(O) =
o.
00, ... 00, (In fact, this is a sufficient condition as well.) (d) Usc the result in (c) to verify that, under the assumptions made in the small type discussion at the end of Section 23.C, when I = 2 no ex post effieient social choice function is truthfully implementable in dominant strategies.
23.E.I" Consider again a bilateral trade selling in which each 0, (i = 1,2) is independently drawn from a uniform distribution on [0, I]. Suppose now that by refusing to participate in the mechanism a seller with valuation 0, receives expected utility 0, (he simply consumes the good). whereas a buyer with valuation 6, receives expected utility 0 (he simply consumes his endowment of the numeraire, which we have normalized to equal 0). Show that in the expected externality mechanism there is a type of buyer or seller who will strictly prefer not to participate.
23.C.II A Consider a quasilinear environment, but now suppose that each agent i has a Bernoulli utility function of the form u,(v,(k, 0,) + m, + I,) with u;(') > O. That is, preferences over certain outcomes take a quasilinear form, but risk preferences arc unrestricted. Verify that Proposition 23.C.4 is unaffected by this change. 23.0.1" [Based on an example in Myerson (1991)] A buyer and a seller arc bargaining over the sale of an indivisible good. The buyer's valuation is O. K 10. The seller's valuation takes one of two values: 0, e {O, 9}. Let I be the period in which trade occurs (I = 1,2, ... ) and let p be the price agreed. Both the buyer and the seller have discount factor lJ < I.
23.E.2A Argue that when the assumptions of Proposition 23.E.I hold in the bilaterai trade selling: (a) There is no social choice function 1(') that is dominant strategy incentive compatible and interim individually rational (i.e., that gives each agent i nonnegative gains from participation conditional on his type 0" for all 0,).
(a) What is the set X of alternatives in this selling? (b) Suppose that in a Bayesian Nash equilibrium of this bargaining process, trade occurs
...
921
922
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESIGN
-----------------------------------------------------------------------------(b) There is no social choice function f(·) that is Bayesian incentive compatible and ex post individually rational [i.e., that gives each agent nonnegative gains from participation for every pair of types (9" 9,». 23.E.3" Show by means of an example that when the buyer and seller in a bilateral trade selling both have a discrete set of possible valuations, social choice functions may exist that are Bayesian incentive compatible, ex post efficient, and individually rational. [Hint: It suffices to let each have two possible types.] Conclude that the assumption of a strictly positive density is required for the Myerson-Sallerthwaite theorem. 23.E.4" A seller (i = I) and a buyer (i = 2) are bargaining over the sale of an indivisible good. Trade can occur at discrete periods t = 1,2, .... Both the buyer and the seller have discount factor ~ < I. The buyer's and seller's valuations are drawn independently with positive densities from [~" 0,] and [~" 0,], respectively. Assume that (Q" 0,)" (~" 0,) # 0. Note that in this selling ex post efficiency requires that trade occur in period I whenever 0, > 0" and that trade not occur whenever 0, > 0,. Use the Myerson-Sallerthwaite theorem to show that, in this selling with discounting, no voluntary trading process can achieve ex post efficiency. 23.E.S" Suppose there is a conrinuum of buyers and sellers (with quasilinear preferences). Each seller initially has one unit of an indivisible good and each buyer initially has none. A seller's valuation for consumption of the good is 0, e [Q" ii,], which is independently and identically drawn from distribution CI>,(') with associated strictly positive density 4>,('). A buyer's valuation f~om consumption of the good is 0, E [Q" 0,], which is independently and identically drawn from distribution CI>,(') with associated strictly positive density 4>,c-). (_) Characterize the trading rule in an ex post efficient social choice function. Which buyers and sellers end up with a unit of the good? (b) Exhibit a social choice function that has the trading rule you identified in (a), is Bayesian incentive compatible, and is individually rational. [Hint: Think of a "competitive" mechanism.] Conclude that the inefficiency identified in the Myerson-Sallerthwaite theorem goes away as the number of buyers and sellers grows large. [For a formal examination showing that, with a finite number of traders, the efficiency loss goes to zero as the number of traders grows large, see Gresik and Sallerthwaite (1989).] 23.E.6" Consider a bilateral trading selling in which both agents initially own one unit of a good. Each agent i's (i = 1,2) valuation per unit consumed of the good is 0,. Assume that 0, is independently drawn from a uniform distribution on [0, 1]. (a) Characterize the trading rule in an ex post efficient social choice function. (b) Consider the following mechanism: Each agent submits a bid; the highest bidder buys the other agent's unit of the good and pays him the amount of his bid. Derive a symmetric Bayesian Nash equilibrium of this mechanism. [Hine: Look for one in which an agent's bid is a linear function of his type.] (0) What is the social choice function that is implemented by this mechanism? Verify that it is Bayesian incentive compatible. Is it ex post efficient? Is it individually rational [which here requires that U,(O,) ~ 9, for all 9, and i = 1,2]? Intuitively, why is there a difference from the conclusion of the Myerson-Satterthwaite theorem? [See Cramton, Gibbons, and Klemperer (1987) for a formal analysis of these "partnership division" problems.]
23.E.7" Consider a bilateral trade setting in which the buyer's and seller's valuations are drawn independently from the uniform distribution on [0, I]. (0) Show that if f(·) is a Bayesian incentive compatible and interim individually rational
EXEIICIIEI923
----------------------------------------------------------------------------------social choice function that is ex post efficient, the sum of the buyer's and seller's expected utilities under f(·) cannot be less than 5/6. (b) Show that, in fact, there is no social choice function (whether Bayesian incentive compatible and interim individually rational or not) in which the sum of the buyer's and seller's expected utilities exceeds 2/3. 23.F.l c Consider the quasilinear setting studied in Sections 23.C and 23.D. Show that if the social choice function f(·) e F' is ex post classically efficient in FlO then it is both ex ante and interim incentive efficient in P. [From this fact, we see that if an ex post classically efficient social choice function can be implemented in a setting with privately observed types (i.e., if it is incentive feasible), then no other incentive feasible social choice function can welfare dominate it. Note, however, that there may be other ex ante or interim incentive efficient social choice functions that are not ex post efficient; for example, you can verify that in Example 23.F.1 there is an ex post classically efficient social choice function that is incentive feasible, but the particular interim incentive efficient social choice function derived in the example is not ex post efficient.] 23.F.2" [Based on Maskin and Riley (1984)] A monopolist seller produces a good with constant returns to scale at a cost of c > 0 per unit. The monopolist sells to a consumer whose preference for the product the monopolist cannot observe. A consumer of type 9 > 0 derives a utility of Ov(.<) - I when he consumes .< units of the monopolist's product and pays the monopolist a total of I dollars for these units. Assume that v'(·) > 0 and VO(.) < O. The set of possible COnsumer types is [q,O] with > Q > 0, and the distribution of types is 4>('), with an associated strictly positive density function 4>(') > O. Assume that [9 - «I - 4>(0))/4>(9))] is nondecreasing in O. Characterize the monopolist's optimal selling mechanism to this consumer, assuming that a consumer of type 0 can always choose not to buy at all, thereby deriving a utility of O.
°
23.F.3 c An auction with a reserve price is an auction in which there is a minimum allowable bid. Suppose that in the auction selling of Example 23.F.2 the I buyers are symmetric and that Q = O. Argue that a second'price sealed·bid auction with a reserve price is an optimal auction in this case. What is the optimal reserve price? Can you think of a modified second·price sealed·bid auction that is optimal in the general (nonsymmetric) case? 23.F.4" Derive the optimal y,{') functions in the auction selling of Example 23.F.2 when the seller's valuation for the object is 00 > O. 23.F.5 B Suppose that a monopolist seller who has two potential buyers has a total of one divisible unit to sell; that is, production costs are zero up to one unit, and infinite beyond that. The demand function of buyer i is the decreasing function x,(p) for i = 1,2. The monopolist can name distinct prices for the two buyers. (a) Characterize the monopolist's optimal prices. (b) Relate your answer in (0) to the optimal auction derived in Example 23.F.2. [For more on this, see Bulow and Roberts (1989).] 23.F.6 c [Based on Baron and Myerson (1982)]. Consider the optimal regulatory scheme for a regulator of a monopolist who has known demand function x(p), with x'(p) < 0, and a privately observed constant marginal cost of production 9. The regulator can set the monopolist's price and can make a transfer from or to the monopolist, so the set of outcomes is X = {(p, I): p > 0 and Ie R}. The regulator must guarantee the monopolist a nonnegative profit regardless of his production costs to prevent the monopolist from shutting down. The monopolist's marginal cost 0 is drawn from [~, 0] with 8 > ~ > 0 according to the distribution function <1>('), which has an associated strictly positive density function 4>(') > O. Assume that
924
CHAPTER
23:
INCENTIVES
AND
MECHANISM
DESION
$(O)/IIJ(O) is nondecreasing in O. Denote a type-O monopolist's profit from outcome (p,l) by "(p, I, 0) = (p - O)x(p)
+ I.
-
EXERCISEI925
-------------------------------------------------------------------------------------------normalized valuation function, that is, a function such that v.(k o) = O. Suppose that kO(.) and the Groves transfers are calculated using these announcements. Does each agent have a unique (weakly) dominant strategy in this normalized Groves mechanism?
(a) Adapl the characterization in Proposition 23.0.2 to this application.
23.BB.IA Consider the dynamic mechanism in Example 23.BB.1.
(b) Suppose that the regulator wants to design a direct revelation regulatory scheme
(a) For each possible preference profile, write down its normal form and identify its Nash equilibria.
(p('), 1(')) that maximizes the expected value of a weighted sum of consumer and producer
surplus,
f-(I,~
xis) ds
(b) For each possible preference profile, identify this mechanism's subgame perfect Nash equilibria.
+ ",,(P(O), 1(0), 0),
where" < I. Characterize the regulator's optimal regulatory scheme. What if " :!: 1?
23.BB.28 Is a social choice function that is implementable in dominant strategies necessarily implementable in Nash equilibrium? What if we are interested in strong implementation instead?
23.F.7c [Based on Dana and Spier (1994)] Two firms, j = 1,2. compete for the right to produce in a given market. A social planner designs an optimal auction of production rights to maximize the expected value of social welfare as measured by
w=
l>J + S + (;' J
23.BB-3c Consider a setting of public project choice (see Example 23.8.3) in which K = {O, II· Let 0, denote agent i's benefit if the project is done (i.e., if k = 1); normalize the value from k = 0 to equal zero. Assume that El, = R. In this setting, the only mechanisms that involve an ex post efficient project choice are Groves mechanisms. Let kO(.) denote the project choice rule in such a mechanism. Also, suppose that I ~ 3. The transfers in a Groves mechanism are characterized by two properties:
I) ~:rJ' J
where I; denotes the transfer from firm j to the planner, S is consumer surplus, "; is the gross (pretransfer) profit of firmj, and;' > I is the shadow cost of public funds. The auction specifies transfers for each of the firms and a market structure; that is, it either awards neither firm production rights, awards only one firm production rights (thereby making that firm an unregulated monopolist), or gives production rights to both firms (thereby making them compete as unregulated duopolists). Each firm j privately observes its fixed cost of production OJ. The fixed cost levels 0, and 0, are independently distributed on [g,O] with continuously differentiable density function('). Assume that $(. )/1/>(') is increasing in O. The firms have common marginal cost c < 1 and produce a homogeneous product for which the market inverse demand function is p(x) = I - x (this is publicly known). If both firms are awarded production rights, they interact as Cournot competitors (see Section 12.C). Characterize the planner's optimal auction of production rights.
°
(i) if kO(O" _I) = kO(O;, 0_,), then I.{O" 0_,) = 1,(0;, 0_,); (ii) if kO(O" 0_,) = I and kO(O;, 0_,) = 0, then 1,(0,,0_,) - 1,(0;,0_,) = 2.J'" OJ. Which, if any, of these two properties must be satisfied by any Nash implementable social choice function that involves an ex post efficient project choice?
23.F.8 A Show that any ex post classically efficient social choice function in Example 23.F.3 has YL = y" = I. 23.F_9 8 Show that in the model of Example 23.F.3: (a) No feasible social choice function is ex post efficient. (b) In any feasible social choice function, y" S YL and I" S I L • (c) In any feasible social choice function, the expected gains from trade of a low-quality seller are at least as large as the expected gains from trade of a high-quality seller; that is, It - 20j't ~ I" - 40y". 23.F.10· Characterize the sets of interim and ex ante incentive efficient social choice functions in the model of Example 23.F.3 when trade is not voluntary for the seller (but it is voluntary for the buyer). 23.AA.1 • Reconsider Exercise 23.C.8. Exhibit a mechanism r = (S., . .. , S" g(')) that is not a direct revelation mechanism that truthfully implements f(') in dominant strategies and for which each agent has a unique (weakly) dominant strategy. 23.AA.2 8 Let K = {ko, k., ... , kN I be the set of possible projects and suppose that, for each agent i, {v,L 0,): Ole El,l = 'V, that is, that every possible valuation function from K to R arises for some 0, eEl,. Do players in a Groves mechanism have a unique (weakly) dominant strategy? Consider instead a mechanism in which each agent i is allowed to announce a
.....
Mathematical Appendix
SECTION
M.A:
MATRIX
NOTATION
FOR
particular, that if M = 1 (so that [(x) E R) then DI(x) is a 1 x N matrix; in fact VI(x) = [D [(x)]'. To avoid ambiguity, in some cases we write DJ(x) to indicate explicitly the variables with respect to which the function [(.) is being differentiated. For example, with this notation, if I: RNH .... RM is a function whose arguments are the vectors x E RN and y E RX, the matrix D.[(x, y) is the M x N matrix whose mn th entry is iJIm(x, y)/iJx •. Finally, for a real-valued differentiable function I: RN .... R, the Hessian matrix D2[(X) is the derivative matrix of the vector-valued gradient function V[(x); i.e., D2 [(x) = D[V[(x)). In the remainder of this section, we consider differentiable functions and examine how two well-known rules of calculus-the chain rule and the product rule---{;ome out in matrix notation. The Chain Rule
Suppose that g: RS .... RN and I: RN .... RAt are differentiable functions. The composite function [(g(')) is also differentiable. Consider any point x E RS. The chain rule allows us to evaluate the M x S derivative matrix of the composite function with respect to x, DJ(g(x» by matrix multiplication of the N x S derivative matrix of g('), Dg(x), and the M x N derivative matrix of I(') evaluated at g(x), that is, D I(y), where y = g(x). Specifically,
This appendix contains a quick and unsystematic review of some of the mathematical concepts and techniques used in the text. The formal results are quoted as "Theorems" and they are fairly rigorously stated. It seems useful in a technical appendix such as this to provide motivational remarks, examples, and general ideas for some proofs. This we often do under the label of the "Proof" of the mathematical theorem under discussion. Nonetheless, no rigor of any sort is intended here. Perhaps the heading "Discussion of Theorem" would be more accurate. It goes without saying that this appendix is no substitute for a more extensive and systematic, book-length, treatment. Good references for some or most of the material covered in this appendix, as well as for further background reading, are Simon and Blume (1993), Sydsaeter and Hammond (1994), Novshek (1993), Dixit (1990), Chang (1984), and Intriligator (1971).
D.[(g(x)) = D[(g(x» Dg(x).
(M.A.I)
The Product Rule
Here we simply provide a few illustrations. (i) Suppose that [: RN .... R has the form [(x) = g(x)h(x), where both g(.) and h(') are real-valued functions of the N variables x = (x" . .. ,XN) (so that g: RN .... R and h: RN .... R). Then the product rule of calculus tells us that D[(x) = g(x) Dh(x)
+ h(x) Dg(x).
(M.A.2)
which, transposing, can also be written as
M.A Matrix Notation for Derivatives
VI(x) = g(x) Vh(x)
+ h(x) Vg(x).
(ii) Suppose that J: RN .... R has the form I(x) = g(x)·h(x) where both g(.) and h(') are vector·valued functions which map the N variables x = (Xl' ... ' x N ) into RM. Then D[(x) = g(x)'Dh(x) + h(x)·Dg(x). (M.A.3)
We begin by reviewing some matters of notation. The first and most important is that formally and mathematically a "vector" in RN is a column. This applies to any vector; it does not matter, for example, if the vector represents quantities or prices. It applies also to the gradient vector VI(x) E RN of a function at a point x; this is the vector whose nth entry is the partial derivative with respect to the nth variable of the real-valued function I: RN .... R, evaluated at the point x ERN. Expositionally, however, because rows take less space to display, we typically describe vectors horizontally in the text, as in x = (Xl' ... ' XN)' But the rule has no exception: all vectors are mathematically columns. The inner product of two N vectors X E RN and y E RN is written as X' Y = L. x.y•. If we view these vectors as N x 1 matrices, we see that X' Y = X T y, where T is the matrix transposition operator. An expression such as "x·" can always be read as "XT"; for example, the expression X' A, where A is an N x M matrix, is the same as
Note that h(x)' Dg(x) = [h(X)]T Dg(x) is a I x N matrix, as is the other term in the right-hand side. Thus, the vector-valued case (M.AJ) implies the scalar-valued formula (M.A.2). (iii) Suppose that I: R .... RM has the form [(x) = a(x)g(x), where a(') is a real-valued function of one variable (i.e., a: R .... R) and g: R .... RM. Then D [(x) = a(x) Dg(x) + a'(x)g(x). (M.A.4) (iv) Suppose that [: RN .... RM has the form [(x) = h(x)g(x) where h: RN .... R and g: RN .... RM. Then DI(x) = h(x) Dg(x)
+ g(x) Dh(x).
(M.A.5)
Note that g(x) is an M-element vector (i.e., an M x 1 matrix) and Dh(x) is a 1 x N matrix. Hence, g(x) Dh(x) is an M x N matrix (of rank 1). Observe also that (M.A.4) follows as a special case of (M.A.5).
xTA.
If I: !\IN .... RM is a vector-valued differentiable function, then at any x ERN we denote by D[(x) the M x N matrix whose mnth entry is iJIm(x)/iJx•. Note, in 926
.....
DERIVATIVES
927
928
MATHEMATICAL
APPENDIX
M.B Homogeneous Functions and Euler's Formula
-
SECTION
M.B:
HOMOGENEOUS
FUNCTIONS
AND
EULER'S
FORMULA
929
---------------------------------------------------------------------------------X,
In this section. we consider functions of N variables, f(x l ••••• x N). defined for all nonnegative values (Xl ••••• x N ):2: O. Definition M.B.1: A function f(x" ... ,xN) is homogeneous of degree r (for r = ... , -1.0,1, ... ) if for every I> 0 we have f(IX" ... ,IXN ) = (,f(x, • ...• x N ).
As an example. f(x i • x,) = XI/X, is homogeneous of degree zero and f(x ,• x,) = (XIX,)I" is homogeneous of degree one. Note that if f(x" ...• x N) is homogeneous of degree zero and we restrict the domain to have XI > 0 then, by taking I = I/x i • we can write the function f(·) as
j(-)
= t' Flgur. M.B.1
f(-) = I X,
f(l. x,/x l ••••• XN/X I) = f(x l ••••• x N). Similarly, if the function is homogeneous of degree one then and the slope of the level set containing point IX for
f(l. x,/x I•...• XN/X I ) = (l/xI)f(x , •...• x N ).
I
An illustration of this fact is provided in Figure M.B.1. Suppose that f(·) is homogeneous of some degree r and that h(·) is an increasing function of one variable. Then the function hU(x" ... , XN» is called homothetie. Note that the family of level sets of hU('» coincides with the family of level sets of f(·). Therefore. for any homothetic function it is also true that the slopes of the level sets are unchanged along rays through the origin. A key property of homogeneous functions is given in Theorem M.B.2.
> O. By the definition of homogeneity (Definition M.S.\) we have f(IX , •...• IX N ) -I'f(x , •... , XN) = O.
Differentiating this expression with respect to x. gives t
> 0 at IX is
_ Of..~X]!~1 = _ I' - , 0 f(x)/ox I 0f(x)/ox, iJf(IX)/OX 2 1'-' of(x)/ox, = - iJf(x)/ox,'
Theorem M.B.1: If f(x" . .. , xN) is homogeneous of degree r (for r = ...• -1.0,1 •... ), then for any n = 1, ... ,N the partial derivative function of (x, , ...• xN)/ox n is homogeneous of degree r - 1. Proof: Fix a
I
Of(IX" .... tXN) ,of(xI.· ... XN) 0 -t = • ox, ox.
Theorem M.B.2: (Euler's Formula) Suppose that f(x" ... ,XN) is homogeneous of degreer(for somer = ... , -1, 0, 1, ... ) and differentiable. Then at any (x" . .. , xN ) we have
so that
of(IX " ... ,IX N ) = t,-I of(X " ...• XN). OX, OX, By Definition M.B.I, we conclude that of(X " ...• xN)/ox, is homogeneous of degree r - I. •
f!.L
of (x, , ... ,xN )
n=l
iJxn
-
-
f(-
xn = r
Xl.···
- )
I
XN •
or, in matrix notation, Vf(x)·x = rf(x).
For example. for the homogeneous of degree one function f(x ,• x,) = (XIX,)I/'. we have of(X I. x,)/ox i = f(X,/XI)I/', which is indeed homogeneous of degree zero in accordance with Theorem M.B.1. Note that if f(·) is a homogeneous function of any degree then f(x ...• XN) = " is, a radial f(x;, ... , x;") implies f(tx , •... , tXN) = f(tx;, ... , tx;") for any t > 0; that expansion of a level set of f(·) gives a new level set of f(·). I This has an interesting implication: the slopes of the level sets of f(·) are unchanged along any ray through the origin. For example, suppose that N = 2. Then, assuming that of(x)/ox, "" 0, the slope of the level set containing point x = (XI' x,) at X is -(of(x)/ox,)/(of(x)/ox,),
Proof: By definition we have
f(lx t .... ,IX N) - t'!(x, .... , XN)
= O.
Differentiating this expression with respect to t gives
Evaluating at
I
= I, we obtain Euler's formula . •
For a function that is homogeneous of degree zero, Euler's formula says that t. A level set offunction 1(,) is a set of the form {x E
R~: I(x)
= k} for some k. A radial
expansion of this set is the set of points obtained by mUltiplying each vector X in this level set by some positive scalar t > O.
....
The level sels of a homogeneous funclion.
930
MATHEMATICAL
-
APPENDIX
As an example. note that for the function f(x i • x,) = XI/X,. we have of(x I> x,)/ox i = l/x2 and of(x I' x,)/ox, = -(xl/(x,)'). and so ~ of(xl.···.XN) _ '~I aX. x.
I -
XI
= .i; XI
-
-
0
(x,)' x, = .
• E C TI 0 N
M.
c:
CON CAY E
AND
~ of(XI' •••• XN) - - f(L.
X,,-
AX,.
(M.C.2) for any collection of vectors XI E A, ...• x that 0: 1 + ... + O:K = I.
- )
X1"",XN'
For example. when f(x l• x,) = (XIX,)I/'. we have of(X I' X,)/OXI of(.x l • X2 )/OX 2 = HXI/X2)1/'. and so
= t(X,/XI)I/'
K
E
A and numbers 0:, ~ O•..• , 0:" ~ 0 such
and Let us consider again the one-variable case. We could view each number 0:, in condition (M.C.2l as the "probability" that x' occurs. Then condition (M.C.2) says that the value of the expectation is not smalier than the expected value. Indeed, a concave function f: R .... R is characterized by the condition that
~o.f'...2(-'xl,,-''_'....c'._X"-,,N) __ I (X,)I/' _ I (XI)I/' _ X - ax. • 2 -XI XI + -2 -X, X2
L.. .~,
= (X IX2)1/2 =
f(X,.
(M.C.3)
x,).
for any distribution function F: R .... [0, I). Condition (M.C.3) is known as J,nsen's inequality.
M.C Concave and Quasiconcave Functions In this section. we consider functions of N variables f(x, •...• x N) defined on a domain A that is a convex subset of RN (such as A = RN or A = R~ = {x E IR N: x ~ O}).' We denote X = (Xl •...• x N ).
The properties of convexily and slriel convexily for a function f(·) are defined analogously but with the inequality in (M.CI) reversed. In particular, for a strictly convex function f('), a straight line connecting any two points in its graph should lie entirely above its graph, as shown in Figure M.C.2. Note also that f(·) is concave if and only if - f(·) is convex. Theorem M.C'! provides a useful alternative characterization of concavity and strict concavity.
Definition M.C.l: The function f: A .... R. defined on the convex set A eRN, is con-
cave if f(rJ.X'
+ (1
- o:)x) ~ o:f(x')
+ (1
- o:)f(x)
(M.C.1)
for all x and x' E A and all 0: E [0, 1]. If the inequality is strict for all x' 0: E (0,1), then we say that the function is strictly concave.
~
x and all
Theorem M.C.l: The (continuously differentiable) function f: A .... R is concave if and only if
Figure M.CI(a) illustrates a strictly concave function of one variable. For this case, condition (M.C.I) says that the straight line connecting any two points in the graph of f(·) lies entirely below this graph. 3 In Figure M.C.I(b). we show a function
fIx
R
,f(·<,)
+ (l
-
-
+ z) S; fIx) + Vf(x)-z
(M.C.4)
for all x E A and z ERN (with x + z E A). The function f(·) is strictly concave if inequality (M.C.4) holds strictly for all x E A and all z ~ o.
Figure M.C.l (a) A strictly concave function. (b) A concave but not strictly concave function.
f('"" + (l
R f(·)
.)"')~
flu' + (I -
.)f(x)~
.)X)}
• f(x') + (I - .)f(x)
_---f()
FIgure M.C.2
R f(·)
fIx')
A strictly function .
fIx) fIx') .f(x') + (I - o:)f(x)
x
x' ax'
(a)
FUN C TID N •
that is concave but not strictly concave; note that in this case the straight line connecting points x and x' lies on the graph of the function. so that condition (M.C.I) holds with equality. We note that condition (M.C.I) is equivalent to the seemingly stronger property that
in accordance with Euler's formula. For a function that is homogeneous of degree one. Euler's formula says that
11"'1
QUA' leo N CAY E
+ (I -
x ax'
.)x
I
+ (I
x'
flax'
+ (1
- o)x)
- .)x
(b)
2. For basic facts about convex sets. see Section M.G. 1 The graph of the function f: A - R is the set «x, yl E A x R: y = fIx)).
931
---------------------------------------------------------------------------------------
x
ax'
+ (I
- .)x
conve~
932
MATHEMATICAL
APPENDIX
-----------------------------------------------------------------R I(x)
__ _V/(x)
+ VI(x)' z......
Flgur. M.C.3
Any tangent to the graph of a concave function lies above the graph of the funclion.
x
Proof: We argue only the necessity of condition (M.C.4) for concave functions. For all <X E (0, I), the condition f(<xx' + (I - <x)x) ~ <xf(x') + (1 - <x)f(x) for all x, x' E A, can be rewritten (think of z = x' - x) as fIx
+ z)
$
f(x)
+ fIx + <xz) - f(x) <X
for all x E A, Z E RN (with x + Z E A), and a E (0, I). Taking the limit as a ... 0, we conclude that condition (M.C.4) must hold for a (continuously differentiable) concave function f(·) . • Condition (M.C.4) is shown graphically in Figure M.C.3. It says that any tangent to the graph of a concave function f(·) must lie (weakly) above the graph
-
SECTION
M.C:
CONCAVE
AND
QUASICQNCAVE
Theorem M.C.2: The (twice continuously differentiable) function f: A ... R is concave if and only if 02f(x) is negative semidefinite for every x E A. If 02f(x) is negative definite for every x E A, then the function is strictly concave. Proof: We argue only necessity. Suppose that f(·) is concave. Consider a fixed x E A and a direction of displacement from x, Z E RN with Z ,;. O. Taking a Taylor expansion of the function (a) = fIx + az), where a: E R, around the point a = 0 gives a2 fIx + az) - f(x) - Vf(x)·(az) = -z,02f(x + pz)z 2 for some p E [0, a). By Theorem M.C.I, the left·hand side of the above expression is nonpositive. Therefore, z'02f(x + pz)z:o;; O. Since a, hence p, can be taken to be arbitrarily small, this gives the conclusion z· 0 2f(x)z $ o.• In the special case in which N = I [so f(·) is a function of a single variable), negative semidefiniteness of 0 2f(x) amounts to the condition that d 2f(x)/dx 2 :0;; 0, whereas with negative definiteness we have d 2f(x)/dx 2 < 0 [to see this note that then z'O'f(x)z = z'(d'f(x)/dx 2»). Theorem M.C.2 tells us that in this case f(-) is concave if and only if d' f(x)/dx 2 $ 0 for all x, and that if d' f(x)/dx' < 0 for all x, then f(·) is strictly concave. Note that Theorem M.C.2 does nOI assert that negative definiteness of 0' f(x) must hold whenever f(·) is strictly concave. Indeed, this is not true: For example, when N = 1 the function f(x) = -x· is strictly concave, but d' f(0)/dx 2 = O. For convex and strictly convex functions the analogous result to Theorem M.C.2 holds by merely replacing the word "negative" with "positive."
off0 The corresponding characterization of convex and strictly convex functions simply entails reversing the direction of the inequality in condition (M.CA); that is, a convex function is characterized by the condition that fIx + z) ~ f(x) + Vf(x)'z for all x E A and Z ERN (with x + Z E A). We next develop a third characterization of concave and strictly concave functions.
The remainder of this section is devoted to discussion of quasiconcave and slrictly quasiconcave functions. Deflnltlon M.C.3: The function f: A ... R, defined on the convex set A c: IRN , is quasiconcave if its upper contour sets (x E A: fIx) ~ t} are convex sets; that is, if fIx)
Deflnltlon M.C.2: The N x N matrix M is negative semidefinite if z'Mz:o;; 0
(M.C.S)
for all z E IRN If the inequality is strict for all z ,;. 0, then the matrix M is negative definite. Reversing the inequalities in condition (M.C.S), we get the concepts of positive semidefinite and pOSitive definite matrices. We refer to Section M.E for further details on these properties of matrices. Here we put on record their intimate connection with the properties of the Hessian matrices 0 2 f(·) of concave functions.·
4. For theorems M.C.2, M.C.3, and M.C.4, the set A is assumed to be open (see Section M.F) so as to avoid boundary problems.
implies that
f(<xx
+ (1 -
a)x') ~ t
(M.C.S)
for any t E R, x, x' E A, and" E [0,1]" If the concluding inequality in (M.C.S) is strict whenever x,;. x' and <X E (0, 1), then we say that f(·) is strictly quasiconcave.
Analogously, we say that the function f(·) is quasiconvex if its lower contour sets are convex; that is, if f(x) $ I and fIx') $ I implies that f(<xx + (I - <x)x') :0;; I for any IE R, x, x' E A, and <X E [0,1). For strict quasiconvexity, the final inequality must hold strictly whenever x ,;. x· and a E (0,1). Note that f(·) is quasiconcave if and only if the function - f(·) is quasiconvex. The level sets of a strictly quasiconcave function are illustrated in Figure M.CA(a);
5. For more on convex sets, see Section M.G.
FUNCTIONS
933
934
MATHEMATICAL
APPENDIX
SECTION
----------------------------------------------------------------------------------
M.D:
MATRICES:
NEGATIVE
(SEMI)DEFINITENESS
AND
OTHER
PROPERTIES
X,
X,
X,
Ii: [(i) ~ [(x)}' Flgu... M.C.4
Ii: [(i) = I} {i:[(i) = I} {i:f(i) = t'}
(al
(b)
(a) The level sets of a strictly q uasiconca ve function. (b) The level sets of a quasiconcave function that is not strictly quasiconcave.
...-----x· Flgu ... M.C.S
Condition (M.C.8).
in Figure M.C.4(b) we show a function that is quasiconcave, but not strictly q uasiconca ve. lt follows from Definition M.C.3 that f(·) is quasiconcave if and only if
flax
+ (I
- a)x') ~ Min (f(x), fIx')}
For a quasiconvex function, we reverse the direction of both inequalities in (M.CS). Theorem M.C.4: The (twice continuously differentiable) function f: A _ R is quasiconcave if and only if for every x E A, the Hessian matrix 02f(x) is negative semidefinite in the subspace (z ERN: Vf(x)·z = O). that is, if and only if
(M.C7)
for all x, x' E A and a E [0, I]. From this, or directly from (M.C6), we see that a concave function is automatically quasiconcave. The converse is not true: For example, any increasing function of one variable is quasiconcave. Thus, concavity is a stronger property that quasiconcavity. It is also stronger in a different sense: concavity is a "cardinal" property in that it will not generally be preserved under an increasing transformation of f(·). Quasiconcavity, in contrast, will be preserved. Theorems M.C.3 and M.C4 are the quasiconcave counterparts of Theorems M.CI and M.C2, respectively.
z'02f(x)z:5: 0
M.D Matrices: Negative (Semi)Definiteness and Other Properties In this section, we gather various useful facts about matrices.
Proof: Again, we argue only the necessity of (M.CS) for quasiconcave functions. If fIx') ~ fIx) and a E (0, I] then. using condition (M.C7), we have that ~
(M.C.9)
For a quasiconvex function, we replace the word "negative" with "positive" everywhere in the statement of Theorem M.C4.
for all x, x' E A. If Vf(x) '(x' - x) > 0 whenever fIx') ~ fIx) and x' -F x. then f(') is strictly quasiconcave. In the other direction. if f(') is strictly quasiconcave and if Vf(x) -F 0 for al\ x E A, then Vf(x)· (x' - x) > 0 whenever fIx') ~ fIx) and x' -F x.
+ x) - fIx)
Vf(x)'z = 0
Proof: Necessity (again, we limit ourselves to this) can be argued exactly as for Theorem M.C2. The only adjustment is that we restrict z to be such that V f(x)' Z = 0 and we resort to Theorem M.C3 instead of Theorem M.C\. •
Theorem M.C.3: The (continuously differentiable) function f: A - R is quasiconcave if and only if (M.C.B) Vf(x)· (x' - x) ~ 0 whenever fIx') ~ fIx)
f(a(x' - x)
whenever
for every x E A. 6 If the Hessian matrix 02f(x) is negative definite in the subspace {z E IR N : Vf(x)·z = O} for every x E A, then f(') is strictly quasiconcave.
Definition M.D.1: The N x N matrix M is negative semidefinite il
o.
z·Mz:5: 0
a
(M.D.1)
for al\ z E IRN If the inequality is strict for al\ z -F 0, then the matrix M is negative definite. Reversing the inequalities in condition (M.D.1), we get the concepts of positive semidefinite and positive definite matrices.
Taking the limit as ex - 0, we get Vf(x)'(x' - x) ~ O. The need for the condition" VfIx) -F 0 for all x E A in the last part of the theorem is illustrated by the function fIx) = x 3 for x E IR. This function is strictly quasiconcave (check this using the criterion in Definition M.C3), but because VflO) = 0 we have Vf(x)·(x' - x) = 0 whenever x = O. • fl
Note that a matrix M is positive semidefinite (respectively, positive definite) if and only if the matrix - M is negative semidefinite (respectively, negative definite). Recall that for an N x N matrix M the complex number;' is a characteristic value (or eigenvalue or root) if it solves the equation 1M - ).II = O. The characteristic values of symmetric matrices are always real.
Theorem M.C.3's characterization of quasiconcave functions is illustrated in Figure M.CS. The content of the theorem's condition (M.CS) is that for any quasiconcave function f(·) and any pair of points x and x' with fIx') ~ fIx), the gradient vector VfIx) and the vector (x' - x) must form an acute angle.
6. See Section M.E for a discussion of the properties of such matrices.
d
935
936
MATHEMATICAL
APPENDIX
Theorem M.D.1: Suppose that M is an N x N matrix. (i) The matrix M is negative definite if and only if the symmetric matrix M + MT is negative definite. (ii) If M is symmetric. then M is negative definite if and only if all of the characteristic values of M are negative. (iii) The matrix M is negative definite if and only if M- 1 is negative definite. (iv) If the matrix M is negative definite. then for all diagonal N x N matrices K with positive diagonal entries the matrix KM is stable.'
--- -
• E C T ION
Proof: Part (i) simply follows from the observation ihat z'(M + MT)Z = 2z'Mz for every z E RH. The logic of part (ii) is the following. Any symmetric matrix M can be diagonalized in a simple manner: There is an N x N matrix of full rank e having e T = e - I and such that CM C T is a diagonal matrix with the diagonal entries equal to the characteristic values of M. But then z· Mz = (Cz)· eMCT(Cz). and for every Z E RH there is a z such that i = Cz. Thus. the matrix M is negative definite if and only if the diagonal matrix CMC T is. But it is straightforward to verify that a diagonal matrix is negative definite if and only if everyone of its diagonal entries is negative. Part (iii): Suppose that M - I is negative definite and let z ¥- O. Then z· M z = (Z'MZ)T = z'MTz = (MTz)'M-I(MTz) < O. Part (iv): It is known that a matrix A is stable if and only if there is a symmetric positive definite matrix E such that EA is negative definite. Thus, in our case, we can take A = KM and E = K- . •
M. 0:
.. A T RIC E . :
NEG A T I V E
( • E .. II 0 E FIN I TEN E ••
AND
0 THE R
Proof: (i) The necessity part is simple. Note that by the definition of negative definiteness we have that every ,M, is negative definite. Thus, by Theorem M.D.!, the characteristic values of ,M, are negative. The determinant of a square matrix is equal to the product of its characteristic values. Hence, I,M,I has the sign of (-IY. The sufficiency part requires some computation. which we shall not carry out. It is very easy to verify for the case N = 2 [if the conclusion of (i) holds for a 2 x 2 symmetric matrix. then the determinant is positive and both diagonal entries are negative; the combination of these two facts is well known to imply the negativity of the two characteristic values]. For (ii), we simply note the requirement to consider all permutations. For example. if M is a matrix with all its entries equal to zero except the N N entry, which is positive. then M satisfies the nonnegative version of (i) but it is not negative semidefinite according to Definition M.D.1. Notice that in part (iii) we only claim necessity of the determinantal condition. [n fact, for nonsymmetric matrices the condition is not sufficient. • Example M.D.I: Consider a real-valued function of two variables, I(x" x,). In what follows. we let subscripts denote partial derivatives; for example, 112(X,. x,) = iJ' I(x I' x,)/iJx, iJx,. Theorem M.C.2 tells us that 1(') is strictly concave if
Xl)]
D'/(x" x,) = [/"(X,, x,) 112(X" I,I(X I • X,) I,,(x l , X,)
is negative definite for all (XI' x,), According to Theorem M.D.2, this is true if and only if
'
For positive definite matrices, we can simply reverse the words "positive" and "negative" wherever they appear in Theorem M.D.1. Our next result (Theorem M.D.2) provides a determinantal test for negative definiteness or negative semidefiniteness of a matrix M. Given any T x S matrix M, we denote by ,M the r x S submatrix of M where only the first t ~ Trows are retained. Analogously, we let M, be the T x s submatrix of M where the first s ~ S columns are retained, and we let ,M, be the t x s submatrix of M where only the first r oS T rows and s oS S columns are retained. Also. if M is an N x N matrix, then for any permutation It of the indices {I, ...• N} we denote by M' the matrix in which rows and columns are correspondingly permuted.
and
/ll(X I • x,) 112(X" X')I > 0, I"(x,, x,)
IIII (x " x,)
or equivalently, if and only if
and
111(X" x,)/,,(x l , x,) - U12(X" x,)]' > O. Theorem M.C.2 also tells us that 1(') is concave if and only if D' I(x" x,) is negative semidefinite for all (XI' x,). Theorem M.D.2 tells us that this is the case if and only if
Theorem M.D.2: Let M be an N x N matrix. (i) Suppose that M is symmetric. Then M is negative definite if and only if (-l),I,M,1 > 0 for every, = 1..... N. (ii) Suppose that M is symmetric. Then M is negative semidefinite If and only if ( -1 )'J,M;I 2: 0 for every' = 1•...• N and for every permutation It of the indices {1 ....• N}. (iii) Suppose that M is negative definite (not necessarily symmetric). Then ( -1 )'J,M;I > 0 for every, = 1....• N and for every permutation It of the indices {1 .... • N}.8
and
/II(X" x,) l , x,)
I111(x
ltiXI,Xl)I2:0. 1,,(xI' x,)
and, permuting the rows and columns of D'/(x l , x,).
1/,,(x l • x,)1
~
0
and
Xl)l . .
I,,(x I , x,) 111(x" O. '" 111(X" x,)
I112(X I, x,)
Thus. 1(') is concave if and only if 111(x I , x,)
7. A matrix M is stable irall or its characteristic values have negative real pans. This terminology is motivated by the ract that in this case the solution or the system or differential equations dx(r)/dl = MX(I) will converge to zero as I _ 00 ror any initial position x(O). 8. A matrix M such that - M satisfies the condition in (iii) is called a P matrix. The reason is that the detenninant of any submatrix obtained by deleting some rows (and corresponding columns)
:s 0,
I"(x,, x,) oS 0, and
•
is positive.
.....
PRO PER TIE'
937
938
MATHEMATICAL
-
SECTION
APPENDIX
-------------------------------------------------------------------------A similar test is available for positive definite and semidefinite matrices: The results for these matrices parallel conditions (i) to (iii) of Theorem M.D.2, but omit the factor (_1)'.9 Theorem M.D.3: Let M be an N x N symmetric matrix and let B be an N x S matrix with S ~ N and rank equal to S.
(-1)'!
,M:
(,B")T
= O}
tSEMI)DEFINITENESS
AND
OTHER
and (performing the appropriate permutations) f"(x,, x,)
f"(x,, x,)
f,(x" x,)
f12(X" x,)
fl1(X" x,)
f,(x" x,) ~ O.
f,(x" x,)
f,(x" x,)
0
-
To characterize matrices that are positive definite or positive semidefinite on the subspace {t e IJl:N: Bz = O}, we need only alter Theorem M.D.3 by replacing the term ( _ I)' with ( _I )s.
0
for r = S + 1, ... , N. (ii) M is negative semidefinite on {z eRN: Bz z ERN with Bz = 0 and z ,;. 0) if and only if
NEGATIVE
2f,(x" x,)f,(x" x,)fdx" x,) - [f,(x" x,)]' f"(x,, x,) - [f,(x" x,)]'fl1 (x" x,) ~ O.
,M, ,B! > 0
(,B)T
MATRICES:
Computing these two determinants gives us the necessary and sufficient condition
(i) M is negative definite on {z eRN: Bz = O} (i.e., z· Mz < 0 for any z eRN with Bz = 0 and z ,;. 0) if and only if
(-1)'!
M.D:
(I.e., z'Mz ~ 0 for any
Theorem M.D.4: Suppose that M is an N x N matrix and that for some p » 0 we have Mp = 0 and MTp = O. Denote Tp = {z eRN: p'z = O} and let if be the (N - 1) x (N - 1) matrix obtained from M by deleting one row and the corresponding column.
,B"I ~0 0
for r = S + 1, ... , N and and every permutation n, where ,8" is the matrix formed by permuting only the rows of the matrix ,8 according to the permutation n (,M; is, as before, a matrix formed by permuting both the rows and columns of ,M,).
(i) If rank M = N - 1, then rank Nt = N - 1. (ii) If z· Mz < 0 for all z e Tp with z,;. 0 (i.e., if M is negative definite on Tp), then z· Mz < 0 for any z eRN not proportional to p. (iii) The matrix M is negative definite on Tp if and only if Nt is negative definite.
Proof: We will not prove this result. Note that it is parallel to parts (i) and (ii) of Theorem M.D.2 with the bordered matrix here playing a role similar to the matrix there. _
Proof: (i) Suppose that rank M < N - I, that is, Mz = 0 for some i e RN -, with i ,;. O. Complete to a vector z e RN by letting the value of the missing coordinate be zero. Then we have that, first, z is linearly independent of p (recall that p» 0) and, second, Mz = 0 and Mp = O. Thus, rank M < N - I, which contradicts the
z
Example M.D.2: Suppose we have a function of two variables, f(x l , x,). We assume that Vf(x) ,;. 0 for every x. Theorem M.C.4 tells us that f(·) is strictly quasi concave if the Hessian matrix D' f(x l , x,) is negative definite in the subspace (z e R': Vf(x)' z = O} for every x = (x" x,). By Theorem M.D.3 the latter is true if and only if fl1(X" x,)
f12(X" x,)
fleX"~ x,)
f"(x,, x,)
f"(x,, x,)
f,(x" x,) > 0,
fleX"~ x,)
f,(x" x,)
hypothesis. (ii) Take a z e RN not proportional to p. For IX, = (P'z)/(P'p) and z* we have z* e Tp and z' ,;. O. Because MTp = Mp = 0, we have then
(iii) This is similar to part (ii). In fact, part (ii) directly implies that M is negative definite if M is negative definite on Tp (because for any ze RN -', z'Mi = z'Mz, where z has been completed from z by placing a zero in the missing coordinate, and if z ,;. 0 this z is by construction not proportional to pl. For the converse, let n denote the row and column dropped from M to obtain M. If for every z' e Tp with z'';' 0 we let z = z' - (z;/P.)p, then z. = 0 and z';' 0 [if z were equal to zero, then we would have z' = (z;/P.)p in contradiction to z'·p = 0]. Moreover, z"Mz' = z·M. = z'Mz < O. _
0
2f,(x" x,)f,(x" X,)f12(X" x,) - [flex"~ x,)]'fdx" x,) - [f,(x" x,)]' fl1(X" x,) > O.
= XIX, we get 2x,x, > 0 confirming that the function is strictly quasiconcave. By Theorem M.C.4, f(·) is quasiconcave if and only if the Hessian matrix D' f(x" x,) is negative semidefinite in the subspace (z e R': Vf(x)' z = O} for every x = (x" x,). By Theorem M.D.3 this is true if and only if
If we apply this test to f(x" x,)
fl1(X" x,)
fdx" x,)
f,(x" x,)
f"(x,, x,)
f,(x" x,) ~ 0,
f,(x l , x,)
f,(xI' x,)
IX,p,
z'Mz = (z· + IX,p)'M(z' + IX,p) = z"Mz' < O.
or equivalently, if and only if
f"(x,, x,)
=z-
Definition M.D.2: The N x N matrix M with generic entry aji has a dominant diagonal if there is (p" ... ,PN) »0 such that, for every i = 1, ... , N, !pjajA > Li*; !Pia;J Dellnltlon M.D.3: The N x N matrix M has the gross substitute sign pattern if every nondiagonal entry is positive. Theorem M.D.S: Suppose that M is an N x N matrix.
0
(i) If M has a dominant diagonal, then It Is nonsingular. (ii) Suppose that M is symmetric. If M has a negative and dominant diagonal then it is negative definite.
9. Recall that M is positive (semi)definite if and only if - M is negative (semi)definite. Moreover,
hM,1 = (-I)'I,M,I·
...
PROPERTIES
939
940
MATHEMATICAL
APPENDIX
(iii) If M has the gross substitute sign pattern and if for some p » 0 we have Mp « 0 and MTp « 0, then M is negative definite. (iv) If M has the gross substitute sign pattern and we have Mp = MTp = 0 for some p » 0, then Nt is negative definite, where Nt Is any (N - 1) x (N - 1) matrix obtained from M by deleting a row and the corresponding column. (v) Suppose that all the entries of M are nonnegative and that Mz« z for some z » 0 (i.e., M is a productive input-output matrix). Then the matrix (1- M)-' exists. In fact, (/- M)-' = L~:cf M*.
-
SECTION
IMPLICIT
FUNCTION
THEOREM
941
x
R
~(ij)
-------7!T--- ~(.)
x' --------=x x' ------- I I I I I I
I
:
: I I I I I
: I I I I I
Figure M.E.1
A locally solvable equation. (a) Solutions or f(x; q) = 0 near (;C, q). (b) The graph or ~(- ).
q' ij q'
f(·;q')
j{-; ij) f(·;q')
(a)
(b)
Suppose that x = (x, •.... xN) E A and ii = (ii, •.. ·• iiM) E B satisfy equations (M.E.I). That is. I.(x. ii) = 0 for every n. We are then interested in the possibility of solving for x = (x, •...• x N ) as a function of q = (q, •...• qM) locally around ii and .ii. Formally. we say that a set A' is an open neighborhood of a point x E IRN if A' = {x' E IR N: IIx' - xII < £} for some scalar £ > O. An open neighborhood B' of a point q E IRM is defined in the same way. Definition M.E.1: Suppose that x = (x" ... ,xN) E A and q = (q" ... ,qM) E B sa~sf! the equations (M.E.1). We say that we can locally solve equations (M.E.1) at (x, q) for x = (x, . ... ,XN) as a function of q = (q" ...• qM) if there a.re open neighborhoods A' c A and B' c B, of x and q, respectively, and N untquely determined "implicit" functions '1,( .), ... , '1N(') from B' to A' such that 'n('1,(q). ... ,'1N(q); q) = 0
for every q
E
B' and every n,
and for every n. In Figure M.E.I we represent. for the case where N = M = I, a situation in which the system of equations can be locally solved around a given solution. The implicit function theorem gives a sufficient condition for the existence of such implicit functions and tells us the first-order comparative statics effects of q on x at a solution.
M.E The Implicit Function Theorem
Theorem M.E.1: (Implicit Function Theorem) Suppose that every equation I n (·) is continuously differentiable with respect to its N + M variables and that we consider a solution x = (x" ... , XN) at parameter values q = (q" ...• qM)' that is, satisfying 'nIx; q) = 0 for every n. If the Jacobian matrix of the system (M.E.1) ~It~ respect to the endogenous variables, evaluated at (x, q). is nonsingular, that IS, If
The setting for the implicit function theorem (1FT) is as follows. We have a system of N equations depending on N endogenous variables x = (x, •...• x N) and M parameters q = (q, •...• qM): XN;
THE
I.~---------~~---------------
Proof: (i) Assume, for simplicity. that p = (I •...• I). Suppose. by way of contradiction. that Mz = 0 for z # O. Choose a coordinate n such that Iz,l ~ Iz,.1 for every other coordinate n'. Then la•• z.1 > L, •• la.,z.1 ~ L,.,la.jzjl. where 0/) is the generic entry of M. Hence. we cannot have Lj a,)z) = O. and so M z # O. Contradiction. (ii) If M has a negative dominant diagonal then so does the matrix M - aI, for any value a ~ O. Hence. by (i) we have (_I)NIM - all # O. Now if a is very large it is clear that (-ltIM - all> 0 (since (_I)NIM - all = (-I)NaNI(Mja) - II and 1-/1=(-1)''). Moreover. since (-I)NIM-aII is continuous in a and (_I)NIM - all # 0 for all a ~ O. this tells us that (-I)NIM - all> 0 for all a ~ O. Hence. (-I )NIMI > O. By the same argument. (-IYI,M,I > 0 for all r. So. if M is also symmetric then by part (i) of Theorem M.D.2 it is negative definite. (iii) The stated conditions imply that M + MT has a negative and dominant diagonal [in particular. note that Mp« 0 and AfT p «0 implies that p.(2a.. ) < - L) •• Pj(aj • + a.) for all n. where 0/) is the generic entry of M]. Because, by the gross substitute property. 0/) > 0 for i # j.this gives us Ip.(2a.. )1 > IL) .. p)(a", + 0,,)1 for all n. Hence. the conclusion follows from part (ii) of this theorem and part (i) of Theorem M.D.!. (iv) If M satisfies the condition of (iv). then the fact that M has the gross substitute sign pattern implies that M does as well and that Mp «0 and MT p «0. Hence. M satisfies the conditions of (iii) and is therefore negative definite. (v) This result was already proved in the Appendix to Chapter 5 (see the proof of Proposition 5.AA.I).
f,(x, •...•
M.E:
q, •...• qM) = 0 (M.E.I)
Of,(x, q)
Of,(x, q)
ox,
oXN
fN(x, •...• xN;q, •...• qM) = 0
# 0,
The domain of the endogenous variables is A c RN and the domain of the parameters is Be RM.'O
OfN(x, q)
OfN(x, q)
ox,
oXN
(M.E.2)
then the system can be locally solved at (x, q) by implicitly defined lunctions 'In: B' _ A' that are continuously differentiable. Moreover, the first-order effects
10. In wha, follows. we ,ake A and B to be open sets (sec Section M.F) so as to avoid boundary problems.
..
942
MATHEMATICAL
- -
APPENDIX
of q on x at (x, 17) are given by Dq'l(q) = -[Dxf(x; qlr'Dqf(x; 17).
(M.E.3)
Proof: A proof of the existence of the implicit functions 'I.: B' -+ A' is too technical for this appendix, but its common-sense logic is easy to grasp. Expression (M.E.2), a full rank condition, tells us that we can move the values of the system of equations in any direction by appropriate changes of the endogenous variables. Therefore, if there is a shock to the parameters and the values of the equation system are pushed away from zero, then we can adjust the endogenous variables so as to restore the "equilibrium." Now, given a system of implicit functions '1(q) = ('1,(q), ... ,'1N(q)) defined on some neighborhood of (x, Ii), the first-order comparative static effects iJ'I.(ii)/iJqm are readily determined. Let f(x; q) = (/, (x; q), ... ,fN(X; q)). Since we have
f('1(q); q) = 0
CONTINUOUS
FUNCTIONS
AND
COMPACT
Definition M.E.2: Given open sets A c ~ and 8 c RM , the (continuously differentiable) system 01 equations f(·; ) = 0 defined on A is regular at e 8 If (M.E.2) holds at any solution x; that is. if fIx; ) = 0 implies that 10/(x;
Theorem M.E.2: (Transversality Theorem) Suppose that we are given open sets A c RN and 8 c RM and a (continuously differentiable) function f: A x 8 _ RN " f(' ; .) satisfies the condition The N x (N + M) matrix Of(x; q) has rank N whenever fIx; q) = 0, then the system of N equations in N unknowns 1('; ) = 0 is regular for almost every
for all q E B',
M.F Continuous Functions and Compact Sets I n this section, we begin by formally defining the concept of a continuous function. We then develop the notion of a compact set (and, along the way, the notions of open and closed sets). Finally. we discuss some properties of continuous functions that relate to compact sets. A sequence in !RN assigns to every positive integer m = 1,2, ... a vector xm E !RN. We denote the sequence by {xm}:::;" or, simply, by {xm} or even xm.
DJ(x; ii)D.'1
•
Note that when N = M = I (the case of one endogenous variable and one parameter), (M.E.3) reduces to the simple expression
Definition M.F.1: The sequence {xm} converges to x ERN, written as Iim m _ 00 x'" = x, or xm ~ x, if for every £ > 0 there is an integer M, such that IIx m - xII < £ whenever m > M,. The point x is then said to be the limit point (or simply the limit) of sequence {xm}.
d'l(ii) = _ iJ f(x-; ii)/iJq dq iJf(x; ii)/iJx The special case of the implicit function theorem where M = N and every equation has the form f.(x, q) = g.(x) - q. = 0 is known as the inverse function theorem.
In words: The sequence {xm} converges to x if xm approaches x arbitrarily closely as m increases.
How restrictive is condition (M.E.2)? Not very. In Figure M.E.2 we depict a situation where it fails to hold. [By contrast, in Figure M.E.I condition (M.E.2) is satisfied.] However, the tangency displayed in Figure M.E.2 appears pathological: it would be removed by any small perturbation of the function J(.; '). An important result, the transversality theorem, makes this idea precise by asserting that, under a weak condition [enough first-order variability of J('; .) with respect to x and q], (M.E.2) holds lIellericalty on the parameters. We present a preliminary concept in Definition M.E.2.
R
".F:
With this definition we then have Theorem M.E.2.
we Can apply the chain rule of calculus to obtain
D.'1(ii) = - [DJ(x; ii)) - I D.f(.x; ij).
SECTION
Definition M.F.2: Consider a domain X eRN. A function f: X .... R Is continuous if for all x E X and every sequence xm -+ x (having xm E X for all m), we have f(xm) -+ f(x). A function f: X -+ IRK is continuous if every coordinate function f*(·) is continuous. In words: a function is continuous if, when we take a sequence of points x', x 2 , • •• converging to x, the corresponding sequence of function values f(x'), f(x 2 ), ••• converges to f(x). Intuitively, a function fails to be continuous ifit displays a "jump" in its value at some point x. Examples of continuous and discontinuous functions defined on [0, I] are illustrated in Figure M.F.1. We next develop the notions of open, closed, and compact sets.
~~M~
Condition (M.E.2) is violated at the solution (X, ii)·
Definition M.F.3: Fix a set X c !RN. We say that a set A c X Is open (relative to X) if for every x E A there is an £ > 0 such that Ilx' - xII < £ and x' E X implies x' EA. It. "Almost every" means that it ror example. we choose 4 according to some nondegenerate multinomial normal distribution in A", then with probability I the equation system f(·; 4) = 0 is
1(:<1)
regular. This is the concept
d
or "genericity" in this context.
SETS
943
944
MATHEMATICAL
APPENDIX
SEC T ION
1M. F:
CON TIN U 0 U S
FUN C T ION SAN 0
COM PAC T
SET
S
945
---------------------------------------------------------------------------- -------------------------------------------------------------------------Property (iii) of Theorem M.F.I is noteworthy because it gives us a direct way to characterize a closed set: a set A is closed if and only if the limit point of any sequence whose members are all elements of A is itself an element of A. Points (in X) that are the limits of sequences whose members are all elements of the set A are known as the limit poims of A. Thus, property (iii) says that a set A is closed if and only if it contains all of its limit points. Given A eX, the interior of A (relative to X) is the open set 13
R
f(·)
Figure M.F.1
Continuous and discontinuous functions. (a) A continuous function. (b) A discontinuous
o (a)
(b)
lntx A = {x
E
A: there is
f,
> 0 such that IIx' - xII <
t
and x' E X implies x' E A}.
The closure of A (relative to X) is the closed set Cl x A = X\Intx (X\A). Equivalently, Cl x A is the union of the set A and its limit points; it is the smallest closed set containing A. The boundary of A (relative to X) is the closed set Bdryx A = Cl x A\Int x A. The set A is closed if and only if Bdryx A c A.
function.
x,
Definition M.F.4: A set A c IRN is bounded if there is , E IR such that IIxll < , for every x E A. The set A c IRN is compact if it is bounded and closed relative to IRN We conclude by noting two properties of continuous functions relating to compact sets. Formally, given a function f: X .... IRK, the image of a set A c X under f(·) is the set I(A) = (YE IRK: y = fIx) for some XE A}.
Figure M.F.2
x, (a)
(b)
Open and closed sets.
Theorem M.F.2: Suppose that f: X .... IRK is a continuous function defined on a nonempty set X c IRN (i) The image of a compact set under f(·) is compact: That is, if A c X is compact, then f(A) = {y E IRK: y = fIx) for some x E A} is a compact subset of IRK. (ii) Suppose that K = 1 and X is compact. Then f(') has a maximizer: That is, there is x E X such that fIx) ~ fIx') for every x· EX.
A set A c X is closed (relative to X) if its complement X\A is open (relative to X).12 If X = IRN we simply refer to "open" and "closed" sets. Given a point x ERN, a set B = {x' ERN: IIx' - xII < t} for some scalar t > 0 is called an open ball around x. With this notion, the idea of an open set can be put as follows: Suppose that the universe of possible vectors in IRN is X. A set A c X is open (relative to X) if, for every point x in A, there is an open ball around x all of whose elements (in X) are elements of A. In Figure M.F.2(a) the hatched set A is open (relative to X). In the figure, we depict a typical point x E A and a shaded open ball around x that lies within A; points on the dashed curve do not belong to A. In contrast, the hatched set A in Figure M.F.2(b) is closed because the set X\A is open; note how there is an open ball around the point x E X\A that lies entirely within X\A [in the figure, the points on the inner solid curve belong to A]. Theorem M.F.I gathers some basic facts about open and closed sets. Theorem M.F.1: Fix a set X c ~. In what follows, all the open and closed sets are relative to X. (i) The union of any number, finite or infinite, of open sets is open. The intersection of a finite number of open sets is open. (ii) The intersection of any number, finite or infinite, of closed sets is closed. The union of a finite number of closed sets is closed. (iii) A set A c X is closed if and only if for every sequence xm .... X E X, with xmEA for all m, we have xEA.
Part (ii) of Theorem M.F.2 asserts that any continuous function f: X .... R defined on a compact set X attains a maximum. We illustrate this result in Figure M.FJ. A maximum is not attained either in Figure M.FJ(a) or in Figure M.F.3(b). In Figure M.F.3(a), the function is continuous, but the domain is not compact. In Figure M.F.3(b), the domain is compact, but the function is not continuous. Given a sequence {x*}, suppose that we have a strictly increasing function m(k) that assigns to each positive integer k a positive integer m(k). Then the sequence x MCI ', x M " ' , • • • (written M (x "')) is called a subsequence of {x M}. That is, {xM(l'} is composed of an (order-preserving) subset of the sequence {x M}. For example, if the sequence {XM} is 1,2,4.16,25,36, ... , then one subsequence of {x is 1.4.16.36.... ; another is 2,4,16,25,36, .... M
}
Theorem M.F.3: Suppose that the set A c liN is compact. (i) Every sequence {x"') with x'" E A for all m has a convergent subsequence. Specifically, there is a subsequence {x"'lkl} of the sequence {x"'} that has a limit in A, that is, a point x E A such that x"'lk, ..... x. (ii) If, in addition to being compact, A is also discrete, that is, if all its points are isolated [formally, for every x E A there is • > 0 such that x' = x whenever x' E A and IIx' - xII < r.], then A is finite.
t 2. Given two sets A and B. the set A\B is the set containing all the elements of A that are not elements of B.
J 3. In what follows in this paragraph, all
..
of the open and closed sets are once again relative to X .
946
MATHEMATICAL
APPENDIX
SEC T ION
-------------------------------------------------------------------------------------------R fIx) =
+ (I
u
R
~
M. Q:
CON V E X
8 E T'
AND
S EPA A A TIN G
- .)x·
Figure M.G.1
Indispensability of the continuity and compactness assumptions for the existence of a
Convex and nonconvex sets. (a) A convex set. (b) A nonconvex set.
(b)
(al
maximizer. (a) A continuous function with no
Convex Hull
maximizer On a
o!------~
noncom pact domain. (b) A discontinuous
x = [0,1]
x = (0.1]
function with no maximizer On a compact domain.
(b)
(0)
Flgur. M.G.2
M.G Convex Sets and Separating Hyperplanes
A nonconvex set and its convex hull.
In this section. we review some basic properties of convex sets, including the important separating hyperplane theorems.
0 1
Definition M.G.l: The set A c IRN is convex if and ex E [0, 1].14
<XX
+ (1
- ex)x'
EA
whenever x, x·
E
A
In words: A set in IRN is convex if whenever it contains two vectors x and x', it also contains the entire segment connecting them. In Figure M.G.I(a), we depict a convex set. The set in Figure M.G.I(b) is not convex. Note that for a concave function f: A .... IR the set {(z, v) E R N + I : v;5; f(z), z E A} is convex. Note also that the intersection of any number of convex sets is convex, but the union of convex sets need not be convex.
a
Xl' ... ' X J
with
x)
E
r
Extreme Points of K
= (0, b, c, d} Flgur. M.G.3
h
A very important result of convexity theory is contained in Theorem M.G.\. Theorem M.G.1: Suppose that BeRN is a convex set that is also compact (that is, closed and bounded; see Section M.F). Then every x E B can be expressed as a convex combination of at most N + 1 extreme points of B.
Figure M.G.2 represents a set and its convex hull. It is not difficult to verify that the convex hull can also be described as the set of all possible convex combinations of elements of B, that is, ex)x): for some
K
The extreme points of the convex set represented in Figure M.G.3 are the four corners.
Definition M.G.2: Given a set B c IRN , the convex hull of B, denoted Co B, is the smallest convex set containing B, that is, the intersection of all convex sets that contain B.
{I
Proof: The proof is too technical to be given here. Note simply that the result is correct for the convex set in Figure M.G.3: Any point belongs to the triangle spanned by some collection of three corners. _
B for all j,
j=l
and some (ex" ... , exJ) <': 0 with
I
We now turn to the development of the separating hyperplane theorems.
ex) = I}.
j'l
Definition M.G.4: Given p ERN with P -F O. and c E R, the hyperplane generated by p and c is the set Hp.c = {ZE RN : p·z = c}. The sets {ZE~: p'z <': c} and {z ERN: p'z;5; c} are called, respectively, the half-space above and the halfspace below the hyperplane Hp.c'
Definition M.G.3: The vector x E B is an extreme point of the convex set B c IRN if it cannot be expressed as x = exy + (1 - ex)z for any y, Z E B and ex E (0, 1).
IXX + (1 - o:)x' is an element of the interior of A whenever (0, I) (see Seclion M.F for a definition of the interior of a set).
14. The set A is sericlly convex if x. x' E A and
0 E
947
Figure M.F.3
X
Co B =
H Y PEA P LAN E S
Hyperplanes and half-spaces are convex sets. Figure M.G.4 provides illustrations .
..
Extreme points of a convex set.
948
MATHEMATICAL
-
APPENDIX
-~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~_
Half·space above 1/
SECTION
M.H:
CORRESPONDENCES
P.'
Flgur. M.G.6 The supporling hyperplane theorem
Jj,lf·sp,ec below II p .•
-0 Figure M.G.4
Hyperplanes and half-spaces.
.1': Point in B Closest
10
Theorem M.G.3: (Supporting Hyperplane Theorem) Suppose that BeRN is convex and that x is not an element of the interior of set B (Le., x f Int B; see Section M.F for the concept of the interior of a set). Then there is p E RN with P ¢ 0 such that p' x ~ p' y for every y E B. Proof: Suppose that x f Int B. The following argument can be followed in Figure M.G.6. It is intuitive that we can find a sequence x~ .... x such that, for all m, xm is not an clement of the closure of set B (i.e., x~ f CI B; see Section M.F for a discussion of seq uences and the closure of a set). By the separating hyperplane theorem (Theorem M.G.2l, for each m there is a pm ¢ 0 and a em E R such that
x
(M.G.I)
B
H,,(
FIgure M.G.S The separating hyperplane Iheorem.
for every Y E B. Without loss of generality we can suppose that II p~ I = I for every m. Thus, extracting a subsequence if necessary (see the small type discussion at the end of Section M.F), we can assume that there is p ¢ 0 and e E R such that p~ .... p and em .... c. Hence, taking limits in (M.G.I), we have P'x~c~p'y
for every ye B. • Theorem M.G.2: (Separating Hyperplane Theorem) Suppose that BeRN is con and closed (se S r M vex . e ec Ion .F for a discussion of closed sets) and that x J 8 Th there IS p E RN 'th ' .. . en WI p ¢ 0, and a value CE R such that p'X > C and P'Y < c for every y E B. More generally, suppose that the convex sets A, BeRN are disjoint (" A n 8 = 0). Then there is peW' with p ~ 0, and a value c e R, such that p' x ~e~ foreverYXEAandp'y:S;cf B' or every ye . That IS, there is a hyperplane that separates A and B, leaving A and B on different sides of it. Proof: We discuss only the fi (. . rst part I.e., the separation of a point and a closed ' convex set) [n Figu M G 5 WI' . . r e . . we represent a closed, convex set B and a point x f B e a so JOdlcate by ye B the point in set B closest to x 15 [f we let _ . and ' - . . p - x - y c - p y, we can then see, first, that p'X > c' [since P'x _ c' = p'x _ p' _ (x - y)·(x - y) = Ilx - yll2 > 0] and, second, that for any Z E B the vectors p ;n~ y, cannot make an. acute angle, that is, p'(z - y) = p'Z - c':s; O. Finally, let - c + c where c > 0 IS small enough for p'X > c' + c = c to hold . •
::=
15. We use Ihe familiar Euclidea d' I . point in B that we . B be n IS ance measure. It IS to guarantee the existence of a closest require to closed.
M9
------------------------------------
Finally, for the important concept of the support function of a set and its properties we refer to Section 3.F of the text.
M.H Correspondences I t is common in economics to resort to a generalized concept of a function called a correspondence. Definition M.H.1: Given a set A eRN. a correspondence f: A .... IRK is a rule that assigns a set f(x) c RK to every x E A. Note that when, for every x E A, f(x} is composed of precisely one element, then f(·) can be viewed as a function in the usual sense. Note also that the definition allows for f(x) = 0, but typically we consider only correspondences with f(x} ¢ 0 for every x e A. Finally, if for some set Y c R" we have f(x) c Y for every x e A, we indicate this by f: A .... Y. We now proceed to discuss continuity notions for correspondences. Given A c IR" and Y c R", the graph of the correspondence f: A .... Y is the set {(x,ylEA x Y:yef(x)}. Definition M.H.2: Given A c IRN and the closed set Y c ilK, the correspondence f: A .... Y has a closed graph if for any two sequences xm .... X E A and ym .... y, with xm E A and ym e f(x m) for every m, we have y E fIx).
950
MATHEMATICAL
SECTION
APPENDIX
II
II
f(il = P, 2}
f(·)
f(')
\
f--L-.~<
Flgur. M.H.1
Closed graphs and upper hemicontinuous correspondences. (a) A closed graph correspondence that is not upper
(b) An upper
o
951
Upper hemicontinuity is only one of two possible generalizations of the continuity notion to correspondences. We now state the second (for the case where the range space Y is compact).
hernicontinuous.
(a)
CORRESPONDENCES
By Definition M.H.3, f(S) is bounded, and so CI f(S) (relative to R<) is a compact set. If, contradicting continuity of the function f('), the sequence (f(XM)} [which lies in the compact set CI f(S)] did not converge to f(x), then by Theorem M.F.3 we could extract a subsequence X MIk ' -+ x such that f(xm(l') -+ y for some), E CI f(S) having y ~ f(x). But then the graph of f(') could not be closed, in contradiction to the upper hemicontinuity of f(-) as a correspondence. _
A = [0,1]
Y = [0, 3] A = [0,1]
M.H:
hcmicontinuous
correspondence.
(b)
Definition M.H.4: Given A c IRN and a compact set Y c IRK, the correspondence I: A -+ Y is lower hemicontinuous (Ihc) if for every sequence xm -+ X E A with xm E A for all m, and every y E I(x), we can find a sequence ym -+ y and an integer M such that ym E I(x m) for m > M.
Note that the concept of a closed graph is simply our usual notion of c10sedness (relative to A x Y) applied to the set {(x, y) E A x Y: y E fIx)} (see Section M.F). Definition M.H.3: Given A c RN and the closed set Y c RK, the correspondence I: A -+ Y is upper hemicontinuous (uhc) if it has a closed graph and the images of compact sets are bounded, that is, for every compact set 8 c A the set I(B) = (y E Y: y E I(x) for some x E B} is bounded. 16 • 17
Figure M. H.2(a) represents a lower hemicontinuous correspondence. 19 Observe that the correspondence is nol upper hemicontinuous-it does not have a closed graph. Similarly, the correspondence represented in Figure M.H.2(b) is upper hemicontinuous but it fails to be lower hemicontinuous (consider the illustrated sequence xM -+ x that approaches x from below and the point y E fIx)). Roughly speaking, upper hemicontinuity is compatible only with "discontinuities" that appear as "explosions" of sets [as at x =! in Figure M.H.2(b)], while lower hemieontinuity is compatible only with "implosions" of sets [as at x = t in Figure M.H.2(a)). As with upper hemicontinuous correspondences, if f(·) is a function then the concepts of lower hemicontinuity as a correspondence and of continuity as a function coincide. Finally, when a correspondence is both upper and lower hemicontinuous, we say that it is conti/luous. An example is illustrated in Figure M.H.3.
In many applications, the range space Y of f(·) is itself compact. In that case, the upper hemicontinuity property reduces to the closed graph condition. In Figure M.H.l (a), we represent a correspondence (in fact, a function) having a closed graph that is not upper hemicontinuous. In contrast, the correspondence represented in Figure M.H.I(b) is upper hemicontinuous. The upper hemicontinuity property for correspondences can be thought of as a natural generalization of the notion of continuity for functions. Indeed, we have the result of Theorem M.H.!. Theorem M.H.1: Given A c RN and the closed set Y c: RK, suppose that I: A -+ Y is a single-valued correspondence (so that it is, in fact, a function). Then 1(') is an upper hemicontinuous correspondence if and only if it is continuous as a function.
Figure M.H.2
Proof: If f(·) is continuous as a function, then Definition M.F.2 implies that f(·) has a closed graph (relative to A x Y). In addition, Theorem M.F.2 teIls us that images of compact sets under f(·) are compact, henae bounded. Thus, f(') is upper hemicontinuous as
f(i) =
(x',
i"
a correspondence. M
A=[O.I]
.(x~.)',~)
: I .- _ .
M
:
A=[O.I] y= [0. J]
y= [0.3]
r')
,.1::. (x:;:)
Suppose now that f(') is upper hemicontinuous as a correspondence and consider any sequence x -+ x E A with x E A for all m. Let S = {x m = 1,2, ... } u {x}. Then there exists an r> 0 such that 11-,'11 < r if x' E S.18 Because S is also closed, it follows that S is compact. M
P:
)
f(
hemicontinuous
Ix'.y')
correspondence that is not upper
~ "" , " ,,, "
16. See Section M.F ror a discussion of bounded and compact sets. 17. It can be verified that Definition M.H.3 implies that the image of a compact set under an upper hemicontinuous correspondence is in (act compact (i.e .• closed and bounded), a property also shared by continuous runctions (see Theorem M.F.2). 18. To sec this. recall that if x.., - x, then for any £ > 0 there is a positive integer M( such that II_,M - xII <, ror all m> M,. Hence, ror any r> Max {lIx'II, ... , IIx"'II, Ilxll + c}, we have 11."11 < r ir x' E S.
o
.'(1 X'" X
(a)
hemicontinuous.
(b) An upper
o
x' x'"
~
hemicontinuous
=x
(b)
19. For another source of examples, note that any correspondence f: A _ Y with an open graph (relative to A x Y) is lower hemicontinuous.
..
Upper and lower hemicontinuous correspondences.
correspondence that is not lower hemicontinuous.
952
....
THE .... TIC .. L
SECTION
"PPENDIX
-------------------------------------------------------A = [0,1] Y = [0,3]
1'""
/
01 fl·'
/
Figure M.H.3
A continuous correspondence.
o
o
/
//
x
/1
o (a)
/
/
/
/
I
[(.)
/
iI I
i i x
/
/
/~I
/
/
/
/)17 [()
x
The logic of Kakutani's fixed point theorem is illustrated in Figure M.1.2(a) for condition we could have cases such as that in Figure M.1.2(b) where no fixed point exists. Finally, we mention a fixed point theorem that is of a different style but that is being found of increasing relevance to economic applications.
Brouwer's fixed point theorem. (a) A continuous function from [0. 1]10 [0, I] has a fixed point. (b) The continuily assumption is indispensable.
Theorem M.I.3: (Tarsky's Fixed Point Theorem) Suppose that f: [0, l]N -+ [0, l]N is a nondecreasing function, that is, fIx') ~ fIx) whenever x' ;;>; x. Then f(') has a fixed point; that is, there is an x E A such that x = fIx). Tarsky's theorem differs from Brouwer's in three respects. First, the base set is not any compact, convex set, but rather a special one-an N-product of intervals. Second, the function is required to be nondecreasing. Third, the function is not required to be continuous. The logic of Tarsky's fixed point theorem is illustrated in Figure M.1.3 for the case N = I. In the figure, the function [(.) is not continuous. Yet, the fact that it is nondecreasing forces its graph to intersect the diagonal.
f(')
/
o (b)
-
THEOREMS
953
/
N = I. Note that the convexity of the set [(x) for all x is indispensable. Without this
Figure M.1.1
/
/
/
/
Theorem M.1.2: (Kakutani's Fixed Point Theorem) Suppose that A c It" is a nonempty, compact, convex set, and that f: A -+ A is an upper hemicontinuous correspondence from A into itself with the property that the set fIx) c A is nonempty and convex for every x E A. Then f(') has a fixed point; that is, there is an x E A such that XE fIx).
The logic of Brouwer's fixed point theorem is illustrated in Figure M.1.l (a) for the easy case where N = I and A = [0,1). In this case, the theorem says that the graph of any continuous function from the interval [0, I] into itself must cross the diagonal, and it is then a simple consequence of the intermediate value theorem. In
/
/
/
/
particular, if we define the continuous function q,(x) = [(x) - x, then q,(0) ~ 0 and E [0,1]; hence, [(x) = x for some x E [0, I). In Figure M.I.l(b) we can see that, indeed, the continuity of [(-) is required. As for the convexity of the domain, consider the function defined by a 90-degree clockwise rotation on the circle S = {x E R2: Ilxll = \}: It is a continuous function with no fixed point. The set S, however, is not convex. In applications, it is often the case that the following extension of Brouwer's fixed point theorem to correspondence is most useful.
Theorem M.I.l: (Brouwers Fixed Point Theorem) Suppose that A c ~ is a nonempty, compact, convex set, and that f: A -+ A is a continuous function from A into itself. Then f(·) has a fixed point; that is, there is an x E A such that x = fIx).
/
/
/
POINT
q,( I) !> 0, and so q,(x) = 0 for some x
In economics the most frequent technique for establishing the existence of solutions to an equilibrium system of equations consists of setting up the problem as the search for a fixed point of a suitably constructed function or correspondence [: A -+ A from some set A c R' into itself. A vector x E A is a fixed point of [(.) if x = [(x) [or, in the correspondence case, if x E [(x»). That is, the vector is mapped into itself and so it remains "fixed." The reason for proceeding in this, often roundabout, way is that important mathematical theorems for proving the existence of fixed points are readily available. The most basic and well-known result is stated in Theorem M.1.1.
/
/
/
FIXED
(a)
M.I Fixed Point Theorems
[(x) =
/
/
/
/
/
/
M.I:
Figure M.1.2
Kakutani's fixed point theorem. (a) A fixed point exists. (b) The
con vex-val ued ness assumption is indispensable.
954
MATHEMATICAL
f(x)
APPENDIX
SECTION
R
----_--·------===---;<.L..-I f(·)/,
=x
/1
~
V //
/
/
/
/
I / / Jr""/ /
/
f(·)
I
I I I I
.-(: Local Maximizer
I
.!:
I I
In this section we consider a function f: RN
-+
Flgur. M.1.3 (left)
Tarski's fixed point theorem.
R
Definition M.J.1: The vector X E RN is a local maximizer of f(') if there is an open neighborhood of X, A eRN, such that fIx) ~ fIx) for every x E A. '0 If f(x) ~ f(x) for every x E RN (i.e., if we can take A = RN ), then we say that x is a global maximizer of f(·) (or simply a maximizer). The concepts of local and global minimizers are defined analogously. In Figure M.J.I, we illustrate a local maximizer x and a local minimizer ~ (with open neighborhoods A and A', respectively) ofa function for the case in which N = I. Theorem M.J.1: Suppose that f(') is differentiable and that x E RN is a local maximizer or local minimizer of f(·). Then
iJxn
-+
R is twice continuously differen-
(i) If x ERN is a local maximizer. then the (symmetric) N x N matrix 02f(x) is negative semidefinite. (ii) If 02f(x) is negative definite. then x is a local maximizer.
MJ Unconstrained Maximization
iJf(x) = 0
does not hold. Consider for example, the function fIx"~ x,) = (x,)' - (x,)' defined on R'. At the origin we have VflO, 0) = (0,0). Thus, the origin is a critical point, but it is neither a local maximizer nor a local minimizer of this function. To characterize local maximizers and local minimizers of f(·) more completely, we must look at second-order conditions. Theorem M.J.2: Suppose that the function f: RN tiable and that fIx) = O.
I
x
UNCONSTRAINED
Local Minimizer
I I
/
M.J:
for every n,
(M.J.1)
or, in more concise notation, Vf(x) = O.
(M.J.2)
Proof: Suppose that x is a local maximizer or local minimizer of f(·) but that df(x)/ox, = a > 0 (the argument is analogous if a < 0). Denote bye' ERN the vector having its nth entry equal to I and all other entries equal to 0 (i.e., having eZ = I and eZ = 0 for II # n). By the definition of a (partial) derivative, this means that there is an r. > 0 arbitrarily small such that [fIx + £e') - f(x))/e > a/2 > 0 and [I(:;; - ee') - f(x)]/e < - a/2. Thus, flo' - ee') < fIx) < fIx + ee'). In words: The function f(·) is locally increasing around x in the direction of the nth coordinate axis. But then .x can be neither a local maximizer nor a local minimizer of f(·). Contradiction. _ The conclusion of Theorem M.J.I can be seen in Figure M.J.I: In the figure, we have af(x)/ox = 0 and iJf(~)/iJx = O. A vector :i' E RN such that Vf(_x) = 0 is called a critical point. By Theorem M.J.I, every local maximizer or local minimizer is a critical point. The converse, however, 20. An open neighborhood of .x is an open set that includes i.
Replacing "negative" by "positive," the same is true for local minimizers. Proof: The idea is as follows. For an arbitrary direction of displacement z E RN and scalar r.. a Taylor's expansion of the function "'(e) = fIx + ez) around e = 0 gives
nx + I:z) -
Figure M.J.1 (right)
Local maximizers and minimizers.
fUe) = eVf(x)·z + le'z'D'f(x)z + Remainder =
!e'z' D' f(.i)z + Remainder,
where I.' E R + and (l/e ) Remainder is small if e is small. If x is a local maximizer, then for I.' small we must have (I/e')[f(x + ez) - fIx)] ~ 0, and so taking limits 2
we get
z'D'f(x)z ~ O. Similarly. if =. D2 f(x)z < 0 for any z # 0, then fIx so x is a local maximizer. _
+ ez) < fIx) for e > 0 small, and
In the borderline case in which D'f(x) is negative semidefinite but not negative delinite, we cannot assert that x is a local maximizer. Consider, for example, the function fIx) = x, whose domain is R. Then D' flO) is negative semidefinite because ci 2 f(O)/cix = 0, but x = 0 is neither a local maximizer nor a local minimizer of this function. Finally, when is a local maximizer x of f(·) (or, more generally, a critical point) automatically a global maximizer? Theorem MJ.3 tells us that a sufficient condition is the concavity of the objective function f(·). Theorem M.J.3: Any critical point x of a concave function f(') [i.e., any Vf(x) = 0] is a global maximizer of f(·).
x satisfying
Proof: Recall from Theorem M.CI that for a concave function we have f(x) :::; - .x) for every x in the domain of the function. Since V fIx) = 0, this tells us that .x is a global maximizer. _
lUi) + V f(·i)· (x
By analogous reasoning, any critical point of a convex function f(·) is a global m;/I;m;=er of f( . ).'1
21. In facl. Ihis follows directly from Theorem MJ.3 because x is a global minimizer of f(·) if and only if it is a global maximizer of - f('), and - f(') is concave if and only if f(·) is convex.
MAXIMIZATION
955
56
MATHEMATICAL
SECTION
APPENDIX
Proof: The role of the constraint qualification is to insure that maximizer in the linearized problem
'ILK Constrained Maximization We start by considering the problem of maximizing a function f(·) under M equality constraints. Namely, we study the problem Max
M.K:
fix)
Max
fix)
+ V f(x)· (x
CONSTRAINED
x is
MAXIMIZATION
957
also a local
- x)
xeRN
s.t. Vg,(x)'(x - x) = 0
(M.K.I)
xeRN
VYM(X)'(X - x) = 0, gM(X) = hM' where the functions f(- ), y, (-), ... ,YM(') are defined on IRN (or, more generally, on an open set A c IR N ). We assume that N ;:: M; if M ;:: N there will generally be no points satisfying all of the constraints. The set of all x E IRN satisfying the constraints of problem (M.K.I) is denoted C = (x
E
hm
IR N : y .. (x) =
for
m = I, ... , M}
and is called the cOlIStrailll set. The definitions of a local constrained or a global cOllSCrailled maximizer arc parallel to those given in Definition MJ.I, except that we now consider only points x that belong to the constraint set C. The feasible point .X E C is a local constrained maximizer in problem (M.K.I) if there exists an open neighborhood of .x, say A c IRN, such that fix) ;:: fix) for all x EAr. C, that is, if .i solves problem (M.K.I) when we replace the condition x E RN by x E A. The point .x is a global constrained maximizer if it solves problem (M.K.I), that is, if fi-i) ;:: fix) for all x E C. Our first result (Theorem M.K.I) states the first-order conditions for this constrained maximization problem. Theorem M.K.1: Suppose that the objective and constraint functions of problem (M.K.1) are differentiable and that x E C is a local constrained maximizer. Assume also that the M x N matrix og,(x)
ag,I" oXN
ox,
[ 'g.l"
1
ogMix)
ox,
oXN has rank M. (This is called the constraint qualification: It says that the constraints are independent at x.) Then there are numbers i. m E IR, one for each constraint, such that OI(x) = [ i. ogm(x) m oXn mo' oXn
for every n = 1, ... , N,
in which the objective function and constraints have been linearized around the point .t. ThIlS, the constraint qualification guarantees the correctness of the following intuitively sensible statement: If .x is a local constrained maximizer, then every direction of displacement z E IRN having no first-order effect on the constraints, that is, satisfying V Y.. (x), z = 0 for every m, must also have no first-order effect on the objective function, that is, must have Vf(x)·z = 0 (see also the discussion after the proof and Figure M.K.I). From now on we assume that this is true. The rest is just a bit of linear algebra. Let E be the (M + I) x N matrix whose first row is Vf(.i)T and whose last M rows are the vectors Vy,(X)T, ... , Vy.,(X)T. By the implication of the constraint qualification cited above, we have {z ERN: Vy .. (x)·z = 0 for all m} = {z E IR N : Ez = OJ. Hence, these two linear spaces have the same dimension M. Therefore, by a basic result of linear algebra, rank E = M. Hence, V f(.x) must be a linear combination of the linearly independent set of gradients Vg,(x), ... , VgM(x). This is exactly what (M.K.3) says. _ I n words, Theorem M.K. I asserts that, at a local constrained maximizer x, the gradient of the objective function is a linear combination of the gradients of the constraint functions. The indispensability of the constraint qualification is illustrated in Figure M.K.l. In the figure, we wish to maximize a linear function f(x" X2) in the constraint set C = {(X"X2)E R 2 :g .. (X"x 2 ) = h.. form = 1,2} [the figure shows the loci of points satisfying g,(x"x 2) = h, and g2(X"X2) = b2, as well as the level sets of the function f(')]' While the point x is a global constrained maximum (it is the only vector in the constraint set!), we see that V fix) is not spanned by the vectors Vy,(.x) and VY2(X) [i.e., it cannot be expressed as a linear combination ofVg,(x) and VY2(';:)]. Note, however, that Vg,(,;:) = - Vg 2 (,;:), and so the constraint qualification Figure M.K.1
x,
Indispensability of the constraint qualification.
(M.K.2)
p." ... , )'M)]'
or, in more concise notation [letting ). = M
Vf(x) =
Lo
m
;'mVgm(x)
'
The numbers i. m are referred to as Lagrange multipliers.
(M.K.3)
f(x)
958
MAT HEM A TIC A l A P PEN D I X
SECTION
-------------------------------------------------------is violated. Observe that, indeed, x is no! a local maximizer of fIx) [= fIx) since f(·) is linear] in the linearized constraint set C
= {(Xt, x,) E IR': Vgm(X)'(X
- x)
=0
for
m
+ Vf(x)' (x -
x)
= I, 2}.
Often the first-order conditions (M.K.2) or (M.K.3) are presented in a slightly different way. Given variables x = (x t , ••• , x N ) and ). = (At, ... , i. N ), we can define the Lagrangian function
Note that conditions (M.K.2) [or (M.K.3)] are the (unconstrained) first-order conditions of this function with respect to the variables x = (x" ... , XN)' Similarly, the constraints g(x) = 0 are the first-order conditions of L(', .) with respect to the variables i. = (i.t, ... , AM)' In summary, Theorem M.K.I says that if x is a local constrained maximizer (and if the constraint qualification is satisfied), then for some values i. t , • •• ,i'M all of the partial derivatives of the Lagrangian function are null; that is, ,1L(x, i.)/ox. = 0 for n = I, ... , Nand oL(x, ).)/Oi'm = 0 for m = I, ... , M. Theorem M.K.I implies that if x is a local maximizer in problem (M.K.I), then the N + M variables (x" ... , XN' i. t , •.. , AM) are a solution to the N + M equations formed by (M.K.2) and gm(x) = iim for m = I, ... , M. There is also a second-order theory associated with problem (M.K.I). Suppose that at x the constraint qualification is satisfied and that there are Lagrange multipliers satisfying (M.K.3). If x is a local maximizer, then D;L(x, i.)
= D'f(x) - I)... Vgm(x)
Inequality Constraints We now generalize our analysis to problems that may have inequality constraints. The basic problem is therefore now Max
fIx)
s.t. gt(X) =
(M.KA)
ii t
gM(X) = h t (X);5; f
5,..
CONSTRAINED
where every function is defined on RN (or an open set A c: R N ). We assume that N 2! M + K. It is of course possible to have M = 0 (no equality constraints) or K = 0 (no inequality constraints). We again denote the constraint set by C c: RN, and the meaning of a local constrained maximizer or a global constrained maximizer is unaltered from above. We now say that the constraint qualification is satisfied at x E C if the constraints that hold at x with equali!y are independent; that is, if the vectors in {Vgm(x): m = I, ... , M} u {Vh.(x): h.(x) = f.} are linearly independent. Theorem M.K.2 presents the first-order conditions for this problem. All of the functions involved are assumed to be differentiable. Theorem M.K.2: (Kuhn- Tucker Conditions) Suppose that x E C is a local maximizer of problem (M.K.4). Assume also that the constraint qualification is satisfied. Then there are multipliers Am E R, one for each equality constraint, and Ak E R+, one for each inequality constraint, such that" (i) For every n = 1, ... ,N, at(ii) =
OXn
I m='
i. °Ym(ii) m OXn
+
f
Ak ohk(x) ,
k-'
ox
(M.K.5)
or, in more concise notation, M
Vf(ii) =
K
I
i· m VYm(ii) +
m""1
I)·. Vhk(ii).
(M.K.6)
k-1
(ii) For every k = 1, ... , K,
i.• (h.(ii)
is negative semidefinite on the subspace {t ERN: Vgm(x)' t = 0 for all mi. In the other direction, if the vector x is feasible (i.e., x E C) and satisfies the first-order conditions (M.K.2), and if D!L(x, A) is negative definite on the subspace {z ERN: Vgm(x)·z = 0 for all m}, then x is a local maximizer. These conditions can be verified using the determinantal tests provided in Theorem M.DJ. Finally, note that a local constrained minimizer of f(·) is a local constrained maximizer of - f('), and therefore Theorem M.K.I and our discussion of secondorder conditions above is also applicable to the characterization of local constrained minimizers.
M.K:
-
c.)
= 0,
(M.K.7)
i.e., A. = 0 for any constraint k that does not hold with equality. Proof: We illustrate the proof of the result for the case in which there are only inequality constraints (i.e., M = 0). As with the case of equality constraints, the role of the constraint qualification is to insure that x remains a local maximizer in the problem linearized around x. More specifically, we assume from now on that the following is true: Any direction of displacement tERN that satisfies the constraints to first order [i.e., such that Vh.(x)·z ;5; 0 for every k with h.(x) = f.] must not create a first-order increase in the objective function, that is, must have Vf(x)'z ;5; O. In Figure M.K.2 we represent a problem with two variables and two constraints for whieh the logic of the result is illustrated and made plausible. The Kuhn-Tucker theorem says that if x is a local maximizer then V fIx) must be in the cone
r
=
{y E IR': y =
).t Vht(x)
+ )., Vh,(x) for some (A" A,) 2! O}
depicted in the figure; that is, Vf(.ii) must be a nonnegative linear combination of Vht(x) and Vh,(x). Suppose now that x is a local maximizer. If starting from x we move along the boundary of the constraint set to any point (x, + Zt, x, + z,) with Zt < 0 and z, > 0, then in the situation depicted we would have ht(x t +
Zt,
X2 + Z,) = Ct ,
h,(x t + Z,' x, + Z,) < f"
t
22. By convention, if there are no constraints (Le., if M = K = 0). then the right-hand side of (M.K.5) is zero.
MAXIMIZATION
95S
,0
MATHEMATICAL
APPENDIX
SEC T ION
... K:
CON S T R A I NED
condition for x. is now
of(x)
ax. and we have the added condition that -A.+,X. = O.
Figure M.K.2
h~(x) :5 (~~
II I (x) :5 (\
Necessity of the Kuhn-Tucker
"
conditions.
and f{·' , + :, ..\, + :,) $ f(·\;). Taking limits we conclude that in the direction z such that 'VII,{.i)·: = 0 and 'VII,{.Xl·z < 0, we have 'Vf(x)'z S; O. Geometrically this means that the vector 'V fe;;) must lie below the vector 'VII,(.i:), as is shown in the figure. By similar reasoning (moving along the boundary of the constraint set C in the opposite direction), if .i: is a local constrained maximizer the vector 'Vf(.i:) must lie above the vector 'VII,(x). Hence, 'V fix) must lie in the cone r. This is precisely what the Kuhn-Tucker conditions require in this case. The above intuition can be extended to the general case. Suppose that all the constraints are binding at x (if a constraint k is not binding, put i., = 0 and drop it from the list). We must show that 'V fix) belongs to the convex cone
r
=
{Y
N E IR :
y=
~ i., 'Vh,(.i)
for some
(i." ... , i. K )
o}
C IRN.
Assume for a moment that this is not so, that is, that 'V fix) ~ r. Then, by the separating hyperplane theorem (Theorem M.G.2), there exists a nonzero vector Z E IRN and a number fiE IR such that 'V f(x) , Z > Pand y' Z < Pfor every Y E r. Since 0 E r we must have fI > O. Hence, 'V f(x)' Z > O. Also, for any Y E we have Oy E for all () 0 and 'Vh,(.i)·z S; 0 for all k, which contradicts the linearization implication of the constraint qualification. _
r
r
It is common in applications that a constraint takes the form of a nonnegativity requirement on some variable x.; that is, x.;:: O. In this case, the appropriate first·order conditions require only a small modification of those above. In particular, we need only change the first·order condition for x. to ()J}i:) S;
ax,.
f ".=1
i.
m
vg m(x2 + ox,.
f 1e""'1
i., ~m(X), ox,.
with equality if X. > O.
(M.K.8)
To see why this is so, suppose that we explicitly introduced this nonnegativity requirement as our (K + I)th inequality constraint [i.e., hK> ,(x) = -x. S; 0] and let i· K ~ 0 be the corresponding multiplier. Note that i' K +, (iJh. + ,(x)/ox.) = -i.• and vh K + ,(x)/ax•. = 0 for /I' ~ n. Thus, if we apply condition (M.K.S) of Theorem M.K.2, the only modification to the first·order conditions is that the first.order
+,
+,
But these two conditions are exactly equivalent to condition (M.K.8). Given the simplicity of the adjustment required to take account of nonnegativity constraints, it is customary in applications not to explicitly introduce the non negativity constraint and its associated multiplier, but rather simply to modify the usual first-order conditions as in (M.K.8). Note also that any constraint of the form h,(x) ;:: c, can be written as - h.(x) S; -c•. Using this fact, we see that Theorem M.K.2 extends to constraints of the form II.(x)
>0
for any x and x' with fix') > f(x).
(M.K.9)
Then if x E C satisfies the Kuhn-Tucker conditions [conditions (i) and (ii) of Theorem M.K.2], and if the constraint qualification holds at X, it follows that x is a global maximizer. ,. Proof: Suppose that this is not so, that is, that fix) > fix) for some x E RN satisfying II,{x) $ C, for every k. Denote Z = x - x. Then, by (M.K.9) we have 'Vf(i)'z > O. If J., > O. then the Kuhn-Tucker conditions imply that h,(x) = c,. Moreover, since 11,(') is quasiconvex and I,,(x) S; c, = h,(x), it follows that 'Vh,(x)'z S; O. Hence, we have both 'V f(.i)· z > 0 and Lk )., 'Vh,(x)' z S; 0, which contradicts the Kuhn-Tucker conditions because these require that 'V fix) = L, A. 'Vh,(x). _
23. More generally, equality constraints are permissible if they are linear. 24. If instead we have V f(x)'(x' - x) < 0 whenever f(x') < fix) and the multipliers have the nonposilive sign that corresponds to a minimization problem. then x is a global minimizer.
.. A X I .. I Z A T ION
961
962
MATHEMATICAL
SECTION
APPENDIX
x,
R f(·)
Vfex)
= i.,Vh,(.<) + i.,Vh,(i)
",(x) ~"
c
fIx)
= fIx)
I,,(x) $ i',
Note that condition (M.K.9) of Theorem M.K.3 is satisfied if 1(') is concave or if I(') is q uasieoneave and VJ(x) #- 0 for all x E RH. The condition that the constraint functions 10 ,(.), ... , ".(.) are quasiconvex implies that the constraint set C is convex (check this)." In Figure M.K.3, we illustrate the theorem for a case in which N = K = 2, M = 0, and J(.) is a quasiconcave function with VJ(x) #- 0 for all x. The indispensability of condition (M.K.9) in Theorem M.K.3 is shown in Figure M.K.4. There we have N = M = I, and the quasiconcave function J(.) is being maximized on the constraint set C = {x E R: h(x) :s OJ. In the figure, the point x satisfies the Kuhn-Tucker conditions for a multiplier value of A = 0, but is not a global maximizer of 1(') on C (the point x· is the global constrained maximizer). Note, however, that condition (M.K.IO) is violated when x = x and x' = x·. We observe finally in Theorem M.KA an important implication of the constraint set C being convex and the objective function 1(') being strictly quasiconcave.
x
Theorem M.K.4: Suppose that in problem (M.K.4) the constraint set C is convex and the objective function f(·) is strictly quasiconcave. Then there is a unique global constrained maximizer.26 Proof: If x and x' #- x were both global constrained maXImizers, then the point x" = ~x + (I - Q:)x' for Q: E (0, I) would be feasible (i.e., x· E C) and by the strict quasiconcavity of I('), would yield a higher value of 1(') [i.e., J(x') > J(x) = J(x')]. • Suppose that in the case in which only inequality constraints are present we denote by C _, Ihe relaxed conSiraint set arising when the kth inequalily constrain I is dropped. The following Iwo facts are oflen useful in applications. (i) Iff(.i:)
M.k:
CONSTRAINED
(ii) Suppose that all of the constraint functions h, (·), .... h.(·) are continuous and quasiconvex and that condition (M.K.9) holds. Then if x is a solution to problem (M.K.4) in which Ihe kth constraint is not binding [i.e., if h,(x) < c,], we have f(x) ;;" f(x) for all x E C,. That is, under the staled assumptions, if a constrainl is not binding at a solution 10 problem (M.K.4), then ignoring il altogether should have no effect on the solution. To see this, suppose otherwise; i.e .. Ihat there is an x' E C_, such thai fIx') > f(x). Then because the constraint funclions h,(·), .. . , hd') are quasiconvex, we know Ihat the point x(a) = ax' + (I - a).< is an element of C _, for all a E [0,1]. Moreover, since the kth constraint is not binding at i, there is an IX > 0 such Ihat h,(x(a)) < l, for all , < 2. Hence. x(a) e C for all a < IX. But Ihe derivalive of f(x(a)) at a = 0 is V f(xHx' - x) > 0 [recall that condition (M.K.9) holds and that. by assumption fIx') > f(x)], Therefore, there must be a point x(a) E C such Ihat f(x(,)) > f(·<)-a contradiction to x being a global constrained maximizer in problem (M.KA).
Figure M.K.3 (left)
With quasiconvex constraint functions and a quasiconcave objective function salisfying Vf(x) # 0
for all x, satisfaction of Ihe Kuhn-Tucker
x implies thai x is a global constrained ma~imizer. conditions at
Comparative Statics In our previous discussion we have treated the parameters b = (b ... , hAt) and " of problem M.K.4 as given. We will now let them vary. Suppose that (b, e) E RAt are parameters for which problem (M.K.4) has some solution i(b, c) and denote the value or this solution by u(b. c) = J(.i(b, el). Under fairly general conditions (see the small type at the end of this section), the value r(b, c) depends continuously on the parameters (b, c). Theorem M.K.5 provides an interpretation for the Lagrange multipliers as the "shadow prices" of the constraints.
c = (c" ... , i\)
+.
Figure M.K.4 (right)
The necessity of condition (M.K.IO) for Theorem M.K.3.
Theorem M.K.S: Suppose that in an open neighborhood of (b, c) the set of binding constraints remains unaltered and that v(b, c) is differentiable at (b, c)." Then for every m = 1, ... , M and k = 1, ...• K we have
iJv(b, c)
.
a b = I 'm
m
and
iJv(b. c) . ~~-=I'k' iJc k
Proof: This is a particular case of the envelope theorem (Theorem M.L.I) to be presented in the next section . • Consider a more general optimization problem. We maximize a function f: RN - R subject to x E C(q) where C(q) is a nonempty conslraint set and q = (q, ... , qs) belongs to an admissible set of parameters Q c RS . Suppose that f(·) is continuous and that C(q) is compact for every q E Q. Then we know [from Theorem M.F.2, part (ii)] that the maximum problem has at least one solulion. Denote by x(q) c C(q) Ihe sel of solutions corresponding to q and by "(q) [= fIx) for any x E x(q)] the associaled maximum value. Theorem M.K.6 concerns the continuily of x(·) and v(·). Theorem M.K.6: (Theorem of the Maximum) Suppose that the constraint correspondence C: Q _ RN is continuous (see Section M.H) and that f(') is a continuous function. Then the maximizer correspondence x: Q - !'IN is upper hemicontinuous and the vatue function v: Q -+ R is continuous.
27. These are simplirying assumptions; a similar result holds more generally but requires the usc of directional derivatives at points or nondifferentiabitity or the runction v(· • . ).
MAXIMIZATION
963
MATHEMATICAL
SECT'ON
APPENDIX
if q, < q"
(CO, q,)}
if q, < q"
x(q) =
X,
+ x,
=
- . . . - - - 1(,"; .) ~r--
Flgur. M.L.1
But note, and this is a key observation, that by the first-order conditions for unconstrained maximization (see Section MJ), we must have of(x(q); ij)/ox = O. Therefore, (M.L.2) simplifies to
dv(ij) dq
In this section, we return to the problem of maximizing a function f(-) under constraints, but we suppose that we want to keep track of some parameters q = (q" ... ,qs) E RS entering the objective function or the constraints of the problem. In particular, we now write the maximization problem as
b, bM •
We denote by v(') the value function of problem (M.L.I); that is, v(q) is the value attained by f(·) at a solution to problem (M.L.I) when the parameter vector is q. To be specific, we suppose that v(q) is well-defined in the neighborhood of some reference parameter vector ij E RS. It is then natural to investigate the marginal effects of changes in q on the value v(q). The envelope theorem addresses this matter. 28 It will be convenient from now on to assume that, at least locally (i.e., for values of q close to ij), the solution to problem (M.L.!) is a (differentiable) function x(q). We can then write v(q) = f(x(q); q). To start with the simplest case, suppose that there is a single variable and a single parameter (i.e., N = K = I) and that there are no constraints (Le., M = 0). Then, by the chain rule,
of(x(ij); ij)
dq
oq
--=
+
of(x(ij); ij) dx(ij) ox
dq
.
of(x(q), ij) oq
(M.L.3)
That is, the fact that x(q) is determined by maximizing the function f(·; q) has the implication that in computing the first-order elfects of changes in q on the maximum value, we can equally well assume that the maximizer will not adjust: The only elfect of any consequence is the direct effect. This result is illustrated in Figure M.L.I, which also motivates the use of the term "envelope." In the figure we represent the function fIx; .) for different values of x. Because at every q we have v(q) = Max. fIx; q), the value function v(·) is given by the upper envelope of these functions. Suppose now that we consider a fixed ij. Then, denoting x = x(ij), we have f(x.; q) s v(q) for all q, and f(x.; ij) = v(ii). Hence, the graph of fIx; .) lies weakly below the graph of v(') and touches it when q = ij. So the two graphs have the same slope at that point. This is precisely what condition (M.L.3) says. We now state the general envelope theorem for a problem with any number of variables, parameters, and constraints. As we will see, its conclusion is similar to (M.L.3), except that Lagrange multipliers play an important role.
(M.L.I)
fIx; q)
dv(ij)
__ I(x"; .)
The envelope theorem.
1.L The Envelope Theorem
g.,(x; q) =
965
if q, = q,.
q.}
s.t. g,(x; q) =
THEOREM
,n
Both the objective function and the constraint correspondence are continuous (you should check the latter). In accordance with Theorem M.K.6, x(·) is upper hemicontinuous. But it is not continuous (there is an explosion along the line q I = q,). On the other hand, suppose that we take Q = [0, I]'. Then the conclusion of the theorem fails: the maximizer correspondence is not upper hemicontinuous [we have x(2£, £) = {(O, 2e)}, but x(O, 0) = {(I, l))]. However, the assumptions also fail: at q = (0,0) the vector (I, I) belongs to the constraint set, but at q = (e, £) no vector x with X, + X, > £ belongs to the constraint set. Hence the constraint correspondence is not continuous once extended to Q = [0,1]'.
Max
ENVELOPE
I(x(ij); .)
and x(q) = {(x" x,) e [0,1]':
THE
R
The result cannot be improved upon. Suppose that we maximize x I + x, subject Q = (0,1)'. Then the
to x, e [0,1], x, e [0,1], and q,x , + q,x, S q,q, for q = (q" q,) e maximizer correspondence is given by x(q) = {(q"O)}
M.L:
Theorem M.l.l: (Envelope Theorem) Consider the value function v(q) for the problem (M.L.l). Assume that it is differentiable at q E W' and that (A." . .. ,A. M ) are values of the Lagrange multipliers associated with the maximizer solution x(q) at q. Then 19 ov(q) = of(x(q); q) _
oqs
(M.L.2)
oqs
~ A. ogm(x(q); q) m-'
m
oqs
for s = 1, ... ,S,
(M.L.4)
29. If we have a case with inequality constraints in which the set of binding constraints remains
unaltered in a neighborhood of ii, then expressions (M.L.4) and (M.L.5) are still valid: Accounting for the nonbinding constraints will have no effect either on the left-hand side or on the right-hand side
28. Formally, we are presenting a case with equality constraints. But note that as long as in a neighborhood of the parameter vector under consideration the set of binding constraints does not change, the discussion applies automatically to the case with inequality constraints.
(because its associated multipliers are zero).
-
966
MATHEMATICAL
APPENDIX
SECTION
M.M:
LINEAR
PROGRAMMING
967
or, in matrix notation, M
L
Vv(ql = Vqf(x(q); 17) -
(M.L5)
Am Vqgm(x(q); 17)·
m=1
Proof: We proceed as in the case of a single variable and no constraints. Let x(·) be the maximizer function. Then v(q) = f(x(q); q) for all q, and therefore, using the chain rule, we have
aV(ij) = af(x(ij); ij) oq, aq,
+
t (af(x(ij); ij) ax.@).
.~,
ax.
aq,
The first-order conditions (M.K.2) tell us that
af(x(ij); Ii) + OX,.
I
Am °Ym(x(ij); ij).
ax,.
m= I
4i., + 7i., = 3
Hence (switching the order of summation as we go),
DV(ij) = af(x(ij); ij) aq, oq. Moreover, since Ym(x(q); q) =
f
+
J.
m
t (iJYm(X(ij); ij) ax.@).
m~'.~'
bm for
ox.
Figure M.M.I represents a linear programming problem with N = 2, the two constraints 2xI + X, S 4 and XI + 3x, S 7, and the objective function x, + X" A most interesting fact about the linear programming problem (M.M.I) is that with it we can associate another linear programming problem, called the dual problem, that has the form of a minimization problem with K variables (one for each constraint of the original, or primal, problem) and N constraints (one for each variable of the primal problem):
oq.
all q, we have
~ (CYm(X(ij); ij) OX.@) = _ aYm(x; ij) L.. ax. iJq. aq.
•=,
for all m = I, ... , M .
Combining, we get (M.LA) . •
M.M Linear Programming
Min
(M.M.2)
(AI •••.• J.,.)~O
Linear programming problems constitute the special cases of constrained maximization problems for which both the constraints and the objective function are linear in the variables (x" ... , x N ). A general linear programming problem is typically written in the form (M.M.I)
Max
or, in matrix notation,
(XI.···.X.,,)~O
Max
c·i.
AeR~
s.t. AT;. ~
Of,
in matrix notation,
Max
f'x
xeRI'i
s.t. Ax S c,
where A is the K x N matrix with generic entry a, .. and are (column) vectors. 30
f
E
IR H,
X E
IR H, and
CE
IRK
30. We say that this is the general form of the linear programming problem because, first, an equality constraint a'x == b can always be expressed as two inequality constraints (a'x ~ band - 0 ' x ~ b) and, second, a variable XII that is unrestricted in sign can always be replaced by the difference or two variables (XII. - x lI _), each restricted to be nonnegative.
f,
where i. E IRK is a column vector. Figure M.M.2 represents the dual problem associated with Figure M.M.1. The constraints are now 2.1, + Al ~ I and i., + 3)'2 ~ I, and the objective function is now 4)., + 7),2' Suppose that X E IR~ and A E R~ satisfy, respectively, the constraints of the primal and the dual problems. Then
f·x S (ATJ.)·x = ;"(Ax) S ).-c =
c,)..
(M.M.3)
Thus, the solution value to the primal problem can be no larger than the solution value to the dual problem. The duality theorem of linear programming, now to be stated, says that these values are actually equal. The key for an understanding of this fact is that, as the notation suggests, the dual variables (A" ... , J. K ) have the interpretation of Lagrange multipliers.
Figure M.M.1 (Iell)
A linear programming problem (the primal). Figure M.M.2 (right)
A linear programming problem (the dual).
968
MATHEMATICAL
SECTION
APPENDIX
Theorem M.M.1: (Duality Theorem of Linear Programming) Suppose that the primal problem (M.M.1) attains a maximum value VER. Then v is also the minimum value attained by the dual problem (M.M.2).
Hence, ). satisfies the constraints of the dual problem (since AT). ~ f) and
vIz)
We can verify the duality theorem for the primal and dual problems of Figures M.M.I and M.M.2. The maximizer vector for the primal problem is x = (1,2), !), yielding a value of I + 2 = 3. The minimizer vector for the dual problem is). = yielding a value of 4(j) + 7m = 1f = 3.
= Max
Theorem M.N.1: Suppose that f: A ..... R is a continuous function such that for every z € A the Bellman equation is satisfied; that is, f(z) = Max
u(z, z')
+ M(z')
(M.N.2)
z'EA
for all z E A. Then the function f(') coincides with v(·); that is, f(z) = vIz) for every zEA.
Proof: Successively applying (M.N.2) we have that, for every T,
T-, f(z) = Max lx, 1[., 0
L
cI'u(x" x,+ ,)
,-0
s.t. x, E A for all Xo
+ ciT f(XT) t ~ T,
= z.
But as T ..... 00, the term cI T f(·) makes an increasingly negligible contribution to the sum. We conclude therefore that f(z) = vIz). -
M.N Dynamic Programming Dynamic programming is a technique for the study of maximization problems defined over sequences that extend to an infinite horizon. We consider here only a very particular and simple case of what is a very general theory [an extensive review is contained in Stokey and Lucas with Prescott (1989)]. Suppose that A c RN is a nonempty, compact, set. l2 Let u: A x A ..... IR be a continuous function and let .1 E (0, I). Given a vector Z E A (interpreted as the initial condition of the variables {x,},";.o), the maximization problem we are now interested in is ~
Ixcl{-.o
+ bv(z').
It is perhaps more surprising that, as shown in Theorem M.N.I, the value function is the only function that satisfies this equation.
a,
Max
u(z, z')
%'''A
(M.MA) Now, by (M.M.3), we know that c''! ~ f·x for all ). E IR~ such that AT). ~ f. Therefore c,)' ~ c'). if AT,! ~ f. So (M.MA) tell us that, in fact, ). solves the dual problem (M.M.2) and therefore the value of the dual problem, c')., equals f'x, the value of the primal problem. _
DYNAMIC
It is fairly clear that for every z € A the value function satisfies the so-called Bellman equation (or the Bellman optimality principle):
x E IRN
be a maximizer vector for problem (M.M.!). Denote by ). = 0" ... ').K) ~ 0 the Lagrange multipliers associated with this problem (see Theorem M.K.2)." Formally, we regard ). as a column vector. Then, applying Theorem M.K.2, we have AT). = f and ).·(c - Ax) = O. Proof: Let
M.N:
L b'u(x" x,+ ,)
(M.N.I)
t-O
s.t. x, E A for every t, Xo
= z.
It is not mathematically difficult to verify that a maximizer sequence exists for problem (M.N.I) and that, therefore, there is a maximum value v{z). The function v: A ..... R is called the value function of problem (M.N.I). As with u(', .) itself, the value function is continuous. If, in addition, A is convex and u(', .) is concave, then v( .) is also concave.
Theorem M.N.I suggests a pr.JCedure for the computation of the value function. Suppose that for r = 0 we start with an arbitrary continuous function fo: A ..... R. Think of fo(z') as a trial "evaluation" function giving a tentative evaluation of the value of starting with z' € A. Then we can generate a new tentative evaluation function f, (.) by letting, for every z E A, f,(z) = Max
u(z, z')
+ bfo(z').
t'EA
If f,(') = fo('), then fo(') satisfies the Bellman equation and Theorem M.N.! tells us that, in fact, fo(') = v{'). If f,O # fo(-), then fo(') was not correct. We could then try again, starting with the new tentative f, ('). This will give us a function f2('), and so on for an entire sequence of functions {I.(. )};"-o. Does this take us anywhere? The answer is that it does: For every z E A, we have I.(z) ..... vIz) as r ..... 00. That is, as r increases, we approach the correct evaluation of z. Suppose that the sequence {x,};"..o is a sequence (or a trajectory) that solves the maximization problem (M.N.I). A fortiori, for every 12: I, the decisions taken at I must be optimal. Examining the sum in (M.N.I), we see that x, must solve
Max u(x,_" x,) + Ju(x" x, • .>.
(M.N.3)
x,,,A
31. For linear programming problems the constraint qualification is not required. Put another
way, 'he lineari'y of the constraints is a sufficient form of constraint qualification. 32. The com pac' ness assump'ion cannot be dispensed with entirely, but it can be much weakened.
Assuming that .i, is in the interior of A, (M.N.3) implies that
ou(x,_" x,) + J ou(x" x,.,) OXNh ax"
= 0
(M.N.4)
PROGRAMMING
969
970
'" A THE'" A TIC A LAP PEN 0 1 X
for every n = I •...• N." The necessary conditions captured by (M.NA) are called the Euler equations of problem (M.N.I).
Index 33. Note that the function u(· • . ) has 2N arguments. the N variables of Ihe initial period and the N variables of Ihe subsequent period. In condition (M.N.4l. the variable x, is the nIh component of the initial period. and the variable X N + is the nth component of the subsequent period. II
REFERENCES Chang. A. C. (1984). Fundamental Methods of Mathematical Economics, 3d ed. New York: McGraw-HilL Dixit, A. (1990), Optimization in Economic Theory, 2d ed. New York: Oxford University Press. Intriligator, M. (1971). Mathematical Optimization and Economic Theory. Englewood Cliffs, N.J.: Prentice-
Hall. Novshck, W. (1993). Mathematics for Economists. New York, NY: Academic Press. Simon, C. P .• and L. Blume. (1993). Mathematics for Economists. New York: Norton. Sydsaetcr. K. and P. 1. Hammond. (1994). Mathematics for Economic Analysis. Englewood Cliffs, N.J.: Prentice Hall. 8
Stokey, N., and R. Lucas, with E. Prescott (1989). Recursive Methods in Economic Dynamics. Cambridge, Mass.: Harvard University Press.
AI."Ceplahk ahernalivcs 10 ACllVity OInalysls 154 flO Acyclicity ROO Additive closure 335 Addili ...e ~p"rahility 99. 735 Adlhtivity 13J Admi"sihic date-events pairs 107 Adverse sek'i:tlon 437. 440. 906 Advcrthing 415 Aggregctte consumer surplus 332, 334 Aggrei!ate consumption 318 Aggregate consumption bundles 551 Aggregate consumption vectors 561 Aggregate cost function 148.321.367 Aggregate demand 105 and aggregate wealth 106 and represcntative consumer 116 and weak alliom 109 Aggregate demand functions 109. t 14. 116, 319. ))4. 335. 820 AggregOile excess demOind. convexiftcalion of 629 Aggreg:ttc ex~s demand function 518, 599. 600. 6O~. 613. 61" Aggregate factor demands 5J2 Aggregate Marsh;atlian surplus 326. 329. 330. )31. )71. )87. 448 as welfare meaSure 334 Aggregate (nel) supply theory 141 Aggregate output index 400 Aggregate producer surplus 333 Aggregate production decisions 661 Aggreg.a.le production function 531-2 Aggregate production sets 141, 553. 561 Aggregau: production vector 564 Aggregate profit 148. 333 Aggregate supply correspondence 147 Aggregate supply function 319 Aggregate surplus 326. 329, 330. 331. 371, 387.448 as welfare measure 328 Aggregate wealth and aggregate demand 106 Aggregation 1"7 regularizing efrects of 122 Aggregator function 136 Aggressive variable 416
AHais paradox 179 common reactions to 180 Allocation 307.312.516 competiti ...e 31". 524 definition 5-46 feasible 312. 516. 524. 546-7. 5SO Pareto optimal 313, 523 supportable 524 Walrasian equilibrium 524 Altruism 736 MAnylhing goes" principle 598, 616, 761 Approximate price taking 66S Arbitrage. pricing by 106 Arbitrage·frce asset price vtctor 702 Arrow-Debreu budget tonstr.int 70S Arrow-Debrcu continsent tommodities 108 Arrow-Oebreu equilibrium 691. 691-9. 704. 705.707 .nd Pareto optimality 692 Arrow impossibility theorem 187. 189. 196799.807.809.812. 837. 852 Arrow-Pratt coefficient of absolute risk aversion 190 Arrow securities 700 Asset allocation 711 risk aversion 188. 192 Asset markets 699 Asset portfolio 112 general problem 189. 192 Asset prices 102. 703. 106. 714 martingale property of 708 Asset returns 714 Asset structure 700. 71" complete 104- 7 Assets derivative 700 examples 700 financial 699 long-term 701 primary 700 real 699 returns of 699 safe (or riskless) 100 short-term 707 Asymmetric information 309, 351, 368. 436. 437.477.716. 719.857 and Pareto inefficiency 440
Auction scttm}; K6J-5. K67. 8M9. 903 Average marginal contribution 681 Average of marginOiI utilities 673 Average payoffs in a repeated game 418. 42:! Average productivity 441 Aversion to inequality condition 826 Axiomatic bargaining approach 838 Backward·bcndmg demand curve 343 Backward induction seneralized procedure 277 in finite sames of perfect information 210 Bargaining problem 839 BarSainins procedures 371 Bargaining solutions 839 and independence of irrelevant ahematives 841 and Pareto property 840 and symmetry property 840 consistency properties 845 egalitarian 839 egalitarian solution 841 independent of utility origins (IUO) 839 independent of utility units (IVU) 840 in ...ariant to independent changes of origins 839 invariant to independent changes of units
840 Kalai-Smorodinsky 839. 844 monotonicity properties 84S Na!lh 839. 842 Paretian 840 partially monotone 854 utilitarian solution 839. 841 Basic welfare properties and equilibrium 545 Bayes' rule 284. 286. 452.459 Bayes' theorem 284 Bayesian game 254. 867, 869. 883 Bayesian implementation 883 Bayesian incentive compatibility with linear utility 887 Bayesian incentive compatible social choice function 883. 904 Bayesian Nash equilibrium 255-6. 858. 883. 896.911.91) revelation principle for 884. 901 971
972
IN 0 E X
Behavior strategy 233 Beliefs and sequential rationality 282 domination-based refinements of 468 nolion of 283 "reasonable" 292. 467 structurally consistent 289 system of 28) 8c:llman equation 969 Bellman optimality principle 969 Ikrgson-Samudson socIal weUare function 117.825 Ikrnoulli utility function 184-7, 194,227, 479.489.697. 703.708.709. 7I6.7IK. 7n. 726. 730.858.861.862.866,869, 870.885.887.
Bloc.:king coalition 659 Boldrin-Monlruccio theorem 761 Borda count 794. 795 Bordered Hessian 95. 9)9 Boundedness 185 Brouwer's 6).cd-point theorem 589.952 Bubbl. 771. 774 Budget balance 449. 880. 882 Budget-constrained utility maximization problems 565 Budget constraints 527. 701. 70). 715. 770 Budget hyperplane 21 Budget line 21. 517. 524, 527 Bud,et sets 10,35,517.554.571.608 sequence or 696 Business stealing 408. 409 Call (European) option 700 Capacity constrained duopoly pricing game 395.432 Capacity constraints )89. 394 Capacity investment 417. 423. 426 Capital overaccumulation 739. 741. 753 Cardinal proP'=rties 9, 50 Cardinal transformations 835 Centipede game 281 Certainty equivalence 186 Chain of justification 244 Chain rule 33. 69. 927. 942 Change-of-Iastes model 8 Chaotic equilibrium trajectories 760. 777 Characteristk function 676 Characteristic value 935 Choice 5 beha·.. ior 9 randomized 231 rules 9 relationship \I.'ilh prderence relations 11
IN 0 E X
structures 9. 12. 13. 15, 16 under unocrtainty 167 Circular city model 399 Clarke (pivotal) mechanism 374. 878 Closed convex hull 64 Closed sets 943-4 Coalition 674 dermition 653, 846 Coalitional bargaining 846 Coase theorem 357 Cobb-Douglas economy 597 Cobb-Douglas preferences 612. 915 Cobb- Doughu production function 130, Ill. 142. 544 Cobb-Dougl.. s utihty (unction 55, 56, 63, 519.612 Coefficient of absolute prudence 211 Commodities 17 Commodity futures 700 Commodity space Ig. 567 Commodity-speclrK: demand functions 24 Commodity talation 323. 821 deadweight loss from 84, 331 Commodity vector 18 Common knowledge 226 Compact sets 943 Comparative statics 24, 322. 963 analysis 322. 616 effects of sales lu 323 short-run and lon,-run 339··40 Compensated law of demand 32. 62 Compensated price ch,tnges 32 Compensated price cross-derivalives 70 Compensating variation (CV) 82 Compensaltan principle 334. 829 Compensatory distortion 820 Competition. structure 0( 660 Competitive allocations. welrare·malimizing property ol 531 Competitive budgets 20 Competiti~c economy 314 Competitive equilibrium 307, 312. 318. 322, 359.361-3.438.439.442.445.446. 528.652 definition 314 long-run 334 non-cxistence of 324 Pareto optimality of 325 uniqueness of 320 wa,e 441 stC'lll$() Walrasian equilibrium Competitive limit 411 Competiti\'C markets 311 Competitive scr«ning model 460, 500 Complementary inputs 683 Complements 71 Complete contingent plan 228. 273 Complete impatience problem 761 Complele information 253. SS8. g67. 912 Complete preorder 6 Completely (or lotally) mixed strategy 284 Completeness 6. 9. 13. 42. 744 Composite commodities 44 Composite function 927 Computation of equilibrium 749 Concave fuooion 376. 9)0 Concave utility function 494 Concavily-convexity assumptions S66
Concavity property 826 Conditional density function 479. 503 Conditional factor demand correspondence 139 Condorcct cycle 813 Condorcet pilradox 8. 678, 797, 800, 802, 813 Condorcet winner 802. 803. 80S Cone 132 Congestion effects 360 Consequenti.. list premise 170 Consistency properties 845-6 Constant elasticity 828. 853 Constant elasticity oC substitution 97 Constant rate of SIIvings 746 Constllnt returns convel technology 609 Constant returns to scale 132. 391 Constrained maximization 956 Constrained Pareto optimal Radner equilibrium 71 I Constrained Pareto optimum 445. 4H7, 499, 710 Constrained social optimum 825 Constraint correspondence 964 Constraint functions 963 Constraint qualifteation 956 Constraint set 956 Consumer 17 Consumer choice: 17 Consumer demand 3 Consumer prererences 319 Consumption allocations 325 Consumption bundle 18. 20. 34 Consumpttan-invcstment 761 Consumplion set 18 Consumption stream 7 766 bounded 744 myopically. or short-run. utility ma).imizing 745 Consumption vector 18.555 Contestable market 411 Continsenl commodities 200. 567 market economy with 688 Contingent commodity vectors 689 Continuity axiom HI. 20) Continuous functions 602. 943 Continuous preference relation 80 I Continuous utility function 58 Continuously differentiable function 931.
,ood
n.
934 Continuum population 671 Contract curve 5ll. 654. 655 Contra,,-'1 design problem 477 Contraction 763 Conversencc: 943 Convex cone 134 Convel consumption sets 19 Convex function 565, 931. 955 ConVel hull 946 Convn objective function 503 Convn preferences 628 Convex ploduc:tion 134 Convex pro,ramming problem 484 Convex sets 22, 946 Conve). technoloJ,ic:s 583 Conve). utility possibility sets 560 Convex-valued average demand correspondence: 123
Convexification of aggregate nce:ss demand 629 Convexity 44. 133 assumptions 42, 43 On preferences 627 in second welfare theorem 557 property 608 Cooperative game theory. de!OCriptive side 679 Cooperative solution and dummy axiom 847 delinition 847 independent of utility origins and of common ch
01472 DA's Brother 238. 254 Date-event 691 DeadweIght loss from commodity talation 84. 332. 334 measures 86 of monopoly )85 tflangle 84. 332 Decision making. Set' Individual d~ision making Decision nodes 222. 227. 283 Decision rule 228 Decre:!sin, returns to scale )89 Demand functions 3, 23. 78. 397. 517. 530. 582.720 Demand law J2 Democracy 799 Depletable externalities 364. 365, 367. 375 alloc-.able 365 Derivatives. matrix notation far 926
Desirability assumptions 42 Determinacy condition 776 Determinacy propert~s of equilibrium 768 Deterministic social choice runctions 859 DiaSonal matrix 9)6 Dictator 799. 808, 811 Dictatorship 791. 807 Diminishing marginal rates of substitution 44 Direct revelation mechanism 868, 878 Directly revealed prc:ferred 604 Di5appointment 181 Discount factor 403. 417. 733. 743. 756. 759, 763.767 Discounted payoffs in a repeated game 421 Disequilibrium 625 Dispersion o( individual preferences 122 Distance function 623 Distortionary redistribution schemes 557 Distortionary taxation deadweight loss of 332 welfare effects of 331 Dislribution functions 123. 125 lotteries 183-4 over nonnegative amounts of money 184 Distribution of utility 827 Diversification 535 Diversification cone 54) Divinity. nolion or 472 Dominant diagonal 939 Dominant strategy 236. 911 strictly 237. 91 I weakly 238. 373. 911 Dominant strategy equilibrium 857. 870 Dominant strate,y implementation 869 Dominated strategy 236 iterated deletion 238 strictly 237 weakly 238 Domination-based refinements of beliefs 468 Dou ble a uction mechanism 918 Dual problem 967 Duality 63 Duality theorem 66. 68 of linear programming 967-8 Duality theorem argument 6H Dummy axiom 847 .. Dutcb books" 181 Dynamic games 267 Dynamic mechanisms 916 Dynamic model 400 Dynamic programming 752.753.968 Dynamic properties or equilibria 759 Economies basic model and defimtions 546 excbange 580. 613. 628. 644. 665, 669, 818 re,ular S91 uniqueness in 613 witb production 629 Edgeworth box 515. 5)). 550. 551. 587. 596. 629.631.634.654.667.692 Pareto optimality 5ll Walrasian equilibrium 519 Effective budget set 661, 664 Efficiency 149. 740. 742 Efficient production 149 across technologies 564
Efficient $Calc 144 Effort choice 479 Elfort level in principal-agent model 479. 491 Egalitarian contribution rule 861 Egalitarianism 680 Elastic demand 428 Elasticity of demand 27 Elasticity of substitution 644 between ,oads 97 Elementary activities 154 Ellsberg paradox 207 Endowment vector 516. 518. 525 Enforceable property rights 356 Engel aggregation 28 Engel function 24 Entrants. equilibrium number 334. 405 Entrepreneurs 475 Entry 334. 405 bias 408 blockaded 426 cost 410 in competitive markets 334 in oligopolistic settings 406 one-stage model 411 two-stage model 406. 411 with Bertrand competition 407 with Cournot competition 407 Entry deterrence 417. 423 Envelope: theorem 964 Envelope: theorem argument 69. 74 Equal-treatment allocation 656.657 Equal-treatment property 656 Equilibrium. St'~ Competitive equilibrium; Nash equilibrium; Perfect Bayesian equilibrium; Price equilibrium with transfers: Price: quasiequilibrium; Sequential equilibrium; Walrasian equilibrium Equilibrium domination 293.470 Equivalent variation (EV) 82 Euclidean preferences 805. 814 Euclidean space 674 Euler equations 749. 769. 970 Euler's formula 929 Events 690 Ex ante actions rrom induced preferences 182 Ex ante classical efficiency 900 Ex ante incentive efficiency 449, 899, 907. 909.910 Ex ante participation (or indiyidual rationality) constraints 893 b post (classical) efficiency 859,900 Ex post participation (or individual rationality) constraints 893 Excess demand 518 beess demand functions 580, 584, 592. 597. 602. 604. 610. 622 additive across 612 for labor S42 Excess demand vector 622 Excess supply 518 Exchange economy 580. 613. 628.644.665. 669.818 Existence problem 584 Existence with production externalities 650
973
978
INDEX
IN 0 E X
Probability premium 186 Problem the commons 380 Producer surplus 333 Product differentiation 389, 395, 4 t3
or
Product rule 927 Production 127 activity 154 (unction 129 plan 128 vector
128, 560
Production economy 582 Production externalities. existence with 650 Production runction 544, 650, 757
s" abo Production ~I Production inclusive cxcess demand correspondence 5M3 Proouction inclusive ClceSS demand (unction 582 Production path 738 bounded 744, 145. 765
efficient 739 reasible 744 myopically, or shan-run. profit maximizing 739 myopically profit ma:a:imiling with respect to a price sequence 740 short-run efficient 743 stationary 7S4 Production plans 128.731,739 trunCation 748 Production possibility set 533 Production sets 128.630,736
aggregate: 133 properties or 130 restricted 131 Production technology 736 Production transformation (unction. See Production set Productive input--output matrix 940 Profit functions 136. 142. 376 Profit maximization 135. 3\4. 548. 565. 572. 739. 742, 743. 758 Proper subgame 275 Proportionality assumption on initial endowments 609 Proportionally one-to-one 605 Public bads 360 Public goods 350, 568 definition 359 desirable 360 efficient supply of 364 equilibrium level 362 exclusion 360 inefficiency of private provision 361 nondepletablc 359 optimality condition 360 private provision of 367 pro... ision or 362 pure 360 Public project. 861,877 Public signals 252 Punishments 420 Pure consumption good 761 Pure exchange economy SIS. 546. 556.
860 Pure strategy 2]1 Nash equilibrium 251. 27] Purification theorems 257
Quanti'1y competition 389, 424 Quanti-ty dynamics 62S Quanti Iy titonncment 625 Quasico(lncave runctions 565, 933 Quasico
Radne rtquilibrium 696.697.699-701, 103-6.709.710.712 Ramsey taxation formula 820 Ramse::, taxation problem 347, 379 Ramsc::y-Solow model 737. 738, 741, 750, 153.756.757.761 Randc)11l variables 200, 6R8, 116 RlI.ndol1liz-I!d choices 231 Ratiomal expectations 4)9 equ ilibrium 724 equ Ilbrium price function 721 Ratio.,al preferenct: relation 14.203. 801. 809.832 Ratioalility 6 Ratiol1lalizablc: strategies 242 Ratio"tlalizing preference 13. 14. 16 deranition 14 Ratio ning rule 395 Rawl!Sl!n social optimum 828 Rawl!Sian social welfare function 841.853 Real·;.asset model 775 Relll· "Ilue function 753.937 Reasonable beliefs 292 Reasonable-beliefs refinements in signaling games 467 ReculTsive utility model 735 Redisiitribution limits 665 Ref\e livity 7 Regrw:t theory 180 Regrel valuation function 213 Reg~lar economies 591 Reg...larizing effects of aggrtgation 122 Repeated games 400. 417 Repealed interaction 400 Reprnentative consumer an d aggregate demand 116 mOillimization problem 399 Repruentative consumer model 399 Repr:-nentative producer 528 Rese:rvation utility constraint 490 Reservation utility level 480. 489. 500 Reservation wage function 443 Resc::rve price 923 Resolution of uncertainty 690 R~Lricted domains 799 Return matrix 70 I Returns 699 Returns to scale 132 Revecaled prc:£erc:nce proof 429 Rev-c:aled prc:£erence relation 11.608 Rev·(lation game 501 Revelation mechanism 373. 493. 868
Revelation principle 493. 857, 868, 869, 896 for Bayesian Nash equilibrium 884,901 for dominant strategies 871 Revenue equivalence theorem 858, 885. 889. 890 Risk and uncertainty 207 elementary increase: in 198 modeling 168 neutral 185 Risk aversion 167. 185.483. ti26 absolute 193 and lotterlc:s 183 asset allocation 188. 192 coefficient of relative 194 comparisons across individuals 191 constant rdative 194 decreasing absolute 193 decreasing relative 194 infinite 494, 499 insuranc:c 187 measurement of 190 more-risk-averse-than rdation 191 nonincreillsing relative 193. 194 nolion of un relilltive 193 strict 186 Risk neutrality 4112 Risky alternatives 16K outcomes of 16t1 Risky IISsets 18K. 194 Robinson Crusoe: economies 526 Roy's identity 73. 117. 118 Rubinstein's bargaining modcJ 298.1\54 Rybcszynski theorem 537. 543
Saddle point 623 SI. Petersburg- Menger paradox 185 Sales tax. comparative statIcs efTects 323 Scale of operation 154 Scitovsky contour 120 Screening 437. 460 pooling equilibria 462 separating equilibria 462 welfare properties of 466 Second-best frontier 824 Second-best market intervention 458 Second-best Pareto optima 824 Second·best policy problems 818 Second-best solutions 368 Second-best welfare economics 557 Second-best welfare issue 710 Second fundamental theorem of welfare economics 151. 308. 325. 327. 350. 436.524. 525. 528. 551. 557. 589. 659. 666.741.748.819 Second-highest valuation 864 Second-order stochastic dominance 167. 195, 197 Second-prict: sealed·bid auction 262. 866. 867.868.879.891.905 Selr-se:lection 437 constraints 495. 907 Self-selective. allOCilltions 666 Separating equilibrium in signaling model 453 in screening model 462
Separating hyperplane: theorem 64, 702, 947-8 Sequence 943 of budget scts 696 of initial endowments 743 Sequential equilibrium 290 weak 283 Scq~nti.1 r»t;nnillity and bel ids 2112 of strategies 283 pnnciple of 2611 $cquential trade 694 Serial millimum deciSion rule 828 Shadow price 563 Shapley allocation 673 Sh,tpley value 6U. 674, 679. 6KI. 847 basic properties 682 Shepard's lemma 141,531 Short-run compuative statics 339-40 Short-run cosl function 146-7 Short-run equilibrium 778 Short-run law of demand in overlapt'ling ~enerations models 779 Shurt-run prolit maximir.ation 748 Short·term production technology 743 Signal functions 717 Signaling 437, 450 .... clfare effects 45 I Signaling games. reasonable-belids refinements in 467 Slmplcx 232 Simultaneous-move games 218. 231. 235. 636 Smgle·consumer economy. Pareto optimality 528 Sing1c:-consumer problem 718 Singlc.crossing property 453. 490. 873 Singlc-good Cournot competition 661 Single-input. increasing returns production function 679 Single-market partial equilibrium analysis 342 Single-output ca.sc: 143 Sing\c·peaked prderences 789, 801 Single·peakedness 799. 803 Single unit of indivisible private good 862,
877 Single· ... alued correspondence 950 Singleton information sets 224 Slutsky compensated price changes 30 Slutsky equation 71.72, 114 Slutsky matrix 34.35.72. 80, 92. 95, 114. 115 Slutsky wealth compensation 30. 73 Smith. Adam 327. 514. 549 Social choice function 807. 808. 859. 863.
864 Bayesian incentive compatible 883. 888. 890.891.895.905.907 dictatorial 808. 810. 813. 876. 914 dominant strategy implementable 810. 913 ex ante elassically efficient 900 ex ante incentive efficient 909. 910 ex ante individually rational 893 ex post (classicallyl effident 861.862.875. 900 ex post individually rational 893 feasible 906 implementable 868
implementation in dominant strategies 810 in Bayesian Nash equilibrium 883 incentive efficient 901 individually rational 80S interim incentive efficient 898. 899 interim individually rational 893, 895 monotonic 807. 809. 873. 875, 914 Paretian 807. 809, 811 strongly implemented in Nash equilihrium 914 truthrully implementable in Bayesian Nash equilibrium 883 truthrully implementable in dominant strategies 871. 873. 877 truthrully implementable (or incentive compatible) 868 welfare analysis 900 Soc:ial choict theory 789 Social optima 825 Social prd'erenc:es 789. 832 over two alternatives 790 relation 833, 834 transiti\'lty of 799 Social rationality. less than full 799 SOCIal utility function. Paretian 851 Social utilny value 832 Social welfare 117.408 aggregator 790. 793 measure of 333 Social welfare functional 190. 80) dictatorial 791 generated from continuous and increasing social wetrart function 83S independent of irrelevant individuals 838 invariant to common cardinal transformations 834 invariant to common chanJl!s or origin or of units 834 IOvariant to common ordinal transformations 838 invariant to independent changes of origin or of units 836 lexically dictatorial 852 majority voting 791. 792 neutral between alternatives 791 on ,iven subset 793 Paretian 794.810,832. 833 positively responsive 791 symmetric among agents 791 Social welrare functions (SWF) 117. 120, 328. 329. 448. 559. 825. 859 Bergson-Samuelson 117 continuous and increasing 835 generalized utilitarian 828 increasing. continuous 837 invariant properties 831 maximin or Rawlsian type 827. 828 neutral 827 purely utilitarian 827 utilitarian 119 Social wei face optima and Pareto optimality 558 Sonnenschein-Mantel-Dc:breu theorem 598, 761 Spanning 714 through optiom 704
Spatial models of product differentiation
399 Spot economy 717 Spot markets 694. 695, 698 Spot prices 717 Stability condition 414. 472 Stackleberg leadership model 426. 433 Stage game 417 Nash equilibrium 410 (State·)contmgent commodity 688 (State·)contingent commodity vector 688 (State-)contlOgent production plan 689 State-dependent preferences 200. 202 State-dependcnt utility 199 insurance 20 I State-independent upected utility representation 201 State of the market 724 State of the world and uncertainty 688 States of nature 168. 199,200 representations of uncertainty 200 Static models of oligopoly )87 Stationarit~· 734 Stationary cquilibrium paths 755 Stationary paths 754 myopically supportt'd by prorortional prict:s 755 Stationary strategics K54 Stt'ady statl·... 754.756.758.760. 7M! Stolper Samuelson theorem 536. 543 Strategic complements 4 t 5 Strategic effects from investment in marginal cost reduction 415 Strategic entry accommodation 426 Strategic entry deterrence 423 Strategic interdependence 217. 219 Strategic precommitments to affect future competition 414 Strategic substitutes 415 Strategies 228 behavior 133 completely mixed 284 dominant. Su Dominant strategy dominated. See Dominated strategy mixed 232. 233, 240. 250 pure 231, 250 sequential rationality of 283 totally mixed 259 Strategy-beliefs pair 285 Strategy profile 229. 638. 870 Strict concavity 186 Slrict conve~ity 801. 813, 931 Strict preferences 6. 793. 832 Strictly conca\·e function 930 Strictly quasiconcave functions 933 Strike price 700. 706 Strong axiom or revealed preference (SA) 91. 115.604 Strong compensation test 830 Strong gross substitute property (SGS) 647 Strongly monotone consumer prererences 581 Subgame, definition 275 Subgame perfect implementation 915 Subgame perfect Nash equilibrium (SPNE) 268.27.3-82.295-9,401-7,412. 419-24.434,443-4.450.461,465.
854
979
980
INDEX
INDEX
Subjective upec1ed utility theorem 206 Subjective probability 168 theory lOS Subsequence 945 Substitutes 71 Substitution etrecls 34. 600 Substitution matrix 34. 599, 600
Sufficient statistic 488. 722 Sunk costs Ill. 146 Sunspot equilibrium 109-10 Sunspot free equilibria 710 Sunspot set 709 Sunspots 709 Sup:raddilivity 676 Supply corrcspondcnu: 136 Supply substitution matrix 139 Support 'unction 64, 949 Supporting hyperplane theorem 948-9 Sure-thinl axiom 204 Symmetric auction 5ening 891 Symmetric bidders 90S Symmetric equilibrium 398.615 Symmetric informalion 691. 716. 717 Symmetric joint monopoly point )92 Symmetric malrix 936. 938 Symmetric profit·miuimizing equilibrium (ocOl1404 System or beliefs 283 System lilability 623 Tangency point 497 Tarsky's filled point theorem 953 Tatonncment dynamics 624, 724, 779 Tatonncment price dynamics 648 Tlitonncment Slability 620. 779 Tax-and-transfer scheme 331, 334 Tax incidence 538 Tax rate 814 Tax revenue 333 Tues 3~. 369, 316 equilibrium with 643 see "bo Distorlionary taxation; Sales tu Taylor's expansion 933. 955 Taylor's rormula 658 Temporal modeling 691 Temporary equilibrium 780 Terminal nodes 222. 221 Theory of nonunique prior beliefs 214 Threat point 839 Threshold price 401 Ttck·Tack·Toe 220, 229 extensive (arm 222 Time aggreg;uion 18 and equilibrium 732 at which a physical commodity is available 691 impatience 133 low discount of 762 of delivery 891 Totally (or completely) mixed strategy 259 Trade-off between returns and risk 197 Tr;&ding equilibrium 660, 665 Trading posts 663, 665 Tr;&ding rule 660, 664 Trans(er paradox 542 Trctnsfcrable utility (TU) games 676. 678 in chanu:terlstic form 676
Transfers. equilibrium with 524 Transformation frontier 128 Transformation (unction 128 Transilivity 6, 9, 13,42.803,810 Transitory shocks 764 Transversalily condition 740. 741, 744. 745. 747. 755. 756 Transversality theorem 595. 596. 942 Trembling·hand penecl Nash equilibria extensive form 292, 299 normal (orm 258 Truthful implementation 868 Truthful revelation mechanisms 493 Truth·leUing 374. 449, 861, 880 constraints 495-7. 907 Turnpike Iheorem 762.763 2 )( 2 production model 529 Pareto sct of the 533 Two-commodity economtes 622 Two·good. two-state. two-consumer pure exchange economy 697 Two-$C(tor model 738 Two-slage entry model4l1 Type allocation 657
Unanimity g;&mc: 686. 855 Uncertainty and risk 201 and state of the world 688 firm behavior in general equilibrium under 713 general equilibrium under 687 resolution of 690 state-o(·nature representations of 200 see "Iso Choice. under uncertainty Uncompensated law or demand (ULD) III Unconstrained mnimiution 954 Under.
Value equivalence theorem 673 Value function 752, 7:53, 764.964. 968 Variable cost (unction 145 Vectors 926 Vetoers 800 Vickrey auction 262. 866 Voluntllry mechanism 891 Voluntary parttcipation 869 von Neumann-Morgenstern expected utility function 173.180, 184,232.241.716. 718 Walking in Downtown Manhattan game 249 Walnls' law 23, 27, 28, 30-2. S2, 54. 59, 75. SO. 87. 109. 582. 585. 589. 599.601. 602. 604. 7SO Walras price dynamics 648 Walrasian allocations 6S9, 665 Walrasian budget sets 21. 22, 24, 661, 665 Walrasian demand and indirect utility function 73 continuity 92 correspondence 23. 5 I. 62. 93 differentiability 94 (unction 27-30. 34. 35, 39, 51. 56, 71, 10. 89-91.116.117.319. B4. 353. 555 Walrasian dynamiQ 625 Walrasian equilibrium 525, 527, 557. 558. 579. 5SO. 582. 589.631.662.744.747. 766. 771. 773 allocation 524. 528. 5SO. 654. 657, 659. 660.672. 673.766 and basic welf.re properties 545 characlerWng through welfare equations
630 computatton of 749 definition S47 determinacy properlies of 768 Edgeworth box 519 existence of 584 factor price ratio 536 general approach to existence of 632 model 578 multiplicity of 715 noncooperative roundations of 660 onc~nsumer case: 743 price 322. 717 price vector 520. 580. 581, 589. 600. 607. 616 locally untqut 590. 592 regular 591 problem. onc<ensumer 767 productions 663 several consume" 765 specialized 534 static properties 616 trajectory 759 uniqueness of 606. 749 weights 768 welfare properttes of 522 with taxC$ 643 with transCcn 52. see Competitive equilibrium Walrasian model of perfect compc=tition S90 Walrasian price vector 659 Walruian quasiequilibrium 632. 639 Walrasian theory of markets 511 Walrasian wealth levels 557
"'SO
We;&k axiom of revealed preference (WA) 5. 10. 28. II 5. 607 and aggregate demand 109 and gross substitution 613 for liggregate excess demand 579, 607 Weak compensation test 830 Weak gross substitution 612 Weak order 6 Weak per(ect Bayesi'lR equilibrium 283, 290. 452 ~trengthenings of 288 Weak preference reversal property 872 Weak sequential equilibrium 283 Weakly dominant strategy 238. 373, 911 Weally dominated strate~y 238
Weakly monotone preference relation 96 Wealth distribution rule 108 Wealth effects 24, 27, 343 Wealth expansion path 24 Wealth levels 516, 555, 557 Wealth redistribution, lump-sum 521. 524 Welfare analysis 80 in partial equilibrium model 328-34 of social choice functions 900 with partial information 87 Welfare economics 149,787 elements 0(817 fundamental theorems 0(, SI'I' First fundamental theorem of welfare
economics; Second fundamental theorem of welfare economics Welfare etrects of dlstortionary tax: 331 Welfare equations. characterizing equilibrium through 630 Welfare loss 385, 387 Welfare-maximizing property of competitive allocations 531 Welfare measure 371 aggregate surplus as 334 Widget market 396. 429 Wilson equilibrium 466 Zame!o's theorem 272 Zero-sum gam('s 221
981
974
INDEX
Expectations 695. 780 self·fulfilled or rational 696 Expected benefits 886
Expected externality mechanism 885-7, 911 hpttted utility as guide to introspection \78
,.
form 167. 173. 174 (rilmcwork 183 funcllon 174
representation 703 theorem 167. 175. 176. 182
theory 168 .~t't· also Extended cxpeclc=d utility Expected utility function 719
EXpt'nditurc funClion 59, 63 and demand 78 and Hicbian demand 68 and preferences 76 Expenditure minimization 57, 55)' 555
Extended ckpectcd utility representation 200, 202
theorem 203 Extended independence all.iom 203
Extensive form representation 221, 233 Extensive form trembling-hand perfect Nash equilibrium 292. 299 External effects 350, 352 EXlernal properttes 768
Externality 350 and missing markets 358
IN 0 E X
Firms 3 equilibrium number of 407. 408 obje<:tives of 152. 713 First-best nonconvexities 822 First-best Pareto frOntier 823 First-best Pareto optima 824 First-best policy problems 818. 819 First fundamentallheorem of welfare economics ISO. 308. 325, 326. 330. 436.124.131.149.611.614.619.739. 741.747.711 First-order conditions ar,ument 68. 74 First-order conditions for Pareto optimality 161 First-order stochastic dominance 167. 195 First-pri~ sealed-bid auction 865, 867.868. 891.901 Fixed point 952 Fixed-point argument 639 Fir.ed-point theorems 952 Brouwer's 589, 952 Kakutani's 261, 585-6. 638. 953 Tarsky's 953 Focal points in ,ames 248 Folk theorem 404. 417. 422 Nash reversion 418. 420 Forward induction 293 Forward markets 694 Framing problem 7 Fr«-disposal condition 131.633 Free-disposal quasicquilibrium 635 Free entry 133. 334 Fr«-rider problem 351. 362. 367 Frobenius' theorem 80 Full implementation 915 Fundamental theorems of welfare economics_ ~~ First fundamental theorem of welfare econom~ Second fundamental theorem of welfare economics Fundamental value 771 Future competition. strategic precommitments to affect 414
as source of multiple local social optima 376 bargaining over 356 consumption 366 definition 351 equilibrium price ]59 generation ]66. 373 generators of 365. 372 marginal costs ]73 measurement of lis level 358. 359 optimal level 373 permits 367 property rights over 376 traditional solutions 354 two
Gains from trade 907 Game bar. 230 Game-theoretic: model of adverse selection
Factor allocations 5]1, 532-3 Pareto set of 533 Factor demands 531 Factor endowments 530 Factor intensity 534. 536 Factor markets 529 Factor price 530. 532. 534. 538. 543 equalization theorem 535 Family size 123 Feasibility 755. 773 Feasible aggregate productions 574 Feasible allocations 312. 635, 653. 656. 657. 666 technical properties of set of 513 Feasible individual consumptions 574 Feasibility requirement 744 Finite games of perfect information 270 Finite horizon bargaining 296 Finitely repeated game 401
Game theory 211 see also Cooperative game theory; Noncooperative game theory Game tr« 221 Games concept of 219 cooperative: theory of 673 dynamic 267 er.tensive form representation 221, 233 finite 228 in Characteristic form 674-5. 846 infinitely repeated 281. 401. 417 normal form representation 228 of complete information 253 of imperfc:c:t information 226 of incomplete information 253 of perfect information 226 outcomes 220 payoffs 220 perturbed 258
443
players 219 rules 219 simultaneous-move 231 zero-sum 221 General equilibrium 661 principal features 511 under uncertainty 687 General equilibrium theory 511. 567 eumples 515 modern classics 5 I 3 versus partial equilibrium theory 538 Generalized medians 1106 Generalil.c:d partition Ml5 Generations. overlappmg 769 Generic determinacy of equilibria 775 Genericity analysis 593 Geometry of cost 143 Gibbard-Satterthwaite theorem 858. 860. 873.874 Giffen goods 26. 29 Global constrained mar.imizer 956. 962 Global convergence 763 Global max.imiler 954 Gloool stability 623 Glove market 681. 683 Golden rule path 755 Golden rule steady state 756. 785 Golden rules 754 Gorman form 119. 120 Gross complements 71 Grolls substitute er.cess demand function 776 Gross substitute sign pattern 9)9 Gross substitution 71, 579,611.618 and weak uiom 613 Groves (-Clarke) mechanism 374. 381. 858. 876. 878. 880. 882.921 Hair-spaces 64, 947 Heckseher-Ohlin theorem 544 Herfmdahl indel of comxntration 43 I Hessian matrir. 79.927,932.938 Hicksian composite commodities 99 Hicksian demand 62. 63 and expenditure function 68 Hicksian demand correspondence (or function) 60, 62. 11.90 Hicksian demand cUrve 71, 83. S4 Hicksian demand function 71,90.343 Hicksian wealth compensation 62 Hidden action principal-agent model 478.
SOl multiple etTort levels in 502 Hidden information principal-agent model 489.101.900 solution using Kuhn-Tucker conditions
SOl with continuous types 900 Homogeneity of degree lero 27. 28. 34. 35 Homogeneous functions 928 Homogeneous-good markets 413 Homothetic prererences 45. 609 Homotheticity 611 Homotopy 597 Hotelling's lemma 138.141 Household production te<:hnology 735 Hybrid hidden action/hidden inrormation principal-agent models 501 Hyperplane 64, 805. 843
Imperfe<:t information 716 Implementation 857 in Bayesian Nash equilibrium 883. 894, 897 in dominant strategies 869 in environments with complete information 912 in Nash equilibrium 913 using extensive form games 915 Implicit function theorem 95. S94. 598, 882, 940
In(.'Cntive compatibility constraints 494, 897 Incentive constrainu 48), 503 Incentive efficiency er. anle 449. 899. 907. 909. 910 mterim 449. 899. 907. 909. 910 Incentive feasible set 898 Incentive scheme "8. 486 Income effects. See Wealth effects Incomplete informatton 253. 859. 867. 869. 883 Incomplete markets 709 Indecomposability condition 650 Inde<:omposable cconomtc:s 633 Independence ar.tom 167. 171 violations of 181 Independent penon-b)'-pcrson monotonicity (IPM) 919 Independently distributed silnals 252 Indeterminateness and nonuniquenns 776 lnder. analysis and uniqueness 615 Index formula 603 Inde .. of a"regale output 400 Inde .. of regular equilibrium 592 Indel theorem 593. 597 Indifference map 764. 765 Indifference relation 6 Indifference set 43 Indirect demllnd function 100 Indirect utility functton 56 and Walrasian de:mand 73 Individual choice behavior. modeling approaches 5 Individual consumers 3 Individual decision making choice-based approach 3 preference· based approach 3 Individulli preferences 789 distribution of 629 Individual spannina661 Individually rational me<:hanism 894 Individually rational payoffs 421 Indivisible private lood. allocation of single unit 862 Induced preferences 181.643 rrom e..-ante actions 182 Induced socilll choice function 816 Industry marginal cost curve 330 InduSlr)' marginal cost function 321. 330 l!1dustr)' suppl)' (unction 321 Inequality condition•• verston to 826 lnequalit)' constraints 958 Inequality problem 961 Infimum 65 Infinite horizon bargainins 298 Infinitely repeated Bertrand duopoly game 401
Infinitely repeated Coumot game 422
Infinitely repeated 'Imes 401. 411. 419. 421 Information and resolution of una:rtainly 690 Information partitions 716. 717 Information rents 903 Inrormation sets 224, 228, 274. 283 Information signal function 717. 723 Inrormatton structures 690 family of 690 Informational asymmetry 309. 351. 368, 436. 437.477.716.719.817 I nitial decision node 222 Input/output 128 se~ olso leonticf input-output model Insurance: risk aversion 181 state-dependenl Ulility lOt Insurance contract 510 Insurance markets 460, 476 Intesrability 75. 78 Interest rates 754 Interim elpeeled utility 893 Interim incentive efficiency 449. 898. 907. 909.910 Interim participation (or individual rationality) constraints 893. 900 Interior equilibrium 533 Intermediate value theorem 953 Inlernal properties 768 Interpersonal comparabilit)' of utility 825 Intertemporal production and efficienc), 736 Intertemporal utilit), 733 Intransitive decisions 8 Intuitive criterion 458. 470 Invariance propert), 853 Inverse demand function 321. 330. 385. 430.
661 Inverse function theorem 592. 942 Inverse supply runction 321 Investment in marginal cost reduction, strategic effects from 415 Invisible hand properly 524. 549 Irrefler.ivity 7 Irreversibility 132 Isoprotit curves 491. 497 hoprofit line 136 isoprotit locus 392 lsoquant 140 Jacobian matrix 615. 941 Jensen's inequalit), 185,931 Joint protit 391 Joint profit-maximizing problem 531 Just perceptible differences 7 Kakutanj's fixed-point theorem 261. 585-6. 638.913 Kalai-Smorodinsky solution 844 Kaldor compensation test 852 Kuhn-Tucker conditions 53. 189. 398.495. SOS. 163. 9S9 Kuhn-Tucker theorem 959 lagrange multiplier 53. 54. 14. 703, 957. 958. 963.961.967 lagran,ian function 958 Large economies 627 Laspeyres quantity index 37
Law of demand compensated 32. 62 uncompensated III law of supply 138, 147 Learning dynamics 780 Lebesgue measure 596 Lemons market 724 Lemons model 437 Length or period 735 Leonticf input-output matrix 156 Leonticf input-output model 155 with no substitution possibilities 155 with substitution possibilities 157 Leonttel' preferences 49, 91 S Lerner index. of monopoly power 42K l.c:ve1 set 928 Lex.ical dktatorship 812 Ler.ical muimum decision rule 827 Lexicographic ordering 85 I Ler.icographic preference relations 46. 47 Lcximin soc:::ial welfare ordering of utility vectors 827 Limit points 945 Lindahl equilibrium 363. 367. 568 Linear activit)' model 154 Linear city model or product differentiation 396 Linear compensation schemes 488 Linear constraints 50) Linear er.penditure system 98 Linear function in probabilities 173 Linear order 793. SOl. 802. 804 Linear programming 966 Linear social welfare function 560. 561. 565 Linear te<:hnology 753 local constrained m31imizc:r 956 local constrained minimizer 958 Local maximizer 954 Local nOMiliation condition 42. S54 local nonsaliation of preferences 549 Local uniqueness 589 Locall)' cheaper consumption condition 555 Locally determinate theory 589 Locally isolated steady state equilibrium determinate: 716 indeterminate 776 Locally nonsatiated consumers' preferences 112 Locally stable equilibrium 622 Logarilhmic concavity 806 Long-run aggregate supply correspondence: 331 Long-run comparative: statics 339-40 Long-run competitive equilibrium 334 nonelistencc 336 Long-run cost runction 146-7 Lotteries 167,859 and risk aversion 183 compound 169. 183 concept of 168 distribution functions 183-4 preference for 170, t 82 reduced 169 simple 168 simplex diagram \79 space 183 stale 206 with monetary payotTs 183, 194
975
INDEX
1N 0 E X
Lower contour sets 43, 872 LoW1:r hemic::ontinuous. correspondences 951 lump·sum taxation 557 lum~sum transfen 331 Lum~sum wealth redistribution 524, 551 lyapunov runctions 624 Machina's paradox 180 Majority voting 67S, 676. 791 Malinvllud prices 739, 741, 743 Manllger-owner relationship. Srr Principal-agent problem Marginal contribution 681. 682 Marginal cost 140, 143,318,385 Mluginal cost price equilibrium with transrers 571 Marginal cost pricing 566. 570 M.v.rJirud cost reductton, strategic effects (rom investment in 415 Marginal externality 355 Marginlll productivity 440, 670, 756 Marginal rate or substitution (M RS) 54,
S64 Marginal ratc of technicalsubititution (M RTS) 129-30 MarginaJ ratc or transrormation (MRT) 129, 743 Marginal revenue 3a5 Marlinal utility or wealth SS Marcinal value 563 Market aeent 6)5 Market-based solutions to externalities 356. 358. 365. 367 Mluket clearing 314, 316 Market-dearing condition 363 Market-clcaring prices 661 Market completeness 5SO Market demand correspondence. S~~ Walrasian demand correspondence Markel demand function. S~~ W.lrasian dem.nd runction Markel economy 17,307 with continlent commodities 688 Market equilibrium 307 Market railure 307, 350 Market intervention 362, 445 Market power 383 Market size 412 Market unraveling and adverse: selection 440 Market value 714.715 Marshallian awepte surplus 326, 328. 339. 385. 408. 429 Marshallian consumer surplus 83, 89 Marshallian demand runction 51 Marshallian dynamics 625 Marshallian partial equilibrium analy,is 316, 341.648 Martingale property or asset prices 708 Matchinl Pennies 220. 253 eJ.lensivc rorm 225 Nash equilibrium or 250 Version B 221 normal ronn 130 SlrateBies in 229 Version C 223 nonnal ronn 230 strategies in 229 Version 0 226
Matrix notation ror derivatives 926 Maximization problem 964 Maximizer amcspondence 964 May's theorem 792 Mean-prescrvinlspreads 197 Mechanism definition 866 desiln problem 857. 858 implementing 867 Median voter 789,802,814 Meeting in New York game 221. 247, 248. 211 Minimax payoff 421 Minimax theorem 264 Mixed strate&ies 232, 233, 240, 250. 302 Modi6ed lolden rule 755. 763 steady st.te 762 Monetary payoff 205 distributions in terms or return and risk
194 Monetary steady state 774 Money cert.inty line 201 Monopolist 383, 384 Monopolistic: competition model 400 Monopolistic price discrimination SOO Monopolistic: saeeninl 488 Monopoly pricinl )84 Monopoly quantity distortion 386 Monopsonistic screenin, Monolone likelihood ratio property 485 Monolone pr.rcccOCC$ 42, 47. 5S6. 703 Monolonicily 790. 808, 810. 845, 854. 874, 915.919 Moral hazard 477. 478 su GUo Hidden aaions Multidimensional ctrort 40 Multilateral eXlemalittcs 3SI, 363. 364 Multip~ effort levels in bidden action model
soo
S02 Multiple Walrasian (or competitive) equilibria 442, 457.519.589.606, 775. 910 Multi ....lued social choice correspondence 915 Myerson-Satterthwaite theorem 369, 858,
894.900 Myopic profit maximization 739 Myopic utility maximization 746 Nash equilibrium 23~, 24S, 246. 361, 380. 381.388-90.392, 398. 414. 430. 63l and r.tionalizable stflte,ies 247 as consequence or rational inrerence 248 as necessary condition ror a unique predicted outcome 248 as selr-enrorcinl .Ireement 249 as stable soci.1 convention 249 concept or 246 definition 246 dilCussion of concept 248 existence 252. 260, 639 eltensive rorm Irembling·hand perfect 292, 299 implementation in 913 in Matching Pennies 250 in Meetinl in New York lame 247 miled str.tegy 250 (norm.1 ronn) trembling-hand perfect 258
or
proper 260 pure strategy 253, 273 SlratelY profile 247 subpme perfect. S~~ Subgame perrect Nash equilibrium (SPNE) Nash reversion 420, 423 definition 418 folk theorem 418. 420 strateg~ 401. 405 Natural monopoly 570 Negati ...e definite m.trix 35,618.619,932. 935 Negative dominant diaSonal 646. 647 Neptive externality 352, 354 IS source or nonconvexit~ 375 Negative semidefinite matrix 34. 35. 70, 618. 932, 93l Net demander 518, 519 Net supplier 518, 519 Netput vector 128 Newton'S method 624 Niche chotec game 278 No-incentive·to--mi5fepresent condition 811 No·information equilibrium 719 Nonconstant returns to scale 394 Nonconvex production technologies ~70 NonconvCllilics 31•• 627 Noncooperative equilibrium. concept oC 660 Noncooperative roundations or Walrasian equilibria 660 Noncooper.tive pme theory 217 basic:: elements 219 Nondccrcasina rearranacment 85 I Nondecrcasin, returns to scale 132 Nondeplelable extern.lity 365. 366, 367. 380 Nondur.bles 136 Nonenvy .llocation 666 Nonequilibrium dynamics 778 Nonincreuin, returN 10 scale 132 Nonlocal shocks 619 Nonmonetary slcady Slate 773 Nonnegative amounts of money, distribution functions over 184 Nonneptive respoNivencss 919 Nonoptimality or competitive outcome 353 Nonpaternalism 82S Nonrational decisions 8 Nonsatiation usumption 5SO Nonsubstitution theorem 159,342 Nontightncsl 743 Nonuniquc prior belids. theory or 214 Nonuniqueness and indeterminateness 776 Normal commodity 25 Normal demand 25 Normal rorm representation 229, 258. 295 Normal vector 64 Normalization 24 Normalive representative consumer 118 Na.tradc equilibrium 773 No-trade stationary Itale 173 No veto power 915 N·replica economy 655 N-sector model 738 Nucleolus solution 8SS Numeraire 325. 334, ).42, 511. 521, 713. 821. 862 Numc:raire commodity 45. 311, 316 Numeraire rca.llocation 325. 328
Objective runction 366, 369, 484. 964 Occupational choice 568 Oiler curve 25, 517 Oligarchy 800 Oligopoly definition 383. 387 repeated game models or 400 static models or 387 One,onsumer case, equilibrium 743, 767 One-consumer, one-producer economy 525 One-period utility function 762 One-stage entry game 410 One·sta~e entry model with Bertrand competition 410 Opcn ball around x 944 Open sets 943~4 Optimal aggregate production levels 325, 56l Orumal allocation of a ....ailable goods across consumers 564 Optimal auctions 903 Optimal Raycsian mechanisms 897 Optimal consumption level 326 Optimal contracting with hidden actions (moral hazard) 47R~~R
with hidden information 4M8 501 Optimal productIOn level 326 Optimal ~hadow prices 327 Optimality of allocation or production aero!» firms 33.1 Oplimi/.atlun problems 625, 963 Options 700 pricing by 706 spanning through 704 Ordinal prererence rl:lation 858 Ordinal properttes 9, SO Ordinary demand correspondence. Se~ Walrasian demand correspondenct Ordinary demand runction. Su Walrasian demand runction Outcome or risky alternatives 168 Ovenlil consumptlon vector 639 Overall profit maximization 783 Q\'erddermination 594 O\'crlapping ,enerations 769 Own rale or interest 7X3 Owner~managcr relationship, See I)rlnclpal~altcnt problem Ownership shares 689 Paaschc quantity index 37 Pairwise independence condilion 789, 794, 795. 798.810. 832.8J3 Pairwise majority voting 789. 802, 803 Parellan property 825 Paretian social welfare functionals 794 Pareto domination 460. 549. 550. 655, 773 Pareto effidency. Sf''' Pareto optimality Pareto rrontier 313, 326. 559, 818. 825, 828 Pareto improvement 381.458.466 Pareto inefficiency 351, 499 and asymmetric inrormation 440 Radner equilibrium 710 Pareto optimality 307. 312. 328. 339. 350, 353.366.368. 17I. 378. 436. 438. 451. 522.547. lll. 6l7. 668. 704. 747. 773. '97
allocation 313, 325, 328, 438, 467. 523-5, 547. 5ll. 554. l56. 562. l64. 630. 66l. 709 and Arrow-Dc:breu equilibrium 692 and social welrare optima 558 conditions for 327. 360 definition 313 Edgeworth box 523 first·order conditions ror 561 of competitive equilibrium 440. 768, 775 problem 327 single-consumer economy 528 uniqueness as implication of 614 Pareto optimum allocation. Su Pareto optimality ParelO ordered equillbna 712 Pareto rankinl 443. 455 Pareto sct 523, 537 Pareto set or ractor allocations 533 Pareto sct of 2 )( 2 production model 53) Partial cooperation 846 Partial equilibrium model 308, 316. 341. 358 welfare analysis in 328-34 Partial equilibrium theory versus general equilibrium theory 538 Participation (or individual rationality) constraints 858. 869. 887. 891 in publie project choice 892 Path·invariance propt:rty 16 Pecuniary eKternality 352 Perfect Bayesi.an equilibrium (POE) 283. 290. 452-l. 468. 471-2 wellk 283. 290 Perrect information, finite games or 270 Perrect recall 224 Permanent income 778 Permanent shocks 764 Perturbed game 258 Physical commodity vectors 689 Pigouvian taxation 355, 379 Pi ...otal mechanism 875 Planninl problem 748-SO. 767, 768 two-step 767 Pleasure runction 57S Policy function 752, 760 Pooled information equilibrium price runction 721. 723 Pooling equilibria in signaling model 453, 456 Pareto-dominated 457 in screening model 462. 463 Portfolio 700 Positive association 114 between endowments and demands 611 Positive definite matriK 932. 9)5 Positive externality 352. 355 as source or increasing returns 375 Positive representative consumer 116 Positive semidefinite matrix 932, 935 POSitive theory or equilibrium 578 Post--cntry oligopoly lame 409 Potential compensation test 120, 126 Potential Pareto improvement 334 Predation game 268 PrcCerence maximization 552. 554 Preference parameters 912 Prderence relations 5, 6. 171. 175, 177.201. 2Q3. l49. 800. 809. 811
977
basic properties 41 continuous 46 continuous, strictly con ...ex. and strongly monotone 581 convex 44 homothetic 45 lel(icographic 46, 47 local nonsatiated 42 quasihnear 45 rational 42 relationship with choice rules II itrictly convex 44 Preference reversal property 872 Preferences 5. 46 and expenditure function 76 convexity of 49 for lolleries 182 Pricc:·clearing runction 661 Price-cost margin 428 Price competition, Bertrand model or 388 Price dynamics 625 Price effects 25. 27 Price equilibrium with transfers 548, 549. 5ll. 552. l54. l56. l66 Price function 719. 723 Price· independent rule 119 Price insensitivity 10 own actions 661 Prlcc·mediated competition models 660 Price quasiequilibrium supporting allocation 555 with transrers 551. 552. 554. 556 Price quoting 715 Price sequence 743, 747. 754. 755. 767 bounded 745 Price system 748 Price-taking assumption 20. 3 I 5. 5S7 Price-tak.ing behavior 330. 548. 5se, 715 Price-taking environment 660 Price titonnement 620 Price vector 314, SS5. 739 Price-wage sequence 756 Price-wealth change 110 Pricing by arbitrage 706 oC options 706 Principal-agent model 477-510 effort level in 479. 491 hidden actions 478-88 hidden information 4RR -501. 900-3 hybrid SOl nonobservable managenal effort 482 observable managerial effort 481 obscn'able state variable 0 490 state 0 observed only by manager 492 unobservable managerial effort and risk-neutral manager 482 Pnnciple of completeness (or universality) of markets 20 Principle of sequential rationality 269 Prisoner's Dilemma 236 Pri ...ate goods 569 Private information. See Asymmetric inrormation Private ownership economy 547, 578. 579 Private signals 252 Private values 859, 905 Probability,linear runction in 173 Probability distribution 284
.II