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!
o o . I.e. assuming l i m H!£i*) = 1 ,
Vt6[0,T]
then at x = MAx = XM, we can impose a boundary condition as follows: u^1
=
a and a < x < <j)b, Y(x) = f(x), the string coincides with the obstacle, the obstacle exerts on the string a reaction force / > 0, thus -Y" = / > 0. Combining (1)—(4), the problem is to find y = Y(x) G C"^], such that J min{-y",Y - tp(x)} = 0, 0, Jn 0,
(0 < x < 1)
\ y(0) = F(l) = 0.
(6.1.19) (6.1.20)
Using the principle of the minimal potential energy, the above equilibrium problem can also be formulated as the following variation problem: find Y(x) € Q, such that J(Y)= min J{y), nO)ef2
where J(y) is the stress energy of the string: J
(y) = \ti(i?dx,
and il = {y(x)\y(x) G ClOil],y(0) = vW = 0-3/W >
120
Mathematical Modeling and Methods of Option Pricing
tinuation region (the part not contacting the obstacle) called separated set, the stopping region called coincident set, and the optimal exercise boundary called the boundary of the coincident set, i.e. the free boundary. (6.1.13)—(6.1.19) and (6.1.17),(6.1.18) are two mathematical formulations of the perpetual American put option pricing. In the former form the free boundary appears explicitly, and in the latter form, the free boundary is implicit. Each form has its own merit. One chooses a form according to the problem under study. Remark Consider the dividend-paying case. If the dividend rate q is timeindependent, then the mathematical model for the perpetual American put option is: find {V(S),S 0 }, such that
£S2fdS?
+
(r~^S!c[5~rV==0'
(SQ<S<°°)
V(S0) = K - So,
(6.1.21) (6.1.22)
V'(So) = - 1 ,
(6.1.23)
V(oo) = 0.
(6.1.24)
Let V = Sa and substitute it into (6.1.21) to obtain a's characteristic equation: 2
—a
+{r-q-
2
—)a-r
= 0.
It has two roots a = a±=oj±Ju2 + ^ , V o"
(6.1.25)
where
-r + g+sl U =
^
•
It is easy to check a- < 0 < a+. Hence the general solution of the equation is V(S) = ASa+ + BSa~. Applying the boundary condition (6.1.24) to determine A = Q. Applying the boundary conditions (6.1.22), (6.1.23), we get BSQ
— I\ — »So,
(6.1.26)
American Option Pricing and Optimal Exercise Strategy
121
a-BS^-'1 = -1. Solve it to determine
a_ U-l/a_y S o
=
r
^-.
(6.1.27)
Substituting into (6.1.26) to get the solution (6.1.28) Remark How do we price perpetual American call options? First, for dividend-paying perpetual American call option, the continuation region Ei is {0 < 5 < So}, the stopping region £2 is {So < S < 00}, where So is the optimal exercise boundary. Notice that So > K, thus perpetual American call option price V = V(S) satisfies the following free boundary problem:
7TS2!dS? + (r~^S!$5~rV
= 0
(°<s<s°)>
(6.1.29)
V(S0) = S0-K,
(6.1.30)
V'(So) = l,
(6.1.31)
V(0) = 0.
(6.1.32)
We can certainly take the same process as in the case of put option, to derive the pricing formula V = V(S) and the optimal exercise boundary So- However, in order to obtain more knowledge, we introduce the following theorem, which states the call and put symmetry between the perpetual American options. Theorem 6.1 Let Vc(S;r,q), Vp(S;r,q), and Sc(r,g), Sp(r,q) be the price and the optimal exercise boundary of the dividend-paying perpetual American call and put options, respectively, then Vc(S;rtq) = j^Vp(^-,q,r), VSc(r,q)Sp(q,r) = K, where K is the strike price.
(6.1.33) (6.1.34)
122
Mathematical Modeling and Methods of Option Pricing
Proof Let V = Vp(y; q, r) and S = Sp(q, r) be the solution of the following free boundary problem: ^y
2
^
+ (
(s p < y
(6.1.35) (6.1.36)
VP(SP) = K-SP, ' v;(Sp) = -l,
(6.1.37)
Vp(oo) = 0.
(6.1.38)
Define w = -vP(y),
(6.1.39)
K2 S=—. y
(6.1.40)
Under the transformation (6.1.40), d o m a i n { S p
=>•
K1
domain{0 < S < - — • } . Op
K2 = S*, the free boundary conditions (6.1.36) and(6.1.37) then become Denote -rr— Op
W(S*) = f-(K-Sp) and
dW ~dlT s=s*
(6.1.41)
=S*-K,
=
(dW dy\ \~3y~~3S)s=s,
-(-f)(-«L-
<6L42>
And the boundary condition (6.1.38) becomes: W(0) = \-Vp(y)]
= 0.
(6.1.43)
Finally let us derive the ODE that W(S) satisfies in {0 < S < S*}. It is easy to find by straightforward computation:
123
American Option Pricing and Optimal Exercise Strategy Thus the differential equation (6.1.35) is transformed to
0
a d
dVp
.
+ {q
= TW ¥
a . dVp
r
" " T)2/^" ~ 9Vp
w fa* d lc,dW.
f
.
2
+ [y + (9-r-y)- 9 ]w}, i.e.
Ts2CSr
+ {r q)S< rW
(6 L44)
=0
- §-
'
-
Combining (6.1.41)—(6.1.44), the pair {W(S), S'} in [0, 5*] is the solution of the free boundary problem (6.1.29)—(6.1.32) for perpetual American call option pricing. By the uniqueness of the solution of the free boundary problem (here we use without proof), we conclude S* = Sc(r,q).
W(S) = Vc(S;r,q),
By transformation (6.1.39),(6.1.40), we can go back to the original variables VP(S; q, r) and Sp{q, r), and arrive at the conclusion of the theorem. Now let us derive the perpetual American call option price formula from the perpetual American put option price formula using Theorem 6.1. We first point out (6.1.45)
a-(q,r) = 1 -a+(r,q). This can be derived from (6.1.25):
Q_(9,r)
= " 9 + ; 2 + ^ - \Mq-r-^y v
G
V
^
+
2^q
-[^± + ^-,-f)- + ~] =
l-a+(r;q).
124
Mathematical Modeling and Methods of Option Pricing
Thus by Theorem 6.1 and (6.1.45), Ve(S;r,q)=\-Vp(y;q,r)\
\a-(q,r)[l-l/a-(q,r)\ _
1 i
/ (1
1 •*•
<x+(r,q) - 1 V
V
\oc+(.r,q) I jsl-a+
]
^
(r,q) qa+(r,q)
a+(r,q)J
and (6.1.46)
_ Corollary
K
a+(r,q) <*+(r,q)-V
When q = 0,
(j
Z
Z
= 1. Thus 5C(T-, 0) = oo. This says for perpetual American call option, if the underlying asset pays no dividend, then the holder should never consider early exercising the option. In this case, option price Vc(S;r,0) = S.
6.2
Models of American Options
To be specific, let us still take non-dividend-paying American put option as example to establish the mathematical models. For an American put option with expiration date t = T, there exist two regions: the continuation region Hi, where: V(S,t)>(K-S)+;
125
American Option Pricing and Optimal Exercise Strategy and the stopping region E2, where: V(S,t) = (K-S)+. Between these two regions lies the optimal exercise boundary T : S = S(t). Similar to the discussions in §6.1, we can conclude Ei = {(S,t)\S(t) <S
E2 = {(5, i)|0 < S < S(t), 0 < t < T} , and S(t)
(0
(6.2.1)
In Ei, using the A-hedging principle and the ltd formula, we can infer: when (5,t) G Ei, the option price V = V(S,t) satisfies the Black-Scholes equation: (6.2.2) On the optimal exercise boundary I"1, V(S{t),t) = K -S(t), 8V — (S(t),t) = -1,
(6.2.3) (6.2.4)
(notice that, because of (6.2.1), we can omit the "positive taking" symbol.) and when S —> 00, V -> 0, (6.2.5) at t = T, V{S,T) = (K-S)+.
(6.2.6)
This means in order to price American put option, we need to look for function pair {V(S,t),S{t)} in Ei , such that they satisfy the PDE problem (6.2.2)— (6.2.4). Since S(t) is a free boundary, it is called a free boundary problem for a parabolic equation. R e m a r k The free boundary condition (6.2.4) indicates the option price's Pt\f derivative (i.e. A = -^-) is continuous at crossing the optimal exercise boundary. This fact, as we pointed out in §6.1, expresses the principle of American option pricing, i.e. the American option value is maximized by an exercise strategy that makes the option value and option derivative continuous. Similar to §6.1, starting from the free boundary problem (6.2.2)—(6.2.4), we can get the variational inequality model for American put option pricing: In domain E : {0 < S < 00,0 < t < T}, look for function V(S,t) G Ci = V(S, t)\V, ^ continuous on E, where E = Ei U E 2 U T, such that
126
Mathematical Modeling and Methods of Option Pricing (1)
in the continuation region E i , V(S,t) >
(K-S)+,
CV = 0, (2)
in the stopping region E2, V(S,t) =
(K-S)+,
CV = C(K - 5) = -TK < 0. (3)
At the terminal time t = T, V(S,t) = (K-S)+.
(4)
When S -> 00,
V(S, t) -> 0. Combining (1)—(4), we get the variational inequality form of American put option pricing: min{-CV,V-{K-S)+} = 0, (E) find {V(S,t) G CE = V(S,t)\V, ^ continuous on £}, such that (6.2.7)
! (0 < S < 00)
V(S, T) = (K-S)+,
(6.2.8)
(S -> cx>) (6.2.9) V -> 0. Remark For American option with a continuous dividend rate q, the model is f min{-£V,V \ V|t=T =
- g(S)} = 0, fl(5).
(6.2.10) (6.2.11)
where (6.2.12) f (A" - S ) + ,
(put option)
(6.2.13)
\(5-AT)+.
(call option)
(6.2.14)
It is natural to ask what relation exist between American call and put option prices. In fact, we can use a similar procedure to extend the perpetual American call-put symmetry (Theorem 6.1) to American options in general. Theorem 6.2 Let VC(S, t; r, q), VP(S, t; r, q) and Sc(t; r, q), Sp{t; r, q) be the prices and the optimal exercise boundaries of the dividend-paying American call
American Option Pricing and Optimal Exercise Strategy
127
and put options with the same expiration date T and strike price K, respectively, then Vc(S,t;r,q) =
^Vp(^-,t;q,r)
and ^Sc(t,r,q)Sp{t;q,r)
= K,
where K is the option's strike price, r the risk-free interest rate, q the dividend rate. The proof of this theorem is almost the same as the proof of Theorem 6.1. Therefore we leave it to the reader as an exercise.
6.3
Decomposition of American Options
Financially, a decomposition can be made to the American option price: a European option price plus another part due to the extra premium required by early exercising the contract. European option is priced by the Black-Scholes formula, and the extra premium obviously depends on the position of the optimal exercise boundary. In mathematics, we want to deduce the equation satisfied by the optimal exercise boundary 5 = S(t) from this idea. In order to deduce the integral equation of S(i), we need to use the fundamental solution of the Black-Scholes equation.
Definition 6.1 G(S,t;£,T)
is called the fundamental solution of the
Biack-Scholes equation , if it satisfies the following terminal value problem to the Black-Scholes equation:
f CV = S£ + £s2 0 + (r - q)S§V - rV = 0, \v(5,T) = 5(5-0,
(6.3.1) (6-3.2)
where 0 < S < o o , 0 < £ < o o , 0 < £ < T , 5(x) is the Dirac function. In order to obtain the expression of G(S, t; £, T), define z = ln|,
r = T-t.
(6.3.3)
Under the above transformation, (6.3.1),(6.3.2) becomes
!
dV
a2 d2V
,„ „
V(s,0) = £*(*), where x € R, 0 < T < T.
(AdV ,
u
n
(6.3.3)
(6.3.5)
128
Mathematical Modeling and Methods of Option Pricing In deriving (6.3.5), we notice that under the transformation (6.3.3),
S(S - 0 = me* ~ 1)) = \s(ex - 1) = itf(i). Similar to derivation in §5.3, make the transformation V = eaT+Pxu,
(6.3.6)
where
P = l-L=rL>
(6-3-7) (6.3.8)
Thus the initial value problem (6.3.4),(6.3.5) is reduced to
{
r\
re o n^
5?-^-^=°
( 6 - 39 )
du
O~ d U
>
(6.3.10)
u{x,Q) = e-l *\8{x) = \8{x). Solution to the PDE initial value problem (6.3.9),(6.3.10) is
Substituting it into (6.3.6) and by (6.3.7),(6.3.8), we get V(X T)
> = J^reXP{-rT
' ^
X +
^ ~ * ~ T^2>-
Going back to the original variables (5, t) by (6.3.8), we get the fundamental solution of the Black-Scholes equation -r(T-t)
G(S,t;£,T) = —^ — . fry/WT-t) ex
2
ln
• P{- J^WT) [ f +
(6.3.n)
2
(r-1-£)(T-t)} }.
Theorem 6.3 / / the fundamental solution G(S, t; £, rf) is regarded as a function of £,r), then it is the fundamental solution of the adjoint equation of the Black-Scholes equation. That is, let v{Z,ri) = G(S,t;Z,Ti),
American Option Pricing and Optimal Exercise Strategy
129
then v(£,r)) satisfies
(C'v = -§jj + 4jp{?v) - (r - q)^v)
-rv = O,
(6.3.12)
(6-3-13)
I «(€>*) = *«-<£).
where 0 < f < oo, 0 < 5 < oo, i < Vh Corollary Theorem 6.3 indicates, if the fundamental solution of equation (6.3.12) is G*(£,T);S,t), then G(5,t;€,»/) = G'tt,ij;S1t)Proof of Theorem 6.3 /•oo
0= /
JO
rn-l
dx
Jt + e
Consider the integral [Gm{x,y;S,t)£G(x,v;Z,TJ)
-
G(x,y,t,V)£*G-(x,y;S,t)]dxdy
Jo
Jt+C idV
2 dxK
dx'
-Ts[4H +(r " ?) S (lC(r) }*Since when x —• 0, oo,
G^(x2G*)^0, xGG' -» 0. Thus / G*{x,v - e;S,t)G(x,r] - e;Z,T))dx = H G*(x,t + e;S,t)G(x,t + t;(,,ri)dx. Jo Jo l
C*v , the adjoint operator of Cu = £ ^ = 1 "ij W a ^ . . + E?=i b^x)^:
+ c(x)u.
is denned by C*v = E " J = 1 g ^ j ( a i j ( ^ ) " ) - E?=i ^"( b i( a: ) t ') + c(x)u. £ and £*u satisfy / {vCu - u£*v)dx = 0,
Vw,t)£ Cfi°(n).
130
Mathematical Modeling and Methods of Option Pricing
Let e -» 0, from the initial conditions (6.3.2) and (6.3.13)
[°° G*(x,T);S,t)d(x-£)dx Jo = / Jo
6(x-S)G(x,t;S,7])dx,
i.e. G*(Z,V;S,t) = G(S,t;Z,V). thus completes the proof of the theorem. Now we can establish the decomposition formula of American option. Theorem 6.4 Let V(S,t) be the American put option price, then V(S,t) = VE{S,t) + e(S,t),
(6.3.14)
where VE(S, t) is the European put option price: d2) - 5e- g ( T " t ) 7V(- d\).
VE(S,t) = Ke~r(-T-t}N(-
(6.3.15)
where d\ and d2 are defined by (5-4-11), (5-4-12). e(S,t) is the early exercise premium, r-T
e(S,t)=
Jt
rS(V)
dV Jo
(Kr-qt)G{S,t;t,T,)dt.
(6.3.16)
where G(S,t; £,77) is the fundamental solution of the Black-Scholes equation. Proof For American put option, since in domain E : { 0 < 5 < o o , 0 < t < T} , the option price V(S, t) has continuous first derivative , and in each region has continuous second derivative, thus V(S, t) satisfies
-^S'* = {t-qs}sTet
(6 3 17)
--
where C is the Black-Scholes operator. Multiply both sides of (6.3.17) by G*(£,r];S,t) and integrate on the domain {0 < f < 00,( + e < r) < T}. Since E 2 = {0 < £ < S(r]),0 < 77 < T}, and S{rj) is
American Option Pricing and Optimal Exercise Strategy
monotonic, therefore /
dr,
Jt+e
(Kr-qQG'{t,T,;S,t)dt
JO
= - f
Jt+e
dr, f°° CTfoTKS.QCVdZ JO
= - I* dr, I" [CT&mS^CV&T,) Jt+e
JO
- V(t,r,)£'Cr(Z,rr,S,t)]dt
T|< I -|K 2 G-))+(--«)|«"O-)}«.
Since when £ —» 0, oo
ZVG* —* 0, Thus we have
r CT{i,t + vtS,t)V(i,t + e)di
Jo
= f"' G*(t,T;S,t)V(£,T)dt Jo
+ /
Jt + e
dr, /
[Kr-qi)G*{i,r,;S,t)dt.
Jo
Let e —» 0, and consider (6.3.13) and the Corollary of Theorem 6.3, we get /•oo
G(S,t;S,T)(K-Z)+dt;
V(S,t)= / Jo
+
rT
Jt
dr,
Jo
fS(V)
(Kr-qt)G(S,t;Z,r,)dt
= VE(S,t) + e(S,t). Q.E.D.
131
132
Mathematical Modeling and Methods of Option Pricing
Expressions (6.3.14)—(6.3.16) indicate: if the optimal exercise boundary S = S(t) is given, then American option price can be determined by (6.3.14)— (6.3.16). How can we find the optimal exercise boundary S = S(t)? As a corollary of Theorem 6.4, we can obtain a nonlinear integral equation for the optimal exercise boundary S = S(t). Theorem 6.5 The optimal exercise boundary of American put option S = S(t) satisfies the following nonlinear Volterra integral equation of the second kind:
S(t) = K + S(t)e-"^N ( - " " 1 ^ 4 f " t ) ) - Ke-«T-»N ( \
~ln^+/Mr-t)\ ay/T-t I
(6.3.18) where 0i=T-q~Y'
(6319)
/32=r--<7+y.
(6.3.20)
Proof By (6.3.16) and the fundamental solution expression (6.3.11), e(S,t) can be expressed as
e(S, t)= Jt
dq Jo
(Kr - qfl
/ a£yj2Tr(r) -1)
f hnl+friv-t)}2}
(6.3.21)
American Option Pricing and Optimal Exercise Strategy
133
Change the variable to
_ In f +Pi(y~t) cr^r] — t -<%
dx
a
ivV ~ t
and (6.3.21) becomes
&Vv — t
- - ^ - fTe^-^-^dv
f°° s
e-^+oJn^ix),
= Krt\-^-A.-4^^-^dn a>Jr) — t
-i fr jr.-^-»[i- J »('°'fej^-' 1 )]*, - ,s /T«-*-> f. - AT f - ^ ^ - ' n i „. (6.3.22) Jt
[
\
ay/v-t
Substituting (6.3.22) into (6.3.14), and taking into account
V(S(t),t) = K-S(t), thus we have K-S(t) = VE(S(t),t) + e(S(t),t),
yj
134
Mathematical Modeling and Methods of Option Pricing
S(t) = K + S(t)e- .<*-«>* - Ke~ '
f^+W-t)} ^T+fc(T-t)\
Q.E.D. While it is very difficult to solve the nonlinear integral equation (6.3.18), we can obtain an asymptotic expression of the optimal exercise boundary S = S(t) near the terminal point t = T. Remark We can use a similar derivation to obtain the expression of the early exercise premium for American call option:
e(S,t) = [ dn [°° (tf - Kr)G{S,t;i,v)di, and a nonlinear Volterra integral equation for the optimal exercise boundary S = S(t). 6.4
Properties of American Option Price
A qualitative understanding of American option pricing is appropriate before we start the numerical studies. In contrast to European option, American option price, except for the perpetual option, does not have closed-form solutions in general. Therefore in studying the properties of American option we rely on the theoretical approach of partial differential equations. As discussed in the previous chapter §5.8, this requires a study on the properties of the derivatives of the solution with respect to the variables or parameters in the boundary-terminal value problem to parabolic equations. But since American option price satisfies a variational inequality, it is not differentiable. Therefore we introduce a penalty function /3e(x), and set up a penalty problem corresponding to the variational inequality as its approximation. Then we will prove the assertion using the maximum principle for the parabolic equations. Moreover, since the payoff function
American Option Pricing and Optimal Exercise Strategy
135
(S — K)+ and (K — S)+ are not differentiable at S = K, we will also need to smooth the payoff function to ensure the differentiability of the PDE problem. Definition 6.2
Function f3e(x) is called penalty function in (—00,00), if MX) G Cf.^y &(z)<0,
(6.4.1)
&(0)<-<7«,
(C e >0)
(6.4.2)
#(i)>0,
(6.4.3)
# ' ( z ) < 0,
(6.4.4)
and
f 0, x > 0, y -00, a; < 0.
£lim/? £ (x)=4
-*°
(6.4.5)
Under the transformation x = \nS,
r = T-t
and
w(a;,T) = K(5,t), the variational inequality (6.2.10)—(6.2.11) becomes (mm{-C0v,v-(K-ex)+}=0,
(6.4.6) (6.4.7)
\v(x,Q) = (K-e')+, where „
CoV=
dv
a dv
,
d^-Y^-(-r-q-Y)di
o . dv
+ rv
-
Definition 6.3 In {x e R,0 < T < T}, the initial value problem to parabolic equations
I
+rv + 0e{v-ne{K-ex))
= O,
x
(6.4.8) (6.4.9)
[v(x,0)=Ilt(K-e ),
is called the penalty problem of the variational inequality (6.4.6), (6.4.7), where
{
V V > e,
/
\y\ < e,
0 y<-e,
(6.4.10)
136
Mathematical Modeling and Methods of Option Pricing n e ( y ) € C°°(R),1
> 0,
>n't(y)
n"(2/)>O,limn £ (2/) = i/+.
(6.4.11)
As a corollary of (6.4.10),(6.4.11), we have |n«(j/)-i/Ili(t/)|<e.
(6.4.12)
Let DT = {{x,t)\a <x
Theorem 6.6( the maximum principle) and \u(x,t)\ < M e Q | x | 2 ~ \
Ifu(x,t) e C 2 ' 1 (D T )nC(i5T), (Q,£>0)
and satisfies the linear parabolic equation in D: du , sd2u ,, ,du — - a{x,t)—^ + b(x,t)~ + c(x,t)u = where
. f(x,t),
a{x,i) > ao > 0, c(x,t) > 0 . x
If f( :t) < 0, -u(x,t) can on/y attain its nonnegative maximum at dpDrRemark The maximum principle also holds for higher dimensional parabolic equations. The proof can be found in textbooks. For the penalty problem (6.4.8)—(6.4.9), we have the following conclusions ([15])Theorem 6.7(Convergence of the penalty problem) Ifve(x,t) is the solution of the penalty problem (6.4-8)—(6.4-9), then as e —^ 0, in any finite and closed domain within DT, vc(x,t) converges uniformly to v(x,t) which is the solution of the Cauchy problem (6.4.6)—(6.4-7). Now let us study the properties of the American option pricing using Theorem 6.7 and the maximum principle(Theorem 6.6).
American Option Pricing and Optimal Exercise Strategy
137
To be specific, we will take American put option as example in performing rigid proof. Conclusions can be extended to American call option using the American call-put symmetry (Theorem 6.2). Let V = V(S,t) = V(X,T) be American put option price, and V€(X,T) be the solution of the corresponding penalty problem (6.4.8)—(6.4.9). Theorem 6.8 For American put option pricing, (1) ifSi>S2,then V(Sut)
(6.4.13)
ifKi>K2,then
0 < V(S, t;Ki)- V(S, t; K2) < Kx - K2.
(6.4.14)
Proof (6.4.13) and (6.4.14) can be proved in a similar way. We only prove the right side part of the inequality (6.4.14): V(S,t;Ki) - V(S,t;K2) < Ki - K2.
(6.4.15)
Define W = Ki-K2-
V1(X,T)+V2{X,T),
where Vi(x,r) = ve(x,r; Ki), (i = 1,2).
Substitute it into equation (6.4.8),
dW a2 d2W , a2sdW r q ) -d7-Y-dx^-^ - -Y ^
+ rW
- Pc(vi - U£(K! - e 1 )) + pe(v2 - ne(K2 - ex)) = r(Ki - K2). Since - /?«(wi - n e (tfi - e x )) + frfa - n£(A"2 - e x )) = " # ( 0 [vi -V2- (Uc(Ki - ex) - Ue(K2 - ex))]
=
A{S){w-(i-n'ttn))(K1-K3)},
where € = 0i(wi - ne(A"i - ex)) + (1 - 0i)(v2 - Ue(K2 - ex)), V = ftj(^i - ex) + (1 - 82)(K2 - ex), (0< 0i,02 < 1)
(6.4.16)
138
Mathematical Modeling and Methods of Option Pricing then (6.4.16) becomes
= [r + #(0(1 - n'fa))] (Ki - K2).
(6.4.17)
The initial value of W is: W\T=0 = {Ki - K2) - (n e (Ki - ex) - Ue(K2 - ex)) (6.4.18)
= (l-U't(71))(K1-K2). By definitions of 0C, II £ in (6.4.3),(6.4.11), we get
r + tf(0>0, r + ^e(0(l-ni(i7))>0, i-ni(»?)>o. Since it is assumed K\ >K2, applying the maximum principle to the Cauchy problem (6.4.17),(6.4.18), we conclude W > 0, i.e. vc(x,T;K1)-vt(x,r;K2)
< Ki - K2.
Let e —• 0, then by Theorem 6.7 V(S, t;Ki)-
V(S, t; K2) < Ki - K2.
thus (6.4.15) is proved. Lemma 6.1 (6.4.9), then
Suppose VC(X,T) is the solution of the penalty problem (6.4-8)— (6.4.19)
Proof
In order to prove the lemma, we assume that /3e is of the form /3t(a) = ~Ut(-a),
(6.4.20)
where Hc(y) is defined in (6.4.10). It is easy to verify that /3e(a) defined above satisfies (6.4.1)—(6.4.4). Then by (6.4.12), 0c(a) satisfies the inequality: |&(a)-Q#(a)|
(6.4.21)
139
American Option Pricing and Optimal Exercise Strategy In fact, by the definition of 0c(a) in (6.4.20), and (6.4.12), we have
| A (a) - a#(a)| = - -Ln e (- a) - ^ l £ ( - a) = -^|n«(-a)-(-a)n' e (-a)|
Now let us prove the inequality (6.4.19). Define
ox
r
It is easy to verify that W satisfies the equation: dW
a2 d2W
a2 ,dW
,
+ #(«) [ ^ + Il'c{K- ex)exj - Pc(a) = -^e, i.e. dW
a2d2W
a2 dW
,
= - V£(l + V e («)) + A(«) - ^ ( a ) k +n'e(»)ex], where a = vt — He(y), y = K — ex. Due to the inequality (6.4.21), and (6.4.3),(6.4.10), we have dW
cT2d2W
S.dW
,.
, .,. .....
= - Vi(l + ^/3=(o)) + & (a) - a#(a) - ^(a) [Uc(y) + We(y)ex] < - V~e{\ + ^P'e(a)) + \R < 0;
(6.4.22)
and at r = 0, W(x,0) = - U't(K - ex)ex - U€(K - ex) - ^
< 0.
(6.4.23)
140
Mathematical Modeling and Methods of Option Pricing
Therefore, applying the maximum (6.4.22),(6.4.23), we get
principle to the Cauchy <0,
W(X,T)
i.e.
dve
. y/e Vc
"a
-
ox Thus completes the proof of the lemma. Theorem 6.9 (1)
•
r
For American put option pricing,
if ri > Ti, then (6.4.24)
V{S, t; r2) > V(S, t,n); (2)
problem
if qi > q2, then
V(S,t;qx)>V(S,t,q2). (6.4.25) Proof Since inequalities (6.4.24) and (6.4.25) can be proved similarly, let us take (6.4.24) as example. Define W = V2(X,T) -
VI(X,T)
+
( n
~
r 2
^ ,
where Vi(x,r) = vc(x,T;ri)(i = 1,2). Substituting W into (6.4.8), a2 d2W
dW
{r2 q
^-T-^-
,
o-2.dW
)
- -Y ^x~
+ r2W
+ /3t{v2 - Ue(K - ex)) - j3e{vi -nt(K=
ir1-r2)V-e+ n
{dv1_
dx
e*))
dv±_ ox
Here /3£(a) is the penalty function as defined in (6.4.20). Thus by (6.4.19) (Lemma 6.1), we get
^ - ^ - t o - ' - ^ + ^ + fi™*
>(r2-n)[^-tn-#]>0, W(x,o)
=
^-2rr2^.
Thus by the maximum principle,
W(X,T)
cannot take the negative minimum
141
American Option Pricing and Optimal Exercise Strategy in {x e R,0 < T < T}. Then from W(x,0) > 0, we infer W(X,T)>0,
i.e. Ve(x,r;n) < vt(x,T-r2) + ( r i ~ r 2 ^ . Let e-tO, and by Theorem 6.7, we get (6.4.24). Lemma 6.2 If V€(X,T) is the solution of the penalty problem (6.4-8)— (6.4.9), then
Proof
^T 1 > 0. or
Denote
(6.4.26)
In domain DT{X € R,0 < r < T}, since VC{X,T), the solution of the penalty problem (6.4.8)—(6.4.9), is sufficiently smooth, W satisfies the Cauchy problem as follows:
/ ^F " i ^ \w{x,0)
~(r-q-
\)^r
+ rW +fc(a)W= 0,
(6.4.27) (6.4.28)
= >{x),
where xi \
9v
\a2d2Vc
e
a2 dve
.1
= y (n':(y)e2x - K{y)e*) - (r - q - ^-)K(y)e* - rUt{y) - P(0), where
y = K-ex.
By (6.4.2) and (6.4.11),
[nt(y) - yn'e(y)} - rKU't{y)
>a-^-rK. We can choose
- &(0) = C€=rK + ^e,
142
Mathematical Modeling and Methods of Option Pricing
so that (j>(x) > 0. Then applying the maximum principle to the Cauchy problem (6.4.27),(6.4.28), we conclude in DT W(X,T)>0,
i.e. (6.4.26) is true. Theorem 6.10 For American put option pricing, (1) ifh>t2, then V(S,t2)>V(S,t1); (2)
ifTi > T2, then when
(6.4.29)
0
V{S,t;Ti)>V{S,t;T2).
P r o o f Conclusion (6.4.29) is a direct corollary of (6.4.26). In order to prove (6.4.30), denote W{X,T)
= u£(x,r;Ti) - V,(X,T;T2).
(T = T 2 - t)
It is easy to check that in domain {x € R,0 < r < T2}, W{x,r) penalty problem
satisfies the
{ W- - ^ U - ^ - 1 - 4)T& + ^ + ^)W = °>
(^31)
\W{X,T)
(6.4.32)
= 4>{X),
where 4>(x) = vc(x, 0; Ti) - v,(x, 0; T 2 ) = v€(x,Q; Ti) - Ut(K - ex). Now back to the original time variable t <j>(x) = vt{x,Tr,T{) -Ut{K - ex) =
vt(x,T2;Ti)-vc{x,Ti;Ti)
> 0. Applying the maximum principle to the Cauchy problem (6.4.31),(6.4.32), we conclude W(X,T)>0,
i.e. VC(X,T;TX)
>vt(x,T;T2).
Let e —» 0, and by Theorem 6.7, we get (6.4.30) .
143
American Option Pricing and Optimal Exercise Strategy Lemma 6.3 then
Proof
Ifve(x,r)
is the solution of the penalty problem (6.4-8),(6.4.9),
33
£-&**
<" >
Denote . '
W{X T)
. =
d2ve dve ^ - - ^ '
Since the solution ve(x,r) of the penalty problem (6.4.8),(6.4.9) is sufficiently smooth in the problem domain, thus W(x,r) satisfies the problem:
+ (r + P'^))W = f(x> T)> (6-4-34)
i^-4^--(r-1-^)^ \w(x,0)
(6.4.35)
= 4>(x),
where
f(x,r) = - A'{a) (?£ +K(y)e*y + faa)!?:(y)e2*, 4>{x) = U"(y)e2x, a = vt —Uc(y), y=
K-ex.
From (6.4.3),(6.4.4) and (6.4.11) we get /(x,r)>0, 4>{x) > 0.
Applying the maximum principle to the Cauchy problem (6.4.34),(6.4.35), we get W{X,T) >0.
Thus the Lemma is proved. Theorem 6.11
If en > cr2, then V{S,t;a1)>V{S,t;a2)-
Proof
Denote W(X,T)
= VI(O;,T) -
V2(X,T),
(6.4.36)
144
Mathematical Modeling and Methods of Option Pricing
where Vi(x, r) = vc(x, r; en), (i = 1, 2). It is easy to check that in the domain DT, W(X,T)
satisfies the equation
(6.4.37) with the initial condition: (6.4.38)
W(x,0)=0.
Prom (6.4.33) and the theorem's assumption, the right side of equation (6.4.37) is nonnegative. Thus applying the maximum principle to the problem (6.4.37),(6.4.38), we get W(X,T) >0,
i.e. vt(x,T;ai) > vt{x,T\O2)-
Let e -> 0, and by Theorem 6.7, we get (6.4.36). By American call-put symmetry
|y,(y,i;«,r),
Vc(S,t;r,q)=
(6.4.39)
we can extend the properties of the American put option price (Theorems 6.8— 6.11) to the American call option. For simplicity of writing, we define points P, Po and Q: P=(S,t;r,q,T,a),
Po = {So, to; ro, qo, To, Co), Q=
(^-,t;q,r,T,a).
Then (6.4.39) can be written as
Vc(P) =
K
^Vp(Po) Po=Q
American Option Pricing and Optimal Exercise Strategy
145
and assuming corresponding derivatives are meaningful, then:
dr
K dq0
PQ=Q
dVc=S_dVp dq
K dr0
dt - K
dt
PQ=Q
W )
-
Ul
&-£&«»** According to the above analysis of American option price, we get the following table: (here "+" indicates the option price is an increasing function of the variable, "-" indicates the option price is a decreasing function of the variable) Variable S(Stock Price) K (Strike Price) r (Risk-free Interest Rate) q (Dividend Rate) a (Volatility) T(Life) i(time)
Call Option Put Option = + + + + + + + ~ +
By comparing this table with the table in §5.8, we conclude: (1) American option price has the same dependence as European option price on S,K,r,q,cr. (2) In contrast to European option price, American option price has a definite dependence on the option lifetime T and time t. Financially, this is because for American options, whether put option or call option, an option with longer life has all the exercise opportunity that options with shorter life have, hence a longlife option must be worth at least as much as the short-life options. As for the time factor, since as t increases, T — t decreases (the terminal time t = T gets
146
Mathematical Modeling and Methods of Option Pricing
closer), the profiteering opportunity from early exercising the option decreases, hence the option becomes less valuable.
6.5
Optimal Exercise Boundary
Since American option has early exercise term in the contract, the determination of the optimal exercise boundary is of special importance to the holder of American options. But as a free boundary, its determination is coupled with the determination of the option price. Therefore, a closed-form expression of the free boundary is not attainable in general. In this section, we will use the PDE theory to make qualitative analysis on the optimal exercise boundary S = S(t), including the position of S(T), the monotonicity of S(t), upper and lower bounds of S(t) and convexity of S(t) etc., and based on these, give an asymptotic expression of S(t) near t = T. All these results will not only give us a good knowledge about the optimal exercise boundary, but will also help the numerical computation of the American option price. First, let us determine the position of the optimal exercise boundary S = S(t) at t = T. Theorem 6.12 Suppose F : {S = S(t), (0 < t < T)} is the optimal exercise boundary for dividend-paying American option, then
min(^,K), I t )
(put option)
! max/^,^}.
(call option)
(6.5.1) (6.5.2)
Proof First let us consider the American put option. (A) If q < r,
(6.5.3)
then we need to show S(T) = K. Suppose otherwise, i.e., S(T) ^ K. As we already pointed out: since V > 0, therefore S(T) < K. Then in the continuation region Ei there would exist a region Ds : {S(t) < S < K,T - 5 < t
CV = 0,
American Option Pricing and Optimal Exercise Strategy
r
—.wsi
0
Then at t = T, S(T)
147
S(T) K
S
<S
£L~[£-£+<->*£-"L = rtf - 9 S. With (6.5.3) and S < K, there would be
9t
t=T
But V(S,T) = (K - S), therefore inside Ds, there must be V(S,t)< (K-S). But this contradicts to the property V > (K — S) + . Thus we conclude S(T) = K (B) If r
D(61}
(6.5.4)
^ .
If S(T) < ^J£. Since |AT < /iT, then in Ei there would exist a region
: J5(i) < 5 < ^ , T - 5 < t < T\, such that CV = 0.
148
Mathematical Modeling and Methods of Option Pricing t
0
S{T) Kr/g
S
Then similar to (A), inside D^1', there would be dV
-sd t
t=T
=
TK
- qS > 0
and K(5,t) < K - 5 . This contradicts to the property V(S,t) > (K - S) +. (b) If S(T) > ^K. Since S(T) > K, and (6.5.4), then in the stopping region of the option E2 there would exist a region
Df) : I -K < S < S(t),T - 5 < t < T\ , such that inside D5 , }
Then in Df
V =
K-S.
there would be -£V = rK - qS < 0.
r
I A
'"JLZ °
Kr/q
SiT)
S
This contradicts to equation (6.2.10). Therefore there must be S(T) = ^ - .
American Option Pricing and Optimal Exercise Strategy
149
Combining conclusions of (A) and (B), (6.5.1) is proved. Similarly one can prove (6.5.2). Remark In the above proof, we made following two a priori assumptions: (1) S = S(t) is continuous; (2)
in addition, both $¥-, ^ - X are continuous up to t = T excluding S — K. OT oS We omitted the proof of these two assumptions (see [15]). Remark Based on the position S(T) (6.5.1) for American put option, using American call-put symmetry (Theorem 6.2), we can derive directly the position of the optimal exercise boundary at the terminal time S(T) (6.5.2) for American call option in Theorem 6.12. In fact, from (6.1.34)
5c(T;r 9) =
'
K2
Sp(T;q,r)
_ f K,
=
K2
min{K,^}
r
IS*. r > q = max(K, -K). Corollary For American call option, if q < r, then S(T) = -K. q
/ / <7 -> 0, then S(T) -> oo. Thus, using the PDE model of American option price, we have proved again : non-dividend-paying American call option should never be exercised prior to the expiration date. Theorem 6.13 Suppose T : {5 = S(t), (0 < t < T)} is the optimal exercise boundary of American option, then ( Sp(t) is monotonic non-decreasing,
(put option)
\ Sc(t) is monotonic non-increasing,
(call option)
and for put option there are estimates as follows ( rK} SP,oo < Sp(t) < min i K, L
(6.5.5)
150
Mathematical Modeling and Methods of Option Pricing
and for call option there are estiamtes as follows max IK, — | < Sc(t) < Sc,oo.
(6.5.6)
Here SPtOO and SC:OO are the optimal exercise boundary for the perpetual American put and call options, respectively. They have explicit expressions (6.1.27), (6.1.46) and (6.1.25). Proof According to American call-put symmetry, we only need to prove for the case of American put option. In fact, suppose S(t) is not monotonic non-decreasing, (for simplicity of writing, we omit the subscript p, i.e. Sp(t),SPtOo will be written as S(t) and 5oo, respectively) i.e. there exists 0 < £i < £2 < T, such that S(h) > S(t2). Then at t = £2, {0 < S < S(t2),t = £2} belong to the stopping region E2, {£(£2) < S < 00,t = £2} belong to the continuation region Si. In particular (5(ti),i2) € Si, i.e. V(S(ti),i2)>(K-S(ii))+. But since (S(£i),£i) G V, i.e. V(S(t1),ti) =
(K-S(t1))+,
thus we conclude: when £2 > ti V(S(t1),t2)>V(S(t1),t1). This contradicts to (6.4.29). Thus we have proved for American put option that the optimal exercise boundary 5 = S(t) must be a monotonic non-decreasing function of £. Due to the monotonicity of S(£), when 0 < t < T, there is S(t) < S(T) = min {K, — \ .
I
1 J
(6.5.7)
Now let us prove the lower bound of S(t). Since American option price as a function of the expiration date T is monotone increasing (see (6.4.30)), financially, among all American options with the same
American Option Pricing and Optimal Exercise Strategy
151
strike price K, the perpetual American option is the most expensive, because it includes all the exercise opportunities that other American options have, i.e. (6.5.8)
V(S,t;T) < V{S;oo),
where V(-; T) denotes the price of American option with lifetime T(0 < T < oo). Let Soo be the optimal exercise boundary of the perpetual American put option, so that when 0 < 5 < Soo, there is V(S,t;T)
-S)+,
V(S,t;T) = (K
-S)+.
thus when 0 < S < S^ This indicates region {0 < S < Soc,0 < t < T} C £2(the stopping region), thus for all t € [0,T], there is S(t) > Soo. (6.5.9) Q.E.D. Remark When q = 0, the optimal exercise boundary of American put option T : S = S(t) is a convex curve, i.e. S"(t) > 0([8]). The proof of this conclusion is not elementary and is thus omitted. Prom Theorem 6.12, Theorem 6.13 and the above remark, although we do not have a quantitative expression of the optimal exercise boundary, we do have a basic understanding of its picture. For American put option (q > 0) (for q = 0, the convexity is proved) t
O
4!
mm(K,Kr/
S
152
Mathematical Modeling and Methods of Option Pricing
For American call option(g > 0) ! IS"—1
T
O
tmxtK,Kr/q)
S
Now we need to find an approximate expression of the optimal exercise boundary, i.e. a quantitatively description of the optimal exercise boundary. (A) Case q < r. Theorem 6.14 When 0 < T - (
~ /
W
* o-V(T-t)\ln(T~t)\.
(6.5.10)
For simplicity we prove the theorem in the case of q = 0 only, since other cases can be treated similarly. Although the proof of the theorem is lengthy, the idea behind the proof is clear. First, define a level curve S = S(i), which is the solution of the following functional equation: VE(S(t),t) = K - S(t),
(6.5.11)
i.e. 5 — S(t) is a curve on which the European put option price equals the exercise payoff. Since (see §5.8)
^
= -[1-^)1.
thus ^[VE-(K-S)}
= N(d1)>0
thus according to the implicit function theorem , the functional equation (6.5.11) has a unique solution S = S(t), which belongs to C^Ty and S(T) = K.
American Option Pricing and Optimal Exercise Strategy
Lemma 6.4
153
When 0
(6.5.12)
Before we prove the lemma, let us first look at its meaning. In fact, (6.5.12) leads to K - S{t) _ K - S{t) K ~ K
+
S(t) - S{t) K
Therefore, the term at the right side of the asymptotic expression (6.5.10) comes mainly from K ' w n e r e ^(*) *s a definite function to be determined by equation (6.5.11). Thus the proof of Theorem 6.14 is reduced to a rather elementary calculus problem. Proof of Lemma 6.4 From (6.3.16) (where q = 0) we know: in domain S : {0 < S < oo, 0 < t < T) V(S,t) > VB(S,t), Thus on the curve 5 = 5(t), V(S(t),t) > VE(S(t),t) = K - S(t), i.e. 5 = S(t) is in the continuation region Ei. Therefore,
s(t)<s{t). we[o,r]. By the Taylor formula: V(S(t),t) = V(S(t),t) + ^ | £ M ( S ( t ) _ S(t))
+ \^Z where ^ e (S(t),S(t)).
(S(t)-S(t)f,
By the free boundary conditions (6.2.3),(6.2.4), we
154
Mathematical Modeling and Methods of Option Pricing
get V(S(t), t) = K- S(t) - (S(t) - S(t))
+ l^Z (S(t)-S(t)f (S(t)-S(t))2.
=K-S(t)+^ By (6.5.11) and (6.4.29), V(S(t),t) - VE(S(t),t) = i | ^
(S(t) - S(t)f
_[%t)-5(t)f[
dV dV]
>-[S(t)-S(t)}2}^-—
.
(6.5.13)
It is easy to check u = jXr satisfies the following PDE problem in Si: ^*
"2" a ^
(
r + a ) 6 g y - 0,
' w|t=T = 0,
(b.5.14) (6.5.15)
.u\s=s(t) = - 1 -
(6.5.16)
Let u# = - ^ # , then in Ei, U£;(5, i) also satisfies the same equation (6.5.14) and the terminal condition (6.5.15), but on 5 = S(t) it satisfies the boundary condition UE s
=s^ ~
dVE(S(t),t) dS '
Since
dvE(s(t),t) dS
-
'
applying the maximum principle to the function uE — u in Si, we get uE — u > 0,
155
American Option Pricing and Optimal Exercise Strategy i.e. in S i , dV_
8VE_
~dS ~ ~~dS~'
Substituting it into (6.5.13), we get
V(S(t),t) - VE(S(t),t) > - [S(t) - S(t)]2 —-§r = [S(t)-S(t)} —= 2
2
f°°
e-Vda,
a V2TT Jdi(t)
(6.5.17) where
ln-jU(r+£)(T-t) Since £ G (S(t), S(t)), thus £
z
Therefore, when 0 < T - t < 1,
V27T 7dl(«)
4
2V2w Jo
Substituting it into (6.5.17), and by (6.3.16), we get 0 < 5(i) - S(t) < 2a^[V(S(t),t)
- VE(S(t),t)]i
< C(T-t)i. Q.E.D. In order to prove: when O < T - ( < 1 , K
~K{t) *°V(T-t)\ln(T-t)\
(6.5.18)
156
Mathematical Modeling and Methods of Option Pricing
is true, let us review the functional equation (6.5.11) for S(t). With the expression of VE(S, t) (the Black-Scholes formula), (6.5.11) can be written as K - S{t) = Xe- r(r -*>(l - N{d2)) - S{t){l - N(dx)), i.e. e -r(T-t)
_ ! = eln iH1+r(T-t)N(di)
_
ff^
(g 5 l g )
where
_ l n f + (r-+^)(T-t) 1
ay/T^l ln%l + (r-^)(T-t)
"2
/^—T
=
•
(TV J - *
Let
then (6.5.19) can be written as l^eriT-t)
= N
m +
{L_
_ eay(t)VT=t+r(T-t)N^t) + ^_ + Z)y/T^t). (7 2
(6.5.20)
Since S(t) < K, therefore 0>aVT^ty(t)=\n^
=
_K_m + 0(\K-S(t)n
In order t o prove (6.5.18), we only need t o prove: a s O < T — £ < ^ 1 , there exists the following asymptotic expansion for y(t),
y(t) « - y/\]n(T-t)\.
(6.5.21)
157
American Option Pricing and Optimal Exercise Strategy
For this, we need the following lemma. Lemma 6.5 Let r = T — t, then lim y(r) = -oo.
(6.5.22)
T—>0
Proof
For simplicity of discussion, let us assume lim 2/(T) = (3
T—>0
(In general, the limit can be replaced by its upper limit). Note that equation (6.5.20) can be rewritten as 1 1 /-S+(7-f)V? 2 — (1-6;^) = — = / e~ — da y/r y/2nT Jy+{~ + i)V7 + - T = ( 1 - erT+a^Fy) V27TT
/
e-Tda.
J_oo
Let r —* 0, we get
It is easy t o see, t h e above equation h a s a single root: j3 = - o o . Q.E.D.
Lemma 6.6
Letr-T-t,
then lim Vry(T) = 0.
Proof rectly.
(6.5.23)
(6.5.23) is verified by the definition of y{r) and S(T) = K di-
Now we can prove the asymptotic expression (6.5.21). Lemma 6.7 When 0 < T - t < 1, y(t) « - y/\\n(T-t)\. Proof
(6.5.24)
Since
N(y(t) + C~ ± | ) \ ^ ^ ) = 7NT(y) + iV'(y)(^ ± \)y/T=t
+ R±,
(6.5.25)
158
Mathematical Modeling and Methods of Option Pricing
where
*± = \N"{V
+ e±Vf^l)(^
±a-)\T-1),
where
thus
I^± I < C(T - t) max N"{y + OVT^t) |0|<£+f
Oy/T=i\e-li*S^.
= C(T-t) flmax Jy+ Due to (6.5.22), thus \R±\=o(T-t).
(6.5.26)
Substituting (6.5.25),(6.5.26) into (6.5.20), we get (1 _ e°Sr=iv+rlT-t))N(y) =
- | V T ~ i ( l + e »VT=IS+r(T-t) )JV / (i|) (6.5.27)
-r(T -t)+ o{T - t).
With the asymptotic expansion of N(x) at x = —oo
N{x)=
+ +o{
7r^H ^ M'
(6528)
--
and by Lemma 6.6, we get 1 _ e«VT=iy+r(T-t)
=
_ ^^f^ly
+ \a(T - t)f
+ 0{(T-t) + (VT^t\y\)3).
(6.5.29)
Substituting (6.5.28),(6.5.29) into (6.5.27), and neglecting higher order terms of T — t, we get
i.e.
American Option Pricing and Optimal Exercise Strategy
159
a Thus we have
i.e. y*-y/\hx{T-t)\. Thus proves the Lemma 6.7. By definition V = —, aVT^t
In —^r-, K
thus
^W« f f > /(r-t)|in(r-t)|. Then by (6.5.12), Theorem 6.14 is true. Remark If 5 = r, then Theorem 6.14 is not true, and the asymptotic expression (6.5.10) will be replaced by
^
^
a vW(T-t)|ln(r-t)|.
The proof is similar. We refer the reader to [39]. (B) Case q > r. As q > r, S(T) = — . q
(6.5.30)
We will prove the following asymptotic expansion for the optimal exercise boundary S = S(t)([13]).
160
Mathematical Modeling and Methods of Option Pricing
Theorem 6.15 Let S = S(t) be the optimal exercise boundary of American put option, and q > r, then we have, as 0 < T — t <^1, & - S(t)
,
a
(6.5.31)
where a is the root of the following transcendental equation,
a 3 eT / e-^dS = 2(2 - a 2 ) . Ja
(6.5.32)
Proof In Si (the continuation region), American put option price satisfies the free boundary problem:
' ^ + £ s 2 f g + (r-«)S?S-rV = 0, V{S(t),t)=K~S{t),
(El} (0
< Vs(S(t),t) =-I,
(0
V(S,T) = (K-S)+,
(Z|C<5 < 0 O )
r
S{T) = -f, where Ei = {(S, t)\S(t) < S < oo, 0 < t < T}. Define Q
z = ln—, K
(6.5.34)
T = Y(T~ *),
(6-5-35)
P = ^(K-S-V),
(6.5.36)
x(r)=ln^,
(6.5.37)
a:(0)= I n - . a
(6.5.38)
The new unknown function {p(x, r), a;(r)} satisfies a free boundary problem as follows:
161
American Option Pricing and Optimal Exercise Strategy
i? = l^ + (fc'-1)^-^+/(a;)' P(X(T),T)
(f]l) (0
= O,
(0
< PX(X(T),T)=O,
p(x, 0) = -{ex - 1)+,
(x* < x < oo)
x(0) = x*, where Si =
{(X,T)\X(T)
< x < oo,Q < T < f},
(6.5.39)
f(x) = {k' - k)ex + k,
(6.5.40)
k=%,k'=2-^l. a a
(6.5.41)
In order to obtain an asymptotic expansion of X(T) at r = 0, let us observe the behavior of the solution {p(x, T),X(T)} near (x*,0). By (6.5.40),(6.5.41), / ( i ) has asymptotic expansion at x = x*:
/(*) = /(x*) + f'(x*)(x - x*) + O(\x - x*\2) = \{k' - k)ex" + jfe] + (k' - k)ex' (x - x*) + O(\x - x*\2) &-k(x-x*). At a; = x*, since q > r, thus p(:E*,0) = - ( e * * - l ) + = -(l-lj
=0.
Therefore there exists a neighborhood of x = x*, in which p(a:,0)=p I (a;,0) = 0. Since p(x,r) and px(a;,T) are continuous up to T = 0, thus near x = x*, neglecting higher order terms of \x - x*\, the free boundary problem can
162
Mathematical Modeling and Methods of Option Pricing
be reduced to
3? = & " k{x ~ x*}'
(X(T)
P(X{T),T)=0,
(r>0)
(6.5.43)
(T>0)
(6.5.44)
(x* <x < oo)
(6.5.45)
« p x (z(r),r)=0, p(x,0) = 0,
~ x K °°' T > 0)
(6 5 42)
-'
x(0) =x*. By dimensional analysis, we know this problem has a self-similar solution: P{X,T)=T*W{£),
(6.5.46)
X{T) = x* - asfr,
(6.5.47)
where I =^
.
(6.5.48)
By straightforward computation, dp
3 iTTr. ,
idW,
»
Substituting into equation (6.5.42), we get 3
i T2
2 ^
I'-XS^ 2
- ^
= r 2
i52W
^
2
-
+
fc(x
- ^
Multiplying both sides by r " , and noticing (6.5.48), we obtain a free boundary problem for an ordinary differential equation: find {W(£),a} such that:
^f
+ i% - iW = -K
(~oo <£<<*), (6.5.49)
< W(a) = 0,
(6.5.50)
. W'(a) = 0.
(6.5.51)
163
American Option Pricing and Optimal Exercise Strategy
We need to add an asymptotic condition for W at £ —> — oo. Noticing that £ —» -ex) is equivalent to r —» 0,x ^ a;*, then by (6.5.45), (6.5.46), we get
lim [W(0 - k£] = lim
[%^ -
k^-^\
=
^
rp(»,r)-p(x>0)_ x ^ x j
=
lim
r" 1 / 2
lim
r-1/2 /
=
r^0,x#x«
/ pT(x, 0T)d6 - k(x* - x)
[JO /•I
J
pxx{x,6T)d9.
Since p(x, r) is the solution of the free boundary problem (6.5.42)— (6.5.45) and p(x, r) = 0 as x > x*, we know from the regularity theory of parabolic equations: as x > x*, pXx(x,0) = pXxt(x,0) = 0. Therefore the limit of the right side of the above equality is 0. Thus as £ —> — oo, W ~ A£.
(6.5.52)
Thus under the transformation (6.5.46)—(6.5.48), the free boundary problem (6.5.42)—(6.5.45) is reduced to a free boundary problem to a second order ODE (6.5.49)—(6.5.52). Now we need to find {W(£), a} that satisfies the free boundary problem (6.5.49)—(6.5.52). The non-homogeneous ODE (6.5.49) has a special solution:
and its general solution is in the form
W(Q = k(; + A(e + 60 + B (£2 + 4)e-T + I(£3+6£)y"*
e-£dsL
(6.5.53) And by the boundary condition (6.5.52), A = 0.
(6.5.54)
164
Mathematical Modeling and Methods of Option Pricing
By boundary conditions (6.5.50),(6.5.51), 0 = ka + BR(a), 0 = k + BR'(&), i.e. B\R(a)-aR'{a)} =0, where
R(0 = (e + 4)e~£ + V + 60 t e-^dS. *•
J-oo
Let Q be the root of the equation R(a) - aR'(a) = 0, i.e. a satisfies equation (6.5.32), then
B
'W)-
< 6 - 555 >
Substituting (6.5.54),(6.5.55) into (6.5.53), we find
Substituting this into (6.5.46),(6.5.47), we get the asymptotic expansion of p(x,T)&t(x*,0):
and X(T)
= X* — a.y/~T.
Then by transformation (6.5.34)—(6.5.38), we get the asymptotic expansion of V(S, t) and S(t) at ( ^ , T). This completes the proof of Theorem 6.15. Remark By American call-put symmetry, we can obtain the expression of the optimal exercise boundary for American call option near the expiration date t — T, as q < r,
S
JL± - JLaJr=i
American Option Pricing and Optimal Exercise Strategy
165
where a is the root of the transcendental equation (6.5.32). Remark By numerical computation, an approximate value of the root a of equation (6.5.32) is determined to be a « 0.9034...
6.6
Numerical Method (I)
Finite Difference Method
American option pricing is a free boundary problem to the Black-Scholes equation. Except for the perpetual American option, in general there is no closed-form expression of the option price. Therefore numerical methods are particularly important for American option pricing. In the next two sections, we focus on the difference method and the line method. The former is based on the discretization of the variational inequality (6.2.7)—(6.2.9), and the latter is based on the discretization of the free boundary problem (6.2.2)—(6.2.6). To be specific, in the following we consider non-dividendpaying American put option only. Other cases can be treated similarly.
(A) Explicit Difference Scheme For variational inequality (6.2.7)—(6.2.9) given in £ : {0 < S < oo,0 < t < T}, make the transformation x=ln|
(6.6.1)
v{x,t) = jfV(S,t).
(6.6.2)
and
Thus it is reduced to (mm{-£0v,v-(l-ex)+}=0, \v(x,T) = (l-ex)+,
(xeR,0
(6.6.3) (6.6.4)
(x&K)
where £oV =
dv
+
a2 d2v
^ T ^
+( r
a2,dv ) r
-T ^- "
In the region {x e R, 0 < t < T}, make a mesh
Q = {(nAt, jAx)\0
, (6 6 5)
'-
166
Mathematical Modeling and Methods of Option Pricing rp
where Z is the set of natural numbers, At = 4T, AX > 0. At each mesh point, define a function vj =v{jAx,nAt),
(6.6.6)
tfj = V{jAx) = (1 - eiAx)+.
(6.6.7)
Let \9t)n+1J-
At
'
vgl - iff?
fdv\
W)n+1J
Ax"
Substituting them into (6.6.3)—(6.6.4), we get the equation for the mesh point (jAx, (n + I)At):
{
vn+1 - iin v
v
i
i
2 vn+1 - 2 D " + 1 4- vn+1
Si ^ + rv?,v?-
Zv
i
2
+v.i-i
2 vn+1 - vn+1 ,
(
a \ vi+i
Ax ^ ' (0
v
j-i
2Si (6.6.8) (6-6.9)
Due to the fact min(^, B) = 0 ^==> min(a.A, B) = 0 ,
(a > 0)
and min(C -A,C-B)
= 0 <^=> C = max(4, B),
thus in equation (6.6.8), let a = At, and we get min{(l + rAt)v] - (1 - w)^" +1 - av^ here CT2At
- cv£f,v?
-
167
American Option Pricing and Optimal Exercise Strategy
u , At
a2
c = w — a.
Then in (6.6.8), let a = l+\^t,
we get
min w
{ " - 1+7** t (1 ~ w)^"+1 + ""ft 1 + CV^ 'w" " ^ } = °-
i.e.
< = m a x { T + W K1 - W K +1 + K+i1 + <»£i] • Vi } ,
(66n)
( 0 < n < i V - l , j eZ) Thus v™ can be computed in the following steps: (1) Define ipj by
Ax2 and 1
a2
then the explicit difference scheme(6.6.9),(6.6.11) is convergent, i.e. lim
v&(x,t) = v(x,t),
where v&(x,t) is the linear interpolation of the lattice function v(jAx,nAt) = v™, and v(x,t) is a viscosity solution of the variational inequality (6.6.3),(6.6.4). The definition of the viscosity solution and the proof of this result involve a lot of PDE theoretical knowledge, and will not be discussed in this
168
Mathematical Modeling and Methods of Option Pricing
book. We refer the reader to [21]. Theorem 6.17 For American put option, the BTM and the explicit difference scheme (6.6.9),(6.6.11) of the variational inequality (6.6.3), (6.6.4) at ui = 1 are equivalent. Proof Similar to the proof of Theorem 5.5, in the domain S : {0 < 5 < oo, 0 < t < T}, make a mesh (j = 0 , ± l , ± 2 , . . . )
Sj=Kui, tn = nAt,
(n =
0,l,...,N)
rp
where u > 0, At = 4r. Then American option valuation formula can be written as
V? = max fyqVr*1 + (1 - g ) ^ 1 ] , ^ } ,
where ud= 1, p— d
q=
—d'
p=l + r At, $ j = (/f-5 j ) + . Let Ax = lnw, Xj — jAx — jlnu = In -£,
Then by (5.7.18), q=\ +
^(r-^)^Ai+O(At).
(6.6.12)
American Option Pricing and Optimal Exercise Strategy
169
Neglecting higher order terms, (6.6.12) can be written as
L
(
(6.6.13)
a2
Comparing (6.6.11) with (6.6.13), in (6.6.11), let w = 1, i.e. <7 2 A* _
Ax 2
'
then in (6.6.11), the coefficient a can be represented as 1 At . a2. 1 1 . a2, r— Thus (6.6.13) and (6.6.11) (as w = 1) are equivalent. From Theorem 6.17 and the convergence theorem of the explicit difference scheme of American option, we get:
Theorem 6.18(Convergence theorem of the BTM of American 1
2
put option) If -K r - \ Ax < 1, then when At -> 0, the BTM of American put option converges to the cohesive solution of the variational inequality (6.6.3), (6.6.4).
(B)
Implicit Difference Scheme
To solve American put option pricing problem (6.6.3),(6.6.4) by implicit difference scheme, first we need to truncate the infinite mesh at —N± and •^2, (Ni,N2 > 0), and add boundary conditions on the two boundaries
{x = -Nx,0 < t < T} and {x = N2,0 < t < T}. From the properties of American option, if N\ is sufficiently large, the boundary {x = —Ni,0
=
(6.6.14)
170
Mathematical Modeling and Methods of Option Pricing
And since v —> 0 as x —> oo, therefore if A^2 is sufficiently large, on the boundary {x — N2,0 < t < T}, there is v(N2,t)=0.
(6.6.15)
Thus on {-Ni < x < N2,0 < t < T}, v(x,t) satisfies the variational inequality
'-t-£0-(r-^)|i+^0' < v > tp(x),
(6-6.16) (6.6.17) (6.6.18)
In the region {-iVi < a ; < A ^ 2 , 0 < t < T } make a mesh (jAx,nAt),
(-rai
<j
where At = 4T, Aa; = -^ = jA. Define lattice function v" = v(jAx,nAt), And discretize the variational inequality (6.6.14)—(6.6.18): let
(dv\
vr1-^
2Ax '
\dx)n]
(&v\ _v?+1-2v? + v?_1 2 \dx JnJ Ax2 Thus we have ' -av?+1 + bv
(6.6.20)
v? > ipj, +1
[-av?+1 + bvj - cv^x - vj ] [v? - ipj] = 0, ' ^ni=^-ni,
(6.6.19)
(6.6.21) (6.6.22)
< 2 = 0,
(6.6.23)
. t ^ = ipj,
(6.6.24)
American Option Pricing and Optimal Exercise Strategy
171
(-rii + 1 < j < n 2 - 1,0 < n < N - 1) where a=
2
+
i^(r"Y)Ax'
c = u) — a,
5=l+w + rAt, a2At
Remark It is not difficult to prove that the results of Theorem 3.5 remain valid for the scheme (6.6.19)-(6.6.24). (6.6.19)—(6.6.21) can be rewritten in the matrix form
{
AV n > 6n,
(6.6.25)
V" > $ ,
(6.6.26) n
n
(AV" - O )j(V - *),• = 0,
(-ni + l < j < n 2 - l )
where ' b -c A =
0•
—a b —c
—a b —c -a b .
.0 V"=
:
&
n2-l
/an
(6.6.27)
172
Mathematical Modeling and Methods of Option Pricing
>r>2-l
jp-ni+i._
where it follows from the boundary conditions (6.6.22), (6.6.23),
0] = v]+1.
(-ni
+
2<j
The algorithm is as follows: (1) Define
{-m <j
(2) By backward induction, we solve the discrete scheme of the variational inequality (6.6.25)—(6.6.27) to find v?(0 < n < N - 1) step by step. If w™+1 is known, how can we solve the discrete scheme of the variational inequality (6.6.25)—(6.6.27) to get v™, {-m + 1 < j < n2 - 1)? We will follow the idea of the Gauss elimination method. To solve the equations, we first reduce matrix A to a lower triangular matrix, then solve the equations step by step. Since it involves inequality, after each elementary operation, we must check the sign of the proportional divisors. The solution-finding procedure goes as follows. Let us start from the last two equations in (6.6.25): -avn_ni+3
+ bvn_ni+2
-avlni+2
- cvn_ni+1
> 9n_ni+2,
+ 6 i £ n i + 1 > en_ni+l.
Let Aa; be sufficiently small, such that 1
- ~2 r~°~W
Aa;>0
thus a > 0 , c > 0.
>
(6.6.28) (6.6.29)
American Option Pricing and Optimal Exercise Strategy
173
(6.6.29) x ^ + (6.6.28), then (6.6.28) becomes -av\1+3
+ (b- j)vn_ni+2
> 6n_ni+2 + -b6n_ni+1.
(6.6.30)
Since r > 0, therefore under above elementary operations, the sign of the inequality is preserved. Denote h
ac
h
O-m+2 = o-
y-
Thus (6.6.30) can be written as -avn_ni+z
+ b.ni+2vn_ni+2
> §n_ni+2,
(6.6.31)
where #_ ni +2 represents the right side of the inequality (6.6.30). In order to repeat the above procedures, we need to make sure that b-m+2 is positive. In fact
b-n1+2 - \{b2 ~™)>\
[(1 + OJ)2 - ^
> °-
Then combine the 3rd-to-the-last equation in (6.6.25) with (6.6.31). And repeat the procedure in a similar way. In general, use above elementary operation to obtain the inequality : -avl+l
(6.6.32)
+ bkvl > 61
and as will be verified soon, bk > 0 . Then combine it with the n\ + k + 1-to-the-last equation in (6.6.25), i.e. consider the inequality (6.6.32) and -awjfc+2 + K + i - <*>k > W+i • (6.6.32) x f + (6.6.33), we get (10
-*
-a< + 2 + ( 6 - - K + 1 > 0 f c + 1 (here we make use of -£- > 0), where an nn i c an ^fc+1 - ^fe+l + b~"fc-
(6-6.33)
174
Mathematical Modeling and Methods of Option Pricing
Denote ac k Ok+l = O- —,
A
If we can prove all Bm defined by Bm+1
=b-—
ac
(1 < m < M)
(6.6.34)
and (6.6.35)
B1 = b
are positive, in view of the the same results as mentioned in Theorem 3.5, then we can conclude: the discrete scheme of the variational inequality (6.6.25)—(6.6.27) is equivalent to the variational inequality
{
A V n > 0n,
(6.6.36)
Vn > # ,
(6.6.37)
(AV"-0"),(V"-$),=O,
( - n i + l < j < 712-1)
(6.6.38)
where A is the lower triangular matrix "6 n 2 _!
0
~a
A= .
,
0 fln=
(6.6.39)
-a6_ni+i. : an
L e m m a 6.8 Ifb2-4ac following solution:
> 0, then equations (6.6.34),(6.6.35) have the
Bm = - ^ - ,
(2<m<M)
(6.6.40)
where Ji = b,
(6.6.41)
American Option Pricing and Optimal Exercise Strategy
175
m
Jm = £ ( « + ) - - > _ ) ' ,
(m > 2)
(6.6.42)
i=0
and b ± \/b2 - Aac Proof
,„..,, (6.6.43)
.
a± =
Multiplying the difference equation (6.6.34) by Bm, we get Bm+1Bm
= bBm - ac.
(m > 1)
Then multiplying both sides by Bm-\ ...B\, and from (6.6.40),(6.6.41) and (6.6.35), we obtain the difference equation for Jm = Yl'tLi BiJm+i - bJm + acJ m _i = 0 ,
(m > 2)
(6.6.44)
with the initial conditions (6.6.45)
Jx = b,
J 2 = JxBi = b(b - ^ ) = b2 - ac. (6.6.46) o In order to solve the difference equation (6.6.44) under the initial conditions (6.6.45),(6.6.46), let T
—p
m
Substituting it into (6.6.44), we get £2 -b£ + ac = 0. This quadratic equation is called the characteristic equation of the difference equation (6.6.44). The two roots are C=
Q±
=
b ± y/b2 - Aac j •
Thus the general solution of (6.6.44) has the form Jm = A(a+)m + B(a-)m. Applying initial conditions (6.6.45),(6.6.46), we get a+A + a-B = b, a\A + c?_B = b2 - ac.
(6.6.47)
176
Mathematical Modeling and Methods of Option Pricing
And solve them to determine Q
A-
+ V 6 — 4ac
R-
~Q-
2
V 62 — 4ac
Substituting them into (6.6.47), we get Jm = - r = = [ ( " + ) m + 1 ~ ( « - ) m + 1 ] yb2— 4ac m
__ a + -a_ y^
^
V o — 4ac j _ 0 m
i=0
Thus (6.6.42) is proved. And by definition of J m ,
B™—.
(m>2)
Thus proves the lemma. Now let us check the condition of Lemma 6.8 b2 - 4ac > 0. By definition, 6 = 1 + u> + r At,
a = | + fly/At, c = | - /?V£*. Thus the condition is
b2 - Aac = (1 + to + rAtf
- 4 (^- - $2At\ > 0,
American Option Pricing and Optimal Exercise Strategy
177
where
a2At Aar
w = -7—5-.
Since ac
w2 [
Az 2
a2 21
Thus if
then a+ > a_ > 0, and by (6.6.41),(6.6.42), we have (1 < m < M)
Jm>0, and Bm = ^ - . Jm-l
(2<m<M)
Thus we have proved: all elements bk (—rai + 1 < k < ni — 1) on the main diagonal of the lower triangular matrix A are positive. Theorem 6.19 The discrete scheme of the variational inequality (6.6.36)—(6.6.38) has solution
i£a_i = m a x ( j - i - ^ _ l l ¥ ) B 2 . 1 } 1 L°n2-1
J
^" = m a x | ^ - [ ^ + a ^ + 1 ] , ^ | , ( - m + 1 < j < n 2 - 2).
(6.6.48)
(6.6.49)
178
Mathematical Modeling and Methods of Option Pricing
Proof Let us first examine the first row of (6.6.36)—(6.6.38), i.e. for j = n 2 - 1 case: f*W-ii£a-i>^a-i. < < 2 - l > ¥>n 2 -l,
I (&n 2 -l^ 2 -l - t2-l)K2-l
~ ¥>na-l) = 0.
Rewrite them in the following form minK 2 _! - 7 o
n2-l
" 2 -lX 2 -l - Vn2-l} = 0,
then the solution can be expressed as follows: < , _ ! = max \
0™
1;
Substituting this result into the 2nd row of (6.6.36)—(6.6.38), i.e. for the case j = n^ — 2, similarly, we get: mM-aitfa-l + fon2-2<2_2 " fl"2-2.«na-2 ~ Vn2-2} = 0. Since w"2_i has been found, thus we have
<2-2 = max {-J— [^2_2 + a ^ . J , ^ ^ } .
lOn2-2 J Repeating the procedure, and using backward induction, we can obtain (6.6.48),(6.6.49). Thus proves the theorem. If 1 — i ( r — ^r)\/At > 0, we can similarly prove the convergence of the approximate solution obtained by the implicit difference scheme, when At, Ax—>0. In contrast to the explicit difference scheme, we do not need to 2 A
assume u> =
6.7
a
^
2
< 1.
Numerical Methods(II)
Line Method
Consider non-dividend-paying American put option modeled as a free boundary problem, i.e. to find {V(S, t),S(t)}, such that in region Si : {S{t) < S < oo,0 < t < T} they satisfy the following free boundary
American Option Pricing and Optimal Exercise Strategy
179
problem:
= O, (E)
'%T + ^fy+rSW-rV
(6.7.2)
V(S(t),t) = K-S(t), (
(6.7.3)
Vs(S(t),t) =-I,
(6.7.4)
V(S,T) = (K-S)+, (S^oo)
V(S,t)-*0,
(6.7.1)
(6.7.5)
S(T) = K.
(6.7.6)
Since obviously we can normalize the strike price K by transformation S = -jr,V = jr,S(t) = ps', without loss of generality, we can assume K = \. Divide [0,T] into N equal intervals: O = t
o
< t
1
<•••
=T,
T tn = nAt,At= —. At each line t = tn, define Vn(S) = V(S,tn) and
Sn=S(tn),
so that they satisfy a system of free boundary problems obtained by discretizing (6.7.1)—(6.7.6) in time t, i.e. to find {Vn(S),Sn},(n = 0,1,...,N), so that
' vn+,(s^-vn{s)
+
^ d ^
+r 5
^ _ r V n = Oj
(677)
(5 n < 5 < oo) Vrn(Sn) = 1 - Sn, ' ^{Sn)
= -l,
K ( 5 ) - • 0,
(S -> oo)
Viv(5) = ( l - S ) + , . 5jy = 1, where 0 < n < TV — 1.
(6.7.8) (6.7.9) (6.7.10) (6.7.11) (6.7.12)
180
Mathematical Modeling and Methods of Option Pricing
71
']
*—rr-**
ij—1—1—i
Vi
•
....
. s
The algorithm for solving the above system of free boundary problems is as follows: (1) Since VN(S) = (1 - S)+, (0 < S < 00) SN = 1. Thus we have found the value of {Vn(S), Sn} when n = N. (2) By induction, if {Vn+i(S), S n +i} is given, and Sn+i < 1, define ( Vn+1(S), Sn+1 < S < 00, Vn+i(S) = I [l-S, 0<S<Sn+1. Obviously Vn+1 (5) €CJ,>OO). (3) In region {Sn < S < 00}, solve the following free boundary problem to second order ODE: CAVn = ^ S
2
^
+ rS^
- (r + -£j)Vn = ^Vn+1(S),
(6.7.13)
(Sn < S < 00) - Vn(Sn) = l-Sn,
(6.7.14)
^ f (Sn) = - 1 ,
(6-7.15)
Vn(S) -> 0,
(5 -> 00)
(6.7.16)
to get{Vn, .!?„}. And repeat the procedure. (4) There are many methods to solve the free boundary problem (6.7.13)—(6.7.16). For this special case, we can obtain a closed-form solution of equation (6.7.13). In the following, we will explain how to find Vn and 5n([6]) by solving the equation directly. How can we find VAT-I(S) and SJV_I? (6.7.13) is a nonhomogeneous second order ODE. First we need to find the general solution of the corre-
American Option Pricing and Optimal Exercise Strategy
181
sponding homogeneous equation: let
Vn(S) = Sa. Substituting it into (6.7.13), we get K{a) = ^-a(a - 1) + ra - (r + -^-) = 0.
(6.7.17)
Equation (6.7.17) is called the characteristic equation of the ODE (6.7.13). It has two roots: p where _ 1 7
2
~2~^' I,
^ V ^
r
0^~~
2^2
+ T^AT
Thus a+ > 0 >
Q_.
Thus homogeneous equation corresponding to (6.7.13) has a general solution of the form:
V(S) = d1Sa+
+d2Sa-.
From properties of the optimal exercise boundary, we can assume SjV-i < 1- Then the right side of equation (6.7.13) is a piecewise polynomial, i.e. equation (6.7.13) can be written as
{
£AV/V_I = 0 ,
(1 < S < oo)
(6.7.18)
C*VN-! =—^(1 - S), (SN^<S<1) (6.7.19) V/v-i, V^_x are continuous at 5 = 1. (6.7.20) Since the inhomogeneous equation (6.7.19) has a special solution V= S
TTr^i- >
182
Mathematical Modeling and Methods of Option Pricing
hence the general solution of the equations (6.7.18)—(6.7.19) is f
f 4 1} S Q + + d£]Sa- + l
V}v-i(5 ) = <
]A
- S, (SN-i < S < 1) (6.7.21)
[ d{^Sa+ + 4 2 ) 5 Q - ,
(1 < S < oo)
(6.7.22)
where the constants d\ (i,m = 1 , 2 ) and SAT-I a r e to be determined by the boundary conditions (6.7.14)—(6.7.16) and the consistency conditions (6.7.20). By (6.7.16), we find
<42) = 0;
(6.7.23)
42)=41)+41)-T^?
(6-7.24)
a _ 4 2 ) = a+d^ + a _ 4 1 } - 1;
(6.7.25)
By (6.7.20), we find
By the free boundary conditions (6.7.14), (6.7.15), we get (6.7.26)
(6.7.27)
By (6.7.24),(6.7.25), we find
2±#>=4»>_4i> + i/ a _. Thus d{1)(a+-a_)/a_ = ^ - -
I
^ ,
i.e. (6.7.28)
183
American Option Pricing and Optimal Exercise Strategy
By (6.7.26) x a_ — (6.7.27) x SN-I, we obtain the equation for the free boundary point SN-I-
i.e. (6.7.29)
It is easy to see: 0 < SV-i < 1- Substituting d[ ',SV-i into (6.7.27), we find d (l) =
_^+ r f (l) 5 a + _-a_ _
(6 73Q)
Then substituting d^ and SP into (6.7.24), we get d^. found {VN^{S),SN^}: ' 1 - S, VN^(S) = J S^ga*
Thus we have
(0 < S < SN-!) + dWSa.
+
^^
df]Sa~,
^ _ g^
(5jy_1
< S < 1)
(1<5
where d{1},d{21},d^ and SN^ are defined in (6.7.28),(6.7.30),(6.7.23) and (6.7.29), and they can be determined by the following procedure:
Since in [0,oo), VN-I(S) consists of a piecewise function, therefore in order to find a special solution of the equation with this function as a nonhomogeneous term on the right side, we need the following lemma. Lemma 6.9 Consider the nonhomogeneous ODE: (6.7.31) where UJ is the single root of the characteristic equation (6.7.17), i. e. K{w) = ^-LJ(CJ _ l) + ro, _ ( r + i _ ) = o,
K\UJ)
? 0;
(6.7.32)
184
Mathematical Modeling and Methods of Option Pricing
Then the equation (6.7.31) has a special solution
Proof
Substituting the expression of W into equation (6.7.31), we
get
£ w
- =7^M^-i)+™-ir+lt»s"ins + (^(2o;-l) + r)Su]
= A«k0 ( £ ( a W - 1 ) + r)fl" "AT '
Now we look for
{VN-2{S),SN-2},
' £AVN_2 = - ^ 4 2 ) 5 " * ~ .
which satisfies
(1 < S < °°)
£ A y JV _ 2 = - ^ (4 1 ) 5 Q + + ^ S 0 - ) - ( T + ^ J
(6.7.33) - 5)^,
(6.7.34)
(SJV-I < S < 1)
£ A y JV _ 2 *
=
-2li(l-5),
V/V_2,^JV_2
are
{SN-2 < S < SN-!)
continuous at 5 = 1,
(6.7.35) (6.7.36)
V/v_2, ^y_ 2 are continuous at 51 = Sjv-i)
(6.7.37)
^ - 2 ( 5 ^ - 2 ) = 1 - SiV-2,
(6-7-38)
V^_ 2 (S^_ 2 ) = - 1 ,
(6.7.39)
Vlv-2(5) -> 0.
(6.7.40)
(5 -> oo)
Equation (6.7.35) has a special solution Wi(S) = T^-rS.
(6.7.41)
185
American Option Pricing and Optimal Exercise Strategy
And by Lemma 6.9, equations (6.7.33), (6.7.34) have special solutions
+
(67 42)
-
(TTW~5'
(6.7.43) Therefore in region {S^ _ 2 < S < oo}, the general solution of equations (6.7.33)—(6.7.35) is ' Wi(S) + 4 1 ) 5 a + + d21}Sa= • W2(S) + ^ 2 ) S a + + d[2)Sa-
VN-2(S)
W 3 (S) + ^ S ^ + S23)Sa-
{SN-2
<S<
(SJV-I
< S1 < 1)
(1 < 5 < oo)
SJV-I)
(6.7.44) (6.7.45) (6.7.46)
The six coefficients d\m\m = 1,2,3, i = 1,2) and the free boundary point SN-2 will be determined from the four consistency conditions (6.7.36), (6.7.37), two free boundary conditions (6.7.38), (6.7.39) and a condition at infinity (6.7.40). By (6.7.40), we find ^ 3 ) = 0.
(6.7.47)
W2(l) + d<2) + d22) = W 3 (l) + 4 3 ) ,
(6.7.48)
W£(l) + a+df] + a_4 2) = W£(l) + a_4 3) -
(6.7.49)
By (6.7.36), we get
Since W2(l) =
^
2
- 1, W 3 (l) = 0, thus from (6.7.48), (6.7.49), we
find
d?) =
J {Q- [ ( T T W " ' ] " m { 1 ) " ^ ( 1 ) ] } • (6-7-50)
Prom (6.7.37), we get
w1(sN.1) + d^s^t, + 4x)^"-i = W2(SN-!) + df )S^+_1 + 42)S°-_1I (6.7.51)
186
Mathematical Modeling and Methods of Option Pricing
= ^(Siv-i) + a+^S^Ll1 + a-^SftLl1, Eliminating d2 and d2', we find
;(1) _ -(2) O-IW^Ar-x) - W2(SN-1)} -
- W^(SN^)]
SN-JIWUSN-!)
b
P N-l
(6.7.52) By free boundary conditions (6.7.38), (6.7.39), we get WI(SAT-2) WI'(SJV- 2 )
+ 4 1 ) ^ - 2 + ^ S ^ = 1 - SJV-2,
+a + 4 1 ^ ^
1
+ a-d£] SaN-J2l = - 1 .
Eliminating d21], and noticing Wi(5) = obtain the equation for SN-2'
1 x+ rA(
(6.7.53)
- 5 , ^ ( 5 ) = - 1 , we (6.7.54)
Substituting (6.7.50), (6.7.52) into (6.7.54), we get a transcendental equation for Sjy-2- It is easy to see, S^-2 < S^-iBack to (6.7.53), we can determine d2 • Since d\ ,d\ ,d2 have been determined, we can find d2 from (6.7.51), then substituting it into (6.7.48) to determine d2 . By now we have determined all coefficients d™ (m = 1,2, 3, i = 1,2) and S;v-2- The procedure goes as follows: j(3)
Uj
.
."(2)
—> Ci
,-(1)
—> Ct^
T(l)
„
—> ON —2 —> &2
1(2)
— ""2
7(3)
— ""2 •
Let us use the backward induction. Suppose when t = tn+i, Vn+i(S) £ CL , and Sn+i(n > 0) are given. The function ^1+1(5) can be defined in the following way. Divide [0,00) into N — n subintervals: [0,5 n+ i],...,[5 Ar _i,l],[l ) oo). In0<5<5n+i,
Vn+1(S) = 1 - S = M^S); In [S'n+i,©©), function Vn+i(S) is defined as follows. [Sn+j-l,Sn+j\, Vn+1(S) = Mj(S),
j =
2,...,N-n+l,
In interval
187
American Option Pricing and Optimal Exercise Strategy
where SN = 1, SN+I
= oo,
n-2
n-2
i=l
i=l
Mj(S) = J2 c\j)Sa+ (inSy + ^ 4J)Sa~ (inSy + e(j)S + f{j). Here cf ,af ,e^\f^ are constants, and a+,a- are the roots of the characteristic equation (6.7.17). Thus in order to find the values of Vn(S) and Sn at t = tn, we first rewrite the free boundary problem (6.7.13)— (6.7.16) as CAVU = - ^ M i ( 5 ) ,
(Sn < S < Sn+1)
(6.7.55)
CAVn = —^M2(S),
(Sn+1 <S< Sn+2)
(6.7.56)
£&Vn = —^MN-n(S),
(SN-i < S < 1)
(6.7.57)
CAVn = --^MN^n+1(S),
(1<5
(6.7.58)
and Vn(S), Vn(S) are continuous at S = Sn+i,
(6.7.59)
Vn(S), V^(S) are continuous at S = 1,
(6.7.60)
and Vn(Sn) = l-Sn,
(6.7.61)
V;(Sn) = - 1 ,
(6.7.62)
V r n (oo)=0.
(6.7.63)
188
Mathematical Modeling and Methods of Option Pricing
Vn(S) as a solution of the free boundary problem (6.7.55)—(6.7.63) can be expressed in the following form: ' Wx (5) + # > Sa+ + 4 1 } Sa-,
(Sn<S<Sn+1)
Vn(S)=< WN.n+1(S)
+ d[N~n+1)Sa+ + d{2N-n+1)Sa~, (1 < S < oo)
where Wi(S)(i = 1,..., N - n + 1) is the special solution (see Lemma 6.10) to the nonhomogeneous ODE (6.7.55)—(6.7.58) with --^Mi(S) at the right side, respectively. The 2(iV — n + 1) unknown constants dj(m = 1,..., N — n + 1; j = 1,2) and the free boundary point S = Sn can be determined by the consistency conditions (6.6.59),(6.7.60) at 5 = 1 , 5JV-I, ..., 5n_|_i, the free boundary conditions (6.7.61)(6.7.62), and the condition at infinity (6.7.63), altogether 2(7V - n + 1) + 1 equations. They can be determined by the following process: J(n+1) _> J(n) _ > . . . _ , J(l) ^ ^ _> JU) ^ JW ^ . . . ^ J(n+1) The actual procedure is rather tedious, but all operations are explicit, and there is no need to solve a large system of equations. This is an advantage of this method. Next we show how to obtain the special solution to the nonhomogeneous ODE with 5 a ± (In S)j at the right side. Lemma 6.10 Consider nonhomogeneous ODE
where w is a root of the characteristic equation (6.7.17), i. e. K{w) = yw(w - i) + rw - (r + ^ ) = 0, K'(LJ) / 0.
It has a special solution H/(5) = ^C,S" i '(ln5) i , i=l
189
American Option Pricing and Optimal Exercise Strategy
where C j ( l
+ l) are given by the following recurrence Ci=
2K'(CJ)
Cj+1 =
Proof
(i =
°i+1'
formula:
1 2
' .---'j)
v + i)K\uy
By straightforward calculation, 2
CziSrQnS)*) = iK'^STQuS)*-1 + —i(i - l)(lnS^S", thus
cA(w) = cA
lJ2CiSU>(lnSY)
= C7 + l)Cj +1 Jf / (a;)(ln5) J 'S a '
+ £ i [^'(wjCi + ^-(i + l)Ci+1l (lnS)'-1^-. i=i
L
z
J
To satisfy the equation £A(Lj) = S"(\nSy, there must be (j + 1)CJ+1K'(UJ)
= 1,
/jr'(a;)Ct- +°—(i + l)Ci+1 = 0 . Q.E.D. 6.8
Other Types of American Options
(A) American Binary Option
(1 < i < j)
(* > 1)
190
Mathematical Modeling and Methods of Option Pricing
American binary option: during the contract's lifetime, whenever the underlying asset price St exceeds the strike price K, the contract can be exercised to receive $1 in cash. The math model for this type of American option is not a free boundary problem, since the underlying asset price at the exercise time is fixed: St = K. Let V(S, t) be the option price, then it satisfies: lfr + !Ts2J^
+ (r-
= 0,
(0 < S < K)
' V(K,t) = 1,
(6.8.1) (6.8.2)
(0 < S < K)
V(S, T) = H(S -K) = 0.
(6.8.3)
Let (6.8.4)
V(S,t) = V0(S)-u(S,t),
where Vo(S) is the price of the perpetual American binary option, i.e. Vo(S') satisfies i^S2^
+ (r-q)S^-rV0
(0 < S < K)
= 0,
\ V0(K) = 1.
(6.8.5) (6.8.6)
Similar to discussion in §6.1, let
K)
'
+B
{K)
^
where a± is the root of the characteristic equation: 2
(6.8.8)
—a(a-l) + (r-q)a-r = 0, i.e. ^ _ -
(
, - , - ^
) ±
^ - , - ^
Thus a+ > 0 > a-. Since V0(0) is finite, B = 0.
+
^
( 6 8 9 )
191
American Option Pricing and Optimal Exercise Strategy
And by boundary condition (6.8.8), A = l. Thus, denoting a+ = a, the solution of (6.8.5)—(6.8.6) is of the form
Fo=(£)Q.
(6.8.10)
Substituting it into (6.8.4), we obtain the initial-boundary value problem for function u(S, t) = VQ(S) - V(S, t): W
+ g S2
T W
+ {r
~ 9)S^~ru
= 0,
(0<S
(6.8.11)
< u(K,t) = 0 , u(S,T)= (j^y
(6.8.12) •
(6.8.13)
(0<S
Change the variable to x = ln—.
(6.8.14)
Then u(x, t) satisfies the one side bounded problem:
f + ^ 0 - ( - ^ - ^ ) i - - = O,(O<,
(0
u(x,T) = e-ax.
(0<x
Suppose u(x, t) can be written as u = e«*+HT-t)Wf where 1 , a
1 2, i
(6.8.15)
192
Mathematical Modeling and Methods of Option Pricing
Then W(x, t) satisfies
•fK
+
£ g f = O,(O<x
< W(O,t) = O,
(0
W{x,T) = e -("+°K (0 < x < oo) Using the image method, we perform an odd extension by defining an odd function
x > Q j
+^ ^ = 0 ,
\ W{x,T) =
(a;ei?,0
(6.8.16)
(x G R)
(6.8.17)
Since
1
f°° /
V-oo
(*-o2 e **(T^
r°° r
(f+o2 i
<7v/27r(r - t) Jo L
, _, >,
J
Substitute it into (6.8.15) to get u(x,t). Then go back to the original variable to get u(S,t). Then substituting into (6.8.4), we get the valuation of American binary option. In particular if q = 0, then a = 1, and
VW)=(§)^W(d3)+(§)tf(di), where
Inf+ (r-g)(r-t) "l =
/rr
.
cry 1 — t
.
(b.8.18)
193
American Option Pricing and Optimal Exercise Strategy
4 =
lnf-(r-j)(T-Q
(6819)
(JyT — t
The derivation is left to the reader. (B) Bermudan Option Bermudan option: early exercise is allowed at certain predetermined times during the lifetime of the contract, and no early exercise allowed, just like European option, at other times. Let 0 < t\ < • • • < tM = T be the allowed early exercise times for a put option. Then
{
VN(S,t),tN-i
:
Vi(S,t), 0 < t < t i .
and in each region {0 < S < oo, U-\
\ Vi(S,ti) - max{Vi+1(S,U), (K - S)+},
(i =
l,...,N)
where VN+1(S,tN) = (K - S)+. (C) American Capped Call Options Add the following term to an ordinary American call option contract: when the underlying asset price exceeds the upper limit L(L > K) as specified in the contract, the option writer has the right to purchase back the option at the price L-K. The motivation behind this term is obvious: the option writer wants to manage the risk inherent in selling American call options. For American call options with above terms, its payoff function is payoff =(min(S t , £ ) - # ) + , Hence its mathematical model is a variational inequality given in region
194
Mathematical Modeling and Methods of Option Pricing
Y,L{O<S
{
min{-£K, V - (S - K)+} = 0, (S L ) V{L, t) = L-K,
(0
V(S,T) = (S-K)+.
(0<S
(D) American Options with Warning Time (Make Your Mind Up) The contract requires a warning period before the option is exercised. Once the decision is made, the option holder cannot change mind, i.e. the holder must exercise the option upon expiration of the warning period in any circumstance. Suppose the holder makes a decision at to and the warning time is ToThen at to + TQ , the payoff for the holder of the American call option is payoff
=S-K.
Note that there is no "positive taking" operation in the above function. This indicates that the holder must exercise the contract even if the asset is out of the money. Let V(S,to + T)(0 < T < T0) be the price of the American call option with warning time, then during the warning time, V satisfies the initial value problem to the Black-Scholes equations:
/ W+ 1 T 0
+ (r
" q)S^5 ~ rV = °' (° " S < °°'° " T " ro) (0 < S < oo)
\ V\T=T0 = S-K. It has a solution V{S,t0 + T) = Se-q(T0-^ -
Ke-r(T°-T).
Hence at T = 0 when the holder makes the decision, the intrinsic value of the option is V{S,t0) = Se-qT0
-Ke~rT0.
Then, the price of the American call option with warning time To must satisfy V{S,t) >Se~qT0
-Ke~rT0.
American Option Pricing and Optimal Exercise Strategy
195
Based on the above analysis, we can formulate the problem as a variational inequality in domain £{0 < S < oo, 0 < t < T}: f min{-£V, V - (Se-*T° - Ke-rT°)} = 0, (£) \v(S,T)
(0<5
= {S-K)+.
Summery 1. The main feature of American option is its early exercise right. During the lifetime of the option, the holder needs to watch closely the asset price to decide whether to hold or exercise the option. According to whether the option's intrinsic price exceeds or equals the early exercise payoff, the domain of V(S, t) is divided into two regions: the continuation region in which V(S, t)>(S-
V(S, t)>(K-S)
K)+,
+
;
(call option)
(put option)
and the stopping region in which V(S,t) = (S-K)+,
(call)
V(S,t) = (K-S)+.
(put)
The interface is called the optimal exercise boundary. Holders of American options need to know "picture" to have an optimal exercise strategy. 2. The math model of American option pricing is a free boundary problem the obstacle problem, in which the obstacle is the payoff function of the option, the coincidence set of the solution and the obstacle is the stopping region of the option, and the boundary of the coincidence set is a free boundary, which is the optimal exercise boundary of American option. This mathematical model has two equivalent formulations: a. Variational inequality. In this form, the solution domain is the whole region S = { 0 < 5 < o o , 0 < t < T}, and the free boundary is implicit in the problem statement.
196
Mathematical Modeling and Methods of Option Pricing
b. Free boundary problem. In this form, the solution domain is restricted to the continuation region Si : Si = {0 < S < S(t), 0
(call)
S x = {S(t) < S < oo, 0 < t < T}.
(put)
The free boundary S = S(t) as the boundary of the problem appears explicitly in the problem statement. 3. As a nonlinear PDE problem, American option price in general cannot be given in a closed form, except for the perpetual American option. Therefore, the study of the properties of the optimal exercise boundary is of great importance for American option pricing. Let Sc(t),Sp(t) and 5CiOO,5PiOO be the optimal exercise boundaries for ordinary American call, put option and perpetual American call, put options with the same strike price, respectively. Then for the American call option, we have: Sc(t) monotonic decreasing, convex; rK max( — ,K) = SC(T) < Sc(t) < 5CiOO, q and its asymptotic expansion at t = T: •K(l + ay/(T-t)\ln(T-t)\), Sc{t)
q > r,
~< K(l + v W C T " t)\HT ~ *)|), Q = r, rK^ + S^^/T^i),
For the American put option, we have Sp(t) monotonic increasing, convex; rK SP,oo < Sp(t) < SP(T) = min( — ,K),
q
American Option Pricing and Optimal Exercise Strategy
197
and its asymptotic expansion at t = T: • K(l - ay/(T-t)\\D(T-t)\), 5j,(t)«,
K(l-V2ay/{T-t)\ln(T-t)\), _ l£(l - 2|Vr=i),
q < r, q = r, q>r,
where a is the root of the transcendental equation
a V 2 / 4 fX e-^'Uuj = 2(2 - a2). Ja
If the optimal exercise boundary S = S(t) is known, or approximately determined, then American option price V(S,t) can be determined or approximately given by (6.3.14)—(6.3.16). 4- American option price in general can only be obtained by numerical methods. Two methods are introduced in this chapter: a. The difference method (explicit and implicit schemes); b. The line method. As a natural extension of §5.7, we have shown that when neglecting the higher order terms, the BTM of American option pricing (described in §3.5) is equivalent to a special case of the explicit difference scheme, thus provides a foundation for proving the convergence of the BTM of American option pricing in the framework of PDE. 5. While the call-put parity does not hold for American options, there exists the American call-put symmetry Vc(S,t;r,q) =
^Vp(^,t;q,r)
and
yJsc(t;r,q)Sp(t;q,r)=K, where Vc, Vp and Sc, Sp are the prices and the optimal exercise boundary of American call option and put option, respectively.
Exercises 1. Consider the dividend-paying (dividend rate q > 0) perpetual American call option. Write out its mathematical model and solve it directly, to find its premium and the exercise boundary.
198
Mathematical Modeling and Methods of Option Pricing
2. Study how the perpetual American option price is affected by the volatility a, and explain the financial meaning. 3. American straddle option is a contract with early exercise term, and its payoff function is payoff = \S - K\
= {S-K)+
+ (K -S)+.
Set up a mathematical model for the dividend-paying perpetual American straddle option, and find the premium and the optimal exercise boundary. Moreover, can its premium be represented as the sum of a perpetual American call option and a perpetual American put option? Why? Compare their premiums. 4. Write out the mathematical model for the perpetual American call option with warning time, and find the premium and the optimal exercise boundary. 5. Prove(Theorem 6.2)American call-put symmetry: Vc(S,t;r,q) =
^Vp{^-,t;q,r)
and y/Sc(t;r,q)Sp(t;q,r)
= K,
where VC(S, t; r, q), VP(S, t; r, q) and Sc(t; r, q),Sp(t; q, r) are the price and the optimal exercise boundary of dividend-paying American (call, put) options with the same expiration date T and strike price K, respectively. 6. Show that inequality (6.4.25) is true. I.e. if q\ > 92, the American put option price V satisfies V(S,t;qi)>V(S,t;q2), If V(S, t; q) is the American call option price, how will the above relation be modified? 7. Show that for the American capped call option, the optimal exercise boundary S(t) is S(t) = min(L,S(i)), where L is the upper limit specified in the contract, S(t) is the optimal exercise boundary for the American call option without cap. 8. Find the premium and optimal exercise boundary for the perpetual American capped call option. 9. Show that among all American call options with the same strike price, the perpetual American call option is the most expensive one. I.e. if Vr(S, t) is the price of American call option with expiration date t = T, Voo(S) is the perpetual American call option price with the same strike price, then for all T > 0, there must be VT(S,t)
American Option Pricing and Optimal Exercise Strategy
199
here Sr(t) and S^ are the optimal exercise boundary for American call option and the perpetual American call option with the same expiration date T, respectively.
Chapter 7
Multi-Asset Option Pricing
The price movement of two or more risky assets can be described by a system of stochastic differential equations. Options derived from two or more underlying risky assets are called multi-asset options. Multi-asset option price satisfies a multidimensional parabolic partial differential equation. Different types of multi-asset options are distinguished by their payoff functions. Hence a given multi-asset option pricing problem can be modeled as a terminal-boundary value problem to a multidimensional parabolic PDE with appropriate terminal condition. Since the dimension of the PDE is determined by the number of the underlying risky assets, which can be very large, even if an explicit solution can be obtained, it will still be very difficult to evaluate the expression. Thus, an important question in multiasset option pricing is whether a given multi-asset option pricing problem can be reduced to a 1-D problem by introducing suitable composed variables?
7.1
Stochastic Models of Multi-Assets Pricing
In order to price multi-asset options, we need first to establish the price movement model for the underlying multi-assets. Let St be the price of the i-th risky asset (i = 1 , . . . , n). Suppose Si satisfies a stochastic differential equation (SDE) of the following form: —i = indt + ciidWu
(t = l , . . . , n )
(7.1.1)
where dWi (i = 1 , . . . , n) are standard Brownian motion which satisfy E(dWi) = 0, 201
(7.1.2)
202
Mathematical Modeling and Methods of Option Pricing
Var(dWi) = dt,
(7.1.3) (» ^ j)
Cov(dWi,dWj) = pijdt.
(7.1.4)
Here Cov(-, •) denotes the covariance. By the definition of covariance and (7.1.2), Cov{dWi,dWj) = E([dWi - E{dWi)\[dWj - E(dWj)]) = E(dWidWj),
(i^j)
The prices of multi-assets can also be modeled as: dS-i
T—^
(i = l , . . . , n )
-=± = (iidt + y2<TijdWj,
Si
U
(7.1.5)
where dWi(i = 1,... ,n) are one-dimensional Brownian motions, i.e. E(dWi) = 0,
(7.1.6)
Var(dWi) = dt,
(7.1.7)
Cov(dWi, dWj) = 0 ,
(i ^ j)
(7.1.8)
SDE (7.1.5) can be written in vector form:
dS = adt + [a]dWt, where "5i"| S=
:
rwSi"| ,5=
.Sn\
:
[Wlt,Wt=
lUnSn]
(TnSi
.PnlSn
. . . (TlmSi
•••
GnmSn.
: [Wmt_
,
203
Multi-Asset Option Pricing
here "5i
0"
.0
V
[S] =
C l l . . . <7i m
M =
;
: •
7.2
Black-Scholes Equation
Let Si,..., Sn be n risky assets (e.g., stock, foreign exchange rate,...), satisfying geometric Brownian motion (7.1.5). Let V be an option derived from the underlying assets Si,..., Sn, as a function of n + 1 variables Si,..., Sn and t: (7.2.1)
V = V(S1:...,Sn,t).
A-hedging principle Choose A, shares of asset 5,(i = 1,... ,n) to hedge against the option V(S\,..., Sn,t), i.e. to construct a portfolio II: n
U = V(Si,...,Sn,t)-^2AiSi,
(7.2.2)
i=\
such that it is risk-free in (t, t + dt). From the Ito formula for the multivariate stochastic process, we have n
n
A dS
dU = dV - J2 i i i=l
IdV n
- 13
A S
i i1idt
» = 1
^y_\
1A «,,
n
n
+ J2 dS~dS* - E AidSi - E ^^ftdt,
(7.2.3)
204
Mathematical Modeling and Methods of Option Pricing
where
aij = Y^Vik<7jk, fc=i
= l,...,n)
(i,j
i.e. A = [dij] =
where <JQ is the transpose of matrix Uo. In order that II is risk-free in (t, t + dt), i.e. n
dU = rUdt = r(V - ^ AiSJdt,
(7.2.4)
A, = ££.
(7.2.5)
we choose
Substituting (7.2.5) into (7.2.3),(7.2.4), and eliminating dt, we get (7.2.6) ij=l
^
i=l
This equation is called the Black-Scholes equation for multi-asset options. Since the quadratic form T
V
AV
= rr
VT? e
Rn
thus A = [ciij] is a symmetrical nonnegative matrix. Therefore (7.2.6) is a multidimensional parabolic equation. Denoting the option payoff function at maturity (t = P(Si,..., Sn), then the mathematical model of the European asset option is: solve PDE (7.2.6) in domain E : {0 < 5, < 1,... ,n);0 < t < T} with the terminal condition V(Su...,Sn,T) 7.3
= P(Slt...,Sn).
equation T) by multioo(i — (7.2.7)
Black-Scholes Formula
The terminal value problem to multidimensional Black-Scholes equation (7.2.6),(7.2.7) has a solution.
205
Multi-Asset Option Pricing
By transformation (7.3.1)
x{= In Si, (7.2.6) becomes 1 .A d2V 1- - > an flt 2 ^ t3dxidx4
A*. 1- > (r-qi *^K
dV
i,j=l
J
¥,
au.dV -) 2'dxi
n
rV = 0.
._ „ n. (7.3.2) y '
i=l
In vector form, ?Y-+l-SJTxAVxV + brVxV-rV
= Z,
(7.3.3)
where A = [oij], >
- ?i -
6=
^
'
:
,
(7.3.4)
j-qn-^T. r
9
i
a \-dxn
J
and b1^ is the transpose of matrix (vector) b. By transformation z = Bx, Z\ "I : .•Zn J
^11 . . . 6ln =
:
^1
:
\_bnl • • • b
n n
: \
,
(7.3.5)
]_Xn_
where B is chosen such that under the new coordinates (z\,... ,zn), the coefficients of the second order derivative term in equation (7.3.3) are diagonalized, and due to Vx = BTVZ, equation (7.3.3) becomes ^
+ i VTZ{BABT)VZV
+ (Bb)TVzV -rV = 0.
(7.3.6)
206
Mathematical Modeling and Methods of Option Pricing
Since A is a symmetric matrix, there must exist an orthogonal transformation B such that BABT is a diagonal matrix, i.e. there exists transformation (7.3.5) satisfying BTB = BBT = I,
(7.3.7)
such that 'Ai T
BAB
0 ' •-.
= A=
.0
,
(7.3.8)
V
where X\,..., Xn are the eigenvalues of the symmetric matrix A, which are the real roots of the characteristic equation of A: det \A - XI\ = 0.
(7.3.9)
Suppose £j is the eigenvector corresponding to the eigenvalue Aj, i.e. (A - XJ)ii = 0,
(7.3.10)
and
&& = U2 = i. Then by transformation (7.3.5), with
B=
£i : Xn J
£n £12 • • • £in : : ,
=
(7.3.11)
L^nl Cn2 • • • £nra.
equation (7.3.6) will be reduced to
% + \ p % + ±M%-rV-B.
,,,12,
Therefore under transformation (7.3.5), the terminal value problem of the Black-Scholes equation (7.2.6),(7.2.7) becomes a Cauchy problem: f Equation (7.3.12), { V(zu.
..,zn,T)
= P(zi,...
,*„),
(7.3.13)
where P{z\,..., zn) = P(eXl,..., eXn). To solve it, we need first to find the fundamental solution of equation (7.3.12), i.e. a solution of (7.3.12)
207
Multi-Asset Option Pricing
with the terminal condition: V(zu where 5(z\,...,
...,zn,T)
4),
= S(Zl -z\,...,zn-
(7.3.14)
zn) is the multidimensional Dirac function S(zi,...,zn)
=6(z1)...S(zn).
Let V = esTg+^T-Vw,
(7.3.15)
where /? is a constant, Q is a n-dimensional vector:
LJ =
'•
Substituting (7.3.15) into equation (7.3.12), we get
-aT + a B ^ + ^l-aT + a i ; ^ - ^ i=l
l
i=l
-{r + P-
-(CDTACJ)
- brBTQ\W
= 0.
Choose
<* = - f ^ T ^ )
(7-3.16)
w = -k~lBb,
(7.3.17)
i.e.
and /? = - r + -LOTAL3 +
VBTQ
= - r + \(Bb)TK-lkK-\Bb) = -r+ ^BbfA-^Bb) =
- brBTA'1Bb
- {Bb)TA-\Bb)
-r-l-{Bb)TA-\Bb). (7.3.18)
208
Mathematical Modeling and Methods of Option Pricing
Then under the transformation (7.3.15), equation (7.3.12) is reduced to dW
d2W
1A
^+2|>^=0-
<7'3-19>
And the terminal condition (7.3.14) becomes W(zi,.
..,Zn,T)
= e-*Tz~06(Zl - z°,...,
zn - z°n),
(7.3.20)
where •*?• zh=
.
'•
.4. According to the theory of the heat equation, the solution to the problem (7.3.19), (7.3.20) is of the form:
^,,,; 2 o). e -^n^ A < ( r _^-"^
eXP
l
2{T-t)
)•
Back to the original function V and variable x, we obtain the fundamental solution T(x,t;x0) to the Cauchy problem (7.3.2), (7.3.13) as follows:
r
i ? e-r{-T-v
i
{B{x-xo))TK-\B{x-xo))
/ •
e X p
\
2(T=t)
- i( J B6) T A- 1 (56)(T -t)-
(A^BbfBix
Since i^A-1!? = B-lk-\B-l)T
= {BTKB)-1 = A~\
- xo)\ .
209
Multi-Asset Option Pricing
thus
[
exp
1
i t e-r(T-t)
{"W^T) i{x ~ f ° + S{T ~l))TJ4-1(f ~ s°+ i{T ~f)]\} (7.3.21)
Then the solution of the Cauchy problem (7.3.12),(7.3.13) has the form /•OO
v(x,t)=
r-OO
r(x,«)P(Ci,..,g
... J— OO
J — OO
Back to the original variables S i , . . . , Sn by (7.3.1), we obtain the European multi-asset option pricing formula: r y(5 i)=
'
1
i f e~r(T-t)
[2n(T~t)\ • /
Jo
•••/
Jo
^7|F exp - • — — - —
f?i,-.-,77n
L
2
(-/ ~ * ) J
\dr]1...drln,
(7.3.22) where 5=
:
,
.an_
and =ln^ +(r-ft-^i)(T-t). (t = l , . . . , n ) (7.3.23) Vi 2 (7.3.22) is called the Black-Scholes formula for European multi-asset options. Remark (7.3.22) indicates that there exists a closed-form expression for any European multi-asset option. However, this is a multiple integral with singularities in the integrand. When a large number of assets are involved, the integral has high multiplicity and is very difficult to evaluate. Thus, finding a closed-form expression is only the first step in solving the pricing problem of the European multi-asset option. We still need to find a j
210
Mathematical Modeling and Methods of Option Pricing
a simple way to get special solution to the problem for each concrete form of the payoff function. 7.4
Rainbow Options
There are three general categories of European multi-asset options: (I) rainbow options, (II) basket options, (III) quanto options. For each category of multi-asset options, we will explain its financial meaning, specify its payoff function upon maturity, and derive the option pricing formula. In particular, we will find out whether it can be reduced to a one-dimensional problem. As a rainbow is a combination of various colors, a rainbow option is a combination of various underlying assets. The value of a rainbow option depends on the performance of the underlying assets. According to the payoff structure, rainbow options have the following forms: (A) Better-of options The holder of better-of option receives exercise payoff associated with the better performer of the underlying assets.
Example An investor considers to invest in stock index A or stock index B, but is not sure which index will bring a higher return. By buying a better-of option, the investor is guaranteed to receive a higher return according to the better of the two indexes at maturity. There are two types of better-of options: (1) According to the price, payoff =max(a 1 Si(T),...,a n S n (T)), where Si(T) is the price of the i-th. risky asset at t = T, and Qj is a coefficient so that all risky asset prices are at the same level. (2) According to the growth rate, payoff (rate) = max(Si(T),..., Sn(T)), where §i(T) is the price growth rate of the i-th risky asset at t = T. Since
-
U j
_ SJ(T) - $(()) _ s^m _ , 3(0)
"5,(0)
i;
211
Multi-Asset Option Pricing
thus max&iT),...,
anSn(T)) - 1,
Sn(T)) = maxiaiS^T),...,
where
In general, the payoff function of a European better-of option can be written as payoff = max(5i(T),..., Sn(T)), where Si(t) can be either the same-level price, or the growth rate, of the i-th risky asset. Mathematical model for better-of options: to solve the multivariate Black-Scholes equation (7.2.6) with the terminal condition: V(SU... ,SnjT) = max{S 1 ) ..., Sn}.
(7.4.1)
Its solution is given by the multi-asset Black-Scholes formula (7.3.22), i.e.
v<s,,..., s_, t) = [ 2 T ( r_ ( ) ]' e"r(r"" det Ml"' Jo
Jo
Vi,---,Vn
I
2(T-t))
where a is defined by (7.3.23). Remark Closely related to the better-of option, another form of rainbow option is the worse-of option. Its exercise payoff is payoff
=mm(S1(T),...,Sn(T)).
(B) Out-performance options The exercise payoff of out-performance option is based on the difference in the performance of two assets. Example An investor holding a position in a risky asset A, not sure whether A will outperform B, can purchase an out-performance option so that payoff will be based on the difference if risky asset B outperforms A. Thus the exercise payoff of the option is max(Sls(T') — SA(T),0), where SA(T),SB(T) can be either the same-level price, or the price growth rate,
212
Mathematical Modeling and Methods of Option Pricing
of the risky assets A, B. Mathematical model of out-performance options: to solve equation (7.2.6) with the terminal condition: (7.4.2)
V(S1,S2,T)=max{S2-S1,0}.
Remark A similar form of option is the spread option. Its exercise payoff function is V(Sl,S2,T)=
max(S2 - St - K, 0),
(7.4.3)
where K is the strike price. Remark Another form related to the out-performance option is the options to exchange one asset to another. By purchasing an option to exchange one asset for another, an investor holding a position in a risky asset A can exchange A for another risky asset B, or trade A for B by adding cash K, when B outperforms A. Mathematical model of options to exchange one asset for another: to solve equation (7.2.6) with the terminal condition (7.4.2) or equation (7.2.6) with the terminal condition (7.4.3). Remark A European better-of (worse-of) two assets option can be decomposed into a risky asset and an option to exchange one asset for another. In fact, since European option pricing is a linear problem, the payoff function of a better-of (worse-of) option can be written as max(5i(T),S2(T))
= 5i(T) + max(S2(T) - S^TJ.O)
mm(S1(T),S2(T))
= S2{T) - max(S 2 (r) - Si(T),0).
and
(C) Maximum call options and minimum call options A maximum call option offers its holder n choices: to buy risky asset Si at strike price Ki, (i = 1 , . . . , n). The option holder can choose the best performer at the expiration date to get the maximum payoff. The payoff at the expiration date is payoff = max{(5i(r) - Kx) +,...,
(Sn(T) -
Kn)+}.
213
Multi-Asset Option Pricing
Mathematical model of the maximum call options: solve equation (7.2.6) with the terminal condition V(S!, ...,Sn,T)
= max{(S1
- Kx)+, ...,(Sn-
Kn)+}.
(7.4.4)
Its solution is given by the Black-Scholes formula (7.3.22). Remark For maximum call options, if
K1 = --- = Kn = K, then the payoff function is reduced to V(S!, ...,Sn,T)=
max{(Si -K),...,(Sn= {m&x{S1,...,Sn}-K)+.
K),0} (7.4.5)
This kind of the maximum call options is often regarded as a generalization of the single asset vanilla call option. Remark Corresponding to the maximum (minimum) call options, we may also consider the maximum (minimum) put options. In particular, if K\ = • • • — Kn = K, its payoff function is payoff =(K-
m a x ( 5 1 , . . . , Sn))+
(7.4.6)
or payoff ={K- min(5 1 ,..., Sn))+. (7.4.7) This is often regarded as a generalization of the vanilla put option on a single asset. In the following, we will discuss which of the rainbow option pricing problems can be reduced to a one-dimensional problem. For some problems, by introducing a new variable as combination of the original ones, the equation, the terminal and boundary conditions, and the domain will depend on this new variable only. The out-performance and better-of (worse-of) options of two risky assets can each be reduced to a one-dimensional problem. Let us begin with the out-performance option. The mathematical model of the out-performance option pricing is the terminal-boundary value problem (7.2.6), (7.4.2): i.e. in domain £ :
214
Mathematical Modeling and Methods of Option Pricing
{(Si,S2, t)\0 < Si < oo,0 < S2 < oo,0 < t < T}, solve the problem 2
dV,l\n
qid
,,„
V
+ (r-qi)S1§^
a
q
d2V
, _
+ (r-q2)S2§^-rV
2
q2d
v]
= 0,
V(5i,5 2 ,T) = max(5 2 - SUO).
(7.4.8) (7.4.9)
Let £= ^ ,
(7.4.10)
u(Z,t) = V(S1,S2,t)/S2.
(7.4.11)
and
Then we have dV _ ~dSi~ dV
ds-2=u
+
du d£ _ du ~di~dS~i ~ ~d£'
2
o
du d£ S2
du
dtds-2=u-^>
c?y_ I d2u 'b~s2~^2'W d2V _ S^ud^_ _ g d2u 2 dSxdS2 ~ d£ 8S2 ~ ~ S2 d£2 ' 82V _ dyLd^_dudi__d%Ld^_ _ g2 d2u dS% ~ d£ W2 ~ MdS2 ~^d? 8S2 ~~ S2 d£2 ' Substituting these into equation (7.4.8), we get du 1, . od2u , i^du — + - [a n - 2a i2 + a22] g 2 -— + (q2 - qi)£— - q2u = 0.
. . (7.4.12)
By transformation (7.4.10),(7.4.11), the terminal condition (7.4.9) becomes u(£,T) = ±-V{S!,S2,T) o2
= ^ m a x ( S 2 - 5i,0) = (1 - 0 + o2
(7.4.13)
215
Multi-Asset Option Pricing
And the problem domain S is transformed to: {0<£
t) = -e-nV-QtNi-di) + e - * ^ - ' ^ - ^ ) ,
(7.4.14)
where %
CL\ =
ln^ + [g2 - gi + I (an - 2a 12 + a 22 )] (T ~ t) .
V(an-2a12-|-a22)(r-t)
,
rf2 = di - \/(aii - 2ai2 + o22)(T - <).
I I A. 1 0 )
(7.4.16)
Back to the original variable (Si,S2,t) by reversing the transformation (7.4.10),(7.4.11), we get V(S1:S2,t) = S2e-^T-^N(-d2)
ui =
- 5ie"' l(r - t) iV(-a T 1 ) >
I n ^ + [ ( f t - ( ? 1 + | ( a i i - 2 a 1 2 + a22)](r-t) x/(an-2ai2
=
+ a 2 2 )(T-i)
,
(7.4.17)
(7.4.10)
(7.4.19)
The better-of two risky assets option pricing model: in domain £ solve equation (7.4.8) with the terminal condition: V(S1,S2,T) = max(Si,S2).
(7.4.20)
By the same transformation (7.4.10),(7.4.11), u(£,t) satisfies equation (7.4.12) in domain {0 < £ < oo, 0 < t < T} and the terminal condition:
u(£,T) = ±-V(SuS2,T) = max(e,l) = (1 - 0 + + f-
(7-4.21)
216
Mathematical Modeling and Methods of Option Pricing
Then, w(£, t) can be expressed as the sum of a vanilla put option (7.4.14) and a function ^e~qi^T^tK Back to the original variable (S\,S2,t), we get
V(S1,S2,t) = S2 [|^(1 - AT(-J1))e-9l(T-t) + e-^ T -"iV(-i 2 )] = Sie-^-^Nidia)
+ S2e">^T~^N{d2,i),
(7.4.22)
where di2 =
In § + [q2 - gi + \ (an - 2a12 + a22)] (T - t) . : , V(«n-2a12 + a22)(T-t) lnfi+[9i-
U2 i =
7.5
—
V(an - 2ai 2 + a22)(T - t)
.
(7.4.23)
(7.4./4)
Basket Options
The payoff of basket options is n
payoff =(£,<XiSi-K)+,
(7.5.1)
i=l
where a, is the portion of the i-th asset Si in the basket. Basket options are usually used in trading two or more foreign currencies. Let Si be the exchange rate of the i-th foreign currency, and K be the weighted average exchange rate based on the exchange rates of all underlying currencies. According to the portfolio theory, the volatility of the basket is in general smaller than that of each individual currency. Therefore the basket option premium is less than the sum of the premiums of all individual options on each underlying currency. Prom the multivariate Black-Scholes formula we can get the pricing formula for the basket options. Is it possible to reduce the basket option pricing problem to a 1-D problem? The answer is negative! The reason is that the basket option price is based on the arithmetic mean of n asset prices, whereas the price
217
Multi-Asset Option Pricing
of each asset makes geometric Brownian motion. However, if the payoff of a basket option is the geometric mean of n underlying assets, i.e. if the payoff function is n
payoff =(Y[S»<-K)+,
(7.5.2)
i=i
where J^Qj = l,a; > 0, then the pricing problem can be reduced to a 1-D problem. By transformation xi=ln5i,
(7.5.3)
equation (7.2.6) becomes (7.3.2), and the payoff function becomes n
/
K +
n
\ +
(JIST - ) = expj^a^i} - K i=i
V
t=i
.
(7.5.4)
/
Change the variables to n (7.5.5)
(, = ^aiXi, i=l
thus, dV _ dV
d2v
d2v ~aiaj~d?-
dxidxj and equation (7.3.2) is reduced to dV
-m
+
l,2d2V
.
+ {r q
2° W
'~
1
2
a )
2aV
^ ~rV = °-
where n 2
a
= ^2 o-ijO'iOij, »,J = 1
n
i=l
^9
(7 5 6
--)
218
Mathematical Modeling and Methods of Option Pricing
and the terminal condition (7.5.4) is reduced to V(£,T) = (eS-K)+.
(7.5.7)
Then by the Black-Scholes formula of the vanilla European call options, we can get the expression of V(t;,t), and back to the original variable (Si,...,Sn,t), we get
V = e-W-Qs?1 ... SZ-Nfa) - e-r^T-^KN(d2),
(7.5.8)
where cal
f
qan
^ l n ^ ^ di =
+
\r-q+^ 7 = —
«21
\(T-t) ,
d2 = dx - ay/T - t. 7.6
(7.5.9)
(7.5.10)
Quanto Options
Quanto options are cross-currency contracts. In addition to the risk due to the fluctuations of the foreign asset price, there is also a risk due to the fluctuations of the exchange rate for that foreign currency. Example Suppose a US investor wants to buy a European call option on a certain European company stock at strike price K (Euro). What should be the premium of the option in USD? Here are two risky assets: one is the European stock price S\ (in Euro), and the other is the USD/Euro exchange rate 52- At option maturity, the exchange rate can either be (1) the spot exchange rate S^i), or (2) a spot exchange rate 52 (t) which will be no less than a guaranteed rate S%. Correspondingly, the payoff function at maturity is either payoff function = S2(T){S1(T) - K)+
(7.6.1)
or payoff function = max(5 2 (r),5^)(5 1 (r) - K)+
= (S2(T) - S%) + (S[T) - K)+ + S%{S[T) - K)+.
(7.6.2)
219
Multi-Asset Option Pricing
Now let us establish the PDE for quanto option price (in USD). Suppose ^p-=Hidt + *idW1,
(7.6.3)
^
(7.6.4)
= n2dt + a2dW2,
where Si is the European stock price (in Euro), 52 is the USD/Euro rate. Also assume the Brownian motion with correlations, E(dWi) = 0, Var{dWi) = dt, Cov{dW1 ,dW2) = pdt.
(i = 1,2) (\p\ < 1)
Let the option price be V (in USD): V = V{SuS2,t). By A-hedging, construct a portfolio II (in USD): U = V-
A1S2S1
(7.6.5)
- A2S2.
Choose A i , A 2 such t h a t II is risk-free in (t,t + dt), i.e.
dU = nlldt,
dU = dV - Aid(5i5 2 ) - A2dS2 - A&Siqdt
(7.6.6)
-
A2S2r2dt,
where r\ is the risk-free interest rate in home country, and r2 is the risk-free interest rate in the foreign country, and q is the dividend rate of the risky asset Si. By the Ito formula, we have
(7.6.7) d{S!S2) = SidS2 + S2dSi + o-1a2Sr152pA.
(7.6.8)
220
Mathematical Modeling and Methods of Option Pricing
Thus from (7.6.6)—(7.6.8), n(V - AiSxS2 - A2S2)dt (8V
+ °lSl^\
1. 2c,2d2V
„ „ 82V
- AK7K7 2 5IS2/I»} dt - A!SiS2qdt dV
- A2S2r2dt + ( — - A1S2)dS1 obi dV + («F- ob2 Choose
Al51
- A2)dS2-
(7.6.9)
AiSi + A ^ — . C/D2
i.e. (7.6.10)
(7.6.11)
Substituting (7.6.10),(7.6.11) into (7.6.9), we have
+ (n - q - <J\a2p)Si— + (n - r2)S2 — dV + (r 2 -n)5i-— -riV Obi
= 0.
221
Multi-Asset Option Pricing
i.e.
dV
In
+
1 f o^&V
2 1 ° ^ as?
d2V
n n
+ 2p
^^
+ a
2ci29
2
Vl
^asl\
dV dV + in - qi)Sijr^ + (n - g 2 ) 5 2 - - - n V = 0,
(7.6.12)
where 9i = n - r2 + g + <Ji<J2P,
(7.6.13) (7.6.14)
q2 = r2. The payoff function att = T (see (7.6.1), (7.6.2)) is either V(S1,S2,T) = S2(S1-K)+
(7.6.15)
V{Si,S2,T) = max(52,5°)(51 - K)+.
(7.6.16)
or
Therefore the mathematical model of quanto options is a two dimensional PDE terminal value problem (7.6.12), (7.6.15) or (7.6.12), (7.6.16). By the multivariate Black-Scholes formula (7.3.22), we can give an explicit expression of the option price. Take the terminal value problem (7.6.12), (7.6.15) as example, V
(Si,S2,t)
r = _ l L==^ e-'»< -*> 27T(i -t) Ox<J2sJ\ - p2
te
E
•jrjr 7 -'[-
<M
^Tag-fK (7.6.17)
where o
2
ax = In — + (r-2 - q - <7Xa2p - ^-)(T - t), a2=ln —+ ( r ! - r 2 - ^ ) ( T - t ) . % 2 Note: in deriving (7.6.17) from (7.3.22), we used
[ e r f poxa-A [paia2 ai J
222
Mathematical Modeling and Methods of Option Pricing det A = (T%a%(l - p2), A-i
7.7
=
1
I" °2
-po-i^"]
American Multi-Asset Options
As in the case of American single-asset options, the mathematical model of American multi-asset options pricing is a parabolic variational inequality. Define £ = { (Si,..., Sn, t)\ (Si,..., Sn) £ R£,0 < t < T} , ( R £ = { ( 5 i , . . . , Sn)\0 <Si
=
l,...,n}.)
And let V = V(Si,..., Sn, t) be the American option price. Then V is the solution to the following problem in E: (min{-^-CV,V-P(S1,...,Sn)}=0, {V(S1,...,Sn)
(R»)
= P(Sl,...,Sn),
(E)
(7.7.1) (7.7.2)
where £ is the multivariate Black-Scholes differential operator given by (7.2.6):
and P ( 5 i , . . . , 5 n ) is the payoff function. Equation (7.7.1) indicates: in the continuation region Si: dt V>P(Su...,Sn); and in the stopping region E2: dt V =
P(Slt...,Sn).
Multi-Asset Option Pricing
223
The interface T between Si and S2 is the optimal exercise boundary of the American multi-asset option. V as a function of n + 1 variables (Si,... ,Sn,t), and its derivatives with respect to S\,... ,Sn , i.e. V and VXV, are continuous in the entire region S = Si I J r | J ^ 2 v As in the case of American single-asset options, we can also formulate (7.7.1),(7.7.2) as a free boundary problem. However, for American multi-asset options, the structure of the optimal exercise boundary (i.e. the free boundary) can be extremely complicated, therefore we need to combine various payoff functions P{S\,..., Sn) to study the behavior of the boundary V. To study the behavior of the optimal exercise boundary F, we need to first study the properties of the option price V = V(Si,... ,Sn,t). Similar to what we did in §6.4, we study the signs of ^rg-, ^jr by introducing a penalty function. Since the procedures are analogous, we only give the conclusions, and leave the proof to the reader. Theorem 7.1 Suppose the payoff function P{S\,..., Sn) satisfies: (1) P(Si,..., Sn) is a Lipschitz continuous function; (2) There exists P£(Si,. ..,Sn)€ C°°(R^), such that \imPe(Su...,Sn)
e-*0
=
P(S1:...,Sn),
and is uniformly convergent in any finite bounded domain within R™; (3) ^>0(or^<0), (7.7.3) and in any finite bounded domain D within R " , there exists a constant Co which depends on D only and is independent of e, such that CPe > CD,
(7.7.4)
then the solution of the American multi-asset option pricing problem (7.7.1),(7.7.2) has the following properties: (1) V(Si,... ,Sn,t) is a nondecreasing (monotone increasing) function ofSi(i = l , . . . n ) , (2) V ( £ i , . . . , Sn, t) is a nonincreasing function oft. Now let us study the properties of the optimal exercise boundary for American multi-asset options with a specific payoff function.
(A) American better-of option on two assets
224
Mathematical Modeling and Methods of Option Pricing
The mathematical model is: ([20])
(E)
(mm{-^--CV,V-P(S1,S2)}=0, \ v | t = T = P(5i,S2),
(7.7.5) (7.7.6)
(SUS2 eRl)
where
CV = 1 \ajSffV + 2Pa1a2S1S2^y5-2 + af Sf | | ] + (r - 9 i ) S i ^ - + (r - <72)S 2 f^ - rV = 0, P ( S i , 5 2 ) = max(5 1 ,5 2 ).
(7.7.7) (7.7.8)
Similar to §7.4, we introduce the transformation (see (7.4.10),(7.4.11)) 4
_*
.K,«, = i 3 % ^ .
O2
(7.7J.)
>J2
Under the above transformation, the problem (7.7.5),(7.7.6) is reduced to the following one dimensional problem: in S = {(£,i)|0<£
min <
(7.7.10)
, U | 4 = T = max(£,l),
(7.7.11)
where a 2 = (T? + 2po-1
(7.7.12)
9i,92 > 0 ,
(7.7.13)
If
then we can write S as S i U f U E 2 ) where Ei is the "continuation region" of the option: u > max((, 1), du
1 ~ od2u
m + 2aeW
s,9u
,
+ {q2 qi)
~ ^rq2U
= 0;
225
Multi-Asset Option Pricing
E2 is the "stopping region" of the option: u = max(£, 1),
f is the interface of Si and £2Theorem 7.2 If (7.7.13) is true, then for the problem (7.7.10), (7.7.11), the continuation region is: Si = {(£,*) \ii{t) < t < 6(*),0 < t < T}, where
6(T) = 6(r) = i, ^i(t) nondecr-easing, £2(4) nonincreasing. J n S i , u satisfies the free boundary problem:
W + ¥2?^z
+ fe - 9i)£gjf - 92« = 0,
u(ei(t),t) = l,
(0
< ^(6(t),*)=0, «(&(*), *)=f.
(Ex)
(7.7.15)
(0
| | ( 6 ( t ) , i ) = l.
(7.7.14) (7.7.16) (7.7.17)
(0
(7.7.18)
Proof Wefirstprove 9Si P| {t = T}, the intersection of the boundary 9Si
of Ei and the line t = T, cannot be an open set. Suppose otherwise, i.e. there exists (a, b) 6 9Si |"| {i = T}. Since u satisfies the parabolic equation (7.7.14) in Ei, thus due to the regularity of the solution of parabolic equation, at t = T, excluding ( = 1, in (a, 6) there is 9
ft |t=T = [-\?e^i =
/ 9i€, € > 1,
- (* - 9 i ) ^ +ft]max«, 1)
226
Mathematical Modeling and Methods of Option Pricing
By assumption (7.7.13), ^ | t = T > 0 within (a,b) (excluding £ = 1). Thus in the neighborhood of t = T, u is strictly increasing, i.e. ««,t)<«« i r) = max(l,O. This contradicts to u > max(l,£). Let 3Si be the boundary of Ei, then by Theorem 7.1, u(£, t) is nondecreasing with respect to £. Thus by similar argument as above, we claim dti n{t = T} = point(1,T), and the following two continuous rays originating from (1,T) are the free boundaries: £ = £i(t) and £ = &(*), where £i(£) < 1 < £2(t). In d£i, u satisfies the free boundary conditions (7.7.15),(7.7.16) and (7.7.17),(7.7.18). Now let us examine the intersection of each line £ = £0 and 9Ei. It follows from the previous arguments, there exists 0 < T(£O) < T, such that the line segment {(£, r) |^ = ^0 , T(^O) < T < T} belongs to the "stopping" region E 2 . Here r(l) = T, r(fo) < T,
(fr 7^ 1)
By Theorem 7.1, V(f,t) is nondecreasing with respect to t, thus the line £ = £o and 9Ei may either (1) intersect once ( at a point or a segment) or (2) not intersect at all. Thus we have proved 6(£)<1<6W and £i(£) nondecreasing, £2(2) nonincreasing. Thus competes the proof of Theorem 7.2. Remark In fact, by the regularity of the solution of parabolic equation and the strong maximum principle, we can further show that the line £ = £0 and 3Ei cannot intersect at a segment, i.e. £i(£) and ^(t) must be strictly increasing and decreasing, respectively.
227
Multi-Asset Option Pricing Back to the original variables (Si,S2,t), better-of option:
we get the price of the American
(S2, 0<§i
[Si,
fc(t)
(7-7.19)
<£<<»,
where the optimal exercise boundary consists of surfaces Fi and F 2 : Fi : Si = ti(t)S2,
(7.7.20 )
T2 : Si = &(t)S2,
(7.7.21)
and
where 6 CO = &(T) = 1,fc(t)T.ftW I •
52J1
sjl
S>/S*
o^-
/
\ y v=s2u(—,o
»~
o**—
'—•
*^
Left figure: At t — T,T\ and F 2 degenerate to a ray: Si = S 2 . Right figure: At t < T, positions of Tu T2 This shows how the optimal exercise boundary T for American better-of options evolves with time t according to the following picture: At t = T, Fi and F 2 coincide to the ray Si = S2. When t < T, the optimal exercise boundary lines separate. At any given time t, the optimal exercise boundary consists of two rays originating from (0,0): Si = £i(t)S2 and Si = $ 2 (t)S 2 . Between them lies the continuation region, and beyond them lies the stopping region. At t = 0, the optimal exercise boundary rays are Si = £i(0)S 2 and Si = £ 2 (0)S 2 . Now we ask: is it possible to estimate the lower bound of £i (0) and the upper bound of £2(0)? To answer this question, we consider the model of the perpetual American
228
Mathematical Modeling and Methods of Option Pricing
better-of option pricing. Similar to §6.1, this is an elliptic obstacle problem:
{
min{- CVco, Vx - max(5i, S2)} = 0,
Voo(0, S2) = S2,
(R^.)
(Si € R+)
(7.7.22)
(7.7.23)
Voo(Sl,0) = Si. (5 2 G R+) Under the transformation V^ = S2W,
(7.7.24) (7.7.25)
t = I1-
(7-7.26)
J2
the free boundary problem (7.7.22)—(7.7.24) is reduced an ODE obstacle problem in domain R+:
min{-\c?e£g- - (92 - qi)^
+ q2W,
W-max(l,0} = 0,
(£6R+)
W(0) = 1,
(7.7.27) (7.7.28)
. W ( 0 ~ ^,
(C - oo)
(7.7.29)
where a is defined in (7.7.12). Similar to §6.1, the obstacle problem (7.7.27)—(7.7.29) is equivalent to the following free boundary problem: find {W(f); £!,£§} such that ^ 2 ^ _
+ {q2_qi)(dW_q2W
=
^
{i0
(7730)
W(&) = 1,
(7.7.31)
< W'(£) =0,
(7.7.32)
W(&) = &
(7.7.33) (7.7.34)
.W'(&) = 1. The genera] solution to equation (7.7.30) is
W = A(h"*+B(ha>, ?1
(7.7.35)
S2
where au(i = 1,2) are the two roots of the characteristic equation: -a2a(a - 1) + (q2 - qi)a - q2 = 0, i.e. at=uj+{-iye,
(i = 1,2)
(7.7.36)
229
Multi-Asset Option Pricing where ^ = 5 + ^(91-92),
W l
+
F + i 2 L 7^ ! ~
(7.7.37)
(7 7 38)
--
It follows from the assumption (7.7.13) that c*i < 0, A,B, ^1,^2 (7.7.34):
ca
a 2 > 0.
n be determined from the free boundary conditions (7.7.31)—
si
^(Ir'
+
+
^HlUo' 51 ?1
R^-i
°
?2
?1
S2 ?2
Solving the above system of equations, we get A,B,£f,$. them into (7.7.35), we get
Then substituting
w(o = a °!!a ft''""1 ^" °2~°1' ^a2
where # = 2*11!. Back to the original variables SUS2
(* = l, 2)
(7.7.39)
and Voo (see (7.7.25), (7.7.26)), we get
230
Mathematical Modeling and Methods of Option Pricing
the pricing formula of the perpetual American better-of option:
Kx>(. Ol,b2)
= <
^
a
a
_
Ti :
5i = ^?5 2 ,
F2 :
5 l = ^2'S'2-
Positions of the optimal exercise boundary Fi and F2 for the perpetual American better-of option
y^^
/ /
^^
V
*»?
a
r~
Since the perpetual American better-of option is the most expensive one comparing to other American better-of options with the same strike price, thus max{5i,52} < V{SuS2,t) < V0o(SuS2). Thus we have
tf < 6W < 6(T) = l = 6(T) < 6W < & Thus for American better-of option's optimal exercise boundary Fi and F2, we have the following two valuation relations for their slopes:
6(«)>(? =
^
^
.
231
Multi-Asset Option Pricing where a i , a 2 , / ? i , / 3 2 are defined in (7.7.36)—(7.3.39).
Remark The above analysis of the optimal exercise boundary for American better-of options is made under the assumption (7.7.13). If one of the dividend rates qi(i = 1,2) is 0, then the behavior of the optimal exercise boundary should be modified. Without losing generality, let us assume 9i > 0,92 = 0.
(7.7.40)
Then under transformation (7.7.9), the problem (7.7.5),(7.7.6) is reduced to
fmin{-^-^V^+9i^,u-max(M)J=0,
(7.7.41)
U(£,T) = max(£,l).
(7.7.42)
Thus corresponding to Theorem 7.2, we have T h e o r e m 7.3 If assumption (7.7.40) (7.7.41), (7.7.42), the continuation region is S i = {(£,*)
is true,
then for the problem
\0
where 6(T) = 1, &{t) nonincreasing, and in Tii, u satisfies the following free boundary problem:
«(&(t),t)=f,
(0
| t e W , ' ) = l,
(0
u(0,t) = l.
(0
i.e. for problem (7.7-41),(7.7.42) there is only one free boundary £ = £2(2), in £ = {0 < £ < 00,0 < t < T}, u is of the form
fiiU,t),o<€
U,
6W
Back to the original variables, the option price is
V(S1,S2,t)={
[Si,
&{t)<^
232
Mathematical Modeling and Methods of Option Pricing
and the optimal exercise boundary is r:Si=6(t)S2.
(7.7.43)
The proof of Theorem 7.3 is left to the reader. Consider the perpetual American better-of option, under assumption (7.7.40), by transformation (7.7.25),(7.7.26), {VF(O,?°} satisfies the free boundary problem as follows:
-¥2?^$-+^*8r
= °> ° <*< &
(7-7-44)
W(&) = £2°,
(7.7.45)
W($) = 1,
(7.7.46)
. W(0) = 1.
(7.7.47)
Its characteristic equation is 1
-a2a(a - 1) - qia = 0, which has two roots a 1,0:2:
ai = u+{-l)%
(« = 1, 2)
where
thus Qi = 0 < a2.
The solution of the free boundary problem (7.7.44)—(7.7.47) is of the form:
w= A +B^y\
(7.7.48)
Applying the free boundary conditions (7.7.45),(7.7.46) and (7.7.47), we get A = l,
A + B = £, a2B _
f° ~ S2
233
Multi-Asset Option Pricing
Solving the system, we get A,B and f°. Substituting them into (7.7.47), we get
ft) _
Q2
Back to the original variables (Si,S2), we get
I Si,
6°
Since
max(Si,5 2 ) < V(SuS2,t)
thus we have an estimate for the slope the of the optimal exercise boundary (7.7.43) under assumption (7.7.40):
Q2
— 1
2qi
Remark If both underlying assets are non-dividend-paying, i.e., qi(i = 1,2) = 0, then as in the case of single-asset option, it is unwise to early exercise the American better-of option. Then American better-of option has the same price as European better-of option, and there is no optimal exercise boundary.
(B)
American call-max options on two risky assets
For American call-max options on two risky assets, the exercise payoff is: P(Si,S2)
= (max(Si,5 2 ) - K)+.
(7.7.49)
It can be regarded as a generalization of the American better-of option discussed in (A), since American better-of option is the K — 0 case of American call-max options on two risky assets. American call-max options on two risky assets (K > 0) is a typical 2-D problem. We will make an in-depth qualitative study of its optimal exercise boundary by mathematical analysis. In particular we will make use of the results of the optimal exercise boundary for American better-of options and of the relation between the optimal exercise boundaries of these two American options to gain a deep understanding of the behavior of the optimal exercise boundary for American call-max options on two risky assets. ([20])
The mathematical model for American call-max options on two risky assets: in domain S: {{Si,S2)eR2+,0
234
Mathematical Modeling and Methods of Option Pricing
solve the variational inequality
(£)
(mm{-^--£V,V-P(SuS2)}=0, \ V \t=T = P(Si, S2), where P(SUS2)
(7.7.50)
(Si, S2) € R+
(7.7.51)
is defined in (7.7.49), and
(7.7.52)
By transformation (i = 1, 2)
Xi = lnSit the problem (7.7.50),(7.7.51) becomes m i n { - ^ -C0V,V-p(x1,x2)\
=0,
(a;i,x2) G R 2 , 0 < t < T, (7.7.53)
ar
a;
K| t =r = p( i> 2),
2
(2:1,2:2) GR ,
(7.7.54)
where
+ (r-V-4)l^
+ (r-<»-4)l£-2-rV,
p(x 1 ,x 2 ) = P(e :ri ,e X2 ) = (max(e X l ,e : C 2 )-K) + .
(7-7.55) (7.7.56)
To show that the option price is monotonic with respect to S\,S2,t,K (i.e. xi,X2,t,K), we need to introduce pe(xi,X2) € C°°(R 2 ) according to the conditions of Theorem 7.1. Therefore we define a function pt(xi, x2): pc(xi,x2) = nc(Ft(xux2)-K),
(e>0)
(7.7.57)
where s,
s > e,
/ , -e<s < t ! 0, s < -t.
(7.7.58)
235
Multi-Asset Option Pricing
and 0 < n«(fl) < 1,
I
n"(«)>o,
(7.7.59)
j
limn£(s) = s+.
e—*0
And define pxl
F £ (x l!a;2 ) =
_L px2
Y
p*\
+
_
2
px2
6
-xl
M
_
2£
e
p *2
),
(7.7.60)
where h(5) = - a r c t a n 5 .
(7.7.61)
7T
It is easy to check: h'(S) > 0,
0<2h'{S) + Sh"(S)< | , 0 < 1 ± [/i(5) + 5/»'(5)] < 2.
(7.7.62)
Then we have lim FAx\,X2) = maxfe^.e 12 ). And therefore limp E (xi,x 2 ) = (max(e xl ,e X2 ) - K)+ = p(xi,x2)Obviously, pt(xi,X2) is monotonic with respect to x\,X2,K. To check condition (7.7.4) of Theorem 7.1, we only need to examine the lower bound of CaPi{xi,X2). Since (F e ) I1Ia = ^ 1 [2h'(S) + Sh"(S)] + ^
(Fe)xlX2 = ^p-
[-2ft'(5) - 5ft"(S)] ,
(FeU*2 = ^ - [2ft'(5) + S/i"(S)] + ^ where e*l
S =
[1 + h(S) + Sh'{S)] ,
_
2e
e ^2
•
[1 - ft(5) - 5^(5)] ,
236
Mathematical Modeling and Methods of Option Pricing
thus by straightforward computation, from (7.7.59),(7.7.62), we get - K) K ( F E X 1 ) 2 + 2pa1a2FcX1FtX2
Cpc = \n"{Ft
+ ol{FlX2)2} + ^ n l ( F £ - K) [2ti(S) + Sh"{SJ] • [a2ie2xi ~ 2po-1a2e^+^ + ale2**}
+ ^-(r ~ qi)n't(Ft - K)(l + h(S) + Sh'(S)) + ~(r
- q2)ri€(Fe - K){1 - h(S) - Sh'(S))
-rne(Fe-K) > r [-ne(F« -K) +rie(F<- K)(Ft - K)j + Yl'e(Fc - K) [rK -
max(qltq2)Ft]
+ - n l ( F £ - K)(r - max(gi, g 2 )) T 4^2 • Define D = {0 < Si < M(i = 1, 2)}, then since maxFe < e M , D
thus from
n £ (s)-sn £ (s)
we get
If
Cpc > --r - max(gi,q 2 ) e M H . 2 _ | ^J | r < 1,M > 1,
then £,Pt > — 1 — 2max(<7i, q2)e . thus we have verified that there exists a e-independent lower bound CD for Cpe (see (7.7.4)). By Theorem 7.1, we infer the following properties of American call-max option price V(S\,S2,t): {\)USi>Si (i= 1,2), then V(SuS2,t)
> V(Si,S2,t);
(7.7.63)
237
Multi-Asset Option Pricing (2) If K > K, then 0
(7.7.64)
< K - K;
(3) If i > t, then (7.7.65)
V(SuS2,t)-V(S1,S2,t)<0.
Now we study the behavior of American call-max options according to (7.7.63)—(7.7.65). Suppose T(t,K) is the optimal exercise boundary for V(Si,S2,t;K), T(t,K)=T1(t,K)DT2(t,K), where = dH2{t,K) n {Si > S2 > 0} ,
(7.7.66)
r 2 (t, K) = dY,2(t, K) n {S2 > 5i > 0} ,
(7.7.67)
Ti(t,K)
where H2(t, K) is the stopping region of the American call-max option at time t, i.e. E2{t, K) = {(Si,S2) | V(SU S2,t- K) = P(SUS2; K), (S1,S2) € R2+} . (7.7.68) Theorem 7.4 If qi > 0, (i = 1,2), then ( Kr T,2{T,K) = I (51,52)1 Si > 52,51 > max(K, — ) ; 5 2 > 5 i , 5 2 > max(K, — ) } . (?2 J
(7.7.69)
Proof If (S°,S°,T) belongs to the continuation region Hi(T,K), and the payoff function P(S2,S2) is differentiate in the neighborhood of (S°,S2), then due to the regularity of the solution to parabolic equations, in the neighborhood of (5i,5 2 ,T), V(Si,S2,t) satisfies equation
Thus dV -Q£ (S°,S°,T) = -£(V\I=T)
(S°,S°) = -£(P(Sl,S2))
(S?,S0).
By straightforward computation, we get £(P(SUS2))
= Kr- Siqi
as Si > S2, 5j > K,
238
Mathematical Modeling and Methods of Option Pricing
Thus in the regions
= Kr-S2q2,
as S2 > Si, S2 > K.
51 > S2 and Si > max(K, — )
(7.7.70)
52 > Si and S2 > max(K, — ) ,
(7.7.71)
and
we have C(P(SuS2))<0, i.e. when (S?,S2,T)
satisfies conditions (7.7.70), (7.7.71), dV ~fa CS«,S»,T) > 0.
But this contradicts to property (3) of V (see (7.7.65)). Thus T,2{T,K) D l(Si,S2) | Si > S2,Si > max (K, ^ \ ; S2 > Si,S2 > max (K, — ) \ . Therefore, in order to prove Theorem 7.4, we only need to point out that all points not belonging to the set on the righthand side of the above expression must belong to the continuation region Ei(T, K). In fact, since for points in region {(Si,S2)\K > m a x ^ i , ^ ) } , P(Si,S2) = 0, thus the set must belong to Ei(T, K). Moreover, if ^ > K, then at t = T, the set / ( S i , ^ ) | Si > S2,K < Si < ^ j must also belong to Ei(T,K). otherwise, i.e. it belongs to E2(T, K), then since in Y,2(t,K)
Suppose
V(SuS2,t) = P(SuS2), then there must be
£at + £V<0.
But from the payoff function at t = T (7.7.49), for those points Si > S2 and
K < Si < ff,
CP(Si,S2) >0.
This results in contradiction. Therefore the region j (Si,S2) Si > S2, K < Si < ^ j must belong to region Y,i(T,K). By the same argument, region < (Si,S2) \ S2 > Si, K < S2 < ^ > must belong to the continuation region Si(T, K). Thus completes the proof of the Theorem.
239
Multi-Asset Option Pricing Theorem 7.5
Ift2>h,
then T,2(t2,K)DT,2(tuK);
(7.7.72)
E2(t,jr 2 )CE 2 ((,J(i).
(7.7.73)
IfK2 > Ki, then Proof The relation (7.7.72) is a corollary of the option price property (3) (see(7.7.65)). In fact, suppose otherwise, i.e. there exist (S°,S°) and t = t°,t2 with t° > t?, such that:
(5?,52°,t?)eE2(t?,#:), but
(slslfye^itlK). Then we have
v(s°us°,t°2) > P(solts°2) =
vislslti).
But this contradicts to (7.7.65). The relation (7.7.73) is a corollary of the option price property (2) (see (7.7.64)). In fact, suppose otherwise, then there exist (Si,S2) and t = to, such that when K2 > K\, (5?,S 2 0 ,i 0 )GS 2 (t,K 2 ), but (s?,s£,to)e£i(t,#i). Due to (7.7.64), we have P{S°X,S^K2) = V{Sl,S^,to\K2)
> P(S^,S^;Ki)-K2
+ Klt
i.e. (max(S?, 5?) - K2)+ + K2 > (max(5?, 5?) - Kx)+ + Kx.
(7.7.74)
Thus there must be max(.S'10,5'2)
(7.7.75)
otherwise, if max(S%,S%) >K2>
Ku
then (max(5?, 5?) - K2)+ + K2 = max(5?, 5^) = (max(5?, 5^) - Kx)+ + Kx. This contradicts to (7.7.74). But (7.7.75) and (5°,5°,to) € E 2 (<,^ 2 ) (see (7.7.73)) are contradictory, therefore (5°,5°, to) must belong to E 2 ( t , # i ) . Thus we have proved (7.7.73).
240
Mathematical Modeling and Methods of Option Pricing Corollary
By (7.7.73), at any given time t,
£ 2 (t, K) C E 2 (i, 0),
\/K> 0.
(7.7.76)
Corollary By ("/•". 7.72), for a given strike price K, the stopping region of the American call-max option starts from t = T, and Y,2(t,K) (see (7.7.69)) narrows down with decreasing t, i.e. when further away from the expiration date. Now we can give a qualitative description of the evolving picture of the optimal exercise boundary of the American call-max options. At t = T, Y\ and T2 partly coincide along Si = S2, and £ 2 ( T , K) is determined by (7.7.69).
maxVK,-^
1—f
1
r,(r>
/ f
|
0
^
roax(#ij£)
/
^
^hV^ Si
When t < T , S 2 ( t , 0 ) is separated and depart from Si = S2, and by (7.7.76), £2(4, .ft") is also separated in the same way, and expands, one branch moving in Si > S2 region, denoted by E 2 (t,K), the other branch moving in Si < S 2 region, and denoted by H22\t,K). Their boundaries in {Si > S 2 } and {S 2 > Si}
Multi-Asset Option Pricing are denoted by Fi(i,K) and T2(t,K),
241
respectively.
Theorem 7.6 American call-max option price V(S\ ,S2,t) is a convex function of Si,S2, i.e. for arbitrary 0 < A < 1, (1 - \)V(Si,S2,
t) + XV {Si + ASltS2
where (Si,S2), {Si + ASUS2
+ AS2, t) > V(5i + AASi, S2 + XAS2, t),
+ AS2) eK%.
For the proof of this theorem, we refer the reader to [20]. As a corollary of the theorem, we have: Theorem 7.7 For American call-max options, the stopping regions E^1 (i) and S^2)(t)(0
Qx = (5? + A(5i - 5?), Sj? + X(S2 - s2), t) Since V(Si, S2, i) is convex with respect to Si,S2, thus 0 < (1 - X)V(S^,S°,t) + \V(SuS2,t) - F(5? + A(5i - 5?), S% + X(S2 - S^), t) = (1 - A)F(5?, S^) + XP{SU S2) - V(5? + A(5i - S?),S% + X(S2 - S$),t). Since Q,Qo € E ^ t ) , thus Si > S2,Si > S2, and Si > max(K, ^ L ) , 5? > max(if, ^ L ) , r r thus P(Sl,S2) =
Si-K,
P(Si,S%) = S?-K.
(7.7.77)
242
Mathematical Modeling and Methods of Option Pricing
Therefore (7.7.77) can be rewritten as V(S? + A(Si - 5?), S2° + A(S2 - S2), t) < ( l - A ) ( 5 ? - K - ) + A(5i-A:) = 5? + A(5i - 5?) - K = P(5? + A(5i - 5?), 5? + A(S2 - S20)). Since V{QX) > P(5? + A(5i - 5?), S§ + A(52 - 5?)), thus V(S? + A(5i - 5?), £2° + A(S2 - $£), t) = P(SOX + X(Si - £?), 52° + A(.92 - S20)), i.e. (O
QAGE2X),
Thus the theorem is proved. Finally, we give the following results for the optimal exercise boundary of American call-max options. Theorem 7.8 For the American call-max option, the optimal exercise boundaries Fi(t) onrfF2(t) have the following properties: (1) There exist
s2 =
T2(t) :
Si =
MSi,t),
(2) ipi(Si,t) and tp2{S2,i) are nondecreasing convex function ofSi,(i = 1,2), respectively; (3)
MSut) _ i s™oo si b(ty D2—+OO
O2
where Si = £i(£)£2 and Si = &(t)S2 are the optimal exercise boundaries of American better-of option. We refer the reader to [20] for the proof of this theorem. Conclusion (3) of Theorem 7.8 indicates: for the American call-max option with an arbitrary strike price K, its optimal exercise boundaries Ti(t,K) and Y2(t, K), have Fi(i, 0) and F 2 (i, 0) (i.e. the optimal exercise boundaries of American better-of option) as their asymptotic lines.
243
Multi-Asset Option Pricing Summary
1. Multi-asset option pricing is modeled as a backward multivariate parabolic PDE terminal value problem. The terminal condition is given by the payoff function of the option. Explicit solutions exist for European multi-asset options. 2. Pricing problem of out-performance options and options to exchange one set for another, whether European or American style, can be reduced to a 1-D problem. 3. Although there is no explicit expression in general for multi-asset American option price, an evolving picture of the optimal exercise boundary can be obtained by PDE analysis. This is very useful for understanding the option pricing and numerical computation.
Exercises Suppose Si,...,Sn are prices of n risky assets, and in [0,T] their movements satisfy the SDE system (7.1.5)—(7.1.8). Show that: 1. If V — V(Si.... ,Sn,t) is the European option price with a payoff at t = T being n
V(Su...,Sn,T) = £*<(&), i=l
then
n
V(Su...,Sn,t) = Y,Vi{Si,t), where Vi(Si, t),
(i = 1,..., n) is the solution to the following problem:
^ +£
^ + (r-*)S.ig-rV, = 0, (0 < Si < oo,0
where »i
— 2^1 "•<•] •
2. Let C = C ( S i , . . . , S n , t) and P = P(SU..., Sn,t) be the prices of the arithmetic-mean basket call and put options, respectively. And their payoffs at maturity (t = T) are given by
C(S1,...,Sn,T)=(YtaiSi-K\
,
P(Si,...,SB,T)=fif-^aiS4)
,
244
where oti > 0,
Mathematical Modeling and Methods of Option Pricing
(i = 1,..., n)
i=l
Find the call-put parity formula for the arithmetic-mean basket options. 3. Let C = C{Si,..., Sn,t) and P = P(SU ..., Sn, t) be the prices of geometricmean basket call and put option, respectively. And their payoffs at maturity (t = T) are
C(S1,...,Sn,T)=(f[S?t-K\ ,
P(Sl,...,Sn,T)=(K-f[sfA , where oti > 0,
(i = 1,..., n,)
Find the call-put parity formula for the geometric-mean basket options. 4. Let C = C(Si,S2,t) and P = P(SuS2,t) be the prices of call-max and put-max options, respectively. And at maturity (t = T), their payoff functions are C(SUS2,T) = (max(Si,S2) - K) +, P(SUS2,T)
= (K- max(Si, 5 2 )) + .
Find the call-put parity formula. 5. Let C = C(Si,S2,t) and P = P(Si,S2, t) be the prices of the quanto call, put options, respectively. And at maturity (t = T), their payoff functions are C(SuS2,T)
=
S2(S1-K)+,
P(5 1 ,5 2 ,T) = 5 2 ( ^ - S 1 ) + . Find the call-put parity formula for quantos options. (Consider two cases: a. the foreign security Si pays no dividend, b. The foreign security Si pays dividend at rate q > 0.) 6. Let V(Si,S2, t) and U(Si,S2, t) be the prices of the better-of and worse-of options of two assets Si,S2, respectively, and their payoffs at maturity (t = T) be V(T) = max(5 1 (T),5 2 (r)), and
[/(T) = min(5 1 (T),5 2 (T)).
Find the pricing formula for these two options.
245
Multi-Asset Option Pricing
7. Let V(Si,S2,t) and U(Si,S2,t) be the prices of call-max and call-min options, and their payoffs at maturity (t = T) be V(Si, S2,T) = max{(51 - K1)+,(S2 and U{SUS2,T)
- K2)+}
= min{(5'i - K{)+, (S2 -
K2)+}.
Find the call-max-call-min parity. 8. Let Si, S2 be two risky assets, with dividend rates gi > 0, q2 > 0. And let V(Si,S2) be the perpetual American out-performance option with a payoff V(S1,S2) =
(S1-S2)+.
Find the option price formula and the position of the optimal exercise boundary. 9. Let V(Si,S2,t) be a quanto call option. Its payoff at maturity (t = T) is: payoff = S 2 ( T ) ( S i ( T ) - / 0 + . Find a hedging strategy for the writer of the option to manage the risk. 10. Let S\, 52 be two risky assets with dividend rates q\ > 0, q2 > 0; and let V(Si,S2,t) be an American out-performance option with the expiration date T. Make a qualitative analysis of the optimal exercise boundary of the option.
Chapter 8
Path-Dependent Options (I) Weakly Path-Dependent Options The European options we have seen in the previous chapters have a common feature: the option's payoff depends on the underlying asset's price on the expiration day only regardless the historical path of the underlying asset's price. In the next two chapters, we will study path-dependent options, whose payoff is determined not only by the asset price on the expiration day, but is also affected by how the asset price changes during the option's lifetime. According to how strongly an option's payoff is tied to the underlying asset's price history, path-dependent options are divided into two classes. In the first class, known as weakly path-dependent options, the payoff depends on whether the asset price has reached certain predetermined price level or levels. In the second class, known as strongly path-dependent options, the payoff depends on the price history in part or all of the option's lifetime. Barrier option is the most important type of the weakly path-dependent options. It has many variations. Indeed, American options also belong to the weakly path-dependent options.
8.1
Barrier Options
Barrier option is a type of European options whose payoff depends not only on the underlying asset's price on the expiration day, but also on whether the underlying asset's price has reached a predetermined price (known as the barrier) during the option's lifetime. Barrier options can be divided into two kinds according to the status of the option when the underlying asset price reaches the barrier. Knock-out option: once the asset price hits the barrier level, the option is extinguished. If the barrier is hit from above, it is called down-and-out option; If the barrier is hit from below, it is called up-and-out option. The other kind is knock-in options: once the asset price hits the barrier, the option is activated. The barrier level is also called the trigger. Similarly, regarding the relation between the spot price and the trigger, there are down-and-in options and up-and-in options. 247
248
Mathematical Modeling and Methods of Option Pricing
Each kind of the above options can be a call or a put. In summary, there are 8 cases of payoff functions for the barrier options as shown below:
barrier option(5lB—trigger) Knock-out ({ST-K)+I{St>sB, down-and-out < {(K -ST)+I{st>sB, ((ST-K)+I{St<SBt up-and-out < { (K -ST)+I{st<sB, Knock-in ((ST~K) down-and-in <
+
(call)
te[0,r]}
te[o,T]}
(put) (call)
te[o,T]}
(put)
[l-I{St>SB,
{(K-ST)+[1-I{SI>SB,
up-and-in
t6[o,Tj}
((ST-K)+[l-IlSt<SB, < [ (K - ST)+[l - I{St<SB,
t6[0iT]} ]
(call) (put)
telo,T}})
te[OiT]} ]
t6[0,T]}]
(call)
(PUt)
where IU(S) (/ u in short) is the characteristic function of the set w,
If we regard (ST — K)+ or (K — ST)+ as payoff of the vanilla option, then it is obvious: payoff of knock-out +payoff of knock-in=payoff of vanilla option Since European option pricing problem is a linear one, the following relation between the price of a barrier option and the price of the corresponding European option holds: ' vanillal,'-') £/ — ' up—and—outv.'-'j ") * * up—and — i n \ ^ i *J — 'down — and—out \^y ") '
/n -i i \
'down—and—in(*^> ")•
The financial meaning of the above equality is clear: when the barrier S — SB is reached, both the knock-in and knock-out (having the same strike price K and expiration time T) are triggered, and the net effect is the same as a vanilla European option. Barrier options are very popular in financial markets. Investors are attracted to barrier options because barrier options are cheaper than European options. This depends on the investor's view of the future price of the underlying asset. Take a down-and-out call as example. If an investor believes once^the asset price falls to the barrier (S = 5 B ) , the chance for the price to return to S > K is so small that the investor would rather give up the possible chance of making profit
Path-Dependent Options (I) — Weakly Path-Dependent Options
249
when the asset price reaches and exceeds the barrier SB at maturity t = T, and renounce the option. Thus the investor can save the option premium, but is at risk of getting no profit if the asset price falls to 5 B temporarily and keeps rising hereafter so that on the expiration day ST > K. In the following example we can see how premiums for various barrier options compare with the premium of the vanilla option. Example Consider a stock call option with 6 month expiration. We assume the spot price is $145, risk-free interest rate r — 6%, dividend rate q = 3%, volatility a = 0.295. option type strike price K barrier Sg vanilla $145 ~ up-and-out $145 $160 up-and-out $145 ~ $190 up-and-in ~ $145 ~ $160 up-and-in $145 $190 down-and-out $145 $130 down-and-out $145 $110 down-and-in $145 ~ $130 down-and-in $145 $110
premium " $12.87 $0.17 $4.46 $12.70 $8.41 $10.48 $12.83 $2.39 $0.04
Partial differential equation model for barrier options: For barrier options, domain, boundary condition at barrier S = SB, and t e r m i n a l condition at t = T are different for different cases: "knock-out" and "knock-in", "up" and "down", "call" and "put". (1) Domain Let D be the domain, going "up", the asset price must be below the barrier: D = {(S,t)\0<S <SB, 0
=Kanilla(SB,t);
(3) Terminal condition at t = T "knock-out call"
V W I 1 -(«-*>•.
( S / l l * . dT.n)
250
Mathematical Modeling and Methods of Option Pricing
"knock-in call" V(ST)
=0
(0<S<SB
up
\
(0<S<SB
up
\
"knock-out put" V(S T) - (K - 1+ V{b,I)-{K A) ,
[
S B
<
down)
S < 0 0
"knock-in put"
(°e-5f?B
V(S,T) = 0;
7 )
\ SB < S < oo
down J
(4) The partial differential equation All kinds of barrier options satisfy the same Black-Scholes equation,
In summary, barrier option pricing is a terminal-boundary problem of the Black-Scholes equation. Comparing to vanilla option pricing, it has one more boundary S = SB and a corresponding boundary condition. Its domain, boundary condition and terminal condition will be defined according to the type of the barrier option. Solution Take the down-and-out call option as example. We will find its closed form solution. Pricing problem of other types of barrier options can be solved similarly. Mathematical model of the European down-and-out call option is:
f^
+ 2 f f 2 5 2 0 + (r-i)S^-rV
[v(SB,t)=0,
= Q,
(D)
(8.1.2)
(SB<S
(8.1.3)
(0
(8.1.4)
where D = {SB < S < oo, 0 < t < T}. Under the transformation
x = \n-f,
(8.1.5)
V = SBu,
(8.1.6)
Path-Dependent Options (I) — Weakly Path-Dependent Options
251
the above problem becomes
§+^2§W-9-£)i-™=0-
(s-1-7)
(x€R+,0
+
(0 < x < oo)
,
u(0,t) = 0, where
(8.1.8) (8.1.9)
(0
K B = •#-.
Introduce a new function VK as follows: u = e°*+f>lT-*)w,
(8.1.10)
where a = -^(r-g-^),
By the transformation (8.1.10), we get a terminal-boundary value problem for W in { i £ R + l 0 < t < T } : (8.1.11) x
x
+
< W(x, T) = e-°' (e - KB) ,
(8.1.12)
[ w ( 0 , t ) = 0.
(8.1.13)
Applying the image method, define (e-ax(ex -KB)+,
x>0, (8.1.14) X<0.
Obviously
+ ^
\w(x,T)
=
0
(8.1.15) (8.1.16)
Since its solution must be an odd function, thus in D W(x,t) satisfies the boundary-terminal value problem (8.1.11)—(8.1.13).
252
Mathematical Modeling and Methods of Option Pricing
The solution of the Cauchy problem (8.1.15),(8.1.16) can be written in the form of the Poisson formula:
%^T-t)[/0V^^(-^ + /
Jo
,
=
e I'^T-V
1
[e'^'HT-t) _ e ~2^(T-t)l e -«<( e « _ KB) + d£-
/
[
cry/2ir(T-t)J0
'
V
(e
' -Ks)d?
;
Back to the original function u(x,t) by (8.1.10), we get -r(T-t)
u(Tlt)
[(*-{)+('•- »-^r)(r-Q]2
r°°
=ff% rrrt/
e
a^2ir(l — t) J\nKB -r(T-t)-^(r-q-^-)x _£ °2 ay/2n(T -1)
-lKBe-rV-t) V27T
r°° l(*+e)-<'-<,--£KT-t)]3 / e 2,HT-t) huKB
»-lnKB + (r-,-^-)(T-l)
f
^TW^f
(ei~KB)d4
^4^
7-oc
_e-»^-r(T-t) , V27T
' ^ ' ^
-n-lnK- B + (r--,+ ^ . ) ( T - t ) 7
7
W
^
^
^
7-00 -.-lnKg + (r- 9 -4)(T-')
+ 4a. e -'-(T-«)-^(r-,-^)« r V27T
y-oc
^V(^iy
g
_^^
By (8.1.5),(8.1.6), back to the original variable (S, t) and function V(S, t), we get V(S,t) = Se-vV-VNWi) - Ke-riT-»N(da) - 5B ( ^ ) ~ ^ < r " 9 ) e-"^T-^N(d3) -Ke—(^) 1 -- ( r -% ( , 4 ) = KanUla(5,t) - ( - ^ ) " ^ ^ ' ^ [^L e~ "^ ^ N (d3) - Ke~^^
N(d<)}, (8.1.17)
Path-Dependent Options (I) — Weakly Path-Dependent Options
253
where
ln-! + (r-*+£)(T-t)
di = —&
.
*
,
aVT^t d2 = di- ayjT
-1,
l n | | + (r-- g +^)(T-t) aVT-t di = dz — o~\/T — t.
The solution (8.1.17) can be written as V down _ and - out (S,i) = V™,iU.(5,0 - ( ^ ) ~ P r ( r " ' ) + 1 Kanilla(^,t). (8.1.18) By (8.1.1) we get Vdown-and-in(S',t)= f — 1
V v a n i l l a (-^, t).
(8.1.19)
Similarly we can obtain the valuation formulas for VUp_and-out(5', t) and 'up—and—in^j t).
In order to obtain the pricing formulas for put barrier options, we introduce the following call-put parity for the barrier options. Theorem 8.1 For down-and-out options, the following call-put parity holds Pdown-and-out(S,£) + SN(d!) = c d o w n _ a n d - o u t (5, t) + Ke~rlT-t)
N(d2)
^y^-^{slN{d3)_Ke-r(T-t)N{d4)l
+
(8.1.20) where
. \n^- + (r-q+^-)(T-t) P *= o ^ ' d2 = di — o~yT — t,
.
d3 =
l n ^ + ( r - , + ^-)(T-t) "
.
as/T-t di = d3 - o-VT - t.
*
,
Proof Let W be the difference between a down-and-out call and a downand-out put: W
=
Cdown—and— out ('-'j t) —Pdown —and—out (^j ^ ) -
254
Mathematical Modeling and Methods of Option Pricing
It is easy to show that W(S,t) satisfies the terminal-boundary value problem in D = {SB < S < oo, 0 < t < T}: OW + ^
S
^
+ ir-q)S™-rW
= 0,
{D)
• W(S,T) = S-K,
(SB<S
W(SB,t) = 0.
(0
Under the transformation (8.1.5) and W = SBU, the problem is reduced to
§ +^
2
0 + (r--ir)f§-™ = O, (z€R+, 0
" u(x,T) = ex-KB,
(0 < x < oo)
u(0,t)=0.
{0
The solution can be found to be
V27T
J_00 * + <.r—g-s£-HT-t)
V2TT6S
7-oo
^ ( r ._, ) a ! _, ( r-o
—
7 =
V2TT
/•
^7^
/
7_oo
y.-rCr-Q-tjKr-,)-!]. r SB
y/2^
^ 5
_ 2^
e
aw
e
_ ^ ^
J-oo
Back to the original variable (5, i) and W, we arrive at the conclusion of the theorem. Similarly we can prove call-put parity for other types of barrier options. By these parity formulas, we can readily write out the valuation formula for a put barrier option from that of the call barrier option.
Path-Dependent Options (I) — Weakly Path-Dependent Options 8.2
255
Time-Dependent Barrier Options
If the barrier changes with time, the option is called time-dependent barrier option. There are two types of time-dependent barrier options: (A) Moving barrier options Let barrier SB be a known function of time t, S = SB(t). In general, the price of this kind of options does not have a closed form expression, unless 5 s is an exponential function of t. Take a down-and-out call as example. In the domain D = {(S,t) \SB(t) < S < oo, 0 < t < T } , we solve the following terminal-boundary value problem of the Black-Scholes equation: + h°2S2^r
^
+ {r-
= 0,
(D)
' V(SB(t),t) = O, V{S,T) = (S- K)+. By the transformation
(8.2.1)
(0
(8.2.2)
(SB{T) < S < oo)
(8.2.3)
(8 L5)
x = ln
'
sk)'
(8.1.6)
V = SB(T)u,
the problem (8.2.1)—(8.2.3) is reduced to the following problem for u(x,t) in domain {x G R+, 0 < t < T}:
JW + 2 C T 2 0 + V-9-^~ ] u(O,t) = O, [u(x,0) = (ex-KB)
+
<*(t)]g| - ru = 0,
(8.2.4) (8.2.5) (8.2.6)
,
where K. - j
^
,
(8.2.7) (8.2.8)
If a(t) = a(a constant),
(8.2.9)
SB(t) = SB(T)e-^T-t).
(8.2.10)
then the solution of
is
256
Mathematical Modeling and Methods of Option Pricing
Then (8.2.4) becomes: (8.2.11)
where q - a + q. Then by (8.1.17), we get V(S,t) = 5e- ( « +tt)(T - t) Af(d]:) „ \ ~-K(r-q-a) + l
(
where
c
Ke-^'^Nidl) ..
S a2 „ _ In— + (r - q - a+ —)(T - t) d*2 = d\ln%l d3
(8.2.12)
ay/T^t, +
(r-q-a+^)(T-t) ay/T-t
'
d\ = d% — CTy/T — t.
(B) Partial barrier options The barrier is set up only during predetermined time windows in the option's lifetime. For an up-and-out partial barrier option, in days when the barrier is set up, the asset price St with respect to the barrier SB can be one of the following three cases: (a) St < SB- the underlying asset price stays below the barrier. (b) During the period, there exists t = to such that St0 = 5 B (the asset price hits the barrier). (c) St > SB'- the asset price crossed 5 s before the period and then stays above the barrier SB during the period. For up-and-out partial barrier options, if case (c) happens, there are two types of barrier options: Type (I): The barrier is active. For this type of options, the payoff function on the expiration date (take a call option as example) is payoff function = (ST - K)+I{St<Sgt
t€n},
where II denotes barrier active period, Iu is the characteristic function of w. Type (II): The barrier is inactive. It mean that according to the contract, the barrier is triggered only if the barrier is crossed, either from above or below, during the period.
Path-Dependent Options (I) — Weakly Path-Dependent Options For this type of options, the payoff function on the expiration date is
payoff function = (ST - K)+[I{St<sB,
ten} + I{st>sB,
ten}]-
Now we set up the partial differential equation models for these two types of barrier options. Consider an up-and-out partial barrier option. Its lifetime is [0, T\. Suppose the barrier is set up in the period II which consists of two intervals: [£i, £2] and [t3,T], where 0 < £1 < t2 < £3 < T, and corresponding trigger values are SB and 5B- Suppose the asset price change as follows:
I,
1
T
h
y
Ml O
<,
Sj,
Sg
S
Then in [£i,£2], for both Type(I) and Type (II) options, the barrier is considered triggered; But in [t3,T], the barrier is considered triggered for type (I) but not for type (II) options. In fact, the difference between type I and type II can be regarded as in the barrier levels. As shown in the figure, for type (II) partial barrier options, barriers are represented by the two vertical line segments at {S = S s , t i < t < £2} and {5 = 5B,£3 < t < T}, where the option price satisfies the knock-out boundary
257
258
Mathematical Modeling and Methods of Option Pricing I
*
J
-
T
(
~
"I I •""*
o
sB
h
n -
' barrier
ss s
Type (II) partial barrier options '
,"___:::i T
,
- barrier
Y_A
.
_s*^ I
^ \ barrier
Type (I) partial barrier options
conditions: V(SB,t) = 0,
(ti < t < t2)
(8.2.13)
V[SB,t) = 0.
(i3 < t < T)
(8.2.14)
For type (I) partial barrier options, in addition to the two vertical line segments {S = 5 B , t i < t < t2} and {S = SB,t3 < t < T}, where option price satisfies the knock-out boundary conditions (8.2.13), (8.2.14), barriers are set also on the two horizontal line segments {SB < S < oo, t = ti} and {5s < 5 < oo, t = £3}, where the option price satisfies the knock-out "boundary" conditions: V(S,U) = 0,
(SB<S
V(S, t3) = 0.
(SB<S<
< 00)
(8.2.15)
00)
(8.2.16)
and Therefore the pricing model for these two types of partial barrier options is: In [t3,T]: Type (I) Di = {0 < 5 < 5 S , £3 < t < T}: (£V = 0,
(Di)
I V{SB,t) = 0,
{ts
{ V{S,T) = (S-K)+;
(0<5<5 B )
Path-Dependent Options (I) — Weakly Path-Dependent Options
oo,
Type (II) I^i = {0 < S < SB, t3 < t < T} and D^ = {SB < S < t3
(A1!)
(cvx = o, |vi(5 B ,t) = 0,
(*s
[v1(S,T) = (S-K) , and
(0<S<SB)
r ^v 2 = o,
(A2I)
< V2(SB,t) = 0,
(h
{V2(S,T) = (S-K) ; I n [t 2 ,* 3 ]: Type (I)
D = {0 < 5 < oo,
(SB<S
t2
t3}:
CV = 0,
! Type (II)
(£»)
(V(S,t3 + O), V(S,t3) = \ I 0;
£>:
(0<S<SB) (5B
< S < oo)
f CV = 0,
(D)
< . (^1(5,^+0), V(S,t3) = < .
[
In {tut2}: Type (I)
(0<S<SB)
[v2(s,t3 + oy,
th. = {0 < S < SB,
(sB<s
h < t < t2}: (Di)
(£V = 0, < V(5 B ,t)=0,
ti
[v(s,t2)r=v(s,t2 + oy, oo,
259
Type (II) Dli = {0 < S < SB, ti < t < t2}:
(o<s<sB)
h < t < t2} and t>n = {SB < S <
(cvi = o,
(A 1 ,)
(h
\v1(S,t2) = V(S,t2 + 0)
(0<S<SB)
260
Mathematical Modeling and Methods of Option Pricing
and
InlO,*!]: Type (I)
f w2 = o,
(A2i)
< v2(sB,t) = o,
(t1
[ V2{S,t2) = V(S,t2 + 0);
{SB<S< oo)
D = {0 < S < oo,
i
CV = 0,
V(st1)
0)
= !V{S'tl
\ D = {0 < 5 < oo,
T y p e (II)
0 < t < ti}:
+ 0)>
0; 0 < t <
(°^5-5B)
( 5 S < S < oo)
CV = 0,
! In the above,
8.3
(£>)
mt l ) = {vr i(5itl+0)> (S,t +0); 2
1
S SB °S <S
Reset Options
Reset option is such a contract, when the underlying asset price reaches a predetermined level, the strike price will be reset, to give the option holder more profit opportunity. Such reset can occur once or multiple times. Reset options can be divided into two types: (I) Reset options with predetermined dates: The strike price can be reset on predetermined dates. Take a call option as example. Let 0 < ti < • • • < tn < T. On the initial day (t = 0), the strike price is set to K. At t = t\, if the asset price S(t\) is below K, then the strike price is reset to S(ti) (otherwise the original price K), and so on. In general, at t = t m (1 < m < n), the reset strike price Km is given by
Km = mm(Km-i,S(tm)), K0 = K. (II) Reset options with predetermined levels: The strike price can be reset to predetermined price levels.
Path-Dependent Options (I) — Weakly Path-Dependent Options
261
Take a call option as example. Let K > K\ > Ki > • • • > Kn. On the initial day (t = 0), the strike price is K. If the asset price falls to a predetermined level K\, then the strike price is reset to K\, and so on. In general, if the asset price falls from Km-\ to Km, then the strike price is reset to Km, (1 < m < n). Why do we discuss reset options with barrier options? In fact, a reset process is equivalent to a combination of two barrier options: knock out a barrier option at the "old" strike price, and simultaneously knock in a barrier option at the "new" strike price. Since the interpretation of the barrier is different for the two types of reset options, their mathematical models are different. For simplicity, we consider reset call options only, and allow only single reset. Type I reset options with predetermined dates. At ii € (0, T), if the asset price S(ti) < K (K is the strike price at initial date t = 0), then the strike price is reset to K = min(K,S(ti)). The barrier setting is illustrated in the figure below {0<S
barrier -—-i i
i
I I I I
0
j
7
In [ti,T], solve f CV = 0, \v(S,T) where CV = ^
(0 < S < oo, h < t < T) = (S-K)+,
(S£R+)
+ ! < T 2 S 2 0 + (r - q)S%& - rV.
The solution is of the form V = V(S,t;K). The asset price at t = ti is a known constant S(ti). If S(ti) > K, then K = K in [ii,T], and the option price is V = V(S,t;K), If 5(ti) < K, then K = S(ti) in [h,T], and the option price is V = V{S,t;S(t!)).
262
Mathematical Modeling and Methods of Option Pricing Thus at t = ti, we have (V(S(ti),tv,K), V(S(t1),t1) = \ lV(5(ti),ti;5(ti)),
S(t!)>K, S(h)
Define a function V(S) :
v<s) = {V{S'tuK)'
S
~K'
\v{S,tr,S),
S
In [t, ti], solve f £V = 0, [V(S,t1) = V(S).
(0 < 5 < oo, 0 < t < ti) (5€R+)
In particular, we can evaluate V(S, 0), since at each step the solution has a closed form expression constructed from the fundamental solutions of the BlackScholes operator.
Type II
reset options with predetermined levels.
If the asset price falls to K\ < K, then reset the strike price to K\. The barrier setting is
{S = KU
0
t< T
°
A:,
K
S
First, in D{0 < S < oo, 0 < t < T}, solve (D)
(£V = 0, +
\V(S,T) = {S-K1) , to find V = V{S,t;Ki).
(R+)
Path-Dependent Options (I) — Weakly Path-Dependent Options
263
Then the price for reset options with predetermined levels satisfies a terminal boundary problem in D{Ki <S
{
CV = 0,
(K! < S < oo, 0 < t < T)
V(K1,t) = V(K1,t;K1), V(S, T) = (S- K)+.
(0
It is easy to see that the reset option price V(S, t) can be decomposed into V(S,t) = Vdown-and-out(S,t;/S:;Kl) + Vdown-and-in (S,t; K\\ K\),
(8.3.1)
where Vdown-and-outfjn^S, t; K; SB) denotes the price of a down-and-out (in) barrier option with strike price K and barrier level 5 s . The financial meaning is that the price of this type of reset options can be decomposed into a knock-out barrier option at strike price K and a knock-in barrier option at strike price K\. Therefore, it has closed form solution. Remark The mathematical model and the solution procedure for the above two types of the reset options can be extended to the multiple reset options, although the expression will be rather complicated. Interested readers can refer [9] and [25].
8.4
Modified Barrier Options
A barrier option's premium is lower than its corresponding vanilla option, and is therefore more attractive to the investors. But barrier options create a number of problems for both option buyers and sellers : 1. The holder of a knock-out barrier option loses the investment if an asset price spike hits the barrier. 2. When the asset price approaches the barrier level, conflicting interests of the option holders and sellers tend to trigger short-term market manipulations. 3. A, the hedging share, is discontinuous at the barrier. When the asset price approaches the barrier, the seller must re-balance the hedging strategy. Various modifications to barrier options have been introduced to overcome these problems. We call them modified barrier options. In this section we will explain three types of modified barrier options. Take a down-and-out call option as example. The basic idea is: when the asset price falls across the barrier SB, the contract does not end abruptly. Instead, it goes through a phaseout process. There are several ways to describe this process.
264
Mathematical Modeling and Methods of Option Pricing
(1) When the asset price falls to the region S < S B , the option payoff decreases at a killing rate p . The payoff function on the expiration date is given by V(S,T) = (ST - K)+e~pTT,
(8.4.1)
where r± is the occupation time, which is the total amount of time in [0, t] when the asset price 5 is below 5 B (including 5 B itself), n = mes{t'|S(t') < SB, 0 < t' < t}, or (8.4.2)
rt = [ H(SB - ST)dT, Jo
where
H(S) = {1:
(8.4.3)
HI
This phaseout process indicates: although the option does not extinguish, its payoff decreases continuously during the occupation time. This type of barrier options is called step option. (2) Once the asset price falls to below S B , a clock starts to measure the time outside the barrier 5 B , SO that when the accumulated occupation time below SB (including SB) reaches a predetermined number of days D, the option extinguishes. The payoff of the option remains the same before the accumulated time reaches D. The payoff function on the expiration day is V(ST,T,TT)
= (ST - K)+I{TT
(8.4.4)
This type of option is called Parasian options. (3) When the asset price falls to SB , again a clock starts to measure the time outside the barrier SB- But the count method differs from (2). If the continuous occupation time below SB (excluding SB) reaches a predetermined time D, the option extinguishes. That means, if the asset price returns to SB before the continuous occupation time reaches D, then the clock will be reset to zero. ft, the continuous occupation time below SB (excluding SB), is given by Tt = t-gt
= t- sup{t'\S(t') >SB
0 < t' < t}.
(8.4.5)
The payoff function on the expiration day is V(ST,T,TT)
= (ST - K)+I{rT
(8.4.6)
This type of option is called Parisian options. Let us examine the difference between the accumulated occupation time Tt and the continuous occupation time ft.
265
Path-Dependent Options (I) — Weakly Path-Dependent Options
In the following figure, rt and ft axe drawn for an imagined asset price curve St and a given barrier level SB-
;IUAZ * M jvy|v[ ii i i i i i i
i
i
i
i
i
"Mil—[—I—i—-
*> M l I I I i i i i i l l I II I I r,J I I X
i I A I /I 1/ I X* '
i 1 1 I '
Remark Step option, Parasian option, Parisian option and barrier option are related to each other as follows:
option, option.
(1) As p —> oo, step option becomes barrier option, (2) As p —> 0, step option becomes vanilla option, (3) As D —> T, Parasian option and Parisian option become vanilla (4) As D —> 0, Parasian option and Parisian option become barrier
Take the down-and-out option as example. From (8.4.2) and (8.4.5), the definition of r t and f4 , it is easy to see drrL_dri_(0, dt dt \ l , Notice that at St = SB, ^ and ^ Suppose the option price is
St>SB, St<SB.
differ greatly (see above figure).
V = V(S,r,t),
(8.4.7) ( ;
266
Mathematical Modeling and Methods of Option Pricing
(here T denotes either Tt or ft.) We will apply the A-hedging technique by constructing a portfolio II:
n =v-
AS,
and choose A, such that II is risk-free in (t, t + dt). Since dU = dV - AdS, and by the Ito formula:
dV = ^dt + §£rfr + ^-S2^dt - (9V
a2 q2d2V\j+
,dVdT,
+ g^dS , dV
,q
therefore we choose Thus to satisfy dU = rlldt,
there must be:
(8.4.8)
As 5 < 5 s ,
(8.4.9)
% + % + &%+«%-"-
•
Next, let us analyze each of the three types of modified barrier options and find the mathematical model in each case (take the down-and-out option as example). (1) Step option Solution form : V = V(S,t,T); Domain: {0 < S < oo, 0 < t < T, 0 < T < T}; PDE: S > SB S < SB
, ,
Eq. (8.4.8), Eq. (8.4.9);
Terminal condition: V(S,T,T) = (S-K)
+
e~pT;
(8.4.10)
Boundary conditions: V(0,t,r) = 0, V ~ 5(as S -> oo);
(8.4.11) (8.4.12)
Path-Dependent Options (I) — Weakly Path-Dependent Options
267
Continuity conditions: V(SB + 0,t,r) = V(SB - 0,t,T), ^(SB+0,t,r)
= ^(SB-0,t,r).
(8.4.13) (8.4.14)
By (8.4.13),(8.4.14), equations (8.4.8) and (8.4.9) can be combined into
£ + w . _„£•£,£ +„£_,„_ a
(s..15)
(2)Parasian option Solution form: V = V(S, t, r); Domain: {0 < S < oo, 0
, ,
Eq. (8.4.8), Eq. (8.4.9);
Terminal condition: V(S,T,T) = (S-K)+;
(8.4.16)
Boundary conditions: V(S,t,D)=Q,
(8.4.17)
(8.4.11), (8.4.12); Continuity conditions: (8.4.13), (8.4.14); By (8.4.13),(8.4.14), equations (8.4.8) and (8.4.9) can be combined into (8.4.15).
(3) Parisian option
Solution form (V(S,t,r), \ V(S, t),
0<S<SB, <S < OO;
SB
(This is the difference between Parisian option and Parasian option in view of the different definitions of the occupation time r : when the asset price enters the region S > SB after crossing SB from below, for Parasian option, the occupation time retains its current value, but for Parisian option, the occupation time is reset to zero. Imagine the occupation time be measured by a stopwatch. When S crosses 5 s from below, for Parasian option, the stopwatch is paused, and for Parisian option, the stopwatch is not only stopped, but also reset to zero. Therefore when S > SB, the Parasian option's price V depends on r, but Parisian option's price V does not depend on T ) . Domain: { 0 < 5 < o o , 0 < * < T , 0 < T < D}\
268
Mathematical Modeling and Methods of Option Pricing PDE: S>SB, S < SB, Terminal condition: Boundary conditions: Continuity conditions:
Eq. (8.4.8), Eq. (8.4.9);
Eq. (8.4.16); Eqs. (8.4.17), (8.4.11) and (8.4.12);
V(SB, t) = V(SB,T, t) = V(SB,O, t),
(8.4.18)
||(SBlt)=fj(SB,<M);
(8.4.19)
Here we see again the difference between Parisian option and Parasian option. Equation V(SB,r,t) = V(SB,0,i) shows: when S returns to 5 s , the occupation time is reset to zero. In addition, only for r = 0, A = ^ - is continuous at S = SB- If T > 0, A is discontinuous in general. Now let us discuss how to find solution for each type of the barrier options. For step options, closed form expression can be obtained but in a rather complex form. Therefore we will discuss the numerical solution. Since step option and Parasian option both satisfy the same PDE (8.4.15), their numerical solution procedures are quite similar. Therefore we only need to study Parasian option's numerical calculation procedure. (A) Splitting method Partition the option's lifetime [0, T] into N equal intervals: 0 = t 0 < £i < • • • < tN = T, where tn = nAt, At = 4 . Parasian option In each interval [tn,tn+i], (0 < n < N — 1), suppose V(S,tn+i,r) is known, we want to find V(S, tn,r). First, solve the first order PDE terminal-boundary value problem(5 is a parameter, 0 < S < oo) in the domain {tn
{
^+H(SB-S)^
= 0,
{tn
V(S,tn+uT) = V(S,tn+UT), V{S,t,D)=0. The characteristic line of this equation is T = H(SB-S)(t-tn+1)
0
(0
For S < SB, the characteristic line is shown in the figure below:
Path-Dependent Options (I) — Weakly Path-Dependent Options
'iri-i
~7
269
-yi
Thus the solution of the above problem is : if T + H(SB - S)(tn+1 - t) < D, V(S,t,r) = V(S,tn+i,T + H(SB -S)(tn+i -*)), itT +
H(SB-S)(tn+i-t)>D, V(S,t,r) = 0.
Next, take V(x,t, T) as the initial value. In the domain {0 < S < oo, tn < t S tn+i}, solve the Black-Scholes equation(r is a parameter):
\v(S,tn+UT)
= V(S,tn,T),
Here ( V(S,tn+i,T + H{SB - S)At), \
0.
(0
(D - H(SB - S)At
Thus we have found the solution in the domain { 0 < S < o o , tn < t < tn+i, 0 < r
and in particular, V\t=tn=V{S,tn,T).
Since V(S,tff,T) = V(S,T,T) = (S-K)
+
,
by the backward induction, we can find the price of Parasian option in the whole domain {0 < S < oo, 0 < t < T, 0 < T < D}.
270
Mathematical Modeling and Methods of Option Pricing
Parisian option For Parisian option, V and A = ^ are discontinuous at the barrier S = SB for T > 0. Therefore the equations (8.4.8) and (8.4.9) cannot be combined into (8.4.15). So there will be a difference when using the splitting method. In each interval [t n ,i n + i], suppose V(S,tn+i)(0 < S < SB) and V(S, tn+i,T)(SB < S < oo) are given. We can rewrite it as V(S, tn+i, H(SB — S)T), where (1, S>0,
H(S) = { \o,
s
(Note the difference between H{S) and H(S) at S = 0.) First, solve the first order differential equation (S is a parameter, 0 < S < SB) in the domain {tn
n
• V(S,tn+uH(SB
- S)T) = V(S,tn+i,H(SB
V{S,t,H{SB-S)D)
~ S)r),
= 0.
The characteristic curve of the equation is T=(t-tn+i)+T*.
(0<5<5B)
Thus if r+(< n + i - t) < D, V(S, t, H(SB - S)T) = V(S, tn+i, H(SB -S){T+ if T + (tn+1
(t n + 1 - t)]),
~i)>D,
V(S, t, H(SB - S)T) = 0.
(tn
tn+i)
Thus we have V+ m«j QW^ jV(S,tn+uH(SB-S)(At V(S,tn,H{SB-S)T) ={ Q
+ r)), (0 < r < D - At) (D-At
Next, solve the Black-Scholes equation (here r is a parameter) in the domain {0 < 5 < oo, tn
l^| t =t B+1
=V(S,tn,H(SB-S)r).
Thus we have found the solution in the domain {0 < 5 < oo, tn < t < t n +i. 0 < T
V = V(S, t, H(SB - S)r).
(tn
tn+i)
271
Path-Dependent Options (I) — Weakly Path-Dependent Options In particular V\t=tn =
V(S,tn,H{SB-S)r).
Since V(S,tN,H(SB-S)T) = (S-K)+, thus using the backward induction, we can find the price of Parisian option in the whole domain {0 < S < oo, 0 < t < T, 0 < r < D}. (B)Finite Difference Method with Characteristic line Under the transformation x = \n—,
SBu(x,t,T) = V(S,t,T).
&B
(8.4.8) becomes du
^
+
<72 d2u
T^
,
+ (r
a1 ,du
-y)^-™
=0
'
...
.
.
.
(8A20)
as x < 0, and (8.4.9) becomes du
Tt
+
du
+
a2 d2u
.
Tr -2 3V
2 + (r
a2 du
" T>to- ru = °'
(8A21)
as a: < 0. Partition the domain {x € R, 0 < t < T, 0 < r < £)} as follows: xi
iAa;,
(i = 0, ± 1 , ± 2 , . . . )
tn=nAt,
(n = 0,1, ...,JV)
T( = eAr,
(e =
=
0,l,...,L)
where Parasian option Since when crossing the barrier (a; = 0, i.e. S = SB), both u and jr^ are continuous, therefore (8.4.20),(8.4.21) can be combined into
OX
(8.4.22) In order to discretize it , we use the finite difference scheme with characteristic line, draw a characteristic line in (tn+i,re) T = H(-x)(t-tn+l) then write the equation (8.4.22) as
+ Tt,
272
Mathematical Modeling and Methods of Option Pricing
-^u(x,t,H(-x)(t - tn+1) + re) r 2 2 L
i
1
OX
(8.4.23)
J T = H ( - j ) ( t - t n + 1 ) + T(
Discretize (8.4.23), denote u"e = u{xi,tn,Tt), ~ «i,< , a iui+i,t ~
"i,£
~^i
+ -2" I
+(r - ° )\uJ+ht
then for x > 0(i.e. i > 0),
Zu
+
i,e
u
i-\,i-\ J
^
A?
u
i-i,ei _
«
= 0
For x < 0(z < 0),
- rul^ = 0,
£)[^\Af-^]
uft = (e iAx - K B ) + ,
(t = 0, ± 1 , . . . ; t = 0 , 1 , . . . , L)
where KB = y - ; (i = 0, ± 1 , . . . ; n = 0 , 1 , . . . , AT)
ulL = 0.
Parisian option Denote u"0 = u(xi,tn,0) = u(xi,tn). Then for x > 0 (i > 0), n+l
"»,o
n
9
--»»,o +, a2
St
r
u
n+1
o n+1 , „ n+1 -|
i+i,Q-2ui,o
A?
T
+ui-i,o
(8.4.24)
for x < 0 (i < 0), n+1
n
A* 4- I r +
and u?e
ff
, f n+1 +
U
I ^+1.1 ~ i - l , <
= (eiAx - KB)+, L
=0,
n+1 , n+1 "I
A?
L
\r~-Tj I
<
^~
o
_,.n
J
(8.4.25)
_ rv
^Sx^ I " rUi'^ ~ °'
(i = 0, ± 1 , . . . ; I = 0 , 1 , . . . , L)
(» = 0 , ± l , . . . ; n = 0,l,...,7V)
(8.2.26) (8.4.27)
Path-Dependent Options (I) — Weakly Path-Dependent Options uo,e = v-o,o-
(1 = 0,1,...,D-l;
n = O,l,...,N)
273 (8.4.28)
Parisian option price can be calculated as follows: If u " / 1 is given, we can solve (8.4.24) to get u"0 (i > 0). Then by the continuity condition (8.4.28) at x = 0, we can find v%t. Finally, solve the equation (8.4.25), and with the boundary condition (8.4.27), we can find u"t, (i < 0).
Summary: (1) Unlike vanilla options, the mathematical model of the European barrier options is a terminal-boundary value problem for a parabolic PDE, where the barriers form the domain boundaries. The boundary conditions are imposed according to the knock-out or knock-in terms. (2) A knock-out barrier option extinguishes when the asset price hits the barrier. To avoid this abrupt change, an occupation time Tt(ft) is introduced in the modified barrier options, and the Black-Scholes equation is replaced by different types of hyperparabolic equations in three variables S,t and T. Since these equations contain only the first derivative of the solution with respect to r, thus it is natural to use the splitting method (to "split" the equation into a first order PDE and a second order equation) and the finite difference method with characteristic line (to form a finite difference scheme along the characteristics line of the first order PDE).
Exercises 1. A financial institute sells a callable European call option. That is, when the stock price reaches S = Sc (S c > K, K is the strike price), the seller has the right to repurchase the option at Sc — K. Valuate this option. 2. A financial institute sells a callable European down-and-out call option, with the barrier at S = SB and repurchase price S = S c , and SB < K < Sc
[K is the strike price)
Model this barrier option, and find the expression of the option price for the case r=\, 9 = 0. 3. Let V(S, t; SB) be the price of a down-and-out call option with barrier level S B . Show that if SB < K, then V(S, t; SB) is a monotone decreasing function of SB- Explain its financial meaning. 4. Formulate the call-put parity for the up-and-in options, and prove it. 5. Let Voo (S; SB) be the price of a perpetual American up-and-out put option, and assume S B > K. Valuate the option and find the optimal exercise boundary S = 7 ( S B ) . Show that If S B l > SBA> K), then Kx>(S;S Bl )> Voo(S;SB2), 7(SSI)<7(SB2).
274
Mathematical Modeling and Methods of Option Pricing
6. Let V(S, £;SB) be the price of an American up-and-out put option, and assume SB > K. Set up its price model, and show that V(S,t;SB)
Chapter 9
Path-Dependent Options (II) Strongly Path-Dependent Options The payoff of strongly path-dependent options depends not only on the underlying asset price on the expiration day, but also on the path of the asset price over the entire or part of the option's lifetime. There are two kinds of strong path-dependence: the option payoff may depend on the average or the maximum (minimum) of the asset price over some certain periods in the option's lifetime. For strongly path-dependent options, in addition to time t and the asset price St, the option price depends also on a path variable Jt, which can be either the path average of the price or the maximum (minimum) price up to time t. That is, V = V(S,J,t). This chapter is organized as follows: 1. Model PDE problems for the strongly path-dependent options; 2. Investigate whether the PDE problem can be reduce to a 1-D problem by a transformation £ = £(S, J,t) so that the option price only depends on t and the new variable £: V = V (£,£); 3. Find solutions: analytical solutions if it is possible, and approximate solutions by using the BTM. 9.1
Asian Options
Asian option is an option whose payoff at maturity depends on the average price of the underlying asset over the option's lifetime. Here the average can be either arithmetic average or geometric average. Let Jt, the path variable, denote the average asset price for the period up to t, then arithmetic average
geometric average
n
discrete case
Jn = ^Yl
St
i
Jn =
( n "=i 5 ti)" = e " E r = 1 ' n S t i
i=i
continuous case Jt = \{ I STd,T
Jo
J t = e « ^o l n S r d T
Correspondingly, there are two kinds of Asian options: arithmetic average Asian options and geometric average Asian options. Furthermore, there are two types of payoffs for the Asian options: 275
276
Mathematical Modeling and Methods of Option Pricing
(1) Fixed strike price(take the call options as example):
payoff =
(JT-K)+.
(2) Floating strike price (take the call options as example): payoff = ( 5 T - JT)+•
Then we have: Asian options with fixed strike price and Asian options with floating strike price accordingly. Thus without distinguishing call and put, there are four types of Asian options: arithmetic average Asian options with fixed strike price. geometric average Asian options with fixed strike price. arithmetic average Asian options with floating strike price. geometric average Asian options with floating strike price. Since the volatility of the average price is always less than the volatility of the individual price series that make up the average, the premium of an Asian option is usually less than that of the corresponding vanilla option. Example Consider a six-month call option of stock. If So = K = $145, r = 6%, q = 3%, a = 29.5%, then the premiums according to the Black-Scholes theory are European vanilla option Asian option (arithmetic/geometric)
Average period Premium $12.87 30days $12.16/12.03 60days $11.30/11.23 180days $7.31/7.13
The above table shows: (1) An Asian option costs less than its vanilla counterpart, and its premium tends to decrease with increasing average period. (2) The price of a geometric Asian option is always lower than its arithmetic counterpart. This is due to the inequality
{f[s{u))± <^£s(ti). Asian options are widely used in the international trading. Consider an import company that imports goods from a foreign country and selling them in domestic market. Suppose the company pays bills in foreign currency monthly, and calculates the earnings in home currency annually. Then the company faces a foreign exchange rate risk. It can choose either of the following plans to avoid the risk: (A) Buy 12 options, each pays one month's bill, at a predetermined currency exchange rate.
277
Path-Dependent Options (II) — Strongly Path-Dependent Options
(B) Buy a 12-month arithmetic average Asian option according to the whole year's need. We will use the following example to compare the premiums (calculated by the Black-Scholes theory): Example A US company imports Japanese products. It pays 1 million Yen to the Japanese company every month. Suppose the US company decides to lock at the current exchange rate 1USD = 85Yen. Assume r = 6%, q = 3%, a = 18%. Then for Plan (A), the premiums for the 12 options are $258, $372, $463, $540, $610, $674, $733, $789, $842, $892, $940, $987, adding up to $8100. For Plan (B), the premium of the 12-month Asian option is $5795. Both plans can help the US company to accomplish its goal, but at different costs. Plan (B) costs $2305 less than Plan (A), a 28.46% saving. For a foreign trade company, if cost in the home country's currency is known, then to avoid currency exchange risk, it usually chooses Asian options with fixed strike price to secure earnings. On the other hand, if cost and earning are both settled in foreign currency, and the concern is in the currency exchange risk at the end of the year when payments are sent to the home country, then it should choose Asian options with floating strike price. In general, Asian options with fixed strike price are more popular than Asian options with floating strike price. Since arithmetic average is simpler than geometric average, arithmetic Asian options are more favored than geometric Asian options in financial markets. However, since the asset price movement is described by the geometric Brownian motion in our price model, from the Black-Scholes theoretical viewpoint, the pricing problems of geometric average Asian options are simpler than those of arithmetic average Asian options.
9.2
Model and Simplification
The price of an Asian option is
V = V(St,Jt,t).
Construct a portfolio
n = v(s, J, t) - AS, where A can b e chosen such t h a t I I is risk-free in (t,t + dt), i.e.,
cffl = rUdt = r(V - AS)dt.
(9.2.1)
278
Mathematical Modeling and Methods of Option Pricing
By the Ito formula,
dn = dV - AdS - qASdt
= (w + 2 " 2 5 2 0 - o^dt + %dS + indJ - Ads
(9-2.2)
We can choose A
d v
A=
dS'
then combining (9.2.1) and (9.2.2), we get
(9.2.3)
where T / STdr, Jt = \
(arithmetic average) (9-2.4)
°ft e -/o
. (geometric average)
Therefore | j[St — Jt],
^P = < *
(arithmetic average)
[ j t [ l n 5 { - I n J t ] (geometric
aVerage)
(9-2.5)
Substituting (9.2.5) into (9.2.3), we get the model for the Asian option pricing:
Arithmetic average Asian option pricing model: a terminal-boundary
problem in the domain {0 < S < oo, 0 < J < oo, 0 < t < T}:
^
+
^%r
{
+
4 s ^
+
(r-q)sW-rV
= 0,
(J — K)+, (call option with fixed strike price) (K - J)+, (put option with fixed strike price)
(S — J)+, (call option with floating strike price) (J — S)+. (put option with floating strike price)
(9.2.6) ,Q 2 „
Path-Dependent Options (II) — Strongly Path-Dependent Options
279
Geometric average Asian option pricing model : a terminal-boundary problem in the domain {0 < 5 < oo, 0 < J < oo, 0 < t < T}: ' W + j\nS-t\nJdV
(
+
^s2^V
+ {r_
q)sdV
_rV
=
^
(J — K)+, (call option with fixed strike price) put option with fixed strike price) (K -J)+,(
(9.2.11)
(9.2.11)
(5 —J) + , ( call option with floating strike price) (J — S)+. ( put option with floating strike price) Equations (9.2.6),(9.2.8) are 2-D hyperparobolic equations. To solve the terminal-boundary problem (9.2.6),(9.2.7) and (9.2.8),(9.2.9), we will reduce them into a 1-D problem by suitable transformation. Fortunately, all Asian option pricing problems can be reduced to 1-D problems. However, only for the geometric Asian options the price has closed-form expressions. For simplicity, we will discuss the call option only. (A) Arithmetic average Asian option with fixed strike price Define ? = ^ , (9.2.10)
(9.2.11) By straightforward calculation, we have dV _ S \dU , dU ( JX\ dV _ SdU (
t\
W~T^ \S)' dV _U , SdU (U -TK\
-5S~T
+
T"5£ i—s2—/ '
82V _ 2 dU (U -TK\ , S 82U (tJ -TK\2
WTWK
s2 ) + T-^{
25 dU (U -TK\
s2 ) ~^W\
5a J'
Substituting these into (9.2.6), we get
(9.2.12) And by (9.2.7), we have the terminal condition:
=^(J_K)+ = (-0+.
U\t=T=^V i
t=T
b
(9.2.13)
Thus, under the transformation (9.2.10), (9.2.11), the arithmetic average Asian option with fixed strike price is reduced to a Cauchy problem (9.2.12),(9.2.13) for a 1-D parabolic equation in the domain {£ € R, 0 < t < T}.
280
Mathematical Modeling and Methods of Option Pricing (B) Arithmetic average Asian option with floating strike price Define
£=f,
(9-2.14)
V = SU.
(9.2.15)
By straightforward calculation, we have dV dV
q\dU
, dU fJ\]
tdU
dV _TI,qdU
( tJ\
d2V _ cd2U (tJ\2 Substituting them into (9.2.6), we get (9.2.16) and by (9.2.7), we have
U\t=T = I V | t = T = ( l - | ^ ) + |t=T = ( l - fj
+
.
(9.2.17)
Thus under the transformation (9.2.14),(9.2.15), the arithmetic average Asian option with floating strike price is reduced to a boundary-terminal value problem (9.2.16),(9.2.17) for a degenerate parabolic equation in the domain {0 < £ < oo,0 < t < T}. (C) Geometric average Asian option with fixed strike price Define i=tlaJ+(T-t)lnSt (921g) (9-2-19)
V = U(Z,t)By straightforward calculation, we have dV _ dU , dU [ l n J
inS]
dV _ dU t
W ~ ~B£TJ' dV _ T-t dU 92V _ (T-t\2
d2U _ T -t dU
Path-Dependent Options (II) — Strongly Path-Dependent Options
281
Substituting them into (9.2.8), we get
dU a2 (T-t\2
82U
cr2\(T-t\dU
(
(9.2.20)
and by (9.2.9), we have U\t=T = V\t=T = (J - K)+\t=T
= (e4 - K)+.
(9.2.21)
Thus by the transformation (9.2.18),(9.2.19), the geometric average Asian option with fixed strike price is reduced to a Cauchy problem (9.2.20),(9.2.21) in the domain {(, £ R, 0 < t < T}.
(D) Geometric average Asian option with floating strike price
Define ^=iln|,
(9.2.22)
V = SU.
(9.2.23)
By straightforward calculation, we have
dV _ qdU ( 1 ^
82V _ q82U M \
2
1 dU
Substituting them into (9.2.8), we get (9.2.24) and by (9.2.9), we have U\t=T = ±V\t=T = (1 - | ) + | t = r = (1 - eT«)+.
(9.2.25)
Thus by the transformation (9.2.22),(9.2.23), the geometric average Asian option with floating strike price is reduced to the Cauchy problem (9.2.24),(9.2.25) in the domain {£ € R, 0 < t < T}. Remark For geometric average Asian options with floating strike price, although under the transformation (9.2.22),(9.2.23), the original boundary-terminal value problem is reduced to a 1-D problem, yet it is not easy to get a closed form solution for the Cauchy problem (9.2.24),(9.2.25), because the coefficients of the first order term in the equation (9.2.24) contains not only variable t, but also variable £. Therefore, sometimes we still prefer using the transformation (9.2.18),
282
Mathematical Modeling and Methods of Option Pricing
(9.2.19), even though this transformation cannot reduce the geometric average Asian option with floating strike price (9.2.8),(9.2.9) to a 1-D problem (due to the condition (9.2.9), V(S, J, t) — (S — J)+ cannot be transformed to a function of £ only), but since the equation has a special structure analogous to a 1-D problem, there is a closed form solution. For a geometric average Asian option with floating strike price, let z = lnS, y=nnJ
+
(9.2.26)
(T-t)lnSt
(9227)
(9.2.28)
V = U(x,y,t). By straightforward calculation, the equation (9.2.8) is reduced to dU , a2 (T-t\2
d2U , 2 (T-t\
d2U
or dU
, a2 \ ( T - t \
d
,
W + ^lMvy
d}2Tr
+ 1*1"
(g229)
and by (9.2.9), we get U\t=T = (e* - ey) + .
(9.2.30)
Thus under the transformation (9.2.26)—(9.2.28), the geometric average Asian option with floating strike price is reduced to a Cauchy problem (9.2.29),(9.2.30) in the domain {(x,y) e R 2 ,0 < t < T}. (E) American—style Asian options with floating strike price Any American-style Asian option with floating strike price, whether arithmetic average or geometric average, can be transformed to a 1-D free boundary problem. (1) Arithmetic average case Consider the dividend paying American-style Asian options with floating strike price. The pricing model is (take the call option as example) an obstacle problem in the domain { 0 < S < o o , 0 < J < o o , 0 < t < T}: ( min{-£V,V -(S - J) + } = 0, +
\v\t=T = (S-J) , where
(9.2.31) (9.2.32)
Path-Dependent Options (II) — Strongly Path-Dependent Options
rv
cv
dV ^S-JdV +
^a2 +
2
2d
s
V
= -m — dJ T W +
283
dV
{r q)s
- ds~rV-
Let €=y,
(9-2.33)
V = SU.
(9.2.34)
Under the above transformation, the obstacle problem (9.2.31), (9.2.32) becomes a 1-D free boundary problem in the domain {£ g R + , 0 < t < T}:
{
min(-£i[/,t/-fl-f)"1"} =0,
(9.2.35)
t/|t=r(l-|.) + ,
(9.2.36)
where z „
£lU
dU , a2,2d2U
=^t+Y^
+
.
[1 {r q
du
TT
- - ^d-rqU-
(2) Geometric average case Consider the dividend paying American-style Asian options with floating strike price, the model is (take the call option as example) an obstacle problem in the domain {0 < S < oo,0 < J < oo, 0
Let
e=Jln|,
(9.2.39)
V = SU.
(9.2.40)
Under the above transformation, the obstacle problem (9.2.37),(9.2.38) becomes a 1-D free boundary problem in the domain {£ G R, 0 < t < T}: j min{-£il/, U - (1 - e^) + } = 0, \[/|t=T = (l-e^)+,
284
Mathematical Modeling and Methods of Option Pricing
where r
TT
dU
a2 1 d2U
lY
1 \dU
For American-style Asian options with fixed strike price, it is not known whether it can be transformed into to a 1-D problem. But this is known: American-style Asian options with fixed strike price cannot be reduced to a 1-D free boundary problem under the transformation (9.2.10),(9.2.11) and (9.2.18),(9.2.19).
9.3
Valuation Formula for European-Style Geometric Average Asian Option
Closed form valuation formulas exist for the European-style geometric average Asian options.
(A) Asian option with fixed strike price
Under the transformation (9.2.18),(9.2.19), the pricing problem becomes a Cauchy problem
( f + ^(^^
+ (' " « - ^ ) ( ^ ) f - rU = 0,
\ U\t=T = (e« - K)+.
(9.3.1) (9.3.2)
Let W = Ue0it),
(9.3.3)
t) = t + a(t),
(9.3.4)
r = 7 (t). Substituting them into (9.3.1), we get
Choose
<*'(t) + (r-q-£)(£f±) P'(t) + r = 0, 7 '(t)
= -(£^)2,
and impose the terminal conditions: a(T) = p(T) = 7 (T) = 0.
(9.3.5)
= 0,
Path-Dependent Options (II) — Strongly Path-Dependent Options
285
The solutions axe
<*(t) = ^r(r-q-5p)(T-t)2, P(t)=r(T-t), l(t)=^I{T-t)\ Under the transformation (9.3.3)—(9.3.5), the terminal-boundary value problem (9.3.1),(9.3.2) is turned into a Cauchy problem for the heat, equation:
{
a2 d2W _ n
dW
W(r,,0) =
(e»-K)+.
Its solution can be written in the form of the Poisson formula: W{T],T)
=
\
(ey - K) + e
/
2a2r dy.
ffVZTTT J - o o
Back to the original variables (S, J, t) and function V by the transformation (9.3.3)—(9.3.5) and (9.2.18),(9.2.19), we get V(S, J,t) = {JtST-t]± e^'+^-V
N(dl) - KN(d*2),
where
l l n ^ dl =
^2
=
a* =
+ (^ + a'y/T-t
V)(T- t ) '
d'l ~ cr*yT — t,
- t) V3T (B) Asian option with floating strike price Under the transformation (9.2.26)—(9.2.28), the problem (9.2.8),(9.2.9) is reduced to a Cauchy problem (9.2.29),(9.2.30) in the domain {(x,y) € R 2 ,0 < t < T}:
dU , a2 [ d | T - t d 1 2 JT ./r U
W + T [Ui + -^Vy-l !
U\t=T = (e x - e«) + =C/ 0 (z, y).
„ a2 \ \ d . T - t d 1 u,, rU rff n =0
+ ^ ~ q - -T) \5i + -TSy-\
-
>
(9.3.6) (9.3.7)
The Cauchy problem (9.3.6),(9.3.7) can be solved by the Fourier transform ([43])
286
Mathematical Modeling and Methods of Option Pricing
as follows. Let /•oo
/*oo
/
U(t,r,,t)=
U(x,y,t)e'l^x+Tly)dxdy,
J — oo J — oo
Applying the Fourier transform on both sides of (9.3.6), we get an ordinary differential equation for U as a function of t (£, 77 are parameters):
f-^ + ^,)V,( r -,-^)( £ + ^,)*-^o, ( ,, 8) Under the Fourier transform, the initial condition (9.3.7) becomes &{T) = Uo(Z,v)-
(9-3.9)
The solution of the ordinary differential equation (9.3.8) with the initial value (9.3.9) is U(£,T),t) = Uo(S,r)) exp{-(di£ 2 + 2d2^V + dijf) + i(d^ + d5r/) - d6}, (9.3.10) where
di = £(T-t),
d2 = 4^i, d4=
(r-q-£\(T-t),
d6 = r(T-t). Performing the inverse Fourier transform to (9.3.10), we get /•OO
/"OO
/
U(x,y,t)=
Uo(a,0)G(x-a,y-(3)dadp,
(9.3.11)
J — oo J — oo
where U0(a,(3) = (ea-ef))+, i
POO />OO
G(x, j/) = ——j / (2TT)
/
J—OO J - O O
.ei(x?+OT)^d7?
(9-3-12)
e x p { - ( d ^ 2 + 2d2(,n + d3v2) + ^(^4$ + d5n) ~ d6} (9.3.13)
Path-Dependent Options (II) — Strongly Path-Dependent Options
287
To evaluate the integral (9.3.13), change the integration variables as follows
Since
di£2 + 2d2£v + <W
thus the integral (9.3.13) can be written as
e" d6 /
G(x,j,)=—3-1
,»
f-
e
V
\/dTdI7
._.
\ \/2dk
y/2dk ) j£
J -00
=
d6
e~ An'JdKh-d* v
/d-4 + x exp
d5+y\
d5 + y\2'
/dj + x
_ V Vdl VT3 ) _ V VTr y/d£ ) (l + -^=) &(!--£=) 8 L V VdldlJ \ Vdld^J J
Substituting it into (9.3.11), and back to the original variables (S,J,t) and the function V(S,J,t) by the transformation (9.2.26),(9.2.28), we find V(S,J,t) = S[e- r(T - e) iV(- 52 ) -
jrSXf±e3°N(-g1)},
where
gi = ( j ( T 3 ^ 3 ) i / 2 ^ n j - ( r - g + ^ ) ( r 2 - t 2 )
+
92 = ^ 3 ^ 1 7 2 ["n ^ - (r - g + ^)(T 2 - t2)].
^ ^ !
]
288 9.4
Mathematical Modeling and Methods of Option Pricing Call-Put Parities for Asian Options
Based on the pricing models of Asian options, we can find the call-put parities for them. (A) Call-put parity for arithmetic average Asian option with fixed strike price Let C(S, J, i) and P(S, J, t) denote the price of an Asian call and put option, respectively. Define W{S,J,t) = C(S,J,t)-P(S,J,t), Then in {0 < 5 < oo, 0 < J < oo, 0 < t < T}, W satisfies
J^ + ^ f f + ^ f ^ - ^ l f - ^ O . +
{ W \t=T= (J - K) ~ (K - J)+ = J - K.
(9-4.1) (9.4.2)
Under the transformation (9.2.10), the function W=^W
(9.4.3)
satisfies the Cauchy problem in the domain { ( £ R , 0 < t < T}:
fd4-+ 4e^f - Kr -te + l]1%-1# = 0 \ w ( € , T ) = -€.
«GR)
(9.4.4) (9.4.5)
Let W(S,t) = a(t)£ + b(J,).
(9.4.6)
Substituting it into (9.4.4), (9.4.5), and comparing the coefficients, we get: ' a'{t) - ra(t) = 0, b'(t) - a{t) - qb(t) = 0, ' o(T) = - l , b{T) = 0. The solutions are (r ^ q) a(t) = - e - ^ - O ,
6(0 = F ^[e- r ( T - t ) -«-' ( T -' ) ].
289
Path-Dependent Options (II) — Strongly Path-Dependent Options Substituting them into (9.4.6), and by (9.2.10) and (9.4.3), we get (r # q): C(S, J, t) - P{S, J, t) = S[a(t)TK^U = [tj l T
_
S _ K]e-r(T-t) (r — q)T '
+
_
h{t)]
S -,(r-t) (r — q)T
(9 4 7) v
(B) Call-put parity for arithmetic average Asian option with floating strike price Define W{S, J, t) = C(S, J, t) - P(S, J, t). Then in {0 < S < oo, 0 < J < oo, 0 < t < T}, W satisfies (9.4.1) and the terminal condition W \t=T= (S - J)+ - (J - S)+ = S - J.
(9.4.8)
By the transformation (9.2.14), the function (9.4.9)
W = ^W satisfies the Cauchy problem in {0 < f < oo, 0 < t < T}:
J 94- + 4
^ + I* - ( r - 1 ) ^ ~1W = O,
(9.4.10) (9.4.11)
[W \t=T=l-^Let
(9.4.12)
W = a(t)£ + b(t), Substituting into (9.4.10),(9.4.11), and comparing the coefficients, we get ' a'{t) - ra{t) = 0, b'(t) + a(t)-qb{t) ' a(T) = - l , . b(T) = 1. Solve them to get (r ^ q) a(t) = ~ ^
T
- * \
= O,
'
290
Mathematical Modeling and Methods of Option Pricing
Substituting them into (9.4.12), and by (9.2.14) and (9.4.9), we get (r ^ q): C(S, J, t) - P(S, J, t) = 5[o(t) y + 6(t)]
= -^-P(r-0 + nhff^
+ 1
1 ~ w=*\ s^T't)-
^
(C) Call-put parity for geometric average Asian option with fixed strike price Let W(S, J, i) = C(S, J, t) - P(S, J, t). Thus in {0 < S < oo, 0 < J < oo, 0 < t < T}, W satisfies the initial value problem:
| dW + jlnS^nJ
dW + £p!£w
\ W \t=r= {J - K)+ -{K-J)+
+ (r
__ q)s§W _ rW
= Q>
{94U) (9.4.14)
(9.4.15)
= J-K.
By the transformation (9.2.18), in {^ 6 R, 0 < t < T} W satisfies :
(W\t=T=ei -K.
(9.4.17)
Let W = o(t)e€ + b(t).
(9.4.18)
Substituting them into (9.4.16),(9.4.17), comparing the coefficients, we get
a'(t) + \ (£f+ya(t) +(r-q-4)
(^-f1) a{$ ~ ra^ = °>
6'(O-r6(O = O, a(T) = 1, 6(T) =
-K.
Solve them to get:
a(t) = e
L r
T
6(0 = - ^ - < - ' » .
J,
291
Path-Dependent Options (II) — Strongly Path-Dependent Options Substituting into (9.4.18), and by (9.2.18), we get C(S,J,t) -P(S,J,i) = a(t)JT'SZ^1 +b(t) = ^e
L
e" r ( T -''.
U T S T - K
(9.4.19) (D) Call-put parity for geometric average Asian option with floating strike price Let W{S, J, t) = C(S, J, t) - P(S, J, t), Thus in {0 < S < oo, 0 < J < oo, 0 < t < T}, W satisfies
(
dW , AnS -\nJ dW , a2 g2d2W TV \t=T= (S - J)+ ~(J-S)+
.
rr
n K
W
,-w _ n
CQ 4 9 m (9.4.21)
= S-J.
Under the transformation (9.2.26)—(9.2.28), the function W satisfies the Cauchy problem in the domain {(x, y) e R 2 , 0 < t < T}: dW , a2 (T-t\2 i o2 d2W , („
d2W , 2 (T-t\ o2\ [T-tdW
d2W , dW]
w
n
(9 4 22)
and by (9.2.9), W\t=T = ex - ey.
(9.4.23)
W(x, y) = a(t)e x - b(t)ey.
(9.4.24)
Let
Substituting it into (9.4.22),(9.4.23), by comparing the coefficients we get ' a'{t) - qa(t) = 0,
« 6 ' w + 4 1 2 1 ^ ) 2 fow+(r -1 - 4 ) l 21 ^ 1 ) &w - r6 w=°a(T) = 1, . b(T) = 1.
292
Mathematical Modeling and Methods of Option Pricing
Solve them to get a(t) = e -«( T - f ),
Substituting it into (9.4.24), and by (9.2.26),(9.2.27), we get C(S,J,t) - P(S, J,t) = S e - ^ - j r s ^ e ^ ^ 9.5
1
^""^'"
2
-"<*•-'>.
Lookback Option
A lookback option allows its holder to "look back" at the underlying asset price history on the expiration day, then choose the lowest (highest) price as the strike price to buy(sell) the underlying asset. Thus its payoff on the expiration date (t = T) is payoff = ST — min St (call option) 0
= max St — ST (put option) o
(9.5.1)
V
y
'
(9.5.2) '
Therefore lookback options are called "buy at the low, sell at the high" options, or standard lookback options. According to our classification of the pathdependent options, they should be called "lookback call(put) option with floating strike price". Options of this type are expensive because of high payoffs. However, the lookback call option is worthless if the asset price hits the minimum at expiration. In this case, the option holder has paid a dear premium for nothing. Example Consider a 6-month stock call option. Suppose So = K = $145, r = 6%,g = 3%, a = 29.5%, the premiums calculated by the Black-Scholes theory are: premium European vanilla option $12.87 lookback option $23.12 In this example, the premium of the lookback option is nearly twice that of its vanilla counterpart! The payoff of lookback options, which is a typical type of the strongly pathdependent options, depends strongly on the underlying asset's price history in the option's lifetime. The path dependency variable J is Jt = min ST or Jt = max ST. 0
0
(9.5.3)
V
'
It follows from the definition of the path variable that, for the lookback call
Path-Dependent Options (II) — Strongly Path-Dependent Options
293
option, there is min ST,
St>Jt=
(9.5.4)
whereas for the lookback put options, there is St < Jt = max ST. ~
(9.5.5)
0
Pricing model(take the put option as example) Let V denote the price of a lookback put option, V = V(S,J,t), where J is defined in (9.5.3). To apply the A-hedging technique, construct a portfolio II:
n = v - AS, Choose A , such t h a t II is risk-free in (t,t + dt), i.e.:
dU = rlldt.
(9.5.6)
By the Ito formula: <m=dV - AdS - qSAdt
- \~dr + 2
as2" ~ g S )
+
"57
+ ( | | - A) dS.
(9.5.7)
As the path dependency variable Jt defined in (9.5.3) is not differentiable with respect to t, in order to find dJt, we need to approximate Jt by Jn{t) as follows:
Jn(t)= ^j\Sr)ndrY
.
Obviously Jn(t) is differentiable with respect to t, =
nJT'it)^
5 r
~tJ"W-
(9-5-8)
Moreover, as n —» oo, since St is a continuous function of t, thus there is lim Jn(t) = max ST = Jt. n —oo
V
'
0
(9.5.9)
294
Mathematical Modeling and Methods of Option Pricing
Using Jn(t) in place of the path variable Jt, assuming S < Jn, thus by (9.5.8), the equation (9.5.7) can be written as
(9.5.10)
Choose A-dV Then by (9.5.10),(9.5.6), we get / 5 \ n-l s Jn dv [XJ ~ dv i a 2d2v aT + nt dX + 2aSW
dv - ds-rV
+ {r q)s
{S < Jn)
= 0(9.5.11)
For a given (J,t), let n —> oo, then by (9.5.9), the coefficient of ^ ¥ - in equation (9.5.11) approaches 0. Thus we have (9.5.12) (0 < S < J < oo, 0 < t < T) By (9.5.1), we get V(S,J,T) = J-S,
(9.5.13)
Since {5 = J, 0 < J < oo, 0 < t < T} is the boundary of the domain, we need a boundary condition: dV -QJ \S=J= 0.
(9.5.14)
Its financial meaning is clear: the option price is not sensitive to the level of the maximum of the asset price. Therefore the price model for the lookback put option is the terminalboundary problem (9.5.12)—(9.5.14) in the domain { 0 < 5 < J < o o , 0< t < T}. Notice that the path variable J does not show up in (9.5.12). But J appears in the boundary condition (9.5.14) and the terminal condition (9.5.13). Similarly we can formulate the price model for the lookback call options, which is the following problem in the domain { 0 < J < 5 < o o , 0
f +^ 0
+ (-^f
- ^ = 0,
(9.5.15)
V{S, J,T) = S-J,
(0 < J < S < oo),
(9.5.16)
\s=j= 0.
(0 < J < oo)
(9.5.17)
^
Path-Dependent Options (II) — Strongly Path-Dependent Options
295
For the European lookback options, the pricing problem can be reduced to a 1-D problem, and has a closed-form solution. To be specific, let us consider the lookback put options. Let z = ln^,
V = Su(x,t).
(9.5.18)
By straightforward calculation, we get
d2V
'dS2'
=
1 \_du , d2u]
~5 [ ^x &?J '
Substituting them into (9.5.12), we get
'ft + £ 0 + («-r-£)fe-*" = °.
<9-5-19>
(0 < x < oo, 0 < t < T) u \t=T = ex - 1, | f |x=o = 0 .
(0 < x < oo) (0 < t < T)
(9.5.20) (9.5.21)
Let (9.5.22)
u = e^+^-^W, where
(3=-q-±z[a2-2{q-r)Y. we get
^
(*6R+,0
+% 0 = ° .
• W \t=T = (e x - l ) e - Q X ,
(x € R + )
. [w+HLo = °-
(0
We will solve it using the image method. However, straightforward continuation by symmetry (odd or even) does not work with the boundary condition (9.5.25). Therefore we define
*(,) = {«-"(«--1)'
*>°.
|_
x < 0.
(9.5.26)
v
'
296
Mathematical Modeling and Methods of Option Pricing
and consider the Cauchy problem:
l^dr
+ g
T^r
(*eR, 0
= °>
[ W _ = *(x).
(9.5.27)
(x G R)
(9.5.28)
We need to choose tp{x), such that the solution W(x,t) to the Cauchy problem (9.5.27),(9.5.28) satisfies the boundary condition (9.5.25) at x = 0
^+alf
L dX
J x=o
=0.
(9.5.29)
If this can be accomplished, then by the uniqueness of the solution of the problem (9.5.23)—(9.5.25), we can assert (x > 0, 0 < t < T)
W(x,t) = W(x, t).
(9.5.30)
By the Poisson formula, the solution to (9.2.27),(9.2.28) can be written as r+oc
W(x,t)=
J — oo
r(x-£,*)$(£)#,
(9.5.31)
where T(x — £,£) is the fundamental solution of the equation (9.5.27):
l
r(x-^,t) =
2a{T-t)
(0
Thus we have
dW
f°°
%£ = ] rx(x - z,t)Q(t)ds = -/ oo r e (i-€,t)*(0de J — oo
= -r(x, t)^(o) + /
J — oo
r(x - e, *)*'(Ode-
Path-Dependent Options (II) — Strongly Path-Dependent Options
297
Then (9.5.29) can be expressed as follows:
= -r(o,*Mo) + r rfct)[$'K) + Q $KM J -oo
= -r(o,t)v(O) + /
>/—oo
r(£,t)fo/(O + M £ M
T(t, t)[(l - « ) e ( 1 " Q ) ? + a e - Q ? + ae (1 — £) - ae- Q? ]df
+ r
= -r(0,*M0)+ / °
ela-1H]dt.
T(Z,t)[
J — oo
Choose ip{x) that satisfies the following equation and boundary condition:
I v(0) = o. If 2a — 1 7^ 0, i.e. r =f= q, then the solution is
^
)
=
2^T[e
Substituting it into (9.5.31),(9.5.30), we get
W(x,t) = J° T(x - £, t) [^-L^e-** - eC-DC)] di + I" T(x - (,t)[e^-aK - e-ai]d4. Jo
In the case r = q,
-xeax,
W{x,t) = - f° T(x-Z,t)ZeaidZ J — oo
+ r T(x - U)[e{1~aK Jo
e~a^.
Back to the original variables (S,J,t) and function V by (9.5.22),(9.5.18), we get the valuation formula for the lookback put options:
298
Mathematical Modeling and Methods of Option Pricing
In the case r ^ q,
V(S,J,t) = Je-^-* U(60 -\{^)9
^(-63)I
In the case r = q, [iV(62) + In (^j N(-b3) - y (T - t)N(-b2)
V(S, J, t) = -Se-^-*
_«_v^ie-b4]+Je-HT-t)N{bi)t V2TT
J
where
ln| + (-r + g + ^)(r-t) (TVJ — t b2 = 61 - <7A/T - t ,
(7
Similarly, we can get the pricing formula for the lookback call options: In the case r 7^ q,
V(S,J,t) = -Je-^ T -" [iV(a2) - i ( ^ " ^ ( - o a ) ] +Se-^-t)[Ar(a1)-^(-a1)]; In the case r — q, V(S, J, t) = S e - r ( T - 4 ) U ( O l ) + ^ ( 7 ) W(-<>3) + y ( T - t)AT(-ai) 2"
r=—e V2TT
where
^
— Je
l
'N(a2),
Path-Dependent Options (II) — Strongly Path-Dependent Options
299
\n£+(r-q+£\(T-t) ai = o>2
=
aVT^t 0,1 — cjyT
'
— t,
. 2(r - q)s/T - t a3 = -01 + - i ^ , a Remark Having studied the lookback options which belong to the pathdependent option with floating strike price, it is natural to ask a question, how to price lookback options with fixed strike price, whose payoff function at maturity is given by payoff = ( max St — K)+
(call option)
or payoff = (K — min St)+• (put option) Its pricing formula can be obtained by using PDE approach. However, since its derivation is rather complicated and this type of options is not popular in the financial market, we will not discuss it here. Remark The lookback options are closely related to the reset options which we discussed in the previous chapter. In fact, it is obvious that if 0 = to < ti < ••• < tff = T are the reset times, and At = max \U+\ — U\, then as At —> 0, the reset option will become the lookback option. Therefore, the reset option pricing can be approximated by a lookback option pricing, as the valuation formula of lookback options is much simpler than that of reset options. American lookback options According to the arbitrage-free principle, the model for American lookback options is (take the put option as example) in the domain {0 < S < oo, S < J, 0 < t < T}, V = V(S, J, t) that satisfies min{-£V; V - (J - S)} = 0, • ^j
_ = 0,
V\t=T = J-S, Here
(0 < 5 < J < oo, 0 < t < T) (0 < J < oo, 0 < t < T) (0 < S < J < oo)
300
Mathematical Modeling and Methods of Option Pricing
This is a 2-D problem. By the transformation (9.5.18), it can be reduced to a 1-D problem. That is, in the domain {0 < a: < oo, 0
{
min{-£ o w,u - (ex - 1)} = 0,
(x € R + , 0 < t < T)
§Lo =0 '
(o
u(x,T) = ex-l,
(xeR.+)
where .
du
CoU=
+ a
m2^
1 2d2u
+
(
o2\ du
q r
{ - -Y)dx--qu-
Its equivalent free boundary problem is: find {u(x,t), X(t)}, such that in the domain {0 < x < X(t), 0 < t < T}, it satisfies: £ou = 0, (0 < x < X(t), 0
-1,
• ||(X(t),t) = e x ( t ', |^(0,i) = 0,
ri
2
2d
u
(
(0
u(x, T) = ex - 1. Since
(0
(0 < x < X(T)) o2\du
ll
thus from the discussion of the properties of the free boundary in §6.5, we know that X{T) = m a x ( l n ^ , o ) and X(£)monotone decreasing. Substituting it into the original variables (5, J, t), thus for the American lookback put options, the optimal exercise boundary T has the following properties: (1) For any given t, Tt is a line: J = (2) In the case t = T, if r > q, TT • J = S, if r
Sex(t);
J = $S;
(3) As T - t increases, the continuation region {0 < 4 < e x ( t > } expands (see the figure)
Path-Dependent Options (II) — Strongly Path-Dependent Options
301
(4) As T — t increases indefinitely, ex^ has an upper limit, which is the optimal exercise boundary of the perpetual American lookback put option. The proof is analogous to what we did in §6.5.
J
° 9.6
r
Numerical Methods
Although some types of the path-dependent option pricing problems (such as geometric average Asian options, lookback options) have closed form valuation formula, numerical methods are often prefered. The most important and widely used numerical method is the binomial tree method (BTM). In this section, we will first introduce the BTM, then the finite difference schemes with characteristic line, and finally establish the equivalence of the two methods.
A. Binomial tree methods for the path-dependent options
Partition the option lifetime [0, T] into N equally spaced nodes a distance At apart: 0 = t0 < ti < • • • < tN = T,
where tn = nAt, (0 < n < N), At = j ^ . Construct the BTM process for the random variables S (the underlying asset price ) and J (the path variable) as follows: Let 5 " = SQUZCT~Z (0 < i < n) denote the possible prices of the underlying asset at tn, J"k, (k € / " ) denotes the possible values of the path variable at tn when the underlying asset price is S?, where / " is the set of indices of the path variable corresponding to 5 " . Suppose at t = tn,S?,J"k are known, what will be their possible values at t = tn+i?
302
Mathematical Modeling and Methods of Option Pricing
By the BTM process(d < 1 < u): qn+l _ qn
qn
s^
S?+1 = Snd.
^ ^
To find out the corresponding changes in "&, we consider the following three cases: Suppose the path of the underlying asset price from So at t — 0 to S" at t = tn is {So,Sl1,...,S^~_\,S^}. In [t n ,£ n +i], the asset price can change from 5 " to either S"^ 1 or 5™+1, thus (1) corresponding arithmetic average
n
1=1
changes to 1 7
i.e.
n+1
<
-
f n 1 ) X~* * a. 9 n + 1 I
Tn-f 1
n
rn
1 Qn+^
,
"/i+l,fcu — n+1 J».k "^ n + l a i + l
(9.6.1) rn+l J
n
in
,
i,fcd — n+l-'i.fc"1"
(2) corresponding geometric average n 1=1
1 pn4-l. n+
l°i
>
Path-Dependent Options (II) — Strongly Path-Dependent Options changes to
<
303
J?+i,k» = Wk) * W+i 1 ) ^ (9.6.2)
^
1
= (^)T*T(^n+1)^T;
(3) corresponding maximum(minimum)(to be specific, take the maximum as example) J^max^
changes to
<
j£1)ku
= max{ J?ik, ST+Y } (9.6.3)
r n + l _ 7» • / t,fc d -
J
i,fc-
Remark Notice that the BTM process of the variable J"fc is pathdependent. For each 5", the corresponding path variable J " = {^"^j^e/" is a 2 " ~ ' dimensional vector in general (there may be overlaps among its components). Once the BTM process of the asset price and the path variable is established, we can derive the BTM algorithm for the path-dependent option pricing using the A-hedging technique, which is based on the arbitrage-free principle. For each one-period and two-state process on+l
rn+l
jn J
i,k
an+1 ^i '
the corresponding process of the option price is
^"^ < ^ ^ ~ \ ^ jn+1 ^•Ji,kd '
304
Mathematical Modeling and Methods of Option Pricing yn+l
V
"k
< ^
^ " ~ \ Vn+l ^
V
i,kd
'
where
| KkV = V(S?+1, J$d\tn+x),
(9.6.4)
Construct a portfolio n = V - A5, and choose A, such that in [ti, U+i], II is risk-free. Then we get
V-n+l
i+i,fc« ~
The solution is
A rtn+1 Ai>
T/-U+1
A rtn+1
i + i = ^i.fc,, - A&i
/T/TI
A c?n\
= p ( K i i f c - A&i ) .
*ft = J K t k + (1 - QWfc1],
(9.6.5)
where 1 9 = ^ . ~ 9 = ^ (9-6.6) it — d w—d Algorithm Before we calculate the path-dependent option price, we first need to form the binomial tree of the asset price and the path variable. In particular, we need to input all of the information about each asset's price S" and corresponding path vector ,/", its components J™k and their corresponding J
i+l,ku
a n aJ
i,kd •
At t = tN = T, according to the given payoff function, {V^}fc€7jv, (0 < i < N) is given. In order to obtain {ViNk~1}keIN-i, (0 < i < N - 1): Step 1 For each path dependent index k € J ^ " 1 at t = tN-i,S = S ^ " 1 , find the indices ku € lf+\ andfed€ i f among the path vectors J^+i and Jj^ corresponding to S = S f and 51 = ^/^j at i = t^Step 2 Among {Vi+i,k}kei^ and {V^} fce/ w, find the components Ji+i,ku a n ( i Ji!kd corresponding to indices ku,kd, and calculate the corresponding prices VHlku and V^kd. Step 3 Calculate V^k~l by the formula (9.6.5)
Path-Dependent Options (II) — Strongly Path-Dependent Options
305
Then by the backward induction, we get all prices of the path-dependent option, especially the premium on the initial date. Remark Although the above algorithm is not very complex, it may require large amount of processor time and memory space. Note that there is a 2 n ~ dimensional vector J " at each node S", and for each vector J™ we must store its components and correspondence information between J ^ 1 and J". When n is large, computation can be very costly. Therefore more efficient modified algorithms are needed, especially for arithmetic average Asian options. We will not discuss it here. Readers can refer to [22] [23].
B. Finite difference scheme with methods of characteristic line To be specific, we only consider the arithmetic average Asian options, i.e. the problem
{% + *¥•% + &*& + **%-" = *> \
(0 < S < oo, 0 < J < oo,
[V(S,J,T) = f(S,J).
<«-^
0
(0 < S < oo,0 < J
(9.6.8)
Partition [0, T] into TV equally spaced nodes a distance Ai apart: 0 = to < ti < • • • < tN - T, where tn = nAt, At = -Jy. In each interval [t n ,t n +i], consider the following first-order partial differential operator
8V_ dt
S-JdV t dJ'
We look for a characteristic line starting from (Jo,in), that satisfies the initial value problem: ( dt _ dJ I J(tn) = JO. Its solution is J = S-j(S-J0). As well-known, along this characteristic line, there is
£**r^->.*-t(S-*,,,
(9.6.9)
306
Mathematical Modeling and Methods of Option Pricing
(because J o = S + ^-(J - S).) Thus the equation (9.6.7) can be rewritten as
(9.6.10) We want to write the terms inside the above brackets in terms of derivatives along the characteristic line. Let -TJJ denote the derivative with respect to S along the characteristic line, then _rf_ _ _9_ dS~8S+
t-tn t
d dJ'
J!L_-(JL dS2 ~ \dS
t tn
~
t
d
V ~dj) '
Thus the equation (9.6.10) can be written as
iv{S,S-tf(S-J0),t) \a2 q2 (r,t-tn
d2V
+
}1o*S*£V+rS%-rV
, (t-tn\2
_rSt-hLdV] 1
aJ
d2v\
= 0
(9.6.11)
lj=S-±f-(S-J0)
Since our goal is to derive the discrete form of the equation (9.6.7), we can assume S2j£Xj,
S
2
^ , sQj are bounded. Thus we have (9.6.12)
where ^ a n d -4n denote respectively derivatives with respect to t and 5 along the characteristic line. Further discretize S. For any given Ax > 0, let u = eAx, and denote Si = ui - eiAx, (i = 0, ± 1 , . . . ) . Thus we get a partition in 0 < S < oo. By the explicit FDS: t = tn+i :
t = tn:
5
*-i
s
&+i
i
s
.
Path-Dependent Options (II) — Strongly Path-Dependent Options
307
Let Jo = t/fc, the characteristic line starting from t = tn, J = Jk is (9.6.13)
J = S-j(S-Jk). Thus along this characteristic line, when t = tn, S = Si V(tn,Si,Jk) = V^k ( V"k is defined in (9.6.4)); when t = tn+i,S = Si+i,Si J?+i,ka) = Vi\-i,k^>
V(tn+i,Si+i, V(t T//.
i,
and Si-i,
*?• Tn+1\ — Vn+1 C
rn+1 \
-irn+l
where jn+l _ -/i+l,fcu -
1 n
+
1
jn+l
c j_ l^i+l +
Q
i^
n
J7
n+\
Tl
k,
j
Consider the discrete form f9dF\ \ ds)s=s,
Fi+1 - Fj-i In5 i+ i-ln5i_i' (sdF\
\°15O~S5) s=s. ~ =
( dF\
In6'i+i -inA'^i - IFj + Fj i
Fi+i
(In5i+i-ln5i)'2 "
Here, according to the assumption, we have
ln^=1
*
l n |±i
Since 54 = 6^*,
l n % i = Ax.
(9.6.14)
thus the explicit FDS for the equation (9.6.7) along this characteristic line can
308
Mathematical Modeling and Methods of Option Pricing
be written as 1/1+1 _ yn v v i,kp i,k .
2 g
9T/n+l
\f7n+l
, l/n+1 l
liXx
I r ~~ °~ / ^ [T/"+I 2Sx
^ i+l,feu
(9.6.15) _ yn+i
1 _ r i / n -I- O
M-l,fedJ
rK
t,fe + ^ ( . i i i ; — U.
After rearrangement, neglecting the higher orders, it can be written as
V& = (1 + r A t ) - { f 1 + ^ r
/
2
(r - ^ ) ] V»+X\u
\i
/
2
\
^
(9-616)
This is the explicit FDS with characteristic line for the equation (9.6.7). Since V^k is known, we can find V£k, (0 < n < N) from the difference equation (9.6.16) by the backward induction. It is easy to show that &*>
(9.6.17)
and 9 6 18
5-2A^-YI>0<-- ) then the scheme (9.6.16) is stable. 2. As Ax, At —* 0, the limit of the difference scheme (9.6.15) is the equation (9.6.7), i.e.the scheme (9.6.16) is consistent with (9.6.7). Then by the Lax's equivalence theorem we assert: under the conditions (9.6.17),(9.6.18), as At, Ax -> 0, the solution by the FDS (9.6.16) converges to the solution of the original terminal-boundary problem (9.6.7), (9.6.8). Lemma / /
gf-l
( 9 .6. 19 ,
and
i-l\r-^\VAt>0,
(9.6.20)
then the FDS (9.6.16) is reduced to
VP,k = -pfeV&lfc,.+ (1 - g)Vf_lU where
(9.6.21)
p = 1 + r-At,
,_i(1 + (,.£)#).
(9.6.22)
Path-Dependent Options (II) — Strongly Path-Dependent Options
309
As we have pointed out in §5.7: u = eaVEi,
d= e-<^,
r
/
2\1
p = erAt j
(9.6.23)
By comparing (9.6.21) and (9.6.5), and by (9.6.23), we have proved: When conditions (9.6.19), (9.6.20) hold, neglecting the higher orders of At, the BTM for the arithmetic average Asian option is equivalent to the explicit FDS with characteristic line for the equations (9.6.7), (9.6.8). Therefore by the convergence of the explicit FDS with characteristic line, we claim that the BTM for the arithmetic average Asian option is a well-posed algorithm. Summery (1) The model of the Asian option pricing is a terminal value problem for a 2-D hyperparabolic equation. All Asian option pricing problems can be reduced to a 1-D problem under suitable transformations. But only the geometric average Asian options have closed form valuation formula. (2) BTM is the most popular numerical method for Asian option pricing. When neglecting the higher order terms, it is equivalent to the FDS with characteristic line (when a,'f — 1 / Notice that for the arithmetic average Asian options, since corresponding to each asset price S™, its path variable J£ is a 2n~l dimensional vector, its computation costs a lot of memory and processor time. (3) The model for the lookback put options is a terminal-boundary problem for the Black-Scholes equation in the domain {0 < S < J < oo,0
Exercises 1. Let Vt denote the price of an Asian option whose payoff at maturity t = T is payoff = ( J ? - * • ) + , where
1 fT JT = — STdr w JT-U
(arithmetic average)
310 or
Mathematical Modeling and Methods of Option Pricing
J£ = e i f?-« l n S T l J r ,
(geometric average)
where w is a given time period, ui < T. Set up the pricing models for these two Asian options. In the case ui = -^, find the valuation formula for the geometric average Asian options. 2. Let S\(t),S2{t) be the prices of two risky assets. Their movement satisfies the stochastic differential equations (7.1.5)—(7.1.8). Let Vt denote the price of an Asian minimum call option on two risky assets, i.e., its payoff at maturity is given by payoB={ndn{jP,J™)-K)+, where j \ is the geometric average of the risky asset price Si(t) jW =c i/o'ln Si (r)dr_
(
•= ^ 2 )
(a) Set up the pricing model for the option. (b) Transform the original 4-D problem to a 2-D problem. 3. Let St be the price of a continuous dividend-paying stock, with the dividend rate q > 0. A Russian option is such a perpetual American option that allows its holder to exercise it at any time t > 0, at the maximum value of the asset price in its path: Mt = max ST0
(a) Formulate the free boundary problem model for the Russian option pricing. (b) Find the price and the optimal exercise boundary for the Russian option. 4. Show that: when neglecting the higher order terms of At, the BTM for the geometric average Asian option is equivalent to the FDS with characteristic line under certain conditions. 5. Formulate the explicit FDS for the lookback options, and show that: when neglecting the higher order terms of At, it is equivalent to the BTM under certain conditions.
Chapter 10
Implied Volatility
Volatility a is a crutial parameter in the Black-Scholes formula. The option price is very sensitive to change in this parameter. In an underlying asset market, investors would definitely like to know how volatile the future price of the asset will be. Indeed, a is in general unpredictable. However, the options market "knows" it. We can obtain and analyze prices of a number of options on a given underlying asset with various strike prices and expiration dates. If the BlackScholes option pricing theory is correct, then the price quotes from the options market should reflect and reveal the price volatility of the underlying asset. The volatility derived from the quoted price of a single option is called the implied volatility. In this chapter, we study how to recover the underlying asset price movement (under the risk-neutral measure) from the information of the options market using the Black-Scholes theory i.e. to infer the implied volatility of the underlying asset price. 10.1
Preliminaries
In deriving the Black-Scholes formula, we have assumed the volatility a of the underlying asset to be constant throughout the option's entire lifetime. By the Black-Scholes formula, the option price is of the form: V = V(S,t;a,K,T). Suppose we have learned from the market that at t = to, the price of an asset is S = So, and the price of its option with the strike price KQ and expiration date To is Vo- Then by the Black-Scholes formula we can get an equation for a: Vo = V(So,to;cr,Ko,To). Since
?>°' da
311
(10.1.1)
312
Mathematical Modeling and Methods of Option Pricing
a can be uniquely determined from (10.1.1). Thus we have inferred the implied volatility
a
<s
k
| O
^
1
K/So
—
^7^
J
Here So is the asset price at t — to. The first figure is called volatility smile curve, and the second figure is called volatility skew curve. For a given strike price K, the implied volatility a varies with the maturity T typically in one of the following ways: A
a
•
T These pictures indicate, the constant volatility assumption in the BlackScholes formula does not reflect the reality.
Implied Volatility
313
A more realistic model is: IT is a function of t and asset price S. Then the stochastic process of the asset price movement under the risk-neutral measure is ^
= (r - q)dt + a(S, t)dWu
(10.1.2)
Correspondingly, the Black-Scholes equation becomes (10.1.3) In general, this type of problem has no explicit solution. It must be solved by numerical methods. With the improved price model of the underlying asset, we ask a question: how can we determine the volatility of an underlying asset price from its option price quotes in the options market? Mathematically, i.e. at t = to, S = So, if V{S0,t0;a,Kk,Tl)
(k = 1,... ,m,l = 1,... ,n)
= Vk,i
(10.1.4)
are given, find the volatility function
(10.1.5)
( 0 < S < o o , 0 < t
(0<S
Suppose at t = t*, (0 < t* < Ti), S ~ S', V(S*,t*;o-,K,T) = F{K,T)
(0 < K < oo,^
is given, find a = a(S,t), (0 < S < oo,Ti < t < T2)
10.2
Dupire Method
Let V = V{S,t;K,T)
(10.1.6)
314
Mathematical Modeling and Methods of Option Pricing
be a European call option price. Define 82V ^=G(S,t;K,T).
(10.2.1)
By (10.1.5),(10.1.6), G satisfies
J W + 2 * 2 ( 5 . * ) 5 2 0 + ( r " )5fS -rG = 0,
(10.2.2) (10.2.3)
{G(S,t) = 8(K-S).
Since 5(x) = 8(—x), thus G(S,t;K,T) is the fundamental solution to equation (10.2.2). By Theorem 6.3: G(S,t;K,T), as a function of K,T(S,t are parameters), is the fundamental solution to the adjoint equation of the problem (10.2.2),(10.2.3), i.e. ~W + \j^(02{K,T)K2G) - {r-q)-^K{KG)-rG = 0, (10.2.4)
(0
(10.2.5)
(0
Substitute (10.2.1) into (10.2.4),(10.2.5), then integrate twice with respect to K in [K, oo]. Since (1) for a given S, if K —> oo, V, KM,a /•oo
(2)
/
JK
/-oo
<% /
J$
JK
(r? - K)S(V - 5)d7?
= /°°(7j - K)+«5(r, - 5)A/ = (5 - K)+, Jo f°° d2V dV
I G(S,t;t,T)« = fK
(4)
(6)
K G) - 0,
/>oo
«(»/ - 5 ) ^ = /
f°°
(3)
(5)
K G, K m , ^(a
J"^V(S,t;t,T)dt /•oo
^ tG(X,t;Z,Tn
W«=-m>
= -V(S,t;K,T), /-oo O 2 T ^
= jK ^Wd4=-K-
au
+
V,
Jx^r ^a*(rl>T)v2G)dv = °2(K,'nKaj£-
315
Implied Volatility Thus the problem (10.2.4),(10.2.5) becomes
f -8V + l t f V ( t f , T ) | ^ - (r - q)K$fe -qV = 0, <
(10.2.6)
(0 < K
[ V\T=t = (S - K)+.
(10.2.7)
{0
From (10.2.6), we get ([14])
a{K,T)=
&£x
\|
5-^
i aV 2 ax2
.
(10.2.8)
Therefore, at t — t*, 5 = 5*, if we can get from the options market the option price quotes with various strike prices and expiration dates, i.e. if we know the function F(K,T): F{K,T) = V(S*,t*;K,T), (10.2.9) then we can calculate a(K,T) by (10.2.8). However, this algorithm is not robust to the real data and is thus not reliable. In fact, for a given F(K, T), in order to calculate a(K, T) by (10.2.8), we need to calculate the derivatives FKK,FK and FT- But a small error in F can result in big changes in its derivatives, especially in its second derivatives. Therefore the algorithm to compute a(K, T) by (10.2.8) is ill-posed. As we pointed out in (10.1.4), in general, F(K, T) is given on a set of discrete points {(Kk,Ti)}(k = l,...,m,l = l , . . . , n ) . Thus interpolation or extrapolation technique would be required to obtain a continuous function F(K,T) in the domain (0 < K < oo,Ti < T < T2) from the values at discrete points. However, naive interpolation and extrapolation tend to incur irregularity and instability in the solution a(K,T). A more robust calibration method is hence needed. Although Dupire method is not practical, nevertheless, we can follow its idea in solving this ill-posed problem. That is, we still want to reduce the implied volatility determination problem to a terminal state observation problem for a parabolic equation. Since the latter is a typical inverse problem, we can always find a well-posed algorithm for it.
10.3
Optimal Control Method
As a first step, let us assume a(S, t) =
316
Mathematical Modeling and Methods of Option Pricing
the function V(S, t*) is not given in the entire domain (0 < S < oo). Instead, V(S*,t*; K) is given as a function of the parameter K (0 < K < oo), at a single point of S ( S = S"). Nevertheless, following Dupire's idea, we have reduced Problem P to a typical terminal state observation problem. Problem P o Let V = V(K, T; a) be a solution to the Cauchy problem
l W = y{K)K2|V
_ (r _ q)KW _ qV>
(10 3 1}
\v(K,0) = (S-K)+,
(10.3.2)
where T = T — t. And suppose at r = T* — T — t*, V(K, r*) is given by V(K,T*;a)
(0 < K < oo)
= F(K),
(10.3.3)
find a = cr(K). Problem Po is a typical terminal state observation problem (here the terminal time is r = r*). By transformation y = \n^,v=^e"TV,
(10.3.4)
problem (10.3.1),(10.3.2) becomes (10.3.5) \v(y,0) = (l-eV)+,
(10.3.6)
(yeR)
where
(10.3.7)
a(y) = l^(K). ProblemQo
Find a(y) € A, such that J(a) = /
(10.3.8)
J(a),
where v = v(y,T;a) is the solution to the Cauchy problem (10.3.5),(10.3.6),
J(") = \f z
\v(y,r*-a)-v*(y)\2dy+^ Z
JR T
y
v*(y)=-^ei 'F(S'e ),
(Function F{K) is defined in (10.3.3)).
[ \Va\2dy,
(10.3.9)
JR
(10.3.10)
317
Implied Volatility
.4 is the admissible set of the variational problem (10.3.8). In order that the problem (10.3.5), (10.3.6) has solution, we choose the set as a(y) < au f \Va\2dy < oo},
A = {a(y)\0
(10.3.11)
JR
where ao,ai and N are constants. Here J(a) is called the cost functional, o = a(y) is called the control variable, and a(y) is called the optimal control or minimizer. The variational problem Qo is called the optimal control problem. By the parabolic equations theory, it can be shown (see[24j) Theorem 10.1 The variational problem Qo has at least one minimizer a(y) € A. Next we establish the necessary condition for the minimizer a(y). Let a(y) be the minimizer, since A is a convex set, thus for arbitrary h(y) £ A, ax(y) = (1 - X)a + Xh £ A, (X € [0,1]). Define a function j(X): j(X) = J((l-X)a
+ Xh),
which attains the minimum at A = 0, thus
0 < j'(0) = [ ^ j \vx(y,r') - v'(y)\2dy +
N
i I |V((l-A)a + A/O|adj/l
dA
jR
where v\(y,r)
,
(10.3.12)
JA=O
satisfies
/ ^ = -Afo)[^-^]-(r-,)^, y
\vx(y,0) = (l-e )
+
dO.3.13) (10.3.14)
.
Denote £\{y,T) = ^ ^ - , by straightforward computation, we get
•fc-«M[0-%]-<'-«>$+[^-fc]l»->. (10.3.15) $A(J/,O)=O.
(10.3.16)
Thus (10.3.12) can be written as / (v(y, T*)-V* {y))£o{y, T*)dy + N [ Va • V(/i - a)dy > 0,
JR
JR
V/i € A (10.3.17)
318
Mathematical Modeling and Methods of Option Pricing
where v(y,r) is a solution to the problem (10.3.5),(10.3.6), where a(y) = a(y). io{y,T*) is the solution to the problem (10.3.15),(10.3.16), where A = 0, i.e.
•*-£-.<,>[0-Si]Hr-.)$-[#-g]<»-.). '
(10.3.18) £0(2/,0) = 0.
(10.3.19)
Let
{
£*<£ = 0,
problem of the problem
(y € R, 0 < r < T*)
V(y,T')=v(y,T*)-v'(v), (yeR) where the differential operator C* is the adjoint operator of C:
d(p d2
^, _ C
=
(10.3.20) (10.3.21)
_ d d
Thus by the Green formula
[T [ ( ^ o - Z0C*v)dydT
JO JR
+ |-(€oa(^) + (r-g)^}d»dr = / [(
[ L* [ S ~ S] {k~ a)dydT = Jjo(-y'T*Wyy)-V*^dy- (10-3-22) Substituting (10.3.22) into (10.3.17), we get
fT I v \ f t - ? 1 (h ~ ^dvdT + N I va • V(A - a)dv > o.
Vh € .A
319
Implied Volatility Denote
/( )=
(
)dr
(10323)
^ 4f^ 0-S '
--
then the above inequality can be written as
[ [Va-V(h-a) + f(y;v,
V/i 6 A
(10.3.24)
Choose a sufficiently smooth \(y), and let 0 < X(y) < 1,
(10.3.25)
and let h(y) = [(1 - X(y))]a(y) + X(y)ai. (i = 0,1) By the definition of the admissible set A (10.3.11), for any X(y) satisfying (10.3.25), h(y) e A. Thus from (10.3.24), for all X(y) satisfying condition (10.3.25), there must be / {VaV[X(y)(ai - o)] + /(»; v,
for all X(y),
then there must be [-a"(v) + f(y,v,
(t = 0,1)
(10.3.26)
Since a 0 < a < oi,
(10.3.27)
thus by (10.3.26) we get \-a"(y) + f(y; v,
(10.3.28)
l-a"(y) + f(y; v, ^)](ai - a) > 0. (10.3.29) Thus we have shown the variational inequality (10.3.24) and (10.3.27)(10.3.29) are equivalent. This is a double obstacle variational inequality problem. Thus we have established the necessary condition of the variational problem Qo. Theorem 10.2 / / a(y) £ A is the minimizer of the variational problem Qo (10.3.8), then there exists a triplet {a(y),v(y,T),
|8-*(v>0+*<*)g + (r-l)g=O,
(10.3.30)
\v(y,0) = (l-eV) + ,
(10.3.31)
320
Mathematical Modeling and Methods of Option Pricing
I -?£ ~ a 7 ( i W ) ~ ilvMvW ~ ( r ~ ^ = °>
(10-3-32)
I
(10.3.33)
and (10.3.27)—(10.3.29). Here v*(y) and f{y;v,(p) are defined in (10.3.4) and (10.3.23). Remark Equation (10.3.30) with the initial condition (10.3.31) is a Cauchy problem to a forward parabolic equation; Equation (10.3.32) with the terminal condition (10.3.33) is a Cauchy problem to a backward parabolic equation; Equation (10.3.27)—(10.3.29) is a variational inequality to a second order ODE. Therefore this is a system of forward-backward parabolic equations coupled with an elliptic variational inequality. The uniqueness of the solution to the above system of equations is investigated in [24]. 10.4
Numerical Method
In order to find the numerical solution to the problem Po, there are two approaches: (1) starting from the variational problem Qo(10.3.8), discretize it, convert it to a convex constraint optimization problem, then find its approximate solution; (2) Starting from the necessary condition of the variational problem, i.e. to find the triplet {a(y),v(y,T),(p(y,T)}, such that it satisfies the forward-backward PDE problem coupled with an elliptic variational inequality (10.3.27)—(10.3.33). In the following we explain the procedure for solving the problem (10.3.27)— (10.3.33). We use iterative procedure as follows: ([26]) (1) Suppose at t = t*, S = S*, the option price quotes with different strike price K and the same expiration date, V(S*,t*;K) = F(K), are obtained from the options market. By (10.3.4) define function «*(*)= ^ e ' T > ( 5 * c » ) ,
(10.4.1)
where T* =T-t*,y = ln Jr(2) Choose initial value a = ao(K) for iteration. It can be the implied volatility determined by the Black-Scholes formula, i.e. tofinda = ao(K) so that it satisfies the following transcendental equation: e-«T-f)s.N
M
+ Ke-rv-nN (*$
y
+
(r-q+^)(T-n\
+ lr-,-*P)V-n\ ao(K)VT -t*
J
=
(10.4.2)
321
Implied Volatility
Due to the strict monotonicity of the option price with respect to a, equation (10.4.2) has a unique solution a =
(10.4.3)
Solve the Cauchy problem (10.3.30), (10.3.31), where a(y) is chosen as: a(y) =ao(y),
where ao(y) is defined in (10.4.3). Thus we obtain a solution vo(y, r). (4) Solve the backward Cauchy problem (10.3.32), (10.3.33), where a(v) =o.o(y),
=vo(y,r)-v*(y),
here v*(y) is defined in (10.4.1). Thus we get a solution (po(y,r). (5) By definition (10.3.23), compute f(y;vo,tfio)'-
/(y; -o, <*) = 1 [' My, r) [ ^ - ^ ] dr.
(10.44)
(6) Solve the variational inequality (10.3.27)—(10.3.29), where f(y; v,
^2an(\n^).
In steps (3),(4),(6), we can use the difference method or other numerical methods to solve the Cauchy problem to parabolic equations and the boundary problem to an ordinary differential equation. Remark The information obtained from the options market is usually in the form of finite discrete points. Thus in order to define the function F(K) = V(S*,t*;K) in 0 < K < oo, we need to use interpolation and extrapolation to convert the discrete data to a continuous function in 0 < K < oo. This conversion introduces errors and sufficient care must be given when performing the calculation.
322
Mathematical Modeling and Methods of Option Pricing
Summary 1. Since the implied volatility obtained by the Black-Scholes formula shows "smile" and "skew", we must modify the price movement model of the underlying asset, i.e., assume the volatility a is a function of S,t. 2. Following Dupire's idea, we convert the implied volatility (function) cr(S,t) determination problem to a terminal state observation problem to a parabolic equation. Then in the framework of the optimal control theory, we have given a well-posed algorithm for seeking the implied volatility (function).
Bibliography
[I] Bachelier, L. (1900) Theorie de la speculation, Ann. Sci. Ecole Norm. Sup., 17, 21-86. [2] Barles, G., Burdean, J., Romano, M. & Samsoen, N.(1995) Critical stock price near expiration, Math. Finan. , 5, 77-95. [3] Black, P.& Scholes, M.(1973) The pricing of options and corporate liabilities.J.polit.Econ., 81, 637-654. [4] Brennan, M. & Schwartz, E.(1977) The valuation of American put options, J.Finan., 32, 449-462. [5] Broadie, M.&Deltemple, J.(1997) The valuation of American options on multiple assets, Math. Finan., 7, 241-286. [6] Carr, P.&Faguet, D.(1994) Fast accurate valuation of American options, Working paper, Cornell University. [7] Carr, P., Jaxrow, R. & Myneni, R.(1992) Alternative charaterization of American put options, Math. Finan., 2, 87-106. [8] Chen, X., Chadam, J., Jiang, L.&Zheng W.(2001) Convexity of exercise boundary of the American put option for no dividend asset, Working paper, Tongji University. [9] Cheng, W.&Zhang, S.(2000) The analysis of reset options, The Journal of Derivatives, 59-71. [10] Chesney, M., Cornwall, J., Jeanhlanc-Picque, M., Kentwell, G. & Yor, M.(1997) Parisian pricing, Risk, 10, (January) , 77-79. [II] Cox, J.&Ross, S.(1976) The valuation of options for alternative stochastic processes. J.Finan.Econ., 3, 145-166. [12] Cox, J., Ross, S.&Rubinstein, M.(1979) Option pricing: A simplified approach, J. Finan. Econ.7, 229-263. [13] Dewynne, J.N., Howison, S.D., Rupf, I.&Wilmott, P.(1993) Some mathematical results in the pricing of American options , Europ. J. Appl. Math., 4, 381-398. [14] Dupire, B.(1994) Pricing with a smile, Risk, 7, (January) , 18-20. [15] Friedman, Avner.(1982) Variational Principles and Free Broundary Problems, John Wiley& Sons, New York. [16] Goldman, M.B., Sozin, H.B.&Gatto, M.A.(1979) Path dependent options:buy 323
324
Mathematical Modeling and Methods of Option Pricing
at the low, sell at the high, J.Finan., 34, 1111-1128. [17] Hull, J. (2000) Options, Furtures and Other Derivatives, Fourth Edition, Prentice-Hall. [18] Jacka, S.D. (1991) Optimal stopping and the American put, Math. Finan., 1, 1-14. [19] Jaillet, P., Lamberton, D.&Lapeyre, B.(1990) Variatinal inequalities and pricing of American options, Acta Appl. Math., 21, 263-289. [20] Jiang, L.(2002) Analysis of pricing American options on the maximum(minimum) of two risk assets , Interface &Free Boundaries, 4, 27-46. [21] Jiang, L.& Dai, M.(2004) Convergence of the explicit difference scheme and the binomial tree method for American options, J.Comp.Math., 22, 371380. [22] Jiang, L.&Dai, M.(2000) On path-dependent options, Mathematical FinanceTheory ^Applications, Higher Education Press, Beijing, 290-317. [23] Jiang, L.&Dai, M.(2004) Convergence of binomial tree method for European/American path-dependent options, SIAM J. Numer. Anal., 42, 10941109. [24] Jiang, L.&Tao Y.(2001) Identifying the volatility of underlying assets from option prices, Inverse Problems, 17, 137-155. [25] Jiang, L., Yang D.&: Zhang, S.(2004) On pricing model of reset option with N predetermined levels, J. System Sci. & Complexity, 17, 137-142. [26] Jiang, L., Chen, Q., Wang, L. &Zhang, J.E.(2003) A new well-posed algorithm to recover implied volatility, Quant. Finan., 3 451-457. [27] Karatzas, I.&Shreve, S.(1998) Methods of Mathematical Finance , Springer Verlag, Berlin, Heidelberg, New York. [28] Kwok, Y.-K.(1998) Mathematical Models of Financial Derivatives, Springer Verlag, Berlin, Heidelberg, New York. [29] Linetsky, V.(1999) Step options, Math.Finan.9, 55-96. [30] Margrale, W.(1978) The value of an option to exchange one asset for another, J.Finan., 33, 177-186. [31] McDonald, R.L., Schroder, M.D.(1998) A parity result for American options, J.Comp.Finan., 3, 5-13. [32] Merton, R.C.(1973) Theory of rational option pricing, Bell J.Econ.Manag. Sci., 4, 141-183. [33] Merton, R.C.(1990) Continuous Time Finance, Basil Blackwell, Oxford. [34] Musiela, M.&Rutkowski, M.(1998) Martingale Methods in Financial Modelling, Springer Verlag, Berlin, Heidelberg, New York. [35] Richtmyer, R.D.(1957) Difference Methods for Initial-Value Problems, Interscience Publishers, Inc., New York. [36] Rogers, L.C.G.&Shi, Z.(1995) The value of an Asian option, J.Appl.Prob., 32, 1077-1088. [37] Robinstein, M.(1991) One for another, Risk, 4, (December) , 30-32. [38] Stulz, R.M.(1982) Options on the minimum or the maximum of two risky assets, J.Finan.Econ., 10 , 161-185. [39] Wan, N., (2003) The asymptotic expression of optimal exercise boundary near expiry for American options when r = g, Working paper, Tongji University.
Bibliography
325
[40] Wilmott, P., Dewynne, J.N. Howison, S.D.(1993) Option Pricing: Mathematical Models and Computation, Oxford Financial Press, Oxford. [41] Wilmott, P.(1998) Derivatives: The Theory and Practice of Financial Engineering, John Wiley &Sons, Inc., New York. [42] Zhang, P.G.(1997) Exotic Options, World Scientific Publishing Co.Pte. Ltd., Singapore. [43] Zhang, Q.(1998) A theory study of path-dependent options with general payoff, Working paper, City Univ. of Hong Kong.
Index
A-hedging, 27
with fixed strike price, 276 arithmetic average Asian options with floating strike price, 276 Asian options with fixed strike price, 276 Asian options with floating strike price, 276 asset-or nothing call, 88
American options, 4 arbitrage, 8 geometric average Asian options with floating strike price, 276 out of the money, 110 put option, 3 risk-neutral world, 30
barrier, 247 barrier options, 247 basket options, 210 Bermudan, 193 Better-of options, 210 binary option, 88 binomial tree method, 25, 301 binomial tree methods for path-dependent options, 301
a call option on a call option, 90 a call option on a put option, 91 a put option on a call option, 91 a put option on a put option, 91 adjoint equation, 128 admissible set, 317 American better-of option on two assets, 223 American call-max options on two risky assets, 233 American capped call options, 193 American multi-asset options, 222 American options with warning time, 194 American-style Asian options with floating strike price, 282 arbitrage opportunity, 10 arbitrage-free, 10 arbitrage-free principle, 9 arbitrageur, 8 arithmetic average Asian options, 275 arithmetic average Asian options
call, 248 call option, 3 call—put symmetry, 48 call-put parity, 14 call-put parity for Asian options, 288 call-put parity for the barrier options, 253 cash-or-nothing call, 88 central limit theorem, 57 chooser options (as you like it), 92 coincident set, 120 compound options, 90 consistent, 97 327
328
Mathematical Modeling and Methods of Option Pricing
contingent claim, 4 continuation region, 47, 113 control variable, 317 Covariance, 70, 202 cumulative probability distribution function of a standard normal distribution, 81 decomposition, 127 delivery price, 2 discounted price, 29 dividend, 39 down-and-in options, 247 down-and-out options, 247 early exercise premium, 130 European down-and-out call option, 250 European options, 4 exercise, 3 exercise price, 3 expectation, 29 expected return rate, 74 expiration date, 3 extrapolation, 315 fair, 37 financial derivatives, 1 finite difference method, 93 Finite difference scheme with methods of characteristic line, 305 finite difference schemes with characteristic line, 301 forward contracts, 2 free boundary, 117 free boundary problem, 113, 117 fundamental solution, 127 fundamental theorem of asset pricing, 38 futures, 2 geometric average Asian options, 275 geometric average Asian options with fixed strike price, 276 geometric Brownian motion, 61
hedger, 6 hedging, 2, 6, 26 ill-posed, 315 image method, 251 implied volatility, 311 in the money, 110 interpolation, 315 investment strategy, 9 killing rate, 264 knock-in options, 247 knock-out options, 247 leverage, 7 long position, 2 lookback call(put) option with a floating strike price, 292 lookback options, 292 martingale, 37 martingale measure, 37 maturity, 2 max(min) put options, 213 maximum call options, 212 maximum principle, 134, 136 minimizer, 317 minimum call options, 212 modified barrier options, 263 moving barrier options, 255 multi-asset options, 201 non-anticipatory, 64 obstacle, 119 obstacle problem, 119 one-period, 26 one-period and two-state model, 26 optimal control, 317 optimal control problem, 317 optimal exercise boundary, 47, 113 options, 2 options to exchange one asset to another, 212 out-performance options, 211 over-the-counter, 3
329
Index Parasian options, 264 Parisian options, 264 partial barrier options, 256 path, 56 path-dependent options, 247 payoff, 2 payoff function, 248 penalty function, 134 penalty problem, 134 perpetual American option, 113 portfolio, 9 positions, 2 premium, 4 probability, 10 probability measure, 28 put, 248
short position, 2 speculation, 1, 7 speculator, 7 splitting method, 268 spread options, 212 stability, 96 standard normal distribution, 57 standard Brownian motion, 74 standard lookback option, 292 step options, 264 stochastic calculus, 55 stochastic differential equation, 68 stopping region, 47, 113 strike price, 3 strongly path-dependent options, 247 symmetry, 48
Quadratic form, 204 quadratic variation, 61 quanto options, 210
tax, 74 time-dependent barrier options, 255 transaction cost, 74 trigger, 247 two-state, 26
rainbow options, 210 random walk, 56 recover, 311 relative price, 29 replication, 30 reset options, 260 reset options with predetermined dates, 260 reset options with predetermined levels, 260 return, 73 risk, 1 risk-neutral world, 74 rounding error, 96 self-financing, 10 separated set, 120
underlying assets, 2 up-and-in options, 247 up-and-out options, 247 vanilla options, 248 vanilla call option, 88 variational inequality, 117 volatility, 68, 74, 216, 311 volatility skew, 312 volatility smile, 312 weakly path-dependent options, 247 wealth, 10 worse-of options, 211