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!
—/i 2 with /i2 > 0), we enter a different phase of the system. Show that the ground state is now (p = const* = v, A^ = 0. What is the photon mass in this phase? Calculate the potential between two static point charges each of value Q. What sets the scale of the screening length? d) Let us now add an external field to the system, 1 M) ~~* M) T ^rfii/rext
'
To see that F£x\ indeed acts like an applied field, show that if one disregards the field
5Criticai whereas for B < JBCriticai it is the phase with
(p'(x') =
C/t) which forbids the transition of an even number of mesons to an odd number. However this is not a symmetry of QCD. More importantly, there are a set of low-energy relations, the Wess-Zumino consistency conditions [WeZ 71], which must be satisfied in the presence of the anomaly and which involve hadronic reactions. The effect of the anomaly was first analyzed by Wess and Zumino who noted that the result could not be expressed as a single local effective lagrangian, and gave a Taylor expansion representation for it.* Witten [Wi 83a] subsequently gave an elegant representation of the Wess-Zumino contribution as an integral over a five-dimensional space whose boundary is physical four-dimensional spacetime. Since the considerations leading to the Wess-Zumino-Witten action can be rather formal, it is best to adopt a direct calculational approach. Fortunately, we are able to employ the familiar sigma model (with fermions) because it contains the same anomaly structure as QCD. That is, it is the presence of fermions having the same quantum numbers as quarks which ensures that the anomaly will occur. The absence of gluons in * Note that if Eq. (2.13) is used to express Eq. (2.14) solely in terms of «7, the result is explicitly scale-independent because otj does not depend on /z. * For a textbook treatment, see [Ge 84]. = = ~ [x(t)] over trajectories one has instead a sum J [d(p(x)] over all possible field configurations. Nevertheless the analogy is rather direct. d^{x)] f[ W , (3.9)
VII-3 The Wess-Zumino-Witten anomaly action
197
the sigma model is not a problem since, according to the Adler-Bardeen theorem[AdB 69], the inclusion of gluons would not modify the result. Since the sigma model involves coupling between mesons and fermions, we can also observe directly the influence of the anomaly on the Goldstone bosons. Although somewhat technically difficult, our approach will clearly illustrate the connection with treatments of the anomaly based on perturbative calculations. Consider as a starting point the lagrangian, Eq. (IV-1.11), of the linear sigma model
C = &pip - gv (i>LUipR + ^ C / V L ) + ... .
(3.1)
We have displayed neither the term containing Tr (d^Ud^U^) and nor any term containing the scalar field S. Such contributions are not essential to our study of the anomaly and will be dropped hereafter. In order to simulate the light quarks of QCD, we shall endow each fermion with a color quantum number (letting the number Nc of colors be arbitrary) and assume there are 3 fermion flavors, each of constituent mass M — gv. Although the original linear sigma model has a flavor SU(2) chiral symmetry, Eq. (3.1) is equally well defined for flavor SU(3). Our analysis begins by imposing on Eq. (3.1) the change of variable
$'L = £tyL , ^ = WR, tf = U ,
(3.2)
like that appearing already in Eq. (IV-7.3). This yields C = $'(i lp -
(?d»z - td^) .
(3 3)
'
For this change of variable the jacobian is not unity, and thus we must write the effective action as eiT(U)
=
f[
= f[d^][df]J
elfd4x P'W-MW
(3.4)
= exp(lnj r ) exp(tr \n(iI/> - M)) . For large M, it can be shown that the tr ln(z Ip — M) factor does not produce any terms at order EA that contain the e^aP dependence char-
198
VII Introducing kaons and etas
acteristic of the anomaly.* Hence the effect of the anomaly must lie in the jacobian J, and it is this we must calculate. It is possible to determine the jacobian by integrating a sequence of infinitesimal transformations. Thus we introduce the extension £ —• £ r , gr = expi T
^ = expiry
,
(3.5)
where r is a continuous parameter and £ = £r=i- Transformations induced by the infinitesimal parameter 6r will give rise to the infinitesimal quantities £$r and 6J, (3-6)
Prom Eqs. (III-3.44), (III-3.47), we find 6J to be 8J = exp(—2i<5rtr ^75) ,
(3-7)
or dlnj dT
(3.8)
r=0
This result should be familiar from our discussion of the axial anomaly in Sect. Ill—3. There remain two steps, first to calculate the regularized representation of tr (^75), and then to integrate with respect to r. To regularize the trace, we employ the limiting procedure tr (^75) = lim tr (^75 exp [-e fT fT\)
{V» = d"+iV^T+iA^)
, (3.9)
with AT and V1^ as in Eq. (3.3), except now constructed from £T and £}. For arbitrary r, we make use of the identities
vT =
{
„i m j
to express Tf)r Tf>T in the form
Tprfr = d^
+a ,
dli = dlt + iVTll + a^Xis = d^ + FTfi , a = -2ATIX + i[(dM + iVTtl) ,X] 75 •
(3.11)
This can be verified by expanding as tr ln(i# - M) = tr ln(-M(l - Up/M)) = tr ln(-M) - tr (z#) 2 /2M 2 + ... . The first term can be regularized as in the text and directly calculated using the techniques described in App. B. The remaining terms vanish for large M.
VII-3 The Wess-Zumino-Witten
anomaly action
199
Prom the heat kernel expansion of App. B, we have*
Carrying out the 'TV' operation, which involves some application of Dirac algebra, yields
= 2iNcTr (^e^lpXXXX)
+•• • ,
(3.13)
where the ellipses denote contributions not involving e^va^ and the factor Nc comes from the sum over each fermion color. Combining the above ingredients, we have for the regulated action
where we recall that Tp = A • (p/(2Fn). This result expresses the effect of the anomaly on the Goldstone bosons. Unfortunately there is no simple way to integrate the entire expression of Eq. (3.14) in closed form. In principle, we could represent each of the axial-vector currents therein (e.g. A^) as a Taylor series expanded about r = 0 and perform the integrations to obtain a series of local lagrangians. Alternatively, however, one can simply express Eq. (3.14) as an integral over a five-dimensional space provided we identify r with a fifth coordinate x§ (defined to be timelike). In this case, we use i *
•
(3-15)
plus the cyclic property of the trace to write )
,
(3.16)
where i , . . . , m = 5,0,1,2,3 with e50123 = +1. This is Witten's form for the Wess-Zumino anomaly function. The r = 1 boundary is our physical spacetime, and the fifth coordinate is just an integration variable. Since each term in the Taylor expansion can be integrated, the result depends * Note the distinction between 'tr' and 'Tr', as in Eq. (Ill—3.48).
200
VII Introducing kaons and etas
only on the remaining four spacetime variables. Observe that vanishes for U in SU(2) due to the properties of Pauli matrices. For chiral 5/7(3), the process K+K~ —• 7r+7r~~7r° is the simplest one described by this action and after expanding Fwzw? it is described by the lagrangian, ^
,
(3.17)
with if = A • cp given in Eq. (1.3). The above discussion has concerned the impact of the anomaly on the Goldstone modes. We must also determine its proper form in the presence of photons or W± fields. For this purpose, we can obtain the maximal information by generalizing the fermion couplings to include arbitrary left-handed or right-handed currents t^r^ 1 ^
(3.18)
-r The calculation of the jacobian then involves the operator V
-d
+U
1 + 75
I ir
1 - 7 5
which generalizes Eq. (3.3). It is somewhat painful to work out the full answer directly, but fortunately we may invoke Bardeen's result of Eq. (Ill— 3.64) for the general anomaly. Using the identities
where ^v,r^v
are given in Eq. (Ill—3.65), we obtain
Fwzw = -^J
drjd4x
+ Va V) 0) ++ (l +^M M^ + 2i,
— —{a^ajyVa^
f^lap)
(3.21)
+ a^Vvadp + v
Note that the first term corresponds to our previous calculation of Eq. (3.14). The WZW anomaly action contains the full influence of the anomalous low energy couplings of mesons to themselves and to gauge fields. By construction, it is gauge invariant. The r integration can be explicitly performed for all terms but the first in Eq. (3.21). However, in the general nonabelian case the result is extremely lengthy [PaR 85]. For
VII-4 The rf (960)
201
the simpler but still interesting example of coupling to a photon field A^, the result is Twzw (U, A^) = Twzw(f^)
+
a/3 d4a; iv(Q(
^^ / K
^
— ie FuVAa Tr I Q (LR + R3) + — \ 2
' (3.22) where R^ = (d^U^)U, L^ = Ud^U^.* We have already used the twophoton portion of this result in Sect. VI-5 for the decay TT° —• 77. We have seen here that whereas the anomalous divergence of the axial current represents the response to an infinitesimal anomaly transformation, the WZW lagrangian represents the integration of a series of infinitesimal transformations. In our analysis of the sigma model, the anomaly has forced the occurrence of certain couplings, among them Tr0 —> 77, 7 —> 3TT and KK —• 3TT. AS noted earlier, although these results are based on an instructional model, the result has the same anomaly structure as QCD because the answer must depend on symmetry properties alone. Indeed, such conclusions were originally deduced from anomalous Ward identities [WeZ 71] without any reference to an underlying model. We regard such predictions as among the most profound consequences of the Standard Model. VII-4 The
T/(960)
Thus far, we have been dealing with hadrons whose presence could be inferred from the symmetry properties of QCD. We have not needed to invoke the quark model to describe the meson spectrum. This procedure is less clear for the rf (960), which is not a Goldstone boson for any of the symmetries of QCD. The least complicated description of the rf is with the quark model. Here one notes that the eight particles (TT"1", TT~, TT°, K+, K~, K°, K° and rjs) havethe same quantum numbers as the respective quark-antiquark pairs (ud, du, (uu—dd), us, su, ds, sJ and (uu+dd—2ss)). However, such QQ pairing suggests a ninth state, the SU(3) singlet 770 ~ (uu+dd+ss). A look through the Particle Data Tables reveals that the lightest candidate for a ninth pseudoscalar is 7/(960). In fact, the 7/(549) and 7/(960) turn out to involve mixtures of 7/8 and 770 (and in principle of other states as well), with the 7/(960) being treated as predominantly the SU(3) singlet. * Witten's original result did not conserve parity, and this was subsequently corrected [PaR85, KaRS84].
202
VII Introducing kaons and etas
One immediate puzzle involving the rf would appear to be its mass, which is almost twice that of the kaon. Of course, the TT, K, r)$ masses must vanish in the limit of zero quark mass, while because of the axial U(l) anomaly there is no such requirement for the TJQ. Moreover, there is the presence of the quark annihilation diagram, Fig. VII-2. This can keep the 770 mass finite in the chiral limit. The exact value of this mass contribution is difficult to calculate within the quark model. At any rate, the proper expectation for the lightest pseudoscalar masses is ml = 2mBo , 2 m 5 m 3 ((22m* + + ™) ) 5 o ,
m2K = (ms + m)Bo , 2 m200 + << = m + -(ms + 2m)B' ,
(4 1)
where rao is the 770 mass in the chiral limit and B1 is in general a constant different from Bo. Here we have temporarily ignored 770-778 mixing. The discussion becomes more interesting when the connection to the axial anomaly is considered. The SU(3) singlet axial current
= ^F^F^
+ 2imuybu + 2imdd7bd + 2ims^s
where Fa" = e ^ ^ i 7 ^ , has the 770-to-vacuum matrix element * ,
(4.3)
in which F^ is the 770 decay constant. By taking the divergence of this equation we obtain
| ^
J, .
(4.4)
i=u,d,s
If the anomaly were not present, there would exist an almost conserved current for the £^4(1) symmetry. If this were dynamically broken along with the SU(3) chiral symmetry, the (pseudo-) Goldstone boson would be the 7/o. The above relation, without the FF term, would then express the result that miQ = O(mq). However, the anomaly contribution (O\FF\r)o) does not in general vanish, so that m^Q remains nonzero in the chiral limit of zero quark mass.
u,d,s
u,d,s u,d,s •
u,d,s G
Fig. VII-2 Annihilation diagram
VII-4 The rjf (960)
203
In fact, it can be argued that the largest contribution to the 770 mass is due to the axial anomaly. The argument is rather subtle, and the 770 mass was long considered to be a problem despite the anomaly. The difficulty arises because there does exist a conserved J7(l) current in the chiral limit. FF is expressible as a total divergence, FF = d^K^1 with K*1 given in Eq. (Ill—5.7). The conserved current is formed by subtracting K^ from the usual current T(0) _
T(0)
K
B
T(0)/Z _
n
(
,
-x
where quark masses are being neglected here. The associated charge will generate UA(X) transformations, and, since UA(1) is not an observable symmetry of the spectrum, it must be dynamically broken. Such reasoning leads one to expect a Goldstone boson. In the matrix element analysis of Eqs. (4.3), (4.4), the same difficulty arises because the identification of FF as a total divergence would seem to imply, by integrating by parts and discarding the surface term at infinity, that fd4xFF
= 0 or
f d4x (0\FF |r?o(p)) = 0 .
(4.6)
The latter matrix element would imply the possibility of a zero momentum state (if the quark masses vanished), apparently producing a Goldstone boson despite the anomaly. The resolution of this apparent paradox is twofold. In the first place, the conserved current JL is gauge dependent due to K^. While this does not invalidate either its use as a U(l) generator or the need for a Goldstone boson, it does allow the possibility that the Goldstone boson need not appear in the physical gauge-invariant spectrum. It can appear rather as a gauge-dependent effect [KoS 75] and need not couple to gauge-invariant operators. (However, neither is it known to be forbidden to do so.) In the matrix element argument, the resolution occurs because there exist topological effects in QCD such that the surface term at the spacetime boundary can prevent Eq. (4.6) from vanishing ['tH 76b] (see the discussion in Sect. Ill—5). This allows the simple analysis of the 7/ mass, which was given in the preceding paragraphs, to be valid. While the proof that precisely these phenomena do occur in QCD is not without controversy, it would appear that this is the route by which Nature has chosen to avoid a ninth Goldstone boson. As a result, the couplings and decays of the rjf are not related by symmetry conditions to those of the octet of pseudoscalars. That is, in QCD there is no 'nonet symmetry' in the standard context of what is meant by symmetry.
204
VII Introducing kaons and etas 770 - rjs mixing
If the u, d, s quarks were massless, the picture developed thus far would imply an octet of massless pseudoscalars plus the massive SU(3) singlet 7/0- When quark mass is introduced in perturbation theory, each squaredmass of the pseudoscalar octet becomes linear in quark masses, and the mass of the T/O is shifted slightly. In addition, SU(S) breaking in the quark mass matrix induces a mixing between the singlet and octet, i.e. between rjo and rjg. This mixing yields the physical eigenstates rj and rf \rj) = cos0\r]g) - sin0 |r/o) ,
|T/) = sin0 |T/8) + cos(9 |r/o) .
(4.7)
The goal is then to calculate 6 theoretically and measure it experimentally. There are two crucial assumptions which go into the mixing scheme described here. One is that it is possible to neglect the other pseudoscalar 7 = 0 states which could mix with both r/g and 770. A check with [RPP 90] shows that the next possible candidate is 77(1279). This is distant from the 77s mass, and hence might not mix significantly with it, but is not extremely far from the 770 mass. The second, somewhat related assumption is that the mixing parameters do not depend significantly on the energy of the state. If they did, a more complicated scheme for finding the mass eigenstates would be required. These assumptions have not yet been justified in QCD. The utility of this scheme seems to be founded on its success in phenomenological applications [GiK 87]. The 77-7/ mixing angle may be determined from experiment. In our opinion, the cleanest phenomenological analysis occurs in the two photon decays of TT0, 77, 7/. All of these decay amplitudes have the same Lorentz structure,
7(k2)
= -§
(4.8)
where the convenient normalization CM — (l> l/>/3, 2>/2/3 ) has been chosen for the unmixed states M = (TT°, 7/8, 7/0). The parameters FM are not a priori known. However, a prediction of the axial anomaly analysis is that Fn = F^, where Fn is the usual pion decay constant. Similarly, the Wess-Zumino-Witten lagrangian of Sect. VII-3 predicts Fm = Fn. We shall explore the sensitivity of this parameter to SU(3) breaking. Finally Fm is not predicted by any symmetry. Quark model prejudice, described below, also suggests Fm ~ Fn. Allowing for 77 — 7/ mixing as in Eq. (4.7),
205
VII-4 The 77'(960)
the decay rates become
1 m* fi^-costf 3^4 Fm
vo
a svn.9
(4.9)
Fvcos6
+
3 ml
The data is very sensitive to 77-77' mixing, largely because of the factor y/8 in the 77 decay amplitude, and in fact requires its presence. In particular, the 'reduced' rates are -3.0 ±0.3 ,
3 ml I V /V = 8
i^sinfl n0
|
F^cosO
(4.10) = 0.61 ±0.05 .
Using the Wess-Zumino-Witten value for FVs, Eq. (4.10) implies 6 = - 1 7 ° ±3° ,
F o = (1.08 ±0.04)2^ .
The dependence on SU(3) breaking in Fm is not large. If we use instead the value JF^ = Fm, where Fm is the usual axial current decay constant calculated in Eq. (2.8), one obtains the compatible values 6 = -22° ±3° ,
~F0 = (1.05 ± 0 . 0 4 ) ^ .
(4.11)
Two-photon decays then require a mixing angle which is around —20°. What is the effect of this mixing? In the quark model, the 778 and 770 states are octet and singlet qq states respectively. The physical mixtures would then be 77
/ _
-
_x
sin^/ _
-
_x
—(uu + dd + ss) v 6 (uu + da — 2ss) ~ 0.58(uu + dd) - 0.5755 , 3 j^[uu + da — 2ss) 2ss) HH ^ v6 v3 ~ 0.40(im + dd) + 0.8255 .
77 =
(4.12)
+ dd + ss)
The mixing is seen to decrease the 55 content of the 77 and increase it for the 77'. It is the added uu content of the 77 which enhances the transition V ~^ 77 due to the larger u quark charge. It is interesting that the parameter FVo is found to be so nearly equal to Fn. In the quark model, if one assumes that the spatial wavefunctions of the quarks in the TT, 778 and 770 are identical, the two-photon amplitudes would be proportional
206
VII Introducing kaons and etas
to the squared quark charge, i.e., + Q2d + Q2s\M°) .
(4.13)
This would produce relative amplitudes TT°: 778 : 770 = 1 : l / \ / 3 : 2^/2/3, which corresponds in the normalization of Eq. (4.8) to Fn = Fm — Fm. This need not work as well as it seems to. The theoretical aspect of mixing is addressed by considering the mass matrix, with off-diagonal element mls = (r/o \Hmass\ Vs) •
(4.14)
This can be estimated by considering the model where we use the quark model description of 770 and 773, with identical wavefunctions. In that case we would calculate 5
Bo =
—rm\
~ -0.9m2K
.
(4.15)
Although this is not a definitive analysis, it does serve to illustrate how mass mixing can be generated. In an obvious notation, the overall 2 x 2 mass matrix is then m 08 m 0 The mass rn^ is completely unknown, but m\ should be close to the GellMann-Okubo value (4mKml)(l
+ 6) ,
(4.17)
where 6 parameterizes those deviations from the Gell-Mann-Okubo relations which can occur even in the absence of 77-7/ mixing. The dependence on 6 is rather strong. If 6 = 0, diagonalization of the matrix gives the values 0 = -10°, mg8 = -0.44ra^, ra0 = 0.95 GeV, while for 6 = 0.25, one finds 6 = -24°, mg8 = 0.93 m\, m0 = 0.90 GeV. These results lie in a reasonable range, consistent with theoretical expectations. Alternatively, one can work backwards from the observed 7]-r]f mixing to determine the parameters in the mass matrix. The choice 9 = —20° ± 2° produces 6 = 0.16 ± 0.04, and mg8 = -(0.81 ± 0.05)ra^. Again, the quark model prediction Eq. (4.15) is found to work surprisingly well.
Problems 1) Other worlds Describe changes in the macroscopic world if the quark masses were slightly different in the following ways:
Problems
207
a) mu = rrtd = 0, b) mu > ra^, c) mu = 0, rrid — m8. 2) r] decay a) Prove that the amplitude for rj —> 3TT vanishes in the limit of isospin symmetry. b) Using the lowest order chiral lagrangian, calculate the rj —• 3TT decay amplitude and slope, A A
K A
1
i
I
u
yv477_).7r+7r-7ro = Alo 1 + ^ I in terms of the quark mass difference rrid ~ mu, where To is the TT°kinetic energy and Q = m^ — Zm^. Compare your results with experiment. This decay remains in disagreement even when a full O(E4) calculation is performed [GaL 85c].
3) T/-?/ mixing and chiral lagrangians
Although the 5/7(3) singlet 770 does not occur directly in the SU(3) chiral lagrangian, its effect can appear indirectly through virtual processes. a) Show that the only lowest order chiral coupling of an 770 to the chiral octet is -iV6Fnm20S
_
° X)
with a normalization constant chosen to reproduce the 77 — 7/ mixing of Eq. (5.10). b) Integrate out the 7/0 field and show that one thereby obtains a term in the chiral lagrangian with 7
64m20(r4_m2)2 •
Evaluate this numerically. Thus the effect of rj — 7/ mixing is represented in the chiral lagrangian by the value of a.7.
VIII Kaons and the AS = 1 interaction
The kaon is the lightest hadron having a nonzero strangeness quantum number. Since the weak interactions do not conserve strangeness, the kaon is unstable and decays weakly into states with zero strangeness, containing pions, photons and/or leptons. There are in fact dozens of allowed modes, and many of these have been carefully studied. In lieu of detailing all of these possibilities, we shall instead concentrate on some of the primary decay channels. VIII-1 Leptonic and semileptonic processes Leptonic decay The simplest weak decay of the charged kaon, denoted by the symbol K^-, is into purely leptonic channels K+ —» /x + ^, K+ —• e+z/e. Such decays are characterized by the constant FK, As discussed previously, because of SU(3) breaking FK is about twenty percent larger than the corresponding pion decay constant Fn. As with the pion, but even more so because of the larger kaon mass, helicity arguments require strong suppression of the electron mode relative to that of the muon. The ratio of e+z/e to / i + ^ decay rates, as in pion decay, provides a test of lepton universality [RPP 90],
r*K . e ^
v
*
= (2.44 ±0.11) x 10"5 ,
(1.2)
K+-+H+V14 expt
in good agreement with the suppression predicted theoretically, m
ml
1 - ml/ml •til •"•K 208
(1 + 8) = 2.41 x 1(T5 ,
(1.3)
VIII-1 Leptonic and semileptonic processes
209
where 6 = —0.04 is the electromagnetic radiative correction including the bremsstrahlung component.* The notation F' indicates that the experimenters have subtracted off the large structure dependent components of > £+i/£^y but have included the small bremsstrahlung component.
Kaon beta decay and V^ The kaon beta decay reactions K+ —> n o£+iy£ and K° —• TT~^ + I^, called g} and K^ respectively, also are important in Standard Model physics. They are each parameterized by two form factors,
[ (1-4) Isospin invariance implies f± = f± * = f±. SU(3) symmetry can be invoked to relate these matrix elements to the strangeness conserving transition TT+ -> iPl+vi, resulting in /+(0) = - 1 and /_(0) = 0. The deviation of /+(0) from unity is predicted to be second order in SU(3) symmetry breaking, i.e. of order (ms — rh)2. This result, the AdemolloGatto [AdG 64] theorem, is proved by considering the commutation of quark vector charges, n
[Q^iQ™]
=QUu~~ss
,
(1.5)
where
j
(
1
.
6
)
Taking matrix elements and inserting a complete set of intermediate states gives
(
2
r
)
•
d-7)
Finally, we isolate the single n~ state from the sum and note that in the SU(3) limit the charge operator can only connect the kaon to another state within the same SU(3) multiplet. This implies^ (1.8) * The dominant term here is the simple contact contribution —3(a/7r)ln(m Sect. VI-L t This is easiest to obtain in the limit p ^ —• oo
At/me)
discussed in
210
VIII Kaons and the AS — 1 interaction
where e is a measure of SU(3) breaking, and we thus conclude that ,
(1.9)
which is the result we were seeking. It is interesting that the SU(2) mass difference mu ^ rrid can modify f+ n (0) in first order despite the Ademollo-Gatto theorem. This can be seen by considering a K+ in the formulae of Eq. (VII-1.1). Now there exist two intermediate states in the same octet as the kaon, i.e. TT°and 7?°,and it is their sum which obeys the Ademollo-Gatto theorem, 1 (0) =1 + O(e2) . (1.10) 4 In the isospin limit, each term must separately obey the theorem because n ~. However, when mu ^ rrid each of the isospin relation f+*n° = f+° form factor in Eq. (VII-1.10) can separately deviate from unity to first order in mu — rrid as long as the first order effect cancels in the sum. Indeed this is what happens, yielding (cf. Prob. VIII. 1)
/f-(0) where IK-K — 0.004 arises from chiral corrections at O(E 4) [GaL 856]. This number can also be easily extracted from experiment by using the ratio of K+ and K° beta decay rates, with the result -^ = 1.029 ± 0.010 , J+
"
(1.12)
\y)
in agreement with the prediction. One of the first applications of current algebra/PC AC techniques was made to K& systems, yielding [CaT 66] <7r°(p) |57Mii| tf+(k)) _ , - - L (o I [Ql
(1.13)
The corresponding condition on the form factors, [f+(q2) + f-(q%2=m2K
=- ^
,
(1.14)
K
is called the Callan-Treiman relation. Interestingly, this does not imply that /-(0) = 1 — FK/FK, as was originally assumed. We can demonstrate this by using the chiral lagrangian of Eq. (VI-2.7) at tree level (loop effects
VIII-1 Leptonic and semileptonic processes
211
are small and will be added later). Expansion of the vector current matrix element yields
(1.15)
This form explicitly satisfies the Callan-Treiman relation off-shell, yet at q2 = 0 yields a positive value for /-(0),
= f+(m2K,ml,0)=-l
,
mj) - 0.13 ,
^ K
§F
(L16)
where we have used the previously determined value Og = (7.7±0.2) x 10~3 (which includes the effects of loops). This prediction is in agreement with the results of K^ experiments
f (0)
f
a35 ± 15 = uw 1 "-0.20 °' ± 0.08 =
{K ]
^' (combined) ,
(L17)
which clearly require the presence of ag contributions as contained in the chiral lagrangian formalism. The q2 variation of the form factors ,2\ _
/o( 9 2 ) =
£ /J2\
,
9 171
K
/
/.2\ ..
-.
A
2
(1*18)
~
follows from the same input parameter as was used to predict rare pionic processes, with the addition of FR- The agreement with experiment as displayed in Table VI-2 is good. The prime importance of the K& process is that it provides the best determination of the weak mixing element V^. Because of the AdemolloGatto theorem, the reaction is protected from large symmetry breaking corrections. In addition, the use of chiral perturbation theory allows a reliable treatment of the reaction. The above study of the form factors indicates that the theory is under control within the limits of experimental precision. The value [LeR 84] Vus = 0.2196 ± 0.0023 +
follows from an analysis of the K° and K decay rates.
(1.19)
212
VIII Kaons and the AS = 1 interaction The decay K —>
7nreue
The final semileptonic process which we consider is K —> TTTT^, labeled X^4. This reaction has the particular interest in providing the only known constraint on the large Nc assumptions sometimes made concerning chiral lagrangians, and is important in determining the low energy constants. As will be explained more fully in Chap. X, the large Nc limit imposes certain relationships between some of the terms in the chiral lagrangian. The particular relationship that is able to be tested in the K& system is the prediction OL
x [(p+ + p_)M h + (P+ - P-V h + (k - P+ - P-)M /a] , <7r + (p + )7r(p_) | l 7 ^ l ^ + ( k ) > = g ^ W ^ ^ + Z
(1.20)
9 •
The form factor f% is essentially unobservable because its effect is proportional to me in ife4, and we shall not consider it further. We have chosen the normalization such that the lowest order chiral predictions are f1 = f2 = g = l [We 66a]. The vector current form factor g = 1 is a prediction of the axial anomaly, as it is related by an SU(3) rotation to 7 —> 3?r. Experimentally, the magnitudes and slopes of the form factors at threshold
+ Ar^]
with fc2 = i ( ( p + + p _ ) 2 - 4 n 4 )
Table vra-i. Experiment and theory for Ku decays. Data /i(0)
/ 2 (0) Ai
A2 9(0) a
1.47 ± 0.04 1.25 ± 0.07 0.08 ± 0.02 0.08 ± 0.02 0.96 ± 0.24
Lowest order0 Order (E4)a 1.00 1.00 0.00 0.00 0.00
1.45 1.24 0.08 0.06 1.00
The lowest order and second order predictions of chiral symmetry.
(1.21)
VIII-2 The nonleptonic weak interaction
213
have been measured with the results [Ro et al 77] given in Table VIII-1. No dependence on the variable q2 = (k — p+ — p-)2 was observed. The enhancement in the threshold form factors and slopes is determined at O(E4:) primarily by 0:1,2,3, which also enters into TTTT scattering. A very consistent pattern emerges. The chiral KM predictions become [Bi 90, RiGDH 91] /i(0) = 1 + Xi + ^
+ {m2K
[32mjal + 4(m2K + 4ml
l
l
\
/2(0) = 1 + X2 - A [(m2. *IT
Ai = n + 8 ^ | [16aJ + 4a£ + 5ar3] ,
X2 = Y2- 8^ar3
,
where the quantities X\ = 0.127, X2 = 0.00, Fi = 0.076, Y2 = 0.045 describe the net effects of loops when evaluated at \x = m^. Since the dependence on a^ is small because of the factor of m^, we shall proceed with c*4 = 0. Phenomenologically, one may either fit /i(0),/2(0), Ai to determine ai?2,3 and then predict A2 and the seven TTTT scattering lengths and slopes, or provide the best determination of ai?2,3 by simultaneously considering all eleven of the TTTT and K^ observables. The results are very similar either way, and therefore we present only the second case. The parameters are those shown in Table VI-1, and the predictions for the observables are contained in Table VI-4 for TTTT physics and in Table VIII-1 for Ku decay. The overall description of the two systems is excellent. We note also that the analysis results in the constraint
which verifies the suppression predicted by the large Nc limit. VIII-2 The nonleptonic weak interaction Thus far in this chapter we have discussed leptonic and semileptonic processes. For these, at most one hadronic current is involved. There exist also nonleptonic interactions, in which two hadronic charged weak currents are coupled by the exchange of W± gauge bosons,
= f f
214
VIII Kaons and the AS = 1 interaction
with V being the KM matrix, given in Eq. (II-4.26). Such interactions are difficult to analyze theoretically because the product of two hadronic currents is a complicated operator. If one imagines inserting a complete set of intermediate states between the currents, all states from zero energy to Mw are important, and the product is singular at short distances. Thus one needs to have theoretical control over the physics of low, intermediate, and high energy scales in order to make reliable predictions. Because this is not the case at present, our predictive power is substantially limited. Let us first consider the particular case of AS = 1 nonleptonic decays. These are governed by the products of currents dT^u uT^s ,
dT^c cT^s ,
dTH iT^s ,
(2.2)
where F^ = 7^(1 + 75) and color labels are suppressed. The first of these would naively be expected to be the most important, because kaons and pions predominantly contain u, d, s quarks. However, the others also contribute through virtual effects. Some properties of the AS = 1 nonleptonic interactions can be read off from these currents. The first product contains two flavor-ST/(3) octet currents, one carrying I = 1/2 and one carrying / = 1, 27 , ( 8 0 8)symm 1 1 ^ 3 1(5 = *2" 2 "2 '
SU(3) : Isospin :
(2.3a) (2.3b)
where the symmetric product is taken because the two currents are members of the same octet. The singlet SU(3) representation is excluded from Eq. (2.3a) because a AS = 1 interaction changes the SU(3) quantum numbers and hence cannot be an 5(7(3) singlet. The other two products in Eq. (2.2) are purely SU(3) octet and isospin 1/2 operators. The currents are also purely left-handed. Thus the nonleptonic hamiltonian transforms under separate left-handed and right-handed chiral rotations as (8L,1R) and (27 L,1R)- These symmetry properties, valid regardless of the dynamical difficulties occurring in nonleptonic decay, allow one to write down effective chiral lagrangians for the nonleptonic kaon decays. The hamiltonian is a Lorentz scalar, charge neutral, AS = 1 operator, and has the above specified chiral properties. At order E2, there exist two possible effective lagrangians for the octet part, viz. £Octet = C$ + C%, where in the notation of Sect. IV-6, £ 8 = g8 Tr (^DJJD»U^
,
£ 8 = 98 Tr (A 6X t/t) + h.c. .
(2.4)
It can easily be checked that both £% and C% are singlets under righthanded transformations, but transform as members of an octet for the left-handed transformations. The barred lagrangian in Eq. (2.4) can in
VIII-2 The nonleptonic weak interaction
215
fact be removed, so that it does not contribute to physical processes. This is seen in two ways. At the simplest level, direct calculation of K —• 2TT and K —> 3?r amplitudes using Cs, including all diagrams, yields a vanishing contribution. Alternatively, this can be understood by noting that in QCD the quantity x appearing in Eq. (2.4) is proportional to the quark mass matrix, x — 2B$mq. Thus the effect of £g is equivalent to a modification of the mass matrix, mq —• mq = mq + g%\§mq
.
(2-5)
This new mass matrix can be diagonalized by a chiral rotation Tr (m'qU) -+ Tr (Rm'qLU) = Tr (mDU)
,
(2.6)
with mo diagonal. The transformed theory clearly has conserved quantum numbers, as it is flavor diagonal. This means that the original theory also has conserved quantum numbers, one of which can be called strangeness. When particles are mass eigenstates, even in the presence of £8? the kaon state does not decay. Hence this Cs can be discarded from considerations, leaving only Cs as responsible for octet K decays [Cr 67]. This octet operator is necessarily A / = 1/2 in character. Another allowed operator, transforming as (27^,1^), contains both A / = 1/2 and A / = 3/2 portions,
where /»(l/2)
jL/97
27
: =
(1/2) x-^1/2 rri
5^97
= ^27
^
/> ^
| \ Ot O r r TT\ \bcs TT TT^ I
i "U
\ ^ OurJ U *• Ou*J C/ I "+" n.C. ,
^ a 6 ^^ l ^ ^ ^ C/' A a^C/ (7 ' ) + h.C. \ /
/O O \
iZ.odi)
(z.ODJ
The coefficients are given by C ly/2
=1
^3/2
_-
°fi-i-«7/9 3 —
x
J
C1/2
= — \/2
^3/2
_ J_
°4+t5/2,l-t2/2 ~~
/o
^1//2 *
The complete classification at order E4 is difficult, but has been obtained [KaMW 90]. We shall apply these lagrangians to the data in Sects. VIII4, XII-6. There we shall see that gs » g^, whereas naive expectations would have octet and 27-plet amplitudes being of comparable strength. This is part of the puzzle of the A / = 1/2 rule. The reliable theoretical calculation of the nonleptonic decay amplitudes, which is tantamount to (1/2)
(3/2)
predicting the quantities gs, #27 and g^7 , is one of the difficult problems mentioned earlier. It has not yet been convincingly accomplished.
216
VIII Kaons and the AS = 1 interaction
The best we can do is to describe the theoretical framework of the short distance expansion, to which we now turn. VIII-3 Short distance behavior At short distances, the asymptotic freedom property of QCD allows a perturbative treatment of the product of currents. The philosophy is to use perturbative QCD to treat the strong interactions for energies Mw > E > /i. The result is an effective lagrangian which depends on the scale fi. Ultimately matrix elements must be taken which include the strong interaction below energy scale /i and the final result should be independent of //. Short distance operator basis The outcome of the short distance calculation can be expressed as an effective nonleptonic hamiltonian expanded in a set of local operators with scale dependent coefficients [Wi 69], .
(3.1)
As in any effective lagrangian, those operators of lowest dimension should be dominant. If the operator O{ has dimension d, its coefficient obeys the scaling property Q ~ M^d. Let us first see how this hamiltonian is generated in perturbation theory. We can later use the renormalization group to sum the leading logarithmic contributions. The lowest order diagrams renormalizing the current product are given in Fig. VIII-1. The process in Fig. VIII-1 (a) corresponds to a left-handed, gaugeinvariant operator of dimension 4, O (d=4) = dlp{\ + 75)5 .
(3.2)
This operator can be removed from consideration by a redefinition of the quark fields (cf. Prob. IV-1). The remaining operators are of dimension 6. Simple ^-exchange with no gluonic corrections gives rise in the short distance expansion to the local operator OA = d7fi (1 + 7 5 ) ^ 7 ^ (1 + 75) s ,
(3.3)
u,c,t
W W (a)
s
u (b)
u
Q
Q (O
Fig. VIII-1 QCD Radiative corrections to the AS = 1 nonleptonic hamiltonian
VIII-3 Short distance behavior
217
with a coefficient CA = 2 in the normalization of Eq. (3.1). The gluonic correction of Fig. VHI-l(b) generates an operator of the form J7 M (l+75)A a u«7 / '(l+75)A o s ,
(3.4)
a
where the {A } are color SU(3) matrices. However, use of the Fierz rearrangement property (see App. C) and the completeness property Eq. (II2.8) of SU(3) matrices allow this to be rewritten in color-singlet form h» (1 + 75) Aa« «
7
M
(1 + 75) Xas = -\oA
+ 20B ,
+ jb)ud^(l
.
where OB = u^(l
+ j5)s
(3.5)
The strong radiative correction is seen to generate a new operator OBPerturbative analysis Consider now the one-loop renormalizations of the four-fermion interaction Fig. VIII-l(b). In calculating Feynman diagrams we typically encounter integrals such as (neglecting quark masses) -
J
16TT2M^
K2
-r iv±w
(3.6)
where we evaluate the integral at the lower end using a scale /x. Clearly presents a natural cutoff in the sense that
The modification of the matrix element to first order in QCD is then ,
(3.8)
where gs is the quark-gluon coupling strength. The gluonic correction to OB must also be examined, and a similar analysis yields OB - OB - JL
In (^)(3OA-
O B) .
(3.9)
We observe that the operators, O± = ^(OA±OB)
,
are form-invariant, O± —• c±O±, with coefficients c±,
(3.10)
218
VIII Kaons and the AS = 1 interaction
where e/+ = —2 and GL = +4. The isospin content of the various operators can be determined in various ways. Perhaps the easiest method involves the use of raising and lowering operators [Ca 66, Li 78], /_i_d — u ,
I+u = —d ,
I-U = d ,
I-d = —u ,
(3.12)
to show that I+O- = 0, implying that O_ is the Iz = 1/2 member of an isospin doublet. With repeated use of raising and lowering operators, one can demonstrate that O_ is purely A / = 1/2 whereas O+ is a combination of A / = 1/2 and A / = 3/2 operators. Prom Eq. (3.11), we see that under one-loop corrections the operator O- is enhanced by the factor
fL^!2.1
(3.13)
where we use a8(fi) ~ 0.4 (AQCD — 0.2 GeV) at // ~ 1 GeV. Similarly O+ is accompanied by the suppression factor
Renormalization group analysis Choosing an even smaller value of // would lead to an even larger correction. However, maintaining just the lowest order perturbation in the QCD interaction would then be unjustified. It is possible to do better than the lowest order perturbative estimate by using the renormalization group to sum the logarithmic factors [GaL 74, AIM 74]. In a renormalizable theory physically measurable quantities can be written as functions of couplings which are renormalized at a renormalization scale /iR. Physical quantities calculated in the theory must be independent of /iR. Denoting ail arbitrary physical quantity by Q, this may be written Q = /(03(MR)>MR)
>
(3.15)
where / is some function of //R and 53 is the strong coupling constant of QCD. Differentiating with respect to /XR, we have O ,
(3.16)
which is the renormalization group equation. It represents the feature that a change in the renormalization scale must be compensated by a modification of the coupling constants, leaving physical quantities invariant. In order to see how this program can be carried out for the effective weak hamiltonian, consider the following irreducible vertex function
VIII-3 Short distance behavior
219
which represents a typical weak nonleptonic matrix element, (0 \
( ^
(y^f (0 where the {<&} are quark fields carrying momenta {pi}. Z2 is the quark wavefunction renormalization constant for the fermion field, and subscripts 'ren', 'unren' denote renormalized and unrenormalized quantities. Choosing the subtraction point yy\ = —/J*^ we require that unrenormalized quantities be independent of //R, r
unren
= 0.
(3.18)
This implies ) \ o)™ = 0,
0
where g$r is the renormalized strong coupling constant, /?QCD is the QCD beta function of Eq. (II-2.57) and JF is the quark field anomalous dimension of Eq. (II-2.69). As we have seen, QCD radiative corrections generally mix the local operators appearing in the short distance expansion, (0 \T (Onqiq2q3q4)\ 0)£n - £
Xnn, (0 \T (On
and the mixing matrix can be diagonalized to obtain a set of multiplicatively renormalized operators (0 \T (OkQlq2q3q4)\ 0>™ = Zk (0 \T {Om
•
(3-20)
If Zk has anomalous dimension 7^, Z f c ~ l + 7fclnMR + -.. ,
(3.21)
then the coefficient functions Ck(^Rx) must satisfy — + /SQCD^— + Ik ~ 47F I ck(fjLRx) = 0 . /iR dgsr ) Prom the above, we have for the operators O± &±
•
(3.22)
(3 23)
-
Having specified the anomalous dimensions of the operators O±, we can solve Eq. (3.22) with methods analogous to those employed in Sect. II—2. That is, because QCD is asymptotically free and we are working at large
220
VIII Kaons and the AS = 1 interaction
momentum scales, we can use the perturbative result (cf. Eq. (II-2.57)),
where 6 = 1 1 — | n / , rif being the number of quark flavors. Upon inserting the leading term in the perturbative expression for as (cf. Eq. (II-2.74)), 12TT M
(325)
one can verify that the solution to Eq. (3.22) is given by / C±(»R)/C ±(MW)
n2
M2
\d±/b
= ^1 + ^ f t l n - f J
.
(3.26)
Note that in the perturbative regime where 1 ^> as, we have C±(/*R)
.
93
M ]n W
c±(Mw) which agrees with our previous result, Eq. (3.11). It is the renormalization group which has allowed us to sum all the 'leading logs'. Of course, at scale Mw one must be able to reproduce the original weak hamiltonian, implying c+(Mw) = c-(Mw) = 1. Taking //R ~ 1 GeV and as = 0.4 as before, we find with c_(/iR) ~ 1.5 ,
c+(/iR) ^ 0.8 .
(3.29)
We observe then a A / = 1/2 enhancement of a factor of 2 or so, which is encouraging but still considerably smaller than the experimental value of A0/A2 ~ 22 discussed in the next section. Two additions to the above analysis must now be addressed. One is the proper treatment of heavy quark thresholds. In reducing the energy scale from Mw down to /iR, one passes through regions where there are
Fig. VIII-2
Penguin diagram.
VIII-3 Short distance behavior
221
successively six, five, four or three light quarks, the word 'light' meaning relative to the energy scale //R. The beta function changes slightly from region to region. A proper treatment must apply the renormalization group scheme in each sector separately. This is a straightforward generalization of the procedures described above. The other addition is the inclusion of penguin diagrams of Fig. VIIIl(c) [ShVZ 77, ShVZ 79b, BiW 84], whimsically named because of a rough resemblance to this antarctic creature. The gluonic penguin is noteworthy because it is purely A/ = ^, thus helping to build a larger A/ = | amplitude, and because it is the main source of CP violation in the AS — 1 hamiltonian. The electroweak penguin, wherein the gluon is replaced by a photon or a Z° boson, also enters the theory of CP violation. The CP conserving portion of the penguin diagrams involves a GIM cancelation between the c, u quarks and hence enters significantly at scales below the charmed quark mass. On the other hand, in the CP violating component, the GIM cancelation is between the t, c quarks and thus this piece is short distance dominated. At lowest order, before renormalization group enhancement, one obtains the following effective interactions for the diagram of Fig. VIII-2, ^Vus In ^
L ^
ln
+ V;dVts In ^ |
m? + ^
^ r^
^R
m
^
^
)
^
c\
(3.30) We have used a scale /iR instead of the up quark mass and have quoted only the logarithmic mt dependence. The quarks q = u,d,s are summed over and Qq is the charge of quark q. Note that since the vector current can be written as a sum of left-handed and right-handed currents, this is the only place where right-handed currents enter Hw. The gluonic penguin contains the right-handed current in an SU(3) singlet, hence retaining the (8L, lfl) property ofHw. However, the electroweak penguin introduces a small (8^,8^) component. The full result can be described with the four-quark AS = 1 operators,
2HC + 2HD , O 5 = d^(l
+
7b)X
2HC - 3HD , O 6 = J 7 / i ( l + 75)5 O7 = |g>y^(l + 75)d
(3.31)
222
VIII Kaons and the AS = 1 interaction
3 O8 = - 2 ^ 7 / i ( l + 75)4/ qj'fil
- ^fb)QqQi i
where q = u,d,s are summed over in 05,6,7,85 * a n d J is the charge of quark q and HA = <*7/i(l + 75)^ ^ ( 1 + 75)5 ,
are
color labels, Qq
# c = d7/x(l + 75)5 ^ ( 1 + 75 K
rf7/i(l + 75)5 vr/Hl + 75)ix , JETD = ^7/i(! + 75)5 57^(1 + 75)5. (3.32) The operators are arranged such that Oi52,5,6 have octet and A / = \ quantum numbers, 03(04) are in the 27-plet with A7 = ^(A7 = | ) , while OY$ arise only from the electroweak penguin diagram. The full hamiltonian is (3 33)
•
-
A renormalization group analysis of the coefficients [BuBH 90] yields = 1.90 - 0.62r ,
c5 = -0.011 - 0.079r ,
c2 = 0.14 - 0.020r ,
ce = -0.001 - 0.029r ,
c3 = c 4 /5 ,
c7 = -0.009 - (0.010 - 0.004r)a ,
c4 = 0.49 - 0.005r ,
c8 = (0.002 + 0.160r)a ,
Cl
(3.34)
with A - 0.2 GeV, /i R ~ 1 GeV, mt = 150GeV and r = -V^Vts/V*^. The number multiplying r has a dependence on mt whereas (within the leading logarithm approximation) the remainder does not if mt > MwThe r dependence in C4 arises only because of the electroweak penguin diagram. This hamiltonian summarizes the QCD short distance analysis and is the basis for estimates of weak amplitudes.
VIII-4 The A/ = 1/2 rule
Phenomenology In the decays K —• TTTT, the 5-wave two-pion final state has a total isospin of either 0 or 2 as a consequence of Bose symmetry. Thus, such decays can be parameterized as
^ o
= Ao ei6° - V2A2ei6*
,
(4.1)
VIII-4 The AI = 1/2 rule
223
where the subscripts 0,2 denote the total TTTT isospin and the strong interaction 5-wave TTTT phase shifts <5j enter as prescribed by Watson's theorem (c/. Eq. (C-2.15). There are, in principle, two distinct 1 — 2 amplitudes A2 and Af2. These are equal if there are no AI = 5/2 components in the weak transition, as is the case in the Standard Model if electromagnetic corrections are neglected. The experimental decay rates themselves imply r+Tr-1 = 5.56 x W-7mK , ^O«O\ = 5.28 x W-7mK , \AK+-+n+*o\ = 3.72 x 10-smK .
(4.2)
The test to see whether Af2 = A2 depends on the imprecisely known TTTT phase difference 80 — 82- For the value 8$ — 82 — 45°favored by fits to 7T7T scattering there appears to exist some AI = 5/2 effect, which would presumably arise from electromagnetic corrections, while for 80 — 82 — 57° there is none. We shall neglect this possibility from now on, and employ just the isospin amplitudes \A0\ = 5.46 x 10-7mK ,
\A2\ = 0.25 x 10" 7 m^ .
(4.3)
The ratio of magnitudes, \A2/Ao\~l/22 ,
(4.4)
indicates a striking dominance of the AI = 1/2 amplitude (which contributes to Ao) over the AI = 3/2 amplitude (which contributes only to A2). This enhancement of Ao over A2, together with related manifestations to be discussed later, is called the AI — 1/2 rule. As we have seen in previous sections, a naive estimate (and even determinations which are less naive!) do not suggest this much of an enhancement. However, the factor 20 dominance of AI = 1/2 effects over those with AI — 3/2 is common to both kaon and hyperon decay. A similar enhancement of AI — 1/2 is found in the K —• TTTTTT channel. In this case, it is customary to expand the transition amplitude about the center of the Dalitz plot. For the decay amplitude K(k) —• 7r(pi) n(p2) ft(ps), the relevant variables are Si
= (k-
(4.5) x
y so
, s
o
where S3 labels the 'odd' pion, i.e. the third pion in each of the final states n+n~n°,nonO7T+,ir+Tr+n~. The large A / = 1/2 amplitudes are
224
VIII Kaons and the AS = 1 interaction
considered up to quadratic order in these variables while the A / = 3/2 amplitudes contain only constant plus linear terms, O^+V-KO
=
OI -
2a3
+
(6i
|
( W +
^ -
= 2(ai + a 3 ) -
^j iS
(4.6)
, i6
+ 2c (V + ^
where ai,6i,c,d are A/ = 1/2 amplitudes, 03,63,623 are A7 = 3/2 amplitudes, and the phases {61} in <5MI (= &M — S\) and 621 (= 62 — f)\) refer to final-state phase shifts in the 7 = 1 , 2 and mixed symmetry 7 = 1 states respectively. Because of the relatively small Q-value for such decays (QnTTTT = ™
a3 = 0.37 ± 0.018 , 63 = -0.646 ± 0.125 , d = -3.39 ± 0.33 .
623 = -2.3 ± 0.31 ,
(4.7) Dominance of the A7 = 1/2 signal is again clear in magnitude and in slope terms, e.g. we find at the center of the Dalitz plot, las/ail-1/26 .
(4.8)
In SU{3) language, the dominance of A7 = 1/2 effects over A7 = 3/2 implies the dominance of octet transitions over those involving the 27plet. This is a consequence, within the Standard Model, of the fact that the A7 = 1/2 27-plet operator contributes relative to the A7 = 3/2 27plet operator with a fixed strength given by the scale-independent ratio of coefficients C3/C4 ~ 1/5 (viz. Eq. 3.34). The 27-plet operator then gives only a small contribution to the A7 = 1/2 amplitudes, with the major portion coming from the octet operators. We shall therefore ignore the A7 = 1/2 27-plet contribution henceforth.
VIII-4 The A/ = 1/2 rule
225
Chiral lagrangian analysis The left-handed chiral property of the Standard Model may be directly tested by the use of chiral symmetry to relate the amplitudes in K —> TTTTTT to those in K —> TTTT. We have already constructed the effective lagrangians for (8/,, 1#) and (27^,1#) transitions. Dropping g^7 , the nonleptonic decays are described by the two parameters gs and g^ at O(E2). Let us see how well this parameterization works, and afterwards add O(E4) corrections. The two free parameters may be determined from AQ and A^ in K —> TTTT decays. Prom the chiral lagrangians of Eqs. (2.4), (2.8b), we find (3/2) (
m2}
<
(4
g)
which yields upon comparison with Eq. (4.3), gs ~ 7.8 x l O " 8 ^ ,
5g
/2)
- 0.25 x l ( T 8 i ^ .
(4.10)
The K —• TTTTTT amplitude may be predicted from these. Because there are only two factors of the energy, no quadratic terms in Eq. (4.6) are present in the predictions, 1 (1/2)
v
U---A
,(3/2)
_
^ 4
(3/2)
_
A 2 m^
'"A
i^
fi
'
7T-
5
[
which correspond to the numerical values (again in units of 10~ 7),
+7r+jr _
= 0.47 + 2 . 3 y .
These are to be compared to the experimental results, T+n_n0
= 9.15 + 14.1 y - 4.85 f + 0.88 x2 ,
^ ^ - ^ = -0.71 + 1 . 3 y , +7r+7r _
(4.12)
= 0.71 + 2.9 y ,
with error bars given previously in Eq. (4.7). This comparison can be seen in Fig. VIII-3, where a slice across the Dalitz plot is given. Also
226
VIII Kaons and the AS = 1 interaction
shown are the extrapolations outside the physical region to the 'soft-pion point' where either p+ —> 0 or j£ —> 0. Predictions at these locations are obtained by using the soft-pion theorem. The chiral relations clearly capture the main features of the amplitude and demonstrate that the K —• 3?r A/ = 1/2 enhancement is not independent of that observed in K —• 2?r decay. However, for the A/ = 1/2 transitions, knowledge of the parameters c and d which accompany the quadratic kinematic terms (cf. Eqs. (4.6),(4,7)) allows us to do somewhat better. The kinematic dependence of x2 or y2 can come only from a chiral lagrangian with four factors of the momentum, and only two combinations are possible: Aquad = 7i& • PoP+ • P- + 72 (k • k+Po ' P- + k • p-p0 - p+)
•
(4.13)
Such behavior can be generated from a variety of chiral lagrangians, (4.14)
g'lTr
+
However the predictions in terms of 7$ are unique. Fitting the quadratic terms to determine 71,72 yields the full amplitude, -*O = (9.5 ± 0.7) + (16.0 ± 0.5) y-4.85 y +0.88 x2 , (4.15) which provides an excellent representation of the data. Final state interaction effects also provide an important contribution and must be included in a complete analysis [KaMW 90]. Note that in the process of determining the quadratic coefficients, the constant and linear terms have also become improved. This process cannot be repeated for A/ = 3/2 amplitudes due to a lack of data on quadratic terms.
-5-4-3-2-1
0
1 2
3 4
5 6
Fig. VIII-3 Dalitz plot
VIII-5 Rare kaon decays
227
Vacuum saturation The discussion of direct calculations of the nonleptonic amplitudes is beyond the scope of this book. Suffice it to say that no treatment is presently adequate. Let us give the simplest estimate, called vacuum saturation, as a convenient benchmark with which to compare the theory. For simplicity we consider only O\ (the largest A/ = 1/2 operator) and O4 (the A/ = 3/2 operator), ,
(4.16)
with c\ ~ 1.9 and C4 ~ 0.5. The vacuum saturation approximation consists of inserting the vacuum intermediate state between the two currents in any way possible, e.g. <7r+(p+)7r-(p_) \drf (1 + 75) uu-f (1 + 75) s = <7T-(P_) \d>fw\ 0) <7T+(P+) \u*f8 + <7r+(p+)7r-(p_) \upfual 0> (0 14 n(p_ - P + In obtaining this result the Fierz rearrangement property den* (1 + 75) uaup^ (1 + 75) sp = daY (1 + 75) spup^ (1 + 75) ua has been used, where a, j3 are color indices which are summed over. In addition, the color singlet property of currents is employed, (0 \dal^sp\ K°(k))= iy/2FKk^
.
(4.18)
Within the vacuum saturation approximation, we see that the amplitudes are given completely by known semileptonic decay matrix elements. Putting in all of the constants, we find that AQ = —V* dVwFn ( m i - ml) cx = 0.84 x 10~7ra# , (4.19) ^2 =
5—-KdV^F,. (m^ - m£) c
4
7
= 0.42 x 10" r
We see that the above estimate of A2 works reasonably well, but that Ao falls considerably short of the observed A/ = 1/2 amplitude. This demonstrates that vacuum saturation is not a realistic approximation. However, it does serve to indicate how much additional A/ = 1/2 enhancement is required to explain the data.
228
VIII Kaons and the AS = 1 interaction VIII-5 Rare kaon decays
Thus far, we have discussed the dominant decay modes of the kaon. There are, however, many additional modes which, despite tiny branching ratios, are the subject of intense experimental activity. We can divide these into three main categories. 1) Forbidden decays - These include tests of the flavor conservation laws of the Standard Model. Positive signals would represent signals of physics beyond our present theory. 2) Rare decays within the Standard Model - These include decays which occur only at one-loop order. Such processes can be viewed as tests of chiral dynamics as developed in this and preceding chapters (e.g., radiative kaon decays) or as particularly sensitive to the particle content of the theory (e.g., K+ —• TT + Z/^ (£ = e^fi^r) probes the top quark mass and the number of light neutrinos). 3) CP violation studies - As will be discussed in the next chapter, the kaon system has thus far provided the only positive information on CP violation.* Any of these have the potential to yield exciting physics. We shall content ourselves with discussing only a small sample of the many possibilities. Consider first the rare decay K+ —• n+U£U£. This mode is often called 'K+ to TT+ plus nothing', in reference to its unique experimental signature. This process can take place only through loop diagrams, such as those in Fig. VIII-4. Calculation of these Feynman diagrams leads to an effective lagrangian of the form [ImL 81, HaL 89]
< • ( ) y^
()••
Fig. VIII-4 The decay The observed baryon-antibaryon asymmetry of the universe also requires the existence of CP violation within the standard cosmological model.
VIII-5 Rare kaon decays
229
and
where Xj =
(4-z) 3 x x D(x) = (5.2) 2 4 4 1— x (1-x) {l-xf 8 In this formula we have neglected mj/rriyy (£ = e,/x, r). The matrix element in this case is simple, being related by isospin to the known charged current amplitude
(?r+(p) [sj^d] K+(k)) = V2
(k)) = f+{q2) {k +p)^ (5.3) with /+(0) = —1, and yields a straightforward prediction for the branching ratio in terms of the number Nu of light neutrinos, Brf
+
(TT°(P) \S^^U\ K
7
(5.4)
W'
= 0.7 x 10T 6 D{xc) + s2 The precise branching ratio depends on the KM angle and top quark mass. Although present estimates suggest only a range of branching ratios, 0.5-8.0 x 10~10, future determinations of the KM angles and mt can provide a very nontrivial test of the Standard Model at one-loop order. This rate seems to be reasonably insensitive to long distance physics. A different class of rare decays consists of the radiative processes K$ —> 77 and KL —» TT°77. These transitions provide interesting tests of chiral perturbation theory at one-loop order. In this case, the long distance process, Fig. VIII-5, is dominant. An important feature is that there is no tree level contribution at order E2 or E4 from any of the strong or weak chiral lagrangians because all of the hadrons involved are neutral. Thus the decays can only come from loop diagrams, or from lagrangians at O(EG). There is also an interesting corollary of this result concerning the renormalization behavior of the loops. Since there are no tree level counterterms at O(E4) with which to absorb divergences from the loop diagrams, and recalling that we have proven all divergences can be han-
a*
(a)
(b)
Fig. VIII-5 Long distance contributions to radiative kaon decays.
VIII Kaons and the AS — 1 interaction
230
died in this fashion, it follows that the sum of the loop diagrams must be finite. This is in fact born out by direct calculation. For Ks —> 77, the prediction of chiral [D'AE 86, Go 86] loops is given in terms of known quantities such as a2m\ i j S f _> 7 7
1-
167T 3 2
Tnz,
mK
mK
In2 Q{z) - 2m In Q(z)]
F(z) = l-z[ir 1 Q(z) =
(5.5)
where g$ is the nonleptonic coupling defined previously in Eq. (2.4b). The theoretical one-loop branching ratio,
g S n = 2-0 x 1(T6 ,
(5.6a)
compares favorably with the recent measurement +77
=(2.4±1.2)xlO-6
(5.6b)
without the need to consider possible contributions at O(Ee). The case of KL —> n0/yj is also instructive. Again, one-loop contributions are finite and [EcPR 88] unambiguous. Indeed, we know that KL —• TT°77 and KL —* 77 are related by the soft-pion theorem in the limit"pfc—»• 0. Explicit evaluation yields ,1/2
dT dz / X
where z =
z—
ml
mK
V
m2
\
rn?KJ
(5.7)
F(z)
and X(a, b) = 1 + a2 + b2 - 2(o + b + ab)
(5.8)
Integration over 2 yields the branching ratio r
(loop)
- 6.8 x 10
(5.9)
For this reaction, there is also an O(E6) correction which could be important [Se 90], viz. the diagram of Fig. VIII-5(c) which shows the effect of p-exchange. Such a contribution would have a very different spectrum from that of Eq. (5.7). However, experiment [Ba et al. 90] reveals a photon energy distribution dT/dz which matches that of Eq. (5.7), but with the rate B r ^ o ^ o ^ ^ (2.1 ± 0.6) x 10~6. No evidence exists for a p-exchange contribution.
Problems
231
Problems 1) Ke3 decay The ratio f+^n° (0)/f+O/ir (0) of semileptonic form factors is a measure of isospin violation. Part of this quantity arises from 7r°-r/g mixing. a) By diagonalizing the pseudoscalar meson mass matrix, show that rrid ^ mu induces the mixing |TT°) = cose\(fs) + sine\(fs) where e ~ V3(md — mu)/[4t(m8 — TO)]and TO = (mu + TO^)/2. b) Demonstrate that this leads to the result (c/. Eq. (1.11))
2) Current algebra/PCAC and K -> 3?r decay The results derived in Sect. VIII-4 with effective lagrangians can also be obtained by means of current algebra/PCAC methods. a) Using PC AC, show that the soft-pion limit of the K —> 3?r transition amplitude is given by qa-*0
a
c
b {
Fn
where Q^ = f d3x Afifa t) is the axial charge, b) Demonstrate that this may be also written as
)h^(rfyqbK°c\Hw(0)\K2)
= y(7riKcqc\[Qa,Hw(0)]\K%) ,
where Qa is an isotopic spin operator, and hence that
1 where / = 5, | signifies question. c) Use a linear expansion Eq. (4.6) with c=d=0) to corrections of order
+ 0)
the isospin component of the quantities in of the K -+ 3n transition amplitude, (i.e. to reproduce the results of Eq. (4.11), up ml.
IX Kaon mixing and CP violation
In our discussion of the electroweak interactions in Chap. II, we saw that the KM matrix contains imaginary couplings which have the potential to violate CP invariance. These arise originally in the Yukawa couplings of quarks to the Higgs bosons, but after the definitions of mass eigenstates, they appear only as a single phase in the W± gauge couplings. In this chapter, we focus on the system of neutral kaons to describe how this phase gives rise to CP violating observables. IX-1 K°-K° mixing It is clear that K° and K°should mix with each other. In addition to less obvious mechanisms discussed later, the most easily seen sourceof mixing occurs through their common TTTT decays, i.e. K°<-> TTTT <-> K°. We can use second-order perturbation theory to study the phenomenon of mixing. Mass matrix phenomenology Writing the wavefunctions in two-component form \
,
)
we have the time development
232
(1.1)
IX-1 K°-K° mixing
233
where, to second order in perturbation theory, the quantity in parentheses is called the mass matrix and is given by*
2 Uj
2mK
(K?\Hw\n (1.3) Here the absorptive piece F arises from use of the identity ^—^
= p(-^-=r)-in6(En-u;)
u — t,n + xe
\u) —
,
(1.4)
bjn)
and hence involves only physical intermediate states 1
n> (n\Hw\K°) 2n8{En - mK)
.
(1.5)
Because M and T are hermitian, we have M21 = M*2 and F2i = T\2> The diagonal elements of the mass matrix are required to be equal by CPT invariance, leading to a general form A p2 where A, p2 and g2 can be complex. The states K° and K° are related by the unitary CP operation,
CP\K°)=tK\K°)
(1.7)
with \£K\2 = 1. Our convention will be to choose £# = — 1. The assumption of C P invariance would relate the off-diagonal elements in the mass matrix so as to imply p = q,
{K°\neS\K°) = (K°KCPJ^CP HeS (CP)-1CP| ^°> = (K°|Weff| K°), (1.8) where (K° \Hes\ K°) is defined in Eq. (1.3). Combined with the hermiticity of M and F, this would imply that Myi and Fi2 are real. In the actual CP-noninvariant world, this is not the case and we have for the eigenstates of the mass matrix, \KL ) = —=j s
V\P\
2
+ kl
2
\p\K°)± q\K0)} ,
(1.9)
* The factors l/2ra/<- are required by the normalization convention of Eq. (C-2.7) for state vectors.
234
IX Kaon mixing and CP violation
where from the above discussion, we have p_
Mn - |Fi2
The difference in eigenvalues is given by 2qp = (mL - ms) - ~{TL - Ts) ( i \1/2/ i \1/2 = 2 ( M12 - ^ r i 2 Mx*2 - - H 2 ~ 2ReM12 - i
(1.11) where the final relation is an approximation valid if CP violation is small (1 » ImMi2/ReMi2). The subscripts in KL and Ks, standing for 'long' and 'short', refer to their respective lifetimes, which differ by a factor of 580. To understand this large difference, we note that if CP were conserved (p = q), these states would become CP eigenstates K± (not to be confused with the charged kaons K±\), \Ks) ^
\K°+) ,
]KL)
0
\K°±) = - L [\K°)T |^ )] ,
_^
]Ko_} f
P=Q
CP\Kl) = ±\K%) .
n 22)
In this limit, which well approximates reality, Ks would decay only to CP-even final states like TTTT, whereas KL would decay only to CP-odd final states. Since the phase space for the former considerably exceeds that of the latter at the rather low energy of the kaon mass, Ks has much the shorter lifetime. The states KS,L, expanded in terms of CP eigenstates, are
- Ts)
(1-13) K°-K° mixing can be observed experimentally from the time development of a state which is produced via a strong interaction process, and therefore starts out at t = 0 as either a pure K° or K°,
(1.14)
9±(t) = I
IX-1 K°-K° mixing
235
where AF = T$ — TL and Am = m^ — 7715, each defined to be a positive quantity. From such experiments, the very precise value Am = (3.522 ± 0.016) x 10"12 MeV has been obtained. Box diagram contribution The Standard Model predicts K°-K° mixing at about the observed level. However, it is difficult to precisely calculate the mass difference Am. There are two main classes of effects, short-distance box diagrams of Fig. IX-1 (a),(b) and long-distance contributions like those in Fig. IX-2, which can generate the mixing amplitude. We turn now to a study of the first of these. When calculated at the quark level, one can envision several Feynman diagrams which mix K° and K° (see Fig. IX-1). The box diagrams in Fig. IX-1 (a),(b) lead to an effective hamiltonian dKs)2 H(xc)m%
+ (V^Vts)2
H(xt)m2m
+ 2 (V&VtaV&Vn) G(x« x t)m2cm] OAS=2 + h.c. , where X{ = mf/M^ and [InL 81] l
9
1
3
1
,
4 + 4T^-2(l^J-2(T^ l n a ; ' (1.16)
v
* ' 4(l-s)(l-y) The factors 771, 772 and 773 are QCD short-distance renormalization factors analogous to the renormalization group coefficients for the AS = 1 hamiltonian discussed in the previous chapter. For rrtt > Mw,
u.c.t
u,c,t
u,c,t
W (a)
<j
s
w s
u,c,t d (b)
Fig. IX-1 Box (a),(b) and other contributions to CP violation.
236
IX Kaon mixing and CP violation
0.2 GeV, they have the values [BuBH 90], 7/i ~ 0.85 ,
% ^ 0.62 ,
773 ~ 0.36 .
(1.17)
Note that the GIM mechanism is at work here, in that the u, c, t intermediate states would exactly cancel each other if their masses were equal. The mass of the ?x-quark has been neglected with respect to heavy quark masses in writing Eq. (1.15). Given present bounds on the KM elements and the t quark mass, the most important contribution to the real part of H^ox is that of the c-quark. In view of this, and noting that H(xc) ~ 1, we then have Re H^T * ^2^Re(F c * d F c s ) 2
Vl
OAS=2 .
(1.18)
At this stage one can provide a rationale for neglecting Fig. IX-l(c). In this diagram the GIM cancelation is logarithmic rather than power law, so no factor of m2 appears. Evidently, all matrix elements of other diagrams should be suppressed compared to the box diagram by factors like fJ?/rn% (where /i is a hadronic scale) or m^/m2. This has been verified for Fig. IX-l(c), and will be discussed below as a 'long-distance' effect. The double penguin diagram of Fig. IX-l(d) has a long-distance part and a short-distance component, obtained when the loop momenta are high. Separating the two is difficult, but the short-distance piece has been studied [EeP 87]. To make contact with phenomenology, one must evaluate the matrix element of OAS=2 between K° and K° states. It is conventional to express the results in terms of the so-called B-parameter, ^
,
(1.19)
where B = 1 corresponds to the simple vacuum saturation approximation described in Sect. VIII-4. A wide variety of models have been applied to determine this amplitude, with estimates for B which vary considerably [PaT 89]. One approach attempts to extract B directly from experiment by using chiral symmetry and SU(3) [DoGH 82]. The point is that both OAS=2 and the A/ = 3/2, AS = 1 operator O4 (see Eq. (VIII-3.31)) are formed from a pair of SU(3) octet left-handed currents, and because of their quantum numbers, each belong to the same 27-dimensional representation of SU(3). This implies the SU(3) relation, = v^2 .
(1.20)
The right-hand side of the above can be related to the K —• 2n matrix element with the aid of chiral symmetry. Using either soft pion techniques
IX-1 K°-K° mixing
237
or the chiral lagrangian given in Eq. (VIII-2.7), we find
(n°(k)\O4\K0(k))=2VlF«
/
+ + 2(n n°\OAI=3/2\K )
, (1.21)
or with on-shell kaons (k2 = rrv^-)^ 8y/2F,
ml
where rji is any one of the QCD factors 771, 772, 773. Expressed this way the result is independent of the scale chosen in QCD short-distance corrections, as it cancels in the ratio of rji/c^. This occurs because the QCD corrections are SU(3) symmetric and treat all members of the 27-plet in an equal fashion. Written in terms of the 5-parameter the result is* B = 0.33[as(fi)]2/9
.
(1.23)
We stress that any model which respects both chiral symmetry and 5f/(3), and also obtains the correct K+ —• TT+TT0 amplitude must lead to this result. Often, model dependent estimates get an answer different from Eq. (1.23) because they incorrectly predict K+ —> TT+TT0. The only significant issues which can affect the above determination of B are the possible breaking of SU{3) symmetry and the contribution of higher-order chiral corrections. One attempt [BiSW 84] to calculate the higher-order effects has examined the leading logarithmic corrections obtained when one calculates loops in chiral perturbation theoryt. In the language of Chap. VI, this consists of keeping the quantities m In(ra 2//i2) which occur in loop contributions, but dropping all factors of ra4 and all effects of the low energy constants occurring at O(E4). The results depend strongly on the 'chiraF scale, which we may write as /xchin that occurs in the chiral logarithmic factor - no corrections are obtained if Mchir — 1/2 GeV whereas a 50% enhancement obtains if /xchir — 1 GeV. Unfortunately, this logarithmic approximation has not been accurate in other settings, and a full chiral perturbation theory calculation cannot yet be done because one does not have sufficient information to pin down the needed low-energy constants at O(E4). A final resolution will require a better understanding of the origins and structure of the weak chiral lagrangian. Lattice QCD computer studies in the quenched approximation (i.e. no fermion loops) presently indicate a value B ~ 0.7, but are also presently obtaining values for the K+ —• TT+TT0 amplitude which are about twice as large as the empirical value. * Here a slight correction has been included for the effects of 77°-7r°mixing which can lead to a simulated A/ = 3/2 contribution to the K —> 2ir decay process. This is discussed in Sect. IX-3. t The paper [BiSW 84] contains an error in the normalization of the B factor, which we have corrected for in the quoted result.
238
IX Kaon mixing and CP violation
If we piece together all of the ingredients to the short-distance predictions, we obtain from the charm contribution to the box diagrams the mass difference, (Am)^ o r y ~r 7 l B(Am) e x p t .
(1.24)
This result is clearly of the right order of magnitude, but there also exist potentially sizable but incalculable long-distance corrections, which preclude a clean prediction of the mass difference. The long-distance contributions are generically of the form of Fig. IX2, with the AS — 1 weak hamiltonian acting twice. As discussed above, a naive analysis suggests that such contributions should be suppressed by a factor ~ m2K/m2c ~ 1/10 with respect to the short-distance effect. However, empirically the AI = 1/2 portion of H^s=1 has a factor of twenty enhancement over the simplest estimates (see Sect. VIII4), such that the naive factor should perhaps be modified to be about (20 7n#/ra c ) 2 ~ 40. Although the effect is not quite that large in reality, estimates of K° <-• (TT0, TJ) <-> K° and K° <-»TTTT <-• K° intermediate states yield long-distance effects comparable to the experimental mass difference. It is unfortunately not possible to be precise about these, but they can be large. For example, the K —• rj —> K° intermediate state by itself yields a contribution to Am of about 1.1 times the experimental result! This is estimated by using SU(3) plus chiral symmetry, and if one allows a 30% uncertainty in the prediction, it produces a large uncertainty by itself in the predicted mass difference. The TTTT contribution also has significant uncertainty. We can only conclude that, while a precise prediction of Am is not possible, the observed size is roughly compatible with that expected in the Standard Model. IX-2 The phenomenology of kaon CP violation The TTTT final state of kaon decay is even under CP provided the strong interactions are invariant under this symmetry. For the 7r°7r° system, this is clear since TT0 is itself a CP eigenstate, CP|TT0) = — |TT°), and the two pions must be in an 5-wave (£ = 0) state, CPITTV)
= ( - l f ( - l ) V V ) = +|7r°7r°).
K°
(a)
(b)
Fig. IX-2 Long-distance contributions.
(2.1)
IX-2 The phenomenology of kaon CP violation
239
The corresponding result for charged pions reflects the fact that ?r+ and 7T~ are CP-conjugate partners, CP|TT ± ) = — |TTT). We have seen that if CP were conserved, the two neutral kaons would organize themselves into CP eigenstates, with only Ks decaying into TTTT. Alternatively, KL decays primarily into the TTTTTT final state, which is CP-odd if the pions are in relative 5-waves. The observation of both neutral kaons decaying into 7T7T is then a signal of CP violation. There can be two sources of CP violation in KL —> TTTT decay. We have already seen that KK mixing can generate a mixture of the CP eigenstates in physical kaons due to CP violation in the mass matrix. There also exists the possibility of direct CP violation in the weak decay amplitude, such that the CP-odd kaon eigenstate \K^_) makes a transition to 7T7T. These two mechanisms are pictured in Fig. IX-3. The Kirn decay amplitudes have already been written down in Eq. (VIII-4.1) in terms of real-valued moduli Ao, A2- This decomposition is a consequence of Watson's theorem, and relies in part upon the assumption of time reversal invariance. However, if direct CP violation occurs, Ao and A
,
A2 = \A2\e^
,
(2.2)
with CP violation in the decay amplitude being characterized by the phases £o and £2- Consequently, the KQ —• TTTT and KQ —> TTTT decay amplitudes assume the modified form
+r- = \A0\e*°eiS°+ i ^ e ^ e * ,
Using the definitions of KL and Ks in Eq. (1.13), a straightforward calculation leads to the following measures of CP violation: _
KL
K+
(a)
/
K,
^
(b)
Fig. IX-3 Mechanisms for CP violation.
240
IX Kaon mixing and CP violation
where e — e + z£o ? ~~
A/2
iei(62-*o)
A2 Ao
-v/2
A2
/1I m A2
Im Ao
Ao VRe A2
Re A o
(2.5) The expression for e can be simplified by approximating the numerical value Ara/Ar = 0.478 ± 0.003 by Am/AT ~ 1/2. This yields the approximate relation, e i7r / 4 1 e y/2 Am '
i 1 Am+^AT
which we shall use repeatedly in the analysis to follow. In addition, since the rate for K —• TTTT is much larger than that for K —> TTTTTT, and K° —• TTTT is in turn dominated by the 7 = 0 final state because of the A / = \ rule, we have .
(2.7)
The above relations allow us to write i
Im M12
\m M12 , ^ _ e'f ^ Im M12 , Im Ap\ Am + *>) ~ y/2 V2Re M 12 + Re Ao) '
l
'
where UJ = Re^/ReAo ~ 1/22. We see that e is sensitive to CP violation in the mass matrix, while ef is a unique manifestation of direct A 5 = 1 CP violation. All CP-violating observables must involve an interference of two amplitudes. In Eq. (2.8), the quantity e expresses the interference of K° —> 7T7T with K° —• K° —• 7T7T, while ef involves interference of the 7 = 0 and 7 = 2 final states. The formulae for e and ef exhibit an important theoretical property. Since the choice of phase convention for any meson M is arbitrary, its state vector may be modified by the global strangeness transformation \M) - • e iA5 |M). For the K° and K° states, this becomes \K°) - • eiX\K°) ,
iX \K°) -> e\K°)
,
(2.9)
which has the effect,
Im Ar _^ Im A/ Re A/ ""* Re A/
'
Im Mu _^ Im M i2 _ Re M i2 ~" Re M i2
'
l
. . ''
We see that the values of e and e' are left unchanged. Various phase conventions appear in the literature. In the Wu-Yang convention, A is
IX-2 The phenomenology of kaon CP violation
241
chosen so that the Ao amplitude is real-valued. This is always possible to achieve by properly choosing the phase of the kaon state. However, it is inconvenient for the Standard Model, where the Ao amplitude naturally picks up a CP-violating phase. We shall therefore employ the convention in which no such phases occur in the definitions of the kaon states. It was in the K —» TTTT system that CP violation was first observed. At present, the status of measurements is \c\ = (2.263 ± 0.023) x 10~3 , f 0.0023 ± 0.0007 [Ba 92] ,
(2.H)
={
[ 0.0006 ± 0.0007 [Wi 92] , f (46.9 ± 2.2)° h_ = phase(r7+_) = < [(43.2 ±1.5)°
[Wi92] ,
= phase(r/00) = (47.1 ± 2.8)°
[Ca et al 90] .
[Ca et al 90] ,
As this book goes to print, there remains a discrepancy as to whether a nonzero value of e; has been found. This is a critical issue in the study of CP violation whose resolution is eagerly awaited. A violation of CP symmetry has also been observed in the semileptonic decays of KL and Kg. These are related to matrix elements of the weak hadronic currents. Since K° must always decay into e^ue7r~ while K° goes to e~Pe7r+, we have
1-6
If the semileptonic decays proceed as in the Standard Model, there is no direct CP violation in the transition amplitude, so that ~ l + 4Ree .
(2.13)
Since Re e = Re e, the above asymmetry is sensitive to the same parameter as appears in the KL —• TTTT studies. Here, measurement gives Re e = (1.635 ± 0.060) x 10"3 = |e| cos(44 ± 3)° ,
(2.14)
which is consistent with the experimental values from K —> TTTT. Finally, although one ordinarily emphasizes the phenomenon of CPviolation in studies of kaon mixing, precision experiments also probe the CPT transformation. For example, a prediction of CPT invariance, the equality of phases <^+_ = <^oo (up to very small corrections from e;), is
242
IX Kaon mixing and CP violation
seen to be consistent with existing data (0.2 ±2.9)°
[CaetoZ.90],
(2.15)
In addition, kaon data can be used to provide a sensitive test of the wellknown CPT prediction that the mass of a particle must equal that of its antiparticle. Specifically, the impressive bound \(mKo —m^o)/m Ko\ < 4x 10~18 has been obtained [Ca et al. 90]. Further study of CPT invariance is left to Prob. XI-3. IX-3 Kaon CP violation in the Standard Model After diagonalization, there can remain a single phase in the KM matrix. This phase generates the imaginary parts of amplitudes which are required for CP violation. It is a physical requirement that results be invariant under rephasing of the quark fields. As a consequence, all observables must be proportional to Im A(4) = A2\er) = cic2c$s\s2szs8 ,
(3.1)
written in the notation of Sect. II—4. In particular, all CP-violating signals must vanish if any of the KM angles vanish. We shall now study the path whereby this phase is transferred from the lagrangian to experimental observables. For kaons, we have seen that the relevant amplitudes are those for AS = 2 K°-K° mixing and for AS = 1 K—> TTTT decays. Tree level amplitudes in kaon decay can never be sensitive to the full rephasing invariant, so that one must consider loops. Typical diagrams are displayed in Fig. IX-4. Experiment can help in simplifying the theoretical analysis. Note that e' is sensitive to AS = 1 physics through the penguin diagram [GiW 79], while e is sensitive to AS = 2 mass matrix physics as well as to AS = 1 effects. However, since experiment tells us that e » e r, it follows that the AS = 1 contributions to e must be small. Likewise, the long-distance contributions of Fig. IX-2 and the contribution of Fig. IX-l(d) must both be small because each also involves the AS = 1 interaction. This
(a)
(b)
Fig. IX-4. (a) Penguin and (b) electroweak-penguin contributions to CP violation in AS = 1 transitions.
IX-3 Kaon CP violation in the Standard Model
243
leaves the box diagrams of Fig. I X - l ( a ) , ( b ) as t h e dominant component of e. Moreover, since t h e KM phase 6 is associated with t h e heavy quark couplings, only t h e heavy quark parts of t h e box diagrams are needed. Hence e is very clearly short-distance dominated.
Analysis of e The evaluation of e follows directly from the treatment of the box diagram in Sect. IX-1, and we find (c/. Eq. (1.15)), G2 —^%
Im M12 = x
2
^-^-r)2m2H(xt)
c \r)iml-
\ p )
- rjsm2cG(xc,
x c)
%
s\
(3.2)
The above relations depend on three principal quantities, viz. mt, the KM angles, and the B parameter. At present, we have only the bound mt > 89GeV, so that not much can be inferred about the influence of the t-quark mass. The information on KM angles allows the range A2X6r] = sls2s3ss < 0.7 x 10~4 ,
(3.3)
where no firm value can yet be given because there is no independent measure of 6. However, by taking t h e representative values mt = 150 GeV, mc — 1.5 GeV, A = 1, p = 1/2, we can express e in t h e useful form
[G(xc, xt)~
(3 4)
'
In this way, the magnitude of e is naturally obtained. Penguin contribution to ef The analysis of ef is rather more involved because one must confront A/ = 1/2 amplitudes in K —• TTTT. These are not yet under theoretical control, so that the predictions of ef must necessarily be uncertain. However, we can review the main ingredients of the analysis, which by now have become clear. As can be seen from Eq. (2.8), the most natural scale for the magnitude of ef/e is just u = Re A2/Rje A$ ~ 1/22. Yet, the experimental value of e'/e is found to be suppressed by at least another order of magnitude. This suppression is a consequence of the CP-violating phase 6 which appears
244
IX Kaon mixing and CP violation
only in the heavy quark sector. Its contribution to the direct K —> TTTT process occurs only as a result of virtual oquark and t-quark loops in the penguin diagram of Fig. IX-4(a) and the electroweak-penguin diagram of Fig. IX-4(b), from which we can extract some general features. The gluonic penguin diagram involves the s —• d transition and is purely A / = 1/2. It therefore generates a phase only in A$. The electroweak penguin would seem at first sight to be suppressed when compared to the usual penguin because of the factor of a/as. However, as will be described below, the electroweak penguin is relevant because it can generate a phase for A2. The full prediction emerges only when we have calculated the phases of both AQ and A*i. Let us first give estimates of various contributions to e' using QCD perturbation theory, but without renormalization group improvements. This serves to identify the important physics and can be adjusted afterward to incorporate the effects from short-distance renormalization. The penguin and electroweak-penguin diagrams yield respectively the CP-odd operators given in Eq. (VIII-3.30). If one wishes to explicitly check the rephasing invariance, it is necessary to recognize that CP-odd interference with other diagrams in general involves quantities like Im (V^VtsV^Vtid) which are invariant under rephasing the quark fields. The other feature to extract from the above operatorsis the symmetry structure. The gluonic penguin involves a lefthanded (ds) current which is a member of an SU(3) octet. The remaining current is a flavor singlet, so that overall the interaction transforms as an octet and hence carries A / = 1 / 2 . More precisely, under chiral SU(3) it transforms as (8^,1^). By contrast the electromagnetic coupling in the electroweak penguin is itself an octet, having both left-handed and right-handed portions,
^
^
(U, sR),
(3.5)
and is a mixture of A / = 1 and A / = 0, __ fQq=
2uu — dd — ss uu — dd = ^ ~ + §
uu + dd — 2ss g •
(3-6)
Thus the symmetry structure of the electromagnetic penguin is more complicated, involving SU(3) octet and 27-plet representations, isospin 1/2 and 3/2 portions, and the chiral properties ( 8 L , 1 # ) , (27^,1^) and
(0 In order to complete the estimate of e', one needs to compute hadronic matrix elements. To illustrate this procedure, we shall use the vacuum saturation method, with the understanding that this approximation could be seriously flawed. One interesting new ingredient, beyond those described in Sect. VIII-4, is the presence of matrix elements of scalar and
^x
IX-3 Kaon CP violation in the Standard Model
245
pseudoscalar operators which arise as a consequence of the Fierz relations, rfn)i(l + 75)sj<&7/i(l - 75)% = -2dj(l - 75)Mfc(l + 75)5j .
(3.7)
For matrix elements of such operators, we have
= Bo ( +
^ )
V2SF (l
,
§)
(3.8)
= -iV2B0F U + J2 where So is a constant, F is F^ in the chiral limit and A defines an energy scale in the momentum dependent terms. To obtain these expressions, we have assumed chiral SU(3) symmetry and, for simplicity, have omitted loops. The momentum dependence is required to obtain a nonzero result, and the constant Bo can be identified from our previous chiral lagrangian studies, where the lowest order relations (ra = (mu + m^)/2), m% = (n\m{uu + rfrf)|7r) ,
m2K — (K\m{uu + dd) + msss\K)
, (3.9)
reveal that Bo = $• = ~^r
•
(3.10)
m +m
2m
However, Bo and the quark masses are not separately obtained from chiral symmetry - only the product rhBo or the mass ratio rh/ms is well-defined. In Chap. XII, we shall obtain a more model dependent estimate of ms from hyperon mass splittings, where a first order treatment yields m\ — mp — (A|ras5s|A) — (p\msss\p) + O (m/ms) = msZm .
(3.11)
The matrix element Zm is estimated to lie in the range 0.5 < Zm < 0.8 within a variety of quark models, and so Bo *
m2 TTlA
*— Z ™
m
.
(3.12)
Note also that the momentum dependence is related to FK/FK, since ms + m
ms + m
(3.13) where the equations of motion have been used in the final relation. From Eq. (3.8), we find _
246
IX Kaon mixing and CP violation
With these ingredients now defined, we can piece together the structure of e'/e. Using the physical values of e and Re Ao, we find Im
(TV+Trf°
Inserting as ~ 0.2 (as is appropriate for heavy quarks), mt ~ 150 GeV, mc — 1.5 GeV and fixing A with the FR/F^ analysis, we obtain finally - = 0.0002 A2r) U + 3.9 Z2m
.
(3.16)
Remember that this result still lacks the renormalization group enhancement. However, it does indicate that one expects a small but non-zero value of e'/e in the Standard Model. Additional contributions to e' In the limit of isospin invariance, the gluonic penguin contributes only to Ao. Since e' is proportional to the difference between the phase of Ao and the phase of A%, we need to be certain also about the phase of Ai- A little thought will convince one that A2 is capable of picking up a sizeable phase from small effects because Re A2 is itself a small number. That is, if there is a small CP-violating process which contributes to A2, the phase IDQLA2/R&A2 can be enhanced because of the smallness of Re^2. We can estimate such an effect. For example, if isospin is broken by quark mass effects, then the decomposition into Ao and A2 will not uniquely represent the A/ = 1/2 and A/ = 3/2 transitions. Some A/ = 1/2 effects will appear in the parameter A2. An example of this is given below. The size of such effects would be roughly yiiso—brk
A
*
A
A
m
d
~~ ™"u
~ °-—^T~ '
/ o i v\
(3>17)
where the second factor is a measure of isospin breaking. In this case the phase of A2 is of order Im A2 Re A2
I m Ao Tiid — TRU Re A2 ms
Im Ao Re Ao vrtd — TUU Re A$ Re A2 ms A 22 = 22 • — ~ 00 7 30 Re Ao 30 Re Ao '
(3-18)
where we have used a first order chiral estimate of the quark mass ratio. In Fig. IX-4(b) we display a second example, the electroweak penguin diagram. This can contribute directly to A2 with an estimated size
IX-3 Kaon CP violation in the Standard Model
247
relative to the penguin of Im A™»
~ — Im ^ e n g ,
lmA2 Re ^2
aImAlengReA0 a s Re Ao Re A2
Im Ao n o ImA 0 Re Ao ~ ' Re ^4o (3.19) We see that because of the extra factor of 22 in Eqs. (3.18), (3.19), even suppressed effects can modify ef markedly if they contribute to A^ The above results are order-of-magnitude estimates only, not real calculations. However, they do indicate that a careful study of isospin breaking and the electroweak penguin is required to fully understand e1. It has become conventional to express e'/e in the form e
l
Re An
1/137 0.2
nn
- '
(3.20)
22
r^rT" Im AQ Re A
Im
To determine the correction factor ft, let us first estimate the isospin breaking correction due to the u — d quark mass difference. One such effect is clearly calculated from chiral perturbation theory. The u—d mass difference induces 7r°-r/ mixing,and this in turn influences the transitions K -» 7r°7r°and K+ - • TT+TT0 through the diagram of Fig. IX-5. The ingredients of the calculation are (i)the weak transition,
^K0^m
= ^=Af
,
(3.21)
which is found using the weak lagrangian Eq. (VIII-2.4b) and where the superscript (0) has been added to indicate that one uses AQ before 7r°-77 mixing, (ii) the amplitude for TTO-778 mixing,
^ ^ ^ K - mD -
K+
Fig. IX-5 Effect of 7r°-rj mixing
(3-22)
248
IX Kaon mixing and CP violation
which is found from the chiral lagrangian of Eq. (IV-5.10). Using these to calculate Fig. IX-5 yields )
~ft > 2
\
2
<
(3.23) We see that, subsequent to mixing, this is equivalent to the modified isospin decomposition A - 4<°)+ lmd-mu Is A
_
4 (0)
1
(Q) m
m d - mu
(3.24)
(Q)
v
^
The change in A$ is small, but the effect in A% is relatively larger, with
^
= _ l ! ^ ^ 4 2 ^ _Q M
A2 3\/2 m s - m A2 For our purposes it is most important that a phase has been induced into A2, with magnitude O? = 0.14 .
(3.26)
There will also be a correction for the rf. Although this cannot be calculated in chiral perturbation theory, it can be estimated from quark model relations plus rj — r( mixing that ^iso-brk ^ % +
(3.27)
We are also interested in the contribution of the electroweak penguin operator to Im A
(3.28)
whereas the (8/,, 1R) and (27^, 1#) operators each require two derivatives. Due to the presence of the electric charge operator Q, the matrix element vanishes for the neutral mode K° —> 7r°7r° but not for the charged mode K° —> 7r+7r~. From this, we can determine the relative contributions of the electroweak penguin to A2, AQ in terms of the K° —> TT+TT" amplitude, 22^2 3
Im (^-\Hewp\K) Im (TT+TTiWlKO) '
(
]
IX-3 Kaon CP violation in the Standard Model
249
Again, the need to calculate a hadronic matrix element hinders one's ability to make a firm prediction. However, in the approximation of vacuum saturation we estimate _
22x/2 a
A2
2B% - {m\ - m2.)
Interestingly, the electroweak penguin operator is chirally enhanced. At the effective lagrangian level this follows because its operator is O(E°) in the energy expansion, while the penguin operator is O(E2). This is reflected in the above matrix element by the factor h? /{m2K — rnfy. Note also that in this approximation fiewp is negative so that it enhances e'. Inserting the numerical factors as ~ 0.2, A = 1 GeV, Zm ~ 3/4, we estimate that fiewp — —0.30. This is a surprisingly large correction for an electroweak contribution, and indicates that it can be an important part of the prediction for e'\ The final ingredient is the calculation of short-distance effects in QCD using the renormalization group. This proceeds along the lines described in Sect. VIII-3. The usual penguin can generate a phase primarily in Oi, O5 and OQ to yield Im Ao = ^ K d K s (TT+TT-I £
Tma Oi \K°) ,
(3.31)
i=l,5,6
and the electroweak penguin contributes to A
ImC
* Oi \K°) .
(3.32)
Present estimates [BuBH 90] yield the values Im c7 ~ -0.004as2sss6 = -0.004aA4A2rj , Im c$ ~ -0.16as2Ssss = -0.16a\4A2r) ,
(3-33)
for mt - 150 GeV, A ~ 0.2 GeV, and fi - 1.0 GeV. In addition, the renormalization group analysis including the electroweak penguin induces a phase in the coefficient of the A/ — 3/2 operator O4. This arises from the purely left-handed portion of the electroweak penguin, with transformation property (27^,1^). The result is a phase in A2, of magnitude /C4
Re c4
This result is independent of the QCD scale /i in the short-distance analysis. Note that by use of Eq. (2.8) and the experimental value of e, this
250
IX Kaon mixing and CP violation
piece by itself generates
°0 1 *'/" = - 0 . 3 X 1 0 - ^ ,
U
(3.35)
C1
which can be significant given the size of the leading contribution in Eq. (3.16). The isospin breaking effect discussed above is unchanged by QCD short-distance corrections. We see then that there are several important contributions to e'. Because of the possibility of cancelations and because we cannot yet calculate the matrix elements with acceptable precision, the only safe statement is that e1 is expected to be nonzero, and to occur in the range probed by present experiments. To summarize, we have outlined the main ingredients for the predictions of e and e'. As this is being written, the experimental results have not settled down, and important theoretical uncertainties lie in the KM angles, the t-quark mass, and the hadronic matrix elements. There is cause for optimism that more will be known about all of these in the near future.
IX—4 Electric dipole moments The condition of T invariance forbids the existence of an electric dipole moment for any elementary particle. This can be seen heuristically in the following manner. The only vectorial quantum numbers associated with a particle are its momentum p and spin s. Thus, for a particle at rest the electric dipole moment must be proportional to the spin, dc = d e l ^ s
ll
.
(4.1)
The dipole interacts with an electric field as = -de-E
=-de^
.
(4.2)
Under time reversal E is unchanged (i.e. think of the E field of capacitors or static charges, which do not change under T), but all angular momenta reverse sign. Hence, i^edm is °dd under T.* For a relativistic spin-1/2 particle, the electric dipole moment contribution has the matrix element (p'\j;ia\p)\edm
= ideu(p')afiUql/l5u(p)
,
(4.3)
with qu = (p1 — p)u, which is equivalent to an interaction density ^
.
(4.4)
* One might be concerned that, since T is an antiunitary operator, this conclusion could be changed by adding a factor of i to He^m. However, hermiticity requires that no additional factors of i appear.
IX~4 Electric dipole moments Since FOi = -E{
251
and
this interaction attains the hamiltonian form in the nonrelativistic limit, #edm = [cPxHedm "> ~de((r) • E .
(4.6)
Note that an electric dipole moment also violates parity. The best existing experimental limits on electric dipole moments are
{
1.2 x 10~25e-cm
(neutrons [RPP 90]) ,
(-2.7 ± 8.3) x 10"27e-cm (electrons [Ab et al. 90]) , (4.7) 23 -3.7 x 10" e-cm (protons [RPP 90]) . The extreme sensitivity of these limits can be seen from the observation that if neutrons were expanded to the size of the earth, the limit quoted would correspond to a dipole with a unit electric charge separated by only a micron! In practice, the dipole moment is a sensitive probe of CP violation, especially for theories beyond the Standard Model. The Standard Model can produce an electric dipole moment either through the 0 parameter of QCD or through W-exchange. The discussion of 6 and the strong CP problem is given in the next section. Some of the possible diagrams involving the electroweak interactions are shown in Fig. IX-6. There emerges the very important property that no electric dipole moment is produced at first order in the weak interaction (i. e.
O(Gp)) in the Standard Model This is because the process does not change flavor. Tree-level processes which do not change flavor always involve the combination of KM matrix elements V^V^ for some flavors i, j , and the loop diagrams of Fig. IX-6 (a) involve ]T\ ^f^ij- Hence these are real and cannot involve the CP-violating phase. Theories which do have first order CP violation tend to produce neutron dipole moments close to the experimental bounds. At second order in the weak interactions, there is enough structure to produce a dipole moment. Interestingly, the single quark line diagrams, of which one example appears in Fig. IX-6(b), sum to zero net effect [Sh 78].
(a)
(b)
(c)
Fig. IX-6 Contributions to an electric dipole moment
252
IX Kaon mixing and CP violation
However, with an extra gluon loop a dipole moment occurs at O{G2Fas). Since as ~ 1 at hadronic scales, this provides no extra suppression. The multiquark interactions of Fig. IX-6(c) can also lead to a neutron dipole moment [GaLOPRP 82]. For the electron, CP violation vanishes if the neutrinos are massless. One can obtain a reasonable estimate of the neutron dipole moment from dimensional analysis alone. There is always a factor of Im A^4^ (cf. Eq. (3.1)) associated with any CP-violating process. In addition, the GIM mechanism would cancel the contributions of degenerate mass quarks, so a contribution of the form (mh — m\)/Myy is also expected. Altogether then, we have de(neutron) - e ^ f - ^ - I m A^/i 3 ~ 10"31e - cm ,
(4.8)
where // is a typical hadronic scale (we use /i ~ 0.3 GeV) which has been included to make the dimensions correct, and we have incorporated factors of n as anticipated from loop diagrams. This estimate is in the middle of the range found in model calculations [BaM 89], which span from 10~33 to 10~30e — cm. It is well out of the reach of experiments in the foreseeable future. IX-5 The strong CP problem The possibility of a 0-term in the QCD lagrangian raises potential problems (see Sect. Ill—5). For 0 ^ 0, QCD will in general violate parity, and even worse, time reversal invariance. The strength of T violation (and hence by the CPT theorem, CP violation) is known to be small, even by the standards of the weak interaction. This knowledge comes from both the observed KL —• 2?r decay and bounds on electric dipole moments. From these it becomes clear that QCD must be T invariant to a very high degree. However, there is nothing within the Standard Model which would force the ^-parameter to be small; indeed, it is a free parameter lying in the range 0 < 6 < 2TT. The puzzle of why 6 ~ 0 in Nature is called the strong CP problem. One is tempted to resolve the issue with an easy remedy first. If QCD were the only ingredient in our theory, we could remove the strong CP problem by imposing an additional discrete symmetry on the QCD lagrangian, the discrete symmetry being CP itself. This wouldn't really explain anything but would at least reduce a continuous problem to a discrete choice. In reality, this will not work for the full Standard Model since, as we have seen, the electroweak sector inherently violates CP. It would thus be improper to impose CP invariance upon the full lagrangian. Moreover, even if one could set #bare = 0 in QCD, electroweak radiative corrections would generate a nonzero value. These turn out to
IX-5 The strong CP problem
253
occur only at high orders of perturbation theory, and are expected to be divergent by power counting arguments, although they have not been explicitly calculated. This divergence is not a fundamental problem because one could simply absorb #bare plus the divergence into a definition of a renormalized parameter 9ren, which could be inferred from experiment. However we are then back to an arbitrary value of 0ren and to the problem of why 9Ten is small. The parameter 0 The situation is actually worse than this in the full Standard Model, as the quark mass matrix can itself shift the value of 6 by an unknown amount. Recall that CP violation in the Standard Model arises from the Yukawa couplings between the Higgs doublet and the fermions. When the Higgs field picks up a vacuum expectation value, these couplings produce mass matrices for the quarks which are neither diagonal nor CP invariant. The mass matrices are diagonalized by separate left-handed and right-handed transformations, and CP violation is shifted to the weak mixing matrix. However, because different left-handed and right-handed rotations are generally required, one encounters an axial U(l) rotation in this transformation to the quark mass eigenstates, and as discussed in Sect. Ill—5, this produces a shift in the value of 0. Let us determine the magnitude of this shift. Denoting by primes the original quark basis, one has the transformation to mass eigenstates given by (cf. Eqs. 11-4.5,4.6) m = S* m'S* ,
tl>L = Slt//L,
1>R = S\tl/R .
(5.1)
Here, we have combined the u and d mass matrices into a single mass matrix. Expressing SL,R as products of U(l) and SU(N) factors, i
^
,
(5.2)
with SL, SR in SU(N), one obtains an axial U(l) transformation angle of (fR — (fL- From the discussion of Sect. Ill—5, this is seen to lead to a change in the 6 parameter, e^0
= O + 2Nf(
(5.3)
where Nf = 6 for the three-generation Standard Model. However, noting that the final mass matrix m is purely real, we have arg(det m) = 0 = arg (det 5^ det m' det SR)
= arg (det sty + arg (det m') + arg (det SR) = 2N (
arg (det m') ,
(5-4)
254
IX Kaon mixing and CP violation
where we have used the SU(N) property, detSn = det SL = 1. The resultant ^-parameter is then 0 = 0 + arg(detm')
,
(5.5)
with m' being the original nondiagonal mass matrix. The real strong CP problem is to understand why 6 is small. One possible solution to the strong CP problem occurs if one of the quark masses vanishes. In this case, the ability to shift 6 by an axial transformation would allow one to remove the effect of 0 by performing an axial phase transformation on the massless quark. Equivalently stated, any effect of 8 must vanish if any quark mass vanishes. Unfortunately, phenomenology does not favor this solution. The u quark is the lightest, but a value mu ^ 0 is favored. Connections with the neutron electric dipole moment The 6 term is not the source of the observed CP violation in K decays. This can be seen because it occurs in a AS = 0 operator, and while this may ultimately generate effects in AS = 1 processes, its influence is stronger in the AS = 0 sector. In particular, it generates an electric dipole moment de for the neutron. Since no such dipole moment has been detected, one can obtain a bound on the magnitude of 6. To determine the effect of 0, it is most convenient to use a chiral rotation to shift the 6 dependence back into the quark mass matrix. A small axial transformation produces the modified mass matrix md
\ ^+ir)^T^
j
= ^LM^R+^RM^L
, (5.6)
where rj is a small parameter proportional to 0 having units of mass, and T is a 3 x 3 hermitian matrix. Consistency requires T to be proportional to the unit matrix. If this were not the case, and instead we wrote T = 1 + \{Ti/2, the effective lagrangian would start out with a term linear in the meson fields, £eff ~ iV IV (rt/T - t/Tt) = 2-^- (T37ro + TsV8 + ...) ,
(5.7)
rather than the usual quadratic dependence. The vacuum would then be unstable because it could lower its energy by producing nonzero values of, say, the TTO field. Thus to incorporate ^-dependence without disturbing vacuum stability, one chooses T = 1. The act of rotating away any depen-
IX-5 The strong CP problem
255
dence on 9 produces a nonzero value of argdet M, and also determines 77, 9 = arg det M = arg \{mu + irj) (rrid + ir)) (ms + irj)] , mumdms r/ ~ 0 (for small 77) Tnvrid + Tnums + rridm such that the mass terms become TUUUU -
msSS
mdmdms
,_
-
_
,
(5.8) v '
(5.9)
murrid + mums The last term is the CP-violating operator of the QCD sector. Note that, as expected, 6 vanishes if any quark is massless. A nonzero neutron electric dipole moment de requires both the action of the above CP-odd operator and that of the electromagnetic current,
(5
*10)
where q — p'—p and we have inserted a complete set of intermediate states {/} in the neutron-to-neutron matrix element. For intermediate baryon states, the matrix elements of Tp^tp are dimensionless numbers of order unity and magnetic moment effects are of order the nucleon magneton, \in. Thus we find for de, de * 9
^ ^ - ^ , (5.11) murrid + mums + mdms AM where AM is some typical energy denominator. Using AM = 300 MeV, we obtain de~0x
10"15 e-cm .
(5.12)
Far more sophisticated methods have been used to calculate this, with results that unfortunately have a spread of a factor of 50 [Pe 89]. Our simple estimate is near the average. In explicit calculations, some subtlety is required because one must be sure that the evaluation correctily represents the U(1)A behavior of the theory [AbGLOPR 91]. However, the precise value is not too important; the significant fact is that bounds on d e ^10~ 2 5 e-cm requires 9 to be tiny, 0<1O~ 10. The strong CP problem does not have a good resolution within the Standard Model. It would appear that the abnormally small value of 0, and of the cosmological constant as well, are indications that more physics is required beyond that contained in the Standard Model.
256
IX Kaon mixing and CP violation Problems
1) CP violation and K-> Sn The Standard Model makes definite predictions for the existence of CP violation within the K —• 3TT system. One such effect is the existence of the decay Ks —• 7r +7r~7r° which one would expect to be forbidden in the limit of CP conservation. a) Actually this expectation is not quite correct. Calculate the amplitude for Ks —• 7T+7r~7r° in the limit of CP conservation and show, using the results of Prob. VIII-2, that
Thus there exists a CP-even piece of the amplitude, but it is small, arising from the A/ = | component of the hamiltonian, and vanishes at the center of the Dalitz plot. b) Now consider the CP-violating component of the amplitude. Using Prob. VIII-2 together with the fact that in the Standard Model, CP violation is purely A/ = \, show that + 7T-7T0
—
V2 c) Defining =
demonstrate that 77+-0 = ^?oo = € — 2ef as first given in [LiW 80]. In this approximation, any deviation of r/+_o — e from zero is small due to the same mechanism which suppresses r/+_ — e (This result can be modified by electroweak-penguin contributions, which have a different chiral structure, and higher order energy dependence, but the result remains discouragingly small). d) Why is experimental measurement of 7/+_o much more difficult than that of r/ + _? 2) Strangeness gauge invariance a) Physics must be invariant under a global strangeness transformation \M) —• exp(iA5)|M), where A is arbitrary. Explain why this is the case. b) Demonstrate that such a transformation has the effect ImMi2 ImMi2 ' ReMi 2 ~* ReM i 2 ~
Problems
257
as claimed in Eq. (2.9b), and that, while unphysical quantities such as e, £o are affected by such a change, physical parameters such as e, ef are.not. 3) Neutral kaon mass matrices and CPT invariance Some of the ideas discussed in this chapter can be addressed in terms of simple models of the neutral kaon mass matrix M which appears in Eq. (1.2). a) Consider the following CP-conserving parameterization as defined in the (If0, If0) basis: A m0 where A is real-valued. Determine the basis states (if_,if+) in which Mo —> M± becomes diagonal and obtain numerical values for 777-0,
A.
b) Working in the (if_,if + ) basis, extend the model of (a) to allow for CP violation by introducing a real-valued parameter <S, rri0
, , , _ / ra_ —id
0 \ 777-4- /
y 2(5
771-1-
and assume there is no direct CP violation. This mass matrix corresponds to the superweak (SW) model. By expressing M± in the (if0, K°) basis, use the analysis of Sects. IX-1,2 to predict (fe ' = phase e and determine 6 from the measured value of |e c) Finally, extend the model in (b) to M±"=(m; * V X * m+ where Re x is a T-conserving, CP-violating and CPT-violating parameter. Show that the states which diagonalize M±" are |*r5} ^
\K+)
-
||K_)
,
\KL) ~ \K.) + £\K+) , where V = {TRL — rns)/2 + iYs/^> Then evaluate r?+_ and 7/00, allowing for the presence of direct CP violation (i.e. e' ^ 0), and derive the following relation between phases,
The result \m^o —m Ko\/mKo < 5 x 10~18 which follows from this relation is the best existing limit on CPT invariance.
X The 1/N C expansion
The Nc * expansion is an attempt to create a perturbative framework for QCD where none exists otherwise. One extrapolates from the physical value for the number of colors, Nc = 3, to the limit iVc —> oo while scaling the QCD coupling constant so that g%Nc is kept fixed ['t H 74]. The amplitudes in the theory are then analyzed in powers of N'1. The hope is that the Nc —* oo world bears sufficient resemblance to the real world to yield significant dynamical insights. There is no magical process which makes the Nc —• oo theory analytically trivial; nonlinearities of the nonabelian gauge interactions are present, and the theory is still not solvable. However any consistent approximation scheme for QCD is welcome, and the large Nc expansion is especially useful for organizing one's thoughts in the analysis of hadronic processes. X—1 The nature of the large N c limit In passing from SU(3) to SU(NC), the quark and gluon representations, originally 3 and 8, become N c and N^ — 1 respectively. The analysis of Feynman graphs at large Nc is simplified by modifying the notation used to describe gluons. As usual, quarks carry a color label j , with ,7 = 1,2,..., Nc. Gluons can be described by two such labels, i.e. A;
-> A*
{A*, = 0) ,
(1.1)
where a = 1,..., N% — 1 and j , k = 1,..., Nc. In doing so, no approximation is being made. The new notation in simply an embodiment of the group product N c x N c —> (Nj? — 1) ^ 1. The quark-gluon coupling is then written
fi , 258
(1.2)
X-l
The nature of the large Nc limit
259
Fig. X-l. Double line notation: (a) quark and (b) gluon propagators, (c) quarkgluon, (d) three-gluon and (e) four-gluon vertices.
and the gluon propagator is
Jd'x e** (0 |r (A*j(x)Aj}(0)) I 0) = (1.3) The term proportional to Nc 1 must be present to ensure that the color singlet combination vanishes, A3- — 0. However, as long as we avoid the color singlet channel, this term will be suppressed in the large Nc limit and may be dropped when working to leading order. Using this new notation, the Feynman diagrams for propagators and vertices are displayed in Fig. X-l. A solid line is drawn for each color index, and each gluon is treated as if it were a quark-antiquark pair (as far as color is concerned). In this double line notation, certain rules which are obeyed by amplitudes to leading order in 1/NC emerge in an obvious manner. Although general topological arguments exist, we shall review these rules by examining the behavior of specific graphs. The power of Feynman diagrams to build intuition is rather compelling in this case. We consider first the familiar quark and gluon propagators. The quark propagator, unadorned by higher order corrections, is O(l) in the Nc —• oo limit. Fig. X-2 depicts two radiative corrections. Fig. X-2(a), the one-gluon loop, is O(l) in powers of Nc because the suppression from the squared coupling g\ is compensated for by the single closed loop, which corresponds to a sum over a free color index and thus contributes a factor of Nc. The graph then is of order g\Nc which is taken to be constant. The graph Fig. X-2(b) with overlapping gluon loops is O(N~2) because, with no free color loops, it is of order g% = (glNc) N~2 ~ N~2. The terms planar and nonplanar are used respectively to describe Figs. X-2(a),2(b), because the latter cannot be drawn in the plane without at least some internal lines crossing each other.
Fig. X-2. Radiative corrections to the quark propagator: (a) planar, (b) nonplanar.
260
X The 1/NC expansion
Four distinct contributions to the gluon propagator are exhibited in Fig. X-3. Figs. X-3(a),3(b) depict in double line notation the quarkant iquark and two-gluon loop contributions. It should be obvious from the above discussion that these are respectively O(N~1) and 0(1). A new diagram, involving the three-gluon coupling, appears in Fig. X-3(c). With three color-loops and six vertices, it is of order (glNc) = O(l). Figure X-3(d) is a nonplanar process with six vertices and one color sum, and is thus O{N~2). The discussion of the gluon propagator indicates why we constrain . to be fixed when taking the large Nc limit. The beta function of QCD is determined to leading order by Figs. X-3(a),3(b). If g\ were held fixed, the beta function would become infinite in the large Nc limit, leading to the immediate onset of asymptotic freedom. The choice g\Nc ~ constant leads to a running coupling constant and is compatible with the behavior for the realistic case of Nc = 3. To summarize, there are several rules which can be abstracted from examples such as these: (i) the leading order contributions are planar diagrams containing the minimum number of quark loops; (ii) each internal quark loop is suppressed by a factor of A^T1; and (iii) nonplanar diagrams are suppressed by factors of N~2. The suppressions in rules (ii), (iii) are combinatorial in origin. Quark loops and nonplanarities each limit the number of color-bearing intermediate states, and consequently cost factors of N~l. X—2 Spectroscopy in the large N
c
limit
In order for the large Nc limit to be relevant to the real world, it must be assumed that confinement of color singlet states continues to hold. In this case, we expect the particle spectrum to continue to be divided into mesons and baryons. Let us treat the mesons first, as the baryons are more problematic. One can form color singlet meson contributions from QQ pairs. To form a color singlet, one must sum over the quark colors. In order to produce a properly normalized QQ state one must therefore include a —
I/O
normalization factor of Nc ' into each QQ meson wavefunction, such
(a)
(b)
(c)
Fig. X-3 Various radiative corrections to the gluon propagator.
X-2 Spectroscopy in the large Nc limit
(a)
261
(b)
Fig. X-4 Mesons in the double line notation. that |Q(*)Q(0)) color
singlet
i^6( Q )t^)t|o) y/Nc
5
(2.1)
where a, /3 are flavor labels, i = 1,..., Nc is the color label, and tf (rf) are the quark (antiquark) creation operators. Meson propagators, as represented in Fig. X-4(a), are then 0(1) in Nc since the factors of (JV«T ' )2 from the normalization of the wavefunction are compensated by a factor of Nc from the quark loop. This leads to the prediction that meson masses are of (9(1) in the large Nc limit, i.e. they remain close to their physical values. Multiquark intermediate states, as in Fig. X-4(b), are suppressed by 1/NC, indicating a suppression of mixing between QQ and Q2Q2 sectors. That is, large Nc plus confinement implies the existence of QQ mesons which contain an arbitrary amount of glue in their wavefunction, but which do not mix with Q2Q2 states. What about the decay widths of QQ mesons? The decay amplitude is pictured in Fig. X-5 (other possibilities involve the suppressed quark loops). This diagram contains three meson wavefunctions and one quark loop and hence is of order (iVc~ ' )3-/Vc = Nc in amplitude or N~l in rate. The large Nc limit thus involves narrow resonances, i.e. T/M —> 0, where F is the meson decay width and M is the meson mass. This is reasonably similar to the real world, where most of the observed resonances have T/M ~ 0.1-0.2 [RPP 90]. Color singlet gluonic states, called glueballs, may also exist. The normalization of a glueball state can be fixed by means of the following argument. Suppose, as will be defined in a gauge invariant manner in Sect. XIII-4, that a neutral meson can be created from two gluons. Then in normalizing this configuration, one must sum over the Nc2 gluon color labels. As a consequence, a normalization factor A^"1 is associated with
Fig. X-5 Strong interaction decay of a QQ meson.
262
X The 1/NC expansion
(a)
(b)
Fig. X-6 Meson-meson scattering. each glueball state. Glueball propagators also emerge as being O(l). There is no physical distinction between two-gluon states, three-gluon states, etc., because all are mixed with each other by the strong interaction. As a result, there need not be any simple association between a specific physical state and gluon number, and thus the concept of a 'constituent gluon' need not be inferred. In glueball decays, however, one must distinguish between glueballs decaying to other glueballs, and those decaying to QQ mesons. Where kinematically allowed, the decay of glueballs to glueballs is 0(1), while that to QQ states is O (1/NC). The lowest lying glueball(s) will then be narrow, while those above the threshold for decay into two glueballs will be of standard, nonsuppressed width. Meson-meson scattering amplitudes are also restricted by large Nc counting rules. Consider the diagrams of Fig. X-6. That of Fig. X-6(a) is of order (N~1/2)4NC ~ TV"1, whereas Fig. X-6(b) is O(N~2) because of the extra quark loop. The scattering amplitudes thus vanish in the large Nc limit, and the leading contributions are connected, planar diagrams. The large Nc limit also predicts that neutral mesons (i.e. Q^Q^ composites with a = j3) do not mix with each other. The possible mixing diagram is given in Fig. X-7, and includes any number of gluons. However, because of the extra quark loop, it is of order N~l, and thus vanishes in the infinite color limit. This means that uu states do not mix with dd or 55, nor do the latter two mix. The large Nc spectrum thus displays a nonet structure with the uu and dd states degenerate (to the extent that electromagnetism and the mu-md mass difference are neglected) and the ss states somewhat heavier. This pattern is reflected in Nature, except that the uu and dd configurations now appear as states of definite isospin, uu ± dd. For example, let us consider the JPC — 1 L 2 + + mesons. For the former, p(770) and a;(783) are interpreted as uu, dd isospin 1 = 1 and
Fig. X-7 Meson-meson mixing.
X-3 Goldstone bosons and the axial anomaly
263
7 = 0 combinations, while y>(1020) is the ss member of the nonet. Including the if*(892) doublet as the us, ds combinations, a simple additivity in the quark mass would imply m
v?(i020) -
m
p(770) = 2 (m#*(892) - m p ( 7 7 0 )) ,
(2.2)
+
which works well. A similar treatment of the 2 + mesons, identifying a2(1320) and /2(1270) as the corresponding uu, dd states and /2(1525) as an ss composite, predicts -
m
a2a(1320) 2(1320) =
22
( mm/q(1430) ~
mm
a2(1320)
which is also approximately satisfied. The fact that p(770), a;(783), /2(1270) and a2(1320) decay primarily to pions, and
(Pj(q)\dllAU0)\0) = Fjmfak .
(3.2)
For the octet of currents, the divergence vanishes for zero quark mass, and as usual leads to the identification of TT, K, rjs as Goldstone bosons. However, for the singlet current the anomaly is present. Even in the
X The 1/NC expansion
264
limit of vanishing quark mass, the current divergence has nonzero matrix elements, in particular, (3.3)
If one repeats the calculation of the anomalous triangle diagram as in Sect. Ill—3 but now allows N c to be arbitrary, one sees that it is proportional to Tr (AaAfe) = 26ab and is therefore independent of Nc. However, by using large Nc counting rules, the matrix element in Eq. (3.3) is seen to be of order g\Nc ~ Nc .* This implies that the gluonic contribution to the axial anomaly vanishes in the large Nc limit. When we take into account the behavior of jfy, we conclude that m2, ~ 1/NC —> 0. The rf is thus massless in the large Nc limit, and we end up with a nonet of Goldstone bosons. To illustrate what happens when the number of colors is treated perturbatively, let us consider the 1/NC corrections to the meson spectrum together with the effects of quark masses. If we first add quark masses, we have, in analogy with the results of Sect. VII-1, the mass matrix \m(uu + dd) + msss\
(3.4)
where we have taken mu = rrid =TO.This leads to a squared-mass matrix (2m ins
0
0
0
0
0
+m
0
0
0
\
(3.5)
§(2m8 \{ rns
0
in the basis (TT, if, r)g,r)o). If this were diagonalized, one would find an isoscalar state degenerate with the pion. This is a manifestation of the U(l) problem which arises when there is no anomaly. However, at the next order in large Nc, the matrix picks up an extra contribution in the SU(3) singlet channel due to the anomaly, yielding /2m 2
m = B0
0
0 ms
i
I
m
0
0
0
0
0
0
| (2ms + m)
\ 0
0
mm-ms)
{m-
\(
\
(3.6)
2rh)
This result depends on the assumption that topologically nontrivial aspects of vacuum structure are smooth in the 7VC —> oo limit.
X-4
The OZI rule
265
where e = O(N®). This mass matrix yields an interesting prediction. The quantities Born and Boms are fixed as usual by using the n and K masses. Also the trace of the full matrix must yield rn^+rn\+mi+m?,, which fixes e = 2.16GeV2. The remaining diagonalization then predicts m^ = 0.98 GeV, 77?^ = 0.50 GeV with a mixing angle of 18°. This is a remarkably accurate representation of the situation in the real world. Although e/N c is suppressed in a technical sense, note how sizeable it actually is. One is hard pressed to imagine any sense in which the physical rj mass can be taken as a small parameter. X - 4 The OZI rule In the 1960's, an empirical property, called the Okubo-Zweig-Iizuka (OZI) rule [Ok 63, Zw 65, Ii 66], was developed for mesonic coupling constants. Its usual statement is that flavor disconnected processes are suppressed compared to those in which quark lines are connected. In the language which we are using here, flavor disconnected processes are those with an extra quark loop. Unfortunately, the phenomenological and theoretical status of this so-called rule is ambiguous. We briefly describe it here because it is part of the common lore of particle physics. The empirical motivation for the OZI rule is best formulated in the decays of mesons. Let us accept that <^(1020) and /2(1525) are primarily states with content ss whereas a;(783) and /2(1270) have content (uu + dd) / \ / 2 . Mixing between the ss and nonstrange components can take place with a small mixing angle, such that
Amp (ss) Amp (\uu + dd)/\/2)
EE
tan 0 ,
(4.1)
with 9 = 9y for the vector mesons and 0 = 6T for the tensor mesons. In both cases, 9 is small. Experimentally, the
= 0.012 ± 0.002 ~ 0.003 x p.s. ,
E^r+p.^
Y ^ = °-004 ± °-001
(42)
- 0-002 x p.s. .
This suggests the hypothesis 'ss states do not decay into final states not containing strange quarks'. Diagrammatically this leads to a pictorial representation of the OZI rule, viz., the dominance of Fig. X-8(a) over Fig. X-8(b). Some scattering processes also show such a suppression. For
266
X The 1/NC expansion
example, we have **-*<"
~ 0.03 ,
(4.3)
which can be interpreted as an OZI suppression. A stronger version of the OZI rule would have the ip/u and /2//2 ratios equal to a universal factor of tan 2 0 (cf. Eq. (4.1)) once kinematic phase space factors are extracted. The narrow widths of the J/ip and T states are also cited as evidence for the OZI rule, since these hadronic decays involve the annihilation of the cc or bb constituents. This can be correct almost as a matter of definition, but it is not very enlightening. Indeed, the small widths of heavy-quark states can be understood within the framework of perturbative QCD without invoking any extra dynamical assumptions. However, perturbative QCD certainly cannot explain the OZI rule in light mesons. It must have a different explanation for these states. There actually exist several empirical indications counter to the OZI rule [Li 84, E1GK 89, RPP 90]. Among the more dramatic examples of OZ/-forbidden reactions, expressed as ratios, are =1.2 ± 0 . 5 , 0 2 3
+0.14
a
i™*+*
=
2.0 ± 0 . 7 ,
0TP+PVK+*-
The universal-mixing model is incorrect more often than not, with counterexamples being = o. 1O ± 0.02 ,
(4.5) _ +0.011 — u.uzy instead of the values 0.03, 0.03 and 0.006 expected from the previous ratios. The empirical 77-7/ mixing angle 6^-^ ~ —20° also violates the OZI rule, which would require a mixing angle of —35°. l v
(a)
^
(b)
Fig. X-8 OZI (a) allowed, (b) forbidden amplitudes.
X-5 Chiral lagrangians
267
There is also an intrinsic logical flaw with the simplest formulation of the OZI rule. This is because OZJ-forbidden processes can take place as the product of two OZ/-allowed processes. For example, each of the following transitions is OZ/-allowed: (4 6)
fi - * VV J
777? - • TTTT .
Hence the OZ/-forbidden reaction f2 —• TTTT can take place by the chains f2 —• KK
- > 7T7T ,
f2 - > 7?77 - > 7T7T .
(4.7)
These two-step processes are in fact required by unitarity to the extent that the individual scattering amplitudes are nonzero. The large Nc limit provides the only known dynamical explanation of the OZI rule at low energies. Although the gluonic coupling constant is not small at these scales and suppressed diagrams have ample energy to proceed, they are predicted to be of order 1/N% in rate because of the extra quark loop. Yet large Nc arguments need not suggest a universal suppression factor of tan 2 0, because there is no need for the 1/NC corrections to be universal. Note that the large Nc framework also forbids the mixing of 77 and 7/ and more generally, the scattering of mesons. Thus, the OZI rule in light-meson systems remains somewhat heuristic. It has a partial justification in large Nc counting rules, but it also has known violations. It is not possible to predict with certainty whether it will work in any given new application. X—5 Chiral lagrangians The large Nc limit places restrictions on the structure of chiral lagrangians [GaL 85a]. To describe these, we must first allow for an enlarged number Nf > 3 of quark flavors. The three-flavor O(E4) lagrangian is expanded as 10
A = $>A
,
(5.1)
where the {Oi} can be read off from Eq. (VI-2.7). Recall that in constructing £4, we removed the O(E4) operator Oo = Tr (D^UD^D^UD^U^
,
(5.2)
because for Nf — 3 it is expressible (cf. Eq. (VI-2.3)) as a linear combination of Oi52,3- However, if the number of flavors exceeds three, one must append Oo to the lagrangian of Eq. (5.1), 10
3
10
aiOi .
(5.3)
268
X The 1/NC expansion
In view of the linear dependence of Oo on Oi,2,3, note that we have needed to modify the coefficients 0:1,2,3 —• /?i,2,3- Upon returning to three flavors, we regain the original coefficients, ai = y + A ,
<*2 = /% + /%,
a 3 = - 2 # ) + /%.
(5.4)
We can now study the large Nc behavior of the extended O(E4) chiral lagrangian. The distinguishing feature is the number of traces in a given O(E4) operator. Each such trace is taken over flavor indices and amounts to a sum over the quark flavors, which in turn can arise only in a quark loop. In particular, those operators with two flavor traces (Oi,2,4,6,7) will require at least two quark loops, while those with one flavor trace need only one quark loop. However, our study of the large Nc limit has taught us that every quark loop leads to a 1/NC suppression. Thus the O(E*) chiral contributions having two traces will be suppressed relative to those with one trace by a power of 1/NC, and provided Ps ^ 0 we can write* Pi
02
OL±
C*6
(xti\
n
^
(
a)
Alternatively, this 7Vc-counting rule implies (provided Po/Ps 7^ 1/2) for the {cti} coefficients of flavor 51/(3), () . (5.5b) as as as The overall power of Nc for the remaining terms can be found by noting that the TTTT scattering amplitude should be of order A^T1, implying ai,2,3 = O(NC). The only exception to the above counting behavior is the operator with coefficient 0:7. This exception occurs because the operator can be generated by an 7/-pole, and the 7/ mass-squared is O(l/Nc). In particular, the coefficient of this term is absolutely predicted in the large iVc limit. This follows if we include the large Nc result for ?TIQ8 from Eq. (3.6) in the 77-7/ mixing analysis of Prob. VII-5. The result of the mixing is
ifif
(5 6)
-
It is the factor of ra~,2 which overcomes the counting rules. Although the double trace suggests that this operator is suppressed in the large iVc limit, we have m~? oc Nc. Thus, at least formally, an extra enhancement would be predicted. The operator Oj presents a special case and is discussed below.
X-6
Weak nonleptonic decays
269
The large Nc limit then predicts the following ordering of the chiral coefficients in £4: a7 = O(N2C) ,
ai, a 2 , a 3 , a 5 , a 8 , a 9 , ai 0 = O{NC) ,
(5.7)
- a 2 , ^4,^6 = C?(1) • We have built these properties into the coefficients appearing in Sect. VI2. The only existing empirical test involves the occurrence of 2OL\ — OLj Nc—>OO yj I
,
(6.1)
where each current in the above is a color singlet. Some added insight can be obtained by considering directly the K —• 2?r nonpenguin diagrams, such as those in Fig. X-9. There is a modification in the large Nc rules when two weak currents interacting via VK±-exchange are involved. Because the currents and the W± are color singlets whose couplings do not involve #3, we see that an extra quark loop connected
n d
(a) Fig. X-9
K —• 7T7T amplitudes to leading N~l order.
270
X The 1/NC expansion
by W± gains an extra factor of Nc from the color sum. Thus Fig. X-9(a) is a factor of Nc larger than Fig. X-9(b). Next consider the case where gluons are exchanged across the interaction vertex as in Fig. X-10(a). The single-gluon exchange amplitude vanishes since a gluon belongs to a color octet whereas both weak currents and mesons are color singlets. Two-gluon amplitudes are expressed in double line notation in Fig. X-10(b). Application of the counting rules to these diagrams indicates they are of O(N~1) compared to those in Fig. X-9(a). Thus in the Nc —• oo limit we are left with only the vacuum saturation diagram of Fig. X-9(a). This diagrammatic exercise explains the nonrenormalization of HwTo leading order, the only gluonic diagrams are those which act on a single current and do not act across a vertex. In QED, the single current vertex is not changed in normalization as a consequence of the Ward identity. Here, one gets more in that the absolute normalization of the kaon amplitudes is also calculable. One finds 76)«| 0)
- ml) ~ 9.8 x l^mK
+ 75)*|
,
=0, 75)«| 0)
^
K°)
(6.2) (6.3) K~)
(6.4)
These results bear no relation to reality either in absolute magnitude or in relative rates (e.g., see Eq. (VIII-4.2)). There is certainly no A/ = 1/2 rule here, because the ratio A2/A0 = l/\/2 implied by the above is far larger than the measured value, A2/A0 ~ 1/22. Since the large Nc limit fails, the origin of the A/ = 1/2 rule must lie beyond leading order in Nc. Unfortunately there is no unique prediction for the next-to-leading corrections, although some nonleading aspects which enhance A/ = 1/2 have been identified. For example, the QCD short distance corrections provide a partial enhancement, and the purely
(a) Fig. X-10
(b) Nonleading K —• TTTT contributions.
Problems
271
A / = 1/2 penguin graphs enter at the first nonleading order in Nc. There has been an interesting attempt [BaBG 87] to identify a set of nonleading corrections in 1/NC and to combine them to obtain the A / = 1/2 rule for kaons. This method involves the use of the large Nc chiral lagrangian at low energy and QCD perturbation theory at high energy. Nonleading corrections enter when one considers one-loop corrections, employing meson loops or quark/gluon loops in the respective domains. An interesting innovation is a serious attempt to match the scale dependence of the QCD coefficients in the weak nonleptonic hamiltonian with the scale dependence of the low energy matrix elements. If matrix elements and coefficient functions are computed with reference to an energy scale //, the key requirement is that the physical nonleptonic amplitude,
Aonlept. = I ^ K d K s J2 Ci(l*)(™\Oi\K) = J2 *(»)Mi(n)
, (6.5)
be /i independent. In [BaBG 87], the scale dependence in the matrix elements Mi(fi) arises when one calculates chiral loops with a high energy cutoff A. The identification \x — A allows a partial cancelation of the scale dependence. Although the results are promising, a full resolution will require a more complete understanding of the relevant hadronic physics at low energy. Problems 1) The large Nc weak hamiltonian Retrace the calculation of the QCD renormalization of the weak nonleptonic hamiltonian described in Sect. VIII-3, but now in the limit Nc —• oo with g\Nc fixed. Show that the penguin operators do not enter and that all short distance effects are of order A^T1, with the operator product coefficients c\ = 1,C2 = 1/5, C3 = 2/15, C4 = 2/3, C5 = ce = O. 2) The strong CP problem in the large N c limit In the large Nc limit, the 770 can be united with the Goldstone octet in the effective lagrangian. Generalizing the chiral matrix to nine fields we write C = Co + CN-i, where Co = ^ Tr (dpUdfU^ + ^-Bo Tr (m(U +
U = exp (iX • tp/F) exp \i\
-—
272
X The 1/NC expansion
a) Confirm that this reproduces the mixing matrix of Eq. (3.6). b) Another way to obtain this result is to employ an auxiliary pseudoscalar field q(x) (with no kinetic energy term) to rewrite £N-i as
Identify the SU(3) singlet axial current and calculate its divergence to show that q(x) plays the same role as FF, i.e. q(x) ~ aFF/87r. Integrate out q(x) to show that this is equivalent to the form of part (a). c) Several authors [RoST 80, DiV 80] suggest adding the 0-term through C = Co + CN-i - 6q{x) . From this starting point, integrate out q(x) and show that a chiral rotation can transfer 9 to argdet m. However, in the sense described in Sect. IX-5, this theory is unstable about U = 1. The stable vacuum corresponds to U^ = 8^exp(i(3j). For small 0, solve for fij in terms of 6. d) Using U — elP/2Ue%P/2, define the fields about the correct vacuum to find the CP-violating terms of the form
identifying a and b and showing they vanish if any quark mass vanishes. Calculate the CP-violating amplitude for rj —> TT+TT".
XI Phenomenological models
QCD has turned out to be a theory of such subtlety and difficulty that a concerted effort over an extended period has not yielded a practical procedure for obtaining analytic solutions. At the same time, vast amounts of hadronic data which require theoretical analysis and interpretation have been collected. This has spurred the development of accessible phenomenological methods. We devote this chapter to a discussion of three dynamical models (potential, bag, and Skyrme) along with a methodology based on sum rules. Although the dynamical models are constructed to mimic aspects of QCD, none of them is QCD. That is, none contains a rigorous program of successive approximations which, for arbitrary quark mass, can be carried out to arbitrary accuracy. Therefore, our treatment will emphasize issues of basic structure rather than details of numerical fits. By using all of these methods, one hopes to gain physical insight into the nature of hadron dynamics. Despite its inherent limitations the program of model building, fortified by the use of sum rules, has been generally successful, and there is now a reasonable understanding of many aspects of hadron spectroscopy. XI—1 Quantum numbers
of QQ and Q3 states
Among the states conjectured to lie in the spectrum of the QCD hamiltonian are mesons, baryons, glueballs, hybrids, dibaryons, etc. However, since practically all currently known hadrons can be classified as either QQ states (mesons) or Q3 states (baryons), it makes sense to focus on just these systems. We shall begin by determining the quark model construction of the light hadron ground states. Much of the material will be valid for heavy-quark systems as well. 273
274
XI Phenomenological models Hadronic flavor-spin state vectors
In many respects, the language of quantum field theory provides a simple andflexibleformat for implementing the quark model. Let us assume that for any given dynamical model, it is possible to solve the field equations of motion and obtain a complete set of spatial wavefunctions, {*/ja(x)} for quarks and {^{x)} for antiquarks, where the labels a and a refer to a complete set of observables. A quark field operator can then be expanded in terms of these wavefunctions, (a)] ,
(1.1)
where a;Q, uJa are the energy eigenvalues, b(a) destroys a quark and eft (a) creates the corresponding antiquark. The quark creation and annihilation operators obey
{b(a),b(a')} = 0 ,
{&(<*),
{d(a),d(ct)} = 0 ,
(1.2)
=0 ,
which are the usual anticommutation relations for fermions. In all practical quark models, an assumption is made which greatly simplifies subsequent steps in the analysis, that the spatial, spin, and color degrees of freedom factorize, at least in lowest order approximation. This is true provided the zeroth order hamiltonian is spin-independent and color-independent. Spin dependent interactions are then taken into account as perturbations. This assumption allows us to write the sets {a} and {a} in terms of the spatial (n), spin (s, ras), flavor (g), and color (fc) degrees of freedom respectively, i.e. a = (n, 5,m s ,g, k). If we are concerned with just the ground state, we can suppress the quantum number n, and for simplicity replace the symbolsfe,d)', etc. for annihilation and creation operators with the flavor symbol q (q = u,d,s for the light hadrons),
Hadrons are constructed in the Fock space defined by the creation operators for quarks and antiquarks. Light hadrons are labeled by the spin (S2,53), isospin (T2,T3), and hypercharge (Y) operators as well as by the baryon number (B). Other observables like the electric charge Qe\
XI-1 Quantum numbers of QQ and Q3 states
275
-•-T,
3
8
3*
10
Fig. XI-1 Some SU(3) flavor representations.
and strangeness S are related to these, Qel = T3 + Y/2 ,
S=
Y-B
(1.4)
Since quarks have spin one-half, the baryon (Q3) and meson (QQ) configurations can carry the spin quantum numbers S = 1/2,3/2 and S = 0,1 respectively. If we neglect the mass difference between strange and nonstrange quarks, then flavor SU(3) is a symmetry of the theory, and both quarks and hadrons occupy SU(3) multiplets. The quarks are assigned to the triplet representation 3 and the antiquarks to 3*. The QQ and Q 3 constructions then involve the group products 3 x 3* = 8 0 1 , (3 x 3) x 3 = (6 0 3*) x 3 = 10 © 8 0 8 0 1
(1.5)
so that baryons appear as decuplets, octets, and singlets whereas mesons appear as octets and singlets. The 577(3) flavor representations 3, 3*, 8, 10 are depicted in Y vs. T3 plots in Fig. XI-1. The circle around the origin for the eight-dimensional representation denotes the presence of two states with identical Y, /3 values. Finally, quarks and antiquarks transform as triplets and antitriplets of the color SU(3) gauge group, and all baryons and mesons are color singlets. Two simple states to construct are the p\ meson and the ££% baryon, lA3+/2> = «' where the superscript and subscript on the hadrons denote electric charge and spin-component, and a summation over color indices for the creation operators is implied. The normalization constants are fixed by requiring that the hadrons {Hn} form an orthonormal set, (Hm\Hn) = 8mn. The other ground state hadrons can be reached from those in Eq. (1.6) by means of ladder operations in the spin and flavor variables. In this manner, one can construct the flavor-spin-color state vectors of the 0~ octet
276
XI Phenomenological models Table XI-1. State vectors of the pseudoscalar octet and singlet mesons
^6 [ « - « ]
-«
7
10)
+ 44 + 44 - 44i
_
7b W - 44 + 44 - 44
44 4 - 44
and singlet mesons and the \ octet baryons displayed in Tables XI-1 and XI-2. A convenient notation for fields which transform as SU(3) octets involves the use of a cartesian basis rather than the 'spherical' basis of Tables XI-1,2. In fact, we have already encountered this description in Sect. VII-1 during our discussion of SU(3) Goldstone bosons where the quantity U = exp(iip • A) played a central role. The eight cartesian fields
Table XI-2. State vectors of baryon spin-1/2 octet
1°)
XI-1 Quantum numbers of QQ and Q3 states
277
{(fa} are related to the usual pseudoscalar fields by 4-
1 /
\
n
(1.7) 1
/
•
N
-={tpa-vpi),
T70
K
which is an alternative way for stating the content of Eq. (VII-2.4). The physical spin one-half baryons p, n , . . . can likewise be expressed in terms of an octet of states {Bi} (i — 1,..., 8) in cartesian basis as ± = - ^ ( B i =F iB2) , ^
E° = Bz , ^
~°= -j=(B6 + iB7),
A= 58 , ,
(1.8)
~- = -L(
In the quark model, hadron observables have simple interpretations, e.g. the baryon number is simply one-third the difference in the number of quarks and antiquarks, etc. Thus, writing quark and antiquark number operators as N(q) and N(q) for a quark flavor g, we have B = [N(u) + N(d) + N(s) - N(u) - N(d) - N(s)}/3 , T3 = [N(u)-N(d)-N(u) Y = [N(u) + N(d) - 2N(s) - N(u) - N(d) + 2N(s)]/3 , Qei = [2N(u) - N(d) - N{s) - 2N(u) + N(d) + N(s)]/3 , and the hadronic spin operator is
Quark spatial wavefunctions Many applications of the quark model require the knowledge of the quark spatial wavefunctions within hadrons. It is here that the greatest variation in the different models can occur, but several general features still remain. Indeed, in many instances it is the general features that are primarily tested. For example, the ground state in all models is a spatially symmetric S-state in which the wavefunction peaks at r = 0. The normalization
278
XI Phenomenological models
0.020
-
0.015
-
0.010
~
0.005
-
0.0
0.2
0.4
0.6
0.8
1.0
1.2
r(fm)
Fig. XI-2 Quark probability density in the bag and oscillator models condition of the quark spatial wavefunction, /
dsx
ensures that the magnitude of t/> will be similar in those models having wavefunctions of comparable spatial extent. This accounts for the agreement which can be found among diverse quark models in specific applications. How does one fix the spatial extent? One approach is to use an observable like the hadronic electromagnetic charge radius, e.g., = 0-87 ± 0.02 fin
= 0-66 ± 0.02 fin
(1.12)
Viewed this way, the bound states are seen to define a scale of order 1 fm. For example, we display two models in Fig. XI-2, the oscillator result with a2 = 0.17 GeV2 and the bag profile, which are each obtained by fitting to ground state baryon observables like the charge radius. Not surprisingly their behaviors are quite similar. Also shown in Fig. XI-2 is an oscillator model wavefunction whose parameter (a2 = 0.049 GeV2) was determined by using data from decays of excited hadrons. The difference is rather striking, and serves to demonstrate that the most important general feature in setting the scale in quark model predictions of dimensional matrix elements is the spatial extent of the wavefunction.* Another aspect of quark wavefunctions involves the issue of relativistic motion. A relativistic quark moving in a spin-independent central * We could obtain a bag result which behaves similarly by employing a charge radius of 0.5 fm rather than the 1 fm value shown.
XI-1 Quantum numbers of QQ and Q 3 states
279
potential has a ground state wavefunction of the form i u(r)x where u, £ signify 'upper' and 'lower' components. As we shall see, in the bag model these radial wavefunctions are just spherical Bessel functions. The above form also appears in some relativized harmonic oscillator models which use a central potential. If we allow for relativistic motion, then the major remaining difference in the quark wavefunctions concerns the lower two components of the Dirac wavefunction. Nonrelativistic models automatically set these equal to zero, while relativistic models can have sizeable lower components. Which description is the correct one? Quark motion in light hadrons must be at least somewhat relativistic since quarks confined to a region of radius R have a momentum given by the uncertainty principle,* p > V3R-1 ~ 342 MeV
(for R ~ 1 fin) .
(1.14)
Since this momentum is comparable to or larger than all the light-quark masses, relativistic effects are unavoidable. A more direct indication of the relativistic nature of quark motion comes from the hadron spectrum. Nonrelativistic systems are characterized by excitation energies which are small compared to the constituent masses. In the hadron spectrum, typical excitation energies lie in the range 300-500 MeV, again comparable to or larger than light-quark masses. Such considerations have motivated relativistic formulations of the quark model. Interpolating fields In the LSZ procedure for analyzing scattering amplitudes the central role is played by interpolating fields. These are the quantities which experience the dynamics of the theory in the course of evolving between the asymptotic instates and out-states. They turn out to be also useful as a kind of bookkeeping device. That is, one way to characterize the spectrum of observed states is to use operators made of appropriate combinations of quark fields ip(x). For example, corresponding to the meson sector of QQ states, one could employ a sequence of quark bilinears, the simplest of which are r/;ip, ^75^, ^7^> >7/i75V>, ^
(1.15)
Any of these operators acting on the vacuum creates states with its own quantum numbers. The lightest states in the quark spectrum will be The <s/3 factor is associated with the fact that there are three dimensions.
280
XI Phenomenological models
associated with those operators which remain nonzero for static quarks, i. e. with creation operators and Dirac spinors of the form
Only the pseudoscalar operators ^75 T/>, ^7075 ^ a n d the vector operators ifr'jiipiipcroiip a r e nonvanishing in this limit. All the other operators have a nonrelativistic reduction proportional to spatial momentum, indicating the need for a unit of orbital angular momentum in forming a state. The interpolating field approach is particularly useful in situations where the imposition of gauge invariance determines whether a given field configuration can occur in the physical spectrum. We shall return to this point in Sect. XIII-4 in the course of discussing glueball states. We now turn to a summary, carried throughout the next three sections, of various attempts to model the dynamics of light hadronic states. XI—2 Potential model The potential model posits that there is a relatively simple effective theory in which the quarks move nonrelativistically within hadrons. In the light of our previous comments on relativistic motion, this would seem to be acceptable only for truly massive quarks like the fr-quark and certainly questionable for the light quarks ix, d, s. However, in the potential model it is assumed that QCD interactions dress each quark with a cloud of virtual gluons and quark-antiquark pairs, and that the resulting dynamical mass contribution is so large that quarks move nonrelativistically. These 'dressed' degrees of freedom are called constituent quarks, and their masses are called 'constituent masses'. Constituent masses are not to be directly identified with the mass parameters occurring in the QCD lagrangian.* Energy levels and wavefunctions are then obtained by solving the nonrelativistic Schrodinger equation in terms of the constituent masses and some assumed potential energy function. The potential model is not without flaws. For light-quark dynamics, it is far from clear that a static potential can adequately describe the QCD interaction. Even with the use of constituent masses, one finds from fits to the mass spectrum and/or the charge radius that quark velocity is nevertheless near the speed of light (c/. Prob. XI-1). Also, although it is possible [LeOPR 85] to make a connection between the lightest pseudoscalar mesons as Goldstone bosons on the one hand and QQ composites on the other, this is not ordinarily done. Such criticisms notwithstanding, * We shall continue to denote the QCD mass parameter of quark qi as rrii, and shall write the corresponding constituent mass as Mi.
XI-2 Potential model
281
the nonrelativistic quark model does provide a framework for describing both ground and excited hadronic states, and brings a measure of order to a spectrum containing hundreds of observed levels. Besides, virtually all physicists are familiar with the Schrodinger equation, and find the potential model to be an understandable and intuitive language.
Basic ingredients One begins by expressing the mass Ma of a hadronic state a as Ma = ^
Mi + Ea ,
(2.1)
where the sum is over the constituent quarks and antiquarks in a. The internal energy Ea is an eigenvalue of the Schrodinger equation = EQi>a
,
(2.2)
with hamiltonian H
= E ^ P * 2 + E ^colorM ,
(2.3)
where r^ = r^ — r^, and the subscript 'color' on the potential energy indicates that the dynamics of quarks necessarily involves the color degree of freedom in some manner. It is standard to assume that the potential energy is a sum of two-body interactions. Although there exists no unique specification of the interquark potential V^oior from QCD, the following features are often adopted: 1) 2) 3) 4)
a spin and flavor independent long range confining potential, a spin and flavor dependent short range potential, basis mixing in the baryon and meson sectors, and relativistic corrections.
We shall discuss specific models of the potential energy function in Sect. XIII-1. They all have in common the color dependence in which the two-particle potential is twice as strong in mesons as it is in baryons,
{
V{rij) 2 V\Tij)
(mesons) , (baryons) .
We shall describe a simple empirical test for such behavior at the end of this section. To appreciate its theoretical basis, note that the quarkantiquark pair in a meson must occur in the 1 representation of color, whereas any two quarks in a baryon must be in a 3* representation (in
282
XI Phenomenological models
order that the three-quark composite be a color singlet), (F(3).F(r) (mesons), oc j F ( 3 ) F ( 3 ) (baryong)
(2.5)
(F 2 (l) - F 2 (3) - F 2 (3*))/2 = - 4 / 3 (mesons) , 2 2 (F (3*) - 2F (3))/2 = - 2 / 3 (baryons) , a where F (R) is a color generator for £J7(3) representation R. Thus, the color dependence in Eq. (2.4) is that which one would naturally associate with the interaction between two quarks or a quark-antiquark pair. Mesons For the two-particle QQ system, it is straightforward to remove the center-of-mass dependence. In the center-of-mass frame the Schrodinger equation becomes /p 2 \ [ _£_ _i_ V(r) I tb (r) — E ib (r)
(2 6)
where r = TQ — TQ and M" 1 = MQ1 + MQ1 is the inverse reduced mass. The LS coupling scheme is typically employed to classify the eigenfunctions of this problem. One constructs the total QQ spin, S = SQ + SQ, and adds the orbital angular momentum L to form the total angular momentum J = S + L. There is an infinite tower of eigenstates, each labeled by the radial quantum number n and the angular momentum quantum numbers J, J 2 , L, S. The QQ states are sometimes described in terms of spectroscopic notation 2S+lLj (JPC), where P is the parity and C is the charge conjugation, P = (_)£+i ,
C = (-)L+5 .
(2.7)
Strictly speaking, although only electrically neutral particles like TT°can be eigenstates of the charge conjugation operation, C is often employed as a label for an entire isomultiplet, like n = (TT+, TT0, TT~). The lowest QQ Table XI-3. Quantum numbers of QQ composites. Singlet
Triplet
0 2 3
1
ZJ 2 (2-+) ^3(3+-)
3 3
r>1|2f3(l—,2—,3—) F 2 , 3 ,4(2++,3++,4++)
XI-2 Potential model
283
orbital configurations, expressed in 2S+1Lj (JPC) notation, are displayed in Table XI-3. The 0 + , 1~, 2 + , . . . series of Jp states is called natural, and has the same quantum numbers as would occur for two spinless mesons of a common intrinsic parity. The alternate sequence, 0~, 1 + , 2~,... is referred to as unnatural. There are a number of Jpc configurations, called exotic whichstates, cannot be accommodated within the QQ construction. For example, the 0 state is exoticbecause any state with J = 0 must have L = 5, and according to the QQ constraint of Eq. (2.7) must therefore carry C = +. Likewise, the CP = — 1 sequence 0 + ~, 1~+, 2 + ~ , . . . is forbidden because the QQ model requires CP = (—)^ +1, implying 5 = 0 and hence J = L. Thus one would obtain P = (—) J+1 in the QQ model and not P = ( - ) J . Baryons Most applications of the quark model for Q 3 baryons involve the light quarks. If, for simplicity, we assume degenerate constituent mass M, the Schrodinger equation is
where the prefactor of 1/2 in the potential energy term follows from Eq. (2.4). It is convenient to define a center-of-mass coordinate R and internal coordinates A and Q by R = (n + r 2 + r 3 )/3 , Q = (n - v2)/V2 ,
(2.9)
A = (n + r2 - 2r3)/V6 . Because it is not possible to remove the three-particle center-of-mass dependence for an arbitrary potential, the following approach is often followed [GrS 76, IsK 78]. The potential V(r^) is rewritten as V(rlj) = Vosc(rij) + U(rij)
,
(2.10)
where VOBC = ^ r i ] ,
U = V-VOSC
.
(2.11)
The Schrodinger equation is solved in terms of the oscillator potential and U is evaluated perturbatively in the oscillator basis. Having removed the center-of-mass coordinate, we are left with the following hamiltonian for the internal energy: Hint
{
+
Q)
+
(
+
A )
(212)
284
XI Phenomenological models
which is just that of two independent quantum oscillators each with spring constant 3k. For later purposes, we write the number of excitation quanta for the two oscillators as Np and N\ (Np,\ = 0,1,2,...) and let TV = Np + N\. The angular momentum for the three-quark system is found in a similar manner as for the QQ mesons, J = L + S. The total quark spin is S = Yl Sj, the orbital angular momentum is given by L = L p + Lj\, and the parity is P = (—)^>+^. The ground state wavefunction has the form v2\3/2
e i P R e x p [-a\Q2
+ A2)/2]
,
(2.13)
where a2 = (3km)1/2. A cautionary remark is in order. One should not misinterpret the use of an oscillator potential - it is not the intent to model the observed baryon spectrum as that of a system of quantum oscillators because such a picture would fail. For example, the oscillator spectrum has EN ~ TV, whereas the baryon spectrum obeys the law of linear Regge trajectories (cf. Sect. XIII-2), E^ ~ N. The oscillator potential provides a convenient basis for structuring the calculation and nothing more.
Color dependence of the interquark potential Short of doing a complete spectroscopic analysis, we can find experimental support in the following simple example for the assertion that the two-particle interquark potential is twice as strong in mesons as it is in baryons. A potential model description for the meson and baryon mass splittings p(770) - TT(138) and A(1232) - N(939) is given by a QCD hyperfine interaction, i?hyp> akin to the delta function contribution in the QED hyperfine potential of Eq. (V-1.16),
Yl
(a = M,B) ,
(2.14)
where the {Wij} are constants and, assuming the color dependence is that given by Eqs. (2.4), (2.5), &M = 1 for mesons and fc# = 1/2 for baryons. We shall discuss in Sect. XIII-2 how this effect could arise from gluon exchange. Although there is ordinarily dependence on quark mass in the {Hij}, it suffices to treat the {Wij} as an overall constant since the hadrons in this example contain only light nonstrange quarks. The point is then to see whether the condition &M — 2fc# is in accord with phenomenology. Noting that for mesons the spin factors yield 2S2-3
f 1/4 = |
(5 = 1 ) , {s
o h
(2.15a)
XI-3 Bag model
285
whereas for baryons one has 4S2-9
/ 3/4 .
(S = 3 / 2 ) ,
-,
,-=1/2),
we find after taking expectation values that nip — mn
2UM |V ; M(0)| 2
2&M (Volume)^
(Volume) M -MI
V)B1"'" -i3/2 2v
(2-!6) V
/
2
.{T )M\
The measured values (c/. Eq. (1.13)) of the proton and pion charge radii imply that /CM/^B — 2. This example, along with others, lends credence to the assumed color dependence of Eq. (2.4). At this point we shall temporarily leave our discussion of the potential model to consider other descriptions of hadronic structure. We shall return to the potential model for the discussion of hadron spectroscopy in Chaps. XII-XIII. XI-3 Bag model A superconductor has an ordered quantum mechanical ground state which does not support a magnetic field (Meissner effect) and which is brought about by a condensation of dynamically paired electrons (Cooper pairs). An order parameter for this medium is provided by the Landau-Ginzburg wavefunction of a Cooper pair. Even at zero temperature, a sufficiently strong magnetic field, S c r , can induce a transition from the superconducting phase to the normal phase. For example, in tin the critical field is 5 c r (tin)~ 3.06 x 10~2 tesla, and the energy density of superconducting pairing (condensation energy ) is Usnper/V ~ 373 J/m 3 . Chromodynamics exhibits similar behavior, and this is the basis for the bag model [ChJJTW 74, Jo 78]. The QCD ground state evidently does not support a chromo-electric field, and is thus analogous to the superconducting state, although a compelling description of the QCD pairing mechanism has not yet been provided. In the bag model, the analog of the normal conducting ground state is called the perturbative vacuum. The vacuum expectation value of the quark bilinear qq (q = u, d, s) plays the role of an order parameter by distinguishing between the two vacua, QCD (O|W|O)QCD < 0 ,
pert (0|w|0) pert = 0 .
(3.1)
Hadrons are represented as color singlet 'bags' of perturbative vacuum occupied by quarks and gluons. The bag model employs as its starting
286
XI Phenomenological models
point the lagrange density [Jo 78] Aag = (£QCD - B) 0(qq)
,
(3.2)
where the ^-function (which vanishes for negative argument) defines the spatial volume encompassed by the perturbative vacuum. B is called the bag constant, and is often expressed in units of (MeV) 4 . Physically, it represents the difference in energy density between the QCD and perturbative vacua. Phenomenological determinations of B yield jgi/4 ^ 150 MeV, which translates to a QCD condensation energy of 34 3 UQCD/V ~ 1.0 x 10 J/m . Although huge on the scale of the condensation energy for superconductivity, this value appears less remarkable in more natural units, B ~ 66 MeV/fm 3 .
Static cavity To obtain the equations of motion and boundary conditions for the bag model, we must minimize the action functional of the theory. We shall consider at first a simplified model consisting of a bag which contains only quarks of a given flavor q and mass m. The equations of motion which follow from the lagrangian of Eq. (3.2) are (ifi - m)q = 0 ,
(3.3)
within the bag volume V and = Q > = 2B
(3.4a) (3.4b)
on the bag surface *S, where n^ is the covariant inward normal to S. Eq. (3.3) describes a Dirac particle of mass m moving freely within the cavity defined by volume V. Since the order parameter qq vanishes at the surface of the bag, the linear boundary condition in Eq. (3.4a) amounts to requiring that the normal component of the quark vector current also vanish at the surface. Thus quarks are confined within the bag. The nonlinear boundary condition represents a balance between the outward pressure of the quark field and the inward pressure of B.
Spherical cavity approximation In principle, the bag surface should be determined dynamically. However, the only manageable approximation for light quark dynamics is one in which the shape of the bag is taken as spherical with some radius R. For such a static configuration, the nonlinear boundary condition becomes equivalent to requiring that the energy be minimum as a function of R. The static cavity hamiltonian is H = f d3x [q\-iot Jv
- V)q + q](3mq + B] .
(3.5)
XI-3 Bag model
287
Observe that B plays the role of a constant energy density at all points within the bag. As in Eq. (1.1), the normal modes of the cavity-confined quarks and antiquarks provide a basis for expanding quantum fields. They are determined by solving the Dirac equation Eq. (3.3) in a spherical cavity. We characterize each mode in terms of a radial quantum number n, an orbital angular momentum quantum number £ (as would appear in the nonrelativistic limit), and a total angular momentum, j . Only j = 1/2 modes are consistent with the nonlinear boundary condition since the rigid spherical cavity cannot accommodate the angular variation of j > 1/2 modes. Such nonspherical orbitals can be treated only approximately, by implementing the nonlinear boundary condition as an angular average or by minimizing the solution with respect to the energy. In addition, since neither pi/% modes nor radially excited sly/2 modes are orthogonal to a translation of the ground state, they must be admixed with some of the j = 3/2 modes to construct physically acceptable excitations. For these reasons, the bag model has been most widely applied in modeling properties of the ground state hadrons rather than their excited states. Let us consider the S\/2 case in some detail. Even with the restriction to a single spin-parity state, there are still an infinity of eigenfrequencies u)n. Each un is fixed by the linear boundary condition, expressible as the transcendental equation n
=
Pn
(n = l , 2 , . . . ) ,
LUji - r TYlrt — 1
(3.6)
where pn = y/ool — m2R2. For zero quark mass, the lowest eigenfrequencies are u = 2.043, 4.611, For light quark mass (mR < 1) the lowest mode frequency is approximated by UJ\ ~ 2.043 + 0.493rai?, and in the limit of heavy quark mass (mR ^> 1) becomes OJ\ —• \/m2R2 + n2. The spatial wavefunction which accompanies destruction of an S1/2 quark with spin alignment A and mode n is
iJ Pr
f /*)X\ ) ,
(3.7)
while for creation of an sy2 antiquark we have
where e = ((un — mR)/(un + mR))1/2, x\ ls a two-component spinor, and x\ = ^a2X\' The full quark field q(x), expanded in terms of the modes, is given by
q(x) = J2 N(un) [il>n(*)e-iUnt/RKn) + ^ ( ^ e ^ ^ d ^ n ) ] ,
(3.9)
288
XI Phenomenological models
where
is a normalization factor which is fixed by demanding that the number operator Nq = Jh d3x q\x)q{x) for quark flavor q have integer eigenvalues. By computing the expectation value of the hamiltonian in a state of N quarks and/or antiquarks of a given flavor, one obtains (H) = NUJ/R + ATTBRS/3 - Zo/R .
(3.11)
In the final term, Zo is a phenomenological constant which has been used in the literature to summarize effects having a 1/R dimension, most notably the effect of zero-point energies, which for an infinite volume system would be unobservable. However, just as the Casimir effect is present for a finite volume system with fixed boundaries, such a term must be present in the static cavity bag model [DeJJK 75]. Unfortunately, a precise calculation of this effect has proven to be rather formidable, and so one treats Zo as a phenomenological parameter. Upon solving the condition d (H) /OR = 0, we obtain expressions for the bag radius R4 = ^(Nu
- Zo) ,
(3.12)
and the bag energy E=
\{AKB)1/A(NUJ
- Z 0 ) 3/4 .
(3.13)
The bag energy E is not precisely the hadron mass. Although the bag surface remains fixed in the cavity approximation, the quarks within move freely as independent particles. Thus at one instant, the configuration of quarks might appear as in Fig. XI-3(a), whereas at another time, the quarks occupy the positions of Fig. XI-3(b). As a result, there are unavoidably fluctuations in the bag center-of-mass position. The bag energy is thus E = I ^/p 2 + M2 \, where M is the hadron mass and p
(a)
(b)
Fig. XI-3 Quarks in a bag
XI-3 Bag model
289
represents the instantaneous hadron momentum. Although the average momentum vanishes ((p) = 0), the fluctuations do not, ((p 2 ) 7^ 0). For all hadrons but the pion, it is reasonable to expand the bag energy in inverse powers of the hadron mass, (3.14) E = M + (p2)/2M + ... . For the pion, one should instead expand as £;=(|p|) + M 2 < | P r 1 ) / 2 + ... . (3.15) One can employ the method of wave packets, to be explained in Sect. XII1, to estimate that (|p|) ~ 2.Si?"1, ( I P I " 1 ) — 0.7R for the pion bag, and ^p2) ~ N(JiR~2 for a bag containing N quarks and/or antiquarks in the Si/2 mode. Gluons in a bag Any detailed phenomenological fit of the bag model to hadrons must include the spin-spin interaction between quarks. One way to incorporate this effect is to posit that gluons, as well as quarks, can exist within a bag. With only gluons present, the lagrangian is taken to be [Jo 78]
- B] # ( - W / 4 - B) ,
(3.16)
and the Euler-Lagrange equations are &F%, = 0
(3.17)
in the bag volume V, and = 0 (3.18a) ^ = -45 (3.18b) on the bag surface S. In the limit of zero coupling, the gluon field strength becomes F*v = d^A^ — dvAa^. The field equations in V are sourceless Maxwell equations with boundary conditions x-E a = 0 and x x B a = 0 on 5, where E a and B a are the color electric and magnetic fields respectively. It is convenient to work directly with the gluon field Aa(x), and with a gauge choice to restrict the dynamic degrees of freedom to the spatial components. In mode n, these obey [V2 + (kn/R)2}Aan = 0 ,
(3.19)
and V • A£ = 0 (3.20) within the bag. The gluon eigenfrequencies kn are determined by the linear boundary condition rx(VxA;)=0.
(3.21)
290
XI Phenomenological models
Restricting our attention to modes of positive parity, we have for the gluon field operator
Aa(x) = J2NG(kn)(ji(knr/R)X1(T(n)aan^
+ H.c.) ,
(3.22)
n,a
where X ^ is a vector spherical harmonic. The gluon normalization factor is obtained, analogously to N(ujn) for quarks, by constraining the gluon number operator to be integer valued and we find [NG(kn)}-2 = [3(1 - sin(2kn)/2kn) - 2(1 + k2n) sin2(kn)]R2 .
(3.23)
The quark-gluon interaction In the following, we shall work with the lowest positive parity mode, for which k\ = 2.744. The quark hyperfine interaction in hadron H can be computed from the second order perturbation theory formula, -E"hyp — \H\Hq-g(Eo — HQ + ie)~ Hq-g\H)
(3.24)
where the unperturbed hamiltonian HQ is given in Eq. (3.5) and Hq-g is the quark-gluon interaction
# q _ g = -gs f Sx 3a(x) • Aa(x) , Jv
(3.25)
defined in terms of the quark color current density J
a
(rr\
n .(rp\^\a
n
.(rp\
^9^
Implicit in Eq. (3.24) is an infinite sum over all intermediate states. In practice the sum can be well approximated by the lowest energy intermediate state, and we find for hadron H ,
(3.27)
where hH = -O.177(H\ Y, <*i ' *j Fz • Fj \H) .
(3.28)
The numerical factor arises from an overlap integral of quark and gluon spatial wavefunctions, and F^,cr^ are respectively the color and spin operators for quark i. It is straightforward to demonstrate that hn — 0.708, IIN — —h/s. = hn/2, and hp = —h n/3.
XI-S Bag model
291
Table XI-4. Results of bag model fit £1/4
Zo
Ro
RN
RA
135.
1.01
2.13
5.5
5.6
m
3.5
33.
Notes: Rl/4 and rh are given in MeV and R0,RA I I ^ A i, Rn are given in GeV- l
A sample fit An example of how to determine parameters of the bag model is provided by quoting a simultaneous fit to JV, A, and n states [DoJ 80]. The SU(2) limit with mu = m^ = m is employed and the constraint of chiral symmetry is implemented by arranging the pion mass to vanish for zero quark mass,TO= 0. Since the fit entails working at different momentum scales, a running effective coupling as(R) is introduced in place of the fixed one. Although the precise form of as(R) is not known for small i?, let us assume its dependence to be logarithmic. A naive identification of the lowest order QCD formula with the bag radius, as(R) = 27r/[91n(jRo/i2)] for three flavors (Ro is the spatial counterpart of the confinement scale A in momentum), is untenable, as the coupling would diverge for i? — Ro. Hadrons, such as the pion and the nucleon, which have hyperfine contributions proportional to — as(R)R~1 would suffer mass instability as a result. This problem is avoided by employing a modified form, as(R) = 2TT/[9 ln(l + Ro/R)]. The bag model then implies the system of equations 2
+ ml)1'2)
= R-l[2un -Zo
4
3 - hNas(RN)] (p2 + ™i)
1/2
) = R^fivA
,
(3.29)
-Zo + 4nBRi/3 - hAas(RA)} ,
which contain a total of seven unknowns, i?, Zo, i?o, RN,A,KI a n d TO. The left-hand side is calculated using the wavepacket method of [DoJ 80], which will be described in Sect. XII-1. Employing Eq. (3.29) in a fit to the TV and A masses and using drriN^/dR \R=RN A = 0 yields four relations. For the pion bag, the energy is expanded in powers of the light quark mass,
En(m) = En(0) + mdE^/drhiO) + . . . .
(3.30)
In addition to fitting to the pion mass and minimizing with respect to Rn, a further constraint follows fromTO2(TO = 0) = 0. Results of the fit are given in Table XI-4. With the bag parameters now determined, the quark wavefunctions are known and various observable matrix elements
292
XI Phenomenological models
can be evaluated. Also, the p meson mass can be predicted, resulting in mp = 704 MeV. XI—4 Skyrme model In Chap. X, we explored the Nc —• oo limit of QCD. In some respects the world thus defined is not unlike our own. Mesons and glueballs exist with masses which are O(l) as Nc —• oo. To lowest order, these particles are noninteracting because their coupling strength is O(N~l). What becomes of baryons in this world? It takes Nc quarks to form a totally antisymmetric color singlet composite, so baryon mass is expected to be O(NC). Note the inverse correlation between interparticle coupling O(N~1) and baryon mass O(NC). This is reminiscent of soliton behavior in theories with nonlinear dynamics. Sine-Gordon soliton An example is afforded by the Sine-Gordon model, defined in one space and one time dimension by the lagrange density, £SG = \{d^)2
- ^ ( 1 - cos 0
(4.1)
where a and j3 are constants. For small amplitude field excitations, an expansion in powers of
£SG = \{dM2 - f v>2 + ^ V + O(PV),
(4.2)
identifies the parameter a as the boson squared mass and (3 as a coupling strength. For j3 —> 0 we recover the free field theory. The Sine-Gordon lagrange density has also a nonperturbative static solution, <po(x) = - tan" 1 (exp( v / ax)) ,
(4.3a)
with energy EQ - 8V^//3 2 •
(4.3b)
This solution is a Sine-Gordon soliton. The natural unit of length for the soliton is a" 1 / 2 , and the energy E$ diverges as the coupling is turned off (/?—•()). The potential energy in this theory has an infinity of equally spaced minima, with (pW = 2im/(3 (n = 0, ±1, ±2,...). As the coordinate x is varied continuously from —oo to +oo, the soliton amplitude (fo(x), starting from the minimum ip^ — 0, moves to the adjoining minimum ?W = 2TT//3. An index AJV, the winding number, counts the number of minima shifted. It can be expressed as the charge associated with a
XI-4 Skyrme model
293
current density,
J" = i
(4.4)
such that AN =
dx J°{x) = — [
(4.5)
^
For
Tr (d^Ud^) ,
(4.6)
where U is an 5/7(2) matrix which transforms as U —> LUR~l under a chiral transformation for L E SU(2)L and R G SU(2)R. Unfortunately £2 cannot support an acceptable soliton, as the soliton would have zero size and zero energy. To see why, recall that the Sine-Gordon soliton has a natural unit of length a" 1 / 2 . Suppose there is an analogous quantity, i?, for the chiral soliton. Then we can write the radial variable as r = fi?, where f is dimensionless. For a static solution, the energy becomes 3
3
E = f d x H = - I d x £=^f
I d3x Tr ( W • W f ) .
(4.7)
Upon expressing the integral in terms of the dimensionless variable r, we find E = aR where a is a nonnegative number. The energy is minimized at R = 0 to the value E = 0. This trivial solution is unacceptable, and thus the model must be extended. The Skyrme model [Sk 61] employs, in addition to £2, a quartic interaction of a certain structure, ^
^
Tr [d^U U\dvU U]}2 ,
(4.8)
where e (not to be confused with the electric charge!) is a dimensionless real-valued parameter. The above chiral lagrangian should look familiar, since it is part of the general fourth order chiral lagrangian used in Chap. VI. In particular, Eq. (4.8) is reproduced if 2oc\ + 2^2 + a% — 0, in
294
XI Phenomenological models
which case (32c 2)"1 = (a2 — 2ai — 0:3 )/4. The comparison with the phenomenology of Chap. VI is not completely straightforward, as the pion physics was treated to one-loop order while the Skyrme lagrangian is used at tree-level. We note, however, that the coefficients in Table VI-1 give 2 Q 1 + 2a2
+
°3 = 0.65 ,
a 2 - 2a! - 03 = 0.0036 .
(4.9)
The latter combination, which is independent of renormalization scale, numerically gives e ~ 5.9. In the following development, we shall follow standard practice by taking the parameter e as arbitrary. We seek a static solution of the Skyrme model. Our strategy shall be to first determine the energy functional of the theory, and then minimize it. Following the procedure leading to Eq. (4.7), we can write the energy as (4.10)
- / •
where X^ = Ud^U^ and X^ = —Xjj,. It is necessary that Xi —> 0 as |x| —• oo in order that the energy be finite. Thus U must approach a constant element of 5C/(2), which we are free to choose as the identity / . For the mesonic sector of the theory, the vacuum state corresponds to C/(x) = / for allx. In this state, both the field variable X{ and the energy E vanish. The form U ~ I+iTZ-r/F^, used extensively in earlier chapters, corresponds to small amplitude pionic excitations of the vacuum. To see that the Skyrme model does support a nontrivial soliton, we cast the energy integrals of Eq. (4.7) in terms of a natural length scale R and find E = aR + bR~1 ,
(4.11)
where a, b are nonnegative. For a, b ^ 0, the energy is minimized at nonzero R and nonzero E. Thus, the quartic term of Eq. (4.8) is seen to have the desired effect of inducing soliton stability. Moreover, for arbitrary U a lower bound on the energy is provided by applying the Schwartz inequality to Eq. (4.11), E > ^
fd3x |Tr EijkXiXjXkl .
(4.12)
It is not hard to show that the integrand of Eq. (4.12) is proportional to the zeroth component of a four-vector current,
XI~4 Skyrme model
295
which is divergenceless, d^B^ = 0, and thus has conserved charge dsx JB°(X) .
(4.14)
• / •
It turns out that the current B^ can be identified as the baryon current density and B is the baryon number of the theory. Note that this is consistent with our prescription U(x) = I for the meson vacuum, where we see from Eq. (4.13) that B = 0. Interestingly, B turns out to have an additional significance. It is the topological winding number for the Skyrme model, analogous to AN for the Sine-Gordon model. The point is, by having associated spatial infinity with a group element of £77(2) to ensure that the field energy is finite, we have placed the elements of physical space into a correspondence with the elements of the compact group SU(2). The parameter space of each set is £ 3 , the unit sphere in four dimensions, and it is precisely the field U which implements the mapping. The mappings from 5 3 to 5 3 are known to fall into classes, each labeled by an integer-valued winding number. In this context, B serves to measure the number of times that the set of space points covers the group parameters of SU(2) for some solution U of the theory.
The
Skyrme soliton
The Skyrme ansatz for a chiral soliton (skyrmion) has the functional form [BaNRS 83, AdNW 83] tfo(x) = exp [iF(r)r • x] .
(4.15)
The unknown quantity is the skyrmion profile function F(r). To specify it, we first determine the energy functional by substituting Uo into Eq. (4.10),
E[F) =
4TT
f Jo
(4.16)
where a prime signifies differentiation with respect to the argument. For a static solution, the minimization of the energy generates an extremum of the action, and is hence equivalent to the equations of motion. The variation 6E/6F = 0 generates a differential equation for F, 4 +
2 s i n ^ V /
2
+
V
+
^
s
i
4
n
2
F
rl
!
0
,
(4.17) as expressed in terms of a dimensionless variable f = r/R, with R 1 = 2eFn. This nonlinear equation must be solved numerically, subject to
XI Phenomenological models
296
certain boundary conditions. The condition U = I at spatial infinity implies F(oo) = 0. The boundary condition at r = 0 is fixed by requiring that the soliton correspond to baryon number 1. For the Skyrme ansatz, the baryon number charge density is
B°(r)= -
(4.18)
2TT 2
and corresponds to a baryon number B = -!- [2F(0) - 2F(oo) - sin2F(0) •sin2F(oo)] . Z7T
(4.19)
This leads to the choice F(0) = n. Although the profile F(r) cannot be determined analytically over its entire range, it is straightforward to show that f 7T — const, r \ const. r~ 2
F(r)
(r —-• 0) , (r —• oo) .
(4.20)
We display F(r) in Fig. XI-4. Insertion of the solution to Eq. (4.17) into the energy functional E[F] yields the mass M of the skyrmion, and from a numerical integration we obtain M ~ 73 F^/e. There is an important point to be realized about the skyrmion - it represents a use of chiral lagrangians outside the region of validity of the energy expansion. Recall that the full chiral lagrangian is written as a power series, C = £2 + £4 + • • • in the number of derivatives. When matrix elements of pions are taken, terms with n derivatives produce n powers of the energy. Hence at low energy, one may consistently ignore operators with large n, as their contributions to matrix elements are highly suppressed. However, in forming the skyrmion one employs only £2 ami a subset of £4. The relative effects of the two are balanced in the minimization of the energy functional, and as a result both contribute
1
I
r
1
I
r
F(r)
Fig. XI-4 Radial profile of the skyrmion
XI-4 Skyrme model
297
equally. In an extended model containing CQ, one would expect the import of CQ to be analogously comparable to £4, etc. Higher-derivative lagrangians thus will contribute to skyrmion matrix elements, and the result cannot be considered a controlled approximation. However, this is not sufficient cause for abandoning the skyrmion approach. It simply becomes a phenomenological model rather than a rigorous method, and thus has a status similar to potential or bag models.
Quantization and wavefunctions The analysis done thus far is at the classical level, and merely shows that the chiral soliton satisfies the equations of motion. To determine the quantum version of the theory, we shall follow a canonical procedure. An analogy with quantization of the rigid rotator may help in understanding the process. A classical solution consists of the rotator being at any fixed angular configuration {0, (p}. To obtain the quantum theory, one allows the rotator to move among these solutions, and describes its motion in terms of the angular coordinates and their conjugate momenta {pe,Pip}- The quantum states are those with definite angular momentum quantum numbers {£,ra},and have wavefunctions given by the spherical harmonics, {9,(p\t,"i)=Ye,m(6,
.
(4.21)
The classical skyrmion solutions consist not only of U$ (cf. Eq. (4.15)), but also of any constant SU(2) rotation thereof, UQ — AU$A~l with AeSU(2). A particularly simple approach to quantization is then to allow the soliton to rotate rigidly in the space of these solutions, U = A(t)U0A-\t)
,
(4.22)
where now A(i) is an arbitrary time-dependent SU(2) matrix. One proceeds to define a set of coordinates {a^}, their conjugate momenta {?!"£ = dC/ddk}, and a hamiltonian constructed via Legendre transformation H = 7rkdk-L
.
(4.23)
We shall presently describe how to choose quantum numbers and determine the associated wavefunctions. Note that this approach is approximate in that it neglects the possibility of spacetime dependent excitations such as pion emission. As such, it would be most appropriate for a weakly coupled theory (as occurs for Nc —> oo) where the soliton rotates slowly, but is only approximate in the real world. In general, an SU(2) matrix like A can be written in terms of three unconstrained parameters {9k} as A(t) = exp(ir • 0) = I cos (9 + ir • 9 sin0 .
(4.24)
298
XI Phenomenological models
However, we can equivalently employ the four constrained parameters, ao = cos# ,
a = 0 sin0 ,
(4.25a)
where 3
l=l •
( 4 - 25b )
Substitution of the rotated quantity U into Eq. (4.7) and evaluation of the spatial integration yields 3
L = -M + \Tr(d0A^d0A)
= -M+
2A^~] a\ ,
(4.26)
k=0
where A = TrA/3esF7r, with A = f drr2
sin2 F [1 + 4(F^ + sin2 F/r2)} ~ 50.9 .
(4.27)
As written in terms of the conjugate momenta 7T& = 4Ad&, the hamiltonian is 1
3
Adopting the canonical quantization conditions (4.29) we see that the canonical momenta can be expressed as differential operators, TTfc = —id/dak- Thus the hamiltonian has the form
H = M-±V24,
(4.30)
where V 2 is the four-dimensional laplacian restricted to act on the threesphere by the constraint of Eq. (4.25b). We can determine the eigenvalues and eigenvectors of H by working in analogy with the more familiar three-dimensional laplacian, | ^
^ | - - 1 L
2
2
.
(4.31)
r or r If constrained to the unit two-sphere by the condition Ylk=i x\ ~ r^ ~ •'» the three-laplacian V3 reduces to —L 2. As is well known, the three components of L are operators Li, L2, £3 which satisfy [Lj, Lk] = i6jkeLe ,
(4.32)
XI-4 Skyrme model
299
and generate rotations in the 2-3, 3-1, 1-2 planes respectively. The underlying symmetry group is SO (3), and the eigenfunctions are the spherical harmonics. The four-dimensional problem is treated by analogy. Upon adding an extra dimension labeled by the index 0, we encounter the additional operators K\, K2, K% which generate rotations in the 0-1, 0-2, 0-3 planes. The full set of six rotational generators can be represented as Kk = aoTTfc — afcTro .
(4.33)
The extended symmetry group is SO (A) and the commutator algebra of the rotation generators is [Lj, Lk) = itjkeLe ,
[Lj, Kk] = iejMKt
,
[Kj, Kk] = iejkeLe . (4.34)
The mathematics of this algebra is well known, underlying for example the symmetry of the Coulomb hamiltonian in nonrelativistic quantum mechanics. By the substitutions T = (L - K)/2 ,
J = (L + K)/2 ,
(4.35)
we arrive at operators T and J which generate commuting 5(7(2) algebras. We associate T with the isospin and J with the angular momentum. The explicit operator representations, Tk = i(-eijkaidj
+ aodk - akdo) ,
Jk — i(—€ij kaiOj — aoO
k
+ akOo) ,
follow immediately from Eq. (4.33), and the Skyrme hamiltonian becomes H = M + (T 2 + J2)/4A .
(4.37) 2
2
It follows from the commutator algebra of Eq. (4.34) that T = J . Thus the quantum spectrum consists of states with equal isospin and angular momentum quantum numbers, T — J. This is no surprise. After all, in the Skyrme ansatz of Eq. (4.15), the isospin and spatial coordinates appear symmetrically, and we expect the quantum spectrum to respect this reciprocity. Our final form for the hamiltonian, H = M + J2/2A ,
(4.38)
has the eigenvalue spectrum E = M + J ( J + 1)/2A ,
(4.39)
where in general J = 0,1/2,1,3/2,.... By analogy with the usual spherical harmonics, the eigenfunctions of H are seen to be traceless symmetric polynomials in the {ak}. However, both {ak} and {—a k} describe the same solution U (cf. Eq. (4.22)). In the quantum theory, eigenfunctions thus fall into either of two classes,
300
XI Phenomenological models
— ak}) = =tV>({afc})- Since fermions correspond to the antisymmetric choice, we select only the half-integer values in Eq. (4.39). In the Skyrme model, the N and A baryons will have wavefunctions which are respectively linear and cubic in the {a^}. To construct such states, it is convenient to employ the differential representations of Eq. (4.36) to prove L3(ai ± za2) = ±(ai db ia2) , ± za3) = ±(a 0 ± za3) ,
^3«o,3 = 0 , #3^1,2 = 0 .
From these and Eq. (4.36), the T3 = J3 = 1/2 eigenstate of a proton with spin up is found to be CA|PT) = ±
(a1+ia2) .
(4.41)
The normalization of this state is obtained from the angular integral over the three-sphere
=
= Jdn3
{a\ + 4) ,
(4.42)
where the angular measure is / » 22 T T
/
dft3 = /
Jo
PIT
P7T
d(p / d6 sin(9 / dx sin 2 * , Jo
(4.43)
Jo
and spherical coordinates in four dimensions are defined by (4.44) a2 == cos sinx\ • aa\3 = = sin sin xx sin cos 66 cos , ? , ao The remaining nucleon states can be found by application of the spin and isospin lowering sperators J_ = [(ai — ia2)c?3 — (a3 + iao)d\ + (—ao + ia 3)d2 + (a2 + it T- = [(ai — za2)c?3 + (—a 3 + iao)c?i + (ao + ia 3 )$ 2 — (a2 + z<
(4.45) where dk = d/dak- The T = J = 3/2 A states are formed by employing analogous ladder operations on is/2 (A\ A++) = - ^ (01 +ia 2 ) 3 .
(4.46)
It is remarkable that fermions can be constructed from a chiral lagrangian which contains nominally bosonic degrees of freedom. However, the presence of a nonzero fermion quantum number can be easily verified by direct calculation. The wavefunctions for the eigenstates (the equivalents of Y^m(0, if) for the rigid rotator) are given by 5/7(2) rotation matrices with half-integer
XI-4 Skyrme model
301
values. These are defined by the transformation properties of states under an SU(2) rotation A [Ed 60], ^
3
)
3
.
(4.47)
The simplest case is then just the T =1/2 representation, which we know is rotated by the matrix A, a
i(ai °+ iaz) " ia2) \ (4 48) a1+ia2) (ao-ia3))&h ' ^AS) Comparison with Eq. (4.41) and with the results of Eq. (4.45) shows that the properly normalized nucleon wavefunctions are
(A\ NT3,Ss) = I ( - f 3 + i / 2 V^]Sa(A)
.
(4.49)
The general case for a nonstrange baryon B of isospin T and spin S (S = T) is given by
[^]
V2
(
X
,
(4.50)
of which the A states are specific examples. Finally, the N and A masses are MN = M + 3/8A = 73Fn/e + If the measured JV, A masses are used as input, one obtains e = 5.44 and Fn = 65 MeV. Alternatively, from the empirical value for Fn and the determination e ~ 5.9 from pion-pion scattering data, the model implies Mjy ~ 1.27 GeV, MA ^ 1.80 GeV. In either case, agreement between theory and experiment is at about the thirty percent level. The next state in the spectrum would have quantum numbers T = J = 5/2 and is predicted by the first of the above fitting procedures to have mass M5/2 = M + 35/8A ~ 1.72 GeV. There is no experimental evidence for such a baryon. Although our discussion has been based on the SU(2) flavor symmetry of isospin, it is possible to extend the analysis to flavor SU(3) [Gu 84]. There, the action consists of the usual quadratic and quartic Skyrme forms, the quark mass term, and also the Wess-Zumino-Witten action of Eq. (VII-3.21). This final contribution was not considered in our previous analysis because it vanishes identically for SU(2). It turns out not to affect the equation for the static classical soliton, but does enter into the quantization procedure and thus contributes to the various quantum currents, etc. Two distinct approaches have been adopted for analyzing
302
XI Phenomenological models
the SU(S) model. Either one expands about zero quark mass and treats the quark mass term perturbatively [Ch 85], or one works in the limit of a large s-quark mass [CaHK 88] and thereby generates equations which describe the motion of kaons in a classical background of the 5(7(2) soliton. We shall not pursue this point further, but shall return to the Skyrme model in Sect. XII-1, where the problem of computing matrix elements is addressed. Although the development of the skyrmion and its quantization have been motivated by large-Nc ideas, we know of no proof that requires the skyrmion to come arbitrarily close to the baryons of QCD in the Nc —> oo limit. An oft-cited counterexample is the existence of a oneflavor version of QCD. Such a theory still contains baryons, such as the A + + . However, it makes no sense to speak of a one-flavor Skyrme model, as an SU(2) group is required for the underlying soliton C/o- The Skyrme model remains an interesting picture for nucleon structure because it is in many ways orthogonal to the quark model, and thus offers opportunities for new insights.
XI-5 QCD sum rules Low energy QCD involves a regime where the degrees of freedom are the hadrons and where it is futile to attempt perturbative calculations of hadronic masses and decay widths. Contrasted with this is the shortdistance asymptotically free limit in which quarks and gluons are the appropriate degrees of freedom, and in which perturbative calculations make sense. The method of QCD sum rules represents an attempt to bridge the gap between the perturbative and nonperturbative sectors by employing the language of dispersion relations [ShVZ 79a, ReRY 85, Na 89]. The existence of sum rules in QCD is quite general, and some might dispute the classification on these sum rules as a phenomenological method. However in practice, to utilize the sum rules involves the introduction of various approximations and heuristic procedures. Like quark model methods, these are motivated by an intuition of the most important physics involved, but are not always rigorous consequences of QCD. As a result, there remains considerable art in their use.
Correlators It is convenient to approach the subject by considering the relatively simple two-point functions. Thus we consider the quark bilinear, Mx)=q1(x)Tq2(x)
,
(5.1)
where F is a Dirac matrix, and analyze the correlator,
i f dAxeiqx(0\T(JT(x)4(0))\0)
.
(5.2)
XI-5 QCD sum rules
303
Such quantities can be expressed in terms of invariant functions H*r(q2) and attendant kinematical factors, e.g. as for the correlators of pseudoscalar currents {Jp) and of conserved vector currents (Jy), UP(q2) = i ( dAx eiq-x(0\T(Jp(x)JP(0))
|0)
(5.3a)
e*x(0\T (J^(x)J^(O)) |0). (5.3b) Analogous structures occur for other currents. There are several means for analyzing a quantity like IIr(
- g)|(O|JF(O)|n)|^ ,
(5.4)
n
where so is the threshold for the physical intermediate states. Such considerations, together with the application of Cauchy's theorem in the complex q2 plane, imply a dispersion relation for IIr(<72),
where the {an} are N subtraction constants.* One attempts to introduce a phenomenological component to the dispersion relation by expressing Im Ilr(s) in terms of measureable quantities, e.g. with cross section data as in the case of the charm contribution ey^c to the vector current, 1
CTe+e-—•charm
$s
where ec is the c-quark electric charge and s is the squared center-of-mass energy. If, as is usually the case, such data is not available, another means must be found for expressing Im Ilr(s) in the range so < s < oc. To approximate the low-5 part of Im lip (5), one usually employs one or more single-particle states. As an illustration, let us determine the contribution to Hp(q2) of a flavored pseudoscalar meson M which is a bound state or a narrow-width resonance of the quark-antiquark pair * The number of subtraction constants needed depends on the behavior of Im Ilr (s) in the s • limit, with Ur{q2) ~ q2N Ing 2 requiring TV subtractions.
304
XI Phenomenological models
i2- In this instance, we take the pseudoscalar current in the form of an axial-vector divergence, Jp —> d^A^L, with 2 ,
= y/2FMm2M
,
where TRM and FM are the meson's mass and decay constant. Eq. (5.4) implies
= 2F2MmlM6(q'1
which yields pp(q2) = 2F^m\I6(q2
Then
-
— m2M) for the spectral function or
Im n P | m eson = 2FlIm\I7r6(s
- m2M)
(5.9)
for the dispersion kernel. Thus, bound state or narrow-resonance contributions give rise to delta-function contributions. It is not difficult to take resonant finite-width effects into account if desired. One or more of these single-particle contributions are then used to represent the low-s part of the dispersion integral. Proceeding to higher s-values in the dispersion integral, one enters the continuum region, where multiparticle intermediate states become significant and the bound-state (or resonance) approximation breaks down. Although, as described below, one ordinarily attempts to suppress the large-5 part of Im Ilr(s) by taking moments or transforms of the dispersion integral, it is common to add to the low-s contribution a 'QCD continuum' approximation, Imllr(s) — • e(s-s large—s
c)ImUcontXs)
,
(5.10)
taken from discontinuities of QCD loop amplitudes and their O(as) corrections. In Eq. (5.10), sc parameterizes the point where the continuum description begins and the form of Im Ilc^t. depends on the specific correlator. Experience has shown that this 'parton' description can yield reasonable agreement of scattering data even down into the resonance region, provided the resonances are averaged over (duality).
Table XI-5. Local operators of low dimension. d:
0
4
4
On'. 1 mqqq G^ G T
6 qTqqVq
6
6
XI-5 QCD sum rules
305
Operator product expansion A representation for correlators which is distinct from the above phenomenological approach can be obtained by employing an operator product expansion for the product of currents,
i fd4x e*xT (Jr(z)jf(O)) = £c£(92)G>n • **
(5.11)
n
The {On} are local operators and the {C£(g 2 )}, called Wilson coefficients, are c-numbers. The {On} are organized according to their dimension, and aside from the unit operator / , are constructed from quark and gluon fields. Table XI-5 exhibits the operators up to dimension 6 which might contribute to the correlator of Eq. (5.2). Although one may naively expect all the operators but the identity to have vanishing vacuum expectation values (as is the case for normalordered local operators in perturbation theory), nonperturbative longdistance effects like those discussed in Sect. Ill—5 generally lead to nonzero values. Most often, the operator product approach contains vacuum expectation values like (^G^Ga^o = (^f-G2)o and (mqqq)o as universal parameters, 'universal' in the sense that the same few parameters appear repeatedly in applications. Calculation reveals that the quantity (^G2)o is divergent in perturbation theory, so the perturbative infinities must be subtracted off if one is working beyond tree-level. In principle, all the vacuum expectation values should be computable from lattice gauge theory once the renormalization prescriptions are specified. At present, the only theoretically determined combinations are the products (msss)0 ^ -Flm\
,
(m(uu + dd))0 ~ -2i^2ra2 ,
(5.12)
which follow from the lowest order chiral analysis in Chaps. VI,VII. Although only the product rrnpt/j is renormalization group invariant and it is difficult to separate out quark masses uniquely, sum rule calculations often adopt the value (qq)0 = -(225±25MeV) 3
(q = u,d,s) ,
(5.13)
corresponding to ms = 180 MeV. In any given application, one must cope with the other local operators as best one can, e.g. the value of (^-G 2)o has been estimated from charmonium data. Of course, use of the short-distance expansion must be justified. We have seen in previous chapters how a given hadronic system is characterized in terms of the energy scales of confinement (A) and quark mass ({mq}). Given these, it is indeed often possible to choose the variable q such that short-distance, asymptotically free kinematics obtain. Two situations which have received the most attention are the heavy quark limit
306
XI Phenomenological models
x (a)
x
>g x
x
(b)
xx (c)
(d)
Fig. XI-5 Contributions to coefficient functions. (m2 » A2,g2) and the light quark limit (q2 » A2 » m2q). Once in the asymptotically free domain, it is legitimate to apply QCD perturbation theory to the C^(q2), with the expansion being typically carried out to one or two powers of as, CTn(q2) = ATn(q2) + BTn(q2)as
+ ... .
(5.14)
Rather extensive lists of Wilson coefficients already appear in the literature. Fig. XI-5 depicts contributions to a few of the Wilson coefficients. Denoting there the action of a current by the symbol c x', we display in (a)-(b) the lowest order and an O(as) correction to operator / and in (c)-(d), the lowest order contributions to (^fG2)o and to (mqq)o respectively. Finally, besides the vacuum expectation values mentioned above, additional parameters which generally occur in the operator product representation are the quark mass mq and the strong coupling as. Since these quantities will depend on the momentum q, one must be careful to interpret them as running quantities whose renormalization is to be specified. Due to asymptotic freedom, they too can be treated perturbatively, e.g. as in the familiar expression Eq. (II-2.78) for as.
Master equation The essence of the QCD sum rule approach is to equate the dispersion and the operator product expressions to obtain a 'master equation', W )
I
Jn
"""U"/
,
_ \
^/nf!7«2\/^i \
(5 jg\
It is important to restrict use of this equation to a range of q2 for which both the short-distance expansion and also any 'resonance + continuum' approximation to Im Ilr are jointly valid. To satisfy these twin constraints, it is common practice to not analyze Eq. (5.15) directly, but rather to first perform certain differential operations leading to either moment or transform representations. The nth moment M^(QQ) is de-
XI-5 QCD sum rules
307
fined as 1
s
f°
,
Imnr(s)
(7TW
(5.16) where, in the spacelike region #2 < 0, one usually works with the variable Q2 = —q 2. By taking sufficiently many derivatives, one can remove unknown subtraction constants from the analysis and at the same time, enhance the contribution of a single-particle state at low s in the dispersion integral. Alternatively, one can express the dispersion integral as a kind of transform. The Borel transform is constructed from the moment M^(Q2) as 2n
2
n Q Ml(Q )
1
—> —
C°°
ds e ^ I m U
n,Q 2 ^oo 7TT JSQ
T(s)
,
(5.17)
where Q2/n = r remains fixed in the limiting process and defines the transform variable r. To obtain the factor e~~s/r in the above dispersion integral, we note
A slightly different version of exponential transform which has appeared in the literature is defined analogously, 1 r° 7T J8o 8o
where the transform variable is now a = n/Q2. The transform method serves to remove the subtraction constants and to suppress the contributions from operators of higher dimension in the operator product expansion. Examples Applications of the QCD sum rule approach generally proceed according to the following steps. 1) Choose the currents and write a dispersion relation for the correlator. 2) Model the dispersion integrals with phenomenological input, usually some combination of single-particle states and continuum. 3) Employ the operator product expansion, including all appropriate operators up to some dimension at which one truncates the series. 4) Obtain the Wilson coefficients as an expansion in as. 5) Use the moment or transform technique to extract information from the master equation. 6) Vary the underlying parameters until stability of output is achieved.
308
XI Phenomenological models
Let us consider three examples, keeping the treatment on an elementary footing to better emphasize the kinds of relationships which QCD sum rules entail. (i) Charmonium spectrum: The correlator for the charm quark vector current J^c m ' = cj^c has the form
?m)(92) = iJdAx e* (5.20)
and the corresponding dispersion relation is
In the following we use the dispersive kernel of Eq. (5.6). Following the original treatment of this system [ShVZ 79a], we work at Q2=0 and employ a moment analysis of the short-distance expansion containing just the identity and gluon contributions. This yields for the n th moment, expanded to first order in as, 1 Q2=0
{
3 ( n - l ) ! ( n + l)
c)
v
^v
s
' 7r2n(n + l)(n + 2)(rc + 3) >n^°A 36 (2n + 5) m^
n
(5 22)
where {a%} = {0.73, 0.71, 0.51, 0.22, -0.14, etc.} for n = 0,1,.... In practice, one analyzes the ratios rn = Mn/Mn-\, expanded in powers of a s , rather than the individual moments. Ratios should be less sensitive to experimental errors and tend to yield a more highly convergent expansion in as. Theory and experiment can then be compared in a plot of rn vs. n. For simplicity, suppose that the strong coupling is fixed from J/i/j decay data, as(s = Am2) ~ 0.2, leaving just the two parameters mc and (^-G2)o to be fitted by sum rules. Since the gluon contribution increases as n 3 relative to the identity operator, the mass parameter mc is fixed from low-n ratios (n < 4), yielding a value mc ~ 1.26 GeV. Proceeding to higher n (n < 10), but keeping only the identity contribution, one notices the agreement between theory and experiment begins to break down. Attributing this to the need for a gluon condensate term, one extracts the value (^-G2)o — (350 MeV)4. Continuing on to even higher n would not be justified in the context of this simple model, because other condensate terms grow even more rapidly with n. Were we to extend the above procedure from c-quarks to fr-quarks, we see from Eq. (5.22) that the relative strength of the gluon contribution would be heavily suppressed by the factor (ra c/rab)4. That is, nonper-
XI-5 QCD sum rules
309
turbative effects are able to compete with radiative corrections on a relatively equal footing for c-quark physics, whereas they are overwhelmed in 6-quark systems. In this sense, charm represents a fertile testing ground for the QCD sum rules. However, care must be exercised in interpreting numerical values of mc and (^-G2)o due to gauge and renormalization dependence. In particular, the above analysis was carried out in Landau gauge. Also, the mass parameter mc involves a determination at spacelike momentum, q2 = —m 2, and is not the same as would be inferred from a timelike e+e~ scattering experiment (for which q2 = +m2).* It is possible to extend and improve upon the above simple development in a number of ways. By taking Q2 ^ 0, one increases the range of n values for which the 'identity plus gluon-condensate' approximation is valid. Instead of using cross section data, one can saturate the dispersive kernel in terms of individual J/ij) states, and by considering different correlators, it is possible to study a variety of partial waves. Finally of course, additional condensates of higher dimension should be included, and alternative transform methods studied. For example, a study using the exponential transform and including all operators of dimension up to and including d = 8 obtains a rather larger gluon condensate, in the range (^-G2)0 ^ (450 -> 510 MeV)4 [Pa 90]. (ii) Decay constant of a heavy meson: We consider the axial-current divergence and correlator associated with a heavy quark Q and a light antiquark q of mass TRQ and m ~ 0 respectively,
UP(q2) =i f
d4x eiq-
A dispersion representation free of subtraction constants is 2
°° , Im Up(s) ds
(JTW '
,
A,
(5 24)
'
Recalling Eq. (5.9), we have in the 'particle plus continuum' approximation, -Im UP(s) = 2F2MMlI6{s - M2M) + 6{s - sc)lmllcont.(s)
(5.25)
7T
for a heavy flavored meson of mass MM and decay constant FM- Upon performing an exponential transform of the type in Eq. (5.19), the sum The relation between the two is m*.\qi=rn2
= (1 + 4In2 as//K)'m%\q2 =
_rri2.
310
XI Phenomenological models
rule can be cast in the form
2F2MMiIe-MM°= - r
ds ds e-^ImnpertW (5.26)
where Imnpert. (s) denotes perturbative contributions to the identity operator in the short-distance expansion and we suppress writing the explicit form of the operators O^Q. Regarding the stability aspect of the fit, one seeks (for a fixed value of the parameter sc in Eq. (5.26)) a minimum F]^m in the value of FM as a is varied. Thus, a stable prediction for FM is found provided one can find a 'plateau of stability' in the response of Fjtfin as sc is varied. The estimates FB — Fp ~ lAFn have been obtained in this manner [Na 87]. Nucleon mass: It is not necessary to restrict oneself to mesonic currents as in Eq. (5.1). Here, we consider a current r)w (and its correlator) which carries the quantum numbers of the nucleon, W =
U(q2) = ni(g 2 ) + #I2(<,2) = i Jd4x
e**(0\T(r,N(x)rjN(0))
|0> ,
(5.27) where C is the charge conjugation matrix. The simplest approximation to the dispersion integral comes from the nucleon pole, n
(92)U = A ^ J ^ , q — ivi
(5.28)
N
where the coupling A^ is proportional to the 'nucleon decay constant', i.e., the probability of finding all three quarks within the nucleon at one point. Upon making a simple approximation to the operator product expansion, ^
^
ln(-q2)
,
(5.29)
and employing a Borel transform, one obtains an amusing relation between nucleon mass and quark condensate [Io 81], MN = (-87T2(qq)0)1/3
+ ... -
1 GeV ,
(5.30)
and implies the vanishing of the former with the latter. However, it should be realized that this result is subject to important corrections in a more careful treatment. Each of the above examples has involved two-point functions. It is possible to apply the method to three-point functions as well, where one
Problems
311
can obtain coupling constant relations. The underlying principles are the same, but some technical details are modified due to the larger number of variables, e.g. one encounters double-moments or double-transforms. QCD sum rules work best when there is a reliable way to estimate the dispersion integral, most often with ground state single-particle contributions. However, the method has its limitations. It is not at its best in probing radial excitations since their dispersion effects are generally rather small. Even having a good approximation to the dispersion integral is not sufficient to guarantee success. For example, the method has trouble in dealing with high-spin (J > 3) mesons because, even with dispersion integrals which are dominated by ground state contributions, power corrections in the operator product expansion become unmanageable. Problems 1) Velocity in potential models Truly nonrelativistic systems have excitation energies small compared to the masses of their constituents. However, fitting the observed spectrum of light hadrons requires excitation energies comparable to or larger than the constituent masses. Assuming nonrelativistic kinematics, consider a particle of reduced mass m moving in a harmonic oscillator potential of angular frequency u. Expressing u in terms of the energy splitting E\ — Eo between the first-excited state and the ground state, use the virial theorem to determine the 'rms' velocities of the ground state (vnns) and of the first-excited state (tins) in terms of E\ — EQ. Compute the magnitude of Vrms/c and Vrms/c using as inputs (i) m/2 — mp ~ 500 MeV for light hadrons and (ii) m^S) ~ mJ/tp — 590 MeV for charmed quarks. Your results should demonstrate that the kinematics of quarks in light hadrons is not truly nonrelativistic. However, one tends to overlook this flaw given the potential model's overall utility. 2) Nucleon mass and the Skyrme model a) Use the Skyrme ansatz of Eq. (4.15) to derive the expression Eq. (4.16) for the nucleon energy E[F]. b) Using the simple trial function F(r) = 7rexp(—r/R), scale out the range factor R to put E[F] in the form of Eq. (4.11), where a ~ 30.8F^ and b ~ 44.7/e2 are determined via numerical integration. c) Minimize E[F] by varying R and compare your result with the value 73Fn/e determined with a more complex variational function.
312
XI Phenomenological models d) Using the numerical value of the nucleon mass, determine e and compare with the value 1 4 1
expected from chiral scaling arguments. 3) A 4QCD sum rule' for the isotropic harmonic oscillator Consider three-dimensional isotropic harmonic motion with angular frequency u) of a particle of mass ra. a) Using ordinary quantum mechanics or more formal path integral methods, determine the exact Green's function G(T) for propagation from time t = 0 to imaginary time t = —IT at fixed spatial point x = 0. G(T) is the analog of the 'correlator' for our quantum mechanical system. b) Prom the representation G{r) = (0, — ir|0,0), use completeness to express G{r) in terms of the S-wave radial wavefunctions {Rn(0)} evaluated at the origin and the energy eigenvalues {En}. What values of n contribute? This representation is the analog of the dispersion relation expression for a correlator in which one takes into account an infinity of resonances. c) Plot the negative logarithmic derivative —d[ln G(r)]/dr for the range 0 < uor < 5 and interpret the large uor behavior in terms of your result in part (b). d) Obtain the first three terms in a power series for —d[lnG(r)]/dr, expanded about r = 0. This is the analog of the series of operator product 'power corrections' to — d[lnG(r)}/dr. Assume, as is the case in QCD, that you know only a limited number of terms in this series, first two terms and then four terms. Is there a common range of UJT for which (i) your truncated series reasonably approximates the exact behavior, and (ii) the approximation for keeping just the lowest bound state in part (b) is likewise reasonable? It is this compromise between competing demands of the resonance and operator product approximations which must be satisfied in sucessfully applying the QCD sum rules to physical systems.
XII Baryon properties
An important sector of hadron phenomenology is associated with the electroweak interactions. Baryons provide a particularly rich source of information, with data on vector and axial-vector couplings, magnetic moments, and charge radii. In Sect. XII-1, we describe the procedure for computing matrix elements in the quark model, and then turn to a variety of applications in the succeeding sections. XII—1 Matrix element computations Much of the application of the quark model to physical systems involves the calculation of matrix elements. The subject divides naturally into two parts. On the one hand, many quantities of interest follow from just the flavor and spin content of the hadronic states. On the other, it is often necessary to have a detailed picture of the quark spatial wavefunction. Flavor and spin matrix elements For the first of these, the quark model is particularly appealing because of the intuitive physical picture which it provides. For example, consider the quark content of the proton state vector, which we reproduce here from Table XI-2,
= -^e^Kul^
- uX->!T] 1°) •
The first two quarks form a spin-zero, isospin-zero pair with the net spin and isospin of the proton being given by the final quark. The prefactor of I / A / 1 8 ensures that the state vector has unit normalization. Calculation reveals that one-third of the magnitude of this normalization factor comes from the u^u^ term and two-thirds from the u^u^di term, i.e. one concludes that 4the proton is twice as likely to be found in the configuration 313
314
XII Baryon properties
with the ?/-quark spins aligned than anti-aligned', (1.2) The 'six parts in eighteen' of the u^u^ configuration arises entirely from the six ways that color can be distributed among three distinct entities. The configuration u^u^dy is twice as large due to the presence of two u^ states. Similar kinds of inferences can be drawn for the remaining baryon state vectors in Table XI-2. We can proceed analogously in deriving and interpreting various matrix element relationships. It is instructive to work at first in the limit of 5/7(3) invariance because more predictions become available. The effect of symmetry breaking is addressed in Sect. XII-2. Let us consider matrix elements, taken between members of the spin-1/2 baryon octet, of the operators Squared charge-radius : / d3x r2ift^Qi/j
oc (Q) ,
Axial-vector current : / d3x ^7375X31P
oc (Xsaz) ,
Magnetic moment : / d3x - ( r x ip^cxQip)^
(1.3)
oc (Qcrz) .
Along with the definition of each operator is indicated the flavor/spin attribute of an individual quark which is being averaged over. For example, a magnetic moment is sensitive to the combination Qaz of each quark within the baryon. Matrix elements will then be products of such averages times quark wavefunction overlap integrals. The flavor/spin averages for the baryon octet are displayed in Table XII-1. To see how these values are arrived at, let us compute the value 5/3 obtained for the proton axial-vector matrix element. For the configuration u^u^di, which occurs with a probability of 2/3, the average value of XzCFz equals (1 + 1 + 1) x 2/3 = 2, whereas for the configuration u^u^ one finds (1 - 1 - 1) x 1/3 = - 1 / 3 . Together they sum to the value 5/3. Table X I I - 1 . Some baryon octet expectation values P
n
0 } 1 -2/3 (A3
a
1
A
£+
0 -1/3
1 1
2/V6 4/3
E~
0 l/3 a 0
-1/3 -4/3
0 -2/3 -1/3
The off-diagonal transition E° -• A has \(Qaz)\ = l/>/3.
-1 -1/3 1/3
XII-1 Matrix element computations
315
Overlaps of spatial wavefunctions The spatial description of quark wavefunctions is less well understood than the spin/flavor aspect of the phenomenology. The most extensive studies of the spatial wavefunctions are associated with matrix elements of currents. Because these are bilinear in quark fields and because of the wavefunction normalization condition, the magnitudes of these amplitudes are constrained to be nearly correct. Dimensional matrix elements are primarily governed by the radius of the bound state. As long as the proper value is fed into the calculation, the scale should come out right. As noted in Sect. XI-1, a relativistic quark moving in a spin independent central potential has a ground state wavefunction of the form e-iEt
,
(1.4)
gnd
where u, £ signify 'upper' and 'lower' components. For the bag model, these radial wavefunctions are just spherical Bessel functions. This form also appears in some relativistic harmonic oscillator models which use a central potential. To characterize different types of relativistic behavior, it is worthwhile to express matrix elements in terms of u and £ without specifying them in detail. The normalization condition for the spatial wavefunction is then ! dsx Vf(x)^(x) = f dsx (u2(r)+£2(r)) = 1 .
(1.5)
In the nonrelativistic regime, the lower component vanishes {£ = 0). Let us consider the size of the lower components which occur in various approaches. In the bag model one obtains for massless quarks the integrated value [ dsx£2(r)~0.26
.
(1.6)
Relativistic effects are often included in potential models by working in momentum space and employing the spinor appropriate for a quark q in momentum eigenstate p,
«(p) = yjE + mA
| .
a p
E + mq
(1.7)
X
In this case the relevant prescription is
I*'
(I 8)
-
316
XII Baryon properties
where the averaging is taken over the momentum space wavefunction of the particular model. Using the uncertainty principle relation of Eq. (XI1.14) to estimate (p 2), we find typical values P
2
x
~ 0.13 -> 0.20
(1.9)
for a confinement scale of 1 fm. Larger effects are found in the harmonic oscillator model if one uses the value a2 = 0.17 GeV2 (see Fig. XI2). Generally, the lower component is found to be significant but not dominant in quark wavefunctions. Connection to momentum eigenstates In all cases except for the nonrelativistic version of the harmonic oscillator model, one cannot explicitly separate out the center-of-mass motion. The result of a quark model description of a bound state is a configuration localized in coordinate space, i.e., a position eigenstate. However, the analysis of scattering and decay deals with the plane waves of momentum eigenstates. The basic assumption made in all quark models is that the bound state with a given set of quantum numbers is related to only those momentum eigenstates of the same type. If we denote |iJ(x)} as a unit normalized hadron state centered about point x and \H(p)) as a plane wave state, then we have
)) = J dsp ¥>(p)e^ x \H(p)) .
(1.10)
We shall give a prescription for obtaining a functional form for (p(p) shortly. Let us normalize the plane wave states for both mesons and baryons as (H(p')\H(p)) = 2cup(27r)3^3)(p' - p) .
(1.11)
The constraint of unit normalization then implies
We can employ the above wavepacket description to derive a general procedure within the quark model for calculating matrix elements [DoJ 80]. Many matrix elements of interest involve a local operator O evaluated between initial and final single-hadron states. Let us characterize the magnitude of the matrix element in terms of a constant g. Then, for baryons in the momentum basis, the spatial dependence is given by (B'(p') |OCr)| £(p)> = g u(p')r O n(p) &-'>* ,
(1.13)
XII-1 Matrix element computations
317
where To is a Dirac matrix appropriate for the operator O. By comparison, one obtains in any bound state quark model (QM) calculation a spatial dependence whose specific form is model dependent, f QM(B \O(x)\B)QM
= f(x) .
(1.14)
Hereafter, let us center all quark model states at the origin. The method of wavepackets then implies QU{B'\Jd
3
x O{x)\B)m=g J dzx J dzp'Sp
(2TT)3
(1.15)
|^(p)| 2 tZ(p)r o «(p) .
For sufficiently heavy bound states the fluctuation in squared momentum (p2) is small, and one may expand about |p| — 0, (1.16) n(p)r o u(p) = «(0)r o «(0) + O (( P 2 )/m|) . A common approach consists of keeping only the leading term to obtain
u(0)rou(0) -
QM (£'|
J dzx O{x)\B)m .
(1.17)
It is interesting to note that this relation, often thought of as fundamental, is in fact only an approximation. As an example, let us perform the complete quark model procedure for the neutron-proton axial-vector current matrix element. We begin by defining as usual
^
(1.18)
For spin-up nucleons the choice /i = 3 gives u(0,1)7375^(0, T) = 2mN ,
(1.19)
yielding for Eq. (1.17) the basic formula, 9A =
QM(P|I
/ d*x u(x)wsd(x)
|n T ) QM .
(1.20)
The field operator for any quark q is expanded as in Eq. (XI-1.1), Qa(x) =
Substituting, we have 9A =
/
d?>x
V>o,s'(x) 7375^o,s(^) ui(s)da(sf)\n^)QM
,
(1.22)
318
XII Baryon properties
where only the n = 0 ground state mode contributes. At this stage, one can factorize the spin and space components by using the general ground state wavefunction of Eq. (1.4). This leads to /
r I d x Xs(^ ^3 — ^ **3
=
d x ipQ s7375'0o s' —
CTo of
I
CL X
I ix v
Ks )
«
Q
and thus dsx (u2 - -i2) QM(PT ^(s.a)^ d(s',a) U^)QM (1-24) J o Finally, upon dealing with the spin dependence in Eq. (1.24), we obtain gA =
Any nonrelativistic quark model, having zero lower components, would simply yield QA — 5/3. If one desires to make relativistic corrections to such a model, the result can be inferred from the above general formula with the appropriate substitution of Eq. (1.8). Clearly, the procedure just given can be extended to matrix elements of any physical observable. The wavepacket formalism also allows for the estimation of the 'centerof-mass' correction. This arises from the (p2) modifications to Eq. (1.16). For the axial current, the zero momentum relation in Eq. (1.19) is extended for nonzero momentum to
where an average over the direction of p has been performed. This expression generalizes Eq. (1.25) to 1
9A\1-
3mnmp 2
3 rap
4
8mn 2
3mn\l
8mpJ\
=
5 f
SJ
3
/
2
\
_ 1
3 (1.27)
where (p )np — 0.5 GeV is a typical bag model value. It is possible to argue that in the transition from the current quarks of the QCD lagrangian to the constituent quarks of the quark model, the couplings to currents should be modified. For example, one might suspect that the coupling of a constituent quark to the axial current occurs not with strength unity, but with a strength g^ such that the nonrelativistic expectation is not g\ = 5/3 but rather g\ = 5gJ[ /3. The choice g£ ~ 3/4 would then yield the experimental value. This is not unreasonable, but if fully adopted, leads to a lack of predictivity. In such a picture, not only can the magnetic moments and weak couplings be
XII-1 Matrix element computations
319
renormalized, but also the spin and flavor structures. That is, in the 'dressing' process which a constituent quark undergoes, there could be 'sea'-quarks, such that the constituent ix-quark could have gluonic, dquark, or s-quark content. Likewise, some of the spin of the constituent quarks could be carried by gluons. One is then at a loss to know how to calculate matrix elements of currents. In practice, however, the naive quark model, with no rescaling of g& or of the magnetic moment, does a reasonable job of describing current matrix elements. It is then of interest to study both the structure and limitations of this simple approach. Calculations in the Skyrme model There are several differences between taking matrix elements in the quark model and in the Skyrme model. To begin, in the quark model a current is expressed as a bilinear covariant in the quark fields (c/. Eq. (1.3)), whereas in the Skyrme model the representation of a current is rather different. As an example, application of either Noether's theorem or the external source method of Sect. IV-6 identifies the SU(2) vector and axial-vector currents to be
±Tr where U = A(t)U$A~l(t) is the quantized skyrmion form and A(t) is an SU(2) matrix. We shall neglect derivatives of A(t), as the quantization hypothesis corresponds to slow rotations. This leads to a result similar in form to Eq. (1.28), but with U -* Uo and ra -» A' 1 (t)raA(t). The answer may be simplified by use of the explicit form of Uo appearing in Eq. (XI-4.15). Let us use Eq. (1.28) to compute the spatial integral of the axial current. After some algebra, we obtain a product of spatial and internal factors,
Jd sin2F 8sin 2 F ,, 8s
4sin 2 F sin2F
320
XII Baryon properties
where a is the isospin component and j is the Lorentz component. This is now suitable for taking matrix elements, such as )aj IPT) = I d°x I dtl3 (p T |A) (JA)aj a iQ j_y
i
i I J~\. J _L1 I /
./a/
X\
) LJ
~2'2
( L3 °)
1 -i I / I J ,
"2'2
where we have used the completeness relation of Eq. (XI-4.42). Upon expressing the trace in Eq. (1.30) as a rotation matrix, Tr (rkAr^A~1)/2 = D\J i we can determine the group integration in Eq. (1.30) in terms of SU{2) Clebsch-Gordan coefficients [Ed 60],
T>
— ( — \^{
V /
-T+m)
9
Z7r
o
s-iT'TT"
(1-31)
fiT'TT"
27"1" -I- 1 ^ m
^Jn
Alternatively, one can work directly with the collective coordinates, e.g. with the aid of Eqs. (XI-4.41-44) we obtain for a = j = 3 r
I
K
2
i
2
2
CLr\ ~r tto — Q-i —
2\/
i
•
\
CLn)[CL\ -\- 1CL2) ^
9
*r^ —^5
(A
•
Qo^
[l.oZj
Before Jone can infer a Skyrme model prediction for gA from this calculation, there is a subtlety not present for the quark calculation which must be addressed. Due to the original chirally invariant lagrangian, the Skyrme model is unique among phenomenological models in being completely compatible with the constraints of chiral symmetry. As a consequence, the near-static axial-vector matrix element is constrained to obey ?,-(p(p')IM?|p(p)) = 0 , and hence must be of the form [AdNW 83],
(q = p-p')
(p(p')\(JA)3MP)) = 2mp9A (Sjk ~ H * ) K ) •
(1.33a)
(1.33b)
The term containing |q|~2 arises from the pion pole, as will be discussed in Sect. XII-3 in connection with the Goldberger-Treiman relation. An angular average of Eq. (1.33b) then yields 2 ^ / 3 , which from comparison with Eq. (1.32) implies gA — G^. Thus in the Skyrme model, the axialvector coupling constant equals the radial integral in Eq. (1.29) which defines G5. Use of the profile given in Sect. XI-4 leads to the prediction gA — 0.61, which is about only one-half the experimental value and constitutes a well-known deficiency of skyrmion phenomenology. Presumably, consideration of a more general chiral lagrangian could modify this result by including higher derivative components in the weak current.
XII-2 Electroweak matrix elements
321
Pions may be added to the Skyrme description through introduction of the matrix £ previously seen in Sects. IV-7,VII-3 [Sc 84], U = £A(t)U0A-\t)ti,
£ = exp [ir • w/(2Fff)]
.
(1.34)
If currents are formed using this ansatz, some terms occur without derivatives on the pion field, while others contain one or more factors of d^n. Since d^n gives rise to a momentum factor q% when matrix elements are taken and soft-pion theorems deal with the limit <$—•(), the lowest order soft-pion contribution will consist of keeping only terms without derivatives. Thus in the process v^ + N —> N + ir + /i the final-state pion is produced by a hadronic weak current and the soft-pion theorem relates the N —> NTT matrix element to the N —> N current form factors. Expanding the currents to first order in the pion field yields
(J
x ^ = b^ [* ( rVrl M u°±w* uo)A) •
b
1
(1-35)
where for notational simplicity we have displayed only the first term in the current. Note the sign flip in the second line. This form is in accord with the soft-pion theorem Urn ( fabc e
(1.36) where the current commutation rules of Eq. (IV-5.14) have been used. XII—2 Electroweak matrix elements The static properties of baryons can be determined from their coupling to the weak and electromagnetic currents. In this section, we shall describe these features in terms of the quark model.
Magnetic moments The generic quark model assumption for the magnetic moment is that the individual quarks couple independently to a photon probe. For ground state baryons where all the quarks move in relative S-waves, the magnetic moment is thus the vector sum of the quark magnetic moments, (2-1)
322
XII Baryon properties
where (T{ is the Pauli matrix representing the spin state of the ith quark and fii is the magnitude of the quark magnetic moment.* Since the light hadrons contain three quark flavors, the most general fitting procedure to the moments of the baryon octet will involve the magnetic moments V>u,V>d, Ms-
It is straightforward to infer baryon magnetic moment predictions in the quark model directly from the state vectors of Table XI-2. For example, we have seen that the proton occurs in the two configurations u^u^d^ and iX|i/|d| with probabilities 2/3 and 1/3 respectively. This can be used to carry out the construction defined by Eq. (2.1) as follows: 2 1 3 v(uXu\dl) + 3 2 1 4 1 g g 3 3 (2.2) and similarly for the other baryons. Experimental and quark model values are displayed in Table XII-2. It is of interest to see how well the assumption of SU(3) symmetry fares. In the limit of degenerate quark mass (denoted by a superbar), the quark magnetic moments are proportional to the quark electric charges, fid = p8 = —zp*
(SU(3) limit) ,
(2.3a)
Table XII-2. Baryon magnetic moments Mode
ME+ I^SOAI A*E-
Ms-
Experiment 0 2.792847386(63) -1.91304275(45) -0.613(4) 2.42(5) 1.61(8) -1.16(2) -1.25(1) -0.69(4)
Quark Model
Fit A 6
Fit B c
V
2.79 -1.86 -0.93 2.79 1.61
2.79 -1.91 -0.61 2.67 1.63
-0.93 -1.86 -0.93
-1.09 -1.44 -0.49
4fJ,d-^JLu
3 3
3 S
3
S
3
a
Expressed in units of the nucleon magneton fij\f = eh/2Mp. SU(3) symmetric fit. c // u ,//^,/i s taken as independent parameters. b
When referring to the 'magnetic moment' of a quantum system, one means the maximum component along a quantization axis (often chosen as the 3-axis). Thus the magnetic moment is sensitive to the third component of quark spin as weighted by the quark magnetic moment.
XII-2 Electroweak matrix elements
323
while isospin symmetry would imply pd = -Ifa
(SU(2) limit) .
(2.3b)
If we determine the one free parameter by fitting to the very precisely known proton moment, we obtain the SU(3) symmetric Fit A shown in Table XII-2. More generally, allowing /iw, /i^, /i s to differ and determining them from the proton, neutron, and lambda moments yields jiu = 1.85 /i^v 5
ftd
— —0.972 /ijv
5
l^s — —0.613 fij\f ,
(2-4)
and leads to the improved (but not perfect) agreement of Fit B in Table XII-2. We see from Eq. (2.4) that the main effect of SU(3) breaking is to substantially reduce the magnetic moment of the strange quark relative to that of the down quark. The deviation of ^d/Hn from the isospin expectation of fid/l^u — —1/2 is smaller and perhaps not significant. Observe that the famous prediction of the SU(2) limit, fin/{\ip = —2/3, is very nearly satisfied. The magnetic moment as derived from the multipole expansion of the electric current is defined by r
3
xrxJem(x)
.
(2.5)
It follows from this expression that the contribution of a nonrelativistic quark V to the hadronic magnetic moment is just the Dirac result,
"• = 55 •
(2 6)
'
where Mq is the quark's constituent mass and Q is its charge. We can use this together with Eq. (2.4) to determine the constituent quark masses, with the result Mu~Md~
320 MeV ,
Ms ~ 510 MeV .
(2.7)
As we shall see in Sect. XIII-1, these masses are comparable to those extracted from mass spectra of the light hadrons. One can also construct models involving relativistic quarks. For these, the magnetic moment contribution of an individual quark becomes
/ / = -Q
(2.8)
Note the absence of an explicit dependence on quark mass. This is compensated by some appropriate dimensional quantity. The inverse radius i?" 1 plays this role in the bag model, and other determinations of R allow for a prediction of the hadronic magnetic moment. For example, the bag model defined by taking zero quark mass (corresponding to the ultrarelativistic limit) and R = 1 fm yields the value fip ~ 2.5 in a treatment
324
XII Baryon properties
which takes center-of-mass corrections into account [DoJ 80]. Although this specific value is somewhat too small, it is fair to say that quark models give a reasonable first approximation to baryon magnetic moments. Semileptonic matrix elements The most general form for the hadronic weak current in the transition Bi -»• B 2£ve is
where the {fi} and {gi} form factors correspond respectively to the vector and axial-vector current matrix elements, and q = p\ — pi is the momentum transfer.* The form factors are all functions of q2 and the phases are chosen so that each form factor is real-valued if time reversal invariance is respected. In practice, the form factors accompanying the two terms with the kinematical factor q^ are difficult to observe because each such contribution is multiplied by a (small) lepton mass upon being contracted with a leptonic weak current. Thus we shall drop these until Sect. XII-4. As regards the remaining form factors, we have already presented the ingredients for performing a quark model analysis. Using the n —+ p transition as a prototype, we have ft*> = (p T | f dsx Wyod |raT) = f dsx (uuud + ejd) = 1 , (2.10a) TTlp -\- fJlfi
J
Z
= lj^r{ujd 97 = (Ptl ff mn + m p
+( \2mn
d3x
+ udeu)
u^lsd
2mpJ
=l^
u +
^ ,
|nT) = || // d3x (uuud - ^^£Jd),
"P +2 g3"P = vni (m| - iJf
(2.10b) (2.10c)
d3x zu-yO-Kd |n T
i ( i i ) .
(2.10d)
Given the context of application, there should be no confusion between the QCD strong coupling 2 constant g% and the axial-vector form factor
XII-3 Symmetry properties and masses
325
In each case, we first give the defining relation, then the general Dirac wavefunction (c/. Eq. (1.4)), and finally the nonrelativistic quark model limit. The vanishing of g^p in the limit of exact isospin symmetry is a consequence of G-parity (cf. Eq. V-3.22)). Predictions for the other baryoiiic transitions are governed by 5/7(3) invariance, amended by small departures from SU(3) invariance as suggested by the quark model, i.e. s —> u transitions are similar to those of d —• u as given above, but with the down quark mass and wavefunction replaced by those of the strange quark. 5/7(3) breaking in the form factors arises from this difference in the wavefunction. As a quark gets heavier, its wavefunction is more concentrated near the origin and the lower component becomes less important. The form factors of the matrix element (JB&| Jc\Ba) evaluated in the 5/7(3) limit at q2 = 0 give for the vector current, /l(0) = ifabc , 1
/2(0) = ifabcf + dahcd , 3
(2.11a)
with f/d = 0.29, and for the axial-vector current, fli(0) = ifabcF + dabcD ,
(2.11b)
9
with F + D = g^ = QA = 1.26. In the above, the indices a,b,c— 1,...,8 label the 5/7(3) of flavor, with c = (1 + i2) for AS = 0 and c = 4 + i5 for AS = 1. There is no SU(3) parameterization for the g u transition. The wavefunction overlaps in g\ lead to a slight increase in the strength of the s —• u transition compared to d —> u because of the reduced lower component of the s quark. For #2, a nonzero but highly model dependent value is generated, typically of order 132/511 — 0-3.
XII-3 Symmetry properties and masses In our discussion of baryon properties, we have given priority to quark models because they are generally simple and have predictive power. However, symmetry methods are also useful, particularly when expressed in the language of chiral lagrangians. We shall combine the two descriptions in this section.
XII Baryon properties
326
Effective lagrangian for baryons We begin by writing effective lagrangians which include baryon fields, using the procedure described in Eqs. (IV-7.3), (VII-3.2). The lowestorder SU(2) invariant lagrangian describing the nucleon and its pionic couplings has the form - m0) N N
—N
~Y U=
= exp [ir •
(3-1)
where N — (^) is the nucleon field, in is the mass matrix for current quarks (with mu — rrid = m), Zo and Z\ are arbitrary constants which parameterize terms proportional to the quark mass matrix, and the constant QA is the nucleon axial-vector coupling constant QA — 1.26 (c/. Prob. XII1). The mass parameter mo represents the nucleon mass in the SU{2) chiral limit. For the full SU(3) octet of baryons, the analog of W is A
E+ A
E-
B = \
1—l
p n
(3.2)
2A
k
"V6
—'
where the phases have been adjusted to match our quark model phase convention of Eq. (XI-1.8). The SU(3) version of Eq. (3.1) becomes - F
- D
, B]))
(3-3)
where the covariant derivative is now T>^B = d^B + ifV^ji?], ^ is the 5C/(3) generalization of the quantity in Eq. (3.1) with r replaced by A, m is the diagonal SU(3) quark mass matrix, m
— 7=(m-ms)\$
rh m.
v3
(3.4)
XII-3 Symmetry properties and masses
327
and mo is the degenerate baryon mass in the SU(3) chiral limit. Consistency of the SU(2) and SU(3) lagrangians requires D + F = gA , dm + f m = 1 , mo = fho + Z\ms - Zoms(fm - dm) . The description thus far is based on symmetry. It includes quark mass, but not higher powers of derivatives. Baryon mass splittings and quark masses The various parameters (m, m s , ZQ etc.) appearing in the chiral lagrangians of Eqs. (3.1), (3.3) can be determined from baryon mass and scattering data. In the nonstrange sector, the nucleon mass is given in the notation of Eq. (3.1) as mN = m0 + (Zo + 2Zi)m .
(3.6)
To isolate the effect of the nonstrange quark mass rh and of the constants Zo, Zi, it will prove useful to define a quantity a, rh(Zo
+
2Z1) .
(3.7)
Shortly, we shall see how this quantity can be determined from pionnucleon scattering data. However, let us first consider the baryonic mass splittings generated by the mass difference ms — rh. Upon using Eq. (3.3) to obtain expressions for the baryon masses and working with isospin-averaged masses, it is possible by adopting the numerical values Z0(ms - rh) = 132 MeV ,
dm/fm = -0.31 ,
(3.8)
to obtain the following good fit: mx-mN
= (fm - dm)Z0(ms - rh) = 251 MeV (Expt. : 254.2 MeV), 4 mx - mA = —d m Zo(m s - rh) = 79MeV (Expt. : 77.5Mev),
mr=-mN = 2fmZ0(ms - rh) = 383 MeV
(Expt. : 379.2MeV). (3.9) Observe that these mass-splittings depend on Zo but not on Z\. The three relations of Eq. (3.9) imply the Gell-Mann-Okubo formula, m s -mN
= -(mE-
mN) + -(m E - mA)
(Expt. : 254 MeV = 248 MeV ) . which displays an impressive level of agreement (c± 3%) with experimental values.
328
XII Baryon properties
The above analysis, based on a chiral lagrangian, can be enhanced by using ideas taken from the quark model. In the limit of noninteracting quarks, the quark model yields for a general spatial wavefunction,* m\ — rriN = m^ — m^ = THE — ra^ — (m
s
— rh) I Sx (u2 — i2) . (3.11)
However, observe that ra^ = m\ (corresponding in the chiral lagrangian description to dm = 0) for noninteracting quarks. Of course, the actual A and S baryons are not degenerate, so additional physics is required. A quark model source of the A — S mass splitting lies in the hyperfine interaction of Eq. (XI-2.14),
where the prefactor of 1/2 is associated with the color dependence of Eq. (XI-2.4). Matrix elements of this operator give rise to the additive mass contributions, m
mjsf = . . . — - W m i ?
1 8
A = . . . — Q^nn ,
1 2
"
1 2
(3 13)
1
'
8 2
where H^ and TYy are related by Hy = 7^ij|^(O)| and the subscripts 'n', V denote an interaction involving a nonstrange quark and a strange quark respectively. For TLnn ^ 7Yns, the S and A will not be degenerate. Treating both quark mass splittings and hyperfine effects as first-order perturbations (e.g. Hss — Hns = Hns — ^nn), one obtains quark model mass relations = (ms -rh) I dsx (u2 - £2) , = l(Hnn-Hns)
,
(3.14)
m* - mN = -(Wnn - Wns) + 2(ms -rh) d3x (u2 - f) " 4 J in accord with the sum rule of Eq. (3.10). These formulae can provide an estimate of quark mass. For the usual range of quark model wavefunctions (encompassing both bag and potential descriptions), the overlap integral One could equivalently use the language of the potential model, where these baryon mass splittings arise from the constituent quark mass difference Ms — M.
XII-3 Symmetry properties and masses
329
has magnitude r
*x{u2-l2)~\^-A
.
(3.15)
To the extent that this estimate is valid, it produces the values ms - m ~ 230 -> 350 MeV ,
m ~ 11 -> 14MeV ,
(3.16)
where the chiral symmetry mass ratio of Eq. (VII-1.6a) has been used to obtain m. In general, quoting absolute values of quark masses is dangerous as one must specify how the operator qq which occurs in the mass term mqqq has been renormalized. It is all too common in the literature to ignore this point by using ms — m = m\ — TUN- The values quoted here are actually current-quark mass differences, renormalized at a hadronic scale using quark model matrix elements. The parameter Z\ which appears in the SU(3) lagrangian of Eq. (3.3) is difficult to constrain in a quark model. For example, one might consider the matrix element (N\msss\N)
= ms \Z\ — Zo(jm — dm)) .
(^-17)
The most naive assumption, that (N \ms'ss\ N) vanishes, would imply Z\ = Zo(fm — dm) ~ 1.9 ZQ. However, one may legitimately question whether such an assumption is reasonable. We shall return to the issue of the 'strangeness content' of the nucleon later in this section. Goldberger-Treiman relation Moving from the study of baryon masses to the topic of interactions, let us consider the coupling of pions and nucleons. The SU(2) lagrangian of Eq. (3.1), expanded to order TT2, becomes CN = N(ip — rriN)N + —N^j^—N 1
~" "
• C^TT
7T x d^K N + -^7T2NNa
+ ... ,
where a is defined in Eq. (3.7). The second term describes the NNn vertex. Upon using Eq. (3.18) to compute the pion emission amplitude N —> NTT1 and comparing with the Lorentz invariant form MN^N7ri
= -ignNNuip'hsr'uip)
,
(3.19)
one immediately obtains the Goldberger-Treiman relation [GoT 58], 9AmN —^— • (3.20)
330
XII Baryon properties
Inserting the experimental value, ff^v/^71" ~ 13.5, for the nNN coupling constant, one finds the Goldberger-Treiman relation to be satisfied to about 5%. There also exist important implications for the g% term in the general expression given in Eq. (2.9) for the axial-current matrix element. In forming the n —• p axial matrix element, one encounters a direct 7^75 contribution and also a pion-pole term which corresponds to pion propagation from the n —> pn~ emission vertex to the axial-current. Making use of Eq. (3.20) and Prob. XII-1, we have gA
«(p)
2™>N9A
= u(pf)
!li
H
(3.21)
1 'K
J
where q = p — p1'. It is this induced pseudoscalar modification which allows the axial-current to be conserved in the chiral limit ra^ —» 0, = 2mN9A \l L /
2
^
J u(p')75«(p)
mA
(3.22)
,2 _
Note that for nonzero pion mass, the above is consistent with the PC AC relation of Eq. (IV-5.7), ; , as both sides have the same matrix element, -t
(3.23) i
q
^ m^
2mNgAml_ , g 2 _ m 2 "(P)75«(P) •
(3.24) The pion-pole contribution of the axial-vector current matrix element has been probed in nuclear muon capture, as will be described in Sect. XII-4. The nucleon sigma term One of the features immediately apparent from the effective lagrangian of Eq. (3.1) is that all the couplings of pions to nucleons, with the exception of the quark mass terms, are derivative couplings. Before turning to the sigma term, which appears in the nonderivative sector, let us briefly consider the expansion in powers of the number of derivatives for pionnucleon scattering. Recall for pion-pion scattering (c/. Sect. VI-4), there
XII-3 Symmetry properties and masses
331
were no large masses and the chiral expansion was expressed in terms of m\ or E\. However, correction terms in the chiral expansion for nucleons will enter at relatively low energies since a term like 2p - q ~ 2mpE7r can get large quickly (it is linear in the energy and has a large coefficient, e.g. En = 250 MeV gives 2mpEn = (700 MeV)2). To combat this difficulty, additional (but still general) inputs such as analyticity and crossing symmetry are often invoked. Fortified with these theoretical constraints, one then matches intermediate energy data to the low energy chiral parameterizations. The low energy chiral results thereby obtained appear to be well satisfied [Ho 83, GaSS 91]. The nonderivative pion-nucleon coupling coming from the quark mass terms in Eq. (3.1) is of particular interest. To determine this contribution from experiment, one must be able to suppress the various derivative couplings. Thus, if one extrapolated in the chiral limit to zero fourmomentum, the derivative couplings would vanish. Not surprisingly then, a soft-pion analysis reveals that the nonderivative coupling can be isolated by extrapolating the isospin-even nN scattering amplitude with Born term subtracted (called D+ in the literature) to the so-called 'ChengDashen point' t = ra2, s — m2N [ChD 71]. It is conventional to multiply the extrapolated amplitude by F% and thus define a quantity S, X = F%D+D
.
(3.25)
To lowest order in the chiral expansion, the measured quantity £ is just the matrix element a defined in Eq. (3.7), 2mN
v
}
It is this isospin-even scattering amplitude D+ which provides a unique window on the nonstrange quark mass ra. Because £ is proportional to the small mass m, it is difficult to determine this quantity precisely, and considerable effort has gone into its extraction. The Cheng-Dashen point lies outside the physical kinematic region, and extrapolation from the experimental region must be done carefully with dispersion relations. Present estimates are f 64 ± 8 MeV S = I [60 MeV
[Ko 82] ,
[
[GaLS90] .
(3.27)
The result a = £ — 15 MeV has been obtained from studies of higher order chiral corrections, implying
(3.28)
332
XII Baryon properties Strangeness in the nucleon
In light of the above discussion, it is tempting to interpret various contributions to the nucleon mass by making use of the energy-momentum trace. Recall the trace anomaly of Eq. (Ill—4.16), muuu + mddd + msss .
(3.29)
Taking the nucleon matrix element gives = TUQ + (7 ,
23
N) - 8 9 4 ± 8 MeV , (3.30)
'
. (N\uu + dd\N) ! a — m-— ! —- ~ 45 MeV This result is already quite interesting in that the largest contributions, the gluon and strange quark terms in mo, appear to be cnon-valence'. At this stage, the separation is essentially model independent. One can learn something about the 'strangeness content of the nucleon' by using an SU(3) analysis of hyperon masses. Thus, we introduce a mass-splitting operator which transforms as the eighth component of an octet, Cm-S —-(m — ms)(uu-\-dd— 2ss) . (3.31) o Since the hyperon mass splittings are governed by this octet operator, we find (p\(ms-m)(uu + dd-2ss)\p) 3 ^ M V ! 6S = — ! — = -(m E - mN) = 574MeV . Zrrip
1
(3.32a) When scaled by the quark mass ratio m/ms, Eq. (3.32a) becomes ^ (N\uu + dd-2ss\N) ! ! o=m 2mN , ( (6.62b) o = o l 9 * 0(m'=-mA)~25MeV (35MeV) , 2m K-m\ where the figure in parenthese includes higher order chiral corrections [Ga 87]. Comparison of 6 and a immediately indicates that they are compatible only if the strange quark matrix element does not vanish. Indeed, one requires (N\ss\N) -0.18 (0.09) . {N\uu + dd + ss\N)
(3.33)
XII-3 Symmetry properties and masses
333
This gives for the constant Z\ of Eq. (3.1) the value Z\ ~ 3.9Z0 (2.9 Zo) to be contrasted with the estimate which follows Eq. (3.17). At the same time, one can separate out the following matrix elements (2mN)-l(N
PQCD pg
N) - 6 3 4 MeV (764 MeV) ,
(3.34)
1
(2mN)- (N \msss\ N) ~ 260 MeV (130 MeV) , where figures in brackets use the corresponding bracketed quantity in Eq. (3.32b). Note the surprisingly large effect of the strange quarks. These results are not without controversy as they rely on the use of SU(3) symmetry. However, the difference between a and the SU(3) value of 6 is large enough that some ss contribution is likely to be required in any case. This analysis does not go well with the naive interpretation of the quark model as embodied, for example, in the proton state vector formula which began this chapter. However, it is not incompatible with a more sophisticated interpretation of the constituent quarks which enter into quark models. In the process of forming a constituent quark, the quark is 'dressed' by gluonic and even ~ss quark fields. It is no longer the naive object that occurs in the QCD lagrangian. It is this dressed object which then may easily generate gluonic and perhaps strange quark matrix elements. Recall that even the vacuum state has gluonic and quark matrix elements. Similar explanations exist in bag and Skyrme models [DoN 86]. Further work is clearly needed to decide on the proper decomposition of the nucleon mass and to understand its significance. Quarks and their spins in baryons The quark model provides a most simple picture of the contents of baryons as systems composed of three constituent quarks and nothing else. Although a description using the quark and gluon degrees of freedom which appear in the fundmental lagrangian may be more complicated, it is nevertheless instructive to explore the constituent picture of the spin structure of the nucleons. For any Lorentz invariant theory, Noether's theorem requires that there be an angular momentum tensor M^a^ which is conserved (d^M^P = 0) and which gives rise to three angular momentum charges associated with rotational invariance, = f dsxMOaP(x)
.
(3.35)
334
XII Baryon properties
In the rest frame of a particle, the {J Q ^} are related to the three components of angular momenta as j
For the example of a free fermion, the above quantities take the form up to total derivatives which do not contribute to the charges, and
J = I dsx \-i^
(x x d) $ - ^ 7 5 J = L + S .
(3.38)
The two contributions in Eq. (3.38) may be labeled the orbital and spin components of the angular momentum. The quarks in the Noether current are lagrangian (current) quarks, not constituent quarks. Nevertheless, in the spirit of the quark model let us apply Eq. (3.38) to the quarks in a spin-up proton. As expressed in terms of upper (u) and lower (£) components (cf. Eq. (XI-1.13)), the orbital and spin contributions are found to be
(L) = 1 1 d*x /V> ,
(S) = / d*x (u2 - ^ 2 ) M .
(3.39)
Aside from the factor 1/2 occurring in a/2, the quark spin contribution to S is just the axial-vector matrix element of Eq. (1.24), whereas the orbital angular momentum contains just the lower component £ because the x x d operator has a nonzero effect only when acting on the a • x factor in the lower component of Eq. (XI-1.13). Observe that the orbital angular momentum is nonvanishing and proportional to the quark spin. The spin and orbital portions for the individual u, d flavors are easily computed to yield
=-\ f
(3.40) A first lesson is that, despite the spin wavefunction of the protons being written entirely in terms of quarks as in Table XI-2, the quark spin averages of Eq. (3.40) do not add up to yield the proton spin. The sum is reduced from the anticipated value of 1/2 by the lower component £ in the Dirac spinor. It is the total angular momentum J which has the expected result, <J> = ^<0-> ,
(3-41)
XII-3 Symmetry properties and masses
335
but the total is split up between the orbital and spin components. Because of the identification of the spin with the axial current, there is data on spin contributions of the individual quark flavors. Thus from nucleon beta decay we have a combination which transforms as the third component of an octet,
With the use of SU(3) symmetry, the eighth component of the same octet can be inferred from hyperon decay data, <5 W
+ sW - 2fl<')> = ^ • ^
§
* \ (0.57 ± 0.05) ,
(3.43)
where we use (3F — D)/(F + D) ~ 0.45 ± 0.04 as implied by empirical fits.* Some caution is required because SU(3) breaking is both expected and observed in fits to hyperon data. The quoted error bar includes our estimate of the uncertainty due to SU(3) breaking, where the purely experimental uncertainty is about half the quoted error bar. If the strange quark matrix element were zero, Eq. (3.43) would be the total spin component and one could extract the content of each separate quark spin. However, other data appear to contradict this assumption. The weak neutral current contains an isoscalar axial component [CoWZ 78] associated with nonvalence quarks, f
(3.44)
Neutral current neutrino elastic scattering experiments give an indication of a possible signal of this operator [Ah et al 87],
(p(P,A')l4/=0)|p(p,A)) =gou(pA'h^5u(p,X)
,
(3.45)
with 50 = 0.15 ±0.09. There are also results from polarized deep inelastic experiments [As et al. 88] which are sensitive to the matrix element*
and which have thus been interpreted as probing quark spin,
(SM + S^ + Si\2
= \x(g2) .
(3.46)
(3.47)
While the experimental error bar is such that the result S(g 2) = 0.01 ± 0.29 at the q2 range of the experiment* is only two standard devia* The quark model predicts 0.50 for this ratio in the SU(S) limit. t Here T,(q2) should not be confused with the use of E in the cr-term discussion, Eq. (3.25). * Because of the axial anomaly, the current in Eq. (3.46) is not conserved, which implies a QCDinduced dependence on q2. Some of the issues in this extraction are discussed in [JaM 90].
336
XII Baryon properties
tions from the result of Eq. (3.43) together with the simplest assumption (Sz ) ^ 0, the central value of this measurement has aroused substantial interest since it would imply that the spin component of light quarks is nearly vanishing. This latter result along with aspects of the analysis remains controversial, and hopefully will be clarified in the near future.
XII-4 Nuclear weak processes One area in which the structure of the weak hadronic current has received a great deal of attention is that of nuclear beta decay and muon capture. Although in some sense this represents simply a nuclear modification of the basic weak transitions n —> p e~~ ve , p —> n e+ ve, the use of nuclei allows specific features to be accented by the choice of levels possessing particular spins and/or parities [Ho 89]. Here, we shall confine our attention to allowed decays (A J = 0, ±1, no parity change) and will emphasize those aspects which stress the structure of the weak current rather than that of the nucleus itself. In particular, nuclear beta decay provides the best determination of Vu^ while muon capture provides the only measurement of the pseudoscalar axial weak form factor predicted by chiral symmetry. Measurement of Vud There are many occurrences in nuclei of an isotriplet of Jp — 0 + states. Examples are found with A = 10,14,26,34,42, Because coulombic effects raise the mass of the proton-rich /3=1 state with respect to that with /3=0, the positron emission process N\(IZ = 1) —• N2(IZ — 0) + e + + ve can occur. These transitions are particularly clean theoretically, and this is the reason why they are important. Since the transition is 0+ —> 0 + , only the vector current is involved, and because of the lack of spin there can be no weak magnetic form factor. The vector current matrix element involves but a single form factor a(g 2), (N2(P2)\Vli\N1(Pl))
= a(
•
(4.1)
This form factor is known at q2 = 0 because the charged vector weak current V^ is just the isospin rotation of the electromagnetic current, [I-,JU = d-fu
•
(4.2)
This relation is often called the conserved vector current hypothesis or CVC, and requires for each of the 0 + —• 0 + transitions, a(0) = y/2 .
(4.3)
XII-4 Nuclear weak processes
337
Table XII-3. Energy release and Tt\j2 values for 0 + -•+ 0+ Fermi decays. Nucleus
£ 0 (KeV)
.Fix/2 (sec)
14Q
1868.44(27) 3209.95(25) 4470.27(18) 5020.49(56) 5403.02(28) 6028.62(69) 6610.01(41) 7220.14(32)
3078.7(3.6) 3071.7(3.4) 3081.4(4.4) 3073.7(5.1) 3081.1(7.1) 3081.6(4.0) 3075.6(5.3) 3077.1(4.2)
26m^ 34
C1
38m j^" 42
Sc
46y 50
Mn Co
54
What is generally quoted for such decays is the Tt\j2 value, essentially the half-life ti/2 multiplied by the (kinematic) phase space factor T [WiM 72]. Theoretically, one expects a universal form
which should be identical for each isotriplet transition. G^ is the weak decay constant measured in muon decay while the logarithmic correction arises from 'hard' photon corrections, as discussed in Chap. VI. The 'soft' photon piece as well as finite-size and Coulombic corrections are contained in the phase space factor T. Much careful experimental and theoretical study has been given to this problem, and the current situation is summarized in Table XII-3 where the experimental Tt\j2 values are tabulated. A fit produces the value Tt\j2 — 3077.3 ± 1 . 5 sec with chi-squared per degree of freedom %2/u = 0.8. This excellent agreement over a wide range of Z-values is evidence that soft photon corrections are under control. Comparison of the experimental Tt\j2 value with the theoretical expression given in Eq. (4.4) yields the determination Vud = 0.9744 ± 0.0010 ,
(4.5)
which makes Vud the most precisely measured component of the KM matrix. The pseudoscalar axial form factor Chiral symmetry predicts a rather striking result for the form factor gs(q2) of Eq. (2.9), namely that it is determined by the pion pole with a coupling
338
XII Baryon properties
fixed by the PC AC condition. One cannot detect this term in either neutron or nuclear beta decay because when the full matrix element is taken, one obtains 53
(4.6)
v(P)
2mN
which is proportional to the electron mass and is thus too small to be seen (effects in the spectra are O(m2/mjsfEe) « 1). However, in the muon capture process /i~p —• i^n, the corresponding effect is 0(171^/171^) ~ 10%. Thus muon capture is a feasible arena in which to study the chiral symmetry prediction. The drawback in this case is that typically one has available from experiment only a single number, the capture rate. In order to interpret such experiments, one needs to know the value of each nuclear form factor at q2 ~ — 0.9 raj;, which introduces some uncertainty since these quantities are determined in beta decay only at q 2 ~ 0. Nevertheless, predicted and experimental capture rates are generally in good agreement provided one assumes (i) the q2 ~ 0 value of form factors from the analogous /3-decay, (ii) q2 dependence of form factors from CVC and electron scattering results, (iii) the CVC value for the weak magnetic term /2, and iv) the PC AC value of Eq. (3.14) for #3. The results are summarized in Table XII-4. Obviously, agreement is good except for 6 Li, for which the origin of the discrepancy is unknown, although it has been speculated that perhaps the spin mixture is not statistical. Table XII-4. Muon capture rates. Reaction
\T
3
He ->
J/M
3
H
Theory (103 sec"1)
Experiment (103 sec"1)
0. 664 ± 0.020a 0. 506 ± 0.0206
0.651 ± 0.057 0.515 ± 0.085 0.464 ± 0.042
1.510 ± 0.040
1.410 ± 0.140 1.505 ± 0.048 1.465 ± 0.067
/*- 6 Li -+ i/M 6 He
0.98 ±0.15
u- i a C _
7.01 ±0.16
w
i2B
012
6.2 ± 0.3 6.7 ± 0.9 -0.75
XII-5 Hyperon semileptonic decay
339
Before proceeding, we should emphasize one relevant point. When PC AC is applied, it is for the nucleon 2mNgi(q2) -
2
^9S(Q2)
= 2F^NN{q2)
2 \ ~1
/
[l - ^J
.
(4.7)
Then at q2 = 0, we have 1.26 = gi(0) ~ F^NN^> = 1.33 (4.8) m which is the Goldberger-Treiman relation. On the other hand, taking similar q2 dependence for gi(q2) and g-KNN{(l2), we find mM <73(-0.9m2) m2 + 0.9m2 "
'
v
J
PC AC is generally applied in nuclei in the context of a simple impulse approximation, and it is this version of PC AC which is tested by the muon capture rates listed in Table XII-4. The direct application of PC AC in nuclei cannot generally be utilized since the pion couplings are unknown. In the case of muon capture on 12C, additional experimental data is available. One class of experiment involves measurement of the polarization of the recoiling 12B nucleus. Combining this measurement with that of the total capture rate yields a separate test of CVC as well as of PC AC. The results ™
^ ~
™ = 8.0 ±3.0 ,
(4.10)
are in good agreement with both symmetry assumptions. In addition, one can measure the average and longitudinal recoil polarizarions in the 12C muon capture, yielding a value for the induced pseudoscalar coupling (-0.9m2) — " which is again in good agreement with PC AC.
'
v~">
XII-5 Hyperon semileptonic decay The goals in studying hyperon semileptonic processes are to confirm the value of V^ obtained in kaon decay and to use the form factors to better understand hadronic structure. These two goals are interconnected. In earlier days when data was not very precise, fits to hyperon decays were made under the assumption of perfect SU{3) invariance in order to extract V^. Presently, the experiments are precise enough that exact
340
XII Baryon properties
SU(3) no longer provides an acceptable fit. The desire to learn about Vus is thus impacted by the need to understand the SU(3) breaking. We have already described in Sect. XII-1 the physics ingredients which lead to SU(3) breaking within a simple quark description. These include recoil or center-of-mass corrections, wavefunction mismatch (in which a normalization condition realized in the symmetry limit no longer holds), and generation of the axial form factor #2- For hyperons, because of the presence of the axial current, SU(3) breaking can occur in first order. This means that hyperon decays are more difficult to use for determining V^s than are kaon decays, where the Ademollo-Gatto theorem reduces the amount of symmetry breaking. Thus at the moment it is probably best to use the value of Vus determined from kaon decay, and require that hyperon decays yield a consistent value. The clearest evidence on SU(3) breaking comes from the S~ —• Ae~i>e rate. Since this is a AS — 0 process, V^ does not enter, and in addition, the vector current matrix element must vanish. Thus the rate is determined by the axial-current contribution alone, for which the theoretical prediction is
where p is a 517(3) breaking factor due to the center-of-mass effect. A bag model estimate yields p — 0.939. Taking p = 1, the best SU(3) symmetric fit to all the data [Ro 90] would require D/(D + F) = 0.629 ± 0.001, and hence gf ~A = 0.647. On the other hand, the data on E~ —> Aez/e requires A = 0.591 ± 0.014, which implies the correction p = 0.914 ± 0.022. gf~ There seems to be no way to avoid this need for 517(3) breaking. The full pattern of SU(3) breaking is more difficult to uncover. One problem is experimental. When the g\ values are extracted from the data, they have generally been analyzed under the assumptions that the /i and J2 form factors have exactly their SU(3) values and that g2 = 0. If these assumptions are not correct, then the values cited in [RPP 90] do not reflect the true g\ but rather some combination of g\, j \ , fa and #2- The correlation with g Ae ^ e , seems likely incorrect as it fits so badly with the remaining patterns. Discarding it, the remaining data can be fit well either by the center-of-mass correction described above, with g2 = 0, or by the full corrections including wavefunction mismatch, with 92/gi = 0.20 d= 0.07 in A —• peve. Without an independent measurement of #2 one cannot decide between these. We note however, that either
XII-6 Nonleptonic decay
341
option yields a value of Vus consistent with that found in kaon decays, Vus = 0.220 ± 0.004 .
(5.2)
XII-6 Nonleptonic decay The dominant decays of hyperons are the nonleptonic B —• Bfn modes. Because of the spin of the baryons and the many decay modes available, the nonleptonic hyperon decays present a richer opportunity for study than do the nonleptonic kaon decays. Phenomenology The B —• B'n matrix elements can be written in the form MB^B'* = u(p')[A + B>K]u{p) ,
(6.1)
with parity-violating (A) and parity-conserving (B) amplitudes. Watson's theorem implies that if CP is conserved, the phase of these amplitudes is given by the strong B'TT scattering phase shifts in the final state S-wave (for A) or P-wave (for B), i.e. A = A0 exp (itff/?r) ,
B = Bo exp (i6%,n) ,
(6.2)
with Ao, Bo real (if CP is conserved). Aside from the nN system, these phase shifts are not known precisely, but are estimated to be ~ 10° in magnitude. The decay rate is expressed in terms of the partial wave amplitudes by \g\(E' where q is the pion momentum in the parent rest frame and we define B = (E1 — TUB'/Ef + rriBf)l^B. Additional observables are the decay distribution W(0), W(9) = 1 + aPB • pB, ,
a= " j ; v ^
~'
,
(6.4)
and the polarization (P#') of the final-state baryon, (a + PB • i>B')PB' + 0 (PB x pBf) + 7 [PB' X (Pfi x PB')] W(9) B) =
V
p
'
(
}
where P # is the polarization of B and p#/ is a unit vector in the direction of motion of Bf. Experimental studies of these distributions lead to the amplitudes listed in Table XII-5.
342
XII Baryon properties Table XII-5. Hyperon decay amplitudes" A amplitudes Thy6
Mode
Expt
A —•
pn~
A —•
UTT0
3.38 3.25 -2.37 -2.39 0.00 0.13 -3.27 -3.18 4.27 4.50 3.14 3.43 -4.51 -4.45
£ + - > 717T+ E + ^j97T 0 S ~ —• 7i7T~ ^ 0 - > ATT° S ~ —• A?r~
B amplitudes Expt
Thy
22.1 23.0 -15.8 -16.0 42.2 4.3 26.6 10.0 -1.44 -10.0 -12.3 3.3 16.6 -4.7
°In unitsof 10- 7 b Lowest-order chiral fit.
The nonleptonic amplitudes may be decomposed into isospin components in a notation where superscripts refer to A / = 1/2, 3/2,
AA_^- = V2A? - Af ,
^ — = f + 43)
(6.6) and X% is of mixed symmetry. Similar relations hold for the B amplitudes. Prom the entries in Table XII-5, it is not hard to see that the A / = 1/2 rule, described previously for kaon decays, is also present here. Table XII-6 illustrates that the dominance of A / = 1/2 amplitudes compared to those with A / = 3/2 holds in the six possible tests in 5-wave and P-wave hyperon decay, at about the same level (several per cent) as occurs in kaon decay.* Thus the A / = 1/2 rule is not an accident of kaon physics, but is rather a universal feature of nonleptonic decays. This makes the failure to clearly understand it all the more frustrating. The assumption that the dominant A / = 1 / 2 hamiltonian is a member of an SU(3) octet leads to an additional formula, called the Lee-Sugawara relation, ^Imo=2AE-^-+AA^FK-
,
(6.7)
* For P-waves, the observed smallness of # £ - _ > n 7 r - indicates that 2?^ is small, presumably accidentally so. In this case the measure of A/ = 3/2 to A/ = 1/2 effects is given by By, ' /Xy^ •
XII-6 Nonleptonic decay
343
which also is well satisfied by the data. In this case, the corresponding formula for the B amplitudes is not a symmetry prediction [MaRR 69], although for unknown reasons it is in qualitative accord there also. Lowest-order chiral analysis Chiral symmetry provides a description of hyperon nonleptonic decay which is of mixed success when truncated at lowest-order in the energy expansion. Given our comiii@nts on the convergence of the energy expansion for baryons made in Sect. XII-3, the need for corrections to the lowest-order results is not surprising. We shall present the lowest-order analysis here, as it forms the starting point for most theoretical analyses. Recalling from Sect. IV-7 the procedure for adding baryons to the chiral analysis, one finds that the two following nonderivative lagrangians have the chiral (8^, 1R) transformation property:
=DbTr ( B 7 5 {&A6£,B}) + FbTV ( [ % ] ) where £,i? are defined in Eqs. (3.1), (3.2) respectively. However, the operator C(P) w must vanish, as it has the wrong transformation property under CP [LeS 61]. That is, a CP transformation implies B -• (i7 2B)
,
£ -* UTJ
,
(6.9)
and including the anticommutation of B and B, Cw is seen to return to itself, but C\y changes sign and hence must vanish. This leaves C\^ as (S)
the only allowed chiral lagrangian at lowest order. Observe that CKW} lacks a 75 factor. Thus its B —> B'TT matrix elements will be parity-violating, leading to only A amplitudes. The parity-conserving B amplitudes in Table XII-6. Ratio of A/ = 3/2,1/2 amplitudes 5-wave P-wave A 0.014 0.006 E -0.017 -0.047 S 0.034 0.023
344
XII Baryon properties TT/
Hw
T7 /
gfl Hv
Fig. XII-1 P-wave hyperon decay amplitudes
B —• B'TT are produced through pole diagrams as in Fig. XII-1, and are proportional to the parity-conserving B —> B1 matrix elements of Dw'. The counting of powers of energy (momentum transfer) in the energy expansion goes as follows. Both the B —> Bf transition and the A amplitudes in B —• B'n are obtained as matrix elements of C\^, which is zeroth order in the energy. The pole diagrams are likewise of zeroth order in the energy, being the product of the C\y vertex (0(1)), a baryon propagator (O(q~1)) and an NNn vertex (O(q)). Since the kinematic part of the pole diagrams, u'^u ~ cr • q, is of first order in g, the B amplitudes themselves are of order B ~ q~l ~ I/Am for the baryon pole. Kaon poles and higher-order chiral lagrangians enter at next order, i.e. having one power of the momentum transfer. The lowest-order chiral SU(3) analysis provides a fit to the data in terms of two parameters, called F and D,
with other amplitudes being predicted by the A/ = 1/2 rule. Use of the numerical values ^ = -0.42,
^ - = 0.92xl(T 7 ,
(6.11)
leads to the excellent fit of the S-wave amplitudes seen in Table XII5. Note that this form has one less free parameter than the general SU(3) structure [MaRR 69]. Thus the prediction of chiral symmetry that ^4x;+^n7r+ — 0 is independent of the D/F ratio (up to A/ = 3/2 effects), and represents a successful explanation of the smallness of this amplitude. In principle, the A amplitudes, together with the strong BB'n vertices, determine the baryon pole contributions to the B amplitudes. There should be no ambiguity since the strong vertices follow from the effective lagrangian of Eq. (2.1) already given for the axial-current matrix elements. These are then parametrized by the same d/f ratio as in the
I* 6 Nonleptonic decay
345
axial current,* e.g. 2/ J
-
Using this parameterization for the pole diagrams, one finds contributions such as w
2mNF7r
Taking d + f = 1.26, d/f — 1.8, one obtains from relations like this the disappointing P-wave predictions quoted in Table XII-6. This failure to simultaneously fit the 5-waves and P-waves is a deficiency of the lowest order chiral analysis. To correct this problem one has various options [DoGH 86]. One must introduce momentum dependent vertices, but this may be done in the 5-wave, the P-wave, or both. A next-order effective lagrangian analysis is not very useful, because although the weak amplitude may be easily fit, there are too many possible effective lagrangians to decide whether the corrections are in the S-wave or P-wave. Although various solutions have been proposed in quark models, such as inclusion of 1/2" pole contributions [LeOPR 79], there is no general agreement among the models. A fair summary of the situation is that, while solutions to the 5-wave/Pwave puzzle do seem possible through next-order corrections, a consensus on the unique solution chosen by Nature has not yet emerged. Quark model predictions For hyperon nonleptonic decay, there is a remarkably simple starting point which yields the A/ = 1/2 rule for at least a portion of the amplitude. In the B —• B' matrix elements, the A/ = 1/2 rule is automatic for ground state baryon state vectors of three quarks which are antisymmetric in color. This in turn implies that the portion of the decay amplitudes described by the lowest-order chiral analysis above will satisfy the A/ = 1/2 rule. The proof makes use of the Fierz rearrangement property (cf. Eq. (C-2.11)) of the product of chiral currents to rewrite the A/ = 3/2 operator O4 as _ 1 O 4 = 4 (<5ij<5ki + <5ii<5kj) diLlffMUjLUk^siL - z-diL^^djLdk^siL
L
i
,
(6.14)
* This statement is the SU(S) generalization of the Goldberger-Treiman relation, Eq. (3.24).
346
XII Baryon properties
where we work with left-handed fields as in Eq. (1-2.3) and i,j,k,l are color indices. Note that O4 is symmetric under the interchange of the colors of u and s (or d and u) in the first term and of d and s in the second. When a matrix element such as
Mfn = (n|O4|A) (6.15) is taken, within the approximation that baryons only contain three quarks the fields u and s must annihilate the quarks in the A. However, in this picture the A is antisymmetric under the color-interchange of any two quarks. Since O4 is symmetric and A is antisymmetric under this interchange, the matrix element MAN must vanish when all quark colors are summed over. As pleasing as this simple result is, it cannot be regarded as completely satisfactory. The B —> B' matrix elements are not responsible for all the B —> B'TT amplitudes, so the A/ = 1/2 rule must be also shown to hold for the remaining contributions. In addition, the 'three-quark' model of baryons is an oversimplified approximation, and we must hope that the 'sea' of quarks and antiquarks does not upset this result. A second prediction which can be derived simply in the three-quark approximation is that, in the absence of the penguin diagram, one must have D/F = — 1. Recall from Sect. VIII-3 that with out including the penguin diagram the hamiltonian can be written entirely in terms of the operators O± = d7M(l + 75)imYi(l + 75)5 ± ^ ( l + 7 5 ) ^ ( 1 + 75)5 . (6.16) Among the B —> Bf matrix elements parameterized by D and F, one finds F) . (6.17) However S~ and S~ contain no i^-quarks, while matrix elements of O+ vanish unless the states contain ^-quarks. Therefore we find . M ^ - ^ s - = 0, implying D/F = — 1. Modification of this result would require sea quarks and/or the penguin diagram, interpreting the latter as a perturbative way to include the effects of nonvalence quarks. As regards the more model dependent issue of the overall magnitude for nonleptonic decay amplitudes, quark models are seen to give approximately the right size for the matrix elements. Consider, for example, the contribution of operator O\. Using the general S-wave quark wavefunctions of Table XI-2, it is a relatively simple calculation to show that = -^V:dVusd
Jd3x
(u2 + £2)2 .
(6.18)
Insertion of bag model wavefunctions with bag radius R ~ 1 fm yields M^+p — —5.5 x 10~ 8 GeV, while the parameterization of Eq. (6.11)
Problems
347
would require M^p = V6(D-F) = -6.0 x 10"8 GeV. This calculation is essentially universal for all quark models, aside from the magnitude of the wavefunction overlap. Any variation of the magnitude may be understood by noting that due to the normalization of the quark wavefunction ipty ~ 1/V (where V is a typical baryon volume), a four-fermion matrix element scales like V~l, /
Sx (VV) ~V~l
.
(6.19)
Those quark models in which the quarks are tightly bound (small V) yield larger matrix elements than those with loosely bound quarks. In summary, the theoretical status of hyperon nonleptonic decays, while far from perfect, is better than that for kaons. Here, we have a simple starting point which comes within factors of two of the desired results. However, additional ingredients are clearly required for a more believable description. Problems 1) The axial-vector coupling Consider the effective lagrangian in Eq. (3.1) for nucleons and pions. For combined left-handed and right-handed transformations of the fields, we have
U -> LUR] , £ -> L£V* = V£rf , N -> VN , where L[R] are the spacetime independent SU(2) matrices corresponding to global transformations in SU(2)L[SU(2)R] and V — V(n(x)) is an SU(2) matrix describing a vectorial transformation of the nucleons. For the lagrangian of Eq. (3.1), use Noether's theorem to generate the SU(2) axial-vector current,
where ^ is the 'square root' of U (cf. Eq. (3.1)), and thereby show that the axial-vector coupling constant for beta decay is given by g\ — gA2) Nonleptonic radiative hadronic decays a) In addition to nonleptonic pion emission B —• B'TT, there exists the nonleptonic radiative mode B —• B'^. Show that the most general form for the radiative decay amplitude is + where C, D are respectively the Ml, El transition moments. TUB
348
XII Baryon properties b) Show that the rate for radiative decay of an unpolarized baryon B is given by = 2a ( m B ^ m B / ) ( m B + mB'){\C c) Demonstrate that if baryon B is polarized with polarization vector P#, the differential decay rate is dvB^Bfl
=
i r^ ^
{1 +
pB.eiaB,
)
where aBi1 = 2Re (C*-D)/[|C|2+|.D|2] is the asymmetry parameter and q is the photon direction. d) Using Eq. (6.8) for the weak BB couplings, show that in the limit of SU(3) symmetry the only nonvanishing contribution to the decay amplitudes comes from the baryon pole piece of the M l amplitude C. What does the experimental result a^ — — 0.83 ± 0.12 from E + radiative decay indicate about the validity of SU(3) invariance in these decays? 3) CP violation and nonleptonic hyperon decay Although the AS = 1 hamiltonian of the Standard Model contains a CP-violating component, there is no practical way to see this in any single hyperon decay mode. Rather, one must compare the decays of hyperons with those of antihyperons [DoHP 86]. In the presence of CP violation, there are two sources of phases in the weak matrix elements, e.g. for the A decay modes, - = A, e*l eiS" + A3 e^
where the isospin (I) subscripts '1,3' stand for A / = 1/2, 3/2, the angular momentum (J) superscripts 'S, P ' stand for *S-waves or Pwaves, Aj are real amplitudes, 6j are strong final state phases and (pj are the weak CP violating phases. Observe that there are three small parameters in these amplitudes - the weak phases (pj, the strong phases 6j ~ 10°, and the ratio of A / = 3/2 to A / = 1/2 effects. To leading order in these quantities, show that one has the CP-odd observables, P-\~/3
——
=
.
/
<
? p \
a + a
= sin (
r
p7r- ~ J W
=
.
/
<
? p \ .
AXAS sin(gf - 6j) sin(yf - yf)
|A| 2
I V I V
"
/ r ?
r P \
_ _ = - sin (
2
^
|S|2
n.p + iftp
Problems
349
A hierarchy is apparent in these three signals. The (3 + (3 asymmetry requires only the weak phase, the a + a asymmetry requires both the weak and final state phases, while F — F has both phases plus a A/ = 3/2 suppression. Present experiments are not sufficiently sensitive to test for CP violation in these observables at the required accuracy.
XIII Hadron spectroscopy
Studies of hadron masses, and of both strong and electromagnetic decays of hadrons, provide insights regarding QCD dynamics over a variety of distance scales. Among various possible theoretical approaches, the potential model has most heavily been employed in this area. We shall start our discussion by considering heavy-quark bound states, which begin to approximate truly nonrelativistic systems and for which the potential model is expected to provide a suitable basis for discussion. XIII-1 The charmonium and bottomonium systems Quarkonium is the bound state of a heavy quark Q with itsantiparticle. Two such systems, charmonium (cc) and bottomonium (bb) have been the subject of much experimental and theoretical study. Due to weak decay of the top quark, the (as yet undiscovered) ti system will almost certainly have rather different properties from these, and thus constitutes a special case (c/. Sect. XIV-2).
Table XIII-1. Nomenclature for S-wave and P-wave states in the cc and bb systems.
L S Charmonium 0 1
1 0
tp(nS)a
Vc(nS)
Vb(nS)
1
XcAnP) hc(nP)
XbAnP) hb(nP)
0 a
Bottomonium
?(nS)
For historical reasons, the spin-one charmonium ground state is called J/X/J.
350
XIII-1 The charmonium and bottomonium systems 4.0
-
?
35
_
(
fl(2S)^
3.0
_ Tl(IS) -+
jPc=
0
-
%<1P)
hc(lP),
/ Jz \] (
/t,«
CD
351
—^XodP) > —
_
"
J
J/y(IS)
r
-
,-
—
0
-
,••
2 *+
Fig. XIII-1 The low-lying spectrum of charmonium. Since the quarkonium systems are quark-antiquark composites, we shall employ the set of quantum numbers n, L, S, J introduced in Sect. XI-2. One generally refers to the individual quarkonium levels with the nomenclature of Table XIII-1, although the nL identification is sometimes replaced by either the degree of excitation or the mass, e.g. I/J(2S) is called ip' or '0(3686). The n2S+lLj spectroscopic notation is also invoked on occasion. Figs. XIII-1,2 give a summary of the lightest observed cc and bb states. At present, there are no firm candidates for lP\ states in either system and only one reasonably firm D-wave assignment, tjj{377Q). The greatest observed degrees of excitation come from the ijj{nS) and Y(nS) radial towers, reaching up to n = 6 for the T system. Excitation energies are relatively small on the scale of the bottomonium reduced mass //& ~ 2.5 GeV, but not that of charmonium /i c ~ 0.8 GeV. By far, the greater part of the theoretical effort on interpreting quarkonium systems has been performed in the context of nonrelativistic quantum mechanics [QuR 79].
2MB
T|K(1S>
Fig. XIII-2 The low-lying spectrum of bottomonium
352
XIII Hadron spectroscopy
Thus, quarkonium mass values are expressed as ™>[nLSJ] = 2 M Q + E[nLSJ]
,
(1.1)
where E\nLSJ] ls obtained by solving the Schrodinger equation for a particle of reduced mass /XQ = MQ/2 moving in the field of an assumed potential energy function. In the following, we shall consider several aspects of quarkonium systems. Lattice studies: The ultimate aim of lattice-gauge studies is to show that the potential picture is a consequence of QCD, and to even specify the quark-antiquark potential itself. Although this program is far from completion, results of lattice simulations are consistent with parameterizing the long-range part of the static QQ potential in pure (i.e. without light dynamic quarks) 5/7(3) gauge theory as [Has 87, Fu 87] V{r) = br--
+ V0 , (1.2) r where a, 6, Vo are constants and the color dependence between quark and antiquark is that in Eq. (XI-2.4). As noted in Sect. XI-2, the linear 'fer' term models a color-flux tube of constant energy density. The coefficient b is commonly described in the lattice-gauge literature as the string tension, in reference to the string model of hadrons, and its value is estimated from a string model relation involving the typical slope OL of a hadronic Regge trajectory (cf. Table XIII-3), b = ( 2 W ) " 1 ~ 0.18 GeV2 .
(1.3)
This is equivalent to a restoring force of about 16 tons! Numerical studies imply a relation between the string tension b and the confinement scale A.j^ of QCD [Fu 87], A
Ws = (0.318 ±0.058)>/6 ^ 0.13 ± 0.02 GeV .
(1.4)
If one goes beyond pure gauge theory by including the effects of a light dynamic quark q, the long-range potential between QQ becomes a shortrange interaction between Qq and Qq [Ze 88]. That is, the original interaction between the color charge carried by a heavy quark-antiquark pair becomes screened by the creation of a light quark-antiquark pair in the color field. Phenomenological potentials: The spectra of quarkonium states already hints at the radial dependence of the QQ potential, with the progression in nL levels suggesting an interaction which lies 'between' coulomb and harmonic oscillator potentials, as depicted in Fig. XIII-3. In practice,
XIII-1 The charmonium and bottomonium systems
ol Oscillator
is Quarkonium
353
*i Coulomb
Fig. XIII-3 Energy levels of various potential functions phenomenological studies of quarkonium are carried out by adopting an assumed potential energy function in accord with this behavior, e.g.* 64?r2
• { [ ^ ( l + ^/A2))]-1}
\ ~ 0.4 GeV} ,
,b~ 0.18 GeV2 \ \ \ , fc~~ 6.87 6.87 GeV 1 cr \d~0.1 J ' (1.5) where J7 {...} denotes a Fourier transform. The first two of the potentials in Eq. (1.5) are commonly called the 'Richardson' [Ri 79] and 'Cornell' [EiGKLY 80] potentials respectively. They are constructed to mimic QCD by exhibiting a linear confining potential at long distances and single gluon exchange at short distances. The Richardson potential even incorporates the asymptotic freedom property for the strong interaction coupling. The third is a power law potential [Ma 81] which, although not motivated by QCD, can be of use in analytical work or in obtaining simple scaling laws. The power law potential also serves as a reminder of how alternative forms can achieve a reasonable success in fitting bb and cc spectra, which after all, are primarily sensitive to the limited length scale 0.25 < r(fm) < 1. From the viewpoint of phenomenologyL it is ultimately more useful to appreciate the general features of the QQ static potential than to dwell on the relative virtues and shortcomings of individual models. In this regard, a study of spin dependence is instructive. .,, , i V(r) = {
, , br-a/r
br — a/r
Spin dependence: In order to analyze spin-dependent effects in quarkonium without detailed a priori knowledge of the inter quark potential, we assume an interaction suggested by the QED interaction of Eq. (V-1.16) * The second and third potentials provide fits only up to an additive constant.
354
XIII Hadron spectroscopy
but allowing for a more general vertex structure [Ja 76], 2
Ti ,
(r = 1,7^,75,7^75,^) ,
(1.6)
where Vi(q2) represents the propagator of an exchanged quantum. The nonrelativistic limit of this expression, expanded in inverse powers of the heavy-quark masses [EiF 81], yields a sum of static and spin-dependent contributions, general = V0 + Fspin = V0 + Vs-o + Ken + V8-a .
(1.7a)
The potential VSp[n is seen to generally contain spin-orbit (Vs-O), tensor (Ken), and spin-spin (V8-8) components, SQ
• r x pg 2M*
Sg • r x ] 2M|"
SQ
H Vten
=
I r
• r x pg — SQ • r x M
Q MQ
r
(1.7b)
* S
Q'SQ
where r = YQ—TQ. The quantity Vsp[n is expressed in terms of Vb, its radial derivative V^, and additional contributions K(r), V({r) (i = 1 , . . . , 4). Referring to Table XIII-2, we see that the QED Breit-Fermi potential represents a special case, with all nonzero contributions expressed in terms of the static coulomb potential. Also appearing in Table XIII-2 are general forms for vector and scalar vertices. The less interesting pseudoscalar vertex (which does not lead to a static potential) and the axial-vector and tensor vertices (which have only a leading spin-spin interaction) are not included. Let us apply Eqs. (1.7a,b) to quarkonium by working in the QQ centerof-mass frame with equal constituent quark masses MQ — MQ = M. The
Table XIII-2 . Spin-dependent potentials
v0
Vx
v2
Vs
v4
7M 7^
—a/r VV
0 0
1
vs
—a/r Vv 0
3a/r 3 V ^ / r - VV 0
2 V 2 VV 0
Interaction r QED Vector Scalar
-Vs
XIII-1
The charmonium and bottomonium systems
355
spin-dependent potential of Eq. (1.7b) then simplifies to
~ 'flL •
s +
b [
S
-
(18)
For the purpose of obtaining mass-splitting relations, we require the computation of various expectation values in the basis of ^[LSJ] states,
(3Po) ,
-2
2(2L - 1)<2L . _,
4
"1-3/4
,
. (1 9)
(%).
'
For example, these expectation values imply the following mass relations for the triplet P states: ™>(*P2) = m-\- ras_o - ^mten , 5 lit \
JL~^ I
ill
Tllig
Q
|" ''1%QJ\ 5
ra(3P0) = fn - 2ra s _ o - 2mten
.
The mass formulae in Eq. (1.10) can be used to test whether the longrange confining potential transforms as a four-scalar (Vs) or instead as a four-vector (VV) [BeCDK 79]. For definiteness, we consider a simple modification of the Cornell model in which the 'scalar vs vector' issue is cast in terms of a parameter £ (0 < £ < 1), Vs = (1 — £)br ,
Vy = £br
.
(1-H)
r It follows from Eq. (1.7b) and Table XIII-2 that the spin-orbit and tensor mass contributions are then given by
¥
(1.12)
and upon defining A = &(r~ 1 )/a(r~ 3 ), we form the ratio _ m(3P2) - mjZPi) " m( 3 Pi) - m(sP0)
2 16 - 19£A - 5A 5 8 + 5£A - A '
(
'
356
XIII Hadron spectroscopy
This can be compared with data from the Upsilon P-wave n = 1,2 states, of
0.
Mo. .70
(L14)
M2P)) .
For vector confinement (f ~ 1), Eq. (1.13) is in accord with the experimental values of Eq. (1.14) only for A ~ 0, whereas scalar confinement produces reasonable agreement for a much larger range, 0.4 < A < 1.0. Moreover, the Cornell model suggests that the latter values for A are rather more reasonable than the former, and so supports the conclusion that confinement is produced by a long-range, four-scalar interaction. Transitions in quarkonium All quarkonium states are unstable. Among the decay mechanisms are annihilation processes, hadronic transitions, and radiative transitions. Roughly speaking, the lightest quarkonium states are relatively narrow, but those lying above the heavy-flavor threshold, defined as twice the mass of the lightest heavy-flavored meson and depicted by dashed lines in Figs. XIII-l(a,b), are rather broader. This pattern is particularly apparent for the SS\ states - below the heavy-flavor threshold, widths are typically tens of keV, whereas above, they are tens of MeV. The primary reason for this difference is that above the heavy-flavor threshold, quarkonium can rapidly 'fall apart' into a pair of heavy-flavored mesons, e.g. T[45] —> BB, whereas below, this mode is kinematically forbidden. In the following, we shall describe only decays which occur beneath the heavy-flavor threshold, and shall limit our discussion to annihilation processes and hadronic decays. Radiative electric and magnetic dipole transitions are adequately described in quantum mechanics textbooks. Annihilation transitions: To motivate a procedure for computing annihilation rates in quarkonium, let us consider the simple case of a charged lepton of mass m moving nonrelativistically with its antiparticle in a 1So state, and undergoing a transition to a two-photon final state.* First we
666 or
66 or yy Q.
(a)
(b)
CO
Fig. XIII-4 Decay of quarkonium through annihilation. The 1S'o (3S±) states have even (odd) charge conjugation, and can therefore give rise to even (odd) numbers of photons in an annihilation process.
XIII-1 The charmonium and bottomonium systems
357
write down the invariant amplitude for the pair annihilation process,
(1.15) for momentum eigenstates. In the lepton rest frame, we are free to choose, the transverse gauge e\ - p- = t\ • p- = 0, i.e. e\2 — 0. Since SS\ states can make no contribution to the two-photon mode, we can compute the squared-amplitude for a ^ o transition by summing over over initial state spins,
^[
+
£ * ! "»']
(1J6)
where u)\£ are the photon energies in the lepton rest frame. Near threshold the photons emerge back to back, and the differential cross section is found to be
Likewise, near threshold a sum on photon polarizations gives - (ej • e^)2)thr = 2 ,
(1.18)
and upon integrating over half the solid angle (due to photon indistinguishability) we obtain the cross section,
* = *P- .
(1.19)
This is the transition rate per incident flux of antileptons. Since the flux is just the antilepton velocity v+ times a unit lepton density, we interpret v+a as the transition rate for a density of one lepton per volume. For a bound state with radial quantum number n and wavefunction \P n(x), the density is |^ n (0)| 2 and the lowest-order expression for the electromagnetic decay rate D ^ p S o ] becomes
^
^|M)| Tib
.
(1.20)
The corresponding rate for 77 emission from xSo states of the bb (T) system is obtained from Eq. (1.20) by including a factor e\ — 1/81, which accounts for the fe-quark charge, and a color factor of three. Determination of the two-gluon emission is found similarly (cf. Fig. XIII-4(a)) provided the gluons are taken to be massless free particles, and is left for a problem at the end of the chapter. Including also the effects of
358
XIII Hadron spectroscopy
QCD radiative corrections, referred to a common renormalization point HR = mb, we have [KwQR 87] 2
5o] =
|vM0)| 2 L - 3.4 , .< 1+4A
—3(^p— L
—^\ •
(L21)
Decays can also occur from the nzS\ states.* The single photon intermediate state of Fig. XIII-4(b) leads to emission of a lepton pair, whereas Fig. XIII-4(c) describes final states consisting of three gluons, two gluons and a photon, or three photons. For such a three-particle final state, there are six Feynman diagrams per amplitude and three-particle phase space to contend with. Upon including QCD radiative corrections, the results are [KwQR 87]
2
" 12 - 6 ^—J '
2187(2m6) 128(7r 2 -9)aa 2 (m 6 )[^ r a (0)| 2 [
I
i
(L22)
^
The QCD contributions in Eq. (1.22) are of interest in several respects. They contribute, on the whole, with rather sizeable coefficients and can substantially affect the annihilation rates. Also, they have come to be used as one of several standard inputs for phenomenological determinations of as. To eliminate the model-dependent factors |\t n (0)| 2 , one works with ratios of annihilation rates, a 5as(mb)
w
9
as(mb)\ V"
Tra2
' "'"
7T J 2 ( Mr\ \2mh)
(1.23) In reality, there are a number of theoretical and experimental concerns which make the extraction of ^(mb) a rather more subtle process than it might at first appear: (i) the contribution of |* n (0)| 2 in Eqs. (1.21),(1.22) * There are annihilations from higher partial waves as well. These involve derivatives of the wavefunction at the origin.
XIII-1 The charmonium and bottomonium systems
359
as a strictly multiplicative factor is a consequence of the nonrelativistic approximation and may be affected by relativistic corrections, (ii) there is no assurance that O(as)2 terms are non-negligible, particularly in the light of the large first order corrections, (iii) experiments see not gluons but rather gluon jets, and at the mass scale of the upsilon system, jets are not particularly well-defined, and (iv) the 7 spectrum observed in the jgg mode is softer than that predicted by perturbative QCD, implying the presence of important nonperturbative effects. Nevertheless, determinations of this type lead to the central value (and its uncertainties) A^- = 16Olgo° MeV as extracted from upsilon data and cited earlier in Table II—3. This example indicates how demanding a task it is to obtain a precise experimental determination of as(q2). Hadron transitions: The transitions V —• V + TT0 and V -* V + 77 involving the decay of an excited 3 5i quarkonium level (V7) down to the S S\ ground state (V) are interesting because they are forbidden in the limits of flavor SU(2) and SU(3) symmetry respectively. Their rates are therefore governed by quark mass differences, and a ratio of such rates provides a determination of quark mass ratios. There is a modest theoretical subtlety in extracting the rates, as degenerate perturbation theory must be used [IoS 80]. The leading-order effective lagrangian for these P-wave transitions must be linear in the quark mass matrix m, ^ IV (m(U = C [{pid — TTlu)—-^ + (2i7ls — TTld —
TYly)—^
+ . . .]
(1.24) where c is a constant. Here, n^ and r)$ are the pure SU(S) states which appear prior to mixing 7T° = cos 9 7T3 + sin 6 % ,
77 = — sin 0 TT3 + cos 9 rjs ,
(1.25)
where tan 9 ~9 = y/3(rrid — mu)/[2(2ms — rrid — mu)] describes the quark mixing. Upon calculating the transition amplitudes and then substituting for the small mixing angle 9, we obtain -u\ = ——(rrid —
2>/2
Mo 2Mp
— mu)9 H
2ms
Tnu) ,
-rrid-mu (1.26)
360
XIII Hadron spectroscopy
where Mo = ic e^^k^kf^p. Q
Ty'^vrj
The ratio of decay rates is found to be 27 m mddm u u -m
16 ms — m
p^
p^
We can extract a quark mass ratio from charmonium data involving ip{2S) - • J / ^ transitions. Prom the measured value Q = 0.037 ± 0.009 [RPP90], wefind md mu
~ = 0.0336 ± 0.004 , (1.28) ms — m which is about 40% larger than the same ratio extracted from pion and kaon masses (c/. Eq. (VII-1.10)). XIII-2 Light mesons and baryons In the quark model, the light baryons and mesons are Q 3 and QQ combinations of the u, d, s quarks. The resulting spectrum is very rich, containing both orbital and radial excitations of the L = 0 ground state hadrons. For mesons, the Q and Q spins couple to the total spins S = 0,1, and each (L, S) combination occurs in the nine flavor configurations of the flavorSU(3) multiplets 8, 1. Analogous statements can be made for baryon states. In the face of such complex spectra, we are mainly interested in the regularities that allow us to extract the essential physics. A tour through the data base in [RPP 90] reveals some general patterns.* Both radial and orbital excitations of the light hadrons appear 0.5 —> 0.7 GeV above the ground states. As pointed out in Sect. XI-1, this indicates that the light quarks move relativistically. Other striking regularities are (i) the existence of quasi-degenerate supermultiplets of particles with differing flavors and equal (or adjoining) spins, and (ii) excitations of a given flavor having increasingly large mass (M) and angular momentum (J) values which obey J = a'M2 + Jo. SU(6) classification of the light hadrons To the extent that the potential is spin-independent and we work in the limit of equal u,d,s mass, the quark hamiltonian is invariant under flavor-SU(3) and spin-*SC/(2) transformations. To lowest order, hadrons are thus placed in irreducible representations of *S/7(6), and quarks are assigned to the fundamental representation 6, 6 = « d T
s"[ u[
d[
si)
.
(2.1)
* Our discussion will focus on hadron masses. Strong and electromagnetic transitions are described in[LeOPR88].
XIII-2 Light mesons and baryons
361
We can also write the SU(6) quark multiplet in terms of the SU(3) flavor representation and the spin multiplicity as 6 = (3,2). Although the SU(6) invariant limit forms a convenient basis for a classification of the meson and baryon states, it cannot be a full symmetry of Nature since the spin is a spacetime property of particles whereas SU(3) flavor symmetry is not. Thus it is impossible to unite the flavor and spin symmetries in a relativistically invariant manner [CoM 67]. Although we shall avoid making detailed predictions based on SU(6), it is nonetheless useful in organizing the multitude of observed hadronic levels. Meson supermultiplets: The L=0 QQ composites are contained in the SU(6) group product 6 x 6* = 35 © 1, where the representations 35, 1 have flavor-spin content 35 = (8,3) © (8,1) ©(1,3),
1 = (1,1) .
(2.2)
The L=0 ground state consists of a vector octet, a pseudoscalar octet, a vector singlet, and a pseudoscalar singlet. For excited states, the meson supermultiplets constitute an SU(6) x O(3) spectrum of particles. The O(3) label refers to how the total angular momentum is obtained from J = L + S, giving rise to the pattern of rotational excitations displayed previously in Table XI-3. Roughly speaking, mesons occur in mass bands having a common degree of radial and/or orbital excitation. Fig. XIII-5 provides a view of the mass spectrum for the lightest mesons. The SU(6) x O(3) structure of the ground state and a sequence of orbitally excited states are observed to the extent that sufficient data is available for particle assignments to be made. Note that the S-wave QQ states are all accounted for, but gaps appear in all higher partial waves. Even after many years of study, meson phenomenology below 2 GeV is far from complete! Baryon supermultiplets: The SU(6) baryon multiplet structure arises from the Q 3 group product (6 x 6) x 6 = (21©15) x 6 = 56©70©70©20, and has flavor-spin content 56 = (10,4) ©(8,2), 70 = (8,4) © (10,2) © (8,2) © (1,2) , 20 = (8,2) ©(1,4).
(2.3)
A three-quark system must adhere to the constraint of Fermi statistics. Each baryon state vector is thus antisymmetric under the interchange of any two quarks. A Young-tableaux analysis of the above group product reveals that the spin-flavor parts of the 56, 70, and 20 multiplets are respectively symmetric, mixed, and antisymmetric under interchange of pairs of quarks. Since the color part of any Q 3 color-singlet state vector
XIII Hadron spectroscopy
362 NONET:
[J
1= 1/2 1 0
_
Cnn, ss) well established good evidence weak evidence
L=4
L=3
L=2
L-1
Fig. XIII-5 Spectrum of the light mesons
is antisymmetric under interchange of any two quarks, the 56-plet has a totally symmetric space wavefunction, with zero orbital angular momentum between each quark-pair. The 70 and 20 multiplets require either radial excitations and/or orbital excitations. Recall the characterization of the baryon spectrum in terms of the basis defined by an independent pair of oscillators (c/. Eq. (XI-2.12)). In this context, a standard notation for a baryon supermultiplet is (R, L^-), where R labels the 577(6) representation, V is the parity, N labels the number of oscillator quanta and L is the orbital angular momentum quantum number (cf. Sect. XI-2). Like meson masses, baryon masses tend to cluster in bands having a common value of N. The first three bands are shown in Fig. XIII6, and effects of SU(6) breaking are displayed for the first two. The lowest lying SU(6) x O(3) supermultiplet is the positive parity (56, 0Q~), having content as in Eq. (2.3). Next comes the negative parity (70,1J~) supermultiplet. This contains more states than the 70-plet shown in Eq. (2.3) because the extension from L = 0 to L = 1 requires addition of
XIII-2 Light mesons and baryons
363
3/2
(56, Oj) (70,
1/2
81/2 83/2
(56, 0 0 )
Fig. XIII-6 The low-lying baryon spectrum. angular momenta, (10, 2) -> (10,4) - (10,2) , (8,4) - (8,6) - (8,4) (8,2) (2.4) (1, 2) -> (1,4) - (1, 2) , (8,2) -> (8,4) - (8, 2) The number of supermultiplets grows per unit of excitation thereafter. There are five SU(6) multiplets in the N = 2 band, (56,2^), (56,0£), (70,2j), (50, Oj), and (20, l£). Recall that the baryonic inter-quark potential was expressed in Eq. (XI-2.10) as V = Vosc + [/, where V^sc is the potential energy of a harmonic oscillator and U = V — Vosc. If the potential energy were purely VOSc^ the supermultiplets within the N = 2 band would all be degenerate. In the potential model, assuming that the largest part of U is purely radial, this degeneracy is removed by the first-order perturbative effect of [/, and the splittings in the N = 2 band are shown at the top of Fig. XIII-6. Aside from choosing the (56, O^) supermultiplet to have the lowest mass, one finds the pattern of splitting to be as shown in Fig. XIII-6, independent of the particular form of U [HeK 83]. trajectories It is natural to classify together a ground state hadron and its rotational excitations, e.g. the isospin one-half baryons ^V(939)j=i/2 (the nucleon), JV(1680)j=5/25 and 7V(2220)j=9/2- Although no higher spin entries have
364
XIII Hadron spectroscopy
been detected in this particular set of nucleonic states (presumably due to experimental limitations), there is no theoretical reason to expect any such sequence to end. The data base in [RPP 90] contains a number of similar structures, each characteristically containing three or four members. Each such collection of states is said to belong to a given Regge trajectory. To see how this concept arises, let us consider the simplest case of two spinless particles with scattering amplitude f(E, z) {i.e. da/dVt = \f{E,z)\2), where E is the energy and z = cos# is the scattering angle. It turns out that analytic properties of the scattering amplitude in the complex angular momentum (J) plane are of interest. One may obtain a representation of f(E, z) in the complex J-plane by converting the partial wave expansion into a so-called Watson-Sommerfeld transform,
f(E, z) = 2^(-Y (2£ + l)a(E, l)Pt(-z)
(2.5)
(bdJ -(2J + l)a(E, J)Pj(-z) , T 2m r where Pa is a Legendre polynomial and C is a contour enclosing the nonnegative integers. Suppose that as C is deformed away from the Re J axis to, say, a line of constant Re J, a pole in the partial wave amplitude a(E, J) is encountered. Such a singularity is referred to as a Regge pole and contributes (c/. Eq. (2.5)) to the full scattering amplitude as sin(7ra[£;])
' "• '
where OL[E] is the energy-dependent pole position in the complex J-plane and j3[E] is the pole residue. Table XIII-3. Regge trajectories. Trajectory
N
Slope a
N
3 3 3 3 2 3 4 4 4
0.99 0.92 0.94 1.1 0.91 0.72 0.84 0.69 0.86
A A E E* 7T
P K K* a
In units of GeV - 2
J-intercept -0.34 0.07 -0.64 -1.2 -0.24 -0.05 0.54 -0.22 0.29
XIII-2 Light mesons and baryons
365
The Regge-pole contribution of Eq. (2.6) can manifest itself physically in both the direct channel as a resonance and a crossed channel as an exchanged particle. Here, we discuss just the former case by demonstrating how a given Regge pole can be related to a sequence of rotational excitations. Suppose that at some energy £"#, the real part of the pole position equals a nonnegative integer £, i.e. Re a[Eji\ = £. Then with the aid of the identity,
i r1 we can infer from Eq. (2.6) the Breit-Wigner resonance form (Rg.-ple.) _
e
fi 1 T/2 7T (a[E] - £)(a[E] + £ + 1) " E - ER + iV/2
l
*}
provided Re OL[ER\ » Im a[En]. A physical resonance thus appears if OL\E] passes near a non-negative integer, and if the Regge pole moves to ever-increasing J values in the complex J-plane as the energy E is increased, it generates a tower of high-spin states. Except in instances of so-called exchange degeneracy, parity dictates that there be two units of angular momentum between members of a given trajectory. In this manner, a single Regge pole in the angular momentum plane gives rise to the collection of physical states called a Regge trajectory. A plot of the angular momentum vs squared-mass for the states on any meson or baryon trajectory reveals the linear behavior, J~a'm2
+ J0 .
(2.9)
A compilation of slopes (a/) and intercepts (Jo) appears in Table XIII-3, with each trajectory labeled by its ground state hadron. Such linearly rising trajectories have been interpreted as a consequence of QCD [JoT 76]. In this picture, hadrons undergoing highly excited rotational motion come to approach color-flux tubes, whereupon it becomes possible to relate the angular momentum of rotation to the energy contained in the color field. This line of reasoning leads to the behavior of Eq. (2.9), and accounts for the universality seen in the slope values displayed in Table XIII-3.
SU(6) breaking effects Although an SU(Q) invariant hamiltonian provides a convenient basis for describing light hadrons, the physical spectrum exhibits substantial departures from the mass degeneracies which occur in this overly symmetric picture. In the following, we shall consider some simple models for explaining the many SU(6) breaking effects observed in the real world. The QCD Breit-Fermi model: If one ascribes the non-confining part of the quark interaction to single-gluon exchange, the nonrelativistic limit
366
XIII Hadron spectroscopy
yields the 'QCD Breit-Fermi potential' [DeGG 75] ^one—gluon — ~~
where as is the strong fine structure constant, r = ry, and k denotes the color dependence of the potential (c/. Sect. XI-2) with k = 1 (1/2) for mesons (baryons). In keeping with the potential model, the mass parameters {Mi} are interpreted as constituent quark masses. Although the QCD Breit-Fermi model incorporates SU(6) breaking by means of both quark mass splittings and spin-dependent interactions, it lacks a rigorous theoretical foundation. One might argue on the grounds of asymptotic freedom that Eq. (2.10) does justice to physics at very short distances (in the approximation that as is constant), but there is no reason to believe that it suffices at intermediate length scales. It also does not account for mixing between isoscalar mesons, so such states must be considered separately. Meson masses: The gluon-exchange model can be used to obtain information on constituent quark mass. In the following, we shall temporarily ignore the minor effect of isospin breaking by working with M = (Mu + Md)/2. To compute meson masses, we take the expectation value of the full hamiltonian between SU(6) eigenstates, specifically the L = 0 QQ states.* Although the form of Eq. (2.10) implies the presence of spin-spin, spin-orbit, and tensor interactions, the spin-orbit and tensor terms do not contribute here because each quark pair moves in an 5-wave, and it is the spin-spin (hyperfine) interaction which lifts the vector meson states relative to the pseudosclar mesons. We can parameterize the non-isoscalar L = 0 meson masses as -\ Q
+ ^ Q Q ( S Q * S Q)
»
( 2-H)
* An analysis of spin-dependence in the L = 1 states is the subject of a problem at the end of the chapter (cf. Prob. XIII-3)).
XIII-2 Light mesons and baryons
367
where n and ns are the number of nonstrange and strange consituents respectively, and HQQ is the hyperfine mass parameter defined in Eq. (XII3.12). One consequence of Eq. (2.11) is a relation involving the mass ratio M/Ms. Fitting the four masses TT(138), K(496), p(770), K*(892) to the parameters in Eq. (2.11) yields nip — m^
=7xgg == M&M nn
~ 0.63 .
(2.12)
s
The origin of this result lies in the inverse dependence of the hyperfine interaction upon constituent quark mass, which affects the mass splitting between S = 1 and S = 0 mesons differently for strange and nonstrange mesons. The numerical value of M/Ms in Eq. (2.12) graphically demonstrates the difference between constituent quark masses and current quark masses, the latter having a mass ratio of about 0.04 (cf. Eq. (VII-1.6a)). In earlier sections of this book which stressed the role of chiral symmetry, the pion was given a special status as a quasi-Goldstone particle. In the QQ model, the small pion mass is seen to be a consequence of severe cancelation between the spin-independent and spin-dependent contributions. However, the parameterization of Eq. (2.11) cannot explain the large 7/(960) mass. In addition to the SU(6) symmetry breaking effects of mass and spin, there is an additive contribution present in the isoscalar channel which is induced by quark-antiquark annihilation into gluons. In the basis of u, d, s quark flavor states, this annihilation process produces a 3 x 3 mass matrix of the form '2M,, + X X X \ (2.13) where for C = +1(—1) mesons, X is the two-gluon (three-gluon) annihilation amplitude, and for simplicity we display just the quark mass contribution (2Mi) as the nonmixing mass contribution. The annihilation process is a short-range phenomenon, so the magnitude of X depends sharply on the orbital angular momentum L of the QQ system. For L ^ 0 waves (where the wavefunction vanishes at zero relative separation), and C = — 1 channels (where the annihilation amplitude is suppressed by the three powers of gluon coupling), we expect Ms — M » X. In this limit, diagonalization of Eq. (2.13) yields to leading order the set of basis states (uu ± dd)/y/2 and ss. Only the L = 0 pseudoscalar channel experiences opposite limit X ^> Ms—M, wherein to leading order the basis vectors are the SU(3) singlet state (uu + dd + ss)/y/3 and octet states (uu — dd)/y/2, (uu + dd — 2ss)/y/6. The overall picture that emerges is one of relatively
368
XIII Hadron spectroscopy
unmixed light pseudoscalar states, and heavily mixed vector, tensor, etc. states. Baryon masses: Applying the one-gluon exchange potential to the ground state baryons of (56, OQ") yields a mass formula analogous to Eq. (2.11),
For the system of l/2 + and 3/2 + (iospin-averaged) baryons, there are eight mass values and since the above mass formula contains five parameters, one should obtain three relations. The additional perturbative assumption Hss — Wns = tins — Wnn for the hyperfine mass parameters yields the well-known Gell-Mann-Okubo relation of Eq. (XII-3.10) for the 1/2+ baryons and the equal spacing rule for 3/2 + states, — 771A = mz* — rriz*
UIQ
ra^*
.
(Z. l o )
(Expt. 153 MeV = 149 MeV = 139 MeV ) A third relation which relates the 3/2 + and 1/2+ masses and is independent of further perturbative assumptions has the form 3raA - raE - 2mN = 2(raE* - raA) (Expt. : 276 MeV - 305 MeV) In addition, one can obtain estimates for M/Ms, among them M
=
2(mE» Q
^
M8 2ms* + ms - 3mA ' ' . ^ M mE* - m s —- = ~ 0.65 , both in accord withMEq. (2.12). s Isospin breaking effects: The above description of SU(6) breaking assumes isospin conservation. In fact, hadrons exhibit small mass splittings within isospin multiplets, arising from electromagnet ism and the u — d mass difference. In the pion and kaon systems, we were able to use chiral £(7(3) symmetry to isolate each of these separately. Unfortunately, this is not possible in general, and models are required to address this issue. There are a few consequences which follow purely from symmetry considerations. Since the mass difference mu — rrtd is A/ = 1, the A7 = 2 combinations mE+ + raE- - 2raEo = 1.7 ± 0.1 MeV , rap+ - mpo = -0.3 ± 2.2 MeV , (2.18)
XIII-2 Light mesons and baryons
369
arise only from the electromagnetic interaction. In addition, both electromagnetic and quark mass contributions satisfy the Coleman-Glashow relation [CoG 64], raE+ - raE- +mn-mp + m E - - mE o = 0 [Expt. 0.4 ± 0.6 MeV = 0 ] . For electromagnet ism, this is a consequence of the [/-spin singlet character of the current, whereas for quark masses it follows from the A/ = 1 and SU(3)-octet character of the current. We proceed further by using a simple model, based on the QED coulomb and hyperfine effects, to describe the electromagnetic interaction of quarks, Ara cou l = v4coul 2^
QiQj
MM
Sl
S<7
'
where Acouh ^hyp axe constants, {Qi} are quark electric charges, and the sums are taken over constituent quarks. In Arahyp, we shall neglect further isospin breaking in the masses and use Mu = M^ = M, and assume electromagnetic self-interactions of a quark to be already accounted for in the mass parameter of that quark. For any values of *AcOui a n d "4hyp> this model contains the sum rule (mn-mp)em
= —(ra E + + m E - -2ra E o) = -0.57±0.03 MeV , (2.21) o leaving the excess due to the quark mass difference, (mn — mp)qm =
• (n\uu — dd\n)
• (p\uu — dd\p)
— mu)(dm + fm)Zo = (mn - mp) - (mn - mp)em = 1.86 ± 0.03 MeV ,
(2.22) where the second line in the above uses the parameterization of hyperon mass splittings given in Eq. (XII-3.8). To the extent that this estimate of quark mass differences is meaningful, one obtains the mass ratio, md
~m" =
ms — m
(m
" "
m
^
rriE — m
m
~ 0.015 ,
(2.23)
E
to be compared to the chiral-symmetry extraction from meson masses which yielded 0.023 (cf. Eq. (VII-1.10)). With further neglect of terms O(a(Ms — M)) in the hyperfine interaction, this exercise can be repeated
370
XIII Hadron spectroscopy
for vector mesons to yield 2 (mK*o - mK*+)em = —-(m p + - mpo) = 0.2 ± 1 . 5 MeV , o
(mK.o - mK*+)qm - (mK*o - mK*+) - {mK,o -
= 6.5 ± 1.9 MeV ,
md
~m" =
7Tls — Vh
mK
'° ~mK'+
mK*+)em
^' '
= 0.053 ± 0.016 .
TYlK* — flip
The additional assumption that the constants *AcOui a n d *4hyp a r ^ the same in the decuplet baryons and the octet baryons, as is true in the SU(6) limit, leads to 5 (mA++ - raAo)em = ^ ( ^ E + + m E - — 2mEo) = 2.8 ± 0.2 MeV , (mA++ - m A o) qm = (mA++ - mAo) - (mA++ - mAo)em = -5.5 ± 0.4 MeV , ^' ' ?7ls — 771
2
±
Of course, the spread of values for the mass ratios raises a concern about the validity of this simple model. However, all methods of calculation agree on the smallness of the ratio (ra^ — mu)/(ms — rh). XIII-3 The heavy-quark limit In the quark description, a heavy-flavored hadron contains at least one of the heavy quarks c, 6, t. It is possible to describe such heavy systems with dynamical models like those employed for the light hadrons [DeGG 75, IzDS 82]. However, while such models are often valuable, it is always preferable to have a valid approximation scheme which follows directly from QCD. In this regard, a study of the heavy-quark limit (rriQ —+ oo) in which the theory is expanded in powers of m^ 1 is proving useful in analytic and lattice studies of meson spectroscopy and in the area of weak decays. Heavy-flavored hadrons in the quark model The spectroscopy of heavy-flavored hadrons should qualitatively follow that of the light hadronic spectrum, with states containing a single heavyquark Q occurring as either mesons (Qq) or baryons (Qqiq2)> The lowest energy state for a given hadronic flavor will have zero orbital angular momentum between the quarks, leading to ground state spin values S — 0,1 for mesons and S = 1/2,3/2 for baryons. The hyperfine interaction will lower the 5 = 0 meson and S = 1/2 baryon masses, and both orbital and radial hadronic excitations of the ground state will be present.
XIII-3 The heavy-quark limit
371
Although it is possible to contemplate extended flavor transformations which involve interchange of the light and heavy quarks, e.g. as in the S77(4) of the light and charmed hadrons, such symmetries are so badly broken by the difference in energy scales MQ » M q and MQ » AQCD as to be rendered useless. However, the SU(3) and SU(2) flavor symmetries associated with the light hadrons are still viable, but multiplet patterns become modified. The mesons Qq will exist in the SU(3) multiplet 3*, whereas in the baryonic Qqiq2 configurations the two light quarks <7i,2 will form the flavor SU(3) multiplets 6 and 3*. For example, the charmed system has the meson ground state 3* : D+[cd], D°[cu], Ds[cs] , which displays the mass pattern of an SU(2) doublet (D^m9, D^se5) and an 577(2) singlet (D{9Q9). The charmed baryon multiplets are 6 : E++[mxc], £ 3* : A+[tufc], S Fig. XIII-7 displays the anticipated charmed meson and charmed baryon levels, including the effect of SU(3) breaking. Heavy-quark constituent mass values can be inferred from the JD* — D and B* — B hyperfine splittings. That the former splitting is about three times the latter is a consequence of M& ~ 3MC and of the inverse dependence of the hyperfine effect upon quark mass. Analogously to Eq. (2.17), we find M Mc
M nip —
- mB mp —
where M = (Mu + M^)/2. These findings depend to some extent on how the fit is done, e.g. with mesons or with baryons, and we leave further study for Prob. XIII-4.
Charmed baryons
Charmed mesons
2.2 M (GeV) 2 _ = D *
2.8 s
2.6
2.4 1 ft
Fig. XIII-7 Spectrum of charmed (a) mesons, (b) baryons
372
XIII Hadron spectroscopy Spectroscopy in the TUQ —>
oo limit
In a hadron which contains a single heavy quark Q along with light degrees of freedom, the heavy quark is essentially static. The best analogy is with atoms, where the nucleus can in the first approximation be treated as a static, electrically-charged source. Likewise, for heavy hadrons the heavy quark is a static source with color charge, and the light degrees of freedom provide a non-static hadronic environment around Q. This scenario can be formalized by partitioning the heavy quark lagrangian as [CaL 86, Ei 88, LeT 88] CQ = ${%$$
mQ) I/J = C0 + £ space - rnQ) ^ ,
£space = - 2 ^
D^
where D^ip is the covariant derivative of SU(3)C. Since the spatial 7 matrices connect upper and lower components, we see that the effect of £Space
is
O(mQ1)'
Observe that the static lagrangian Co of Eq. (3.2) is invariant under spin rotations of the heavy quark Q. In the world defined by £0, with both O(AQCD/MQ) effects and O((XS(MQ)) effects (associated with hardgluon exchange) ignored, heavy hadronic energy levels and couplings are constrained by the SU{2) spin symmetry. It is helpful to visualize the situation. A heavy flavored hadron of spin S will contain a static quark Q having a constant spin vector SQ (with SQ = 1/2) and light degrees of freedom having a constant angular momentum vector J^ = S — SQ.* For a meson of this type, we assume that J(> behaves as it does in the quark model, with J# = 1/2 in the ground state and Jt — L ± 1/2 for L > 0 rotational excitations. Prom the decoupling of the heavy quark spin, it follows that there will be a two-fold degeneracy between mesons having spin values S = Ji ± 1/2. The meson L = 0 ground state will have Ji = 1/2 and thus degenerate states with S = 0,1. The L = 1 first rotational excitation with Ji = 1/2 will give rise to degenerate S = 0,1 levels, whereas for Ji = 3/2 one obtains degenerate levels having S = 1,2. Moreover, the energy differences between different levels should be independent of heavy quark flavor. Analogous conditions hold for heavy flavored baryons and hadronic transitions between levels of differing L can be similarly analyzed. Let us explicitly demonstrate that the splitting between the J p = 1~ and Jp = 0~ states of a Qq meson must vanish in the limit of infinite quark mass. We note that the mathematical condition for spinAlthough the light degree(s) of freedom in the simple quark model is an antiquark q for mesons and two quarks (71*72 for baryons, the physical (i.e. actual) light degrees of freedom could entail unlimited numbers of gluons and/or quark-antiquark pairs.
XIII-3 The heavy-quark limit
373
independence is "n = 0 ,
(3.3)
where S® is the generator of spin rotations about the 3-axis for quark Q and Ho is the hamiltonian obtained from Co- Since the action of S3 on a 0~ state produces a 1~ state, i.e. \MX~) = 2S3 |M 0 -), we then have = 2SsQHo\Mo-) = mo-\M1-) ,
Ho\M1-)=m1-\M1-)
(3.4)
implying that mx- — m0- —• 0 as TTIQ —• 00. Another consequence of working in the static limit of Co is that the propagator, Soo(x,y), of the heavy quark in an external field can be determined exactly. Prom the defining equations, -y)
(A) = do+ igsAo-X/2) , (3.5)
one has the solution 500(0:, y) = -iP(x0, yo)6{3)(x - y)
(3-6) where P(xo, yo) is the path-ordered exponential along the time direction,
P(x0, w) = Pexp \& r dt\. Ao(x,t)l .
(3.7)
In this approximation, the heavy quark is static at point x and the only time-dependence is that of a phase. This discussion can be generalized to a frame where the heavy quark is moving at a fixed velocity v, described by a velocity four vector v^ = pV/rriQ, with v^v^ = 1. One can define projection operators Tv±=l-{l±i})
,
(3.8)
where T^± = Tv±, Tv±Tvzp = 0, and Fv+ + Tv- = 1. The Tv± generalize the usual projection of 'upper' and 'lower' components into the moving frame. A quark moving with velocity v will have the leading description of its wavefunction contained in the 'upper' component described by a field hv [Ge 90, Wi 91], Tv ^
= e-imQv'xhv(x)
,
(3.9)
where the main dependence on the quark mass has been factored out, and hv obviously satisfies Tv+hv = hv. Substituting into the Dirac lagrangian,
374
XIII Hadron spectroscopy
neglecting lower components, and using Tv+Iprv+ = v • D yields - mQ)
- mQ) t/> ~
= hviv • Dhv , (3.10)
which generates the lowest order equation of motion v • Dhv = 0. This approximation can be systematically improved by inclusion of a 'lower' component for the heavy-quark field [EiH 90a,b, Lu 90, GeGW 90], V V = e-imQvx£v(x)
,
(3.11)
with Tv-£v = £v. The equations of motion allow us to solve for £v by following the sequence of steps, 0 = (iIp - mq) tf> = {ilp - mQ) e-imQv'x
[hv + lv] -xIp) [hv + £v]
x
[{-2mQ
which yields lv and ip as
ilphv]
,
(3.12)
(3.13) e-imQvx
=
j
Inserting these forms into Eq. (3.10) and using Tv+hv = hv and Eq. (Ill— 3.50) for^Z) yields L,v — tiv
m
Q
v
D
•
-
(3.14) which is the desired expansion in terms of the heavy quark mass. Because the last term in this expression is second order in v • D and noting that v • Dhv = 0 to lowest order, it will not contribute to matrix elements at order 1/rriQ and can be dropped. The lagrangian of Eq. (3.14) corresponds to a quark moving at fixed velocity. Antiquark solutions can be constructed with the mass dependence e+tmQv'x^ with the result
-
i
v
•
D
-
(3.15) where the field kv satisfies Tv-kv = kv. It is legitimate to neglect the production of heavy QQ pairs. However, one should superpose the la-
XIII-4 Nonconventional hadron states
375
grangians for different velocities in a Lorentz invariant fashion,
C = J d4v 6{v^ - l)%0) [/# + £$]
=f^
The nature of the approximation at this stage is more of a classical limit rather than a nonrelativistic limit. To be sure, for any given quark one can work in the quark's rest frame, in which case the quark will be nonrelativistic. However, when external currents act on the fields, as will be considered in Sect. XIV-2, transitions form one frame to another occur for which Av is not small. On the other hand, the result can be said to be classical because quantum corrections have not yet been included and these can renormalize the coefficients in Ly. Also, diagrams involving the exchange of hard gluons can produce nonstatic intermediate states. Such corrections can be accounted for in perturbation theory [Wi 91].
XIII-4 Nonconventional hadron states Many suggestions have been made regarding the possibility of hadronic states beyond those predicted by the simple quark model of QQ and Qs configurations. The study of such states is hampered by the fact that we still have very little idea why the quark model works. QCD at low energy is a strongly interacting field theory, and we would expect a very rich and complicated description of hadronic structure. That the result should be describable in terms of a simple QQ and Q3 picture as even a first approximation remains a mystery. Quark models have been popular because they seem to work phenomenologically, not because they are a controlled approximation to QCD. This weakness becomes all the more evident when one tries to generalize quark model ideas to new areas. Much of the theoretical work on nonconventional states has involved the concept of a constituent gluon G, analogous to a constituent quark Q, and we shall cast our discussion with respect to this degree of freedom.* It is clear that there should be a cost in energy to excite a constituent gluon. The energy should not be extremely large, else it would be difficult to understand the early onset of scaling in deep inelastic scattering. However, it cannot be less than the uncertainty principle bound on a massless particle confined to a radius R ~ 1 fm of E = p>\/3/R ~ 342 MeV (cf. Sect. XI-1). Calculations have tended to use an effective gluon 'mass' in the range 0.5 < MG (GeV) < 0.6. The basic idea of confinement is that only color-singlet states exist as physical hadrons. If we identify those states which are color singlets * However, it should be understood that such a concept has not been shown to follow rigorously from QCD, nor indeed is a configuration of definite numbers of consitituent gluons a gauge invariant entity (cf. Sect. X-2).
376
XIII Hadron spectroscopy
and which contain few quark or gluon quanta, we can easily find other possible configurations besides QQ and Q3. Some of the more well-known examples are 1) Gluonia (or glueballs) - quarkless G2 or G3 states, which we shall discuss in more detail below, 2) Hybrids - color-singlet mixtures of constituent quarks and gluons like QQG mesons or Q3G baryons, 3) Dibaryons - six-quark configurations in which the quarks have similar spatial wavefunctions rather than two separate three-quark clusters, 4) Meson molecules - loosely bound deuteron-like composites of mesons. A convenient framework for describing the quantum numbers of possible hadronic states is obtained by considering gauge-invariant, colorsinglet operators of low dimension [JaJR 86], as was discussed in Sect. XI1. Table XIII-4 lists all such operators up to dimension five which can be constructed from quark fields, the QCD covariant derivative, and the gluon field strength, denoted respectively by q, X>, and F. Also appearing in Table XIII-4 is the collection of JPC quantum numbers associated with each such operator. Particular spin-parity values are obtained from these operators by choosing indices in appropriate combinations.
Gluonia The existence of a gluon degree of freedom in hadrons is beyond dispute, with evidence from deep inelastic lepton scattering and jet structure in hadron-hadron collisions. However, trying to predict the properties of a new class of hadrons whose primary ingredient is gluonic is nontrivial. Hypothetically, if quarks could be removed from QCD the resulting hadron spectrum would consist only of gluonia (or 'glueballs'). Gluonic configurations should be signaled by the existence of extra states beyond the expected nonets of QQ hadrons. However, mixing with QQ hadrons is generally possible (cf. Sect. X-2). Although predicted by Table XIII-4. Gauge-invariant color-singlet interpolating fields. Operator Dimension qTq qTVq FF grgF FVF
3 4 4 5 5
JPC 2++,2"± 0 + + 2 + + 0—^ 2—*" 0±+,0±-,l±+,l±-,2±+,2±
XIII-4 Nonconventional hadron states
377
the 1/NC expansion to be suppressed, such mixing effects serve to cloud the interpretation of data vis-a-vis gluonium states. Referring to the interpolating fields mentioned above, we see that for gluons the gaugeinvariant combinations pa.Fx
™ piLv
1
£
[iv a
r
5
fiX
1
pa
av •>
c
ILV
L
fi»v a
?
Fa^Fx
L
[i\Lav
(AW \^'LJ
can be formed out of two factors of a gluon field-strength tensor F^v or its dual F%v. The spin, parity, and charge conjugation carried by these these l l operators are respectively JPC = 0 + + , 2~~~ ~, 0~+, 2~ + , and are thus the quantum numbers expected for the lightest glueballs,* i.e. such operators acting on the vacuum state produce states with these quantum numbers. Although there is no a priori guarantee that one obtains a single particle state (e.g. a 2 + + operator could in principle create two 0 + + glueballs in a .D-wave), the simplicity of the operators leads one to suspect that this will be the case. There is one, somewhat controversial, construct missing from the above list. Two massive spin-one particles in an S-wave can have JPC = 1~+ as well as JPC = 0"f+,2+"l~, and some models predict such a gluonium state. However, a 1~+ combination of two massless onshell vector particles is forbidden by a combination of gauge invariance plus rotational symmetry [Ya 50]. The lack of a 1~+ gauge-invariant, two-field operator is an indication of this. Aside from a list of quantum numbers and some guidance as to relative mass values, theory does not provide a very clear profile of gluonium phenomenology. Lattice-gauge methods offer the most hope for future progress. At present, they predict in a quarkless version of QCD that the lightest glueball will be a 0 + + state of mass 1.2 ± 0.3 GeV and that the 2++ glueball is 1.5 ± 0.1 times heavier [BaK 89]. Gluonium states would be classified as flavor-5C/(3) singlets. One process expected to lead to direct production of glueballs is the radiative decay of J/ift, which takes place in QCD through the annihilation process cc —> jGG (as in Fig. XIII-4(c)). The two gluons emerge in a color singlet configuration, and by varying the energy of the final-state photon, all masses between m = 0 and m ~ Mj/^ can be probed. A glueball should thus be revealed as a resonance M in the decay J/ip —• 7M. Of course, QQ states can be produced as well. In order to distinguish gluonia from such neutral quark states, a measure SM? whimsically called the 'stickiness' of meson M, has been introduced [Ch 84], T
JI^M
p.s. [77 -> M]
Gluonic operators with three field-strength tensors produce states with JPC — 0 ± + , , 2 ± + , 1 ± + , 2 ± - , 3^"~. Because of the extra gluon field, one expects these states to be somewhat heavier.
378
XIII Hadron spectroscopy
where 'p.s.' stands for the available phase space. A glueball would be expected to couple strongly to GG but not to_77, and thus to have a high relative value of stickiness compared to a QQ composite. Let us consider two states, among others under active investigation, which are strongly produced in J/ip —• 7M and have provoked much attention as possible anomalous states. These are the /2(1720) and 7/(1440), also called 0(1720) and ^(1440) respectively. Evidence to date suggests that /2(1720) (i.e. 9) has the properties of a non-QQ system. It is a resonance seen primarily in J/ij; —> 7/2(1720) —> jKK and has been assigned the quantum numbers JPC = 2 + + . The TTTT and 7777 decay modes have also been observed, with the quoted branching ratios ™ - 1.00 : 0.47 : 0.10
.
(4.3)
The dominance of KK is even more striking when one notes that D-wave phase space favors the TTTT mode by a factor ~ 2.6. The QQ 2++ ground states are /2(1270) and /£(1525), with dominant components (uu + dd)/\/2 and ss as deduced from their decays. If 9 were a QQ state, it would be a radial excitation of the j ^ and fy. However, the 9 is too close in mass to the /2(1525) to be its radial excitation, and the 9 decay pattern requires its interpretation as largely ss, so that it cannot be the radial excitation of the /2(1270). Besides, there is some experimental evidence identifying the radial excitation of /2(1270) at 1800 MeV. The 9 has not been seen in other hadronic reactions where /2(1270) and /2(1525) stand out strongly, and its stickiness is remarkable, Sf : Sr : Se = 1 : 3 : > 20 .
(4.4)
Even if one accepts that 9 is not a QQ hadron, a firm identification of its primary content is still difficult. The high stickiness and lack of any known multiplet partners favor a glueball interpretation, but its decays do not seem SU(3) symmetric. Additional knowledge of the JPC = 2 + + spectrum and further experimentation will be required to clarify this issue. Indeed, a new analysis [Du 92] reports evidence that JPC = 0 + + for the 9. If this turns out to be correct, the difficulties of accommodating the 9 as a QQ state may not be as severe. The other interesting glueball candidate is the iota, ^(1440). It is the JPC = 0~+ state with the largest branching fraction in J/ip radiative decays, but appears as a rather obscure resonance in hadronic reactions. In comparison with ry(549) and 7/(960), it has stickiness ratios S^ : Srj> : SL = 1 : 4 : > 45 ,
(4.5)
which would favor a glueball interpretation. However, the situation regarding hadronic experiments which probe the JPC — 0~ + spectrum in the range 1.0-1.7 GeV is presently so confused that even QQ states cannot be identified with any certainty. As with all gluonium states, the
XIII-4 Nonconventional hadron states
379
situation will continue to be distressingly vague without a good deal of additional experimental and phenomenological guidance. Additional nonconventional states There is a widespread belief that gluonium states must appear in the spectrum of the QCD hamiltonian. For other kinds of nonconventional configurations, it is far more difficult to reach a meaningful consensus, although experimental efforts to detect such states continue. Let us briefly review several such possibilities. (i) Hybrids: Prom Table XIII-4, we see that among the QQG meson hybrids is one with the quantum numbers JPC = 1~+. This wouldbe hadron is of particular interest because comparison with Table XI-3 reveals that it cannot be a QQ configuration. Model calculations suggest that the lightest such state should be isovector, with mass in the range 1.5-2.0 GeV, and that such states may largely decouple from L = 0 QQ meson final states. A study of Q^G baryon hybrids reveals that none of the states is exotic in the sense of lying outside the usual Q 3 spectrum [GoHK 83]. (ii) Dibaryons: The most remarkable aspect learned yet about the dibaryon states is how much six-quark configurations are restricted by Fermi-Dirac statistics. Table XIII-5 lists the possible six-quark SU(3) multiplets along with their spin values [Ja 77]. Of this collection of states, most attention has been given to the spinless 5C/(3)-singlet state, called the H-dibaryon. This particle, which has strangeness S = — 2 and isospin / = 0, is predicted to be the lightest dibaryon, and to be unstable to weak decay. Although evidence for the H is limited to observation of a neutral object decaying to p + E~ [ShSKM 90], experimental searches continue. (iii) Hadronic molecules: Together with the glueball candidates discussed earlier, another possible interpretation of observed particles as Table XIII-5. Spectroscopy of six-quark configurations. 5/7(6) of color-spin 5(7(3) of flavor Spin 490 896 280 175 189 35 1
1 8 10 10* 27 35 28
0 1,2 1 1,3 0,2 1 0
380
XIII Hadron spectroscopy
nonconventional hadrons occurs with the isovector ao(98O) and isoscalar /o(975) mesons. Nominally, these particles have the quantum numbers of "Hie L = 1 sector of the QQ model, and their near equality in mass suggests an internal composition similar to that of the p(770) and u;(783), i.e. orthogonal configurations of nonstrange quark-antiquark pairs. However, among properties which argue against this are their relatively strong coupling to modes which contain strange quarks, their narrower-thanexpected widths, and their 77 couplings1 The proximity of the KK threshold and the importance of the KK modes has motivated their interpretation as KK molecules [Wil 83]. Unfortunately, interpretation of scattering data near the 1 GeV region is confused, with even the very number of isocalar states in question [AuMP 87]. Problems 1) Power law potential in quarkonium Consider an interquark potential of the form V(r) — crd. a) Use the virial theorem to determine (T)/(V) for the ground state. b) Given the form E2s - Eis = f{d)M-dl^2+d\ where M is the reduced mass, determine d from the observed mass differences in the cc and bb systems, using Eq. (3.1) to supply heavy-quark mass values. c) Assuming this model is used to fit the spin-averaged ground state cc and bb mass values, determine v 2 /c 2 for each system. 2) Quarkonium annihilation from the 1So state Modify Eq. (1.20) to obtain the leading-order contributions appearing in Eq. (1.21). 3) Spin-dependence and light P-wave mesons a) Numerically determine the quantities m, ms_o,rnten of Eq. (1.10) using n = 1 T P-wave mass values. b) Repeat this evaluation for light P-wave mesons, but include a lP\ state in your analysis. Thus, you must generalize Eq. (1.10) to include a spin-spin contribution mss. Choose mass values according to the assignments
f
(3P2) (3Pi) (3P0) .
Comparison of the results with the QCD Breit-Fermi interaction of Eq. (2.10) reveals the real world to have a smaller spin-orbit effect than is (at least naively) anticipated from this model.
Problems
381
4) Mass relations involving heavy quarks a) Repeat the analysis of Eq. (3.1) but using the masses of the charmed/strange mesons DS,D*S instead. Infer a value for M/Mc by referring to the result obtained in Eq. (2.17). Compare with the determination of Eq. (3.1). b) Extend the procedure of Eqs. (2.20-2.25) to isospin-violating mass differences of c-flavored and 6-flavored hadrons.
XIV Weak interactions of heavy quarks
Heavy quarks provide a valuable guide to the study of weak interactions. Measurements of decay lifetimes and of semileptonic decay spectra of heavy, flavored mesons* yield information on elements of the KM mixing matrix, as does the observation of particle-antiparticle mixing like that in the Bd-Bd complex. Finally, detection of CP-violating signals in heavyquark systems has the potential to become the breakthrough that has been sought since the discovery of this phenomenon in the kaon system. XIV—1 Heavy-quark lifetime and semileptonic decays Of the weak interaction properties associated with heavy quarks, the lifetimes and semileptonic decays are the most amenable to theoretical analysis, and it is with these that we begin. The spectator model Consider the weak beta decay, Q —> qeve, of an isolated heavy quark Q into a lighter quark q. By analogy with muon decay, this proceeds with decay rate (if radiative corrections are ignored) _
5 G\^F" mlQ
/0VmQ) > f(x) = 1 - 8x2 + 8x6 - xs - 24x 4 lnx ,
(1.1)
where j{x) is the phase space factor the first terms of which were already encountered in our discussion of muon decay in Sect. V-2. Under what circumstances would this be a good representation for the beta decay of a heavy meson containing quark Ql For it to be accurate, the final state must develop independently of the other (so-called spectator) quark * Note that the conventions of [RPP 90] imply that the quantum numbers of the neutral mesons are K° = {ds),D° = {cu),B° = (db) and B° = (s6).
382
XIV-1 Heavy-quark lifetime and semileptonic decays
383
in the heavy meson. Experience with deep inelastic scattering suggests that this occurs when the recoiling quark q carries energy and momentum larger than typical hadronic scales, i.e. in the range Eq > 1-1.5 GeV. In D decays, the average light-quark energy is (Eq) ~ mo/S ~ 0.5 GeV, so that this approximation is suspect. It should be considerably better in B decays, but still not perfect. Let us first explore the consequences of adopting the spectator model. If we neglect KM suppressed modes, the main decay channels for b quarks are b —> cud, ccs, c£i?£ (£ = e,/i, r), while for c-quarks they are restricted to c —• sdu, SJLV^, seve. Relative to the lepton modes, each hadronic decay channel picks up an additional factor of 3 upon summing over the final state colors. Two of the J3-meson final states (ccs and CTUT) have significant phase space suppressions (reducing them to about 20% of the cud mode) due to the heavy masses involved. The simplest spectator model then predicts branching ratios Sv e X
=* r ^
3 + 2
= 0.2 , x
(1-2)
3 x (1 + 0.2)+ 2 + 0.2 where X denotes a sum over the remaining final state particles. Also, this picture predicts the absolute rates of the D and B decays to be
TD = [s^fer I*y2/(*c)l
- 1-1 x 10"12s , 1
.„
__io
(l-3a) 0.05"
vch
(1.3b)
where f(xc) ~ 0.7 and /(#&) ^ 0.5 are phase space corrections. For definiteness, we have taken mc — 1.5 MeV and nib = 4.9 GeV in the above. However, note the quintic dependence on quark mass; the B lifetime prediction would be 10% lower if ra& = 5.0 GeV were used! For D decays, the D+ and JD°lifetimes differ by a factor of 2.5, rD+ = (10.62 ±0.28) x 10~ 13 sec , 3 2 + = 2.52 ± 0 . 0 9 , rDo
(L4)
whereas the spectator model requires them to be equal. This failure is not surprising, as the D meson mass lies in the region of strong hadronic resonances and final state interactions seriously disturb the spectator picture. Thus we expect the spectator model to reveal only gross features of the D system. It is remarkable, given its simplicity, that the spectator model predicts (roughly) the correct magnitudes of the lifetime and of
384
XIV Weak interactions of heavy quarks
the inclusive branching ratios, B r D o ^ e X = (7.7±l.l)%,
BxD+^ei?eX = (19.2 ± 2.2)% . (1.5) +
We see that the decays of the D correspond more closely to the spectator predictions than do those of the D°. For B mesons, the spectator model is likely to be close enough to reality that some small extra effects can be added in order to render it more realistic. These include QCD corrections of two kinds. First, the nonleptonic hamiltonian picks up short distance corrections of the form described in Sect. VIII-3. These are smaller in magnitude for b decays, because the strong coupling as is evaluated at a higher mass scale. In addition, there are also the QCD radiative corrections associated with the decay of the heavy quark, including bremsstrahlung from the quarks. Besides the QCD correction, one can add bound state corrections to account for the fact that the b quark is not sitting at rest in the B meson, but has some spread in its momentum space wavefunction. The combined effects of all these corrections produce small modifications in the predictions, which are not easy to express analytically. However, for a numerical example we cite a model [AlCCMM 82] which employs a Gaussian momentum distribution for the motion of the bound quarks and includes QCD radiative corrections, obtaining 2
0.05 1.2 x 10"12sec . BrB-+epex ^ 13% , and rB =
(1.6)
Observe how close this QCD corrected version of the spectator model is to the free quark result of Eq. (1.3b). The leptonic branching ratio is close to the experimental value* [St 90] BiB->evex = (10.8 ± 0 . 5 ) % , while the measured lifetime [Dan 92] r*xpt = (1.26 ± 0.07) x 10"12 sec
(1.7) (1.8)
can be used in Eq. (1.6) to extract a value for the KM matrix element, |F cb | -0.049 .
(1.9)
Beyond the spectator model What are the possibilities for making an estimate which is better than the spectator approximation? For the total rate, there is absolutely no * There remains some model dependence in the experimental result due to the need to separate the signal from the B —• DX, D —*• Xev background. The number quoted uses the [AlCCMM 82] model in this process.
XIV-1 Heavy-quark lifetime and semileptonic decays
385
hope of reliably calculating and summing all the individual nonleptonic decays. We must be content either with the spectator model or with very crude calculations of some two-body modes. For semileptonic decays, the situation is somewhat better. The data show that the quasi-one-hadron states, i.e. D —> Keve, K*eve and B —• Deve, D*eue^ form the largest component of the semileptonic rates [Fu et al. 91],
( L 1 0 )
These transitions can be addressed by quark model calculations, so that we have an independent handle on such decays. The hadronic current matrix elements are described by form factors such as (K-(p')\s7,c\D0(p))
= / + ( p + j/) / J + / - ( p - j / ) M ,
(K*-(p') \-S1,c\ D°(p)) = ige^pe" {p + p'f {p - p'f , (K*-(pf)
\s7li75c\ D°(p)) = / l€* +e*-q
(1.11)
[f2 (p + p')fi + / 3 q,] ,
with analogous definitions for the B decays. All form factors are functions of the four-momentum transfer q2 = (p — pf)2. The physics underlying these form factors is two-fold: 1) If the final state meson does not recoil, the amplitude is determined by a simple overlap of the quark wavefunctions, as described in Sect. XII2; 2) As the final meson recoils, the wavefunction overlap becomes smaller, so that the form factors fall off with increasing recoil momentum. For D decays, the KM element is known to a high degree of accuracy from the unitarity of the KM matrix. In this case, the quark model calculations serve to check whether the experimental rate can be reproduced. Most calculations do well for D —> Keve, while D —> K*eve occurs at a rate about one-half of theoretical expectation [St 90]. For B decays involving the b —• c transition, one may assume that the models continue to be valid and thereby extract the value of V^b- Fortunately, various models seem to agree with each other and with the relative amounts of D*/D production, leading to the value [WiSB 85, IsSGW 89, K6S 88], \Vch\ =0.046 ±0.007 ,
(1.12)
where the range of model dependence has been folded into the quoted error.* A preliminary attempt [Ne 91] using the heavy quark symmetry * Implicit to this analysis is the assumption that Br [T(45) —> BB) = 100%.
386
XIV Weak interactions of heavy quarks
relations, to be discussed in Sect. XIV-2, also yields an identical result. This value agrees with the estimate obtained previously from the lifetime. It can be used to imply the following constraints on the KM and Wolfenstein parameterizations: s3 + s2ei6
=0.046 ±0.007
and
|A| = 0.95 ±0.14 .
(1.13)
In the case of nonleptonic B, D decays, we have considerably less confidence in our ability to predict the decay amplitudes. This is especially true in D nonleptonic decay because the rescattering corrections required by unitarity can play a major role. Unitarity predicts (cf Eq. (C-3.14)) for the D —> / matrix element of the transition operator, Uf%^D
,
(1.14)
where n are the physically allowed intermediate states. The scattering matrix elements are evaluated at the mass of the Z?, which happens to lie in an energy range where many strong resonances lie. The scattering elements Tn-+f are therefore expected to be of order unity, implying that rescattering can mask the underlying pattern of weak matrix elements. This makes calculation of D decays particularly suspect. In view of the large B mass, the situation may be better for B decays since scattering amplitudes for decay into individual modes fall off at high energy. A model based on the vacuum saturation method [BaSW 87] may thus prove useful in predicting nonleptonic B decays. Inclusive vs exclusive models for b —> ceue As can be seen from the equality of |Vcb| as extracted either from the spectator model or from quark model estimates of individual semileptonic modes, the two types of analysis give very similar results for the b —> ceve transition. At first this might seem surprising, since the models are quite different. However, the following observation [ShV 88] lends plausibility to the agreement. Consider the semileptonic decay of a heavy quark into another heavy quark, Q\ —• Q2zve, such that their mass difference Am is small compared to the average of their masses ((mi +rri2)/2 ^> Am), yet large compared to the QCD scale (Am » AQCD)- Because of the second condition, one might use the spectator model result,
-m-HlW ,
(1-15)
where V12 is the appropriate weak mixing matrix element. However, if the first condition is satisfied, the quark recoil will be nonrelativistic. This leads to a nonrelativistic calculation of the transitions from a pseudoscalar Q\q state to pseudoscalar and to vector Q2q states. In this limit,
XIV-1 Heavy-quark lifetime and semileptonic decays
387
> ty^x is proportional to the normalization operator, while the axial current ^27i75^i —* tyfti^x is proportional to the spin operator. For states normalized as one then has
(7
\fc\
°~)= 2m , = 2m e}(p') ,
where m is either mi or m2. This translates into invariant form factors = (P + P')» , (1.18) which are the correct relativistic results. Using these to calculate the semileptonic decays, one finds G2
Comparing these, one sees that the sum of the pseudoscalar and vector widths exactly saturates the spectator calculations. In this combined set of limits, it seems that both types of calculations can be valid simultaneously! Application of this insight to b —> ceue decays is somewhat marginal, as the nonrelativistic condition is not well satisfied. A velocity as large as v = 0.8c is reached in portions of the decay region, although on the average a lower value is obtained. However, it is likely that the near equality of spectator versus quark model results is a remnant of the situation described above. These ideas will be extended in Sect. XIV-2. The b —> ueue transition The physics of the b —• u semileptonic transition is not as simple due to the large energy release. In the b —• uev e transition, the allowed range of the energies of the electron and the up-quark are shown in the Dalitz plot of Fig. XIV-1. The energy release is such that the free quark picture should be a reasonable approximation. The only questionable place in the plot occurs in the lower right-hand corner, where the relative momentum of the spectator quark and of the recoiling up-quark is smaller than 1 —• 1.5 GeV/c 2. Here one might expect that bound state effects could become important. Spectator models have a smooth hadronic mass
388
XIV Weak interactions of heavy quarks
distribution instead of having resonances in the low mass region. The effect of the spectator assumption is to smooth over the resonances which occur in this region of phase space. Because Vch ^> Kib> it is difficult to experimentally observe the b —> ueve transition. There are two possible strategies. In one, only the highest energy electrons are looked for. Above Ee = 2.4 GeV, there can be no electrons from b —> ceve because the heavier c-quark mass kinematically rules out these energies. Thus the endpoint region Eo = 2.4 —> 2.6 GeV is uniquely sensitive to the b —> ueve transitions. Alternatively, one can look for the exclusive noncharmed decays B —• 7rePe, peue, 7T7rei/e, etc. In employing this latter technique, comparison with exclusive quark model calculations is necessary. Unfortunately, the exclusive quark model b —» u predictions do not agree with each other as well as do the b —> c results, e.g. the range of discrepancy in the B —• 7rePe transition is a factor of 7. For the method using the electron endpoint region, the exclusive quark model calculations are not sufficient. Note from Fig. XIV-1 that the endpoint region involves both large and small values of u-quark recoil. These configurations occur when the u and ve are produced nearly collinearly, recoiling against a high energy electron. For small u recoil, the final state may well be largely a single meson, with the u-qnavk combining with the spectator quark. However, for large recoil, one expects significant inelastic nonresonant contributions, of which B —• 7T7reue is the simplest example. Quark models cannot hope to individually calculate and then sum these contributions, any more than they could sum the individual hadronic components of a quark jet. However, the spectator model can be applied to the endpoint spectrum. By definition, this model sums over all final states. The QCD and bound state corrections have been applied to the b —• uev e case [A1CCMM 82] and in particular to the endpoint region. As noted above, this method is most reliable for the largest i/-quark recoil, when it is most realistic to have its production independent of the spectator quark. Alternatively, a hybrid approach has u ev
m—.—•D
Fig. XIV-1
I
Q
Kinematically allowed energies in b —> ueve
XIV-1 Heavy-quark lifetime and semileptonic decays
389
been developed [RaDB 90] which uses the exclusive B —> Meue (M = meson) calculations for energies below 1.5 GeV and a spectator model for larger energies. This is an attempt to use both frameworks in the region in which they are valid, and thus provides a check on the importance of low energy effects on the endpoint spectrum. The hybrid method and the pure spectator approach yield endpoint spectra which are in agreement in shape and magnitude to about 30%, indicating that the basic theoretical uncertainties in the endpoint region are not large. We have seen that there are two methods of approaching the b —> ueve measurements, and two corresponding theoretical styles of analysis: (i) exclusive decays such as B —> peue can only be calculated using exclusive quark models, while (ii) inclusive measurements of the electron endpoint region can only be addressed in inclusive models such as the spectator or hybrid models. At the time of this writing, there is progress on both approaches. A few exclusive b —> u events have been reported, although not yet enough to extract rates. In addition, there are indications of b —> u events using the endpoint method [Fu et al. 90], [Al et al. 90]. Prom these, one extracts the ratio -0.10 ±0.03 , (1.20) Tr Kb where the model dependence of the two inclusive methods has been folded into the error estimate. Thus a constraint is obtained on the KM angles, 3 + S2 e%6
= 0.45 ±0.14 ,
(1.21)
or equivalently on the Wolfenstein parameters, v/p2 + r/2 = o.45±O.14 .
(1.22)
The latter is a particularly interesting constraint on the CP-violating parameter 77. The top quark Compared to the other 'heavy' quarks, the top quark will present a rather novel decay pattern. Because nit > M\y +ra&and the KM element |Vtb| is near unity, the dominant decay is the semiweak transition t —• b + W+. The amplitude and transition rate for this process are
llVtb12 g(xw, xb) = (1 - (xw + xfc)2)1/2(l - (xw -
390
XIV Weak interactions of heavy quarks f(xw, xb) = (1 - x2h)2 + x2w(l + x2h) - 2x$v , an
=
(1.23)
where xw = Mw/mt d ^6 m^/rrit. For the range of top quark masses, rrit = 100, 150, 200 GeV, the width is quite large, Tt^bW+ = 0.093, 0.87, 2.4 GeV respectively and corresponds to a lifetime of r = (70., 7.6, 2.7) x 10~25 sec. For such a large t-quark mass, the emitted VF+'s will be predominantly longitudinally polarized, exceeding production of transversely polarized W + 's by a factor ~ m2/M^r. This is a reflection of the large Yukawa coupling of the t quark to the (unphysical) charged Higgs scalar which becomes the longitudinal component of the W+. Other decay modes of the t quark will be highly suppressed by weak mixing factors, e.g. for the mode t —• s + W+ the suppression amounts to | Vts/Vtb>|2 — 2 x 10~3. However, decays of the top quark would be quite sensitive to the presence of physics beyond the Standard Model, such as the occurrence of sufficiently light charged Higgs bosons in models with an extended Higgs structure. An interesting consequence of the large t —• b + W + quark decay rate is that for large values of m*, there will not be sufficient time for the top quark to form bound state hadrons. In view of the large top quark mass, the ti system (toponium) is nonrelativistic and sits in an effectively coulombic potential, V = — 4a s /3r. In the ground state, one finds the quark velocity vrms = 4a s /3 and atomic radius r0 = 3/(2asmt). A characteristic orbital period is then t = 27rro/frms — 97r/(4:a2smt). Using as(r0) = 0.12, we estimate t = (32, 22, 16) x 10~25 sec for mt = 100, 150, 200 GeV. In contrast, the toponium lifetime would be one half the t lifetime given above, since either t or t could decay first. These comparisons imply that a heavy top quark has an appreciable probability of decaying before completion of even a single bound state orbit. An equivalent indication of the same effect is the observation that the toponium weak decay width (twice that of a single top quark) becomes larger than the spacing between energy levels, such as E2s ~ Eis = a2mt/3 = 0.50, 0.75, 1.0 GeV for mt_= 100, 150, 200 GeV. The production cross section, say in e +e~ —• tt, will not then occur through sharp resonances. Instead, there will exist a rather broad and weak threshold enhancement, due to the attractive nature of the coulombic potential [FaK 88, StP 90]. This permits the production and decay of top quarks to be analyzed perturbatively, with I\ serving as the infrared cutoff. A heavy top quark can then provide a new laboratory for perturbative QCD studies. Our lack of understanding of the large top mass illustrates how little we actually know about the mechanism of mass generation. If all fermion masses arise from the Yukawa interaction of a single Higgs doublet, then the Yukawa coupling constants must vary by the factor gt/ge — rut/me > 1.7 x 105. There is nothing inconsistent about such a variation, but it is so
XIV-2 Weak decays in the heavy-quark limit
391
striking as to beg for a logical explanation, one which is presently lacking. On a more practical level, we shall see later in this chapter and also in Chap. XVI how the top quark mass contributes as a parameter to many of the flavor-changing decays and radiative corrections in the Standard Model. Given the present range of top quark mass estimates, various predictions sensitive to rrtt are typically quoted as bands of numerical values. Upon discovery of the top quark and measurement of its mass, it will be possible to test the Standard Model with greater rigor. XIV-2 Weak decays in the heavy-quark limit The discussion of the previous section leaned heavily on the use of models to describe quark weak decay. However, some aspects of weak transitions can be obtained in a model independent fashion through the use of the rrtQ —> oo limit which was introduced in Sect. XIII-3. While the full consequences and limitations of this method are not yet clear, it provides a variety of qualitative and quantitative insights of considerable value. The heavy-quark approximation manages to justify many results which have become part of the standard lore of quark models. For example, consider the decay constant of a Qq pseudoscalar meson M, (0 \q(x)r-y5Q(x)\ M(p)> = iV2FMp» e-^x
.
(2.1)
In the quark model one finds that FM OC (TTIM) ' • This follows from the normalization of momentum eigenstates, <M(p')|M(p)> = 2Ep6^(p - p') ,
(2.2)
such that (0|97075Q|M(0)) = , _
,_
,_
(decay const defn.) ,
(quark model rein.) , (2.3) where i/>(0) is the Qq wavefunction at the origin and Nc is the number of colors. Since as TUQ —> oo, the Qq reduced mass approaches a constant value fi —• mq, we expect that ^(0) itself approaches a constant in this limit,* and the scaling behavior FM OC (UIM)~1/2 then follows immediately from Eq. (2.3). Alternatively, the dependence of FM on TTIM can be derived using the wavepacket formalism introduced in Chap. XII. * For example, in the nonrelativistic potential model, the S-wave wavefunction at the origin is related to the reduced mass by |^(0)| 2 = fi{dV/dr)/2irh2.
392
XIV Weak interactions of heavy quarks
This quark model result can be validated in the heavy-quark limit [Ei 88]. Consider the contribution of meson M to the correlation function
C(t) = ISx (0 |i4o(*,x)4(O)| 0) ,
(2.4)
where AQ = qjolsQ- Inserting a complete set of intermediate states and isolating the contribution of meson M, we have
C(t) = Jd3xj
{2^2E
(0 |4>(*,x)| M(p))(M(p) 14,(0)1 0)+... (2.5)
where the ellipses denote other intermediate states. From the definition of FM, one finds C(t) =
MmMe-imMt
+
_
^
^ ^
2rriM
Alternatively, the heavy quark develops in time in this correlation function according to the static propagator of Eq. (XIII-3.6), C(t) = - ^ 6 - ^ ( 0 1 ^ , 0 ) 7 0 7 5 ^ , C ) ) ( l +75)70759(0)|0) ,
(2.7)
with all the dynamics being contained in the light degrees of freedom. The matrix element is independent of m ^ , and the scaling behavior FM oc (m M )" 1 / 2
(2.8)
follows immediately. This technique is applicable to lattice theoretic calculations of FM- There one considers euclidean (t —> —ir) correlation functions, and identifies the M contribution by the e~mMT behavior. At present, lattice calculations attempting to obtain physical results from the rriQ —• oo limit and from the light-quark limit do not agree in regions of overlap. We therefore feel that it is premature to quote theoretical values of FD, FB- Another piece of quark model lore which can be justified by this correlation function is that the mass difference UIM — TTIQ approaches a constant value in the UIQ —• oo limit. This can be inferred by comparing the exponential time dependences in Eq. (2.6) and Eq. (2.7), and noting that the difference must be be independent of the heavy quark. The heavy-quark limit also makes predictions [IsW 89] for transition form factors between two heavy quarks (which for definiteness we shall call b and c). Recall the lagrangian developed in Eq. (XIII-3.15), the leading term of which is
Cv = hfiiv • D hfi + h®iv - D h® .
(2.9)
This lagrangian exhibits an SU{2) flavor symmetry involving rotation of hy and hv'. It is also spin independent, and thus contains an additional SU{2) spin symmetry. The two 5f/(2)'s may be combined to form an
XIV-2 Weak decays in the heavy-quark limit
393
5(7(4) flavor-spin invariance. Physically, the internal structure of hadrons containing a heavy quark and moving at a common velocity is seen to become independent of the quark flavor and spin. This property leads to many relations between transition amplitudes. An example of a process appropriate for the heavy-quark technique is the weak semileptonic transition B —• D induced by a vector current. For a static matrix element [i.e. both B and D at rest), the weak current transforms quark flavor b —• c, but leaves the remaining contents unchanged, resulting in unit wavefunction overlap. This can be seen calculationally by noting that the time component of the spatially integrated current is the conserved charge of the SU(2) flavor group mentioned above, d3x (D(p') |c(:r)7o&0r)| B(p)) = 6®(p - p') y/AmDmB S
= (P - P') [/+(*m) (rnD + mB) + f-(tm)
(2.1UJ
(mB - mD)} ,
where tm = {TUB — mo)2 is the value of t = (p — p')2 at the point of zero recoil, and the general decomposition of a vector current matrix element, (D(p')\c7lib\B(p)) =f+(t){p + p % + /_(t) ( p - p % -
(2-H)
has been used in the second line of Eq. (2.10). We have seen results similar to Eq. (2.10) in the discussion of the Shifman-Voloshin limit in the previous section. However, there the restriction ms—mi)
(PD + pi,)^ ,
where /#(0) = /D(0) = 1. Let us consider the momentum transfers ts and tsD = (PB — P'D)2 i n terms of the velocities, using p^ = mj pB = m#v and pr> = m^v, we have
{ppDf
l(l-vv')
,
= {PB - PD)2 = (ms - rno)2 + 2mBm,D (l - v • v') .
(2.13)
394
XIV Weak interactions of heavy quarks
If each transition has common velocity factors, the various momentum transfers are related by
to = ^ t o = — (to* - «m) . mzB
(2.14)
TUB
In view of the normalization convention of Eq. (2.2), one must divide the state vector of particle i by \/2rrii (assuming mi » |p|) before applying the b <-• c symmetry. Upon doing so and requiring the resulting expressions to be identical functions of the velocities v and v' leads to the identifications
{D(p'D)\cliC\D{pD)) _ (B(p'B)\Hb\B(pB))
_ (D(p'D\c7ib\B(pB))
After simple algebra, this results in the form factor relations
(2.15)
<2l6)
Although consistent with Eq. (2.13), this manages to separate out /±. The results are expressible in terms of a single function of velocity. It is notationally simpler to express the kinematic dependence using v • v1 instead of t, i.e. fi(t) —• fi(v • v'). Thus, we have
(V • v') = fD(v • vf) = J———f V mB + m D
+(v
where, aside from the constraint £(1) = 1, the function £(v-v') is unknown and must thus be determined phenomenologically. If we exploit the full SU(4) flavor-spin symmetry, then all of the weak current form factors involving B,B*,D and D* can be expressed in terms of £(v • vf), e.g., (D*(p'D) |c7^6| B(pB)) = (D*(p'D) |C7M75&| B(pB)) = JmD*mBZ(v • vr) [(1 + v v% - e* • v t/J . (2.18)
XIV-3 B°-B° and D°-D° mixing
d(s)
w
395
b
Fig. XIV-2 Box diagram contribution to B meson mixing. The symmetry language is appropriate here because, similarly to the symmetry relations detailed in the first part of this book, we have related different processes even though there remains an uncalculable ingredient to be determined from experiment. The limits of validity of these expressions are not yet clear. For example, at very large recoil {i.e. v • v' large), hard gluon exchange could introduce spin dependence into the transition. These limits will be explored more fully experimentally and theoretically in the near future. The real value of the heavy-quark method is that it forms a consistent approximation scheme. The leading order predictions of the type described above appear to bear some resemblance to reality and may end up being directly applicable. However, this will only become certain as the important job of working out the finite-rag corrections is accomplished. There are perturbative corrections which come from interactions of hard gluons [Wi 91] plus 1/m corrections which arise from the expansion of the fields described in Eqs. (XIII-3.14), (XIII-3.15). As with any controlled approximation, it is the size of such corrections which will determine the limits of validity of the results. This process of analyzing corrections is now starting [Lu 90, GeGW 90]. It is likely that the theoretical insights thus gained will strongly influence the phenomenology of B decays, which until now has been dominated by the use of models. XIV-3 B°-B° and D°-D° mixing Just as K°-K° mixing can occur due to the weak interactions, so can mixing exist in the B®-B®, B®-B® and D°-D° systems. We shall cast our discussion of heavy-quark mixing in terms of B° mesons, and return at the end to the D® case. The formalism is the same in all situations and can be taken directly from the discussion of K°-K° mixing in Sect. IX-1. B°-B° mixing The mixing occurring in B° mesons is short-distance dominated. This is because (i) the dominant weak coupling of the 6-quark is to the t-quark, and (ii) the short-distance box diagram (Fig. XIV-2) grows roughly with the squared mass of the intermediate state quarks. Thus the very heavy mass of the top quark enhances this contribution.
396
XIV Weak interactions of heavy quarks
The effective hamiltonians for B® and B® mixing are given by m2tH(xt)VBOBd
^
+ h.c. ,
) 2 m%H(xt)T,BOB> + h.c. ,
(3.1)
OBd = d-y^l + lh)bdrf{\ + 75)6 , 0^^(1
+75)
where TJB — 0.9 is the QCD correction and H(x t) is given in Eq. (IX1.16). The matrix elements of OBd and OBs can be parameterized analogously to that used in kaon mixing,
16 where the pseudoscalar decay constants are normalized as
(0 Idy-fc&l Bd(p)> = iV2FBdPP ,
(3.3)
with a similar definition for Bs. This corresponds to the normalization Fn ~ 92 MeV. Of course our knowledge of FB is not very precise. It is known for a very heavy meson M of mass THM that FM OC (raM)~1//2, but it is neither clear what the constant of proportionality is nor at what mass this scaling law is valid. Model dependent estimates lie in the ranges (74MeV)2 < FldBBd < (200MeV)2 , FB - P - = 1.2±O.1 . FBd
(3.4)
The KM elements which contribute can be estimated using either one of the standard parameterizations, or else the unitarity relations
= - (vchv;d + KbKa)
= - (vchv;s + vuhv^) ~ -vchv;s . The prediction for the overall magnitude of mixing depends strongly on the mass of the top quark. Recall the relation of Eq. (IX-1.11), derived from the diagonalization of the mass matrix, = 2\M12\ ,
(3.6)
which is valid up to exclusion of the factor T12/M12 « 1. This approximation is valid here because Fi2, coming from real intermediate states, does not receive any contributions from the top quark. The ratio of mass differences is independent of the top quark and is largely determined by
XIV-3 B°-B° and D°-D° mixing
397
the KM angles. In the Wolfenstein parameterization, one has
(3.7)
~ (0.034 ±0.009) \l-p-it]\2
.
Since we have constraints on p and r\ from Eq. (1.22), we find > 12 .
(3.8)
It is clear from Eqs. (3.7) (3.8) that the Standard Model requires Bs mixing to be much larger than Bd mixing. Bd-Bd mixing has been observed, with magnitude = 0.70 ± 0.12
(3.9)
Because of the dependence on the top-quark mass and on KM angles, one does not have a firm prediction at present for x&. However, one may verify that the result is consistent with what is currently known about these quantities. For example, the parameter set mt = 160 MeV, p = 77 = — 0.35 and F^ dBsd = F% yields exactly the experimental value for Xd provided the experimental value of F#d is used. This parameter set produces xs = 18. Are there long-distance contributions which can significantly modify Am for B mesons, analogous to those in the K°-K° system? There are several arguments against this possibility. The 'long-distance' effects for Bd are shown in Fig. XIV-3. They involve only the charm or up quarks in the intermediate states. Hence they do not have the factor 7TT| H(xt), growing roughly as m^. This implies that the long-distance effects should be suppressed roughly by the factor (rriB/mt)2 ~ 4 x 10~3. In kaon decays the corresponding mass ratio of (mK/mc)2 is not as important and is in fact overcome by the large A/ = 1/2 enhancement of nonleptonic kaon decays. There is no similarly large nonleptonic enhancement in
D (a)
TT
(b)
Fig. XIV-3 Long-distance contributions to B meson mixing.
XIV Weak interactions of heavy quarks
398
the B system, since the nonleptonic branching fraction is very close to its perturbative estimate. Hence we expect the above estimate of the long-distance suppression to be roughly correct. It is also the case that, in contrast to the kaon system, long-distance intermediate states occur only at the Cabibbo suppressed level for B%. Finally we note that the c and u contributions are accounted for in perturbation theory in the box diagrams. The B mass is heavy enough that the perturbative estimate should not be wildly in error. Given these considerations, we expect the short distance approximation for the B system to be valid. D°-D° mixing The analysis of D°-D° mixing is considerably less clear. The corresponding box diagram is shown in Fig. XIV-4(a), and some possible long distance contributions are given in Fig. XIV-4(b). The GIM cancelation in the intermediate state is between two light quarks d, s (the b effect is suppressed by KM angles). However, there is no compensating large mass factor here, and long-distance and short-distance effects come out at the same order of magnitude. No reliable predictions of Amp can be made. However, the Standard Model clearly requires that ATTID/TD « 1 for the D° system because Amp is twice Cabibbo suppressed (i.e. Amp = O(X2)) while F has no such suppression. Hence, counting KM factors and noting that the GIM cancellation is a measure of the breaking of SU(3) symmetry leads one to estimate that /Ara\
~ A2 x (SU(3) breaking) ~ 10"
(3.10)
Smaller values often emerge in specific calculations. In concluding this section, it is interesting to observe the rather different patterns of behavior which occur in the meson systems exhibiting flavor mixing. The theoretical ratios of long-distance and short-distance
c
_
w
_ u K,(TT)
d,s,b
i K,(TT)
W
(a)
(b)
Fig. XIV-4. Short-distance (a) and long-distance (b) contributions to D meson mixing.
XIV-4
The unitarity triangle
399
Table XIV-1. Patterns of meson-antimeson mixing. I Amiongl
Am
I Ambox I
K°-K° D°-D° B°d-B° B°s-B°
O(l)a O(102)a smalla small0
0.477 ± 0.002 <0.09 0.70 ±0.12 large0
° Theoretical expectation
(box) contributions to Am, and also the magnitude of Ara/F, are summarized in Table XIV-1. XIV-4 The unitarity triangle The B meson transitions form a complex system and provide much of our information on the pattern of weak mixing. The overall B lifetime and b —>• c semileptojiic decays are governed byV^b, the suppressed b —• u modes by Vn\>, B®—B® mixing by 14d> and B^—Bg mixing by V\&. Together with the V^g element, these form all of the 'interesting' sectors of weak mixing. There is a useful pictorial representation of the constraints of unitarity on these elements. Consider the effect of the unitarity relation Of the components to this equation, Vud, Vtd and Vcd are known up to corrections of second order in A = |V^g|, yielding Vuh-\Vch
+ V*d = 0 .
(4.2)
If we treat these elements as complex vectors, this relation is equivalent to a triangle in the complex plane. In the Wolfenstein parameterization the various elements are Vch = -Vts = AX2
- ir,) ,
. A^3u_-e2
Vtd = XSA(1 - p - u/) . (4.3)
^
Fig. XIV-5 The unitarity triangle
400
XIV Weak interactions of heavy quarks
Two possible triangles are shown in Fig. XIV-5 according the assumption that p < 0 (a choice which helps make Xd large) or p > 0. Note that the unitarity triangle can be constructed knowing only the magnitude of the elements | V^bU iKibl a n d l^tdl- The existence of such a closed triangle is independent of the parameterization. Other unitarity triangles, corresponding to the other unitarity constraints, also exist but are either less useful than this one or are equivalent to it [Ja 89]. The unitarity triangle has an interesting connection with CP violation. If the CP-violating parameter rj vanishes, the triangle is reduced down to a line since all the angles go to either 0° or180°. In fact, the area XQA27] of this triangle is exactly the unique rephasing invariant measure of CP violation. The angles y?i, <^25 <^3 are themselves indicators of nonconservation of CP and play a role in the B studies to be described in the next section. It is reasonable to expect that the KM elements can be measured to sufficient accuracy to demonstate that the unitary triangle is not flat. Semileptonic B decays could yield accuracies to ±15% on \VC\>\ and |Vub|? provided some further understanding of the model dependence is achieved. In some sense Vtd is already 'measured' from B^-Bj mixing, but the dependence of mt and FB are too great for a meaningful extraction. However, B®-Bg mixing has essentially the same uncertainties, and they largely cancel in the ratio in Eq. (3.7). If we were able to understand the SU(S) breaking in F^s/F^d^ a measurement of Bs mixing would complete the measurement of the unitarity triangle. XIV-5 CP violation in B mesons The decays of B mesons can exhibit a rich variety of CP-violating signals, some of which are rather large. It is clear that the future will bring experimental efforts to measure these observables. We consider these crucial to the ultimate verification (or falsification) of the origin of CP violation within the Standard Model. Recall that the value of e cannot be regarded as a prediction of the Standard Model because there is an unknown parameter, the KM phase <5, which must be adjusted to fit experiment. In principle the value of e'/e could be regarded as a test, but theoretical and experimental uncertainties are presently too large for this to be practical. There are other candidate theories which could also accommodate the CP-odd signals in kaons without being in contradiction with any existing data. However, the Standard Model, with its single CPodd parameter, makes clear predictions for the patterns of CP violations in B decays, the observation of which would be a major triumph for the Standard Model. There is an important division in the study of CP violations for B mesons into processes which proceed using B°-B° mixing and those which
XIV-5 CP violation in B mesons
401
do not. We shall discuss those involving mixing first, and in the most detail, as some of these are the most reliable predictions of the Standard Model. CP-odd signals induced by mixing General formalism: The time evolution of a B° or B°meson parallels that of a neutral kaon. Denoting Am = ra# — TUL and AF = T# — T^, where H (L) refers to the heavier (lighter) of the neutral B CP-eigenstates, one obtains for states that start out at t = 0 being either B° or i?0,
P= g±(t)
JM12 - iT12 = }.
The strategy for observing CP-violating asymmetries is to compare the decay B°(t) —> / , where / is some given final state, to that of B°(t) —>• / , where / is the CP-conjugate of / ,
\I)=CV\f) .
(5.2)
Let us define the matrix elements
Af) = (f\nw\B°) ,
A(f)= (f\nw\B°) ,
A(f) = (f\nw\B°) ,
A(f)= {f\Hw\B°) ,
and their ratios,*
The decay rates for the two processes are easily found to be [BiKUS 89] ^f oc [a + /?e~ An + 7 e~^An cos Am t
1
\A(f)f|2 - 1 +
+ R« IPU) V
* We caution the reader not to confuse the notation for these ratios with the KM element p in the Wolfenstein parameterization of Eq. (II-4.24)
402
XIV Weak interactions of heavy quarks
"=w)rmi
21
V
- R e | ^ ( / ) j I , (5.5a)
and
~Art
7 e~?Art cos Am t
2"
1+ 2"
1+
(5.5b)
Any observed difference between these two quantities would indicate the presence of CP violation. Before considering some examples, there is a simplifying approximation which it is useful to make. As seen in the previous section, one expects that Myi » Fi2, so it is a good approximation to neglect Fi2 (and hence AF) in almost all cases. The one exception is the semileptonic asymmetry to be discussed below. Even if AF/F were as large as 10%, the exponential factor exp(AFt) would differ significantly from unity only after about ten lifetimes, at which time there would be extremely few particles remaining. In this approximation q/p becomes a pure phase, q/p = el(p, so that \q/p\ = 1. Decays to CP-eigenstates: The most striking processes are those where the final state / is a CP-eigenstate, |/) = ± | / ) , such as / = ipKs, I/JKL, D+D~, 7r+7r~. In this case one has p(f) = l/p(f). Upon considering the total decay asymmetry obtained by integrating the time development
XIV-5 CP violation in B mesons
403
from t = 0 to t — oo, one finds f™dt Af =
dt
(5.6)
1+ where x = Ara/F. This could be non-zero for either (or both) of two
reasons, (i) \*p(f)\\22 ^ 1, (it) Im [*p(/)] ^ 0. The cleanest analysis occurs when |p(/)| = 1, i.e. \A(f)\ = \A(f)\. In this case we find that P
•x*
(5.7)
An example is B® —• ipK®, which proceeds dominantly through b —* ccs, so that p(f) is basically a pure phase
v*vcb
(5.8)
Note that upon neglecting the quantity AF, one has q P
(5.9)
If we then define (5.10) the asymmetry is seen to become Af =
2x si
(5.11)
Thus for a given Ara/F, the prediction is independent of hadronic uncertainties and depends only on the phases in the KM matrix. Certainly this is a remarkably clear result which deserves to be tested. At this stage we can categorize the decays of neutral B mesons to CPeigenstates. For this purpose it is most convenient to use the Wolfenstein form of the KM matrix. In this parameterization, the elements Vtb, Kb, Vts, Vcs are all almost purely real. The Bd and Bs decays can proceed either through the ifM-favored transition b —• ccs or the if M-suppressed transitions b —» uud, b —• ccd, b —• uus. In the former category are included B% —• rfrKs and also B® —> i/j(p, iprj, DfDj. The Bs decays pick
404
XIV Weak interactions of heavy quarks
up no phase since
However, the Bd decay does have a phase since
implying (5.13b) Im pp(^lfs)l = - " ' " * r' 9 7^ 0 . IP J (1 ~~ P) ~^~ V Note that this latter number is exactly sin2
b
= -sin2(^i ~ -0.5 .
(5.14)
J
Also, since Xd is close to unity, the final asymmetry in the decay Bd —> ipKs can be large. This, together with the relative observational 'cleanliness' of the ipKs mode, has justifiably made it the favored decay channel for seeking a CP-violating signal as a probe of the Standard Model. One might at first expect that \p(f)\2 = 1 is automatic if / is a CPeigenstate. However, it is possible to obtain \p(f)\ ^ 1 if there are two different ways to reach the same final state. For example, one could have the decay_B° —> TT+TT" either directly through b —• uud or indirectly through B° —• D+D~ (using b —• ccd) with a rescattering D+D~ —> 7r+7T~. In this case, we obtain I pi&n Ic
, (TT'TT')!
_ . ~ H
_i_ V*,Vu \A- J I piSl ' v cd v cb I ^-lnd | c 5
___\eiS° + VcjVX\Am*\eiS' ,
,w „ ,
——
where <5j and 6D are the strong interaction phase shifts. We see that there are three conditions to have |p(/)| ^ 1, viz. there must be two different paths to the same final state, these paths must have different strong interaction final state phases, and the two paths must also have different weak phases (i.e., arising from the KM matrix).
XIV-5 CP violation in B mesons
405
Table XIV-2. Standard model pattern for CP violation in B decays. Transitions Examples
Im (q/p)p(t) a
b-
sin 2y?i
-> CCS
b- -» ccd b- -+ ?md
Brf —•
ipKs
Bs ~^ ^(f Bd ^ DD Bs
—>•
Bd
- • 7T+7T-
i ^ s —> 7T
b- -> i m s
^ ^ 5 A s
sin 2?i ~0 sin 2y?2 sin 2(^2 sin 2(/?2
Ba-+ir°
The KM suppressed decays can also be analyzed in terms of the angles which appear in the unitarity triangle, and are given in Table XIV-2. Note that the b —> ccd and b —> uud modes can mix with each other, producing the possibility that |p(/)| ^ 1. Unfortunately, reliable predictions of these effects are not presently available. The predictions in Table XIV2 are given for the case |p(/)| = 1. It should also be pointed out that under all circumstances, asymmetries for J5j are suppressed because xs is large and all asymmetries fall off as l/xs for large xs. The physical origin of this result lies with the rapid oscillations in the B® <-> B® system. Regardless of whether one starts out at t — 0 with B® or J3j, after a few oscillation lengths one has roughly equal amounts of B® and B®. This washes out the asymmetry, as expressed mathematically in Eq. (5.6). Decay to non-CP-eigenstates: There may also exist CP violation in final states which are not CP-eigenstates. Consider for example the final state / = D°Ks- Both B% and B® can make transitions to this state, but as shown in Fig. XIV-6 different KM factors contribute. In this case, p
(a)
(b)
Fig. XIV-6 CP violation in B -> D°KS decay.
406
XIV Weak interactions of heavy quarks
and p become respectively
p(D°Ks) = I (5.16) where the ratio A1/A2 of the two reduced matrix elements is not likely to have unit magnitude and will most likely contain a large final state interaction phase. Note that \p(f)\ = \p(f)\> In this case, we have the integrated asymmetry
/o°°* A
D°KS
=
+D°KS
dt
(5.17)
{l-\p(&>Ks)\*) This prediction is not as clear as that for the ipKs mode because unknown hadronic matrix elements contribute. However, a sizeable asymmetry is likely, and the precise prediction could become more certain as one learns more about matrix elements occurring in B decay. Semileptonic asymmetries: For a final example involving mixing, let us consider CP violation in semileptonic decays. In much of our previous analysis, we have neglected the quantity F ^ . However for semileptonic decays, the whole effect vanishes if we neglect Fi2, so we must include it. For this case, only the transitions B° —• £+i/£X, B° —• QrviX (£ = e^/jt^r) can occur. The 'wrong sign' transitions in the time developments, B°(t) —> £~viX, B°(t) —> f+vgX, are then uniquely due to mixing. The appropriate formulas can be obtained from our general result Eqs. 5.5(a), 5.5(b) by the substitutions A(e~) - > 0 A(e+) - > 0
A(e-)p(e~) A(e+)p(e+)
(5.18)
A(e+) = A(e~)
The integrated rate is Cdt\
L
1 p 1 p
• ' _
2 q 2 q
2 2
(5.19)
This sort of CP violation is thus solely sensitive to mixing in the mass matrix, as was the semileptonic K^ asymmetry. Unfortunately it is small
XIV-5 CP violation in B mesons
407
for reasons which have little to do with CP violation. Expanding in powers of Fi2 and defining (fr = arg (F^/M^), one has Fl2
- 2 ~-
(5.20)
Since |Fi2/Mi2| « 1, this asymmetry is always suppressed. For B®, where Fi2 might be non-negligible, there is a further suppression in the Standard Model because the dominant contributions to Fi2 (cc intermediate states) and Myi (ti intermediate states) share the same phase. This is most easily seen using the Wolfenstein parameterization, Eq. (II-4.27), where both are made real. Thus (pr is also suppressed. CP-odd signals not induced by mixing Situations where CP violation occurs without the presence of mixing can occur in B*1 decays through the interference of different decay mechanisms. The requirements are the same as we saw previously in a different context, i.e., there must be two different paths to the same final state, these paths must have different strong interaction final state phases, and the two paths must also have different weak phases. Consider for example the decays B± —> K^TT0, which can take place both through the penguin diagrams of Fig. XIV-7(a) and also through the usual decay process b —• uus of Fig. XIV-7(b). The amplitudes, including the possibility of final state interaction phases, have the form
(5-21) i=u,c,t
where D is the direct amplitude, Pi are the (real) penguin direct amplitudes, and 6U6D are strong phase shifts. Then there can be a CP-
(a)
(b)
Fig. XIV-7 Weak transitions of the (a) penguin, (b) direct type.
408
XIV Weak interactions of heavy quarks
violating decay asymmetry of the form
Im (VisV^sVuh)
sin (S{ - 6D)
\A(B~ -+K-ifl)\2 + \A(B+All of the ingredients for this asymmetry are easily present. However, due to our inability to calculate reliably the amplitudes and phase shifts, no real prediction can be made. That is the generic problem with this class of CP tests. It may well be that a happy accident of two paths of similar magnitude and differing phases will occur, and thus large CP violation signals will be found. However if such effects are not found, we may presently hide beneath our inability to calculate rather than taking the nonobservation as an indictment of the Standard Model. To summarize, we have discussed thus far a variety of possible tests for CP-violating signals in the system of B mesons. The partial rate differences can be quite large. At first, this seems to go against the general dictum that all CP violations in the Standard Model must be proportional to a single, numerically small product of KM angles. However B decays satisfy this stricture in the sense that the mixing and decay of B mesons are in themselves proportional to small KM angles. Overall, the product of mixing, decay and CP violation does turn out to be proportional to all of these KM angles. However, in forming the asymmetry by dividing out the rates themselves, one is canceling the small KM angles, thus leaving a rather large effect. This argument also explains why there is little CP violation in D decays in the Standard Model. The CP observables must be small due to the usual product of KM angles. However, the overall decay rate itself has no small angles, so that the signal remains small. For the complex of Bd mesons, the use of mixing in channels such as ipK$ or D+D~ seems the most promising approach. However, it is still worthwhile to check all possible decay modes. There is a great deal of interest in exploring these decays experimentally, despite a realization that it will not be easy to reconstruct and measure the two-body modes for which predictions have been made. Time-integrated correlations: There is one related issue which is worth discussing as it involves interesting quantum mechanical effects [BiS 81]. We have been discussing the B mesons as if their content at t = 0 were known. This can be accomplished in pair production experiments by watching for the decay of the 'other' B meson which is produced at the same time. By tagging its identity with some characteristic mode, such as b —• c£~ve, one can infer the identity of the B meson. However, let us consider this method in detail for the process e +e~ —• 7* —• B°B°,
XIV-6 Rare decays of B mesons
409
which offers a way to produce B°'swhen running an e + e~ collider at the energy of the T(4iSr), just above threshold. In this case the BB pair must be produced in a P-wave state. If one discusses the system in terms of the mass eigenstates BJJ and BL, Bose statistics require that it be produced in the combination BHBL, because BHBH or BLBL can never be interchanged symmetrically in a P-wave. Phrased in terms of B° and i?°, the state must be a coherent superposition of the form \BB) = - ^ [|£0(p)J3°(-p)> - \B°(-p)B°(p))]
.
(5.23)
Thus correlations are built into the wavefunction, and both B° and B° mesons can oscillate. If one works out the rates for one B to decay into tj;Ks and the other to be tagged by decaying into b —> c£~~U£ or b — one finds using the same assumption as before that
oc sinAm(t -t)Im
(±p(il>Ks)
The correlations in the wavefunction have forced it to be odd under t*-*i. This means that if one were to integrate over all t and t the asymmetry would vanish! This happens for most methods for tagging B mesons in P-wave configurations. However, there are several ways to avoid this problem. One would be not to integrate over time, but rather to directly observe the time dependence. This couldbe done at an asymmetric e+e~ collider which could produce a moving BB center-of-mass such that the B and B decay vertices would be separately visible. Another method would be to go far above threshold, where the two B mesons are produced incoherently. Alternatively one could work near threshold with the reaction e+e~ —• £ £ 7 , so that the BB pair are no longer in a P-wave state. This interesting feature influences experimental search strategies for B meson CP violation. XIV-6 Rare decays of B mesons The system of S-decays is so rich that any single mode will be 'rare' in the sense of having a small branching ratio, and thus will be difficult to observe. Despite this, there are certain classes of decays that are of sufficient interest to warrant both theoretical and experimental study. Specifically, considerable attention has been given to modes that proceed only at one loop, as in Fig. XIV-8. The original hope was that, by measuring the transition rates of such processes, one could extract the mass of the top quark or perhaps observe deviations due to new physics. Those prospects seem less favorable now due to hadronic uncertainties in the transition matrix elements. However, since prediction of rare decays
410
XIV Weak interactions of heavy quarks
involves many of the techniques which we have developed for calculating weak transitions, these decays can provide a nontrivial test of our ability to apply the Standard Model. We shall focus here on one of the most accessible transitions, b —> 57. The determination of other one-loop amplitudes would proceed analogously. The quark transition b —•
57
The process b —> 87 is described by the magnetic dipole transition
x iZ(pa)
(6.2)
with Xi = mf/Myy and x
The flavor content of F
^7
s
t,C
(a)
(b)
Fig. XIV-8 One-loop decay amplitude
XIV-6 Rare decays of B mesons
411
be expressed in the simple form 3Q|F2|
2
^
where f(x) is the phase space factor given in Eq. (1.1), and factors of m s / m 6 arising from phase space and from the amplitude of Eq. (6.1) have been dropped. For mt = 150 GeV, these formulae yield F2 = 0.35 and a relative rate of 0.8 x 10~3, a rather large value. The free quark calculation can of course be improved by QCD shortdistance corrections. These produce a surprisingly large modification to the analysis of b —• 57, and the reason is instructive. The t quark is so heavy that at all scales relevant to the weak decay, its effect may be treated as a point 657 vertex, with renormalizations as in Fig. XIV-9(a). However, the c quark is light on all scales from Mw to ra& so that in its renormalization one must also include the diagrams of Fig. XIV-9(b), where the dot represents the b —> ccs weak Hamiltonian. That is, there is mixing between the b —> 57 vertex and the b —> ccs transition. The renormalization group procedure is similar to that described in Chap. VIII, and has been carried out by several groups [BeBM 87], [DeLTES 87], [GrOSN 90], [GrSW 90]. The result in the leading log approximation is 1 rj? 2
m
t
•
In the last expression, we have employed the form of as for five flavors. With A = 0.2 GeV and p = 1.86, we find F2 = 0.65 [F2(m2t/M^) + 0.79] ~ 0.74 .
(6.6)
The charm contribution has become dominant, and the amplitude has increased by a factor of 2. A consequence is that the dependence on the top quark mass is reduced, resulting in an amplitude change of only ±10% for mt = (150 ± 50) GeV. Unfortunately, to observe the inclusive decays B —> Xs^ at the required branching ratio appears difficult. There are many photons present
t.c
(a)
Fig. XIV-9
9
(b)
Corrections to the b —»
57 vertex
412
XIV Weak interactions of heavy quarks
from background processes and it is hard to detect all states carrying strangeness. In practice one must turn to exclusive channels. The hadron transition B —> K*j At the hadronic level, the quark transition b —> 57 would be observed in channels such as B —• Kirj, KnTT'y, etc. The simplest final state occurs when the Kn system forms a resonant Jp = 1~ state, the if* (890), which in the quark model is treated as a us or ds bound state.* The difficulty lies in relating the quark-level process to real meson transitions. If the spectator model is valid, we do not expect a large fraction of b —> 57 events to end up in B —• K*j. Since the K* emerges at high recoil, |p^* | = 2.6 GeV, the spectator quark must be turned around from sitting nearly at rest to become a part of the final hadron. This is less likely than the production of a kaon plus non-resonant pions. K* production involves either the exchange of a hard gluon to transfer momentum to the spectator or the high momentum tail of the bound state wavefunctions. There is an additional complication. Besides the b —> 57 transition within the hadron, with the largest contribution shown in Fig. XIV-10(a), there can also occur new diagrams, one example of which is shown in Fig. XIV-10(b). Since the low mass charm plus gluon intermediate state was important in the enhancement of F2, this diagram, where the gluon connects to the spectator quark, may also be significant. Unfortunately, this diagram is most likely non-local, and does not appear in the b —• 57 calculation. The effects of it and others like it need to be estimated before a final conclusion can be made about B —> K*j. The B(p) —• K*(\a) + 7(q) vertex can be written in terms of two structure factors Gy and - m2K.) - (p + k)^ e*{k)
w K*
Y
(a)
B |
|G
y (b)
Fig. XIV-10 Contributions to B -> K*j. * The B —• Kj transition is forbidden because it is a spin zero to spin zero transition.
Problems
413
where Gy (GA) is the parity-conserving (parity-violating) amplitude. If the transition is due to the b —> 57 amplitude of Eq. (6.1), these two quantities can be related using the identities
v
w
(6.8)
k) - e [k){p + k)») , to obtain GA = The calculation of the amplitudes is very similar to the B —> pev e transition of semileptonic decay at the highest recoil (lowest q2). Because of this, we should expect that specific model predictions will likely vary over a wide range, since the range of predictions of exclusive channels in semi-leptonic b —> u transitions is large. In fact, one finds that [Wy 89] r
^ X * 7 ~ 0.05 -> 0.46 .
(6.9)
If we accept this rough magnitude, then there is hope that the transition can be seen experimentally. However, the spread in such theoretical estimates is presently too large to expect firm conclusions regarding the extraction of the b —• 57 amplitude from weak decay data. Because of the similarity in the quark model between the processes B —> if*7 and B —• pev e at large p recoil momenta, there is some hope that a comparative study of these two modes may be able to reduce the hadronic uncertainties (cf. Prob. XIV-2). Problems 1) Patterns of CP violation All signals of CP violation involve the interference of two or more amplitudes. Identify the origin of the interference in partial rate asymmetries for the decays (a) Bs —> (pep, (b) Bs —• p±7rZf, (c) Bj —• K*Q(p, (d) B± -> p±7T°, (e) B± - • i^Tr 0 . 2) Amplitude relations in the heavy-quark limit
In the heavy-quark limit, a static b quark in a B meson can be described in terms of just the two upper components of its fourcomponent Dirac field. This can simplify various matrix elements or be used to relate them. Use this feature to show that the B —• if*7 matrix element of the a^v operator, (K*(e, k)\sa^b\B(p)) - d""*? [A e\pp + B e^fy + e+ • p C
414
XIV Weak interactions of heavy quarks
can be related to the vector and axial-vector form factors of B —> p£v£,
= %D e " " a V 4 ^ , = E e^ + e* • p [Fp» + Gfe"] ,
through A = -(E-komBD)/mB , B =-mBD , C = (D + G)/mB , under the assumptions of a static 6 quark and of SU(3) symmetry. In this relation, all form factors must be evaluated at the same momentum transfer, q2 = (p — k)2.
XV The Higgs boson
A central feature of the Standard Model is the spontaneous symmetry breaking in the electroweak sector which gives mass to fermions and to the W± and Z°gauge bosons. The sole physical remnant of this process is the Higgs boson. Although its couplings to the other particles in the theory are fully specified, the Higgs mass is undetermined. As a consequence, efforts to detect this particle cover the widest possible range of mass values [GuHKD 90, Ei 91].
XV—1 Mass and couplings of the Higgs boson Although a complex doublet of Higgs fields is initially introduced in the Weinberg-Salam model, there remains following spontaneous symmetry breaking precisely one physical Higgs state, a neutral scalar particle if0. That is, if we define the number of degrees of freedom for Higgs and gauge boson states respectively as NJJ and NQ, then before the symmetry breaking we have NJJ = 4, NQ = 8 whereas afterwards we find NJJ = 1, NQ = 11. To obtain these values, recall that massive vector particles have three spin components whereas massless vector particles have just two. Although the total of Higgs and gauge-boson degrees of freedom remains fixed (NJJ + NQ = 12), there is a transfer of three states from the Higgs sector to the gauge-boson sector. These Higgs states become the longitudinal spin modes of the W±, Z°particles. This transfer can be displayed analytically by first performing a contact transformation to cast the two complex Higgs states <£>°,?+ in terms of four real fields H° and x = {Xi} (i = 1,2,3)
where U(X) = exp(i* • T/V) , 415
(1.2)
416
XV The Higgs boson
and we recall that v = 1/\/21I2GF — 246 GeV. One completes the procedure with the gauge transformation,
(1-3)
for all fermion weak isodoublets I^L and weak isosinglets I/JR. Within this unitary gauge, the physical content of the theory is manifest, and the quantity <£' is seen to contain a single Higgs field H°. In the following, we shall employ this gauge but with the primes in Eq. (1.3) suppressed. Upon expressing the Higgs potential of Eq. (II—3.19) in terms of the field H°, we find ^
+ \H±
.
(1.4)
The first term in V is a constant energy density which can be interpreted as a contribution to the vacuum energy, (
Although the parameter fx is unknown, we can get a feeling for the scale of the Higgs vacuum energy by supposing ji ~ 0.1 TeV. The Higgs vacuum energy contributes to the cosmological constant of general relativity an amount |AHiggs| = 87rGNewtonf^Higgs = 0.068 cm"2. Such a term would have a remarkable effect on the geometry of spacetime, manifesting itself over a distance scale of |AHiggs|~1//2 — 4 cm! The present limit on the cosmological constant is |A| < 3 x 10~48 cm"2 [RPP 90]. There must then be some important and non-trivial physics which forces the suppression or cancellation of the vacuum energy by roughly fifty orders of magnitude. Also in Eq. (1.4) is the H° mass term, resulting in (1.6) The Higgs mass M# is not fixed because only the quantity v, but not A, is phenomenologically determined. The present lower bound, M# > 57 GeV, comes from measurements at the e +e~ collider LEP [Dav 92]. As shall be described in Chap. XVI, it is possible in principle to constrain the Higgs mass by studying the Higgs contribution to electroweak radiative corrections. Unfortunately, power law contributions of the (unknown) top quark mass tend to overpower the logarithmic Higgs contributions. There exist a number of elegant results on couplings and bounds of very light Higgs [GuHKD 90], but these are now of only historical interest for
XV-1 Mass and couplings of the Higgs boson
417
the minimal Standard Model. There is presently no universally agreed upon upper bound for M#. As we shall describe in Sect. XV-3, if the Higgs mass is large then the theory becomes strongly interacting. This is not a true bound on the mass, but rather a limit on our ability to calculate perturbatively. There are more subtle attempts to bound MH by using the so-called triviality of \(p4 theory [Ca 88]. This uses the result that if one renormalizes the quartic coupling A (cf. Eq. (1.4)) of tp4 theory by using a cut-off, and lets the cut-off become infinite, then one obtains a vanishing renormalized coupling, i.e. Xr —> 0, which implies MH —> 0. In the Standard Model, this final conclusion would be removed by nonzero gauge and Yukawa couplings, but A must not become too large. While extremely interesting, the possible drawback with this approach as regards phenomenology is that it requires assumptions about the presence or absence of new physics beyond the Standard Model. The nature and form of the cut-off depend on higher energy scales, which have not been probed experimentally. It is of course possible, and indeed even expected, that there will be new physics beyond the realm of the Standard Model. Therefore it makes sense phenomenologically to consider all masses up to energies where the Higgs is so broad that the very concept of an isolated resonance fails. The Higgs potential also contains cubic and quartic self-interactions. Of more immediate phenomenological interest is the coupling between the Higgs particle and any fermion / . Prom Eq. (1.3) and Eq. (II-3.20), we deduce the interaction Mw *
(1.7)
where the latter approximate form is a consequence of using 2sin# w ~ 1. The catalog of Higgs particle interactions is completed by presenting its couplings to the W± and Z° bosons. Both trilinear and quadrilinear terms are present, CWWH
= -^W'W^ [Hi + 2vH0] ^ WfW» [2eMwH0
n
L-ZZH =
Si ~l~ 92 ry
g
6^
ryti \ rj2
|/*0 "
(1.8) where we have again employed 2sin# w ~ 1. The preceeding equations reveal the key to phenomenological searches for the Higgs particle at accelerator energies, that its coupling strength generally depends on the mass of the particle with which it interacts.
418
XV The Higgs boson
The naturalness problem Radiative corrections to the Higgs mass raise a question of the 'naturalness' of the minimal Standard Model. For contrast, consider first one-loop electromagnetic corrections to the electron mass. If we impose a cutoff Ae on the momentum flowing through the loop, the mass shift, l + - - l n — 5 - + ... , (1.9) rae,o J 2 7T is obtained. The magnitude of the first order correction is quite modest. Even if we take for Ae the entire mass of the observable universe, Ae ~ 1079 GeV, we obtain only the modest mass shift me ~ 1.7rae?o. This teaches us that, with logarithmic behavior, the renormalization program of absorbing divergences into renormalized parameters is not implausible. However, radiative corrections to the Higgs mass are not as tame. We display in Fig. XV-1 two of the self-energy processes which shift the Higgs boson mass. Considering for definiteness diagram (a), which involves a Higgs loop with quartic self-coupling, we have
This expression is quadratically divergent,* £ # ~ A^, and leads to a shift of the Higgs mass,
If AH is as large as, say, the Planck mass #pianck — 1019 GeV, then in order to obtain a renormalized mass governed by the electroweak scale (MH = O(v) < 1 TeV), the parameter M # o must be negative and have a magnitude which is fine-tuned up to 30 decimal places! While technically possible, this is surely unnatural. This 'unnaturalness problem' has led many physicists to search for alternatives to a fundamental Higgs field, and to suggest that new physics
(a)
(b)
Fig. XV-1 Some quadratically divergent Higgs self-energies. A quadratic divergence also occcurs for the fermion-antifermion loop in Fig. XVI-l(b).
XV-2 Production and decay of the Higgs boson
419
must exist at the TeV scale. For example, one of the proposed models of new physics is that of supersymmetry, in which the quadratic divergence in the Higgs self-energy is removed by cancelations which occur between the contributions of fermionic and bosonic supersymmetric partners [HaK 85]. Interestingly, many supersymmetric models appear to require the presence of two complex Higgs doublets. Of the original eight real fields in the two doublets, 3 become the longitudinal degrees of freedom of the Z° and W± gauge bosons and 5 give rise to physical Higgs quanta. The latter include a charged pair (if ± ), 2 CP — 1 neutrals (H, /i), and 1 CP = — 1 neutral (A). At present, there is no evidence for these quanta as physical particles (LEP data yield mass limits of about 40 GeV on both neutral and charged particles [Dav 92]) or as virtual particles [BaHP 90]. XV-2 Production and decay of the Higgs boson The present mass limit on the Higgs boson (c/. Table 1-4) excludes the region MH < 57 GeV. A bound of this type is obtained by comparing experimental data with theoretical expectations regarding both production and decay of the Higgs particle. In the following, we shall consider the theoretical analysis which underlies this process. Decay Each Higgs particle which is produced in an experiment will eventually decay. Because the Higgs interaction with matter is known, it is generally straightforward to compute the decay rate into any given mode. In principle, one can therefore generate a reliable profile of Higgs decay patterns. Let us begin by considering the decay of a Higgs boson to a fermionantifermion pair. It follows from Eq. (1.7) that the transition amplitude H° —» / + / is expressible in terms of the fermion mass mj and the Higgs parameter v ~ 246 GeV as A
fi(pMq)
,
(2.1)
and leads to the decay rate,
where Xf = mf/Mff and we employ the color factor Nc = 1 for leptons and Nc = 3 for quarks. We see in Eq. (2.2) the expected quadratic dependence of the decay rate on the fermion mass. The decay width for fermion emission is always a small fraction of the Higgs mass for all but the top quark mode, 1 » THQ_^JJ/MH (f ^t).
XV The Higgs boson
420
If sufficiently massive, the Higgs particle can decay into the W± and Z° gauge bosons. For VF-emission, the trilinear coupling of Eq. (1.8) gives rise to the decay rate 1 16TT
Mfj v2
(2.3)
H- For Z° emission, we find a similar but smaller
where xw = expression,
124
- 44
)
(2.4)
where xz = An overall profile of the anticipated Higgs decay width as a function of the Higgs mass appears in Fig. XV-2, where all quarks are taken as free particles. One sees a recurrent pattern, a series of shoulders associated with the opening up of new decay modes and a concomitant rapid increase in the width. Between successive shoulders, there is a steady increase as the Higgs mass increases. At the highest mass considered in Fig. XV-2 (MH — 1 TeV), the Higgs decay width has increased to a value almost comparable with its mass.
Production Of all possible Higgs production mechanisms, we shall focus on just three: (i) emission by the Z° gauge boson, (ii) gluon-gluon fusion, and (iii) fusion. W+W-
1
1
1
1
1
1
1
1
^
0° )
'
D"5
-
10- 1 0
-
10'- 1 5 10
r
T" 1 2
1
10" 1
1
-
-
1
10°
1
1
101
1
1 10 2
1
103
Higgs Mass (GeV)
Fig. XV-2 Higgs decay width as a function of Higgs mass.
XV-2
Production and decay of the Higgs boson
421
Fig. XV-3 Amplitude for Z° -+ H°+ / + / . The production process which has had the greatest impact in the search for the Higgs boson is e + + e~ —• Z°—• H° + / + / ,
(2.5)
where / (/ = i^d, e, ve,...) is any kinematically allowed fermion. The amplitude associated with the production and subsequent decay of H°is displayed in Fig. XV-3. A favorable aspect of this process is the presence of the rather large ZZH vertex. For example, we use the field interactions of Eqs. (1.7), (1.9) to determine the relative strength with which the Higgs boson couples to either an / / pair or a pair of Z°bosons, yielding* 9_ZZH^4MW>>1
QffH
v3 rrif
for all known fermions save the kinematically forbidden top quark. The Z° ~* ffH° transition rate, expressed relativeto that for Z°—> / / , is then computed to be a 4TT sin2 0W cos2 0W J2r
(y — 4r 2 ) + Y L
iZ
6
(2.7) J
where y = 2EH/MZ, r = MH/MZ, 7 = Tz/Mz, and fermion mass is ignored. The magnitude of the ratio in Eq. (2.7) decreases with ME, equaling approximately 10~2 for MH — 0 and 10~4 for MJJ — 40 GeV. This mode has been searched for at SLC and LEP e+e~ colliders, and its non-observation has led to the bound on the mass of the Higgs quoted earlier. Future attainment of greater collider luminosity should allow the Higgs search to continue for even larger mass values. * Specifically, we have taken gffH ^ errif /M\y and have also denned a dimensionless coupling by extracting a factor of Mz and including a factor of 2 to allow for the gzzHMzZ^Z^H0 two ways that the Z°fields can be contracted.
XV The Higgs boson
422
At higher energies, yfs > MZ + MH, the process of Fig. XV-3 can again lead to a search for the Higgs. However, now it is the Z° coupled to e+e~ which is virtual, while the final Z° is on its mass shell. The cross section, - 4 s i n 2 f l w ) 2 ] p2 [p4 sin 4 0 W cos 4 9M A
2
2
2
(s-
(2.8) m
2
with p = (s — m H — m z) — Am Hm z, provides an efficient way to search for the Higgs at higher energy e +e~ colliders. For Higgs production using light hadrons, the coupling between light quarks and the Higgs is so weak that the cross section is extremely suppressed. Surprisingly, however, the gluon-fusion reaction gg —> H is much more favorable. This occurs through the diagram of Fig. XV-4, which includes heavy quarks in the loop. The result is independent of the quark mass for mq ^> MJJ (cf. Prob. XV-1) and can be expressed as a lowenergy effective lagrangian of the form (2.9) the number of
with G" being the gluon field strength tensor and heavy quarks. The decay rate for this process,
(2.10)
72TT 3
is not large enough for gluon emission to be a useful mode of observation. For quarks lighter than MJJ , a local lagrangian does not occur, and the decay rate of Eq. (2.10) is modified to a
i,
(2.11) —
H
Fig. XV-4 Higgs formation via gluon fusion
/
i
XV-3
The possibility of a strongly interacting Higgs sector
423
For the current range of allowed MJJ and mt values, it is the top quark which provides the dominant contribution. Gluon-gluon fusion is of special interest as a mechanism for Higgs production in a collider experiment, and should be the dominant Higgs production mechanism at hadron colliders for M# < 1 TeV. For the scattering oifree gluons at center-of-mass energy s, one obtains the Breit-Wigner resonance cross section (res)
~ " y-~
• -y
~—n~
n
fff - S)
' n~^yy
/ ^ -i c\\
+
2
where q = s/4 and J = 0 for a Higgs resonance. This expression must be folded together with the gluon distributions in the incident hadrons in order to predict the actual yield at a given collider. Numerical studies can be traced through the literature cited in [GuHKD 90]. The W+W~ fusion process W+W~ —> H° is a tree-level reaction (in contrast to gluon-gluon fusion) and is describable using the interaction of Eq. (1.8). The free W-boson cross section would be of the same form as Eq. (2.12) but with THo^ww taken from Eq. (2.4) and now using q2 = (s — 4M^)/4. Again, the W+W~ distributions must be supplied depending on the specific reaction. This is generally calculated in the 'effective W approximation [GuHKD 90] in which Ws are treated as partons within a hadron, being generated by the perturbative coupling to quarks. WW fusion provides an important production mechanism at e+e~ colliders, and overtakes the gluon-gluon fusion at hadron colliders for MH > 1 TeV. XV—3 The possibility of a strongly interacting Higgs sector As the mass of the Higgs boson becomes very large, the analysis of the Higgs sector undergoes a qualitative change. Consider for example the Higgs quartic self-coupling A, which can be expressed as (cf Eq. (1.6)) (31) The quartic coupling is seen to grow as the square of the Higgs mass and is scaled by the energy v. For M# » i/, we enter the domain of strong coupling. A perturbative expansion in A is no longer a sensible procedure, and new methods must be employed. What are the indications that the theory has become strongly interacting? One is the width of the Higgs itself. For large MH , the and Z°Z° decay modes combine to yield the tree level decay rate
424
XV The Higgs boson
When MH ^ 1 . 4 TeV, the Higgs width is equal to its mass, a feature which surely indicates strong interaction and which calls into question even the identification of the Higgs as a resonance. Note that MH — 1.4 TeV corresponds to A ~ 16. Another indicator of strong coupling is the so-called 'unitarity violation'. Actually, any S-matrix element must satisfy the unitarity constraint for all values of the coupling constants and masses. However, this will only occur if the theory is treated to all orders. A perturbative approximation need not itself obey unitarity at a given order. Later in this section (c/. Eq. (3.21)) we shall display a calculation for which, if s ^> M\ ^> M ^ , the tree level T-matrix element for S-wave scattering of longitudinal Z° bosons becomes
This would violate the unitarity constraint (c/. Eq. (VI-4.5)),
|T |2
<3'4)
« < jrzSf '
for MH — 1 TeV , s ^> M | . While this indicates a flaw with the tree level calculation rather than with the full theory, it does suggest that perturbation theory fails at these energies.
The equivalence theorem One might expect that when the Higgs sector becomes strongly interacting, there would be large corrections to low-energy observables. This turns out not to occur. There is a general 'screening' theorem [Ve 77b] which shows that low-energy observables are shielded from large corrections arising from a large Higgs mass. For example, the one-loop correction to the gauge boson mass ratio is only logarithmic, and is made smaller by a factor of OL
12TT 12TTTT TT
( \M\V
In practice, there is no low-energy observable which, at present sensitivities, can detect the effect of Higgs masses up to the strong coupling regime. In order to directly observe a strongly interacting Higgs sector, one needs to use the system most intimately connected with the Higgs doublet, viz. the longitudinal gauge bosons. Recall that in the original symmetric theory, the gauge bosons have only two (transverse) degrees of freedom. The longitudinal component arises only after symmetry breaking, when the gauge bosons absorb three of the four components of the
XV-3 The possibility of a strongly interacting Higgs sector 425 complex Higgs doublet* (3.6) as was displayed in the unitary gauge in Eq. (1.3). In quite a physical sense, the longitudinal components of (VF+, W~, Z°) are due to the unphysical Higgs particles ((^+,
.
(3.7)
This is the key to probing the strongly interacting Higgs sector. Let us use the simple example of H —> W+W~ to illustrate how the equivalence theorem works. The relevant coupling is given in Eq. (1.8), and yields the invariant amplitude MHo^w+w-
= g2MwelfM • e ? .
(3.8)
The longitudinal (e^) and two transverse (e^) polarization vectors are ,±) = - p ( 0 , l , ± i , 0 ) . (3.9) Note the factor of l/Mw in the longitudinal mode. For large Higgs mass, this leads to a great difference in the respective longitudinal and transverse decay amplitudes,
where A, A' are helicity labels, and also in the decay rates, - = {l ~
2x
w)
F
0,
where xw = Mw/MJJ- As a check, observe that these sum to the total rate previously given in Eq. (2.4). The enhancement of the longitudinal/transverse ratio, as scaled by the factor Mjj/Myy, is a general feature * Comparing this representation to that in Eqs. (1.1-1.3), we have in the small \ limit the correspondences \3 —> <^3 and x^ -^ i^-
426
XV The Higgs boson
since the transverse modes are of standard perturbative strength, while for large M# the longitudinal couplings are strongly interacting. In order to test the equivalence theorem, we next calculate the H —•
(3.12)
We then find in the notation of Eq. (3.11), r
#°^+y?- = r 0 .
(3.13)
Up to the presence of M^r/s (i.e. M^/M^j) corrections, this conforms with the equivalence theorem. The equivalence theorem describes a situation almost like that covered by Haag's theorem as described in Chap. IV. Either W^ or y^, properly normalized, can serve as a good field variable for the longitudinal modes; the renaming does not change the physics. The difference with the usual application of Haag's theorem is that more than one field is involved (c/. Eq. (1.3)), and one must show that the extra field only produces suppressed contributions. The most careful proof [ChG 85] is quite involved. However, the essence can be exposed in 't Hooft-Feynman gauge [LeQT 77], wherein the constraint added to fix the gauge (see Sect. XVI3) is dflW£ + iMw
(3.14)
One can define the longitudinal component in momentum space by
W£(*) = e£(*)Wj(fc),
(3.15)
with e^ given in Eq. (3.9). However, at high energy, since
M
+o
we see that
where in the second line we have used the momentum space version of the gauge constraint. This displays the nature of the connection between the longitudinal W's and the Higgs doublet. Note that the equivalence theorem holds at high energy whether or not the symmetry breaking sector is strongly interacting. It may be useful even if a light Higgs boson is found, although special care may be needed in this case [BaS 90].
XV-3 The possibility of a strongly interacting Higgs sector
427
Scattering of longitudinal gauge bosons The best probe of the symmetry breaking sector would be the discovery and study of the Higgs boson. However, if this particle is so heavy that its observation is difficult or even impossible, the use of longitudinal gauge bosons can be valuable. While reactions of these particles can of course be calculated directly, the computations are quite involved and contain subtle cancelations. It is generally easier, and ultimately more instructive, to transfer the problem via the equivalence theorem to the Higgs sector. If we neglect perturbative corrections due to the gauge coupling, the interaction of longitudinal gauge bosons at high energy is governed by the Higgs potential of Eq. (II-3.19). Moreover, since the Higgs potential is in exact correspondence (upon ignoring fermions) with the linear sigma model of Eq. (IV-1.4),* it is seen to have the symmetry structure SU(2)L x SU(2)R (cf Eq. (IV-1.5)). This invariance is larger than the gauge symmetry SU(2)L X C/(l)y, containing an extra custodial SU(2) [SiSVZ 80]. As a result of the SU(2)L X SU(2)R chiral symmetry, we can adopt the effective lagrangian framework which was originally constructed for application to low energy processes to describe the scattering of longitudinal gauge bosons. The primary modification is the replacement Fv ~ 92 MeV —• v = 246 GeV. Note, however, that the Higgs application would if anything be closer to the symmetry limit since
0.015 >
=
f ^ y = 0.0007.
(3.18)
\47TVJ
v
J
2
The O(E ) predictions of the theory are summarized by the effective lagrangian 2
(3.19) where U = exp (ir • (p/v). This result is actually useful in an intermediate range of energies - the energy must be sufficiently greater than M\y for the equivalence theorem to apply, yet below MJJ to allow an expansion in powers of M^2. The simplest manifestation of the strong coupling between longitudinal gauge bosons would occur in WLWL scattering. These results are exactly given by Eq. (VI-4.2,4.7) with m^ —> 0. Written out for the individual * Note that the strict equivalence theorem limit, entailing the neglect of O(mw /E) effects, corresponds to the m^ —> 0 limit of the sigma model.
428
XV The Higgs boson
channels, they are -L ur/+Tj/ —
w+w-^w+w-
- -tf
T
wtz^wizL L
L
L
u
TzLzL-+zLzL = 0 ,
' T
=^ ' _ L
s
wiw--,zLzL = ^ .
(3-20)
V
It is instructive to examine a specific calculation. In the full theory, the leading contribution to elastic ZZ scattering is s
t
u
as a consequence of Higgs exchange in each channel. For Mjj ^> s,t,u, this reduces to + t + u+
—
2
++ . J
(3.22)
The first three terms in this expression sum to give the zero of Eq. (3.20) because kinematically s + t + u = 4 M | which is neglected with respect to s in the usage of the equivalence theorem. The next set of terms of order s1 give the O{E^) corrections to the lowest order effective lagrangian, and would be reproduced through the O(E4) effective lagrangian of the linear sigma model, previously calculated in Eq. (IV-2.10). For example, the amplitude for TW±ZO^\Y±Z° ^S the direct analog of the pion-pion amplitude in Sect. IV-1. Note also that, even though the results of Eq. (3.20) agree with the tree level perturbative calculation, they in fact are the correct results to all orders in the strong coupling when expressed in terms of the renormalized value of v. This is guaranteed by the chiral symmetry analysis, which is fully non-perturbative. The process of WLWL scattering is thus seen to share many of the phenomenological properties of the analysis of TTTT scattering, scaled up in energy by a factor of v/Fn ~ 2700. The lowest order result will violate the simplest consequences of unitarity at energies above 1.7 TeV. This can be corrected order by order in the energy expansion of chiral perturbation theory by including effective lagrangians with more derivatives. Since we are neglecting masses here and using the group /SC/(2), the program involved is even simpler than the hadronic physics described earlier in the book. If it turns out that Nature has indeed chosen a strongly interacting symmetry breaking sector, this program should prove to be extremely useful [Ch 88]. Note that the formalism extends beyond the Standard Model. The lowest order results are constrained by the custo-
Problems
429
dial SU{2) symmetry to be universal.* However, at O(E4) the pattern of the correction to the lowest order results depends on the underlying theory (see Sect. VI-6) and we could hope to distinguish the Standard Model from the symmetry breaking schemes by careful study of longitudinal gauge boson scattering [DobH 89, DoR 90]. Problems 1) Higgs-gluon coupling Calculate the H°—> gg effective lagrangian of Eq. (2.9) using a heavy top quark in the loop. Show that the loop integral is finite and falls with increasing quark mass for mt » ra#. However, show that the result is independent of the quark mass because the coupling is proportional to mt. This violates the naive expectation of decoupling because of the growth of the coupling constant as ra^ —> oo. 2) Equivalence theorem One can see the equivalence theorem at work in the decay t —» 6VF+, which was described earlier in Sect. XIV-1. Assuming a very heavy top quark, calculate the t —> bW^, t —• bW£ and t —> btp + decay rates. Show that the equivalence theorem works for this decay and calculate the O(Mw/E) corrections. Determine the ratio of WL to production.
* The results can also be extended to theories without the custodial symmetry [ChGG 87], but the answers reduce to Eq. (3.20) in the limit p —• 1.
XVI Physics of the W and Z bosons
Because it is a renormalizable theory, the Standard Model is, in principle, capable of making predictions to any level of accuracy. The process of testing the electroweak sector is now fully under way and will continue for many years, much like the systematic exploration of Quantum Electrodynamics. We shall keep our treatment of this already vast field at a relatively simple introductory level, with the intent of helping the reader obtain an overall grasp of the main issues and techniques. XVI-1 Neutral weak currents at low energy Early studies of the weak interactions were confined to processes, like nuclear beta decay and muon decay, which concern just the charged weak current. Starting from the mid-1970's, the field of weak interaction phenomenology was broadened by experiments involving neutral weak currents, most notably [Am et.al. 87] 1) 2) 3) 4) 5)
deep inelastic neutrino scattering from a variety of targets, neutrino-electron scattering, neutrino-proton elastic scattering, parity violation in atoms, and cross sections and asymmetries in ee reactions.
Important work on these low energy {M^vz » |g2|) experiments continues. In addition, electroweak phenomenology has recently been enriched by higher energy (|#2| ~ ^wz) experiments which directly probe the massive gauge bosons, particularly the very accurate Z°studies. Before turning in the next section to a discussion of W± and Z°physics, we shall first review 'traditional' neutral weak current phenomenology. 430
XVI-1 Neutral weak currents at low energy
431
Neutral current effective lagrangians Recall that the neutral weak interaction between the gauge boson Z° and a fermion / is given at tree-level by* 75JJ (/)
02
n (/)
w3
J/) ,
T
(1.1)
(/)
Examples of individual g^Q and <7a0 appear in Eq. (II—3.41). To describe neutral current interactions at low energies, it is convenient to employ an effective four-fermion lagrangian, analogous to the Fermi model of charged current interactions. At tree-level, the Z°-mediated interaction in the low-energy limit is IVI
/,/ where po is the tree-level rho-parameter, C
w,0
Z,0
(1.2)
IVI
Z,0
Comparing the second of the relations in Eq. (1.2) with Eq. (V-2.1), we see that po governs the relative strengths of the neutral and charged weak current effective lagrangians. In the Standard Model, it has the tree-level value unity, PQ ^ = 1. The rho-parameter is important because it is sensitive to the possible presence of physics beyond that encompassed by the Standard Model. Alternative choices for the Higgs structure can lead to different values for po (c/. Prob. XVI-1). Experiments (l)-(5) listed earlier involve neutrino-electron, neutrinoquark, and parity-violating electron-quark interactions. There is an effective lagrangian for each of these, among them
Cvq = - ^ Vtf{\ + Js)ut [e^kaM1 + 75)
432
XVI Physics of the W and Z bosons
where the index a = u,d,... denotes quark flavor. Of course, contributions other than neutral weak effects must not be forgotten, e.g., parityconserving eq scattering also experiences the electromagnetic interaction. In Eq. (1.4), we have implicitly included the effect of radiative corrections and thus omit the subscript '0'. Table XVI-1 gives a compilation ([RPP 90]) of the radiatively corrected coefficients. Observe the presence in Table XVI-1 of quantities pi and /^. At treelevel, they reduce to unity, i.e. p^o — ^i,o — 1- The pi are overall multiplicative factors and the K{ multiply the weak mixing angle, which itself has become renormalized, s^ 0 —> s\. The presence of such quantities in the effective lagrangian can be traced back to the underlying neutral current couplings,
Although the pi and Ki are generally process dependent, they also contain certain common (or universal) contributions, including terms quadratic in the top-quark mass (O(G^,m^)). It is this class of electroweak corrections which shall be emphasized in the discussion to follow. Another unknown and potentially large mass parameter, M#, enters only logarithmically (e.g. <9(ln[M^/Aff ])) at the one-loop level. Determination of the weak mixing angle 0W It is convenient to parameterize neutral weak phenomena in terms of s^, and the value of s^ is often cited as the result of a given experiment. We shall use two examples to illustrate this process. Table XVI-1. Radiatively corrected coefficients. Coefficient
General Form^
nu ^1 °1
( I I I 2 n Peq ^ 2 ' 3 ecl w Peq \2 S^eq^wy
^ Small additive terms are omitted.
XVI-1 Neutral weak currents at low energy
433
(i) Deep-inelastic neutrino scattering from isoscalar targets: Here, one measures the ratios of neutral-to-charged current cross sections, _
a
vN
(1.6)
cc
vN
Under the conditions of 'deep-inelastic' kinematics ([BaP 87], [Fi 89]), theoretical calculations of Rv and Rv are carried out in terms of quark, and not hadronic, degrees of freedom. Although details such as quark distribution functions are beyond the framework of this book [Ro 91], it is plausible that by working with ratios like those in Eq. (1.6), theoretical uncertainites associated with hadron structure tend to cancel. At tree level, Rv and Rv are straightforward to compute if scattering from an isoscalar target is assumed and antiquark contributions are ignored. One then obtains the simple forms (cf. Prob. XVI-2), 1 5 2 4 5 = o ~ w,0+ o(l+ r 0)s W j 0 , wu 4,o
= 2x - ~ ' ' 9 (
1+ r
(1.7)
5
°) w,o >
where r = r 1 = ^ v / a S v a r e measurable quantities with tree-level values ro = f^1 = 3. Radiative corrections produce the modifications [MaS 80]
5
^
('•»>
R i?,0 where KUN and p2vN are functions of the lepton momentum transfer and we have omitted non-factorizable contributions. When comparing the Standard Model with experiment, it is common to find in the literature a variety of values attributed to a given quantity such as s\. This has several causes, e.g. the ever-changing data base, differences in the choice of theoretical parameters like mt* etc. Such is the case for the weak mixing angle, and we cite one just determination One example [Ca 92] of fits to low-energy and high-energy electroweak data which constrain by assuming correctness of the Standard Model gives mt = 144l26l2i GeV , where the latter error bars reflect uncertainty in the Higgs mass and a central value 300 GeV is assumed. Assuming a smaller value of MH lowers the fitted value of mt.
434
XVI Physics of the W and Z bosons
here [RPP 90],
/ °'233± °-003 ± °-005
(mt = 10 ° G e V ) >
inei. - j Q 2 3 ( ) ± ±0 .003 ± 0.005 (mt = 200 GeV) , where the error bars are respectively experimental and theoretical and a Higgs mass MJJ = 100 GeV is assumed. For deep inelastic neutrino scattering, the O{G^ml) effects are somewhat suppressed by a cancelation between the pv^ and KVN contributions. A relatively large theoretical uncertainty arises from the contribution of the c-quark threshold to the charged-current cross section. The (intrinsically nonperturbative) threshold analysis is expressed in terms of a c-quark mass parameter mc. Fits to muon and dimuon production data yield a rather imprecise determination for this quantity, mc = 1.31 ^QAS lFo et al 90 ]* (ii) Atomic parity violation: The Z°-mediated electron-nucleus interaction contains a part which is parity-violating. Consider, for example, the effect in atomic cesium. Because of the weak neutral interaction, the single valence electron in cesium contains small admixtures of P-wave in its 65 (ground) and IS (excited) states. As a consequence, there occurs a measurable IS —> 6S electric-dipole transition from which one can extract information regarding the parity-violating component in the electron wavefunction [NoMW 88]. It is convenient to cast the electron-quark interaction in terms of a hamiltonian density defined in the electron spin space,
£
,
(1.9)
where 75 signals the presence of parity violation and Pnucl(^) reminds us that the electron feels the effect only where the nuclear density is nonzero. The quantity Qw is the 'weak nuclear charge' to which the electron couples, and is given to lowest order by
Qwfi = -2(NuC?fi + Ndpffi)
= Z(l-4slfl)+N
,
(1.10)
where Z and N are respectively the nuclear proton and neutron number. The fact that s^ 0 ~ 0.25 suppresses the proton contribution, leaving the coupling of the atomic electron to neutrons as dominant. Thus, a study of the various isotopes of a given element should prove informative. The C\"d are coefficients whose tree-level and radiatively corrected forms are
\
|4
(1.11)
XVI-1 Neutral weak currents at low energy
435
The extraction of Qw from experimental studies ([NoMW 88]) of Cs133 depends crucially on a theoretical understanding of atomic structure. As such effects become better understood, the central value and estimated uncertainty are modified, e.g., _ f -69.4 ± 1.5 ± 3.8 ~\ -71.04 ± 1.58 ± 0.88
Qw
[NoMW 88] , [MaR 90] ,
where the uncertainties refer respectively to statistical and theoretical contributions. A determination of the weak mixing angle obtained in this manner is [RPP 90] _ f 0.215 ± 0.007 ± 0.017 2 wlatom. par. viol. ~ j Q 2 Q 4 ± Qm? ±
5
(mt = 100 GeV)
where again MJJ = 100 GeV is assumed.
Definitions of the weak mixing angle Thus far, we have stressed the role that the weak mixing angle plays as a parameter in neutral current phenomenology. What is its status in a more formal development of electroweak theory? We shall see in Sect. XVI-3 how the weak mixing angle is ordinarily treated as a derived quantity, expressible in terms of gauge boson masses or electroweak coupling constants. Such relations, introduced at tree-level, are subject to the effects of radiative corrections, and must therefore be considered in the context of a specific renormalization program. We shall briefly describe two such approaches, the on-shell renormalization scheme and the class of renormalizations wherein the weak mixing angle has the status of a running coupling constant. (i) On-shell renormalization: On-shell renormalization [RoT 73, Si 80] fixes the three bare electroweak parameters pi5o, #2,0 &nd ^0 in terms of the physical quantities Mw,Mz, and a (cf. Sect. XVI-3). The on-shell weak mixing angle is then defined in terms of the physical gauge boson masses,
sl = \-M>wlM\ .
(1.12)
Thus, the on-shell weak mixing angle can be experimentally determined not only as a parameter in various neutral weak processes, but directly from Mw and Mz- For the remainder of this chapter, the notation 's^' will be reserved for on-shell renormalization. A great deal of effort has gone into obtaining a precise determination of the weak mixing angle, and the process is continously being updated. The gauge boson mass values in Table 1-4 imply [Ab et al. 90] 4
Z
=
0.227 ±0.006 ,
(1.13a)
436
XVI Physics of the W and Z bosons
where the uncertainty is due to the W± mass. The value cited in [RPP 90], based on an analysis of all data from low energy and high energy experiments and assuming MJJ = 100 GeV, is (mt = 100 GeV) , (mt = 200 GeV) . " \ 0..2189 ± 0.0002 ± 0.0004 Use of the full data base reduces the uncertainty in s^, but introduces dependence on the top quark mass via electroweak radiative corrections. If one treats po as a free parameter in fitting the data by first extracting electroweak radiative corrections and assuming MJJ = 100 GeV, the value _ f 0..2305 ± 0.0002 ± 0.0004
_ f 1.003 ± 0.004 (mt = 100 GeV) , °~~ \ 0.993 ± 0.004 (mt = 200 GeV) . ^'^ is obtained [RPP 90]. This reveals no inconsistency with the basic structure of the Standard Model, so we shall take po = 1 hereafter. Let us return to the subject of electroweak radiative corrections. Contributions to the factors pi and /c» can be classified as either independent of the external fermions (universal) or explicitly dependent on the fermion flavor (nonuniversal), P
« = l+ A p + ( A p ) 2 , where Ap and An denote universal pieces. It should be apparent that W± and Z° propagator corrections, like those in Fig.XVI-1, occur independently of the external fermions and are thus 'universal'. The universal effects are of special interest because they are the primary source of the O(G^ml) radiative corrections [Ve 77a, BiH 87], and in the following we shall approximate (cf. Sect. XVI-5) Ap = (Ap)t + ... ,
AK = 4(Ap)t + • • • ,
(1.16)
where
b
T
(a)
(b)
Fig. XVI-1 Top-quark corrections to the (a) W±, (b) Z° propagators.
XVI-1 Neutral weak currents at low energy
437
Observe in Eq. (1.16) that AK is proportional to Ap. This is a result of the Standard Model; in general, these quantities are independent. In addition to Ap and AK, there is a third electroweak correction, Ar w , which is worthy of mention. This quantity modifies the tree-level relation between the gauge boson masses Mw and M\)
M\
y ^ M K l - Ar w ) '
U WJ
'
and in the Standard Model, one finds (c/. Sect. XVI-5)
Arw = - 4 A p .
(1.19)
S
W
By convention, Ar w does not contain photonic corrections of low frequency (q2 < M|), whose effects instead are contained in G^ and in the running fine structure constant a(Mz)> Like AK, it is proportional to Ap in the Standard Model, but becomes an independent quantity in more general contexts. (ii) The weak mixing angle as a running quantity. A popular approach of this type is modified minimal subtraction (MS). Here, one denotes the weak mixing angle as s^(q2) (or sw(q2)jjg) and adopts the definition [Ma 79, MaS 81]*
We have already seen (cf. Sect. II—1) how modified minimal subtraction can be implemented in dimensional regularization for the electric charge e(q2), and one proceeds accordingly for the coupling #2(2)To fix the running weak mixing angle, it is appropriate to employ measurements taken at the Z° peak, e.g. as produced in the ee reaction ee —• Z° —• / / , where / is a final state fermion.t Denoting all measurements carried out at the Z° peak with a super-bar, one thereby obtains a quantity s^,
4 = 4(Mf) * sl{Ml) ,
(1.21)
and writes for the weak neutral coupling constants,
* Yet another running coupling, s1(q2), appears in the literature [KeL 89]. In the approximation of ignoring the 'non-universal' corrections adopted here, the quantities s^(q2) and s2c(q2) are equivalent. t We exclude the case f = b here. See Chap. XVI-5.
438
XVI Physics of the W and Z bosons
Thus, the Z° resonance amplitude evaluated at energy Mz will yield information on the combination of factors,
V^(JS - 24JW) • y/pj(J£ - 2s2w4Q) .
(1.23)
In the approximation of studying just universal radiative corrections, the quantity pf is independent of fermion type and still contains the O(G^m^) dependence, ( p / W ^ p = 1 + (Ap)t + ... •
(1.24)
However, no analog of the correction factor K (C/. Eq. (1.5)) appears in Eqs. (1.22), (1.23) because it has been absorbed into the definition of s^. Since the relative amount of weak isospin (Tw^) and electric charge (Qei) occurring in the weak neutral current is measureable, it must follow from renormalization scheme independence that Using Eqs. (1.15), (1.16) to solve for K;, and keeping only terms of first order in an expansion in powers of (Ap)t, we obtain M2 4 ~ sl + c w(Ap)t ~ l - - r ^ , (1.26) pM\ to be contrasted with the on-shell version in Eq. (1.12). The introducion of p, itself containing O(G^m2) dependence, leads to a reduced dependence on the top-quark mass in s^. Precision LEP data plays an important role in the determination of s^ [Al 92], 2
_2 Sw
f 0.2335 ±0.0012 \ 0.2333 ± 0.0008
{LEP) (world ave.) .
[
'
}
A tantalizing application of the running weak mixing angle involves relating physics at the Z° scale with that of a 'grand unified' theory defined at an energy EQUT- The latter scale signals the existence of a gauge group undergoing spontaneous symmetry breaking to SU(3)C x SU(2)L x C/(l)y. The condition 91=92 = 93
(E = EGVT)
(1.28)
leads to a prediction [GeQW 74] for the weak mixing angle at the scale EGXJT- In the grand unified theory of SU(5) [GeG 74, La 81] and its supersymmetric extension (SUSY-SU(5)), the MS weak mixing angle obeys (1.29)
XVI-2 Phenomenology of the W^ and 2°gauge bosons
439
At the much lower energy scale /x = Mz, this becomes reduced by a calculable amount,* (1.30)
where a = a(Mz), M j is the mass scale of the superheavy gauge bosons, and C is a constant which depends upon the number n# of Higgs doublets, (SU(S)) (1.31)
(SUSY-SU(5)) . The 5/7(5) extension of the Standard Model has n# = 1, whereas the minimal supersymmetric model takes n# = 2. The 'bare bones' SU(5) model turns out to be unacceptable. It is well known to give rise to an unacceptably short proton lifetime, and recent precision data indicates that the three coupling constants of the Standard Model disagree with a single unification point if evolved according to 577(5) [AmBF 91]. Interestingly, the SUSY extension appears to succeed in both respects. The rate at which s^ 'runs' is decreased due to contributions of supersymmetric partners ('sparticles') of the known particles, and the unification scale is raised to a level (Mx — 1016 GeV) consistent with the observed proton stability. The unification condition of Eq. (1.28) is also satisfied. Studies are underway to see whether a careful analysis of supersymmetry-breaking yields insights regarding masses of the long sought SUSY 'sparticles'. XVI—2 Phenomenology of the W ± and Z°gauge bosons Experiments at SLC and LEP have provided accurate determinations of the Z°mass and decay modes. Facilities like Tevatron and LEP2 are expected to do the same for the W± bosons. In the following, we shall consider a few of the many aspects of W± and Z°physics [A1KV 89]. The emphasis will continue to be phenomenological, with effects of strong and electroweak radiative corrections included but not explicitly calculated. * Actually, Eqs. (1.30), (1.31) represent a simplification in that (i) lowest-order estimates for the renormalization group coefficients are employed, (ii) supersymmetry-breaking effects are ignored, and (iii) the fact that rat > Mz is also ignored.
440
XVI Physics of the W and Z bosons
Decays of W± into fermions The decay of a W-boson into a lepton pair is governed by the lagrangian of Eq. (II-3.31),* ^g
^ l + ^ ) i
+ h.c. .
(2.1)
It is a straightforward exercise to compute the tree-level decay width,
where x = mj/M^ and we have employed Eq. (II-3.43) in obtaining the x —> 0 form. There are also decays W —» q^qd) into quark modes (the superscripts i,j = 1,2,3 are generation labels), induced by
4 f = - ^ KV* tfV(l + 75)^ + h.c. ,
(2.3)
where Vy is a K M matrix element, and the index k labels color. The lowest order decay width for quark emission is
color
T
1
where x,x are mass ratios defined as above, and we assume that all emitted quarks eventually convert to hadrons. Since the t-quark is too massive to be a product of VF-decay, a sum over accessible quark flavors yields J2i,j l^j| 2 = 2. QCD radiative corrections modify Eq. (2.4) by a multiplicative factor <5QCD5 (2.5) = 1 + 0.038 + 0.002 + ...~
1.04 .
The numerical value is estimated from as(M%) = 0.118 ± 0.008 [Ca 92] and rif = 5. There are also smaller electroweak corrections which we shall not discuss. * Although we shall denote tree-level decay widths, cross sections, etc. with a zero superscript in this section, for the sake of notational simplicity, we shall suppress the zero subscript for bare parameters.
XVI-2 Phenomenology of the W^ and 2° gauge bosons
441
If all final state masses are ignored, the predicted total width for W± decay into fermions is r (tot)
_
r (had)
r (lept)
_ -.
^M—W
nA
,
n
*
3
—
(2.6)
(^f) -
An average of current UA1,UA2,CDF data [El 92] yields the value ^ 2.13 ± 0.11 GeV, which is consistent with the prediction of Eq. (2.6). In the limit of masslessfinalstate particles, the branching ratio for decay into a lepton pair Ivi is (Br)^ ~ 1/9 (£ = e, /i, r), while inclusive decay to a mode containing a positively charged quark q (q=u, c) gives (Br)9 ~ 1/3.
,-29
10 10'
1
10
10
10
10
Ecm(GeV)
Fig. XVI-2 Resonances in ee collisions.
442
XVI Physics of the W and Z bosons
Decays of Z° into fermions The collection of resonances observed in ee collisions as a function of the total center-of-mass energy is displayed in Fig. XVI-2. It is the weak interaction which dominates physics for s ~ M | , with the strong and electromagnetic effects merely supplying modest corrections. To lowest order, the decay of a Z° boson into a fermion pair / / can be conveniently expressed as (2-7)
where / = u,d, ve, e, Upon defining y = m^/M^ for fermion mass m, we obtain for the lowest order transition rate to a pair / / , p(0) JL
rrC\
_
NCGUM%
e T —
6TT
KJV
y/2
+
(2.8) For decay into quarks, these lowest order partial decay widths are modified by the QCD factor of Eq. (2.4) [CeG 79, ChKT 79, DiS 79]. There exist also electroweak radiative effects, which we can take into account by employing the modified neutral weak coupling constants #v and §£' of Eq. (1.22). Upon including both strong and electroweak corrections, the tree-level relation of Eq. (2.8) is replaced in the limit of massless final state fermions by C
6TTA/2
Various Z° decay widths measured at LEP are listed in Table XVI-2. Overall, the agreement with Standard Model predictions is seen to be impressive.
Asymmetry measurements For the reaction e~e + —• / ~ / + , the forward-backward asymmetry, AFB 5 refers to the relative difference between the / " traveling forward or backward relative to the incident e~ direction, ^FB = ^ F ~ ^ B ,
(2.10)
where dcosO ——
dto
,
7_i
dcosO ——
dn
.
(2.11)
XVI-2
Phenomenology of the W^ and 2° gauge bosons
443
We shall consider the restricted set of final state leptons / = fi,r, and shall take the incident electron beam to be unpolarized. If the final state fermion mass is neglected, then the tree-level differential cross section at center-of-mass energy EQM = s can be written in terms of direct channel photon and Z° exchange as [KiiZ 89]
+(9ie)2 + 9ie)2)(9if)2 + Si/)2)IX|2](1 + cos2 0) + Sg^ g^ g^ g^ \X\2}
where x(s) is proportional to the Z° propagator, i
^2
For energies s ~ M | , formulae involving x(s) become simplified since Re x(^f|) — 0- Upon performing the integrals in Eq. (2.12) and including the effects of electroweak radiative corrections, one obtains*
3-4
~
, Ofi
"
IQI 1
^
(/)
""•
y/iMz
2.14)
M
where
The dependence on the effective weak mixing angle s^ will occur mainly in the on-resonance (s = M|) asymmetries rather than in the slopes. Note, however, that all on-resonance asymmetries are suppressed by the presence of the small quantity g^ . Measurement of A^ for a final state lepton £ provides a distinct probe of the couplings gv , 3a [B6H 89]. A combination of decay width and forward-backward asymmetry data yields the values given in Table XVI2. Observe, however, that the relative phase of g\,L is not fixed by such measurements. In addition to the leptonic forward-backward asymmetry Apg, one can probe the analogous quark asymmetry Apg. Here, the * For final state quarks, QCD radiative corrections must also be included [DjKZ 90].
444
XVI Physics of the W and Z bosons Table XVI-2. Z°data from LEP [Ca 92].a Decay width Tee 1
MM
rhad Fhad/riept p. 1
inv
rtot
gie)2 gie)2 a
Experiment
Standard Model prediction6
83.0 ±0.5 83.8 ± 0.8 83.3 ±1.0 1740 ± 9 20.92 ±0.11 496.2 ± 8.8 2487 ± 10 0.0012 ± 0.0003 0.2492 ± 0.0012
83.52 - • 83.78 83.52 -> 83.78 83.52 -> 83.78 1731 -> 1742 20.71 -> 20.84 496 -> 497 2484 -+ 2496 0.0011 -> 0.0013 0.2513 -> 0.2518
Decay widths are expressed in units of MeV. For definiteness, the central values mt = 150 GeV and as{Mz) = 0.118 are assumed, and the band of predicted values corresponds to the range of Higgs masses, 50 < M#(GeV) < 1000.
b
preceding formulae describe physics at the parton level, but ultimately it is the charge asymmetry of hadrons which is measured. More can be learned by working with polarized fermions. For example, with s ~ M | the lowest order differential cross section for producing tau leptons of longitudinal polarization PT is dQ
a2\X\2 4s
e)2
+ gi )(giT)2 +
W
cos9 + g^2) cos*)]
(2.16) where again we have ignored (9(F 2 /M|) terms. Working at the Z° peak and integrating over all production angles, we can isolate the observable (T) PO1
^
-
^
a P r = + i +
_^ S^M|
V
^
The LEP determination, A^x
= -0.134 ± 0.035 ,
(2.17b)
reveals that ^v and ^a have the same phase, in accord with the Standard Model. Yet another experiment involves the forward-backward polarization asymmetry which combines the forward-backward and polarization measurements defined above and, if carried out over the angular
XVI-2 Phenomenology of the W^ and 2° gauge bosons
445
range | cos#| < c, gives
Finally, a polarized incident beam of electrons can provide the means to measure the left-right asymmetry, ^
^
,
(2.19)
where ox (<JR) denotes the cross section for an incident left-handed (righthanded) polarized electron, as well as to allow an extended test of the forward-backward asymmetry ^4FB- Neglecting the (Tz/Mz)2) contribution, we have for the left-right and forward-backward asymmetries evaluated at the Z° peak,
A L R ^f e ,
AFB c Ifi [e^ep
,
(2.20)
where Pe is the electron polarization. The number of light neutrino generations Data from Z°-decay has been used to obtain an accurate determination of the number of 'light' neutrinos [Tr 89]. This can be accomplished in a variety of ways. The most straightforward consists of simply comparing the total measured Z° width rt o t with the Standard Model prediction r(SM) 1
tot ->
r& M) = 3 ^ + ^
+1^
,
(2.21)
where Fiept is the width for decay into charged leptons. If we attribute any disagreement between F^f and Ftot to the presence of additional neutrinos light enough to appear in Z° decay and which couple to Z° identically to the known neutrinos, then we have
r£tM) - r[ot = (3 + A A g r ^ ) + rlept + r had = ( ) (2.22) A more incisive battery of tests depends on singling out Fi nv, which is the part of the decay width associated with 'invisible' (i.e. undetectable) particles, which is ascribed to neutrino emission, Tinv = r to t — Thad — Tiept = NUTVV . (2.23) If indeed only neutrinos contribute to Finv and each neutrino contributes alike to the Z° decay width, then the number of light neutrinos is
iv, = r i n v / r w .
(2.24)
446
XVI Physics of the W and Z bosons
For the sake of reference, let us refer to the Standard Model tree-level value for Z°-decay into a neutrino pair, 12TT
v-
-0.1658 GeV .
(2.25)
For the conventional value Nv = 3, this implies Finv ~ 0.4974 GeV. There are different strategies for determining Finv depending on the particular mix of theory and experiment to be employed. Let us describe two. If, along with Mz and Ftot, the ratio TV = Fhad/r^ is measured in the vicinity of the Z° peak, then the assumption of lepton universality together with the theoretical input F ^ ^ allows one to write •p.
-p
-p(SM)/o I fDf\
(<2 Of\\
An alternative method which is somewhat less dependent upon theory is, in addition to the above 7?/, Mz and Ftot, to use the measured hadronic cross section at the Z° peak, crj^ . Beginning with the Breit-Wigner formula (2J + 1) 7T FinFout B W = and assigning S\ = S2 = 1/2 for the incident spin values, J = 1 for the Z° spin and k2 = Af|/4 for the squared momentum, we obtain a
Deak l^71" r e eFhad had "-" 772"—F2
'
/o
OQ\ y Z' M )
From lepton universality, it follows that / n-?eak
r inv —— tot r \-Mz\ L
(2.29)
•»- i n v — J-
Methods such as these have been employed at the LEP facility to obtain the important result [Ca 92] Nu = 2.99 ± 0.05 .
(2.30)
The VWW vertex A sufficiently high energy ee collider can be used to produce pairs. Since this can arise from a virtual Z° or photon intermediate state*, infomation can be obtained regarding the VWW exit vertex, with V = Z°,j. The ZWW interaction is significant because it arises from * At tree-level, there is also neutrino exchange.
XVI-2 Phenomenology of the W^ and 2°gauge bosons
447
the nonabelian gauge structure of the theory. The jWW vertex provides information on the electromagnetic structure of the W boson. Consider the transition V(q, cr) —> VF~(p_, A_)W+(p+, A+) describing creation of a W± pair by a neutral spin-one quantum V. From Lorentz covariance, it follows that [HaHPZ 87]
f = (p. - P+r [fag"? - hvjf]
(2.31)
w
where gyww is the coupling strength and the {fiv(q2)} are form factors. Anticipating Standard Model predictions, we take
(232)
l Discrete symmetries constrain the form factors as /4V = /5V = 0 /5V = ftsv = frv = 0 / 4V = fay = fiy = 0
(C—invariance) , (P-invariance) , (CP—invariance) .
(2.33)
The decomposition in Eq. (2.31) allows for the existence of a magnetic moment /iw and an electric quadrupole moment Qw» Accordingly, it is convenient to express the form factors {fiv} (i = 1,2,3) in terms of quantities Av and KV, 2
/iv = 1 + ^ 2 " A v , /2V = Av , /2V = 1 + ^v + Av •
(2.34)
which are directly related to /iw and Qw> 6
(l +
A
6
+ )
Q
(K7 -
A
l) •
( 2-35)
To lowest order in the Standard Model (SM), one has 4 M = 1,
A^M = 0 ,
(V = 7,-Z°) ,
(2-36)
as well as fffi = 0 f o r i = 4,5,6,7. The corresponding tree-level forms for the W-boson magnetic moment and electric quadrupole moment are .
(2.37)
448
XVI Physics of the W and Z bosons
In principle, it is possible that the VF-boson has static properties which violate at least some of the discrete symmetries. An electric dipole moment dw or magnetic quadrupole moment Qw would be parameterized as 6
~
6
Searches for a neutron electric dipole moment can be used to place the bound \dw\ < 10~19 e-cm on the W electric dipole moment [MaQ 86]. XVI—3 The quantum electroweak lagrangian In the following three sections, we shall give a simple description of how electroweak radiative corrections are calculated. We begin in this section by quantizing the classical electroweak lagrangian to obtain certain of its Feynman rules. There is also an expansion of earlier comments made in Sect. XVI-1 regarding on-shell renormalization. Classical electroweak theory of three fermion generations is defined by an SU(2)L X U(l)y gauge-invariant lagrangian,* ) ,
(3.1)
where $ is the Higgs doublet and the collection {/$} of Higgs-fermion coupling constants is flavor-nondiagonal. With spontaneous symmetry breaking, all particles but neutrinos and the photon become massive and diagonalization of the neutral gauge boson mass matrix occurs in the basis of the photon A^ and massive gauge boson Z® fields as given at tree-level by Eq. (II-3.30). In addition, diagonalization of the fermion mass matrix for the three-generation system involves three mixing angles {0i} and one phase angle 6. The physical degrees of freedom of the gauge and Higgs sectors become manifest in unitary gauge (cf. Sect. XV-1),
4 $ = 4w} ({V- 0 } , Wj,Z%Ap,Ho ;{mf},Mw,Mz,MH,e)
, (3.2)
where the {9i} , 6 parameters are included in the {rrtf}. Gauge-fixing and ghost fields in the electroweak sector The quantum electroweak lagrangian £^ will contain, in addition to the classical lagrangian of Eq. (3.1), both gauge-fixing and ghost-field contributions,
* We have replaced the Higgs parameter /J,2 by the equivalent quantity v2.
XVI-3 The quantum electroweak lagrangian
449
Observe that mixing between gauge fields and unphysical Higgs fields occurs in the covariant derivative of the Higgs doublet (c/. Eq. (11-3.18)),^ 2
(3.4) & + h.c.+ ... . One can arrange the gauge-fixing term to cancel such mixing contributions. Expressing the complex Higgs doublet in terms of the physical field Ho, unphysical fields x+> Xs a n d the vacuum expectation value v as (3.5) we write the gauge-fixing contribution in the form, 2
2
1
(3.6)
- — [d^ + It is not hard to see that cancelation of the unwanted Higgs-gauge mixing terms occurs for arbitrary values of the gauge-fixing parameters £+,3,0Even with this cancelation, there remain in £ew~ quadratic terms containing the unphysical Higgs fields, and such terms will contribute to the propagators of these fields. As explained in App. A-6, once the gauge-fixing is specified as in Eq. (3.6), the structure of the Faddeev-Popov lagrangian d%' of ghost fields is then determined. For the electroweak sector, it turns out that there are four ghost fields, 4 ^ = 4SWW,CB) .
(3.7)
These are associated with the four gauge fields W^, B^ which appear in the original SU(2)i x U(l)y symmetric lagrangian. A subset of electroweak Feynman rules The full set of electroweak Feynman rules is rather lengthy and we refer the reader to the detailed discussions in [B6HS 86, AoHKKM 82] or to the summary in [Ho 90]. A few of the more useful rules, expressed in terms of bare parameters are* t Mixing also occurs, of course, between the neutral gaugefieldsB^, W^. * For notational simplicity, we suppress the zero subscript in the following discussion.
450
XVI Physics of the W and Z bosons
fermion W-boson vertex: 9
ftj-*
2—•
a,\
(3.8a) fermion Z-boson vertex:
(3.8b) W-boson propagator iD^ '(q):
g2 - M ^ + ie [
y
^
q Z + ^
l
(3.8c)
Z-boson propagator iD\J{q):
(3.8d) unphysical charged Higgs propagator iA^
ql - £+M^ + ie (3.8e) In the above, (Vy) is a matrix element for quark-mixing, fl^i) are given in Eq. (II—3.41), and £z is defined by expressing the gauge-fixing in the form of Eq. (3.6) but using the physical neutral fields. As seen in Eqs. (3.8c), (3.8e), each boson propagator is explicitly gauge dependent, and in particular, the propagator of the unphysical x+ van ~ ishes in the £+ —> oc limit of the unitary gauge. This is as expected, because only physical degrees of freedom appear in unitary gauge. In fact, the absence of unphysical degrees of freedom in unitary gauge would appear to be an appealing reason for carrying out the computation of radiative corrections in this gauge. However, there is a 'hidden cost'. In unitary gauge, the W± propagator of Eq. (3.8c) becomes = iunitary
9* " ^
+
XVI-3
The quantum electroweak lagrangian
451
and the high energy behavior produced by the q^q^/M^r term makes this a questionable choice for doing higher order calculations. Instead, as the price for acceptable high energy behaviour, many opt to accept the presence of unphysical fields. One popular choice of gauge-fixing is the 't Hooft-Feynman gauge, defined by setting all the gauge-fixing parameters equal to unity, ^ = 1. In this gauge, the lowest order propagators for the physical gauge bosons and unphysical Higgs and ghost fields have poles at either M ^ or Mz. This condition can be maintained in higher orders by a suitable renormalization of the gauge-fixing parameters. On-shell determination of electroweak parameters Two sets of electroweak parameters appear in the classical lagrangians of Eqs. (1.1), (1.2), • , . . f {fi},91,92,\,v 2 (Eq. (3.1)), Classical parameter sets =< , , (Eq. (3.2)) . \{mf},Mw,Mz,MH,e Considered as bare (input) parameters to the quantum theory, these obey the simple tree-level relations ni
(3.10) At this stage, there are several possible (equivalent) expressions for the bare weak mixing angle, e.g., I
z,o
4,o = -Jfj-
( 3 - llb )
•
Radiative corrections will generally modify tree-level relations, and as a result, necessitate a precise definition of the weak mixing angle. Following the analysis in [Si 80], let us compare the parameter subsets (g\5o, 52,0 ? ^0) and (eo, M^o? ^z,o)- Each of these bare quantities will experience a shift, i,o = 9i-
891 ,
92,0 = 92- $92 ,
^
vl = v2 - 6v2 0
(3.12) In on-shell renormalization, the theory is specified in terms of e, M\y and Mz. Moreover, the following relations are arranged to hold order by
XVI Physics of the W and Z bosons
452
order, &)
These equations constrain the effects of radiative corrections upon the parameters. By differentiating the three relations in Eq. (3.13), one finds after a modest amount of algebra the conditions, (3.14) Also, in on-shell renormalization one defines the weak mixing angle in terms of the masses M\y,Mz as in Eq. (1.12). Since this relation is to be maintained to all orders, the bare value s^0 of Eq. (3.11a) will be modified by shifts in the W and Z masses, 5
w,0 — * ""
+
(3.15)
Ml)\
\
Finally, there is a technical point worth noting. For any renormalizable field theory, it makes sense to express results in terms of the most accurately measured quantities available. Thus, it is preferable in the electroweak sector to replace Mw by G^ and work with a slightly modified parameter set, a~l = 137.0359895(61) , Physical parameter set = { G^ = 1.16637(2) x 10~5 GeV~2 Mz = 91.175(21) GeV .
(3.16)
To accomplish this, the relationship G^ — G^(a :Mw,Mz,...) can be used to replace Mw by G^. At tree level, the replacement is simply 1+1-
r\ V2" 2v27TQ;o \
(3.17)
We shall discuss the effect of radiative corrections upon this relation in Sect. XVI-5. XVI-4 Self-energies of the massive gauge bosons It is is evident from Eq. (3.14) that the parameter shifts <5e2, SM^r and <5M§ play an important role in the study of electroweak radiative correc-
XVI~4 Self-energies of the massive gauge bosons
453
tions. We have already determined from our analysis of QED (cf Eq. (II1.33)) that
where the photon vacuum polarization H(q2) appears in Eq. (II—1.26). In this section, we shall compute the portion of SMyy and <5M| arising from the fermionic vacuum polarization contributions to the W± and Z° propagators. As a consequence, we shall be able to identify the origin of the O{G^m2) propagator contributions. The charged gauge bosons W± The radiative correction experienced by a W± gauge boson propagating at momentum q is expressed in terms of a self-energy function, ^ 2 •
(4-2)
Although a vector-boson propagator iD^v{q) generally contains terms proportional to g^v and to q^qv, it will suffice to study just the g^v part. As indicated at the end of Sect. II—3, the q^qv dependence is absent if the gauge boson couples to a conserved current or will give rise to suppressed contributions if the external particles have small mass. Thus, we have for the W-propagator in 't Hooft-Feynman gauge, q2
q2 -
(4.3) where we have substituted for the bare W mass using Eq. (3.12). Let us now calculate the loop contribution of a fermion-antifermion pair /1/2 to the self-energy Aww(q2). We begin with hh
(4.4)
where Vf1f2 = ^c|^/i/ 2 1 2 f°r the case when the fermions are quarks. Aside from the occurrence of the 1 + 75 chiral factor and the nondegeneracy in fermion masses rai,ra2, the above Feynman integral is identical to the photon vacuum polarization function of Eq. (II—1.20). It is thus
454
XVI Physics of the W and Z bosons
straightforward to evaluate this quantity in dimensional regularization, and we find for the ga@ part,
3 f, 2
2^ T1
7
,
n-1
f1 , , , 2 ,
(4.5) where M 2 = mlx + m^l — x). Since the VF^ boson is an unstable particle with decay rate T\y, the function Aww(q2) must be complex-valued. Let us consider its real and imaginary parts separately. It is clear from Eq. (4.5) that Re Aww(q2) is divergent. One can construct a finite quantity Aww(q2) by defining the field renormalization, W^o = (^2^)1/^2W/i, and constraining 6M^ and 6Z^ to cancel the ultraviolet divergence in Re A ww(g2), + 6Zf{q2-M2v)
.
(4.6)
It follows from Eq. (4.5) that the mass shift 6M$y is fixed by ^ = Re A W ( M ^ ) ,
(4.7)
and the /1/2 contribution to the field renormalization which ensures that Aww(Myy) = 0 is easily found to be (4.8) To obtain a relation for the imaginary part of the self-energy, we simply recall that instability in a propagating state of mass M is described by the replacement M —> M — zF/2. This produces the following modification of a propagator denominator,
q2 -M2
(4.9)
q2-M2-
where we ignore the O(Y2) term. Comparison with Eq. (4.5) then immediately yields Im A W (A%) = MWTW
.
(4.10)
We can use Eq. (4.7) to check this relation by setting q2 = M^. If, for simplicity, we neglect the masses of the fermion-antifermion pair /1/2, then the imaginary part comes from the logarithm contained in the first
XVI-4 Self-energies of the massive gauge bosons
455
of the integrals in Eq. (4.7), Im / dx x(l-x)ln Jo and we obtain
= fi -ie
—> - - , q^=M^ 6
:
z
=
T
1 f-.fr, J1J
2
7= r*
^VF
(4.11)
5
/c%
(4.12) ^
'
DV27T
where we have substituted g\ = 4\/2M^G^. This result agrees with the result of our earlier decay width calculations of Eq. (2.1) for leptons and Eq. (2.4) for quarks. The neutral gauge bosons Z°, 7 The system of neutral gauge bosons is treated analogously to the charged case except that a 2 x 2 propagator matrix appears, and the issue of particle mixing arises. Although the neutral channel was already diagonalized at tree-level (cf Eq. (II-3.30)), interactions reintroduce nondiagonal propagator contributions at higher orders. The g^ part of the neutral channel inverse propagator D r L I/(^f2), diagonal at tree level,
has the renormalized form,
(4.14) Upon taking the inverse, we obtain for the individual neutral boson renormalized propagators, q2
+ j^tf) _ 4 z ( g 2 ) / ( g 2 _M2+
_
M
|
+
^27(g2)] _
^ 2 )
Observe that there is indeed a particle-mixing propagator, Dtf%, proportional to the the reduced self-energy Aiz(q2). It might appear from Eq. (4.15) that Z°-photon mixing gives rise to a photon mass contribution. However, one arranges as a renormalization condition that Alz(0) = 0, and the photon remains massless under electroweak radiative corrections.
456
XVI Physics of the W and Z bosons
If we consider only the vacuum polarization loop contribution due to a fermion of mass ra, we obtain for the Z° self-energy,
m2)
(4,6,
where m is the fermion mass, Nc is a quark color factor, and
( 4 - 17 )
The quantities in Na@ are just those expected from the coupling of fermion / to the neutral weak current. We then obtain from dimensional regularization,
- 3 / dxx(l- x) In
Jo
2iim{t f 2
z
, nr~ + lnV^--
2
V
\
J
1 f1 , , m2 - q2x(l dx In ?-+ l Jo V>
x)]] J -}\ . JJ (4.18) It is also easy to demonstrate that the photon-Z° self-energy A$P is proportional to Atf^ for the case of a fermion loop contribution, + 4m gi > \---jIe
7
^r
where Qf is the electric charge of the fermion.
XVI—5 Examples of electroweak radiative corrections All electroweak electroweak amplitudes will be affected by radiative corrections. In this section, we shall continue to emphasize the role of O{G^ml) contributions by computing their occurrence in the parameters Ap, Ar and in the Z°bb vertex.
XVI-5 Examples of electroweak radiative corrections
457
The O{GtlvrVf) contribution to AQ The quantity Ap introduced in Sect. XVI-1 can be defined as a correction to the rho-parameter of Eq. (1.3),
Dz(q2 = 0)
_J =
MZ + 6MZ
or
Observe that Ap is finite since the singular terms in Eqs. (4.5), (4.18) cancel. If we set mi = mt and m
,
[TO?,
ra?
9i
xm
t]
_ g\N g\ c 647T2 rv
Substitution of Nc = 3 and G^/y/2 = g^/SM^ yields the result shown in Eqs. (1.16), (1.17). This quadratic dependence on the heavy quark mass is in striking contrast with the behavior observed for the photon self-energy (cf. Eq. (II1.26)). In the heavy fermion limit, the photon vacuum polarization exhibits instead the decoupling result O(m^2). A technical factor which brings about this difference is that QED is a vector theory, whereas the charged and neutral weak interactions are chiral. Indeed, one can show (cf. Prob. XVI-3) that the decoupling expected of a vector interaction results when left-handed and right-handed self-energies are averaged. However, equally important is the fact that as mt grows while mt is kept fixed, the weak doublet is being split in mass. Thus, decoupling of the top quark in the large mt limit should not be expected because if we were to integrate out the top quark, we would no longer have a renormalizable theory - the remaining low energy theory would have an incomplete weak doublet. As noted above, if both members of the doublet are taken to be equally heavy (mt =ra&—• oo), there would be no quadratic dependence on the heavy quark mass, and the decoupling theorem (cf. Sect. IV-2) would be satisfied. It is the large splitting in the weak doublet which leads to the observable violation of decoupling.
458
XVI Physics of the W and Z bosons The OtG^rrit) contribution to Ar
The relation between the Fermi constant and the vector boson masses is modified by a quantity Ar, sometimes called the 'quantum correction' [Si 80, BuJ 89]. The tree-level relation of Eq. (II-3.43) becomes modified by the radiative corrections of Fig. XVI-3,
GR_jlo l It is to be understood in Eq. (5.4) that 'GM' is determined from the muon lifetime with the photonic corrections described in Sect. V-2 already taken into account. Thus Ar contains only the remaining electroweak effects. To trace the origin of the quantum correction, we observe first the effect of the W± self-energy on the bare relation in Eq. (5.4), A w (0) V2
1
J' K
Ml
8 92 -
]
where we have taken q2 ~ 0. Next, we replace the bare parameters ^ and M ^ o by their physical forms as in Eq. (3.12). Comparison with Eq. (5.4) directly yields
Upon using Eq. (3.14) to substitute for bg\, we can rewrite this as
4 Recalling that the W± and Z° mass shifts can be related to the selfenergy functions AWW(M^V) and AZZ(M|), it should be clear that Eq. (5.5)
(c)
(d)
Fig. XVI-3. (a)-(b) Vertex, (c) propagator, and (d) mass-shift counterterm corrections to muon decay
XVI-5 Examples of electroweak radiative corrections
459
expresses Ar entirely in terms of calculable quantities.* Although each of the terms in Eq. (5.7) is divergent, the overall combination is finite. A number of rearrangements and algebraic steps can be used to isolate the leading contributions, and one finds Ar = Aa + Arw + (Ar) rem ,
(5.8)
where Aa
= a(Mz)
a
OL
„ n(Mf) , and Arw = -^fAp
.
(5.9)
Sw
A/9 is defined in Eq. (4.17), and (Ar) rem contains smaller finite contributions. The largest contribution to Ar is Aa, the shift in the fine structure constant. Although we have previously expressed the variation in a(q2) in terms of fermion masses (c/. Eq. (II—1.38)), the difficulty in precisely determining quark masses would appear to undermine an accurate evaluation of Aa. However, one can use dispersion relations to relate the hadronic contribution to the vacuum polarization, nhad(
The imaginary part of Uhad(q2) is expressible in terms of cross section data evaluated at invariant energy g2,
Im UhM)
= $Btf) 6
with Rfo>) = ^ - h a d r o " S ) cr{ee —• fifj)
.
(5.n)
Thus, we obtain a dispersion relation for the subtracted quantity IIhad(
- n had (o) 3TT IJimii
Jsos J
s{s -q2-
ie)
where so denotes the point at which data becomes unavailable. For energies above so, a perturbative representation is used to approximate R(s). The result of Eq. (5.12), when added to the lepton contributions, implies a value for oTx{M\) [BuJPV 89, Pe 90]^ a" 1 (Ml) = 128.77 ±0.12 .
(5.13)
* There are additional radiative corrections, such as the 'box' diagrams, which we shall not discuss, t There are minor differences in various evaluations cited in the literature, depending on how the perturbative estimate is performed or on the particular renormalization scheme.
XVI Physics of the W and Z bosons
460
w -b
~~—
*—b
(a)
(b)
Fig. XVI-4 Top-quark corrections to the Z°bbvertex. Our analysis of Ar performed thus far is only to first order in electroweak corrections. What is the result if higher orders are incorporated? For the running fine structure constant, the incorporation of higher orders proceeds according to a renormalization group analysis [Ma 79] and amounts to summing a geometric series. Although the generalization of Eq. (5.8) to all orders has not been rigorously given, a plausible construction exists based on evaluation of the quantum correction to second order. The second order result is suggestive of employing two independent geometric series [CoHJ 89], 1
i+AQ - 4 4
1-Aa
i+
.
(5.14)
As a consequence, the relation between M ^ and the parameters G^, a, and Mz becomes 1 V2 '
r-
i =^
1+
1- 2\/27ra
1 + 1 -
I 1/2'
2\/2-
•KOi
-
(5.15)
Arr,
where p — 1/(1 — Ap), and for the sake of comparison, we display the tree-level result on the first line. The Z —> bb vertex correction The preceding analyses of Ap and Ar could very well be carried out for any other electroweak observable. In most cases, we would again find important O{G^m^) radiative corrections. Thus, for example, the Z° width for decay into lepton £ {£ — e, /i, r) has the form
Tzo^a = r SL^-1 1 + (A^)* + • • -1 >
(5-16)
and grows quadratically with increasing mt [AkBYR 86]. The origin of this effect, the one-loop ti contribution to the Z° propagator, is identical to that discussed earlier.
XVI-6 Beyond the Standard Model
461
Interestingly, however, a more complete calculation reveals a slight decrease to occur in the decay rate r z o _ ^ as rrit grows. This is because, although the decay amplitude contains a (universal) propagator contribution proportional to (Ap)t, an even larger effect, the vertex correction of Fig.XVI-4, contributes with opposite sign [AkBYR 86, DjKZ 90],*
r z0 _^ = I ^ _ ^ |"l + i | ((Ap)* + (Av^)t^j + .. .1 ,
(5.17)
where the Z°bb vertex correction is given by ,* (h\,
20 / A
x
130 a ,
m?
(5.18)
The dJ, ss modes also contain virtual t-quark vertex corrections, but they are greatly suppressed by the tiny accompanying KM factors |Vti|2 (z = d, s). Recalling the characterization given in Sect. XVI-1 of radiative corrections as either 'universal' or 'nonuniversaP, one may interpet the Z°bb effect as a nonuniversal term which contributes as V^P/nonuniv
(5.19) Although O(rrif) corrections are the most important, O(In(TOJ/M|) logarithmic dependence has been included in Eq. (5.19) because it has a non-negligible numerical impact. XVI-6 Beyond the Standard Model In this book, we have focussed on the consequences of the Standard Model. Despite intense experimental scrutiny, this theory has displayed no experimental inconsistencies to date. However, the Standard Model does have several features which many physicists consider unsatisfactory. On the one hand, there exists no understanding of the number of families, the origin of quantum number assignments, or the large number of arbitrary parameters. On the other, there are a set of problems with the 'naturalness' of the theory. We have already detailed the difficulties of obtaining small enough values of the cosmological constant, the strong CP-violating parameter 0, and the quadratic radiative corrections to the mass of the Higgs boson. It is necessary to look beyond the Standard Model if we are to achieve a resolution of these issues. In order to make progress, an experimental signal of new physics will be needed. There are a number of on-going programmatic attempts to * Due to cancelations, the vertex correction turns out not to affect asymmetry phenomena, such as the 6-quark forward-backward asymmetry Ap^.
462
XVI Physics of the W and Z bosons
probe the limits of the Standard Model. Listing the main areas from low energy to high energy, we have 1) 2) 3) 4)
searches for neutrino mass and oscillation, rare decays of muons and kaons, CP violation, especially in B mesons, precision measurements of both W±,Z° properties and low energy neutral current phenomena, and 5) searches for new particles at high energies, particularly the Higgs boson or other indicators of the symmetry-breaking sector.
Of course, while new discoveries might arise from areas outside this list, the above items are particularly powerful probes. Let us consider each briefly. (1) Neutrino mass and oscillation: Strictly speaking, neutrino mass and oscillation phenomena can be accommodated within the Standard Model [KaGP 90], [BoV 87], [MoP 91]. One simply adds a right-handed neutrino field having very small Yukawa couplings. However, the need to postulate neutrino masses small enough to satisfy present bounds makes this an implausible strategy. Actually, there are models of new physics which naturally predict small neutrino masses. One such scheme involves the so-called 'seesaw mechanism' - if a right-handed neutrino is added to the theory, SU(2)L X U(1)Y allows there to be a (presumably large) Majorana mass* TUM as well as a (presumably normal sized) Dirac mass In the (yj,, VR) basis, the mass matrix is )
,
(6.1)
where the crucial zero in the diagonal matrix element arises because VL sits in an SU(2)i doublet and hence is prevented by local SU(2)L invariance from forming a direct Majorana mass. This mass matrix has a light eigenvalue mv — m2D/mM<, which would imply mv ~ 10~12 eV for the choices rap — rne and TUM — 1014 GeV. (2) Rare decays: Many models of new physics at high energy violate the flavor quantum numbers which are otherwise conserved in the Standard Model. These violations are thus anticipated in rare processes, and the decays of kaons and muons are generally the most sensitive probes. The evidence of nonzero rates for forbidden processes at low energy would provide evidence for new physics at the TeV scale. We have described techniques relevant to this topic in Sect. IV-9 and in Sect. VII-5. * Recall (cf. Sect.V-4) that a Majorana mass term arises from the Lorentz invariant, but lepton number violating, coupling of a neutral fermion to its conjugate.
XVI-6 Beyond the Standard Model
463
(3) CP violation: CP nonconservation enters the Standard Model through the Higgs sector by means of complex Yukawa couplings. However, this portion of the theory is in many ways the weakest, and alternate mechanisms of CP violation are being actively explored. If the Higgs sector is changed, the observed CP violation must arise in some other fashion [Le 74, Wo 86, DoHV 87, PaT 89]. The predictions of alternative models may be checked in rare kaon decays, hyperon decays, and most significantly in B decays. In certain channels, notably Bd —> J/ipKs, the theoretical prediction is firm enough (c/. Sect. XIV-5) that a demonstrated lack of CP violation could be used to falsify the minimal Standard Model. (4) Precision measurements ofW±,Z° properties: In the study of the W ,Z® gauge bosons, we have the fortunate confluence of reliable theoretical methods and precise data. The Standard Model unambiguously predicts the properties of these particles. It is possible that experiments may soon find deviations from the Standard Model predictions. Should such deviations be found, it may even prove possible to identify their origin by considering the pattern of deviation. This point is sufficiently important to warrant further comment. Physics associated with a large energy scale A should affect the gauge boson self-energies — iU1 (q) (i = 77,72, WW,ZZ). For A2 ^> g2, one would expect rapid convergence of an expansion for — iUl (q) in powers of #2/A2, yielding the following effective low energy description, ±
-iYl^(q)=g^Ai
+ q2Afi) + ...
.
(6.2)
This description involves eight free parameters, A 7 7 , . . . , Afzz. However, the conditions n 77 (0) = n 7Z (0) = 0 reduce this number to six. An additional three parameters can be absorbed into the renormalization of , which experience the shifts [BaFGH 90], 6a 6_a_ a ~
, ^ '
£G> , GM " ww '
M2Z "
M\
The three remaining parameters may be chosen to be the same quantities, A/9, AK, and Ar w , that we have discussed in previous sections regarding the Standard Model. Actually, it is convenient to first replace AK by a quantity AK/ defined by gv/g& = 1 - 4(1 + A«>j) ,
(6.4)
where
l4
VM2z
,
(6.5)
464
XVI Physics of the W and Z bosons
and then to introduce the set [A1BJ 92]* d = Ap , s2
e2 = c20Ap + - 2 - ^ - j A r w - 2S20AK' , c
(6.6)
0 "" S0
These quantities are related to the self-energy parameters of Eq. (6.2) by* 61 =
j^
j^
=
~MJ ~ M ^
^2 ~ -^ww
^33
6
^33 *
3 =
A'vx A
Ml ' \^*' /
J
In the limit of large mt and M#, only ei experiences power law radiative corrections, and one has for the Standard Model
3G>
SG.Mlsj
(MH\,
4 \MZ)
G,M2W (MH\ 3
12^^
\MZ)
...,
(6.8)
G,M2W
(mt\
2
\MZ)
6TT V2
One can obtain a determination of the three parameters ei52,3 from
a minimal data set consisting of Mw/Mz,
^z^th
an(
^ ^FB ky using
Table XVI-3. Determination of the {c»} [A1BJ 92]. Parameter €l €2 €3
Minimal data set ( - 0 ..05 ± 0. 51) x 10 - 2 ( " 1 ..07 ± 1.00) x 10 - 2 (+0..07 ± 0. 86) x 10 - 2
Full data set ( - 0 .02 ± 0 .37) (-o .71 ± 0 .89) ( - 0 .31 ± 0,.62)
X X X
10 - 2 10 - 2 10 - 2
An equivalent set appearing in the literature is [PeT 90]
The subscript '3', as in ^33 or A3^ refers to the field W3 = c^Z + s w A (c/. Eq. (II-3.30)).
XVI-6 Beyond the Standard Model
465
Eqs. (1.18), (2.9), (2.14)) as definitions. Alternatively, by folding in additional theoretical structure one can enlarge the description to encompass the full available data set. Results from using both approaches appear in Table XVI-3 [A1BJ 92]. The error bars are seen to be relatively large. Roughly speaking, the analysis of [A1BJ 92] reveals no discrepancy between the Standard Model and existing experimental data. Future work should reduce the experimental uncertainties, and discovery of the t-quark would eliminate the largest theoretical uncertainty associated with Standard Model predictions. Thus, the Standard Model could be found to agree with experiment to new levels of precision, or perhaps signals arising from new physics could emerge. For example, one goal of precision studies is to check whether the Higgs mechanism is really the origin of weak symmetry breaking. In practice, effects of the Higgs boson are shielded in many electroweak processes, and are thus hard to measure. However, alternative symmetry-breaking schemes, such as Technicolor [FaS 81], may not be so well hidden. All such schemes involve Goldstone bosons which become the longitudinal components of the W± and Z° bosons, but the remaining physical degrees of freedom need not be the same. If the other particles are light (which in this context means comparable in mass to the W±,Z°), they will presumably soon be discovered directly. If they are heavier, they may still be seen indirectly through their effect on the W±, Z° properties. As we have emphasized throughout this book, the effects of heavy particles can be analyzed theoretically by using effective lagrangians. These must respect the SU(2)L X U(1)Y symmetry, but may or may not include the extra custodial SU{2)L X SU(2)R invariance of the Higgs sector with doublet Higgs fields. The symmetries can be implemented by using an SU{2) matrix U for the Goldstone fields (as described in Chap. IV) with transformation U —• LUR) and covariant derivative DpU = dfj,U + ig<2W\x' f U + igiUr^B^ .
(6-9)
±
As an example of modelling new physics in the W ,Z° system, let us consider the effective lagrangian [PeT 90] 5132 s "*
BllVFi
^ (
v
V]\
+
I
167T
(6.10) where B^ ', F^ are respectively the field strength tensors defined in Eqs. (II—3.11), (II—3.12), and S, T are parameters introduced earlier. The first of the above effective interactions breaks the custodial symmetry (c/. Sect. XV-3), and the second exhibits an SU(2)L X SU(2)R symmetry. In fact, the second is just the aio operator of Eq. (VI-2.7) v
modified by aio —* —S/16TT,
L^V
—• g2F^v-T
and R^v —>• g\B^vT^.
Among
466
XVI Physics of the W and Z bosons
the effects of the lagrangian £new are the self-energy contributions 6Alz{q2) = —m 2 , SAZZ(O) = -Mhx , cw which in turn manifest themselves in the radiative corrections Ar w = AriSM) + e3 . Ap = Ap(SM) + 6i ,
(6.11)
(6.12)
Of course, different models of symmetry breaking predict different values for the Ci (or equivalently of S and T), and in some cases are close to being ruled out. For example, by scaling up the QCD value of aio, it has been estimated [PeT 90, HoT 90, MaR 90] that SU(4) technicolor with a single generation of technifermions has S = 2 (i.e. 63 = 1.6 x 10~2). Measurements of W±,Z° properties at LEP and SLC and comparison with low energy processes can hope to reveal such effects, if present in Nature. (5) High energy searches: While it may be possible to obtain indications for new physics from low energy probes, there is no substitute for directly producing the new degrees of freedom and studying their properties. This can be achieved at the high energy limit of existing accelerators, such as the Tevatron, or with future, very high energy machines such as the SSC or LHC now being planned. Within the Standard Model, the top quark and the Higgs boson must be found. New physics associated with alternate symmetry breaking schemes would likely enter at the TeV energy scale. The most pressing physics goal of the very high energy colliders is the discovery of the Higgs, or the conclusive demonstration that it is not present up to 1 TeV. We have outlined in Sect. XV-3 the fascinating possibility of a strongly interacting symmetry-breaking sector, which could in principle be explored by W^-Wj, scattering. Such a theory should have a potentially rich resonance spectrum, and our experience with QCD would be called upon to sort out the physics of this new strongly interacting sector. If the theory remains weakly coupled, there may also be new particles, perhaps those predicted by supersymmetry. The Standard Model is a remarkable theory, representing the culmination of modern scientific attempts to understand the laws of Nature. While appreciating the power of the Standard Model, it is nonetheless appropriate to look forward to the discovery of new physics from the next layer of reality.
Problems
467
Problems 1) The rho-parameter a) Show that for an arbitrary number of Higgs multiplets ((
b) Given two Higgs fields, respectively with quantum numbers 7W= —7 W3=l/2 and J w = l , 7W3=0, and having the nonvanishing vacuum expectation values (^1/2) and (<^i), obtain a bound for \(ipi)/(
Appendix A Functional integration
In this appendix we outline the basis of functional methods which are employed in the text. Path integral techniques appear at first sight to be rather formal and abstract. However, it is remarkable how easy it is to obtain practical information from them. Very often they add insight or new results which are difficult to obtain from canonical quantization. A.I Quantum mechanical formalism Before attempting to address the full field theoretic formalism we first review the application of such techniques within the more familiar setting of nonrelativistic quantum mechanics in one spatial dimension. Unless otherwise specified we hereafter set ft = 1. Path integral propagator Simply stated, the functional integral is an alternative way of evaluating the quantity D(xf,tf;xuti) = (xfle-^f-^lxi)
= (xf,tf\xuU) .
(1.1)
This matrix element, usually called the propagator, is the amplitude for a particle located at position X{ and time U to be found at position xj and subsequent time tf. The propagator can also be written as a functional integral
D(xf,tf;xi,U) = J-D[x(t)]eiSW» ,
(L2)
where the integration is over all histories (i.e. paths) of the system which begin at spacetime point Xi,t% and end at #/,£/. The paths are identified by specifying the coordinate x at each intermediate time t, so that the symbol V [x(t)] represents a sum over all such trajectories. The contribution of each path to the integral is weighted by the exponential involving 468
Functional integration
469
the classical action
S [x(t)} = J*f dt (jx2(t)
- V (x(t)))
(1.3)
which, since it depends on the detailed shape of #(£), is a functional of the trajectory.* Although the validity of the path integral representation, Eq. (1.2), may not be obvious, its correctness can be verified by beginning with Eq. (1.1) and breaking the time interval tf — U into N discrete steps of size e = (tf — U)/N. Using the completeness relation / JJ—oo — oc
dxn \xn)(xn\
,
one can write Eq. (1.1) as J-oo \ cly /v I O
I c*-/ /v
J-oo
1 /
\ «*-/ /V — 1 I v>
(1.4)
I **/ J\f
/ / • • • \ **/ 1 I ^
/
I *i ii /
where xo = Xi, XN = Xf. In the limit of large iV the time slices become infinitesimal, implying I
I -iHf\
{x£\e
%n€
\
\x£-i)
I
I
= {xe\e ^
^
^
)
+ O(e2) .
Inserting a complete set of momentum states and introducing a convergence factor e~Kp for the resulting integral over momentum, we have i-i) = Hm f
^
eHtxi-xt-i)-iep>I7m-Kp>
y 2 7 r
(1.6)
It is important to understand the difference between the concept of a function and that of a functional. A real-valued function involves the mapping from the space of real numbers onto themselves Reals <— [/ : Reals] . On the other hand, a real-valued functional such as 5 [#(£)] is a mapping from the space of functions x(t) onto real numbers Reals <— [S : x(t)] .
470
Appendix A
Upon taking the continuum limit we obtain
( rn \ N hm ijr—r)
N^oo \27Tl€/
n/
(1.7)
n=l J~c
It is clear then that we can make connection with Eq. (1.2) by identifying each path with the sequence of locations ( x i , . . . , #AT-I) at times e, 2e,..., (iV — l)e. Integration over these intermediate positions is what is meant by the symbol JT> [x(t)], viz. N-l
Each trajectory has an associated exponential factor expi5 [#(£)], where the quantity ^ * ' * ' - ^ e=i \
-V{xt)) /
(1.9)
becomes the classical action in the limit N —> oo. We have thus demonstrated the equivalence of the operator (Eq. (1.1)) and path integral (Eq. (1.2)) representations of the propagator.* It is important to realize that in the latter all quantities are classical - no operators are involved. The path integral propagator contains a great deal of information, and there are a variety of techniques for extracting it. For example, the spatial wavefunctions and energies are all present, as can be seen by inserting a complete set of energy eigenstates {| n)} into the definition of the propagator given in Eq. (1.1), x^e-^nVf-U)
^
1 0 )
n=0 For completeness, we note that by combining Eqs. (1.5)-(1.8), one can also write the propagator in a corresponding hamiltonian path integral representation f,tf;xi,ti)=
lim /
-^-dx1-^-
dp N-l
This form is useful when one is dealing with non-cartesian variables or with constrained systems.
Functional integration
471
In addition, other quantum mechanical amplitudes can be found by use of the identity t
, tf\T (x(ti)... x(tn)) \xi, U) =
where T ' is the time-ordered product.
External sources An important technique involves the addition of an external source. In the quantum mechanical case this is added like an arbitrary external 'force' j(t), (xf,tf\xuti)m
= I V[x(t)]e Jti
.
(1.12)
The amplitude is now a functional of the source j(t). Prom this quantity one can obtain all matrix elements using functional differentiation, which can be defined by means of the relation
= jdt'S(t-t')j(t') => ML=S{t-t)
(1.13)
and yields the result we seek, (xf,tf\T(x(t1)...x(tn))\xi,ti)
For many applications it is necessary Drily to consider matrix elements between the lowest energy states (vacuum) of the quantum system. This can be accomplished in either of two ways. First it is possible to explicitly One can prove this relation by choosing a particular ordering, say U
,
and noting that
fak^f^tflXnitn) Xn (xn, tn \xn- i, t n _ i ) Xn- i . . . XX (^ where we have used completeness and have defined x^ = x(tk) (k = 1,2,..., n) . The amplitudes (xfc^fcl^fc — i>*fc — 1) are simply free propagators as in Eq. (1.1), and can be evaluated by means of the time-slice methods outlined above. Thus the above expression is identical to the righthand side of Eq. (1.11). In the case of a different time ordering the same result goes through provided one always places the times such that the later time always appears to the left of an earlier counterpart. However, this is simply the definition of the time ordered product and hence the proof holds in general.
472
Appendix A
project out this amplitude using the ground state wavefunction (x,t\0)=th(x)e-iEot ,
(1.15)
which implies /»OO
(0\T(x(t1)...x(tn))\0)
=
J—oo
/»OO
dxf
J—oo
dxii>*0(xf)ei
(xf, tf\T (x(h)... x(tn)) \Xi,
(1.16) However, this amplitude can be isolated in a simpler fashion. If we consider the amplitude (x/,£/|aJi,ii) in the unphysical limit tf —> — zr/, U -^ +in we find for large r/ + T{,
(xf,tf\xuU)
Y, (1.17)
Generalizing, we have eiE0(tf-ti)
which is operationally a much simpler procedure than Eq. (1.16). The generating functional We may combine all these techniques in the so-called generating functional, defined by (x/,t/|ii)^(t) .
(1.19)
This has the path integral representation W[j] =
lim
tf^-ioo J
V[x(t)]e
Jt
i
.
(1-20)
t^—^ioo
Noting that for U = iri and tf = — ir/, E { + )
(1.21)
we find that ground state matrix elements as in Eq. (1.16) can be given in terms of the generating functional W[j],
(0|T(x(tl) ... xfa)) |0) - ( -
i r
^m
)
6
" gjM WVl\^ • (1-22)
Functional integration
473
It often happens with path integrals that formal procedures are best defined, as above, by using the imaginary time limits t —• ±ioo. However, in practice it is common instead to express the theory in terms of Minkowski spacetime. Thus, the generating functional will involve the real time limits t —> ±00. Does the dominance of the ground state contribution, as in Eq. (1.21), continue to hold? The answer is 'yes'. At an intuitive level, one understands this as a consequence of the rapid variation of the phase e-iEnt - nfaelimit t —> 00. The more rapid phase variation accompanying the increased energy En of any excited state washes out its contribution relative to that of the ground state. In a more formal sense, the real-time limit is defined by an analytic continuation from imaginary time. To properly define the continuation, one must introduce appropriate cie' factors into the Greens functions in order to deal with various singularities. Beginning with the next section, we shall often employ the Minkowski formulation and thus explicitly display the 'ie' terms in our formulae. The prescription given in Eq. (1.22) represents a powerful but formal procedure for the generation of matrix elements in the presence of an arbitrary potential V(x). Unfortunately, an explicit evaluation is no more generally accessible via this route than is an exact solution of the Schrodinger equation. In practice, aside from an occasional special case, the only path integrals which can be performed exactly are those in quadratic form. However, approximation procedures are generally available. One of the most common of these is perturbation theory. Suppose that the full potential V(x) is the sum of two parts V\(x) and V2(x), where V\(x) is such that the generating functional can be evaluated exactly while V2(x) is in some sense small. Then we can write W[j] =
hm
/ V[x(t)]e
tf-+-ioo J
= t / fe»
c
wy
"
"W
(1-23)
t4—• i o o
n\ where w(O)[-l_ L«y J —
i: 11111
t y —>• — ioo
PJti
f VW(tW I
X-^lay^c/IJ
o
™ *
[2
~
W
^v~w, .-VA/WJ
^ I
2 4
x
LZiil
J
t^ — • i o o
is the generating functional for V\(x) alone. Obviously Eq. (1.23) defines an expansion for W[j] in powers of the perturbing potential V2(x).
474
Appendix A
A.2 The harmonic oscillator It is useful to interrupt our formal development by considering the harmonic oscillator as an example of these methods. This treatment turns out to reproduce known oscillator properties with the use of functional methods, which are very similar to corresponding field theory techniques. It is most convenient to address the problem by employing Fourier transforms, /»OO J771
f- e~iEtx(E) ,
x(t) = / J-00
(2.1)
27r
whereby for U = — 00 and tf = +00,
Sj [x(t)} = £ ^ dt (jx\t)
(2
= L t {? * -
(2.2) with the definition x'{E) = x(E) + j(E)/ (mE2 — mu)2 + ie). An infintesimal imaginary part ie has been introduced to make the integration precise. Upon taking the inverse Fourier transform
x'(t)= I
— e~iEtx'(E) = x(t) + — I m
J—oo ^
dt'D(t-t')j(t')
J—oo
, (2.3)
where 1
D(t -1') = f" ^ e - ^ - ^ ) ^ we have /rn
/»oo
-—
:
,oc
/
2m J-00
= -J-e-Mf-t\
,
(2.4)
ze m
dt
,00 J-00
r1
\
dt'j(t)D(t-t')j(t') .
(2-5)
Finally, changing variables from x(t) to x'(t) we obtain the generating functional r
/°1
/T-f)j(t')
Functional integration
475
Note that the above change of variables has left the measure invariant
J
f
We can use this result to calculate arbitrary oscillator matrix elements. Thus for £2 > ti, we have for the ground state
(0\T(x(t2)x(h))\0) ={-if
W[0] (
2
(2.7)
j=0
ti)
^
which, in the limit t £1, reproduces the familiar result
<0|^|0)
JL
(2.8)
Although only ground state expectation values have been treated thus far, it is also possible to deal with arbitrary oscillator matrix elements with this formalism by generalizing the operator relation (2.9) where
is the usual creation operator. First, however, it is convenient to use the classical relation p = mx to rewrite the operator a^ as
In a simple application, we calculate that
|x|l) = lim J^- (1 - I £•)
(0\x(t2)x(tl)\0)
u: at, (*-*•*)
mu ( = hm W—- 1 t 2 ^+ V 2 V
i
d \ i
— 1 —D(t u dtij m
2
-
2rnuj
which agrees with the result obtained by more conventional means,
(0|x|l) - J-t— (o\ (a + aA \l) = -7L= . x
' ' '
V 2muo \ ' V
/ ' /
J2mu
(2.13)
476
Appendix A
More complicated matrix elements can also be found, as with
W» = IT ,iLT (> + ~. IT) I1 V2H
^ at J V
w at2
64 i\2
,.
lim
A
i
8 \ /
1 + - -r—
1
i
5
T—-
x [/?(«! - t2)D(0) + 2D(ti - t)D(t - t2)} =
2muj
(2.14)
which agrees with
In this manner, arbitrary oscillator matrix elements can be reduced to ground state expectation values, which in turn can be determined from the generating functional W\j\. The ground state amplitude in the presence of an arbitrary source j(t) contains all the information about the harmonic oscillator. One should note the analogy of the above methods to those of quantum field theory. The cone-particle' matrix elements involving | 1) have been reduced to vacuum matrix elements by use of Eq. (2.9). This is similar to the LSZ reduction of fields. As a result, all that one needs to deal with are the vacuum Green's functions. The generating functional is ideal for this purpose, as we shall see in our development of functional techniques in field theory. A.3 Field theoretic formalism One of the advantages of the functional approach to quantum mechanics is that it can be taken over with little difficulty to quantum field theory. An important difference is that instead of trajectories x(i) which pick out a particular point in space at a given time, one must deal with fields
Functional integration
477
Path integrals with fields The formal transition from quantum mechanics to field theory can be accomplished by dividing spacetime, both time and space, into a set of tiny four-dimensional cubes of volume 6t6x6y6z. Within each cube one takes the field
(3.1)
as a constant. Derivatives are defined in terms of differences between fields in neighboring blocks, e.g., ,Vji zkM + St) - ip fa,yj, zk,ti))
•
(3.2)
The lagrangian density is easily found, £(?, dnV)\xiM,zhM
-
C
iff (x*> Vj>zk> *i)> Q\M (**> Vi* zk> U))
,
(3-3)
and the action is written as
S ~ ^2 Sx8y6z6t C (
(3.4)
The field theory analog of the path integral can then be constructed by summing over all possible field values in each cell /»O
II /
(3.5)
i,3,k,r-
Formally, in the limit in which the cell size is taken to zero this is written as
f[d
.
(3.6)
By analogy with the quantum mechanical case (cf. Eq. (1.18)), it is clear that, since the time integration for S in Eq. (3.4) is from — oo to +oo, this amplitude is to be identified with the vacuum-to-vacuum amplitude of the field theory, (0|0) = N f[dtp(x)] e i 5 ^ ) ' ^ ( * ) l
.
(3.7)
Generally, quantum field theory is formulated in terms of vacuum expectation values of time ordered products of the fields G< n >(xi,..., xn) = (0\T (ipix,)...
(3.8)
i.e., the Green's functions of the theory. By analogy with the quantum mechanical case, one is naturally led to the path integral definition
GW(X1,...,
xn) = N J [d
478
Appendix A
where N is a normalization factor. Again we emphasize that all quantities here are c-numbers and no operators are involved. In terms of a functional representation, we then have from Eqs. (3.7), (3.9),
Generating functional with fields These Green's functions can most easily be evaluated by use of the generating functional
W[j] =N f [d
(3.12)
6cp(x) '
which lets us obtain (c/. Eq. (3.9)) (3.13)
W[0] 6j(
j=o
As an example of this formalism consider the free scalar field theory £(°)(x) = -dptpd^ip - —(f 2
.
(3.14)
In general we have 00
{o)
{o)
w [j]=w [o]
jTl \
n
POO
^ 2 U— n /
dxk
n=0 ' lk=l J-°°
rtxk>> (3.15)
where the generating functional W^ [j] is given by = N
[d
V2 ^
*
2
*
J
V
.
(3.16)
There exist two common ways in which to handle the issue of convergence for such functional integrals, i.e. to ensure acceptable behavior for large ?2. One is to give the mass to give the mass an infinitesimal negative imaginary part, m2 —• m2 — ie. This is the approach we shall employ in the discussion to follow. The second involves a continuation to euclidean space by means of t —• — ir wherein the functional integral becomes ,
(3.17)
Functional integration
479
and is now convergent due to the negative argument of the exponential. Continuation back to Minkowski space then yields the desired result. Integrating by parts, we have from Eq. (3.16) W (0) [j] = N f [dip] g
= N f \d(p'] (3.18)
where Ox — Hx + m2 — ie and (pf(x) = (p(x) + / d% AF(x - y)j J (2?r)4
fc2
— m2 + ze
( n x + m 2 )A F (* - y) = -£ ( 4 ) (* - 2/). Note that we have used invariance of the measure (/ [d(p] = J [dip']). Finally, we recognize a factor of W[0] in Eq. (3.18), thus leading to the expression .
(3.20)
We can now determine the Green's functions for the free field theory, e.g.. L - X2)
,
3=0
(~i)4 W[0] 6J(
3=0
(3.21) More interesting is the case of a self-interacting field theory for which the lagrangian density becomes
C{x) = \d^d^
- ^ m V + £ i n t M = 6«\v) + £int(ip) .
(3.22)
The theory is no longer exactly soluble, but one can find a perturbative solution by use of the generating functional W[j] = N f[dip(x)} J
(3.23)
480
Appendix A
As before, the Green's functions of the theory are given by W[0]
16 FT rr ~ 1
3=0
(3.24) For most purposes one requires only the connected portions of the Green's function, i.e. those diagrams which cannot be broken into two or more disjoint pieces. This is illustrated in Fig. A-l which can be found by dividing the full Green's function ... tp(xn))\0)
(3.25)
into products of connected particle sectors and dividing by the vacuumto-vacuum amplitude (0|0) in each sector. Mathematically one eliminates the disconnected diagrams by defining W[j] = eiZ^ .
(3.26)
Then one can show that Z[j] is the generating functional for connected Green's functions, = J2~\n J dXl... J dxnj(xi). -°° n=o - -°°
..j{xn)G^ (3.27)
where 6n
(3.28) j=o
A.4 Quadratic forms The most important example of a soluble path integral is one that is quadratic in the fields because, at least formally, it can be solved exactly. Let us consider an action quadratic in the fields, S=
f dAx
(4.1)
(c)
Fig. A-l. Contributions to the four-point Green's function in tp4 theory: (a)-(b) connected, (c)-(d) disconnected.
Functional integration
481
where O is some differential operator which may contain fields distinct from (p within it. The general result for the quadratic path integral is given by
/quad = J[Mx)]e~ifdlx
^ ) O ^ ( X ) = iV[det O}'1'2
,
(4.2)
where det O is the determinant of the operator O. In order to prove this, one can expand
n
(4-3)
n
where (pn(x) satisfies O(pn(x) = Xn(pn(x)
and
/ dAx
(4.4)
The sum over all field values can then be performed by summing over all values of the expansion coefficients a n , r
poo
= N IJn /
d(ln
(4.5)J
„ J-oo 'OO
v
poo
= N
Tl
dan
e-iKa™ = Nf (det O)~ 1/2
OO and where TV, TV' are normalization constants
det O = J J \n
(4.6)
n=l
denotes, as usual, the product of operator eigenvalues. In general, some effort is required to evaluate the determinant of an operator. One valuable relation, easily proven for finite dimensional matrices and generalizable to infinite dimensional ones is* det O = exp(tr lnO) .
(4.7)
This trace now denotes a summation over spacetime points, i.e. tr lnO= fd4x (x | lnO \ x) , which is the most commonly used form in practice. * For a discrete basis, this follows from the result exp(tr lnO) = e x p Y ^ l n A n =TTexp(lnA n ) = n
n
where An, are the eigenvalues of the operator O.
(4.8)
482
Appendix A
Background field method to one loop We can illustrate one use of this result by constructing an expansion about a background field configuration (which satisfies the classical equation of motion) and retaining the quantum fluctuations up to quadratic order. Consider a scalar field theory with interaction C-mt (?(#)). We define (p as a solution to ( • + m 2 )
(4.9)
Writing
(4.10)
leads to the generating functional W\j] = e ^ ^ / [dS
S [
+ £ int (4.12)
Integration by parts gives
H where Ox = nx + m2-£^((p(x))
.
(4.14)
The functional integration can then be performed (cf. Eq. (4.5)) and we obtain W[j] = const, (det Ox)~1/2 e{*SMx)]+if#xj(x)
(4 15)
It is convenient to normalize the determinant somewhat differently by defining O0x = nx + m2 .
(4.16)
Then, suppressing the x subscript, we write (detO)~ 1/2 = const, (detO^ l O)~ 1 ' 2
,
(4.17)
where const. = (det O 0 )" 1 / 2
,
(4.18)
and O0-1O = l + A F C t (
.
(4.19)
Functional integration
483
Using Eq. (4.2) we have
The generating functional for connected diagrams can now be identified immediately as
Z[j] = S [cp] + Jd4x j(x)(p(x) + l- IV In (l + A F C t (
^j(x)
(4.21)
The trace 'Tr' includes the integration over spacetime variables and can be interpreted as follows, Trln
- a;)C t (^) ,
(4.22)
IV [ A F C t (^) A F C t (^)] = J d*x J d4y AF(x - y)C'U {Cp{y)) x AF(y - x ) C In this manner, one-loop diagrams containing arbitrary numbers of C'(nt ((p) factors are generated. The physics associated with this approximation can be gleaned from counting arguments. The overall power of ft attached to a particular diagram can be found by noting that associated with a propagator and a vertex are the powers % and %~l respectively. There is also an overall factor of ft for each diagram. Then with the relation No. internal lines — No. internal vertices = No. loops — 1 , we see that this approximation corresponds to an expansion to one loop. The classical phase generates the tree diagram (O(h0)) contribution and the determinant yields the one-loop [O{ft1)) correction to a given amplitude. A.5 Fermion field theory Thus far, our development has been performed within the simple context of scalar fields. It is important also to consider the case of fermion fields where the requirements of antisymmetry impose interesting modifications on functional integration techniques. The key to the treatment of
484
Appendix A
anticommuting fields is the use of Grassmann variables. Thus, while ordinary c-number quantities (hereafter denoted by roman letters a, 6,...) commute with one another, [a,a] = [a,6] = [a,c] = . . . = 0 ,
(5.1)
the Grassmann numbers (hereafter denoted by Greek letters a, /?,...) anticommute, {a,a} = {a,(3} = {a,7} = ... = 0 .
(5.2)
It follows that the square of a Grassmann quantity must vanish, a
2=/?2=72
=
_
=
Q
>
( 5 3 )
and that any function must have the general expansion f(a) = /o + ha ,
g(a, P)=go + g\a + g2p + g^ap .
Differentiation is defined correspondingly via d*_dp_ _ d0_da_
(5.4)
_
so that in the notation of Eq. (5.4) we have
£(«) = / i ,
^(«,/?)=»-«.«•
(5-6)
Second derivatives then have the property d2
(57)
°
We must also define the concept of Grassmann integration. If we demand that integration have the property of translation invariance
Jdaf(a) = Jdaf(a + (3) ,
(5.8)
[dafi(3 = 0 or
(5.9)
it follows that f da = 0 .
The normalization in the diagonal integral can be chosen for convenience,
fdaa = l,
fdaf{a) = f1 .
(5.10)
Let us extend this formalism to a matrix notation by considering the discrete sets a = {a\,..., an} and a = {a\,..., an} of Grassmann variables. A class of integrals which commonly arises in a functional framework is W[M] = fdan...
dax dan... daxe^Ma
.
(5.11)
Functional
integration
485
As an example, t h e simple 2 x 2 case is calculated to be
W[M] =
J
da2 da\
1 + iaiMijOLj
(5.12) M22 - M\2M2i)]
.
Only the final term survives the integration, and we obtain W[M] =detM .
(5.13)
This result generalizes to the n x n system [Le 82] yielding essentially the inverse of the result found for Bose fields, W^[M]Fermi = / dan ... doi\ dotn ... da\etaMa = det M , a*n.. ,da\dan..
.da\e
a
* M a oc (det M)
l
.
We can now extend this formalism to the case of fermion fields i^{x) and ^{x). Since such quantities always enter the lagrangian quadratically, the functional integral can be performed exactly to yield W[O] = j [dip] [di>] jJ*A*^x)Oi>{x)
= NdetO
.
(5-15)
The remaining development proceeds parallel to that given for scalar fields. Given the free field lagrangian density £o($,ii>)='ii>(x)(ip-m)tl>(x) ,
(5.16)
the generating functional for the noninteracting spin one-half field becomes
Wfafj] = / where Ox = iif)x — m+ie and f){x), rj(x) are Grassmann fields. Introducing the change of variables ij)'(x) = ip{x) - / d4y SF{x,y)r)(y) , $'(x) = rp(x) - / d4y fj(y)SF(y,x)
,
486
Appendix A
we find that an alternative form for the generating functional is
W[TJ,f}} = ( [<#'] [dtf] e> 01 e (5.19) Thus the generating functional for connected diagrams is
Zhf}} = -fd4xJd4yfj(x)SF(x,y)V(y)
,
(5.20)
and the only non-vanishing connected Green's function is Xo)
(_
-
^ - m + le
(2TT) 4
which is the usual Feynman propagator. A.6 Gauge theories For our final topic, we examine gauge theories within a functional framework. We shall employ QED as the archetypical example, for which the action is 1
f
A
1 f A
— — - / d x F F^v — - / d x (A n A^ — / J
J
(6.1)
where the second line follows from the first by an integration by parts and
Of = gVnx-d£dZ •
(6-2)
In the presence of a source j ^ , the generating functional is then
J\ = N f[dA^]eiS[A» ]+ifd4x
j A
» " .
(6.3)
Due to the bilinear form of Eq. (6.1), it would appear that one could perform the functional integration as usual, resulting in
where the inverse operator DF^(x,y)
is defined as -y)
.
(6.5)
Functional integration
487
However, this is illusory since the inverse does not exist. That is, acting on Eq. (6.5) from the left with the derivative d* yields 0xDFfll/(x,y) = d^4\x-y)
,
(6.6)
implying that Dp^v must be infinite. An alternative way to demonstrate that OxV is a singular operator is to observe that
O£/d£a = 0 .
(6.7) V
Thus any four-gradient d*a is an eigenfunction of Ox having eigenvalue zero, and an operator having zero eigenvalues does not possess an inverse.
Gauge fixing The occurrence of such a divergence in the generating functional of a gauge theory can be traced to gauge invariance. For QED, any gauge transformation of vector potentials (cf. Eq. (II—1.3)), A^x) -> A'^x) = A^x) + -d^x)
,
(6.8)
leaves the action invariant, '
x)] .
(6.9)
If we partition the full field integration [dA^\ into a component which includes only those configurations which are not related by a gauge transformation and a component [da] which denotes all possible gauge transformations, then we have
[[dAJ e i 5 ^> = f[dAp] eiS^
x f[da] .
(6.10)
But J[da] is clearly infinite and this is the origin of the problem. The solution, first given by Faddeev and Popov [FaP 67], involves finding a procedure which somehow isolates the integration over the distinctly different vector potentials A^x). In order to understand this technique, we shall first examine a finite-dimensional analog [Ra 89]. Consider the functional
Z[A] = f n r dx\ e~^^XkAklXl , L / J
(6.11)
where A is an N x N matrix. Suppose that A is brought into diagonal form A D by linear transformation R, .
(6.12)
Appendix A
488
Letting y = Hx denote the coordinates in the diagonal basis, we have N rN
a2 AD
e -LA I
.i=lJ
(6.13)
Suppose that the last n of the N eigenvalues belonging to A vanish. The exponential factor in Eq. (6.13) is then independent of the co-ordinates yN-n+u • - • 5 UN and the corresponding integrations / dy^-n+i • • • / dyjy diverge. This is reflected in the vanishing of detA, and causes the quantities in Eq. (6.13) to diverge. The infinity is removed if the integration is restricted to only variables associated with nonzero eigenvalues, in which case we obtain the finite result (6.14) It is possible to express ^ [ A ] as an integral over the full range of indices 1 < i < N by defining variables Z%
\ arbitrary (AT - n + 1 < i < N) , and writing for the generating functional rN Zf[A] =
(6.15)
(6.16)
.2=1
Upon tranforming back to an arbitrary set of coordinates {#;}, we obtain the useful expression Zf[A] =
'N
r
1 1 / dxi
det
dz
N
f[
6(ZJ(X)) e
^K
. (6.17)
j=N-n+l
Let us now return to the subject of gauge fields, broadening the scope of our discussion to include even nonabelian gauge theories. By analogy, corresponding to the variables 2#_ n +i,..., ZN will be the gauge degrees of freedom and the prescription of Faddeev and Popov becomes for generic gauge fields A^x), 6{Gb{A%)) det \SGb/6aa\
(6.18)
b=l
where the {a a } are gauge transformation parameters (cf. Sect. 1-4) and the {G^(A^)} are functions which vanish for some value of A^(x). Since
Functional integration
489
the {Gb} serve to define the gauge, such contributions to the generating functional are referred to as gauge-fixing terms. The variation 6Gb/6aa signifies the response of the gauge-fixing function Gb to a gauge transformation parameter aa. For any gauge theory, there are a variety of choices possible for the gauge-fixing function G. In QED, one defines the axial gauge by G(A^ = n ^ , (6.19) where n^ is an arbitrary spacelike four-vector. Due to the presence of the four-vector nM, one must forego manifestly covariant Feynman rules in this approach. Thus, one often employs a covariant gauge-fixing condition such as G(AtA) = dfiAtM-F , (6.20) where F is an arbitrary constant. Under the gauge transformation of Eq. (6.8), we find G(AJ -> G(AJ + Ha ,
(6.21)
6G/6a = U .
(6.22)
so that Referring back to the general formula of Eq. (6.18), we see in this case that det \6G/6a\ is independent of the gauge field and thus may be dropped from the functional integral. The QED generating functional then becomes
(6.23) Note that, as promised, this result is finite and leads to a photon propagator in Feynman gauge XV
''
W[0}6ju(x)6jx(yyj»= 0
%
] (2TT) (2TT)4
(6.24) The result is independent of the choice of F. Consequently, even if the constant F is evaluated to the status of a field F(x), one can functionally integrate over F(x) with an arbitrary weighting factor since this will only affect the overall normalization of the generating functional. A common choice is
J[[dF] 6(d»A, -
F(x))e~%fd4x
F
^
= «T* !"*
{d A )2
"»
,
(6.25)
490
Appendix A
where £ is a real-valued parameter. In this case, the generating functional becomes ^
.
[O.ZO)
The integrand of the above spacetime integral can be regarded as the effective lagrangian of the theory, and the gauge-fixing term appears as one of its contributions. At this point, the functional integration can be carried out with impunity to obtain v)
?
( 6>27 )
where D^f is defined as
{Ux3r - (1 - C 1 ) ^ ) DFvX{x -y) = 8^\x
- y) .
(6.28)
We find in this way the form of the photon propagator in an arbitrary gauge, as appearing in Eq. (II—1.17).
Ghost fields In the path integral formalism, if the generating functional can be written in purely exponential form, then one can read off the lagrangian of the theory from the exponent. However, the general formula in Eq. (6.18) for a gauge-fixed generating functional contains a seemingly nonexponential factor, the determinant factor det \6Gb/8aa\. A fruitful procedure, due to Faddeev and Popov, for expressing the determinant as an exponential factor is motivated by the identity (cf. Eq. (5.15)), det M = N f[dc][dc] eimc
,
(6.29)
where c, c are Grassmann fields. This identity suggests that we replace the determinant factor with an appropriate functional integration over Grassmann variables. For QED, the generating functional can then be written in the concise form
= NJ[dA,}[dc][dc} e* (6.30) As pointed out earlier, for this case the integration over c, c yields only an unimportant constant and may be discarded. However, for nonabelian gauge theory Eq. (6.30) generalizes to
W\fi\ = [ ^ a,b,d
(6.31)
Functional integration
491
where repeated indices are summed over. The quantities (,3 2 ) will generally depend upon the fields A^ themselves. Thus, the fields {ca} , {ca} will appear as degrees of freedom in the defining lagrangian of the theory. However, although coupled to the gauge fields A^ through cMc, they do not interact with any source terms and therefore can only appear in closed loops inside more complex diagrams.* Since these Grassmann quantities are unphysical, they are often called Faddeev-Popov ghost fields. They are scalar, anticommuting variables which transform as members of the regular representation of the gauge group, e.g. for the gauge group SU(ri), there are n 2 — 1 of the {ca} and {ca} fields. To complete the discussion, let us determine the ghost-field contribution to the QCD lagrangian. We choose F^ = d^A^ and note the form of a gauge transformation (cf. Eqs. (I-4.12),(I.-4.17) with aa infinitesimal),
A* - 4 f = A% + -d^ab - fbaeA%ae
.
(6.33)
Then we find from a direct evaluation of dFi,/dac followed by the rescaling -g^cc - • cc, gsfbaeAfiCe .
(6.34)
Upon performing an integration by parts in the first term and relabeling the indices in the second, we obtain the ghost contribution to the QCD lagrangian of Eq. (II-2.25).
Problems 1) The van Vleck Determinant The semiclassical approximation to the propagator (valid as h —• 0) can be derived by expanding about the classical path. Writing
x(t) = xcl(t) + 6x(t) , we have
where
6x(t)6x(tf)
\dt2
dx2cl(t) J
Such loops must include a multiplicative factor of —1 to account for the anticommuting nature of these variables.
492
Appendix A
and we have dropped the term linear in 8x(t) by Hamilton's condition. Performing the path integration we have then S S
f trx- t) - N (det
1
2
where N is a normalization constant and the quantity inside the square root is called the van Vleck determinant a) Show that this can be written in the form 6x(t)6x(tf)\ [2ni dxfdx{ J Hint: The following argument is hardly rigorous but leads to the correct answer. Write D(xf,tr,Xi,U)
= A(xf,Xi;tf
-ti)eiS«( x''Xi*'-u)
and use completeness to show that at equal times S(xf — xi) = D(xf, ti\ Xi, ti) — / dx A(xf, x; T)A*(xu x; T)e^s^xf^T^s^Xi^T^
,
where T is an arbitrary positive time. Now define p(xi,x;T) = dSc\{xi,x]T)/dxi so that , x\ T) - Sc\(xu x\ T) ~ (xf - Xi)p(xu x; T) . Finally, change variables from x to p and compare with the free particle result to obtain
b) Show that rxf . Sc\(xf, x{\ T) = -ET + / dx^/2m(E - V{x)) J Xi
and verify that [27rti c i(t / )i d (ti) /*/ dx i^ 3 (x) J Hint: Recall that t is an independent variable, so that dt dt
Functional integration
493
We thus have the result for the semiclassical propagator -i 1/2 \
m
f
iQ,
s
l dxxcl (x)\ which is identical to that found from WKB methods. 2) Propagator for the Charged Scalar Field The lagrangian density for a charged scalar field
\X , JL i —
can be written as DF(x'\x) = -iix'WDn
+ m2 - ie)~l\x) .
Suggestion: This is a quadratic form. Use the generating functional to integrate it. b) By expanding Dp{xf\x) as a power series in A^{x)^ show that an alternative representation for the propagator is oo
3) Functional Methods and cp4 Theory Consider a scalar field theory with the self-interaction
a) Show that the generating functional can be written as
where the free field Feynman propagator iA^(x, y) is as in Eq. (C2.12). b) Evaluate the two-point function to O(A2). Associate a Feynman diagram with each term of this expansion and separate the connected and disconnected diagrams. c) Calculate the connected generating functional via ] = Z0[j] - i In
494
Appendix A where z
o[j] = -jd4xl
d4y j(x)iAF(x,y)j(y)
.
d) Compare the connected diagrams found in parts (b) and (c).
Appendix B Some field theoretic methods
B.I The heat kernel When using path integral techniques one must often evaluate quantities of the form H{x,r) = {x\e-rV\x) ,
(1.1)
where V is a differential operator and r is a parameter. In this section, we shall describe the heat kernel method by which H(X,T) is expressed as a power series in r. For example, if in d dimensions the differential operator T> is of the form V = n + m2 + V ,
(1.2)
where V is some interaction, then the heat kernel expansion for H(x, r) is ^
H{X,T)
= ——^
e-Tm
—^
2
[ao(x) + ai(x)r + a2(x)r2 + ...] .
(1.3)
where ai{x) are coefficients which will be determined below. Let us begin by citing the two most common occurrences of H(x,r). One is in the evaluation of the functional determinant detP = etTlnV = Jd4x ^W^W
,
(1.4)
where ' Tr' is a trace over internal variables like isospin, Dirac matrices, etc., and ' t r ' is a trace over these plus spacetime. The (generally singular) matrix element (x|lnD|x) appearing in Eq. (1.4) can be expressed in a variety of ways. For example, in dimensional regularization one can use the identity a
Jo
x 495
Appendix B
496 to write
(x\ \nV\x) = - f°° — (x\e~rV\ x) + C , ./o r
(1.6)
where C is a divergent constant having no physical consequences. Substituting Eq. (1.3) into the above yields (x\lnV\x)-C = -
(1.7)
The divergences in the series representation arise from the F-function and are restricted in four dimensions to the terms ao(x), a\ (x), a
1
X
(1.8) 5=0
r(«) Jo The penultimate equality in Eq. (1.8) is obtained from repeated formal differentiation of Eq. (1.6) with respect to V. Upon expanding the H(x, r) term in £p(#>s), one arrives at the desired power series expansion of (x\ lnT>\x). This usage is applied in the next section. The other main use of the heat kernel is in the regularization of anomalies. Often one is faced with making sense of Tr (x \O(x)\ #), where O is a local operator. Although such quantities are generally singular, they can be defined in a gauge-invariant manner by damping out the contributions of large eigenvalues, Tr (x \O(x)\ x) = lim Tr (x \O{x)e~eV\ x)
(1.9)
where V is a gauge-invariant differential operator. Again it is only the low-order coefficients, generally those up to a2(x), which contribute in the €—• 0 limit. We employ this technique in Sects. 111-3,4. As an example of heat kernel techniques, let us consider the following operator defined in d dimensions: V = dyp + m2 + a(x)
(d»
+
M))
(110)
where T^(x) and a(x) are functions and/or matrices defined in some internal symmetry space. In particular, neither FM nor a contains derivative operators. Employing a complete set of momentum eigenstates {\p)}
Some field theoretic methods
497
allows us to express the heat kernel as H{X T) =
'
J Wjd e-^e-^e^
,
(1.11)
where in d dimensions, use is made of the relations
x
'~x)=6{d){x
~x>)'
(L12)
Prom the identities d^^e^ip.
+ d,),
we can then write H(X,T)=
- /
f,
\p2-m2]f,-T[d-d+a+2ip-d]
The first exponential factor is simply the free field result, while all the interesting physics is in the second exponential. The latter can be Taylor expanded in powers of r, keeping those terms which contribute up to order r 2 after integration over momentum. Note that each power of p2 contributes a factor of 1/r. Thus we obtain the expansion H(X,T)
= r
-^[(d-d + a){d -d +
cr)-4p-dp-d] (1 15)
4T 3
— \p-dp-d(d-d
+ a) +p
• d(d-d +a)p - d
v
'
'
16r4 where terms odd in p have been dropped and we have displayed only those (9(T 3 ) and O ( T 4 ) terms which contribute to H at order r 2 after p is integrated over. To perform the integral, it is convenient to continue to euclidean momentum PE — {pi,P2>P35P4 = — ipo}- Then with
498
Appendix B
the replacement p^ o^ —> —
\p^p^\ = —p\, we obtain
1 <*>
27r d/2
j
e -m^-
d
2-Td/2
T(d/2) (27r)
JR
/ C / (*^Tr\d
1
ft
T) JP
/
O
)
Vr
T%
1
e" m 2 T
( 47r )d/2
r d/2
~~~
d
d 2
(4?r) /
TW2)
'
rf 2+1
r /
T(d/2 + l) (1.16) T(d/2)
2 4 w+i e- r '
ON
v -^
Employing these relations to evaluate Eq. (1.14) gives (to second order inr), H(X,T) =
x
[l-
or in the notation of Eq. (1.3), OQ(X) = 1 ,
O2(x) = ^
2
a\{x) = -a , + ^ [ d ^ p ^ + ^ [dM, [^,or]] .
(1 18)
'
Fermions are treated in a similar manner. For example, the identity
i
(1.19)
allows the same technique to be used for the operator Iplp. In particular let us consider the case where
Some field theoretic methods
499
With some work, one can cast this into the form of Eq. (1.9) with the identifications
75
The values of ai{x) appearing in Eq. (1.18) can also be used in this case. The heat kernel coefficients have been worked out for more general situations [Gi 75]. B.2 Chiral renormalization and background fields In this section, we illustrate the method described above while also proving an important result for the theory of chiral symmetry. The goal is to demonstrate that all the divergences encountered at one loop can be absorbed into a renormalization of the coefficients of the O(E4) chiral lagrangian and to identify the renormalization constants. The technique used here, the backgroundfieldmethod, is of considerable interest in its own right [Sc 51, De 67, Ab 82, Bal 89] and is applicable to areas such as general relativity [BiD 82]. The basic idea of the background field method is to calculate quantum corrections about some nonvanishing field configuration Tp^
(2.1)
rather than about the zero field,* and to then compute the path integral over the fluctuation 6ip(x). The result is an effective action for (p. This effective action can be expanded in powers of Tp and applied to matrix elements at tree level, resulting in a description of scattering processes at one-loop order. In the case of the chiral lagrangian, one expands the full chiral matrix U = U + 8U ,
(2.2)
where U satisfies the classical equation of motion. Upon integration over 6U, one obtains the one-loop effective action for U. This contains a great deal of information. In particular, U can be expanded in the usual way in terms of a set of external meson fields U = exp(iAa<pa/F) * See the discussion in Appendix A-4.
(a = l,...,8) .
(2.3)
Appendix B
500
Contained in Ses(U) is the effective one-loop action for arbitrary numbers of meson fields. Upon identification of renormalization constants, all processes become renormalized at the same time. Our starting point is, in the notation of Sect. IV-6, the O(E2) lagrangian £ 2 = ^L Tr
+ ^L Tr
+
(2.4)
The procedure to follow is rather technical, so let us first quote the end result of the calculation. Upon performing the one-loop quantum corrections, the effective action will have the form Here the lagrangians in S^611, S™ n are the ones quoted in Sect. VI-2, but now with renormalized coefficients. In particular 5Jen is the sum ST4en = Sj 8 " 5 + Sfv where, in chiral SU(3) and employing dimensional regularization, S$w is given by
Tr
16
Tr (x T C/ + + Ux1) I lL[Tr(XUi \ Tr 4
(2-5)
+ Ux<)]2 + ^Tr
(L^D^UD^ V
+
+ RurLrrfiyu) - 7Tr /
4
with
The terms in S ^ are all of the same form as the terms in the bare lagrangian at order E4. Therefore, all the divergences can be absorbed into renormalized values of these constants. The finite remainder, 5fmte, cannot be simply expressed as a local lagrangian, but can be worked out for any given transition. When S$1V is added to the O(E*) treelevel lagrangian of Eq. (VI-2.7), the result has the same form but with coefficients ,
(2.7)
Some field theoretic methods
501
where the {ji} are numbers which are given in Table B-l for both the case of chiral SU(2) and SU(3). Thus the divergences can all be absorbed into the redefined parameters and these in turn can be determined from experiment. Let us now turn to the task of obtaining this result. In applying the background field method, there are a variety of ways to parameterize 8U, and several different ones are used in the literature. The prime consideration is to maintain the unitarity property UW = 1 = (W + 6tf) (U + SU) along with WO = 1. We shall take U = UeiA ,
(2.8)
with A = AaAa representing the quantum fluctuations. This choice is made to simplify the algebra in the heat kernel renormalization approach, which we shall describe shortly. Another possible choice is U = £eir>Z
(2.9)
with 77 = \arf and ££ = {/. These two forms are related by 77 = Since in the path integral, we integrate over all values of A (or 77) at each point of spacetime, these two choices are equivalent. The expansion of the lagrangian in terms of U and A is straightforward, and we find
Tr
(D^UD^UA
= Tr (D^UD^UA
-2iTr
+ Tr ID^AD^A + U^DJJ (AD^A - D^A A)1 , iTr ( A (2.10) where D^Asd^A + i ^ . A ]
.
(2.11)
Table B-l. Renormalization coefficients i SU(2)
7<
SU(3)
7<
1
2
3
4
5
6
7
8
9
10
1 12 3 32
1 6 3 16
0
l 8 1 8
l 4 3 8
3 32 11 144
0
0
0
5 48
I 6 1 4
1 6 1 4
0
502
Appendix B
Since U satisfies the equation of motion, there is no term linear in A. One may integrate various terms in the action by parts to obtain ^
(2.12)
where
rjf = - i Tr ([Aa, A6] (p%U + iU%U + ir aab = 1 Tr ({ Aa, A6} (xf£/ + U]x) + [Aa,tff£>„£/] [A6, (2.13) The action is now a simple quadratic form, and the path integral may be performed. The only potential complication is the question of interpreting the integration variables. This is referred to as the 'question of the path integral measure'. The integration over all the unitary matrices U can be accomplished by an integration over the parameters in the exponential
= N j[dAa],
(2.14)
where AT is a constant which plays no dynamical role. With this identification one obtains eiZiooP =
ndAa]eifd*x^A«(d^+<7)" bAb
= (det [d^
+ a])"
172
( 1 1 = exp I - - tr In {d^df + a)\ .
(215)
Here ' t r ' indicates a trace over the spacetime indices as well as over the SU{N) indices a, b. The identification of divergences is most conveniently done by using the heat kernel expansion derived earlier in this appendix, where it is shown that all the ultraviolet divergences are contained in the first few expansion coefficients. The relevant terms are Z\ooP = ^ tr In [d^
where
t
+ <J)
Some field theoretic methods
503
For Nf flavors, the operator part of the first term in Eq. (2.16) is
+ ^ ^ IV (x]U + &X) .
Tr a = ^ TV (pfiirffl)
(2.18)
The above two traces are just those which appear in C2 , so that they can only modify Fn and ra2. The remaining terms can be worked out with a bit more algebra. Using the identity J,U]dvu\
,
(2.19)
we find for the field strength, Tab = - T r {\\a
Xb1 (\WD
U U^Djytf] +iU^L
UU
+ iR j \ \ (2.20)
This produces, for Nf flavors in chiral SU(Nf),
Tr (F^F^) = -^Tr (lu^DM.U^D^u] \u^D^U,U]Duu\] 8
+ iNfTr
VL
i
J L
J/
/
(RfU/& U^ff U + .
- Nf Tr {L^UR^U^ - ^j- Tr {L^W + o 1r / -+\i2 1 / -A Tr a1 = - Tr (DJJD^W 1 + - Tr (D U UD U W ) Tr p 8 L V / J 4 \ / + - ^ Tr (D^UD^WDJ.UD^W)
+
^
[Tr fx^T +
/ + - Tr (D^UD^U]\ Tr (\U] + f7xf)
(2.21) + !~ Tr which is not of the same form as the basic O(E4) The only operator lagrangian occurs in the first term of Tr F 2 . However, by use of Eq. (VI2.3) for SU(3), it can be written as a linear combination of our standard forms. For Nf — 3, these add up to the result previously quoted in Eq. (2.5). Here the divergence is in the parameter A. For convenience in applications, we have added some finite terms to the definitions of A. The results for Nf — 2 are also quoted in Table B-l, although some of the operators are redundant for that case.
504
Appendix B
The reader who has understood the above development as well as the standard perturbative methods presented in the main text will be prepared for the use of the background field method in the full calculation of transition amplitudes. This procedure consists of writing Vo = • + m 2
(2.22)
where m2 is the meson mass-squared matrix. The one-loop action is then expanded in powers of the interaction V Zloop = \ tr i
\
1
= - tr pnPo + V^V - -V^VV^V
1 .. .j . +
The first term is an uninteresting constant which may be dropped, and the remainder has the coordinate space form [AF(x - x)V{x)}
--J
cTWyTr [AF(x - y)V(y)AF(y - x)V(x)\ + ... .
(2.24)
When the matrix elements of this action are taken, the result contains not only the divergent terms calculated above, but also the finite components of the one-loop amplitudes. The resulting expressions are presented fully in [GaL 84,GaL 85a]. This method allows one to calculate the one-loop corrections to many processes at the same time, and in practice is a much simpler procedure for some of the more difficult calculations.
Appendix C Useful formulae
C.I Numerics Conversion factors (h = c = &B = 1): 1 GeV" 1 = 6.582122 x 10~25 s
1 GeV = 1.16 x 10 13 K = 1.78 x 10~24 g
= 0.197327 fm Physical Constants: GM = 1.16637(2) x 10~5 GeV~ 2
G^ 1 / 2 = Mpi = 1.2 x 10 19 GeV
a " 1 = 137.0359895(61)
sin2 9V = 0.226(5)
mw = 80.6(4) GeV/c 2
mz = 91.175(21) GeV/c 2
me = 0.51099906(15) MeV/c 2
mp = 938.2723(3) MeV/c 2
F w = 92.4(2)MeV
FK = 112.7(2.1) MeV
\r)+-\ = 2.268(23) x 10"
3
|fX»| = 2.253(24) x 10~ 3
CKM Matrix Elements: = 0.220(2)
\Vud\= 0.9744(10)
IFUSI
|Vcd| = 0.204(17)
\Va\ = 1.02(18)
\Vch\ = 0.046(7)
\Vuh\ = 0.005(2)
C.2 Notations and identities Metric tensor: 1 0 0 -1 0 0 0 0
0 0 -1 0 505
0 0 0
(2.1)
506
Appendix C
Totally antisymmetric four-tensor:
{
eiU,«f> ev'«'P' =
+1 {/i, v, a, /?} even permutation of {0,1,2,3} — 1 odd permutation 0 otherwise g^'gw'gW + g^'g^'g^' + g^'gaa'g^ (2.2)
Totally antisymmetric three-tensor:
{
+1 —1 0
eOijk = -eOijk
{z,j, k} even permutation of {1,2,3} odd permutation otherwise = eijk = eijk
(2.3)
Pauli matrices: ^k
k
ie^lal
( j , k 9 l = 1, 2,3) a
, 6, C, d = 1, 2)
Dirac matrices: • n i
2 s
75 = -*7 7 7 7 (2.5)
7°rj7°= - r ,
(r* = 75)
Trace relations: Tr (7") = 0 Tr (75) = 0 Tr ( 7 "7 I/ ) = 4
=0
(2.6)
Useful formulae
507
Plane wave solutions: The Dirac spinor tx(p, s) is a positive-energy eigenstate of the momentum p and energy E = >/p2 + ra2. Antifermions are described in terms of the Dirac spinor v(p, s). The adjoint solutions are denoted by u = 1^7° and v = v^j°. Note that our normalization of Dirac spinors behaves smoothly in the massless limit.
(p - m)u(p, s) = 0 u(p,s)(fi-m) = 0 U> + m)v(p, s) = 0 v(p, s)(p + m) = 0 ii(p, r)w(p, s) = 2m^rs v(p, r)v(p, 5) = -2mSrs ^ f (p,r)^(p,5) = 2E6rs
(2.7)
S
Gordon decomposition for a fermion of mass m: S(pVb'Mp,
- S ( p » (&±&. + ^
r t
' ) »(P,») (2.8)
Dirac representation: -.
, , u(p, s) = y/E + m I
X a
p
0
\
)
v(p, 5) =
V Xs
(2.10)
Fierz relations: The anticommutativity of fermion fields and the algebra of Dirac matrices imply the (particularly useful) Fierz relations, + 75)^2 (2-11) 75)^2
.
508
Appendix C
Propagators: The propagators associated with fields
(x) = (O\T(M i D F X u ( x ) = (0\T
(
= f dpr)
w
\
)
A
2
4
J (27T)
e
2
2
+ ie)
2
p - M + ie (2.12)
where £ is a gauge-dependent parameter. C.3 Decay lifetimes and cross sections Parameters of choice for quantum fields: The literature reveals a variety of conventions employed in quantum field theory. We can characterize all of these with certain parameters of choice, Ji, Ki, Li (i = J5,F distinguishes bosons from fermions), occurring in the normalization of charged spin zero and spin one-half fields, H3k
^f- (a(k)e-ihx + a\k)eikx)
/
J
f ,3
5
s
>(*) = E J 7 ^ W P > M P > > ~
ipx
dt
s eip
+ (P>*MP' ) ") '
(3.1)
s
in momentum space algebraic relations, e.g.,
,r),&V, s )} =KF6rsSz(p-p')
(3 2)
,
'
and in the normalization of single-particle states
tf .
(3.3)
It is convenient to introduce an additional parameter NF to characterize the choice of fermion spinor normalization, u\p,r)u(p,8)
= NF2Ep6r8 .
(3.4)
For uniformity of notation, we also define NB = 1. The constants Ji,Ki,N{ are constrained by the canonical commutation or anticommu-
Useful formulae
509
tation relations to obey
Using the above, one can express the single-particle expectation value of the quantum mechanical probability density as (i = B,F)
(3.6)
.
The conventions employed in this book, together with the implied normalization for boson or fermion single-particle states, are LB
= LF
= NB
= NF
= 1 ,
J B = J F = KB
= KF
= 2£(2TT) 3 ,
- p) ,
(3.7)
where r, s are spin labels. This choice, although somewhat unconventional for fermions,* has the advantages that bosons and fermions are treated symmetrically throughout the formalism, the zero mass limit presents no difficulty, and matrix elements are free of cumbersome kinematic factors. Lifetimes: Prom the decay law N(t) = iV(0)e~*/r, the mean life r is seen to be the transition rate per decaying particle, F = r" 1 = —N/N. For decay of a particle of energy E\ into a total of n — 1 bosons and/or fermions, the S-matrix amplitude can be written in terms of a reduced (or invariant) amplitude M& as
(f\S - l\i) = -i(27r)4^4)(pi -P2...-Pn)
II n
/
„.
\l/2
(3.8) where the index k labels the individual particles as to whether they are bosons or fermions. The inverse lifetime is computed from the squared Smatrix amplitude per spacetime volume VT and incident particle density pi, integrated over final state phase space. The choice of phase space is already fixed by our analysis. Thus, defining a parameter of choice A(p) for the (momentum) phase space per particle, Phase space per particle * Another book sharing this convention is [ChL 84].
- /
AQT)
510
Appendix C
the application of completeness to Eq. (3.7) yields (p'|p) = / ^ ( p ' | k ) ( k | p )
=* A = KL2 = (2nfp .
(3.10)
The inverse lifetime (or decay width) is then given by
T _ 1=r =
H I (fr d\k \ \S-\\\ VT
fil2 (3.11) where Z = fj • rtjl is a statistical factor accounting for the presence of rij identical particles of type j in the final state, and the sum 'int' is over internal degrees of freedom such as spin and color. Cross sections: For the reaction 1 + 2 —> 3 + . . . n, the cross section a is the transition rate per incident flux. The incident flux /i n c can be represented as fine = PiPalvi - v 2 | = £^r[(pi • n? - m\ml}^
,
(3.12)
and the cross section becomes 1 1
int
(3.13) Watson's theorem: The scattering operator S is unitary, S^S = 1. Thus the transition operator T, defined by S = 1 - iT, obeys i(T - 7^) = T^T. With the aid of the relation (f\T^\i) = (i|T|/)*, we obtain the unitarity constraint for matrix elements, ,
(3.14)
where Tfi = (f\T\i). This constraint implies the existence of phase relations between the various intermediate state amplitudes. For example, consider a weak transition followed by a strong final-state interaction for which there is a unique intermediate state identical to the final state, A —> BC — • BC , weak
strong
(3.15)
Useful formulae
511
i.e. i = A,n = / = BC. In this circumstance, time-reversal invariance of the hamiltonian implies Tfi = Tjf, so the left-hand side of the unitarity relation reduces to —2Im7if and both sides of Eq. (3.14) are real-valued. Denoting the weak and strong matrix elements as |Tw|e**w and |Ts|e**s, it then follows that <5w = <$sC.4 Field dimension We consider a limit in which the theory is invariant under the set of scale transformations x1* —• \x^ (A > 0) of the spacetime coordinates. Associate with each such coordinate transformation a unitary operator U(X) whose effect on a generic quantum field $ is given by U(\)Q>(x)U^(\) — \d*${\x), where d$ is the dimension of the field $. Prom the canonical commutation relation obeyed by a boson field (p or the canonical anticommutation relation obeyed by a fermion field ipa, M> 3 (*) >
(4-1)
it follows that the canonicalfielddimensions are d^ = 1 and d$ — 3/2. Composites built from products of these fields carry a dimension of their own, e.g. all fermion bilinears xpT^ (F is a 4 x 4 matrix) have canonical dimension 3. Unless protected by some kind of algebraic relation, a field dimension will generally be modified from the canonical value by interaction-dependent anomalous dimensions. Field dimensions are particularly useful in ordering the terms contained in a short-distance expansion, A(x)B(0) X
where A, £, On are local quantum fields. From the scale invariance of the short-distance limit, it follows that cn(x) ~ xd°n~dA~d^. Thus the fields On of lowest dimension have the most singular coefficient functions. C.5 Mathematics in d dimensions Dirac Algebra: It is conventional to employ a metric g^ corresponding to a spacetime of continuous dimension d, but to maintain certain d = 4 properties of the Dirac matrices such as the trace relations of Eq. (2.6).* In the following, Id is a diagonal d-dimensional matrix with TV Id = 4 and e = 4 — d.
Since there exists no well-defined continuation of the 75 matrix torf-dimensions,we restrict all 75 relations to d = 4.
512
Appendix C
= dld -yvj>-f = (€-2)j>
(5.1)
t> k 7^ = 4p • qld ~ € j> fi
= 2p • qf + 2q • r$ - 2p
Feynman parameterization:
1 anbm
_ r(n + m) Z*1 - r(n)T(m)
r1
i
^^(l-g)^ 1
[ax + 6(1 - x)]n+™
Jo
r1
( n
'm>
I — = 2 / a; dx / dy r — -^ a6c 7o Jo [ a ( l " ^) + ^ + cx(l - y)}3 Integrals: For the following integrals, we define the denominator function \
ml(l-x)-q2x{l-x)-ie
,
}
(5.2) v
y
(5.3)
constrain ni,ri2 > 1, and denote ie as the infinitesimal Feynman parameter.
J WY [{p-qy-m\ {
'
r ddp
(47r)d/2
+ ie]n*{p*-ml + ie]n2 r(m)r(n2) Jo i?
r(m)r(n2)
[_*£.
d
O.
J (27T) [(p - q)2 - m 2 + i c ] "i [p2 _
m 2 + i£]
"2
( 54c)
^
C
)
Useful formulae
513
Solid angle: d 2
/
dO2 sin 02
ft2 = 2TT, ft3 = 4TT, ft4 = 2TT2, . . .
d6x =
i \d/2)
.
. (5.5)
Gamma, psi, beta, and hypergeometric functions:
T(z) =
oo
dt e~f tz~l
(Re z > 0) , 7o T(z + 1) = zT(z) = z(z - l)T(z - 1) = . . . = z\ ,
(-n + I) = ^ p [^ + V(« + 1) + O(e)l
(n integer) ,
dT(z)/dz = r ^ ) ^ ^ ) where V(^ + 1) = V>(-z) + l/« , + £ + . . . + - - Inn J ~ -0.5772 , 2
dip(z + l)/dz = ip'(z + 1) = tp'(z) - 1/z D,v
'
)
with ^'(1) = TT2/6 ,
(5.6)
..A _ r(n)r(m) _ y T(n + m)
f(a 6;C;z) =
'
n
2
i i r M ) I dtt^a-ir^Hl-**)" 0 (Rec>Re6>0),
„, , . r(c)r(c-a-6) , dF(a,6;c;z) a6 „ . n , Fo,6;c;l)= ) ; \ ^ and L = — F a +1,6 1 (c — ajl (c — b) az c
References
In the text, papers are cited by using the first two letters of the surname of the first author plus the first letter of that of each remaining authors plus the year of publication. For example, the second of the references below is cited as [AbGLOPR 91]. The only exception is the Review of Particle Properties, Phys. Lett B239, 1 (1990) which is cited as [RPP 90]. Experimental results given without citation can be found in [RPP 90]. Abbott, L. (1982). Introduction to the background field method, Ada Phys. Pol B13, 33. Abada, A., Gavela, M.B., Le Yaouanc, A., Oliver, L., Pene, O. and Raynal, J.-C. (1991). On the chiral estimate of the strong CP neutron electric dipole moment (Orsay preprint LPTHE 90/53 [revised]). Abdullah, K.C. et al. (1990). New experimental limit on the electron electric dipole moment, Phys. Rev. Lett. 65, 2347. Abe, F. et al. (CDF collab.) (1990). Measurement of the W-hoson mass, Phys. Rev. Lett. 65, 2243. Ademollo, M. and Gatto, R. (1964). Nonrenormalization theorem for the strangeness-violating vector currents, Phys. Rev. Lett. 13, 264. Adkins, G., Nappi, C. and Witten, E. (1983). Static properties of nucleons in the Skyrme model, Nucl. Phys. B228, 552. Adler, S.L. (1969). Axial-vector vertex in spinor electrodynamics, Phys. Rev. 177, 2426. Adler, S.L. (1970). Perturbation theory anomalies, in Lectures on Elementary Particle Physics, ed. S. Deser, M. Grisaru and H. Pendleton (MIT Press, Cambridge, Mass.). Adler, S.L. and Bardeen, W.A. (1969). Absence of higher-order corrections in the anomalous axial-vector divergence equation, Phys. Rev. 182, 1517. Adler, S.L. and Dashen, R. (1968). Current Algebras and Applications to Particle Physics (Benjamin, New York). Ahrens, L.A. et al. (1987). Measurement of neutrino-proton and antineutrino-proton elastic scattering, Phys. Rev. D35, 785. 514
References
515
Aitchison, I.J.R. and Hey, A.J.G. (1982). Gauge Theories in Particle Physics (Adam Hilger, Bristol). Akhundov, A.A., Bardin, D.Yu., and Reimann, T. (1986). Electroweak one-loop corrections to the decay of the neutral vector boson, Nucl. Phys. B276, 1. Albrecht, H. et al. (1990). Observation of semileptonic charmless B-meson decay, Phys. Lett B234, 409. Alitti, J. et al. (UA2 collab.) (1990). A precise determination of the W and Z masses at the CERN pp collider , Phys. Lett. B241, 150. Altarelli, G. (1989). Experimental tests of perturbative QCD, Ann. Rev. Nucl. Part. Sci. 39, 357. Altarelli, G. (1992). Constraints on new physics from precision measurements, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Altarelli, G. Barbieri, R. and Jadach, S. (1992). Toward a model-independent analysis of electroweak data, Nucl. Phys. B369, 3. Altarelli, G., Cabibbo, N., Corbo, G., Martinelli, G. and Maiani, L. (1982). Leptonic decays of heavy flavors: A theoretical update, Nucl. Phys. B208, 365. Altarelli, G., Kleiss, R. and Verzegnassi, C. (ed.) (1989). Z Physics at LEP1 (CERN 89-08, Geneva). Altarelli, G. and Maiani, L. (1974). Octet enhancement of nonleptonic weak interactions in asymptotically free gauge theories, Phys. Lett. B52, 351. Amaldi, U., Bohm A., Durkin L.S., Langacker, P., Mann, A.K., Marciano, W.J., Sirlin, A. and Williams, H.H. (1987). A comprehensive analysis of data pertaining to the weak neutral current and the intermediate vector boson masses, Phys. Rev. D36, 1&85. Amaldi, U., de Boer W., Fiirstenau, H. (1991). Comparison of grand unified theories with electroweak and strong coupling constants measured at LEP, Phys. Lett. 260, 447. Anderson, P.W. (1984). Basic Notions in Condensed Matter Physics (Benjamin/Cummings, Menlo Park, Calif.). Antipov, Yu.M. et al. (1985). Experimental estimation of the sum of pion electric and magnetic polarizabilities, Z. Phys. C26, 495. Aoki, K., Hioki, Z., Kawabe, R., Konuma, M. and Muta, T. (1982). Electroweak theory, Suppl. Prog. Theo. Phys. 73, 1. Appelquist, T. and Bernard, C. (1981). Nonlinear sigma model in the loop expansion, Phys. Rev. D23, 425. Appelquist, T. and Carrazone, J. (1975). Infrared singularities and massive fields, Phys. Rev. D l l , 2856. Ashman, J. et al. (1988). A measurement of the spin asymmetry and a determination of the structure function g\ in deep inelastic muon proton scattering, Phys. Lett. 206B, 364. Au, K.L., Morgan, D. and Pennington, M.R. (1987). Meson dynamics beyond the quark model, Phys. Rev. D35, 1633.
516
References
van Baal, P. and Kronfeld, A.S. (1989). Spectrum of pure glue theory, Nucl. Phys.(Proc. Suppl.) 9, 227. Bagger, J. and Schmidt, C. (1990). Equivalence theorem redux, Phys. Rev. D41, 264. Bahcall, J.N. (1990). Neutrino Astrophysics (Cambridge University Press, Cambridge). Bahcall, J.N. and Bethe, H.A. (1990). A solution of the solar neutrino problem, Phys. Rev. Lett. 65, 2233. Balachandran, A.P., Nair, V.P., Rajeev, S.G. and Stern, A. (1983). Soliton states in the QCD effective lagrangian, Phys. Rev. D27, 1153. Ball, R.D. (1989). Chiral gauge theory, Phys. Rep. 182, 1. Barbieri, R., Frigens, H., Giuliani, F., and Haber, H.E. (1990). Precision measurements in electroweak physics and supersymmetry, Nucl. Phys. B341, 309. Bardeen, W.A. (1969). Anomalous Ward identities in spinor field theories, Phys. Rev. 184, 1848. Bardeen, W.A., Buras, A.J. and Gerard, J.M. (1987). The K -» 2TT decays in the large Nc limit, Nucl. Phys. B293, 787. Barger, V., Hewett, J.L. and Phillips, R.J.N. (1990). New constraints on the charged Higgs sector in two-Higgs models , Phys. Rev. D41, 3421. Barger, V. and Phillips, R.J.N. (1987). Collider Physics (Addison-Wesley, Redwood City, Calif.). Barr, G. (1992). New Results on e/e' from NA31, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Barr, G. et al. (1990). Observation of the decay KL -+ TTO77, Phys. Lett. 242B, 523. Barr, S.M. and Marciano, W. (1989). Electric dipole moments, in CP Violation, ed. C. Jarlskog (World Scientific, Singapore). Bauer, M., Stech, B. and Wirbel, M. (1987). Exclusive nonleptonic decays of £>, Ds, and B mesons, Z. Phys. C34, 103. Beavis, D., Chu, S.-Y., Desai, B. and Kaus, P. (1979). Charmonium and T systems with small hyperfme splittings, Phys. Rev. D20, 743. Bell, J.S. and Jackiw, R. (1967). A PC AC puzzle: TT°—> 77 in the sigma model, Nuovo Cim. 60A, 47. Berman, S.M. (1958). Radiative corrections to pion beta decay, Phys. Rev. Lett. 1, 468. Bertolini, S.,Borzumati, F. and Masiero, A. (1987). QCD enhancement of radiative B decays, Phys. Rev. Lett. 59, 180. Bigi, I.I., Khoze, V.A., Uraltsev, N.G. and Sanda, A.I. (1989). The question of CP noninvariance as seen through the eyes of neutral beauty, in CP Violation, ed. C. Jarlskog (World Scientific, Singapore). Bigi, I.I. and Sanda, A.I. (1981). Note on the observability of CP violation in B decays, Nucl. Phys. B193, 85.
References
517
van der Bij, J.J. and Hoogeeven, F. (1987). Two-loop correction to weak interaction parameters due to a heavy fermion doublet, Nucl. Phys. B283, 477. Bijnens, J. (1990). K14 decays and the low energy expansion, Nucl. Phys. B337, 635. Bijnens, J., Sonoda, H. and Wise, M.B. (1984). Validity of chiral perturbation theory for K° - JC° mixing, Phys. Rev. Lett. 53, 2367. Bijnens, J. and Wise, M.B. (1984). Electromagnetic contribution to e'/e, Phys. Lett. B137, 245. Birrell, N.D. and Davies, P.C.W. (1982). Quantum Fields in Curved Space (Cambridge University Press, Cambridge). Bjorken, B.J. and Drell, S. (1965). Relativistic Quantum Fields (McGraw-Hill, New York). Boehm, F. and Vogel, P. (1987). Physics of Massive Neutrinos (Cambridge University Press, Cambridge). Bogoliubov, N.N. and Shirkov, D.V. (1983). Quantum Fields (Benjamin/Cummings , Menlo Park, Calif.). Bohm, M. and Hollik, W. (1989). Forward-backward asymmetries, in Z Physics at LEP1, ed. Altarelli, G., Kleiss, R. and Verzegnassi, C. (CERN 89-08, Geneva). Bohm, M., Hollik, W. and Speisberger, H. (1986). On the one-loop renormalization of the electroweak standard model, Fort. Phys. 34, 688. Bollini, C.G. and Giambiagi, J.J. (1972). Dimensional renormalization: The number of dimensions as a regularizing parameter, Nuovo Cim. 12B, 20. Braaten, E. (1988). QCD predictions for the decay of the r lepton, Phys. Rev. Lett 60, 1602. Buchalla, G., Buras, A.J. and Harlander, M.K. (1990). The anatomy of e'/e in the standard model, Nucl. Phys. B337, 313. Buchmiiller, W. and Wyler, D. (1986). Effective lagrangian analysis of new interactions and flavor conservation, Nucl. Phys. B268, 621. Burckhardt, H., Jegerlehner, F., Penso, G., and Verzegnassi, C. (1989). Uncertainties in the hadronic contribution to the QED vacuum polarization, Zeit. Phys. C43, 497. Burgers, G. and Jegerlehner, F. (1989). Ar, or the relation between the electroweak couplings and the weak vector boson masses, in Z Physics at LEP I, ed. Altarelli, G., Kleiss, R. and Verzegnassi, C. (CERN 89-08, Geneva). Cabibbo, N. (1963). Unitary symmetry and leptonic decays, Phys. Rev. Lett. 10, 531. Caldwell, D.O. et al. (1987). Limits on neutrinoless 0(3 decay including that with majoron emission, Phys. Rev. Lett. 59, 419. Callaway, D. (1988). Triviality pursuit: Can elementary scalars exist?, Phys. Rept. 167, 241. Callen, C.G., Coleman, S., Wess, J. and Zumino, B. (1969). Structure of phenomenological lagrangians II, Phys. Rev. 177, 2247.
518
References
Callen, C.G., Hornbostel, K. and Klebanov, I. (1988). Baryon masses in the bound state approach to strangeness in the Skyrme model, Phys. Lett. B202, 269. Callen, C.G. and Treiman, S.B. (1966). Equal-time commutators and K meson decay, Phys. Rev. Lett. 16, 153. Carosi, R. et al. (1990). A measurement of the phases of the CP violating amplitudes in K° —> 2n decays and a test of CPT invariance, Phys. Lett. B237, 303. Carruthers, P. (1966). Introduction to Unitary Symmetry (Wiley Interscience, New York). Carter, J.R. (1992). Precision tests of the Standard Model at LEP, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Caswell, W.E. and Lepage, G.P. (1986). Effective lagrangian for bound state problems in QED, QCD and other field theories, Phys. Lett. B167, 437. Celmaster, W. and Gonsalves, R.J. (1979). Analytic calculation of higher order QCD corrections in e + e~ annihilation, Phys. Rev. Lett. 44, 560. Chanowitz, M.S. (1984). Resonances in photon-photon scattering, in Proc. 6th Int. Workshop on Photon-Photon Physics, ed. R.L. Lander (World Scientific, Singapore). Chanowitz, M.S. (1988)* Electroweak symmetry breaking: Unitarity, dynamics and experimental prospects, Ann. Rev. Nucl. Part. Sci. 38, 323. Chanowitz, M.S. and Ellis, J. (1972). Canonical anomalies and broken scale invariance, Phys. Lett 40B, 397. Chanowitz, M.S. and Gaillard, M.K. (1985). The TeV physics of strongly interacting W's and Z's, Nucl Phys. B261, 379. Chanowitz, M.S., Golden, M. and Georgi, H. (1987). Low energy theorems for strongly interacting W's and Z's, Phys. Rev. D36, 1490. Chemtob, M. (1985). The three-flavour Skyrme model applied to the baryon octet and decimet, Nuovo Cim. 89A, 381. Cheng, T.-P. and Dashen, R. (1971). Is 517(2) x SU(2) a better symmetry than S77(3)?, Phys. Rev. Lett. 26, 594. Cheng, T.-P. and Li, L.-F. (1984). Gauge Theory of Elementary Particle Physics (Clarendon Press, Oxford). Cheterkyn, K.G., Kateev, A.L. and Tkachov, S.A. (1979). Higher order corrections to cr(e+e~ —> hadrons) in QCD, Phys. Lett. B85, 277. Chodos, A., Jaffe, R.L., Johnson, K., Thorn, C.B. and Weisskopf, V.F. (1974). New extended model of hadrons, Phys. Rev. D9, 3471. Close, F.E. (1979). An Introduction to Quarks and Partons (Academic Press, London). Coleman, S. (1985). Aspects of Symmetry: Selected Erice Lectures (Cambridge University Press, Cambridge). Coleman, S., and Glashow, S.L. (1964). Departures from the eightfold way, Phys. Rev. B134, 671.
References
519
Coleman, S. and Mandula, J. (1967). All possible symmetries of the S-matrix, Phys. Rev. 159, 1251. Coleman, S., Wess, J. and Zumino, B. (1969). Structure of phenomenological lagrangians I, Phys. Rev. 177, 2239. Coleman, S. and Witten, E. (1980). Chiral symmetry breakdown in large-Nc chromodynamics, Phys. Rev. Lett. 45, 100. Collins, J., Duncan, A. and Joglekar, S. (1977). Trace and dilation anomalies in gauge theories, Phys. Rev. D16, 438. Collins, J., Wilczek, F. and Zee, A. (1978). Low energy manifestations of heavy particles: Application to the neutral current, Phys. Rev. D18, 242. Commins, E.D. and Bucksbaum, P.H. (1983). Weak Interactions of Quarks and Leptons (Cambridge University Press, Cambridge). Consoli, M., Hollik, W. and Jegerlehner, F. (1989). The effect of the top quark on the Mw — Mz interdependence, Phys. Lett. B227, 167. Cornwall, J.M., Levin, D.N. and Tiktopoulos, G. (1974). Derivation of gauge invariance from high-energy unitarity bounds on the S-matrix, Phys. Rev. D 10, 1145. Creutz, M. (1983). Quarks, Gluons and Lattices (Cambridge University Press, Cambridge). Crewther, R. (1972). Nonperturbative evaluation of the anomalies in low energy theorems, Phys. Rev.Lett. 28, 1421. Crewther, R. (1978). Effects of topological charge in gauge theory, Ada Phys. Austriaca(Proc. Suppl.) 19, 47. Cronin, J.A. (1967). Phenomenological model of strong and weak interactions in chiral 17(3), Phys. Rev. 161, 1483. D'Ambrosio, G. and Espriu, D. (1986). Rare decay modes of the K meson in the chiral lagrangian, Phys. Lett. B175, 237. Danilov, M. (1992). Heavy flavour physics, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Dashen, R. (1969). Chiral 5(7(3) x SU(3) as a symmetry of the strong interactions , Phys. Rev. 183, 1245. Davier, M. (1992). Searches for new particles at LEP, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Davis, R., et al. (1989). Report on the chlorine solar neutrino experiment, in Proc. of XIX Int. Conf. on Neutrino Physics, Neutrino 88, ed. J Schneps et al. (World Scientific, Singapore). DeGrand, T., Jaffe, R.L.,Johnson, K. and Kiskis, J. (1975). Masses and other parameters of light hadrons, Phys. Rev. D12, 2060. DeRujula, A., Georgi, H. and Glashow, S.L. (1975). Hadron masses in a gauge theory, Phys. Rev. D12, 147. Deshpande, N.G., Lo, P., Trampetic, J., Eilam, G. and Singer, P. (1987). Prediction of B —• K*i as a test of the Standard Model, Phys. Rev. Lett. 59, 183. Devlin, T.J. and Dickey, J.O. (1979). Weak hadronic decays: K —• 2?r and K -+ 3TT, Rev Mod. Phys. 51, 237.
520
References
DeWitt, B. (1967). Quantum theory of gravity: II, III, Phys. Rev. 162, 1195, 1239. Dine, M. and Sarpinstein, J. (1979). Higher order QCD corrections in e + e~ annihilation, Phys. Rev. Lett. 43, 668. DiVecchia, P. and Veneziano, G. (1980). Chiral dynamics in the large Nc limit, Nucl. Phys. B171, 253. Djoudi, A., Kiihn, J.H. and Zerwas, P.M. (1990). b-jet asymmetries in Z decays, Zeit. Phys. C46, 411. Dobado, A. and Herrero, M.J. (1989). Phenomenological lagrangian approach to the symmetry breaking sector of the standard model, Phys. Lett B228, 495. Donoghue, J.F., Golowich, E. and Holstein, B.R. (1982). The AS = 2 matrix element for K°-K° mixing, Phys. Lett 119B, 412. Donoghue, J.F., Golowich, E. and Holstein, B.R. (1986). Low-energy weak interactions of quarks, Phys. Rep. 131, 319. Donoghue, J.F., He, X.G. and Pakvasa, S. (1986). Hyperon decays and CP nonconservation, Phys. Rev. D34, 833. Donoghue, J.F. and Holstein, B.R. (1989). Pion transitions and models of chiral symmetry, Phys. Rev. D40, 2378. Donoghue, J.F., Holstein, B.R. and Valencia, G. (1987). Survey of present and future tests of CP violation, Int. J. Mod. Phys. A2, 319. Donoghue, J.F. and Johnson, K. (1980). The pion and an improved static bag, Phys. Rev. D21, 1975. Donoghue, J.F. and Nappi, C. (1986). The quark content of the proton, Phys. Lett. B168, 105. Donoghue, J.F. and Ramirez, C. (1990). Symmetry breaking schemes and WW scattering, Phys. Lett B234, 361. Donoghue, J.F., Ramirez, C. and Valencia, G. (1989). Spectrum of QCD and chiral lagrangians of the strong and weak interaction, Phys. Rev. D39, 1947. Donoghue, J.F. and Wyler, D. (1992). Quark masses beyond leading order, Phys. Rev. D45, 892. Dunwoodie W. (1992). Spectroscopy of light hadrons and glueballs, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Earle, E.D. et al. (1986). A D2O Cerenkov detector for solar neutrinos, in 2 nd Conf. on Intersections between Particle and Nuclear Physics, ed. D.F. Geisaman (Amer. Inst. of Physics, New York). Ecker, G., Gasser, J., Pich, A. and de Rafael, E. (1989). The role of resonances in chiral perturbation theory, Nucl Phys. B321, 311. Ecker, G., Pich, A. and de Rafael, E. (1988). Radiative kaon decays and CP violation in chiral perturbation theory, Nucl. Phys. B303, 665. Eden, R.J., Landshoff, P., Olive, D.I. and Polkinghorne, J.C. (1966). The Analytic S-matrix (Cambridge University Press, Cambridge). Edmonds, A.R. (1960). Angular momentum in quantum mechanics (Princeton University Press, Princeton, N.J.).
References
521
Eeg, J.O. and Picek, I. (1987). Short distance double-penguin-like diagram, Nucl. Phys. 119B, 745. Eichten, E. (1988). Heavy quarks on the lattice, Nucl. Phys. (Proc. Suppl.) 4, 170. Eichten, E. and Feinberg, F. (1981). Spin dependent forces in QCD, Phys. Rev. D23, 2724. Eichten, E., Gottfried, K., Kinoshita, T., Lane, K. D. and Yan, T.-M. (1980). Charmonium: Comparison with experiment, Phys. Rev. D21, 203. Eichten, E. and Hill, B. (1990). Static effective field theory: 1/ra corrections, Phys. Lett. B243, 427. Einhorn, M.B. (1991). The Standard Model Higgs Boson (North-Holland, Amsterdam). Elliott, S.R., Hahn, A.A. and Moe, M.K. (1987). Direct evidence for two-neutrino double-beta decay in 82Se, Phys. Rev. Lett. 59, 2020. Ellis J. (1992). Status of the electroweak interaction, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Ellis, J. and Fogli, G.L. (1990). New bounds on vfit and first bounds on MH from precision electroweak data, Phys. Lett. B244, 543. Ellis, J., Gabathuler, E. and Karliner, M. (1989). The OZI rule does not apply to baryons, Phys. Lett. B217, 173. Faddeev, L.D. and Popov, V.N. (1967). Feynman diagrams for the Yang-Mills field, Phys. Lett. 25B, 29. Fanchiotti, S. and Sirlin, A. (1990). Accurate determination of sin2 0w(Mz), Phys. Rev. 41, 319. Farhi, E. and Susskind, L. (1981). Technicolor, Phys. Rep. C74, 277. Fadin, V.S. and Khose, V.A. (1988). Production of a pair of heavy quarks in e + e~ annihilation in the threshold region, Sov. J. Nucl. Part. Phys. 48, 309. Feynman, R.P. and Hibbs, A. (1965). Path Integrals and Quantum Mechanics (McGraw-Hill, New York). Field, R.D. (1989). Applications of Perturbative QCD (Addison-Wesley, Redwood City, Calif.). Flamm, D. and Schoberl, F. (1982). Introduction to the Quark Model of Elementary Particles (Gordon and Breach, New York). Foudas, C. et aZ.(CCFR) (1990). Neutrino production of opposite-sign dimuons at Fermilab Tevatron energies, Phys. Rev. Lett. 64, 1207. Frampton, P.H. (1987). Gauge Field Theories (Benjamin/Cummings, Menlo Park, Calif.). Fritzsch, H. and Gell-Mann, M. (1972). Current algebra: Quarks and what else?, in Proc. XVI Int. Conf on High Energy Physics, ed. J.D. Jackson and A. Roberts (National Accelerator Laboratory, Batavia, 111.). Fujikawa, K. (1979). Path integral measure for gauge invariant field theories, Phys. Rev. Lett. 42, 1195. Fujikawa, K. (1981). Energy momentum tensor in quantum field theory, Phys. Rev. D23, 2262.
522
References
Fukugita, M. (1987). Present status of numerical quantum chromodynamics, in Lattice Gauge Theory using parallel processors, ed. Li, X., Qiu, Z. and Ren, H. (Gordon and Breach, New York). Fulton, R. et al. (CLEO collab.) (1990). Observation of B meson semileptonic decays to noncharmed final states, Phys. Rev. Lett. 64, 16. Fulton, R. et al. (CLEO collab.) (1991). Exclusive and inclusive semileptonic decays of B mesons to D mesons, Phys. Rev. D43, 651. Gaillard, M.K. and Lee, B.W. (1974). Al = 1/2 rule for nonleptonic decays in asymptotically free gauge theories, Phys. Rev. Lett. 33, 108. Gasser, J. (1987). Chiral perturbation theory and effective lagrangians, Nucl. Phys. B279, 65. Gasser, J. and Leutwyler, H. (1984). Chiral perturbation theory to one loop, Ann. Phys. (NY.) 158, 142. Gasser, J. and Leutwyler, H. (1985a). Chiral perturbation theory: Expansions in the mass of the strange quark, Nucl. Phys. B250, 465. Gasser, J. and Leutwyler, H. (19856). Low energy expansion of meson form factors, Nucl. Phys. B250, 517. Gasser, J. and Leutwyler, H. (1985c). rj —» 3TT to one loop, Nucl. Phys. B250, 539. Gasser, J., Leutwyler, H. and Sainio, M.E. (1991). Sigma term update, Phys. Lett. B253, 252. Gasser, J., Sainio, M.E. and Svarc, A. (1988). Nucleons with chiral loops, Nucl. Phys. B307, 779. Gavela, M.B., Le Yaouanc, A., Oliver, L., Pene, O., Raynal, J.-C. and Pham, T.N. (1982). CP violation induced by penguin diagrams and the neutron electric dipole moment, Phys. Lett. 109B, 215. Gavrin, V.N. (1990). Results from SAGE, in Proceedings of the XXV International Conference on High Energy Physics, ed. K.K Phua (World Scientific, Singapore). Gell-Mann, M. and Levy, M. (1960). The axial vector current in beta decay, Nuovo Cim. 16, 705. Gell-Mann, M. and Low, F.E. (1954). Quantum electrodynamics at small distances, Phys.Rev. 95, 1300. Gell-Mann, M., Oakes, R. and Renner, B. (1968). Behavior of current divergences under SU(3) x S77(3), Phys. Rev. 175, 2195. Georgi, H. (1990). An effective field theory for heavy quarks at low energy, Phys. Lett. B240, 447. Georgi, H. (1982). Lie Algebras in Particle Physics: From Isospin to Unified Theories (Benjamin/Cummings, Menlo Park, Calif.). Georgi, H. (1984). Weak Interactions and Modern Particle Theory (Benjamin/Cummings, Menlo Park, Calif.). Georgi, H. and Glashow, S.L. (1974). Unity of all elementary particle forces, Phys. Rev. Lett. 32, 438. Georgi, H., Grinstein, B. and Wise, M.B. (1990). A& semileptonic form factors for mc ^ oo, Phys. Lett. B252, 456.
References
523
Georgi, H., Quinn, H.R. and Weinberg, S. (1974). Hierarchy of interactions in unified gauge theories, Phys. Rev. Lett 33, 451. Gerstein, I., Jackiw, R, Lee, B.W. and Weinberg, S. (1971). Chiral loops, Phys. Rev. D3, 2486. Gilkey, P. (1975). The spectral geometry of a Riemannian manifold, J. Diff. Geom. 10, 601. Gilman, F.J. and Kauffman, R. (1987). 77 — 7/ mixing angle, Phys. Rev. D36, 2761. Gilman, F.J. and Wise, M.B. (1979). The A/ = 1/2 rule and violation of CP in the six-quark model, Phys. Lett B83, 83. Glashow, S.L. (1961). Partial symmetries of weak interactions, Nucl. Phys. 22, 579. Goity, J.L. (1986). The decays Kg —• 77 and K^ —• 77 in the chiral approach, Zeit. Phys. C34, 341. Goldberger, M. and Treiman, S.B. (1958). Conserved currents in the theory of the Fermi interaction, Phys. Rev. 110, 1478. Goldstone, J. (1961). Field theories with superconductor solutions, Nuovo Cim. 19, 154. Goldstone, J., Salam, A. and Weinberg, S. (1962). Broken symmetries, Phys. Rev. 127, 965. Golowich, E. (1987). Addendum to AS = 1 chiral lagrangian revisited, Phys. Rev. D36, 3516. Golowich, E., Haqq, E. and Karl, G. (1983). Are there baryons which contain constituent gluons?, Phys. Rev. D28, 160. Gottfried, K. and Weisskopf, V.F. (1986). Concepts in Particle Physics I,II (Oxford University Press, Oxford). Grigjanis, R., O'Donnell, P.J., Sutherland, M. and Navelet, H. (1990). Regularization in low energy standard model calculations, Phys. Lett. B237, 252. Grinstein, B., Springer, R. and Wise, M.B. (1990). Strong interaction effects in weak radiative B meson decay, Nucl. Phys. B339, 269. Gromes, D. and Stamatescu, I.O. (1976). Linearly rising quark potential and nonstrange baryon spectrum, Nucl. Phys. B112, 213. Gross, D.J. and Wilczek, F. (1973a). Ultraviolet behavior of nonabelian gauge theories, Phys. Rev. Lett. 30, 1343. Gross, D.J. and Wilczek, F. (19736). Asymptotically free gauge theories I, Phys. Rev. D8, 3633. Guadagnini, E. (1984). Baryons as solitons and mass formulae, Nucl. Phys. B236, 35. Guberina, B., Peccei, R.D. and Riickl, R. (1980). Dimensional regularization techniques and their uses in calculating infrared safe weak decay processes, Nucl. Phys. B171, 333. Gunion, J.F., Haber, H.E., Kane, G.L. and Dawson, S. (1990). The Higgs Hunters Guide (Addison-Wesley, Menlo Park, Calif.).
524
References
Haag, R. (1958). Quantum field theories with composite particles and asymptotic conditions, Phys. Rev. 112, 669. Haber, H.E. and Kane, G.L. (1985). The search for supersymmetry: Probing physics beyond the Standard Model, Phys. Rep. 117C, 75. Hagelin, J.S. and Littenberg, L. (1989). Rare kaon decays, Prog. Part. Nucl. Phys. 23, 1. Hagiwara, K., Hikasa, K., Peccei, R.D., and Zeppenfeld, D. (1987). Probing the weak boson sector in e+e" -> W+W~, Nucl. Phys. B282, 253. Halzen F. and Martin, A.D. (1984). Quarks and Leptons: An Introductory Course in Modern Particle Physics (Wiley, New York). Hasenfratz, P. (1987). Nonperturbative methods in quantum field theory, in Proceedings of the XXIII International Conference on High Energy Physics, ed. S.C. Loken (World Scientific, Singapore). Haxton, W.C. (1987). Analytic treatments of matter-enhanced solar-neutrino oscillations, Phys. Rev. D35, 2352. Hebbeker, T. (1992). QCD studies at LEP, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Hey, A.J.G. and Kelly, R.L. (1983). Baryon spectroscopy, Phys. Rep. 96, 71. Hirata, K.S. et al. (1990). Results from 1000 days of real-time directional solar neutrino data, Phys. Rev. Lett. 65, 1297. Hohler, G. (1983). Pion-nucleon scattering, in Landolt Bornstein new series l-9b2, ed. H. Schopper (Springer, Berlin). Holdom, B. and Terning, J. (1990). Large corrections to electroweak parameters in technicolor theories, Phys. Lett. 247, 88. Hollik, W.F.L. (1990). Radiative corrections in the Standard Model, Fort, der Phys. 38, 165. Holstein, B.R. (1989). Weak Interactions in Nuclei (Princeton University Press, Princeton, N.J.). Hughes, R.J. (1981). More comments on asymptotic freedom, Nucl. Phys. B186, 376. Iizuka, J. (1966). A systematics and phenomenology of meson family, Prog. Theo. Phys. Suppl. 37-38, 21. Inami, T. and Lim. C.S. (1981). Effects of superheavy quarks and leptons on low energy weak processes K\ —» /2/x, K+ —• TT+PZ/, and K° — K°, Prog. Theo. Phys. 65, 297. Ioffe, B.L. (1981). Calculation of baryon masses in QCD, Nucl. Phys. B188, 317. Ioffe, B.L. and Shifman, M.A. (1980). The decay ipf -> J/ip + 7r°(r]) and quark masses, Phys. Lett. 95B, 99. Isgur, N. and Karl, G. (1978). P-wave baryons in the quark model, Phys. Rev. D18, 4187. Isgur, N., Scora, D., Grinstein, B. and Wise, M.B. (1989). Semileptonic B and D decays in the quark model, Phys. Rev. D39, 799. Isgur, N. and Wise, M.B. (1989). Weak decays of heavy mesons in the static quark approximation, Phys. Lett. B232, 113.
References
525
Isgur, N. and Wise, M.B. (1990). Weak transition form factors between heavy mesons, Phys. Lett. 237, 527. Itzykson, C. and Zuber, J.-B. (1980). Quantum Field Theory (McGraw-Hill, New York). Izatt, D., DeTar, C. and Stephenson, M. (1982). Spectroscopy of hadrons containing one heavy quark, Nucl Phys. B199, 269. Jackiw, R. and Rebbi, C. (1976). Vacuum periodicity in a Yang-Mills quantum theory, Phys. Rev. Lett 37, 172. Jackson, J.D. (1976). Lectures on the New Particles, in Weak interactions at high energy and the production of new particles, ed. M.C. Zipf (SLAC Summer Inst. 1976, Stanford). Jaffe, R.L. (1977). Perhaps a stable dibaryon, Phys. Rev. Lett. 38, 195. Jaffe, R.L., Johnson, K. and Ryzak, Z. (1986). Qualitative features of the glueball spectrum, Ann. Phys. (NY.) 168, 334. Jarlskog, C. (1985). Commutators of the quark mass matrices in the standard electroweak model and a measure of maximal CP violation, Phys. Rev. Lett. 55, 1039. Jarlskog, C. (1989). Introduction to CP violation, in CP Violation, ed. C. Jarlskog (World Scientific, Singapore). Jegerlehner, F. (1986). Hadronic contributions to electroweak parameter shifts, Zeit. Phys. C32, 425. Johnson, K. (1978). A field theory lagrangian for the MIT bag model, Phys. Lett. 78B, 259. Johnson, K. and Thorn, C.B. (1976). Stringlike solutions of the bag model, Phys. Rev. D13, 1934. Kambor, J., Missimer, J. and Wyler, D. (1990). The chiral loop expansion of the nonleptonic weak interactions of mesons, Nucl. Phys. B346, 17. Kane G.L. (1987). Modern Elementary Particle Physics (Addison-Wesley, Redwood City, Calif.). Kaplan, D.B. and Manohar, A.V. (1986). Current mass ratios of the light quarks, Phys. Rev. Lett. 56, 2004. Kaymakcalan, 6., Rajeev, S. and Schecter, J. (1984). Nonabelian anomaly and vector meson decays, Phys. Rev. D30, 594. Kayser, B., Gibrat-Debu, F. and Perrin, F. (1990). The Physics of Massive Neutrinos (World Scientific, Singapore). Kennedy, D.C. and Lynn, B.W. (1989). Electroweak radiative corrections with an effective lagrangian in four fermion processes, Nucl Phys. B322, 1. Kiesling, C. (1988). Tests of the Standard Theory of Electroweak Interactions (Springer-Verlag, Berlin). Kinoshita, T. (1959). Radiative corrections to n — e decay, Phys. Rev. Lett. 2, 477. Kinoshita, T. and Yennie, D. (1990). High precision tests of QED: An overview, in Quantum Electrodynamics, ed. T. Kinoshita (World Scientific, Singapore). Kobayashi, M. and Maskawa, T. (1973). CP violation in the renormalizable theory of weak interactions, Prog. Theo. Phys. 49, 652.
526
References
Koch, R. (1982). A new determination of the nN sigma term, Z. Phys. C15, 161. Kogut, J. and Susskind, L. (1975). How quark confinement solves the 77 —• 3?r problem, Phys. Rev. D l l , 3594. Korner, J.G. and Schuler, G. (1988). Exclusive semileptonic decays of bottom mesons in the spectator quark model, Z. Phys. C38, 511. Kuhn, J.H. and Zerwas, P.M. (1989). Heavy flavors, in Z Physics at LEP1, ed. Altarelli, G., Kleiss, R. and Verzegnassi, C. (CERN 89-08, Geneva). Kwong, W., Quigg, C. and Rosner, J.L. (1987). Heavy-quark systems, Ann. Rev. Nucl Part. Sci. 37, 325. Langacker, P.W. (1981). Grand unified theories and proton decay, Phys. Rep. C72, 185. Lee, B.W., Quigg, C. and Thacker, H.B. (1977). Weak interactions at very high energies: The role of the Higgs-boson mass, Phys. Rev. D16, 1519. Lee, B.W. and Swift, A.R. (1964). Dynamical basis of the sum rule 2SI = A_ + V3EJ, Phys. Rev. B136, 228. Lee, T.D. (1974). CP nonconservation and spontaneous symmetry breaking, Phys. Rep. C9, 148. Lee, T.D., (1982). Particle Physics and Introduction to Field Theory (Harwood, New York). Leibrandt, G. (1975). Introduction to the technique of dimensional regularization, Rev. Mod. Phys. 47, 849. Leite Lopes, J. (1981). Gauge Field Theory: An Introduction (Pergamon, Oxford). Lepage, G.P. and Thacker, B.A. (1988). Effective lagrangians for simulation of heavy quark systems, Nucl. Phys.(Proc.SuppL) 4, 199. Leutwyler, H. and Roos, M. (1984). Determination of the elements Vus and Vud of the Kobayashi-Maskawa matrix, Z. Phys. C25, 91. Le Yaouanc, A., Oliver, L., Pene, O. and Raynal, J.-C. (1979). Nonleptonic hyperon decays and the baryon spectrum , Nucl. Phys. B149, 321. Le Yaouanc, A., Oliver, L., Pene, O. and Raynal, J.-C. (1985). Quark model of light mesons with dynamically broken chiral symmetry, Phys. Rev. D31, 137. Le Yaouanc, A., Oliver, L., Pene, O. and Raynal, J.-C. (1988). Hadron Transitions in the Quark Model (Gordon and Breach, New York). Li, L.F. and Wolfenstein, L. (1980). Current algebra analysis of K —• 3?r in the six quark Weinberg-Salam model, Phys. Rev. D21, 178. Lichtenberg, D. (1978). Unitary Symmetry and Elementary Particles (Academic Press, New York). Lipkin, H.J. (1984). The theoretical basis and phenomenology of the OZI rule, Nucl. Phys. B244, 147. Liu, K.F.(ed.) (1987). Chiral Solitons (World Scientific, Singapore). Luke, M. (1990). Effects of subleading operators in the heavy quark effective theory, Phys. Lett. B252, 447.
References
527
Marciano, W.J. (1979). Weak mixing angle and grand unified gauge theories, Phys. Rev. D20, 274. Marciano, W.J. and Pagels, H. (1978). Quantum chromodynamics, Phys. Rep. 36C, 137. Marciano, W.J. and Querjeiro, A. (1986). Bound on the W boson electric dipole moment, Phys. Rev. D33, 3449. Marciano, W.J. and Rosner, J. (1990). Atomic parity violation as a probe of new physics, Phys. Rev. Lett 65, 2963. Marciano, W.J., and Sirlin, A. (1980). Radiative corrections to neutrino-induced neutral-current phenomena in the SU{2)L X U(l) theory, Phys. Rev. D22, 2695. Marshak, R.E., Riazuddin and Ryan, C.P. (1969). Theory of Weak Interactions in Particle Physics (Wiley, New York). Martin, A. (1981). A simultaneous fit of bb, cc, ss and cs spectra, Phys. Lett. 100B, 511. Martin, B.R., Morgan, D.M. and Shaw, G. (1976). Pion-Pion Interactions in Particle Physics (Academic Press, New York). Mikheyev, S.P. and Smirnov, A.Yu. (1986). Resonant amplification of neutrino oscillations in matter and solar neutrino spectroscopy , Nuovo Cim. C9, 17. Mohapatra, R.N. and Lai, C.H. (1981). Gauge Theories of Fundamental Interactions (World Scientific, Singapore). Mohapatra, R.N. and Pal, P.B. (1991). Massive Neutrinos in Physics and Astrophysics (World Scientific, Singapore). Mueller, A.H.(ed.) (1989). Perturbative Quantum Chromodynamics (World Scientific, Singapore). Nachtman, O. (1990). Elementary Particle Physics: Concepts and Phenomena (Springer, Berlin). Nambu, Y. and Lurie, D. (1962). Chirality conservation and soft pion production, Phys. Rev. 125, 1429. Narison, S. (1987). Decay constants of the B and D mesons from QCD-duality sum rules, Phys. Lett. 198B, 104. Narison, S. (1989). QCD Spectral Sum Rules (World Scientific, Singapore). Neubert, M. (1991). Model independent extraction of Vcb from semileptonic decays, Phys. Lett. 264B, 455. Noecker, M.C., Masterson, B.P. and Wieman, C.E. (1988). Precision measurement of parity nonconservation in atomic cesium, Phys. Rev. Lett. 61, 310. Okubo, S. (1963). (p meson and unitarity symmetry model, Phys. Lett. 5, 165. Okun, L. (1982). Leptons and Quarks (North-Holland, Amsterdam). Ovrut, B. and Schnitzer, H. (1980). Decoupling theorems for effective field theories, Phys. Rev. D22, 2518. Pak, N.K., and Rossi, P. (1985). Gauged Goldstone boson effective action from direct integration of Bardeen anomaly, Nucl. Phys. B250, 279. Parker, J. (1990). Quark and gluon condensates from available data in QCD sum rules (Univ. Vienna preprint 90-18).
528
References
Paschos, E.A. and Tiirke, U. (1989). Quark mixing and CP violation, Phys. Rep. 178, 145. Peccei, R.D. (1989). The strong CP problem, in CP Violation, ed. C. Jarlskog (World Scientific, Singapore). Perkins, D.H. (1982). Introduction to High Energy Physics (Addison-Wesley, Reading, Mass.). Peskin, M.E. (1990). Theory of precision electroweak measurements, in Physics at the 100-GeV Mass Scale, ed. E.C. Brennan (SLAC Summer Institute 1989, Stanford). Peskin, M.E. and Takeuchi, T. (1990). New constraint on a strongly interacting Higgs sector, Phys. Rev. Lett. 65, 964. Pilkuhn, H. (1967). The Interactions of Hadrons (North-Holland, Amsterdam). Pokorski, S. (1988). Gauge Field Theory (Cambridge University Press, Cambridge). Politzer, H.D. (1973). Reliable perturbative results for strong interactions?, Phys. Rev. Lett. 30, 1346. Politzer, H.D. (1974). Asymptotic freedom: An approach to strong interactions, Phys. Rep. 14C, 274. Quigg, C. (1983). Gauge Theories of the Strong, Weak and Eletromagnetic Interactions (Benjamin/Cummings, Menlo Park, Calif.). Quigg, C. and Rosner, J.L. (1979). Quantum mechanics with applications to quarkonium, Phys. Rep. 56, 167. Raghavan, R.S. and Pakvasa, S. (1988). Probing the nature of the neutrino: The boron solar-neutrino experiment, Phys. Rev. D37, 849. Ramirez, C, Donoghue, J.F. and Burdman, G. (1990). Semileptonic b —• u decay, Phys. Rev. D41, 1496. Ramond, P. (1989). Field Theory: A Modern Primer (Addison-Wesley, Menlo Park, Calif.). Reinders, L.J., Rubenstein, H. and Yazaki, S. (1985). Hadron properties from QCD sum rules, Phys. Rep. 127, 1. Renton, P. (1990). Electroweak Interactions: An Introduction to the Physics of Quarks and Leptons (Cambridge University Press, Cambridge). Richardson, J. L. (1979). The heavy quark potential and the T, J/ty systems, Phys. Lett. 82B, 272. Riggenbach, C, Gasser, J., Donoghue, J.F. and Holstein, B.R. (1991). Chiral symmetry and the large Nc limit in K^ decays, Phys. Rev. 43, 127. Roberts, R.G. (1991). The Structure of the Proton (Cambridge University Press, Cambridge). Roos, M. (1990). SU(3) breaking in hyperon beta decay, Phys. Lett. 246, 179. Rosen, S.P. and Gelb, J.M. (1986). MSW enhancement of oscillations as a possible solution to the solar-neutrino problem, Phys. Rev. D34, 969. Rosenzweig, C, Schechter, J. and Trahern, C.G. (1980). Is the effective lagrangian for QCD a a model, Phys. Rev. D21, 3388. Ross, D.A. and Taylor, J.C. (1973). Renormalization of a unified theory of weak and electromagnetic interactions, Nucl. Phys. B51, 25.
References
529
Rosselet, L. et al. (1977). Experimental study of 30,000 Ke4 decays, Phys. Rev. D15, 574. Ryder, L.H. (1985). Quantum Field Theory (Cambridge University Press, Cambridge). Sakurai, J.J. (1969). Currents and Mesons (University of Chicago Press, Chicago). Salam, A. (1969). Weak and electromagnetic interactions, in Elementary Particle Theory; Nobel Symposium No.8, ed. N. Svartholm (Almqvist and Wiksell, Stockholm). Schnitzer, H.J. (1984). The soft-pion Skyrmion lagrangian and strong CP violation, Phys. Lett. 139B, 217. Schulman, L.S. (1981). Techniques and Applications of Path Integration (Wiley, New York). Schwinger, J. (1951). On gauge invariance and vacuum polarization, Phys. Rev. 82, 664. Schwinger, J. (1954). The theory of quantized fields, Phys. Rev. 93, 615. Sehgal, L. (1990). Rate and spectrum of KL -• TT°77, Phys. Rev. D41, 161. Shabalin, E.P. (1978). Electric dipole moment of the quark in a gauge theory with left-handed currents, Soviet J. Nucl. Phys. 28, 75. Shahbazian, B.A., Sashin, V.A., Kechechyan, A.O. and Martynov, A.S. (1990). The observation of a stable dibaryon, Phys. Lett. B235, 208. Sheck, F. (1983). Leptons, Hadrons and Nuclei (North-Holland, Amsterdam). Shifman, M., Vainshtein, A. and Zakharov, V. (1977). Nonleptonic decays of K mesons and hyperons, JETP 45, 670. Shifman, M., Vainshtein, A. and Zakharov, V. (1979a). QCD and resonance physics: I,II,III, Nucl. Phys. B147, 385,488,519. Shifman, M., Vainshtein, A. and Zakharov, V. (19796). Nonleptonic decays of strange particles, Nucl. Phys. B120, 316. Shifman, M. and Voloshin, M. (1988). On production of D* and D mesons in B meson decay, Sov. J. Nucl. Phys. 47, 511. Shore, CM. (1981). On the Meissner Effect in gauge theories, Ann. Phys. 134, 259. Sikivie, P., Susskind, L., Voloshin, M. and Zakharov, V. (1980). Isospin breaking in technicolor models, Nucl. Phys. B173, 189. Sirlin, A. (1972). Radiative corrections to the Goldberger-Treiman relation: High frequency contributions, Phys. Rev. D5, 436. Sirlin, A. (1980). Radiative corrections in the SU(2)L X C/(l) theory: A simple renormalization framework, Phys. Rev. D22, 971. Skyrme, T.H.R. (1961). A non-linear field theory, Proc. R. Soc. Lon. A260, 127. Stone, S.L. (1990). Session summary - Heavy quark decay, Nucl. Phys. (Proc. Suppl.) B13, 286. Strassler, M.J. and Peskin, M.E. (1991). The heavy top quark threshold: QCD and the Higgs, Phys. Rev. D43, 1500.
530
References
Sutherland, D. (1967). Current algebra and some non-strong meson decays, Nucl. Phys. B2, 433. Taylor, J.C. (1976). Gauge Theories of Weak Interactions (Cambridge University Press, Cambridge). Terent'ev, M.V. (1973). Pion polarizability, virtual Compton effect and the decay TT —> ei/7, Sov. J. Nucl. Phys. 16, 87. 't Hooft, G. (1974). A planar diagram theory of the strong interactions, Nucl. Phys. B72, 461. 't Hooft, G. (1976a). Symmetry breaking through Bell-Jackiw anomalies, Phys. Rev. Lett. 37, 8. 't Hooft, G. (19766). Computation of the quantum effects due to a four-dimensional pseudoparticle, Phys. Rev. D14, 3432. 't Hooft, G. and Veltman, M. (1972). Regularization and renormalization of gauge fields, Nucl. Phys. B44, 189. Treiman, S.B., Jackiw, R., Gross, D. and Witten, E. (1986). Current Algebra and Anomalies (World Scientific, Singapore). Trentadue, L. (1989). Neutrino counting, in Z Physics at LEP1, ed. Altarelli, G., Kleiss, R. and Verzegnassi, C. (CERN 89-08, Geneva). Tsai, Y. (1971). Decay correlations of heavy leptons in e + + e~ —> t+ + £~, Phys. Rev. D4, 2821. Veltman, M. (1967). Theoretical aspects of high energy neutrino interactions, Proc. Roy. Soc. A301, 103. Veltman, M. (1977a). Limit on mass differences in the Weinberg model, Nucl. Phys. B123, 89. Veltman, M. (19776). Second threshold in weak interactions, Acta Phys. Polonica B8, 475. Vetterli, D. et al. (1989). Effects of vacuum polarization in hadron-hadron scattering, Phys. Rev. Lett. 62, 1453. Weinberg, S. (1966a). Current commutator calculation of K14 form factors, Phys. Rev. Lett. 17, 336; 18,1178E. Weinberg, S. (19666). Pion scattering lengths, Phys. Rev. Lett. 17, 616. Weinberg, S. (1967). A model for leptons, Phys. Rev. Lett. 19, 1264. Weinberg, S. (1968). Nonlinear realizations of chiral symmetry, Phys. Rev. 166, 1568. Weinberg, S. (1973). New approach to the renormalization group, Phys. Rev. D8, 3497. Weinberg, S. (1979). Phenomenological lagrangians, Physica A96, 327. Weinstein, J. and Isgur, N. (1983). qqqq system in a potential model, Phys. Rev. D27, 588. Wess, J. and Zumino, B. (1971). Consequences of anomalous Ward identities, Phys. Lett. B37, 95. Wilkinson, D.T., and Marrs, R.E. (1972). Finite size effects in allowed beta decay, Nucl. Inst. Meth. 105, 505. Wilson, K. (1969). Nonlagrangian models of current algebra, Phys. Rev. 179, 1499.
References
531
Winstein, B. (1992). e'/e from E731, in Proc. LP-HEP 91, ed. S. Hegarty, K. Potter and E. Quercigh (World Scientific, Singapore). Wirbel, M., Stech, B. and Baur, M. (1985). Exclusive semileptonic decays of heavy mesons, Z. Phys. C29, 637. Wise, M.B. (1991). New symmetries of the strong interaction (Proceedings of the 1991 Lake Louise Winter Institute and Caltech preprint CALT-68-1721). Witten, E. (1979). Current algebra for the UA(l) 'Goldstone boson', Nucl. Phys. B156, 269. Witten, E. (1983a). Global aspects of current algebra, Nucl. Phys. B223, 422. Witten, E. (19836). Current algebra, baryons and quark confinement, Nucl. Phys. B223, 433. Wolfenstein, L. (1978). Neutrino oscillations in matter, Phys. Rev. D17, 2369. Wolfenstein, L. (1983). Parametrization of the Kobayashi-Maskawa matrix, Phys. Rev. Lett. 51, 1945. Wolfenstein, L. (1986). Present status of CP violation, Ann. Rev. Nuc. Part. Sci. 36, 137. Wyler, D. (1989). Rare decays of heavy quarks, Nucl. Phys. (Proc. Suppl.) 7A, 358. Yang, C.N. (1950). Selection rules for the dematerialization of a particle into two photons, Phys. Rev. 77, 242. Yang, C.N. and Mills, R.L. (1954). Conservation of isotopic spin and isotopic gauge theory, Phys. Rev. 96, 191. Yndurian, F.J. (1983). Quantum Chromodynamics (Springer, New York). Zerwas, P.M. (1988). Hadron masses, chiral symmetry breaking and the interquark potential in lattice QCD, in Proceedings of the XXIV International Conference on High Energy Physics, ed. R. Kotthaus and J.H. Kuhn (Springer-Verlag, Berlin). Zweig, G. (1965). Fractional charged particles and SU(6), in Symmetries in Elementary Particle Physics, ed. A. Zichichi (Academic Press, New York).
Index
abelian gauge symmetry, 14, 25 Ademollo-Gatto theorem, 209-211, 325 Adler-Bardeen theorem, 182, 196 anomalous dimension, 49, 219 anomaly, 18, 75, 76 7T° -> 77, 76, 181, 186, 201 and rjf mass, 202 axial anomaly, 76, 86, 95 Bardeen formula, 86, 200 in large Nc limit, 263 in weak decays, 176, 212 trace anomaly, 88, 96, 331 Wess-Zumino-Witten action, 196 anomaly cancellation, 53, 87, 96 antisymmetric tensor e^VOL^, 505 approximate symmetry, 12 asymptotic freedom, 47 atomic parity violation, 430, 434 axial U(l) transformation, 72, 75, 85 axial-vector coupling g&, 317, 325
background field method, 187, 481, 499 bag model, 285 bag constant, 286 gluons, 289 static cavity, 286 baryon-antibaryon asymmetry, 228 baryons effective lagrangian, 117, 325 matrix elements, 314, 317 quark model, 283, 368 state vectors, 275 SU(3) properties, 277, 314 beta decay hyperon, 339 nucleon, 336 beta function QCD, 47, 90, 219 QED, 35, 36 beyond the Standard Model effective lagrangians, 121, 465 electroweak properties, 463 rare decays, 123, 228, 462 bottomonium, 350 box diagram B meson, 395 D meson, 398 kaon, 235 Breit-Fermi interaction QCD, 365 QED, 126, 129, 354
b-quark B -+ iT*7, 412 B° - B° mixing, 395 b -> 57, 409 bottomonium, 351 CP violation, 400 heavy quark method, 393 lifetimes and branching ratios, 384 B-parameter, 236, 237 532
Index bremsstrahlung muon decay, 139 soft photons, 135 c-quark _ D° - D° mixing, 398 charmonium, 350, 360 D decays, 383 spectrum of charmed hadrons, 370 Cabibbo angle, 64 Cabibbo matrix, 63 canonical field dimension, 26, 511 Casimir invariant, 38 center-of-mass correction in wavepacket method, 318 in bag model, 288 charge commutation rules, 113 charge conjugation C, 6, 149 charge quantization, 25, 54 charge radius baryon, 314 pion, 173, 278 proton, 278 chiral loops, 106 chiral perturbation theory, 164 general O(E2) lagrangian, 158, 189 general O(E4) lagrangian, 165 large Nc limit, 267 low energy constants, 166, 167, 173, 176, 195, 207, 213, 268 pion-pion scattering, 177, 178 radiative transitions, 176 renormalization, 166, 169, 499 renormalization coefficients, 501 underlying physics, 183 chiral symmetry, 71, 115, 157 chiral perturbation theory, 164 explicit breaking, 158 chirality, 4 Coleman-Glashow relation, 368 color, 24 connected Green's functions, 480 conserved vector current hypothesis CVC, 336 constituent gluon, 375
533 conversion factors, 505 Cornell potential, 353 correlator, 302 cosmological constant, 416, 461 Coulomb interaction, 127 covariant derivative SU(2)L x J7(l), 55 abelian, 15, 25 chiral, 117 definition, 14 mixed case, 17 nonabelian, 16, 37 CP violation, 66, 228, 238 B° - B° mixing, 401 K -+ 3TT, 256 KL -> 7T7T, 238 e, 239, 241, 243 e', 239, 241, 243, 246, 249 B decay, 400, 413 nonleptonic hyperon decay, 348 semileptonic decay, 241, 406 CPT CPT theorem, 7 invariance, 241, 257 creation operator, 274 cross sections, 508 current axial-vector, 6, 12, 110, 146, 159, 319, 324, 347 charged weak, 58 electromagnetic, 58 isospin, 10 neutral weak, 58, 430 vector, 5, 11, 110, 146, 147, 319, 324 weak charged, 62 current algebra, 110, 210, 231 custodial symmetry, 426, 465 D mesons lifetimes and branching ratios, 383 Darwin potential, 128 Dashen's theorem, 190 decay constants B meson FB, 396
534 kaon FK, 193 pion Fn, 111, 159 rho meson gp, 147 decay rates, 508 decoupling theorem, 102, 457 deep inelastic lepton scattering, 50, 430, 432 dibaryon, 376, 379 dilation current, 88 dimensional regularization, 29, 107, 139, 167 Dirac algebra, 511 integrals, 512 solid angle, 512 dimensional transmutation, 50 Dirac algebra, 511 Dirac form factor, 132 Dirac matrices, 506 Dirac spinor, 507 discrete symmetries, 6, 71 dispersion relations, 155, 303 dual field strength F, 76 FF, 93, 202 duality, 304 effective lagrangian electroweak interaction, 467 external sources, 114 formalism, 105 general O(E2) chiral lagrangian, 158 general O(E4) chiral lagrangian, 165 neutral current, 430 new physics, 121, 123, 465 QED, 119 sigma model, 97 symmetry breaking, 109 W,Z system, 465 electric dipole moment, 250, 251, 254, 255 electron, 126 electroweak interactions, See Weinberg-Salam-Glashow model electroweak matrix element, 321
Index energy-momentum tensor, 12, 88, 91, 331 equal spacing rule, 368 equivalence theorem, 424, 429 Euler-Heisenberg lagrangian, 120, 124 exotics, 282 external sources, 189 path integrals, 471 effective lagrangians, 113 generating functional, 73 Fermi couplings (GF,G definitions, 59, 139, 162 measurement, 138 Feynman parameterization, 512 Feynman rules electroweak interactions, 449 QCD, 41 QED, 28 field dimension, 511 field strength abelian, 15 nonabelian, 16, 36, 54 Fierz transformation, 137, 507 fine structure constant, 31 form factor baryon, 324 pion, 170 forward-backward asymmetry, 442 ft value, 336 functional differentiation, 471, 478 Furry's theorem, 119 g-factor, 133 G-parity, 149, 324 gamma function, 513 gauge symmetry, 14, 67 gauge-fixing, 41, 487 't Hooft-Feynman gauge, 450 Feynman gauge, 27 Landau gauge, 27 axial gauge, 489 covariant gauges, 489 electroweak, 448 Feynman gauge, 139 unitary gauge, 450
Index Gell-Mann matrices, 37 Gell-Mann-Okubo relation, 159, 190, 194, 327, 368 general relativity, 67 generating functional, 72, 81, 472, 478 generations, 3 generators of group, 16 ghost fields, 41, 44, 448, 490 GIM mechanism, 236 global vs local symmetry, 14 gluon-gluon fusion, 420 gluons, 36 and bag model, 289 gluonia (glueballs), 261, 376 large-iVc limit, 258 Goldberger-Treiman relation, 329 Goldstone bosons, 20, 115 7]' and the large Nc limit, 263 longitudinal components of W,Z, 465 pions and kaons, 21, 157, 188 Goldstone's theorem, 20 Gordon decomposition, 507 grand unification, 438 Grassmann variables, 41, 82, 483 Haag's theorem, 101, 111, 112, 426 hadronic molecule, 379 heat kernel, 89, 198, 495 fermionic operator, 498 heat kernel coefficients, 498 heavy-quark limit, 370 1/m corrections, 374, 395 rare B decay, 413 spectroscopy, 372 symmetry, 392, 394 weak decay, 391 helicity suppression, 160, 208 Higgs boson, 55, 415 couplings, 417, 429 decay, 419 production, 420 two-doublet model, 418 width, 420, 423
535 Higgs mechanism, 24, 56, 415, 465 hybrid, 376, 379 hyperfine interaction QCD, 284, 354, 366, 370 QED, 128, 369 infrared problem, 134 integrating out heavy fields, 102, 103 interpolating field, 279, 376 isospin breaking, 246 isospin symmetry, 9, 19, 21, 70, 157 jacobian, 82, 197 axial anomaly, 85 trace anomaly, 89 Wess-Zumino-Witten action, 197 jets, 51, 358 kaons KL, Ks, 234 CP violation, 238, 242 decay constant F ^ , 193 leptonic decay, 208 mass, 189 mass matrix, 232 mixing, 232 long distance effects, 238 nonleptonic decay, 222 rare decays, 227 semileptonic decay, 209 Kobayashi-Maskawa matrix, 63, 64, 505 Kb, 384, 385 Kb, 389 Kd, 336 Ks, 211, 340 Wolfenstein parameterization, 65, 385 Lamb shift, 134 large Nc limit, 258 rjf as Goldstone boson, 263 chiral lagrangian, 213, 267 double line notation, 259 Skyrme model, 292, 302 spectroscopy, 260
Index
536 strong CP problem, 271 weak decays, 269 lattice-gauge theory, 352, 377 leading log summation kaon decay, 220 muon decay, 138 Lee-Sugawara relation, 342 lepton, 2, 126 lepton number, 3, 123 lepton universality, 145, 208 Lie algebra, 16 longitudinal gauge boson scattering, 426 longitudinal polarization, 444 low energy expansion, 105 LSZ reduction formula, 111 magnetic moment, 322, 323 baryons, 321 electron, 133 mass constituent, 280, 323, 367 diagonalization, 40, 60, 253 gauge boson, 4, 57, 59, 460 heavy quark, 351, 381 Higgs, 416 lepton, 2 Majorana, 122, 154 nonstrange quark, 329 quark, 2, 158, 188, 191, 327 ambiguity in, 193 quark mass ratio, 189, 196, 359, 369 strange quark, 329 massless limit, 4, 95, 158, 160 Meissner effect, 23, 285 mesonic molecule, 376 mesons 7/(1440) (a.k.a. t), 378 7/(549), 188, 190 T/(960), 188, 201, 263,
367
p(770), 146, 174, 180, 184 ai(1260), 148, 185 /2(1770) (a.k.a. 0), 378 quark model, 282 state vector, 275 SU(3) assignments, 276
metric tensor (Minkowskian), 505 Michel parameter, 146 mixing
B° - 5 ° , 395 D°-D°, 398 K° -K°, 232
r) - 7/, 195, 203, 207
7T° - 77, 359 neutrino, 149 quark, 60 muon, 136 decay QED radiative correction, 139 tree-level, 136 muon capture, 338 natural parity, 282 naturalness, 417, 461 neutrinos, 149 mass, 462 mixing, 149 number of generations, 445 oscillation, 150, 151 Noether current, 8, 76, 116, 334 Noether's theorem, 8, 74, 333, 347 nonabelian gauge symmetry, 15, 36 triple gauge coupling, 446 nonleptonic interaction, 213 A/ = 1/2 rule, 220, 222, 223, 270, 342, 345 chiral lagrangian, 343 effective lagrangian, 214, 215, 224 hamiltonian, 214, 216, 221 hyper on decay, 341 large 7VC limit, 269 quark model, 345 radiative decay, 347 nuclear weak processes, 336 nuclei beta decay, 336 muon capture, 338 Okubo-Zweig-Iizuka (OZI) rule, 265 operator product expansion, 304 parity P, 6
Index partial wave amplitude, 178 path integrals external sources, 471 fermions, 483 field theory, 476 functional differentiation, 471, 478 gauge fields, 486 generating functional, 472, 478 harmonic oscillator, 473 perturbation theory, 473 propagator, 468 quadratic forms, 480, 485 wavefunctions, 470 Pauli form factor, 132 Pauli matrices, 506 PCAC, 110 applications, 210, 231, 330, 337 charge commutation rules, 113 soft-pion theorem, 112 penguin diagram, 221, 243, 346 perturbative vacuum, 285 phase shift, 178 physical constants, 505 pions 7T —> ei/, //i/, 160
as Goldstone bosons, 21, 157 decay constant F^, 111, 159 form factor, 170, 184, 187 mass, 158 pion-nucleon interactions, 329, 330 radiative transitions, 174 scattering, 177, 178 planar diagram, 259 Poincare algebra, 22 polarizability, 175, 187 potential model, 280 power law potential, 353 propagators, 508 pseudoscalar axial form factor, 337 QCD sum rules, 302, 312 Borel transformation, 307 correlator, 302 moments, 306, 308 operator product expansion, 304
537 vacuum condensates, 305 quadratic forms, 480, 485 quantum chromodynamics a*, 50 0-term, 40 beta function, 47 Feynman rules, 41 lagrangian, 36 Ward identities, 43 quantum electrodynamics, 24 Feynman rules, 28 gauge symmetry, 25 lagrangian, 24 on-shell renormalization, 31 running coupling constant, 33 Ward identity, 32 quark constituent, 280, 318 mass estimates, 2, 329 mass ratios, 189, 360, 369 quantum numbers, 2 quark mixing, 63 quark model, 311 bag model, 285 potential model, 281 relativistic wavefunction, 278 spatial wavefunctions, 277, 314 quark potential, 352, 380 scalar vs. vector, 355 spin dependence, 354 quarkonium, 350 annihilation decays, 356 hadronic transitions, 359 radiative corrections, 357 spectroscopy, 350 radiative corrections beta decays, 162, 337 electroweak, 436, 448, 456 S,T,E7, 463 Ap,A«,Ar w , 436, 456 €1,2,3, 4 6 3 >
4 6 4
heavy quark decay, 357 muon decay, 139 pion decay, 160, 162, 186
538 weak decay of heavy quarks, 384 rare decay B mesons, 409 kaons, 227 pions, 174 Regge trajectory, 363 Regge-pole, 364 renormalizability, 26 renormalization, 35 minimal subtraction, 31, 32 modified minimal subtraction, 31, 32 on-shell, 31, 32, 129, 435, 451 renormalization constants chiral perturbation theory, 167, 169, 174 QCD, 42, 46 QED, 28, 32 renormalization group, 47, 218, 222, 249
renormalization point, 35 rephasing invariants, 65, 242, 400 of fields, 63, 240 representation independence, 100 rho-parameter, 122, 431, 467 Richardson potential, 353 running coupling constant QCD, 49 QED, 33, 35, 459 scale invariance, 12, 13, 72 scattering length, 178 second class currents, 149 seesaw mechanism, 462 self-energy fermion, 67, 141 photon, 28 W and Z, 452 semileptonic decay b -> uev, 387 heavy quark, 382, 386 heavy-quark method, 393 hyperon, 339 kaon, 209, 231
Index tau, 143, 146 sigma model explicit symmetry breaking, 109 exponential representation, 99, 101 linear, 10 anomaly, 197 effective lagrangian, 97, 104, 116, 184 explicit breaking, 13 spontaneous symmetry breaking, 21 linear representation, 98, 100 nonlinear, 99 square root representation, 99, 100 symmetry breaking, 21 sigma term, 91, 327, 330 Sine-Gordon model, 292 Skyrme model, 292, 319 lagrangian, 293 nucleon mass, 311 quantization, 297 skyrmion, 295 soft-pion theorem, 22, 111 solar neutrinos, 151 soliton, 292, 293 spectator model, 382 spectroscopic notation, 282, 350 spin matrix element, 313 spin-orbit interaction QCD, 354 QED, 127 spontaneous symmetry breaking SU(2)L x 17(1), 56 sigma model, 21 static quark, 372 stickiness, 377, 378 strangeness content of the nucleon, 332 string tension, 352 strong CP violation, 40, 96, 252, 271 and 0, 253 strongly interacting Higgs sector, 423 SU(2) SU(2)L, 51 chiral coefficients, 167
Index chiral SU(2), 167, 191 conjugate representation, 56 isospin, 9 SU(2) L xU(l), See Weinberg-Salam-Glashow model SU(3), 37, 67 fabc,dabc coefficients, 38 chiral SU(3), 115, 167, 188, 191 flavor SU(3), 70, 275 fundamental representation, 37 regular (adjoint) representation, 39 SU(6) classification scheme, 360 supermultiplet, 360, 361 supersymmetry grand unified theories, 438 two Higgs doublets, 418 unnaturalness problem, 418 superweak model, 257 Sutherland-Veltman theorem, 182 symmetry classification of, 18, 69 dynamical breaking, 19, 117, 157, 188 spontaneous breaking, 19, 21, 56, 69 t-quark electroweak radiative corrections, 436, 456 mass estimate, 433 toponium, 390 weak decay, 389 tau lepton, 143 technicolor, 465 tensor interaction, 354 QED, 128 theta-term (0-term), 40, 94, 252, 254 theta-vacuum, 92, 94, 95 time reversal T, 6 invariance, 250 trace anomaly, 88, 96, 331 trace relations, 506 transverse gauge, 357 triviality of \
539 Uehling potential, 129, 130, 134 unitarity, 173, 187, 423, 510 unitarity triangle, 399 unnatural parity, 282 vacuum polarization, 29, 30, 43, 120, 155 gauge boson, 453, 455 vacuum saturation, 226, 244 van Vleck determinant, 491 vector dominance, 185 vertex correction, 130 Z -> 66, 460
QCD, 45 QED, 130 W bosons W+-W- fusion, 420 decays to fermions, 439 mass, 4, 59, 460 self-energies, 453 triple gauge coupling, 446 Ward identities, 32, 43, 77, 80, 201 Watson's theorem, 239, 510 Watson-Sommerfeld transformation, 364 wavepacket method, 289, 291, 316, 318 weak hyper charge, 24, 51, 52, 54, 55 weak isospin, 24, 51, 52, 55 weak mixing angle (Weinberg angle), 57 definition, 435, 437, 452 determination, 432 Weinberg power counting theorem, 107 Weinberg-Salam-Glashow model, 51 Feynman rules, 449 lagrangian, 54 neutral current phenomenology, 430 quark mixing, 60 radiative corrections, 139, 451 symmetry breaking, 56 W,Z decays, 439 weak mixing angle, 57
540 Wess-Zumino-Witten anomaly action, 196, 200, 301 Wilson coefficients, 216, 222 winding number, 92 Wu-Yang convention, 240 Yang-Mills theory, 15 Yukawa coupling, 56 Z boson, 57
Index decays into fermions, 441 forward-backward asymmetry, 442 mass, 4, 57 neutral current interaction, 431 self energy, 455 vertex correction, 460 W3-B mixing, 57 zero-point energy, 90 zeta-function regularization, 496