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!
") 2 1 (as we shall see, at saturation the occupation numbers are parametrically of order l / a s ) , so their mutual interactions can be treated in the classical approximation, i.e., by solving the classical field equations 6S/6A1* = 0, or, more explicitly: {DvF^)a{x) where Dv = dv- , ebe <<^>ebe
/ *V
InW—. V^
(121)
This allows to relate via wc ~ {n^L) w c ln JQ the characteristic gluon energies (85) and (112) which set the energy scale in the multiple soft and single hard scattering limits. For realistic values [/i > AQCD and Q < E say], In JQ
232
A. Kovner and U. A.
Wiedemann
an arbitrary number of n gluons which carry away a total energy AE:9 „... u n=0 Li=l
aw
Then we discuss how this probability can be used to calculate the medium modification of hadronic observables. 5.1.
Properties
of Quenching
Weights
In general, the quenching weight (122) has a discrete and a continuous part, 51 P(AE)
= p0 6(AE) + p(AE).
(123)
The discrete weight po emerges as a consequence of a finite mean free path. It determines the probability that no additional gluon is emitted due to in-medium scattering and hence no medium-induced energy loss occurs. In order to determine the discrete and continuous part of (123), it is convenient to rewrite eq. (122) as a Laplace transformation 9 P(A£?) = [ ±,p(v)e»*B,
(124)
™—p[-jf*^ ('--")
(125)
Here, the contour C runs along the imaginary axis with Rev = 0. For the further discussion, it is useful to treat the medium-induced gluon energy distribution w ^ in eq. (69) explicitly as the medium modification of a "vacuum" distribution 52 dj(tot) d/ (vac) dI oj—-. =u>— +u>—. (126) aw auj aw From the Laplace transform (124), one finds the total probability /•OO
p(tot) ( A £ )
/•OO
/ Jo
dE P
( A £ _ E) p( v a c ) (E).
(127) Jo This probability p( t o t ) (AE) is normalized to unity and it is positive definite. In contrast, the medium-induced modification of this probability, P(AE), is a generalized probability. It can take negative values for some range in AE, as long as its normalization is unity, =
/.OO
dE P{E) = po + / Jo
dEp(E) = 1.
(128)
Gluon Radiation
and Parton Energy Loss
233
We now discuss separately the properties of the discrete contribution po and the continuous one p(E). A CPU-inexpensive Fortran routine 52 is available for the calculation of these quenching weights.
1.2
-—
1 1
*-'"* "**» ^^s. \\
—
gluon jet quark jet
0.8 0.6 0.4 0.2
!.
Y>^ 10
10
10
10
R Fig. 7. The discrete part po of the quenching weight calculated in the multiple soft scattering limit as a function of R. Figure taken from 5 2 .
5.1.1. Discrete part of the quenching weight The discrete part of the quenching weight is the n = 0 term of eq. (122). It can be expressed in terms of the total gluon multiplicity, po = lim V{v) = exp[-N(uj
= 0)] ,
(129)
where the multiplicity N(w) of gluons with energy larger than u emerges by partially integrating the exponent of (125), N(w)
, dl{w') du) du'
(130)
234
A. Kovner and U. A.
Wiedemann
For the limiting case of infinite in-medium path length, the total multiplicity N(u) diverges and the discrete part vanishes. In general, however, Po is finite. A typical dependence of po on model parameters is shown in Fig. 7 for the radiation spectrum calculated in the multiple soft scattering limit. A qualitatively similar behavior is found in the opacity expansion. Remarkably, po c a n exceed unity for some parameter range, since the medium modification w ^ to the radiation spectrum (126) can be negative. The value po > 1 then compensates a predominantly negative continuous part p(AE) and satisfies the normalization (128). This indicates a phase space region at very small transverse momentum, into which less gluons are emitted in the medium than in the vacuum. This effect is more pronounced for gluons than for quarks. 5.1.2. Continuous part of the quenching weight The continuous part p(AE) of the probability distribution (123) can be calculated numerically from the Mellin transform (125). To facilitate the numerical calculation, one subtracts from P(v) the discrete contribution po which dominates the large-f behavior. Fig. 8 shows the continuous part of the quenching weight, calculated in the multiple soft scattering limit. In the opacity expansion, it looks qualitatively similar. As expected from the normalization condition (128), the continuous part p(AE) shows predominantly negative contributions for the parameter range for which the discrete weight po exceeds unity. With increasing density of scattering centers (i.e. increasing R = \qL3) the probability of loosing a significant energy fraction AE increases. The energy loss is larger for gluons which have a stronger coupling to the medium. This broadens the width of p{AE) for the gluonic case. In the multiple soft scattering approximation, an analytic estimate for the quenching weight can be obtained 9 in the limit R —» oo from the small-w approximation LJ^J OC -7= PB*DPMS(C) = \^<*P
na
" —
e J
-
2a?C?
wherea=^^u,c.
(131)
This reproduces roughly 51 the shape of the probability distribution for large system size, but it has an unphysical large e-tail with infinite first moment J dee PemAsi6)- -^ n alternative analytic approach 2 aims at fitting a two-parameter log-normal distribution to the numerical result for P(AE).
Gluon Radiation
0.5
and Parton Energy Loss
- J5
R=100 R=1400 R=10 4
- i K '
— •!* '• \ —• • • *
• R=
I '. . » *, 1
0 o
. 0.8 _ *•
0.6 —.
Q.
\
—!""*••
0.4
•A
0.75
0.25
0.5
0.75
•-
0
-0.4
0.5
— **\
0.2
-0.2
•
0.25
R=1 R=10 R=20 R=50
Ld
S
oo
- — - ^ ^ • t . , . . , . i i . . !"l
-r,,,
\
235
\, ' I •
i
,
.
1 i
0.25
^..'*'" i
,
,
!
i
0.5
i
i
i
1 i
i
0.75
i
•
1 i
1
AE/0)c
AE/coc
Fig. 8. The continuous part of the quenching weight (123), calculated in the multiple soft scattering limit for a hard quark (upper row) or hard gluon (lower row). Figure taken from 5 2 .
5.2. Quenching
factors
for hadronic
spectra
Assume that a hard parton looses an additional energy fraction AE while escaping the collision region. The medium-dependence of the corresponding inclusive transverse momentum spectra can be characterized in terms of the
236
A. Kovner and U. A.
quenching factor Q
Wiedemann
9
da™(p±)/dpl
J
V
^JdAEP^{^TAE)n-
da^ip^/dpl
W
Here, the last line is obtained by assuming a power law fall-off of the p±spectrum. The effective power n depends in general on p±. It is n ~ 7 for the pj_-range relevant for RHIC. Alternatively, instead of the quenching factor (132), the medium modification of hadronic transverse momentum spectra is often characterized by a shift factor S(p±), da™d(p±) dpi
~
_da™(p±+S(p±)) dpi '
{U6)
which is related to the shift S(p±) by Q(p±)
exp|-^-.5(px)}.
(134)
Most importantly, since the hadronic spectrum shows a strong power law decrease, what matters for the suppression is not the average energy loss (A.E) but the least energy loss with which a hard parton is likely to get away. One concludes that S(p±) < (AE) and depends on transverse momentum 9 . Fig. 9 shows a calculation of the quenching factor (132) in the multiple soft scattering limit. A qualitatively similar result is obtained in the opacity expansion. In general, quenching weights increase monotonically with p±_ since the medium-induced gluon radiation is independent of the total projectile energy for sufficiently high energies. At very low transverse momenta, the calculation based on (69) is not reliable and the interpretation of the medium modification of hadronic spectra in nucleus-nucleus collisions will require additional input (e.g. modifications due to the Cronin effect). Fig. 9 suggests, however, that hadronic spectra at transverse momenta p± > 10 GeV, can be suppressed significantly due to partonic final state rescattering. To quantify the sensitivity of the calculation to the low momentum region, Baier et al. 9 introduced a sharp cut-off on the R —» oo gluon energy distribution which was varied between wcut = 0 and wCut = 500 MeV. However, phase space constraints (i.e. finite R) deplete the gluon radiation spectrum in the soft region, see Fig. 2. As seen in Fig. 9, this decreases
Gluon Radiation
and Parton Energy Loss
237
q*=1 GeV 2 /fm n=4, L=2fm
100 0 p,(GeV)
n=4, L=5fm
100 Pt(GeV)
Fig. 9. The quenching factor (132) calculated in the multiple soft scattering limit. Upper row: calculation in the R —» co-limit but with a variable sharp cut-off on the infrared part of the gluon energy distribution. Lower row: the same calculation is insensitive to infrared contributions if the finite kinematic constraint R = wcL < oo is included. Figure taken from 5 2 .
significantly the sensitivity of quenching factors to the uncontrolled infrared properties of the radiation spectrum.
238
A. Kovner and U. A.
5.3. Medium-modified
Wiedemann
fragmentation
functions
For an alternative calculation of the medium modification of hadronic spectra, one may determine the dependence of fragmentation functions on partonic energy loss. In general, hadronic cross sections are calculated by convoluting the parton distributions of the incoming projectiles with the product dah(z,Q2) of a perturbatively calculable partonic cross section aq and the fragmentation function Dh/q{x,Q2) of the produced parton, dah(z,Q2) = ( ^ f ) dyDh/q{x,Q2)dx. Here, x = Eh/Eq, y = Eq/Q and z = Eh/Q denote fractions between the virtuality of the hard process Q, and the energies of the produced parton and resulting hadron. If the produced parton loses with probability P(e) an additional fraction e = ^ of its energy due to medium-induced radiation, then the hadronic cross section is given in terms of the medium-modified fragmentation function 5 4 , 2 8 D$?\*,Q2) = [
deP(e) ^
D^^Q2).
The hadronized remnants of the medium-induced soft radiation are neglected in the definition of (135). However, these remnants are expected to be soft, and their inclusion would thus amount to an additional contribution to D{™d)(x,Q2) for x > 0.1 say. Fig. 10 shows a calculation of the parton fragmentation functions Ar/g(£>Q 2 ) from (135) using the quenching weights of Fig. 8 and the LO KKP 31 parametrization of Dh/q{x,Q2). For this calculation, the virtuality Q of Dh/q{x,Q2) is identified with the (transverse) initial energy Eq of the parton. This is justified since Eq and Q are of the same order, and Dh/q(x>Q2) n a s a weak logarithmic Q-dependence while medium-induced effects change as a function of e = ^j- ~ ^ ( i ) - For a collision region expanding according to Bjorken scaling, the transport coefficient can be related to the initial gluon rapidity density 4 ' 2 7 ,
That's what is done in Fig. 10. Interestingly, eq. (136) indicates how partonic energy loss changes with the particle multiplicity in nucleus-nucleus collisions. This allows to extrapolate parton energy loss effects from RHIC to LHC energies 51 . In principle, the medium modified fragmentation function should be convoluted with the hard partonic cross section and parton distribution
(135)
Gluon Radiation
and Parton Energy Loss
239
KKP, no medium dN'/dy=350 dN'/dy=1700 dN'/dy=3500,
1
x
Fig. 10. The LO KKP 3 1 fragmentation function u —» 7r for no medium and the mediummodified fragmentation functions for different gluon rapidity densities (see eq. (136)) and L = 7 fm. Figure taken from 5 1 .
functions in order to determine the medium modified hadronic spectrum. For illustration, however, one may exploit that hadronic cross sections weigh ^h/g (x> Q 2 ) b y t n e partonic cross section daq/dp\ ~ l/p^s'p±' and thus 20 effectively test xn^'p^D(^d\x,Q2) . The value n = 6 characterizes 20 the power law for typical values at RHIC (y/s = 200 GeV and p±, ~ 10 GeV). Thus, the position of the maximum :r max of x6D^ (x,Q2) corresponds to the most likely energy fraction xm&yLEq of the leading hadron. And the suppression around its maximum translates into a corresponding relative suppression of this contribution to the high-px hadronic spectrum at p± ~ xmiLXEq. In general, the suppression of hadronic spectra extracted in this way is in rough agreement 52 with calculations of the quenching factor (132).
240
A. Kovner and U. A.
Wiedemann
6. Appendix A: Eikonal calculations in the target light cone gauge. In section 2 of this review we have used the light cone gauge A~ = 0. In this gauge the gluon and quark distributions of the projectile wave function are simply expressible in terms of the gluon and quark number operators and we will therefore refer to it as the projectile light cone gauge (PLCG). The standard light cone gauge used in DIS calculations on the other hand is A+ = 0, which facilitates simple expressions of the target distribution functions. In this appendix we show how the spectrum of emitted gluons is obtained in this standard light cone gauge, which we will refer to as the target light cone gauge (TLCG). Recall, that in PLCG the target is described as an ensemble of gluon fields with dominant component A+. The eikonal S-matrix for the propagation of a charged parton through the target is given my the Wilson line eq.(3) Wfc)
= Vexp{i
fdz-TaAt(xi,z-)}
(137)
In the TLCG the A+ component of the vector potential vanishes. Instead the chromoelectric fields in the target are given in terms of the transverse components Ai. The A1 are obtained from A+ by the gauge transformation from PLCG to TLGT 3 7 A(xi,a:_) = iVi(xi,x-)diV(xi,x-)
(138)
where V(xi,x-)=Vexp{i
f
dz-.TaAZ(xitz-)}
(139)
J—oo
Note that the TLGT condition A+ = 0 does not fix the gauge unambiguously, but only up to residual gauge transformation which does not depend on x~. The choice of the lower limit of the z- integration in eq.(139) is equivalent to fixing this residual gauge freedom by imposing the condition djAj(xi,X- —> —oo) = 0 37 . With this choice we have V ( x i , x _ - > - o o ) = l,
V(x.i,x-
-> oo) = W(xj).
(140)
Eq.(138) defines the vector potential Ai as two dimensional pure gauge, diA] - djAf - fabcA^A^ = 0. Moreover, at X- -> +oo the vector potential is genuinely (and not just two dimensionally) pure gauge.
Gluon Radiation
and Parton Energy Loss
241
Let us now consider scattering of a projectile with the wave function eq.(l) on the target described by ensemble of fields At of the form eq.(138). Since the A+ component of the vector potential vanishes, in eikonal approximation the wave function of the projectile does not change while it propagates through the target. The outgoing wave function therefore is equal to the incoming one *out=
Yl
^({ai,xJ)|{ai,Xi}}.
(141)
{"i,x»}
This however does not mean that no scattering takes place. To calculate the scattering amplitude one has to project the outgoing wave function into the Hilbert space orthogonal to the wave function of the freely propagating system far "to the right" of the target, that is at X- —> +co. In the PLCG the target gauge field vanishes at both x~ —* — oo and X- —» +00 and the freely propagating wave functions are identical "to the left" and "to the right" of the target. However in TLGT this is not the case. At X- —> +00 the target vector potential does not vanish, but is instead a pure gauge, eq. (138). Therefore the freely propagating wave function at x- —> +00 is not identical to that at X- —> —00, but is rather its gauge transform with the gauge transformation generated by the Wilson loop eq. (137). In particular the fields "to the right" of the target, A are related to the fields "to the left" of the target, a by A = W^aW + iWidW.
(142)
Thus the free Fock basis " to the right" of the target is related to the free Fock basis " to the left" of the target by \{a,xi})R
= Wl0(Xi)\{0,xi})L
.
(143)
The outgoing wave function eq.(141) is given in the basis |{a, Xj})x,. On the other hand all observables at late time after scattering, including the number of emitted gluons must be calculated with respect to the basis |{a, Xj})j{. It is thus convenient to rewrite ^0ut using eq.(143) as *out=
Y,
V>({a<,Xi})n^(Xi)aiftl{ft.Xi}>«-
{aj.Xj}
(144)
i
However the same projectile freely propagating to the right of the target would have the wave function Vfree
=
JZ {«i,Xi}
^ ( { a i , Xj}) | { a i , X i } ) f l .
(145)
242
A. Kovner and U. A.
Wiedemann
Thus we see that the outgoing wave function indeed differs from the freely propagating one, and thus the scattering is nontrivial and gluons are emitted in the final state. It is a somewhat curious feature of the TLCG that the nontrivial scattering amplitude appears entirely due to rotation of the free particle basis between early and late times Nevertheless it is obvious that the results of the calculation in this gauge are identical to those in PLCG as they should be. In fact from this point on all calculations are identical to those presented in section 2, as the interesting part of the outgoing wave function is given by eq.(5) with ^ / r e e of eq.(145) substituted for ^in and the "free" gluons are defined as states in the R Hilbert space, |{a,x})/i. 7. Appendix B: Path integral formalism for the photon radiation spectrum In section 3.3, we discussed how to derive the non-abelian gluon radiation spectrum from the non-abelian Furry approximation (44). In this appendix, we present in more detail the derivation of the abelian analogue from the abelian Furry wave function (50). The QED photon radiation cross section in terms of Furry wave functions reads 57 dh ° d{\nx)dpdk
Mfi=
Qem
(2TT)4
\Mfi\\
(146)
fd4x^-\x,p2)a-ee-e^eikx^+(x,Pl)
.
(147)
Here, e _ £ ' 2 ' is the adiabatic switching off of the interaction term at large distances, discussed below eq. (64). To simplify (147), we perform the following steps: (1) Rotation of coordinate system Choose the momenta p\ and p2 of the incoming and outgoing electron in the frame in which the longitudinal axis is taken along the photon: —1 Pi = — k , X
p2=P
1— x k,
(148)
X
(2) z-dependent phase: The z-dependent phases of the Furry wave function (50) combine in the radiation amplitude (147) to an inverse photon
Gluon Radiation
and Parton Energy Loss
243
formation length xm2 k
" = ^-^- =W^)E[-
(149)
(3) Simplifying the spinor structure: The spinor structure Tr = y/T-rxu*{-p2)D*2a-tD1u{vl)
(150)
in the amplitude (147) can be simplified on the cross section level. The spin- and helicity-averaged combination Tr T*, take the simple form fP f;, = [4 - Ax + 2x2} - ^ • J j - + 1m\x2 .
(151)
(4) in-medium average: The cross section (146) contains products of the Green's functions (58). These are averaged over the distribution of scattering centers in the medium, see (52) for the non-abelian case. These averages can be written in terms of the dipole cross section (14): /eXplijdt[U(r(t),t)-U{T>(t),()]
= expj-iy"d£n(0
(152)
For the products of Green's functions, this leads to
x Jvrbexp
1^ f
i2 - J* E ( £ , p , + xrb) 1 ,
(153)
where /i = p (1 — a;) x and P / ( 0 = P 2 ^ + P ! ^ .
(154)
After inserting the Furry wave function (50) into the radiation amplitude (147). The radiation probability (|M/j| 2 ) in terms of averages of Green's
244
A. Kovner and U. A.
Wiedemann
functions reads {\Mfi\2) =
TTJ Re / dri dr dpdx2 dr[ dr' dp' dr'2
2
dz
dz'
-ip 2 -(r 2 -r' 2 )+ipi-(r 1 -ri)-ig(z'-z) -e(|z| + |z'|)
xe x (G(r 2 , z+; p, z'\p2) G*(r'2, z+; r', z'\p2)}
xf_ r f;, (G(p, z'; r, z\p2) G V > z'; p!, z\Pl)) x(G(r,z;n,2_|pi)G*(p/,z;ri,z_|pi)>.
(155)
Equation (155) is rearranged such that the photon emission in the amplitude Ma occurs prior to emission in the complex conjugate amplitude M^. The opposite contribution is accounted for by taking twice the real part. Averages (...) of pairs of Green's functions effectively compare the paths r(£) and r'(£) of the electron in amplitude and complex conjugate amplitude, see eq. (153). Using this average, one arrives at the final expression 63 ' 57 d3a
_ aem
2 2
d(lna;) dk ~ (2TT) E (1 - a;)2
xRej
dzj z_
2
dz'expl[-i2iiXm^E^z'-z)-e(\z\
+ |z'j)j
z
x / dri exp {-ik • r i } exp < - / d£ n(£) a(x r i ) 4-4x
+ 2x2 d Ax2 dr\
dv2
mix21 2
JC(z',r2 =0;z,ri\n)
, (156)
where the abelian path-integral takes the form IC(z',rc(z')\z,rc(z)\n
= Ei(l - x)x)
= / Vrc exp
(157)
The expression (156) provides a path-integral formulation of the LandauPomeranchuk-Migdal radiation spectrum 39 ' 40,45 ' 30 . Graphically, as can be seen from eq. (155), this radiation spectrum can be represented as
Gluon Radiation
and Parton Energy Loss
245
path in Mfl
difference between the two paths path in Mjj
\ Thus, for propagation from Z- = —oo up to the longitudinal photon emission point z in Alfi, the electron propagates with initial energy E\ in both amplitude and complex amplitude. Then it propagates with momentum E2 = Ei(l — x) in MB. but with E\ in M^ up to z', the photon emission point in the complex amplitude, etc. The average (153) ensures that if propagation in MR and M^ occur with the same energy, then the relative distance between the paths r(£) and r'(£) does not change: physically, the radiation cross section counts only those changes in r(£) — r'(£) which amount to phase shifts between amplitude and complex conjugate amplitude. Those are accumulated between the photon emission points z and z' in MR and M^, respectively.
Acknowledgments Both authors thank the Institute for Nuclear Theory at the University of Washington for its hospitality and the Department of Energy for partial support during the completion of this work.
References 1. F. Arleo, Phys. Lett. B 532 (2002) 231 [arXiv:hep-ph/0201066]. 2. F. Arleo, JHEP 0211, 044 (2002) [arXiv:hep-ph/0210104]. 3. R. Baier, Y. L. Dokshitzer, A. H. Mueller, S. Peigne and D. Schiff, Nucl. Phys. B 483 (1997) 291 [arXiv:hep-ph/9607355]. 4. R. Baier, Y. L. Dokshitzer, A. H. Mueller, S. Peigne and D. Schiff, Nucl. Phys. B 484 (1997) 265 [arXiv:hep-ph/9608322]. 5. R. Baier, Y. L. Dokshitzer, A. H. Mueller and D. Schiff, Phys. Rev. C 58 (1998) 1706 [arXiv:hep-ph/9803473]. 6. R. Baier, Y. L. Dokshitzer, A. H. Mueller and D. Schiff, Nucl. Phys. B 531 (1998) 403 [arXiv:hep-ph/9804212].
246 A. Kovner and U. A. Wiedemann 7. R. Baier, Y. L. Dokshitzer, A. H. Mueller and D. Schiff, Phys. Rev. C 60 (1999) 064902 [arXiv:hep-ph/9907267]. 8. R. Baier, Y. L. Dokshitzer, A. H. Mueller and D. Schiff, Phys. Rev. C 64 (2001) 057902 [arXiv:hep-ph/0105062]. 9. R. Baier, Y. L. Dokshitzer, A. H. Mueller and D. Schiff, JHEP 0109 (2001) 033 [arXiv:hep-ph/0106347]. 10. R. Baier, arXiv:hep-ph/0209038. 11. I. Balitsky, Nucl. Phys. B 463 (1996) 99 [arXiv:hep-ph/9509348]. 12. J. D. Bjorken, Fermilab-Pub-82/59-THY, Batavia (1982); Erratum, unpublished. 13. G. T. Bodwin, S. J. Brodsky and G. P. Lepage, Phys. Rev. D 39 (1989) 3287. 14. S. J. Brodsky, A. Hebecker and E. Quack, Phys. Rev. D 55 (1997) 2584 [arXiv:hep-ph/9609384]. 15. W. Buchmuller, T. Gehrmann and A. Hebecker, Nucl. Phys. B 537 (1999) 477 [arXiv:hep-ph/9808454]. 16. A. Del Fabbro and D. Treleani, Phys. Rev. D 66 (2002) 074012 [arXiv:hepph/0207311]. 17. M. Djordjevic and M. Gyulassy, arXiv:nucl-th/0302069. 18. Y. L. Dokshitzer and D. E. Kharzeev, Phys. Lett. B 519 (2001) 199 [arXiv:hep-ph/0106202]. 19. K. J. Eskola, K. Kajantie, P. V. Ruuskanen and K. Tuominen, Nucl. Phys. B 570 (2000) 379 20. K. J. Eskola and H. Honkanen, Nucl. Phys. A 713 (2003) 167 [arXiv:hepph/0205048]. 21. L. V. Gribov, E. M. Levin and M. G. Ryskin, Phys. Rept. 100 (1983) 1. 22. J. F. Gunion and G. Bertsch, Phys. Rev. D 25 (1982) 746. 23. X. F. Guo and X. N. Wang, Phys. Rev. Lett. 85 (2000) 3591 24. M. Gyulassy and X. N. Wang, Nucl. Phys. B 420 (1994) 583. 25. M. Gyulassy, P. Levai and I. Vitev, Phys. Rev. Lett. 85 (2000) 5535 [arXiv:nucl-th/0005032]. 26. M. Gyulassy, P. Levai and I. Vitev, Nucl. Phys. B 594 (2001) 371 [arXiv:nucl-th/0006010]. 27. M. Gyulassy, I. Vitev and X. N. Wang, Phys. Rev. Lett. 86 (2001) 2537 [arXiv:nucl-th/0012092]. 28. M. Gyulassy, P. Levai and I. Vitev, Phys. Lett. B 538 (2002) 282 [arXiv:nucl-th/0112071]. 29. J. Jalilian-Marian, A. Kovner and H. Weigert, Phys. Rev. D 59 (1999) 014015 [arXiv:hep-ph/9709432]. 30. S. Klein, Rev. Mod. Phys. 71 (1999) 1501 [arXiv:hep-ph/9802442]. 31. B. A. Kniehl, G. Kramer and B. Potter, Nucl. Phys. B 582 (2000) 514 [arXiv:hep-ph/0010289]. 32. B. Kopeliovich, arXiv:hep-ph/9609385. 33. B. Z. Kopeliovich, A. V. Tarasov and A. Schafer, Phys. Rev. C 59 (1999)
Gluon Radiation and Parton Energy Loss 247 1609 [arXiv:hep-ph/9808378]. 34. Y. V. Kovchegov and A. H. Mueller, Nucl. Phys. B 529 (1998) 451 [arXiv:hep-ph/9802440]. 35. Y. V. Kovchegov, Phys. Rev. D 60 (1999) 034008 [arXiv:hep-ph/9901281]. 36. Y. V. Kovchegov, Nucl. Phys. A 692 (2001) 557 [arXiv:hep-ph/0011252]. 37. A. Kovner, J. G. Milhano and H. Weigert, Phys. Rev. D 62 (2000) 114005 [arXiv:hep-ph/0004014]. 38. A. Kovner and U. A. Wiedemann, Phys. Rev. D 64 (2001) 114002 [arXiv:hep-ph/0106240]. 39. L. D. Landau and I. Pomeranchuk, Dokl. Akad. Nauk Ser. Fiz. 92 (1953) 535. 40. L. D. Landau and I. Pomeranchuk, Dokl. Akad. Nauk Ser. Fiz. 92 (1953) 735. 41. Z. W. Lin and R. Vogt, Nucl. Phys. B 544 (1999) 339 [arXivihepph/9808214]. 42. I. P. Lokhtin and A. M. Snigirev, Eur. Phys. J. C 21 (2001) 155 [arXiv:hepph/0105244]. 43. M. Luo, J. w. Qiu and G. Sterman, Phys. Rev. D 49 (1994) 4493. 44. M. Luo, J. W. Qiu and G. Sterman, Phys. Rev. D 50 (1994) 1951. 45. A. B. Migdal, Phys. Rev. 103 (1956) 1811. 46. M. Gyulassy, I. Vitev, X. N. Wang and B. W. Zhang, contribution to this volume, arXiv:nucl-th/0302077. 47. A. H. Mueller and J. W. Qiu, Nucl. Phys. B 268 (1986) 427. 48. N. N. Nikolaev and B. G. Zakharov, Z. Phys. C 49 (1991) 607. 49. N. Paver and D. Treleani, Nuovo Cim. A 70 (1982) 215. 50. J. W. Qiu and G. Sterman, arXiv:hep-ph/0111002. 51. C. A. Salgado and U. A. Wiedemann, Phys. Rev. Lett. 89 (2002) 092303 [arXiv:hep-ph/0204221]. 52. C. A. Salgado and U. A. Wiedemann, arXiv:hep-ph/0302184. 53. X. N. Wang, M. Gyulassy and M. Plumer, Phys. Rev. D 51 (1995) 3436 [arXiv:hep-ph/9408344]. 54. X. N. Wang, Z. Huang and I. Sarcevic, Phys. Rev. Lett. 77 (1996) 231 [arXiv:hep-ph/9605213]. 55. X. N. Wang, Phys. Rev. C 63 (2001) 054902 [arXiv:nucl-th/0009019]. 56. E. Wang and X. N. Wang, arXiv:hep-ph/0202105. 57. U. A. Wiedemann and M. Gyulassy, Nucl. Phys. B 560 (1999) 345 [arXiv:hep-ph/9906257]. 58. U. A. Wiedemann, Nucl. Phys. B 582 (2000) 409 [arXiv:hep-ph/0003021]. 59. U. A. Wiedemann, Nucl. Phys. B 588 (2000) 303 [arXiv:hep-ph/0005129]. 60. U. A. Wiedemann, Nucl. Phys. A 690 (2001) 731 [arXiv:hep-ph/000824lj. 61. B. G. Zakharov, JETP Lett. 63 (1996) 952 [arXiv:hep-ph/9607440]. 62. B. G. Zakharov, JETP Lett. 65 (1997) 615, [arXiv:hep-ph/9704255]. 63. B. G. Zakharov, Phys. Atom. Nucl. 62 (1999) 1008 [Yad. Fiz. 62 (1999) 1075] [arXiv:hep-ph/9805271].
248
A. Kovner and U. A.
Wiedemann
64. B. G. Zakharov, Phys. Atom. Nucl. 61 (1998) 838 [Yad. Fiz. 61 (1998) 924] [arXiv:hep-ph/9807540]. 65. B. G. Zakharov, J E T P Lett. 70 (1999) 176 [arXiv:hep-ph/9906536]. 66. B. G. Zakharov, JETP Lett. 73 (2001) 49 [Pisma Zh. Eksp. Teor. Fiz. 73 (2001) 55] [arXiv:hep-ph/0012360].
T H E COLOR GLASS C O N D E N S A T E A N D HIGH E N E R G Y SCATTERING IN QCD
Edmond Iancu Service de Physique Theorique, CEA Saclay, 91191 Gif-sur-Yvette cedex, France
Raju Venugopalan Physics Department, Brookhaven National Laboratory and RIKEN-BNL Research Center, Upton, NY 11973, USA
At very high energies or small values of Bjorken x, the density of partons, per unit transverse area, in hadronic wavefunctions becomes very large leading to a saturation of partonic distributions. When the scale corresponding to the density per unit transverse area, the saturation scale Qs, becomes large (Qs 3> AQC£>), the coupling constant becomes weak (as(Qs) *C 1) which suggests that the high energy limit of QCD may be studied using weak coupling techniques. This simple idea can be formalized in an effective theory, the Color Glass Condensate (CGC), which describes the behavior of the small x components of the hadronic wavefunction in QCD. The Green functions of the theory satisfy Wilsonian renormalization group equations which reduce to the standard linear QCD evolution equations in the limit of low parton densities. The effective theory has a rich structure that has been explored using analytical and numerical techniques. The CGC can be applied to study a wide range of high energy scattering experiments from Deep Inelastic Scattering at HERA and the proposed Electron Ion Collider (EIC) to proton/deuterium-nucleus and nucleus-nucleus experiments at the RHIC and LHC colliders.
249
250 E. Iancu and R. Venugopalan
Contents 1
Outstanding Phenomenological Questions in High Energy QCD 251 1.1 Introduction 251 1.2 Light cone kinematics and dynamics 252 1.3 High energy behavior of total cross-sections 253 1.4 Multi-particle production in QCD 254 1.5 Deep inelastic scattering 257 1.6 Nucleus-nucleus and proton-nucleus collisions 261 1.7 Universality of high energy scattering 264 2 The Effective Theory for the Color Glass Condensate 265 2.1 The hadron wavefunction at small x 265 2.2 The McLerran-Venugopalan model for a large nucleus 266 2.3 The color glass 269 2.4 The classical color field 273 2.5 The gluon distribution 275 2.6 Gluon saturation in a large nucleus 277 2.7 Dipole-hadron scattering at high energy 281 3 The Quantum Evolution of the Color Glass Condensate 285 3.1 The BFKL evolution and its small-x problem 285 3.2 Non-linear evolution for the CGC 290 3.3 The Balitsky-Kovchegov equation 295 3.4 Saturation momentum and geometric scaling 297 3.5 Gluon saturation and perturbative color neutrality 302 3.6 A Gaussian effective theory 309 4 Deep Inelastic Scattering and the CGC 311 4.1 Structure functions in the color glass condensate 312 4.2 The Golec-Biernat-Wusthoff model 314 4.3 Geometric scaling in DIS 316 4.4 The Proissart bound for dipole scattering 318 4.5 Saturation and shadowing in deep inelastic scattering 324 4.6 Probing the CGC with an electron ion collider 326 5 Melting the CGC in Nucleus-Nucleus and Proton-Nucleus Collisions . . 331 5.1 Classical picture of nuclear collisions 332 5.2 Numerical gluodynamics of nuclear collisions 335 5.3 Melting the color glass condensate at RHIC 348 5.4 Equilibration and the quark-gluon plasma 352 5.5 Proton-nucleus and peripheral nucleus-nucleus collisions at RHIC and LHC 355 References 358
The Color Glass Condensate
and High Energy Scattering
in QCD
251
1. Outstanding Phenomenological Questions in High Energy QCD 1.1.
Introduction
The advent of a new generation of high energy collider experiments, beginning with HERA and the Tevatron in the early 1990's to RHIC and LHC in the new millenium (and others in the planning stages) have launched a new era in the study of the strong interactions. Questions which have been around since the early days of the strong interactions, such as the behavior of cross-sections at high energies, the universality of hadronic interactions at high energies, the nature of multi-particle production and the possibility of creating thermalized states of strongly interacting matter have acquired fresh vigor. For instance, it is often believed that little could be learned about the high energy limit of QCD since the physics is assumed to be entirely non-perturbative. On the other hand, we have learned from HERA that parton densities are large at high energies or equivalently, at small values of Bjorken x. The large densities of small x, or "wee" partons suggest, as we will discuss at length later, that semi-hard scales may be present which allow one to describe the physics of this regime using weak coupling techniques. QCD at high energies can thus be described as a many-body theory of partons which are weakly coupled albeit non-perturbative due to the large number of partons. We will call this system a Color Glass Condensate (CGC), for the following reasons: • "Color", since the gluons are colored. • "Glass" because of the strong analogy of the system to actual glasses. A glass is a disordered system which evolves very slowly relative to natural time scales: it is like a solid on short time scales and like a liquid on much longer time scales. Similarly, the partons of interest are disordered and evolve in longitudinal momentum in a manner analogous to a glass. • "Condensate" because it contains a very high density of massless gluons whose momenta are peaked about some characteristic momentum. Increasing the energy forces the gluons to occupy higher momentum states (due to repulsive interactions) causing the coupling to become weaker. The gluon density saturates at a value of order l/as 3> 1, corresponding to a multiparticle Bose condensate state. We will argue in the following that the Color Glass Condensate is the effective theory describing high energy scattering in QCD. We will outline the rich structure of the theory and discuss how it provides insight into outstanding conceptual issues in QCD at asymptotically high energies. The
252
E. Iancu and R.
Venugopalan
theory can be applied to study phenomena at a number of existing and upcoming high energy collider facilities. We will discuss applications of the CGC to study the initial conditions and equilibration in heavy ion collisions and to describe heavy ion phenomenology. We will discuss applications to Deep Inelastic Scattering, to the current experiments at HERA but also for future experiments being discussed at HERA and at Brookhaven. In addition, we will discuss applications of the CGC to proton/deuteron-nucleus scattering experiments planned at RHIC and at the LHC and to peripheral nucleus-nucleus scattering experiments at RHIC. 1.2. Light cone kinematics
and
dynamics
The appropriate kinematics to discuss high energy scattering are light-cone (LC) coordinates. Let z be the longitudinal axis of the collision. For an arbitrary 4-vector v^ = {v°,vl,v2,v3) (v3 = vz, etc.), we define its LC coordinates as v+ = ^(v°+v3), v2
v- = ^(v°-v3), v2
v± = (v\v2).
(1.1)
In particular, we shall refer to x+ = (t + z)/y/2 as the LC "time", and to x~ = (t — z)/y/2 as the LC "longitudinal coordinate". The invariant scalar product of two four-vectors reads: p • x = p~x+ + p+x~ — p± • x± ,
(1.2) +
which suggests that p~ should be interpreted as the LC energy, and p as the (LC) longitudinal momentum. In particular, since p^ = (l/y/2)(E±pz) with E = (m 2 + p 2 ) 1 / 2 , the light cone dispersion relation takes the form P
lpj+m2 _ lmj ~ 2 p+ ~ 2 p+ '
[1 d)
-
2
where the transverse mass m± is denned as m\ = p\ + m . The momentum space rapidity is further simply as: 1 , p+ 1 , 2p+2 yy = - I n — = - l n - * - 5 - . 2 p~ 2 m\
, ,x v(1.4) '
These definitions are useful, among other reasons, because of their simple properties under longitudinal Lorentz boosts: p+ —> np+, p~ —* (l/n)p~, where K is a constant. Under boosts, the rapidity is just shifted by a constant: y —» y + K. The utility of light cone kinematics is not merely that of a convenient coordinate transformation. The Hamiltonian dynamics of quantum field
The Color Glass Condensate
and High Energy Scattering
in QCD
253
theories quantized on the light cone have several remarkable features as was first pointed out by Dirac 1. Firstly, the LC Hamiltonian p~ (which is the generator of translations in the light cone time x+) can be written in the form p~ = p^ + V, where p$ is the free Hamiltonian, corresponding to a complete set of non-interacting Fock eigenstates, and V is the light cone potential. Second, the LC vacuum is trivial, namely, the vacuum state is an eigenstate of both the free and the full Hamiltonian. As a consequence of these properties, multi-parton Fock states can be constructed as eigenstates of the QCD Hamiltonian. Thus, in LC quantization (and in the light-cone gauge A+ = 0) the quark-parton picture of QCD becomes manifest. Finally, we note that the apparent non-relativistic structure of the light cone Hamiltonian suggested by the dispersion relation in Eq. (1.3) is not accidental but is a consequence of an exact isomorphism between the Galilean subgroup of the Poincare group and the symmetry group of two dimensional quantum mechanics 2 . Thus in LC quantization, the Rayleigh-Schrodinger perturbation theory with off-shell energy denominators can be used instead of the more usual Feynman rules. For a more detailed discussion of the light cone formalism and its application to high energy scattering, we direct the reader to Ref. 3.
1.3. High energy behavior of total
cross-sections
We now return to the outstanding phenomenological questions we mentioned in the introduction. Clearly, computing total cross-sections as E —> oo is one of the great unsolved problems of QCD. Unlike processes which are computed in perturbation theory, it is not required that any energy transfer become large as the total collision energy E —> oo. Computing a total cross-section for hadronic scattering therefore appears to be an intrinsically non-perturbative procedure. In the 60's and early 70's, Regge theory was extensively developed in an attempt to understand the total cross-section. The results of these analyses were, to our mind, inconclusive, and at any rate, certainly cannot be claimed to be understood from first principles in QCD. On the basis of very general arguments invoking unitarity, analyticity and crossing, Froissart has shown that the total cross-section for the strong interactions grows at most as fast as In E as E —> oo 4 ' 5 . Several questions arise in this regard. Is the coefficient of In2 E universal for all hadronic processes? Can this coefficient be computed from first principles in QCD? How do we understand the saturation of the unitarity limit dynamically in QCD?
254
E. Iancu and R.
Venugopalan
Or is the Proissart bound an intrinsically non-perturbative phenomenon? 1.4. Multi-particle
production
in
QCD
Can we compute N(E), the total multiplicity of produced particles as a function of energy in QCD? By this we mean not only the multiplicity of particles in jets (which is fairly well understood in perturbative QCD) but also the total number of particles, at least, for semi-hard momenta. Consider the collision of two identical hadrons in the center of mass frame, as shown in Fig. 1. The colliding hadrons are ultrarelativistic and therefore Lorentz contracted in the direction of their motion. Furthermore, we assume that the typical transverse momenta of the produced particles is large compared to AQCD- We know from experiments that the leading particles (valence partons) typically lose only some finite fraction of their momenta in the collision. The produced particles, which are mostly mesons, are produced in the "wake" of the nuclei as they pass through each other. In light cone
large p Fig. 1.
small p
large p
A hadron-hadron collision. The produced particles are shown as circles.
coordinates, the right moving particle ("the projectile") has a 4-momentum Pi = {pt>Pi'°±.) w i t n Pi - y^Pz and pi ~ M2/2y/2pz (since pz » M and m±_ = M, with M = the projectile mass). Similarly, for the left moving hadron ("the target"), we havepJ = pf andp^T = p*• The invariant energy squared is s = (pi +P2) 2 = 2pi • pi ~ IpXpZ — kp\, and coincides, at it should, with the total energy squared {E\ + E2)2 in the center of mass frame.
The Color Glass Condensate
and High Energy Scattering
in QCD
255
Consider a pion produced in this collision and which is moving in the positive z direction. For such a pion, we define the longitudinal momentum fraction, or Feynman x, as: v+ x = qj:
(right mover),
(1.5)
Pi which implies m±/y/2pf
< x < 1. The rapidity of the pion is then
1 v+ 1 2p+2 1 M Y= - l n % = - l n ^ - = y p r o j - m - + l n — , 2 PK 2 mj_ x m±
(1.6)
(yProj = ln(\/2Pi"/M) ~ ln(^/M)), and lies in the range 0 < y < yproj + ln(M/mj_). For a left moving pion (j>% < 0), we use similar definitions where p+ and p~ are exchanged. This gives a symmetric range for y, as in Fig. 2. All the pions are produced in a distribution of rapidities within this range. In Fig. 2, dN/dy is the number of produced particles (say, pions) per unit rapidity. The leading particles are shown in the solid line and are clustered around the projectile and target rapidities. For example, in a heavy ion collision, this is where the nucleons would be. In the dashed line, the distribution of produced mesons is shown.. dN dy
" yproj
Fig. 2.
y
pr°j
The rapidity distribution of particles produced in a hadronic collision.
Several theoretical issues arise in multiparticle production. Can we compute dN/dy? Or even dN/dy at y = 0 ("central rapidity")? How does the average transverse momentum of produced particles {p±) behave with energy? What is the ratio of produced strange/nonstrange mesons, and corresponding ratios of charm, top, bottom etc at y = 0 as the center of mass
256
E. Iancu and R.
Venugopalan
energy approaches infinity? Does multiparticle production as s —» oo at y = 0 become simple, understandable and computable? Note that y = 0 corresponds to particles with pz = 0 or p+ — mj_/\/2, for which x = mj_/(V2pf) = m^/^/s is small, x
dN dy
Fig. 3. Feynman scaling of rapidity distributions. The two different lines correspond to rapidity distributions at different energies.
The Color Glass Condensate
and High Energy Scattering
in QCD
257
of freedom, and the renormalization group is generated by integrating out degrees of freedom at relatively large x to generate these sources. With this understanding, one may be able to compute the number and distribution of particles produced at central rapidities. 1.5. Deep Inelastic
Scattering
In the previous section, we discussed hadron-hadron collisions in which large numbers of particles are produced. Here we will discuss Deep Inelastic Scattering (DIS) of a lepton scattering off a hadronic target 6 . Fewer particles are produced in DIS, so this provides a relatively clean environment to study QCD at high energies. In Fig. 4 is shown the cartoon of a DIS experiment. electron
^^^^
photon
quark
s'
,/ S /
Fig. 4.
) hadron
Deep inelastic scattering of an electron on a hadron.
To describe quark distributions, it is convenient to work in a reference frame where the hadron has a large light-cone longitudinal momentum P+ » M ("infinite momentum frame"). In this frame, one can describe the hadron as a collection of constituents ("partons"), which are nearly on-shell excitations carrying some fraction x of the total longitudinal momentum P+. (The correct mathematical formulation of this picture involves the light cone quantization mentioned previously: the hadron can then be expressed in a "quark-parton" basis.) Thus, the longitudinal momentum of a parton is p+ = xP+, with 0 < x < 1. For the struck quark in Fig. 4, this x variable ("Feynman x") is equal (modulo target mass corrections) to the empirical Bjorken variable XBJ, which is defined in a frame independent way as XBJ = Q2/2P • q. In this definition, Q2 = —q^q^, with q^ the (space-like) 4-momentum of the ex-
258
E. Iancu and R.
Venugopalan
changed photon. The Bjorken variable scales like XBJ ~ Q2/s, with s = the invariant energy squared. Thus, in deep inelastic scattering at high energy (large s at fixed Q2) we measure the quark distributions dNquark/dx at small x (x
(1.7)
between the quark and the hadron increases 7 . In Fig. 5, the ZEUS data
xG(x,Q2)
io' 4 Fig. 5.
io"3
iff2
10'
The Zeus data for the gluon structure functions.
The Color Glass Condensate
and High Energy Scattering
in QCD
259
Low Energy
High Energy
Fig. 6.
Saturation of gluons in a hadron. A view of a hadron head on as x decreases.
for the gluon distribution are plotted for Q2 = 5 GeV 2 , 20 GeV2 and 200 GeV 2 7 . The gluon distribution is the number of gluons per unit rapidity in the hadron wavefunction, xG(x, Q2) — dNgiuons /dy . Experimentally, it is extracted from the data for the quark structure functions, by analyzing the dependence of the latter upon the resolution Q2 of the probe. The growth seen in Fig. 5 appears to be more rapid than T or x 2 . Perturbative considerations of the high energy limit in QCD by Lipatov and colleagues lead to an evolution equation commonly called the BFKL equation 8 which suggests that distributions may grow as an exponential in T 8 ' 9 . Alternatively, the double logarithmic DGLAP evolution equation 10 predicts a less rapid growth, like an exponential in %/T. Both of these evolution equations would predict asymptotically a growth of the distributions which would exceed the Froissart unitarity bound discussed previously. How do we understand in QCD the problem of the rapid rise of gluon distributions at small x? Consider Fig. 6, where we view the hadron head on. The constituents are the valence quarks, gluons and sea quarks, all shown as colored circles. As we add more and more constituents, the hadron becomes more and more densely populated. If one attempts to resolve these constituents with an elementary probe, as in DIS, then, at sufficiently small
260
E. lancu and R.
Venugopalan
x (for a given transverse resolution), the density of the constituents becomes so large that one cannot neglect their mutual interactions any longer. One expects such interactions to give "shadowing" by which we imply a decrease of the scattering cross-section relative to what is expected from incoherent independent scattering. More precisely, we shall see later that, as a effect of these interactions, the parton distribution functions at fixed Q2 saturate, in the sense of showing only a slow, logarithmic, increase with i/x 1 1 ' 1 2 ' 1 3 , 1 4 , 1 5 , 1 6 . (See also Refs. 17, 18, 19, 20 for recent reviews and more references.) For a given Q2, this saturation occurs if x is low enough, lower than some critical value x s (Q 2 ). Converserly, for given x, saturation occurs for transverse momenta below some critical value Q 2 (x), defined as Q2(X) = a
s
N
c
^ ,
(1.8)
where dN/dy is the gluon distribution at y = Yhadron — ln(l/x). Only gluons are included, since, at high energy, the gluon density grows much faster than the quark density, and is the driving force towards saturation. This explains why in the following we shall focus primarily on the gluons. In Eq. (1.8), TTR2 is the hadron area in the impact parameter space (or transverse plane). This is well defined provided the wavelength of the probe is small compared to R, which we assume throughout. Finally, asNc is the color charge squared of a single gluon. Thus, the "saturation scale" (1.8) has the meaning of the average color charge squared of the gluons per unit transverse area per unit rapidity. Since the gluon distribution increases rapidly with the energy, as the HERA data suggests, so does the saturation scale. For high enough energy, or small enough x,
Q f r ) » AQCD , 2
(1-9)
and as(Q ) -C 1. This suggests that weak coupling techniques can be used to study the high energy regime in QCD. However, weak coupling does not necessarily mean that the physics is perturbative. There are many examples of nonperturbative phenomena at weak coupling. An example is instantons in electroweak theory, which lead to the violation of baryon number. Another example is the atomic physics of highly charged nuclei. Although the electromagnetic coupling constant is very weak, aem
The Color Glass Condensate
and High Energy Scattering
in QCD
261
to compute the non-perturbative properties of high Z atoms. Yet another example is QCD at high temperature T » A Q C D ; this is a weakly coupled quark-gluon plasma, but exhibits nonperturbative phenomena on large distances r ~^> 1/T due to the collective behaviour of many quanta 21 . Similarly, the small-x gluons with transverse momenta Q2 < Q2(x) make a high density system, in which the interaction probability a x n (where a ~ a3/Q2 is the typical parton cross-section and n is the gluon density, n = ^/nR2) is of order one n - 1 2 ' 1 3 (cf. Eq. (1.8)). That is, although the coupling is small, as(Q2)
1.6. Nucleus-Nucleus
and Proton-Nucleus
Collisions
In Fig. 7, we plot a cartoon of the space-time evolution of a heavy ion collision 24 . Imagine we have two Lorentz contracted nuclei approaching one another at the speed of light. We choose coordinates such as the collision takes place at z = t = 0, or x+ = x~ = 0. Since the two nuclei are well localized in the longitudinal direction, they can be thought of as sitting at z ~ t (x~ = 0) for the right mover, respectively at z ~ — t (or x+ = 0) for the left mover. To analyze this problem for t > 0, namely, after the collision takes place, it is convenient to introduce a time variable which is Lorentz invariant under longitudinal boosts a r — y/t2 — z2 (the "proper time") and a space-time rapidity variable 1 , ft + z\
1 , x+
,
"This should not be confused with the rapidity variable introduced previously, in Eq. (1.7), and which will not appear in this subsection.
262
E. Iancu and R.
Fig. 7.
Venugopalan
A space-time figure for ultrarelativistic heavy ion collisions.
For free streaming particles with velocity vz (z = vzt = j^-t), the spacetime rapidity equals the momentum space rapidity (1.4): r] = y. We shall see later that this identification remains approximately true also for the offshell quantum fluctuations (the partons) with relatively large longitudinal momenta. At high energies, in the central rapidity region, particle distributions vary slowly and it should be a good approximation to take them to be rapidity invariant and therefore also independent of rj. Therefore distributions are the same on the lines of constant proper time r, which are shown in Fig. 7. An outstanding problem is to formulate the initial conditions for a heavy ion collision and to study the subsequent evolution of the produced partons. Can one argue from first principles that the partonic matter will thermalize into a quark-gluon plasma? There are two separate classes of problems one has to understand for the initial conditions. Firstly, the two nuclei which are colliding are coherent quantum mechanical wavepackets. Therefore, for some early time, the degrees of freedom must be quantum mechanical. This means, in particular, AzApz
> 1,
(1.11)
which particularly constrains the small-x gluons, which are delocalized over
The Color Glass Condensate
and High Energy Scattering
in QCD
263
large longitudinal distances Az ~ l / p z , and thus overlap with each other. Such degrees of freedom cannot be described by semi-classical transport theory for particles. (Classical particles are characterized by a distribution function f(p, x, t), which is a simultaneous function of momenta and coordinates.) However, fortuitously, quantum coherent states can be described as classical fields because they have large occupation numbers ~ \/as :» 1. Heisenberg commutators between particle creation and annihilation operators become negligible in this limit: [ak, 4 ] = 1 < a\ak = Nk .
(1.12)
Classical field theory is also the appropriate language to describe another important feature of the initial conditions, namely the classical charge coherence. At very early times, we have a tremendously large number of particles packed into a longitudinal size scale of less than a fermi due to the Lorentz contraction of the nuclei. We know that such particles cannot interact incoherently. For example, if we measure the field due to two opposite charges (a dipole) at a distance scale r large compared to their separation, the field falls off as 1/r2, not 1/r. On the other hand, in parton cascade models, interactions are taken into account by cross-sections which involve matrix elements squared leaving no room for classical charge coherence. These models should therefore not be applied at very early times. As an effective theory at small x, the Color Glass Condensate can be applied to study the initial stages of a heavy ion collision 29>30>33. The only scales in the problem are the saturation scale Qs and the transverse size of the system R. On a time scale r ~ 1/QS, the initial energy and number distributions of gluons can be computed by solving classical field equations. At later times r 3> 1/Q a , the system becomes dilute and the classical field approximation breaks down, but one in terms of transport equations becomes valid. There is presumably an overlap region where the two descriptions can be both correct 3 5 . At early times, the distribution of gluons in momentum space is primarily transverse. As the system becomes dilute, the gluons begin to scatter "off the transverse plane". Baier, Mueller, Schiff and Son 34 have argued that 2—> 3 processes are those which lead to the most efficient thermalization. Whether on not the initial non-equilibrium gluon distributions thermalize to form a quark gluon plasma at the energies of interest is of outstanding phenomenological interest at RHIC and LHC energies. It is also of interest to understand how much of the RHIC data can be understood purely from initial state effects 36 > 37 ' 38 ' 39 a s opposed to the final state effects which are important for thermalization. Proton (or
264
E. Iancu and R.
Venugopalan
deuterium) — nucleus collisions should be very helpful in helping to isolate initial state effects from final state rescattering effects. In a proton-nucleus collision, one does not expect final state rescattering to dominate the measured particle spectrum. Initial state effects, on the other hand, should be especially important in the proton fragmentation region and have been computed recently within the CGC framework 27 - 40 ' 41 . 42 . The upcoming experiments at RHIC on deuterium-gold collisions will hopefully clarify the role of initial/final state effects in high energy scattering. 1.7. Universality
of High Energy
Scattering
In pion production, it is observed that, with the exception of globally conserved quantities like the energy and the total charge, the rapidity correlations are of short range. If the theory is local in rapidity, the only dimensionful parameter which can determine the physics at a given rapidity is Qg(x). For an approximately scale invariant theory such as QCD, a typical transverse momentum of a constituent will therefore be of order Q2S. If Q2 » l/R2, where R is the radius of the hadron, the finite size of the hadron becomes irrelevant. Thus at small enough x, all hadrons become the same — specific properties of the hadrons (like their size or atomic number A) enter only via the saturation scale Q2(x, A). Hence there should be some equivalence between nuclei and protons: when their Q2 values are the same, their physics should be the same. Eq. (1.8) suggests the following empirical parametrization of the saturation momentum: 41/3
Q2s(x,A) ~ - J -
(1.13)
where the value A ~ 0.2 — 0.3 seems to be preferred by both the data 7 ' 4 3 and the most recent theoretical calculations 4 4 . There should thus be the following correspondence: • RHIC with nuclei ~ HERA with protons • LHC with nuclei ~ HERA with nuclei Estimates of the saturation scale for nuclei at RHIC energies give Qs ~ 1 - 2 GeV, and at LHC 2 - 3 GeV. This further suggests that for relatively simple processes like deep inelastic scattering (at least), the observables should be universal functions of the ratio between the transferred momentum Q2 and the saturation scale Q2(x, A). This feature, called "geometric scaling", is rather well satisfied by the proton structure functions measured at HERA, for all values of
The Color Glass Condensate
and High Energy Scattering
in QCD
265
Bjorken x smaller than 0.01 and in a broad region of Q2 (between 0.045 and 450 GeV 2 ) 45 . Interestingly, the observed scaling extends up to relatively large values of Q2, well above the saturation scale Q2(x), which suggests that the phenomenon of gluon saturation at Q2 < <5^(x) has a rather strong influence also on the physics at much larger Q2 4 6 . The previous considerations suggest how to reconcile unitarity with the growth of the gluon distribution function at small x. The point is that the smaller x is the larger Qs(x) is and thus and the typical partons are smaller. Therefore, when decreasing x, although we are increasing the number of gluons, we do it by adding in more gluons of smaller and smaller size. A probe of transverse size resolution Ax± ~ \/Q will not see partons smaller than this resolution size. Thus when Q < Qs, newly created partons will not contribute to the cross-section. 2. The Effective Theory for the Color Glass Condensate 2.1.
The hadron wavefunction
at small x
As discussed in Sect. 1.2, a relativistic quantum field theory quantized on the lightcone has a simple structure in terms of the bare quanta or "partons" of the theory. For instance, in QCD, the eigenstates of the light-cone Hamiltonian PQCD can be expressed as a linear superposition of the eigenstates of the non-interacting part of the Hamiltonian PQCD,O = PQCD ~ VQCD, where VQCD is the LC potential. The proton wavefunction for instance can be written in this parton basis as \ip >= ci\qqq > +c2\qqqg > + ••• + cn\qqqgggg...qqgg > H
. (2.1)
It is a priori not obvious what the advantage of such a decomposition is. The magic of the parton model is that, at high energies, it is apparent from the structure of VQCD that the interactions of the partons with each other are time dilated. The very complicated picture of the scattering of the proton off an external potential can be replaced with the simple picture of individual partons scattering off the external potential. (In the eikonal approximation, each of these partons acquires a simple phase in the scattering 47 .) At lower energies, the Fock states involving large numbers of partons are not very important but at higher energies they are increasingly important, and involve predominantly gluons. Since the total LC momentum of the proton P+ is divided among a large number of partons, a typical parton will carry a momentum k+ « P+, i.e., only a small fraction x = k+/P+ of the proton total momentum. Understanding the physics of QCD at high energies or
266
E. Iancu and R.
Venugopalan
small x thus requires that we understand the properties of the n-gluon components of the hadron's LC wavefunction. A very interesting approach to computing the properties of the n-gluon component of the hadronic wavefunction is Mueller's color dipole approach for heavy quarkonia 48 . This is valid in the limit where the number of colors Nc is large, so the gluons can be effectively replaced by qq pairs ("color dipoles"). It will be shown later that this approach and the CGC formalism give identical results for the evolution of distributions in x, in this large-A^c limit. The CGC approach that we shall follow here is to construct a coarse grained effective theory for the small-x component of the hadron LC wavefunction. We shall first consider a large nucleus, for which this construction is most intuitive. Then, we shall argue that, at sufficiently large energy, or small enough x, a similar theory can be constructed for any hadron, via weak coupling calculations in QCD.
2.2. The McLerran-Venugopalan
model for a large
nucleus
Consider a nucleus in the infinite momentum frame (IMF) with momentum P+ —> oo. We will assume that the nucleus is of nearly infinite transverse extent with a uniform nuclear matter distribution. As we will discuss later, the model can be extended to include realistic nuclear density profiles. In the IMF, partons which carry large fractions of the nuclear momentum ("valence" partons), are Lorentz contracted to a distance ~ 2.R/7, with 7 = P+/mp and mp the mass of the proton. The "wee" partons with momentum fractions x
The Color Glass Condensate
and High Energy Scattering
in QCD
267
the latter appear to live forever. The valence parton sources are thus static sources of color charge. Since their momenta are large, they are unaffected by absorbing or emitting soft quanta: they are recoilless sources of color charge. In this "eikonal" approximation, the wee parton cloud couples only to the "plus" component of the LC current, which, from the discussion here, can be written as (see also Sect. 2.3 below): J^a = S>1+S{x-)pa(xL),
(2.2)
a
where p (x±) is the valence quark color charge density in the transverse plane. The 5-funtion in x~ assumes an infinitely thin sheet of color charge. The assumption can and must be relaxed; namely, pa(x±) —> pa(xy,x~) as we will discuss later. Note that pa is static, i.e., independent of the LC time x+, for the reasons explained previously. We now turn to the color charge density pa(x±) and how it is generated. We assume that the nucleus is interacting with an external probe which can resolve distances of size Axj_ in the transverse plane that are much smaller than the nucleon size ~ AQCD- Now, in the longitudinal direction, the small probe which has x -C A~1^ simultaneously couples to partons from nucleons all along the nuclear diameter. Since its transverse size is much smaller than the nucleon size, it sees them as sources of color charge. If the density n = NcA/irR2A ~ k2QCDA1/3 (RA = RoA1/3 is the radius of the nucleus) of the valence quarks in the transverse plane is large, n 3> A Q C D , and if n _ 1 -C ASj_
(QaQa)
= g2CfAN
= AS±
9 Cf
^°A,
(2.3)
•KR\
where we have used the fact that the color charge squared of a single quark is g2tata = g2Cj. One can treat this charge as classical since, when A7V is large enough, we can ignore commutators of charges: | [Q", Qb]\ = \ ifabcQc | < Q 2 . Let us introduce the color charge density pa(x~, x±) in such a way that: a p (x±) = f dx~pa(x~,xj_). Then, Qa = [ JAS±
d2x_Lpa(xJL)
= / JAS±
d2xx
Idx' J
pa{x~,x1_),
(2.4)
268
E. Iancu and R.
Venugopalan
and eqs. (2.3) imply (recall that Cf = (N2 - 1)/2AQ: {pa{x±)pb(y±))A = SabS(-2'>{x± {pa{x~,xj_)pb(y~,y±))A
I' dx-\A(x-)
N
2
= 6ab5(-2){x± - -y±)6(x~
= tx2A.
2
3
2
^
-y~)\A(x~),
(2.5)
Here, p?A ~ A1/3 is the average color charge squared of the valence quarks per unit transverse area and per color, and A^(a; _ ) is the corresponding density per unit volume. The latter has some dependence upon x~, whose precise form is, however, not important since the final formulae will involve only the integrated density p?A. There is no explicit dependence upon x± in fiA or XA(x~) since we assume transverse homogeneity within the nuclear disk of radius RA. Finally, the correlations are local in x~ since, as argued before, color sources at different values of x~ belong to different nucleons, so they are uncorrelated. All the higher-point, connected, correlation functions of pa(x) are assumed to vanish. The non-trivial correlators (2.5) are generated by the following weight function 14,25 (with x = (x~,x±)):
»™-"->H/'**»}•
(26>
'
which is a Gaussian in pa, with a local kernel. This is gauge-invariant (since local), so the variable pa in this expression can be the color source in any gauge. The choice of a gauge will however soon become an issue when we shall study the dynamics of the gluons radiated by this random distribution of color charges. The local Gaussian form of the weight function in Eq. (2.6) is valid, by construction, for a large nucleus, and within some restricted kinematical range that we spell here again, for more clarity. As already discussed, this is correct for a transverse resolution Q2 = 1/AS± within the range AQCD -C Q2 <£i A'QCDA1^3. But, clearly, the assumption that the valence quarks are uncorrelated must fail for transverse separations of order i?o ~ 1 / A Q C D or larger, since the 7VC = 3 valence quarks within the same nucleon are confined in a color singlet state. Thus, the total color charge, together with its higher multipolar moments, must vanish when measured over distances of the order of the nucleon size Ro, or larger. As emphasized by Lam and Mahlon 49 (see also Ref. 50 for an earlier discussion), the requirement of color neutrality can be included in the Gaussian weight function by replacing the <5-function in Eq. (2.5) with (pa(x~,x±)pb(0)) = X(x~,xj_)Sab, where X(x~,x±) is
The Color Glass Condensate
and High Energy Scattering
in QCD
269
such that its Fourier transform X(x , k±) vanishes rapidly at momenta k± %> A-QCD-
Consider also the validity range of Eq. (2.6) in longitudinal momenta. As explicit in the previous analysis, the color fields that we are computing have small values of x
2.3. The Color
Glass
Once the weight function for the classical color charge configurations associated with the large-x partons is known, it is possible to write down an effective theory for the small-x gluons. The generating functional for the correlation functions of the small-x gluons reads 23 :
J
\
JA
VA5(A+)e^[A,P\
J
where the external current Ja i s a formal device to generate Green's functions via differentiations, and S[A, p] is the action that describes the dynamics of the wee gluons in the presence of the classical color charge p (see below). The path integral over the gluon fields is written in the light cone (LC) gauge A+ = 0, since this is the gauge which allows for the most direct
270
E. Iancu and R.
Venugopalan
partonic interpretation. Correspondingly, pa is the color charge density in the LC gauge. Note the dependence upon the intermediate longitudinal momentum cutoff A + in the integrals in Eq. (2.7). As explained at the end of the previous subsection, A + is the scale which separates the 'fast' partons (p+ > A + ), which have been 'integrated out' and replaced by the classical color charge pa, from the 'wee' gluons (k+ < A + ), for which the effective theory is meant, and which are still explicit in the path integral, as the gauge fields A%. Since obtained after integrating out the modes with p+ > A + (see Sect. 3), the weight function W^+[p] depends upon the separation scale A + . Clearly, A + must be chosen such that A + / P + > x, with x the longitudinal fraction of interest. For instance, for a large nucleus at not so high energies (e.g., a gold nucleus at RHIC) we have x ~ 10 _ 1 — 10~ 2 , so we can neglect the quantum evolution in a first approximation, and identify p with the color charge of the valence quarks. Then, we can use the MV model described previously, in which case WA+[P] = WA[/»] is the Gaussian weight function given by Eqs. (2.6) and (2.5). But for smaller values of x (say x < 1 0 - 3 ) , as relevant for DIS at HERA, and also for nucleus-nucleus scattering at LHC, quantum effects are essential, so p must include the color sources generated by quantum evolution down to A + . As we shall see in Sect. 3, these sources are predominantly gluons. Thus, in what follows, we shall restrict the quantum evolution to the gluonic sector; that is, the only fermions to be included among the color sources are the valence quarks in the initial condition. If the scale k+ = xP+ of interest is of order A + , or slightly below it, the correlation functions at that scale can be computed in the classical approximation, i.e., by evaluating the path integral over A*1 in Eq. (2.7) in the saddle point approximation. There are two reasons for that: a) The quantum corrections due to gluons in the intermediate range k+
= 6»+pa(x),
igAaJTa with (Ta)bc = -ifabc-
(2.8)
The Color Glass Condensate and High Energy Scattering in QCD
271
Note however that, precisely because the occupation numbers are so large, the corresponding classical fields are strong {Ax ~ \jg at saturation), and thus the classical non-linearities must be treated exactly. In particular, we need the exact solution to the classical equations of motion (EOM) (2.8), that we shall construct in the next subsection. Once this solution is known as an explicit functional of p, the correlation functions of interest are obtained by averaging over p, with weight function WA+[/>]- For instance, the 2-point function is computed as: (AZ(x+,x)A»b(x+,y))A+
= JvPWA+[p}A^)Al(y),
(2.9)
where A% = A^[p] is the solution to Eq. (2.8), and is independent of the LC time x+ (because so is the source pa(x)). This means that only equaltime correlators can be computed in this way; but these are precisely the correlators of interest for small-x scattering. The remaining question is: What is the weight function W A + H for A +
= T exp J ig f
dz+A-{z+,x)
\ ,
(2.10)
272
E. Iancu and R.
Venugopalan
with T denoting the ordering of the color matrices in the exponential w.r.t. their x+ arguments. The action generating the EOM with current (2.10) reads 22 ' 23
S[A,p] = -jdix^F^Fr
+ -^-Jd3xTT{p(x)W[A-}(x)},
(2.11)
where W[yl - ](a;) is given by Eq. (2.10) with x+ —> oo. This action is gauge-invariant indeed 23 . (Another gauge invariant generalization for pA~, namely Tr {p(x) ln(W[A~])(x)}, has been checked in Ref. 51 to give equivalent results.) The mathematical structure of the average over p in Eqs. (2.7) and (2.9) is that of a Color Glass 23>50. Note indeed the special structure of the 2-point function that follows from Eq. (2.7):
(r^(x)A-M)A+ - fv,wkM J ^ y ^ M . ^ " ! . (,12) J
[
f
VA etSlA>p]
J
This is not the same as: fVp
WA[p] fAVAA»(x)A»(y)
eiSlA-A
fVPWA\p] fAVAeiSlA>p\ The physical reason for this is the fundamental separation in time between the rapidly varying wee gluons and the comparatively 'frozen' large-x partons. One thus solves the dynamics of the wee gluons at a fixed distribution of color charges, and only then averages over the latter. There is no feedback from the evolution of the sources on the wee gluon fields. And there is no interference between successive configurations of the color sources. These features, together with the large fluctuations in the color charge density, are the ultimate reasons for treating the large-x partons as forming a classical random distribution. The prototype of a glass is the "spin glass" — a collection of magnetic impurities randomly distributed on a non-magnetic lattice. The dynamical degrees of freedom, which are rapidly varying, are the magnetic moments of the impurities (the "spins"), while the slowly varying disorder refers to the positions of these spins in the host lattice. To study the thermodynamics of such a system, one first computes the free-energy (= the logarithm of the partition function) of the spin system for a fixed disorder (namely, for a given spatial configuration of the impurities), and subsequently makes an average over all such configurations, with some weight function. The final average over the configurations is not a thermal one: what is averaged is the free-energy computed separately for each configuration.
The Color Glass Condensate
and High Energy Scattering
in QCD
273
Similarly, the connected correlation functions of the small-x gluons in the present effective theory are obtained from the following generating functional:
F\tt\ = Jvp WA[p] In (jAVAS(A+)
eiSlA,P]-iSi-A )
(2 .i 4 )
where the logarithm is taken inside the integral over p. That is, the freeenergy reaches its extremum as a function of the external source j for a fixed distribution of the color sources. The measured free-energy (or correlation function) is finally obtained by also averaging over p. Note that the presence of a non-trivial color charge background breaks gauge symmetry explicitly. But this symmetry is restored in the process of averaging over p provided the weight function is gauge invariant, which we shall assume in what follows. 2.4. The classical
color field
In this subsection, we shall construct the solution to the classical EOM (2.8). We note first that, for a large class of gauges, one can always find a solution with the following properties 20 : F? = 0 ,
A~= 0,
A+, Ai : static,
(2.15)
where "static" means independent of x+. This follows from the specific structure of the color source which has just a "+" component, and is static. Since Fli = 0, the transverse fields A% form a two-dimensional pure gauge; that is, there exists a gauge rotation U(x~,x±) € SU(N) such that: Ai(x~,x±)
= -U(x-,x±)diUi(x-,x±).
(2.16)
(in matrix notations appropriate for the adjoint representation: A1 = AlaTa, etc). Thus, the requirements (2.15) leave just two independent field degrees of freedom, A+(x) and U(x), which are further reduced to one (either A+ or U) by imposing a gauge-fixing condition. We consider first the covariant gauge d^A11 = 0. By eqs. (2.15) and (2.16), this implies diA1 = 0, or U = 0. Thus, in this gauge, A%(x) = ^+Oia(x^,x±), with aa(x) linearly related to the color source pa in the COV-gauge: - V i a „ ( i ) = pa(x).
(2.17)
274
E. Iancu and R.
Venugopalan
The only non-trivial field strength is the electric field T^% Eq. (2.17) has the solution: aa(x~,x±)
= / d2y±(x±\—^-\y±)
!
'
•
=
—diaa.
pa(x~,y±)
i^^^-Lv^''^'
(2 i8)
-
where the infrared cutoff p, is necessary to invert the Laplacian operator in two dimensions, but it will eventually disappear from (or get replaced by the confinement scale AQCD m ) our subsequent formulae. We shall need later the classical solution in the LC-gauge A+ — 0. This is of the form A£ = S^Ai with Ai(x~,x±) a "pure gauge", cf. Eq. (2.16). The gauge rotation U(x) can be most simply obtained by a gauge rotation of the solution in the COV-gauge: A" = U(Ail + -dll)U\ 9 where the gauge rotation U(x) is chosen such that A+ = 0, i.e., U\X~,XA_)
= Pexpiig /
dz~ aa{z~,xL)Ta
\ .
(2.19)
(2.20)
The lower limit XQ —> - c o in the integral over x~~ in Eq. (2.20) has been chosen such as to impose the "retarded" boundary condition: Ai(x)
-> 0
as
x~ -> - c o ,
(2.21)
which will be useful in what follows. (Note that the "retardation" property refers here to x~, and not to time.) Together, eqs. (2.16), (2.18) and (2.20) provide an explicit expression for the LC-gauge solution A1 in terms of the color source p in the COVgauge. This is sufficient for the purpose of computing observables since the average in Eq. (2.9) can be re-expressed as a functional integral over the covariant gauge color source p by a change of variables: (A'(x+, x)A*(x+, y) • • •) A + = J Vp WA+ [p] A\[p] A{[p] • • • . (2.22) Up to now, the longitudinal structure of the source has been arbitrary: the solutions written above hold for any function pa(x~). For what follows, however, it is useful to recall, from Sect. 2.2, that p has is localized near x~ = 0. More precisely, the quantum analysis in Sect. 3 will demonstrate
The Color Glass Condensate
and High Energy Scattering
in QCD
275
A1
Fig. 8. The longitudinal structure of the color source p and of the classical field solution A1 for the effective theory at the scale fc+ . As functions of x~, a and T+% are as localized as p.
that the classical source at the longitudinal scale A + has support at a; within the range 0 < x~ < 1/A + . From Eq. (2.18), it is clear that this is also the longitudinal support of the "Coulomb"-field a. Thus, integrals over x~~ as that in Eq. (2.20) receive their whole contribution from x~ in this limited range. Any probe with momenta q+ -C A + , and therefore a much lower longitudinal resolution, will not be able to discriminate the internal structure of the source. Rather, it will see a source/field structure which is singular at x~ = 0: pa{x~,x±) w S(x~)pa(x±) (see Fig. 8). In particular: A\x-,Xl_)
-V(diVi){x±),
« 6{x-)
where V and V* are obtained by letting x V^(a;j.) = Pexp
(2.23)
—> oo in Eq. (2.20):
dz~ a(z~,x±)
(2.24)
>.
distribution 2
We denote by G(x, Q )dx the number of gluons in the hadron having longitudinal momenta between x P + and (x + dx)P+, verse size Ax± ~ 1/Q. In other terms, the gluon distribution the number of gluons with transverse momenta k± ;$ Q per (see Refs. 19, 20 for more details): rQ2 2 KG(X, Q ) = f d k± k+ 2
I
3
2
dN dk+d2k±
wavefunction and a transxG(x, Q2) is unit rapidity
k+=xP+ +
d ke(Q -kl)x5{x-k+/P )
dN — ,
(2.25)
276
E. Iancu and R.
Venugopalan
where Q(x) is the step function, k = (fc + ,kx) and rIN -^ = (a$(x+,k)aic(x+,k))
1k+ = -^(Ai(x+,k)Ai(x+,-k))
, (2.26)
is the Fock space gluon density, namely, the number of gluons per unit of volume in momentum space. The difficulty is, however, that this number depends upon the gauge, so in general it is not a physical observable. Still, as we will shortly argue, this quantity can be given a gauge-invariant meaning when computed in the light-cone gauge A+ = 0. Using the fact that, in this gauge, F^+(k) = ik+A\{k), one obtains (with k+ = xP+ from now on): xG(x, Q 2 ) = 1 1 | ^ 9 ( Q 2 - kl)(F?(k)F?(-k)),
(2.27)
which so far does not look gauge invariant. A manifestly gauge invariant operator can be constructed by appropriately inserting Wilson lines 23>20. In LC gauge, this gauge invariant expression reduces to Eq. (2.27) once the residual gauge freedom of the transverse components of the gauge field is fixed by imposing the "retarded" boundary condition (2.21) 20 . This particular gauge fixing in the classical field problem has important consequences for the quantum calculation in Sect. 3, in that it fixes the ie prescription to be used for the 'axial pole' in the LC-gauge gluon propagator 23 . We shall need later also the gluon density in the transverse phase-space (also referred to as the "unintegrated gluon distribution", or the "gluon occupation number"). This is denned as:
- ^
^—^
, (2.28)
where r = ln(l/x) = \n(P+/k+). Up to the factor 47r3, this is the number of gluons of each color per unit rapidity per unit of transverse phase-space. (As before, we assume a homogeneous distribution in the transverse plane, for simplicity.) For illustration, let us compute the gluon distribution of a nucleus in the MV model. We start with the low density regime, valid when the atomic number A is not too high, so the corresponding classical field is weak and can be computed in the linear approximation. By expanding the general solution (2.16) to linear order in p, or, equivalently, by directly solving the linearized version of Eq. (2.8), one easily obtains: ^
( f c )
"-fc
+
+ /a(Tl'fc±)'
^W^i^Paik),
(2.29)
The Color Glass Condensate and High Energy Scattering in QCD
277
which together with Eq. (2.5) implies: (Fa+(k)ra+(-k))A
~ -±- (Pa(k)pa(-k))A
= nRA(N?
- 1)&.
(2.30)
By inserting this approximation in Eqs (2.28) and (2.27), one obtains the following estimates for the gluon density and distribution function:
(2.31) (iVe2 - l)R\ ^
Q ij2 f HA J A
dfei
OCD '•QCD
= fc
-L
a^JVcC/ ^
ln
Q2 ^QCD
(with as = g2/A-K). The integral over k± in the second line has a logarithmic infrared divergence which has been cut off at the scale AQCD since we know that, because of confinement, there is color neutrality on the nucleon size Ro ~ 1/AQCD 49- We will argue later that, after taking into account quantum evolution, the actual scale for the screening of the infrared physics is not AQCD but the saturation scale Qs. Eqs. (2.31) are in fact the expected results, which could have been obtained also by a direct analysis of the gluon radiation by a single quark, together with the assumption that gluons radiated by different quarks do not interact with each other, so that the total gluon distribution is simply the sum of independent contributions from the Ax Nc valence quarks. This is the Weizsacker-Williams approximation for radiation off independent quarks.
2.6.
Gluon saturation
in a large
nucleus
According to Eq. (2.31), the gluon density in the transverse phase-space is proportional to A1/3, and becomes arbitrarily large when A increases. This is however an artifact of our previous approximations which have neglected the interactions among the radiated gluons, i.e., the non-linear effects in the classical field equations. To see this, one needs to recompute the gluon distribution by using the exact, non-linear solution for the classical field, as obtained in Sect. 2.4. By using ^ l ( ^ ) = Uab(—dlab), one can express the relevant LC-gauge field-field correlator in terms of the color field in the COV-gauge: {FrW^mA
= ((Ulbdiab)s
(Ul&a')g)A.
(2.32)
278
E. Iancu and R.
Venugopalan
One can show that the RHS of this expression can be written as ({Ulbdiab)s
{Ul&a^g)
= (diab(x)diaD(if))
(ul(x)Uca(y))
2
= 6(x- - y-)(TrUHx)U(y))(-
V
±lA(x~,x±
- y±)),
(2.33)
where we have used U\c = Uca in the adjoint representation. Here we have made use of the following correlation function (aa(x)ab(y))A
= Sab5(x~ -y~)jA(x~,x±
lA{x~,kj_)
-y±),
= -j^-XA{x~),
(2.34)
which follows from pa(x~,k±) = k\aa{x~,k±) together with Eq. (2.5). Eq. (2.33) can be proven 20 using rotational symmetry, the path ordering of the Wilson lines in x~, and the fact that the 2-point function of the color fields, Eq. (2.34), is local in x~. The trace in Eq. (2.33), SA(x-,x±-y±)
= j^—^(TrU^(x-,x±)U(x-,y±))A
can be explicitely computed as
,
(2.35)
26 20
'
SA(x-,r±)=exp{-g2Nc[
dz~ [fA{z~ ,0 L) - jA(z~,r±)}},
(2.36)
J — CO
where (cf. Eq. (2.34)) ifik i _ ^
1 t l
_ei*XTx].(2.37)
/ The above integral over k±_ is dominated by soft momenta, and has a logarithmic infrared divergence which, in this classical context, can be screened only by confinement at the scale AQCD- TO leading-log accuracy, i.e., by keeping only terms enhanced by the large logarithm ln(l/'r\A?Q CD ), the precise value of the infrared cutoff is not important, and we can also expand the integrand as: f d2kL l - e ' J
(2TT)
2
f c
k\
^
l/ri f d2k± ~ J
This gives, with fiA(x~) = / ^
1 (k± • r±)2 2
(2TT) k\
2
r\ ~ 167T
1 n
r 2 A2QCD
'
dz~XA(z~),
SA(x-,r±) ~ e x p i - ^ r i / ^ O O l n - y - ^ — 1, 4 r lv [ ± QCD )
(2.38)
The Color Glass Condensate
and High Energy Scattering
in QCD
279
which together with Eq. (2.33) can be used to finally evaluate the gluon density in Eq. (2.28). After simple manipulations, one obtains 2 6 , 2 7 ,
1 - exp{ - \
T\Q\
In ^ — }
where Q\ = OLsNcix\ = asNc j dx-\A(x-)
~
A1/3.
(2.40)
To study the /cj_-dependence of Eq. (2.39), one must still perform the Fourier transform, but the result can be easily anticipated: Let us first introduce the saturation momentum QS(A) which, as we shall see, is the scale separating between linear and non-linear behaviours. This is denned by the condition that, for r± = 2/Qs(A), the exponent in Eq. (2.39) becomes of order one, which gives: Q2S(A) ~ asNci?A l n f f ^
~ A 1 / 3 In A.
(2.41)
"•QCD
Note that this is larger than QA, Eq. (2.40), since we work in the hypothesis that QA » A-QCD- Then we distinguish between two regimes: i) At high momenta k± » QS{A), the integral is dominated by small r±
^(fc±) * ^Nc ^ = f
f r k±>>QA
°
-
(2 42)
'
ii) At small momenta, k± <§: Qs (A), the dominant contribution comes from large distances r± » l / Q s ( A ) , where one can simply neglect the exponential in the numerator and recognize 1/r^ as the Fourier transform b of lnfc^ :
for
k±^QA.
(2.43)
Unlike the linear distribution (2.42), which grows like A1/3, and is strongly infrared dominated (as it goes like l/k\), the distribution in Eq. (2.43), which takes into account the non-linear effects in the classical Yang-Mills equations, rises only logarithmically as a function of both A and l/k\. This b
T h e saturation scale provides the ultraviolet cutoff for the logarithm in Eq. (2.43) since the short distances r±
280
E. Iancu and R.
Venugopalan
is saturation. At saturation, the gluon occupation factor is parametrically of order l/as, which corresponds to a Bose condensate, and is the maximum density allowed by the repulsive interactions between the strong color fields A1 = ^/(AiAi) ~ 1/g. When increasing the atomic number A, the new gluons are produced preponderently at large transverse momenta ^ QS(A). where this repulsion is less important. This is illustrated in Fig. 9. K
A
n
Fig. 9. The gluon phase-space density ¥M(fcx) of a large nucleus (as described by the MV model) plotted as a function of k± for two values of A. Notice the change from a l/k2x behaviour at large momenta k±_ > Qs to a logarithmic behaviour at small momenta
To clarify the physical interpretation of the saturation scale, note that, at short-distances r±<€.l/Qs(A),
A m"
1 2 A A2
_L QC£)
OC
xG(x, 1/rj) {N* -
(2.44)
l)irR2A
is the number of gluons (of each color) having tranverse size r± per unit of transverse area (cf. Eq. (2.31)). Since each such a gluon carries a color charge squared (gTa)(gTa) = g2Nc, we deduce that a3Ncu?A In -^-A— r
J_AQCD
is the average color charge squared of the gluons having tranverse size r± per unit area and per color. Then, Eq. (2.41) is the condition that the total color charge squared within the area occupied by each gluon is of order one. This is the original criterion of saturation by Gribov, Levin and Ryskin n , for which the MV model offers an explicit realization.
The Color Glass Condensate
and High Energy Scattering
in QCD
281
Let us finally compute the (integrated) gluon distribution in the saturation regime, i.e., for Q2
This shows strong nuclear shadowing: it scales like A2/3\nA, linear result (2.31), which scales like A. 2.7. Dipole-hadron
scattering
at high
unlike the
energy
Although the previous definition of the gluon distribution in terms of the Fock space gluon density is useful for a conceptual discussion of saturation, it is on the other hand less clear whether it corresponds to something that could be directly measured in experiments. Recall that, in the lowest-order analysis of deep inelastic scattering where one neglects non-linear effects in the hadron wavefunction, the gluon distribution is related to the scaling violation in the hadron structure function F2: d i ^ / d l n Q 2 oc asxG(x, Q2). It is therefore interesting to compute this quantity also in the presence of non-linear effects, and identify some measurable consequences of saturation. This is what we shall do starting with this section, first in the framework of the MV model, then by including the effects of the quantum evolution in going towards smaller values of x (in Sects. 3 and 4). As a general conclusion, we shall find that saturation effects in the hadron wavefunction correspond to unitarity effects in the high-energy virtual-photon-hadron collision, whose total cross-section is related to F2 via:
(2.46)
As we have seen in Sect. 1.5, when viewed in the infinite momentum frame (IMF) of the hadron, DIS appears as the scattering of the virtual photon off a quark (with longitudinal fraction equal to the Bjorken x of the collision) in the hadron wavefunction. At very small x, this quark is typically not a valence quark, but rather a see quark which is emitted, most probably, off the small-x gluons. It is then convenient to disentangle this final quark emission from the quantum evolution which involves mostly gluons. This can be done by performing a Lorentz boost in such a way to pull the j*qq vertex out of the hadron. That is, instead of the hadron IMF, it is preferable to use the so-called dipole frame 19 (and references therein) in which most
282
E. Iancu and R.
Venugopalan
of the energy is still carried by the hadron (so that the high density effects are again associated with the hadron wavefunction), but the virtual photon moves in the negative z (or positive a: - ) direction with enough energy to dissociate before scattering into a quark-antiquark pair in a color singlet state (a color dipole), which then scatters off the hadron. This sequential picture of DIS is appropriate at high energy, since the lifetime of the qq pair is much larger than the interaction time between this pair and the hadron. More precisely, if r = ln(l/x) = yu — y7* is the (boost-invariant) rapidity gap, with r > 1 at small x, then the dipole frame corresponds to choosing y7»
/ dz / d 2 r J .|tf(z,r ± ;Q'')|''<7dipoie(T,rx).
' ' - / * / *
(2.47)
Here, 9(z,r±;Q2) is the light-cone wavefunction for the photon splitting into a qq pair with transverse size r± and a fraction z of the photon's longitudinal momentum carried by the quark 57 ' 58 . Furthermore, o-
d2b±MT(r±,b±).
(2.48)
At high energy, the dipole-hadron scattering can be treated in the eikonal approximation 57 ' 60 ' 52 . This amounts to neglecting the recoil of the quark (or the antiquark) during its scattering off the color field in the target: the whole effect of the scattering consists in a color precession. Then, the scattering amplitude reads AfT(x±,y±) = l — ST(x±,y±), with the following 5-matrix element: ST(x±,y±)
= ^(tT(VHx±)V{y±)))T,
(2.49)
where V^ and V are the Wilson lines describing the eikonal interaction between the quark (or the antiquark) and the color field at rapidity T due to color sources within the hadron: V^(xj_) = Pexp (ig /
dx~A+(x~,xj_)ta
J
(2.50)
The Color Glass Condensate
and High Energy Scattering
in QCD
283
That is, this is the same as Eq. (2.24), but rewritten in the fundamental representation. The average in Eq. (2.49) is over all the configurations of the color fields in the hadron. In the CGC formalism, where A% = aa, cf. Eq. (2.17), this average is computed as in Eq. (2.22). In what follows we shall focus on the computation of the 5-matrix element (2.49), which encodes all the information about the hadronic scattering, and thus about the non-linear and quantum effects in the hadron wavefunction. [Once this is known, F2 can be immediately obtained by using Eqs. (2.47) and (2.46).] In the MV model, to which we shall restrict in the remaining part of this subsection, this S-matrix element is already known, as obvious when comparing Eqs. (2.49) and (2.35). By translating Eq. (2.38) to the fundamental representation (Nc —> CF = (N% — 1)/2NC) and letting x~ —> 00 there, one obtains: SA(r±)
~ exp{ - l I ^ l n — L — l I
4
r 1V
L QCD
(2.51)
)
where QA = asCFn\ differs only via a color factor from Eq. (2.40). As in the previous discussion of the gluon distribution, we distinguish between a small-r_i_ and a large—r± regime, with the separation between the two regimes given by the saturation scale Q23{A), defined by analogy with Eq. (2.41). i) A small dipole, with r± <^.1/QS(A), is only weakly interacting with the hadron: AfA{rx)
= 1 _ sA(rx) « \ r\Q\
In - ^ —
«
1,
(2.52)
a phenomenon usually referred to as "color transparency" 113 . ii) A relatively large dipole, with r±_»l/Qs (but r±_
284
E. Iancu and R.
Venugopalan
Note that, according to Eq. (2.51), the multiple scatterings are higher-twist effects, i.e., their contributions are of higher order in r\. We see that, remarkably, it is the same scale — the saturation momentum — which sets the critical transverse size for both gluon saturation and the unitarization of the dipole-hadron scattering. This conclusion, that we have found here in the framework of the MV model, will be seen in Sect. 3 to remain valid after including the quantum evolution. Physically, this can be understood as follows: Both saturation and unitarization (when the scattering is seen in the dipole frame) require strong color fields in the hadron wavefunction, such that g fdx~A+ ~ 1. Qs is the critical transverse scale at which this strong field condition begins to be satisfied. What is specific to the present MV model (and, more generally, to any approximation in which the color sources are only weakly correlated with each other, like the gluonic sources in the DGLAP approximation 10 - 12 ) is that the dipole scatters independently off the color sources in the hadron (here, the valence quarks). This is best seen by noticing that Eq. (2.51) can be rewritten as a Glauber formula:
SAr,) - «p{ - aS± | ^
M1
'^"i)}.
PUS)
where XGN(X,Q2) is the gluon distribution of a nucleon, and is given in the present approximation by the second line in Eq. (2.31) with A —> 1. As we shall see in the next section, the previous picture changes quite substantially after including quantum corrections, due to the fact that the evolution towards small x induces correlations among the color sources. As a result, not only the general Glauber formula (2.53) becomes inapplicable (the successive scatterings are not independent any longer), but even its linearized 'leading-twist' approximation, corresponding to a single scattering, fails to apply when l/r± is close enough to Qs, while still above it. This is the BFKL regime where 'higher-twists' effects appear already in the linear evolution. Most interestingly, we shall see that gluon saturation at small x holds independently of the non-linear effects in the classical Yang-Mills equations. Rather, this is the consequence of the correlations among the color sources induced by non-linear effects in the quantum evolution. These same correlations will be shown to ensure color screening already over the perturbative scale 1/Qs <£. 1 / A Q C D , which thus eliminates the infrared sensitivity of the classical MV model to the non-perturbative physics of confinement (see, e.g., Eqs. (2.39) or (2.51)).
The Color Glass Condensate
and High Energy Scattering
in QCD
285
3. The Quantum Evolution of the Color Glass Condensate In this section, we shall explain how to construct the CGC effective theory at small x by integrating out the gluons with x' > x in perturbation theory, in the presence of high density effects. The central result of this analysis will be a renormalization group equation for the weight function VKA+[P] in Eq. (2.7), which generalizes the BFKL equation by including non-linear effects, and has important physical consequences among which gluon saturation. 3.1. The BFKL evolution
and its small-x
problem
Within perturbative QCD, the enhancement of the gluon distribution at small x proceeds via the gluon cascades depicted in Fig. 10. Fig. lO.a shows the direct emission of a soft gluon with longitudinal momentum k+ = xP+
asNc
pi
•K
p+ _ x0 In 7 +x = a3 In — K
(3.1)
X
relative to the tree-level process in Fig. 10.a. This correction is enhanced by the large rapidity interval A T = ln(xo/x) available for the emission of the additional gluon.
P+ k + « P"'
a) Fig. 10. a) Small-x gluon emission by a fast parton; b) the lowest-order radiative correction; c) a gluon cascade.
d
A t the same level of accuracy, a complete calculation must include also the appropriate virtual corrections (self-energy and vertex renormalization); but for the present, qualitative purposes, it is sufficient to consider the real gluon emission.
286
E. Iancu and R.
Venugopalan
A similar enhancement holds for the gluon cascade in Fig. 10.c, in which the succesive gluons are strongly ordered in longitudinal momenta: p+ » Pi ^ P2 ^ ""' ^ Pn ^ ^ + • This gives a contribution of relative order
A(a,.„5!)\
(3.2)
where the factorial comes from the ordering in p+. Clearly, when x is so small that ln(xo/x) ~ l / a s , all such quantum "corrections" become of order one, and must be resummed for consistency. A calculation which includes effects of order (as ln(l/x))" to all orders in n is said to be valid to "leading logarithmic accuracy" (LLA). The gluon cascades in Fig. 10 contribute all to the production of (virtual) gluons with longitudinal fraction x. Thus, by resumming these cascades, one can compute the number of such gluons per unit rapidity, i.e., the gluon distribution (2.25). One can recognize in Eqs. (3.1)-(3.2) the expansion of an exponential. Therefore: dN — = xG(x,Q2) ~ eu*-T = x~ua- , (3.3) dr with uj a pure number. We have tacitly assumed that all the gluons in the cascade have transverse momenta of the same order, namely of order Q. A more refined treatment, based on the BFKL equation 8 , allows one to compute UJ and specifies the Q 2 -dependence, and also the subleading rdependence (beyond the exponential behaviour shown in Eq. (3.3)) of the gluon distribution. To describe the effects of the BFKL evolution in more detail, it is instructive to consider the dipole-hadron scattering introduced in Sect. 2.7. With increasing energy, the gluon fields change in the hadron wavefunction, and therefore so does also the cross-section for the dipole which couples to these fields (cf. Eq. (2.49)). Specifically, to LLA, and in the linear regime where one can neglect multiple scattering, the amplitude NT(x±,y±) = 1 - ST(x±,y±) = Afxy obeys to: dr
J
2TT {X± -z±)2(y±
- z±)2
l
y>
which is the coordinate form of the BFKL equation 8,4S . The physical interpretation of this equation depends upon the Lorentz frame that we choose to visualize the process. When using the dipole frame of Sect. 2.7, the quantum evolution is put entirely in the wavefunction of the hadron (which is boosted to higher and higher energies with increasing r ) , while the dipole remains a simple qq pair. In this frame, Eq. (3.4) describes
The Color Glass Condensate
and High Energy Scattering
in QCD
287
fast partons (p)
Fast hadron
small-x gluons (A[pj) Dipole Fig. 11.
Deep inelastic scattering in the dipole frame
the dipole-hadron scattering as the exchange of a BFKL ladder; the dipole couples to the last gluon (with the smallest value of x) in a gluon cascade which develops fully inside the hadron. (See also Fig. 11.) Alternatively, by a change of frame, one can use the increase in the total energy to accelerate the dipole, and study the evolution of its wavefunction with r. Under an increment dr such that asdr ~ 1, the dipole evolves by emitting one gluon (from either the quark or the antiquark), and the ensuing qqg state scatters off the hadronic target. It is convenient (although not necessary) to view this evolved state in the large-Nc limit, in which the radiated gluon is effectively replaced by a qq pair in a color octet state. Then, the evolution looks like the splitting of the original dipole into two new dipoles, each of them made of a quark (or antiquark) from the initial dipole and an antiquark (or a quark) from the emitted gluon. From this perspective, the various terms in Eq. (3.4) have a simple interpretation: the quantity
(x± - y±f 2?r (X±_ - z±)2(y±
- z±)2
(3.5)
is the differential probability for the initial dipole {x±,y±) to decay into a pair of dipoles (x±,z±) and (x±,z±_), while Afxz and Mzy are the amplitudes for the scattering between any one of these final dipoles and the target. Finally, the negative contribution proportional to —Mxy represents the decrease in the scattering amplitude of the original dipole due to its dissociation (this term is necessary for the conservation of the probability). One should mention here that this different perspective, in which the quantum evolution is put in the dipole and studied in the large-A^ limit, lies
288
E. Iancu and R.
Venugopalan
at the basis of an approach originally developed by Mueller 48 — the 'Color Dipole approach' —, in which the wave function of a very energetic color dipole (an "onium") is constructed in the BFKL approximation. In this approach, the scattering between two "onia" (physically, this corresponds, e.g., to the 7*7* scattering) can be treated as the product of the number of dipoles in each onium times the dipole-dipole cross-section. The ensuing scattering amplitude has been shown to satisfy equation (3.4). The BFKL equation (3.4) can be solved by standard techniques 9 . At high energy, aar 3> l n ( l / r ^ A g C D ) , and for a homogeneous target (e.g., for a large nucleus, and impact parameters near the center of the target), the solution reads (with r± = x± — y±)
^^^-H^}-
<36)
where the reference scale Qo is introduced by the initial conditions at low energy, and thus is of order KQCD (for a nucleus, this carries the dependence on A). Furthermore, w = 41n2 « 2.77 and /?=28C(3) « 33.67. In writing Eq. (3.6) we have assumed a fixed coupling as, as appropriate at leadingorder BFKL accuracy. The modifications due to the running of the coupling will be discussed in Sect.??. Eq. (3.6) exhibits two essential features of the BFKL approximation, which eventually provoke its failure in the high energy limit: (a) Violation of the unitarity bound: The solution (3.6) increases exponentially with r, that is, as a power of the energy. At high energy, such a behaviour violates both the unitarity bound MT(r±,b±) < 1 on the scattering amplitude at fixed impact parameter, and the Froissart bound ^dipoie(s)
We mean here the unitarity of the scattering amplitude at fixed impact parameter. The discussion of the total cross-section is more involved, and deferred to Sect. 4.4.
The Color Glass Condensate
and High Energy Scattering
in QCD
289
multiple scatterings, i.e., by keeping terms of all orders in gA+ in the Wilson lines in Eqs. (2.49)-(2.50). Both Eqs. (2.52) and (3.6) correspond to a single scattering approximation — they are obtained by retaining terms which are, at most, quadratic in gA+ in the expansion of the Wilson lines in Eq. (2.49).) However, unlike in the MV model, at small x we expect an additional source of non-linearities which arise from the interactions among the gluonic sources. Such interactions lead to the fusion of gluons from different parton cascades ("gluon recombination"), a phenomenon which should tame the rapid growth of the number of partons. Thus saturation arises from the competition of two effects: the growth of the gluonic density due to radiation and its depletion due to recombination effects 11-12. Since now the dominant color sources are themselves gluons, the saturation we speak of here refers simultaneously to the source and the fields radiated by them. This is because what we call "sources" and "radiated fields" is only relative, as it depends upon the scale A + = xP+ at which we consider the effective theory. Both the multiple scattering and gluon recombination mechanisms are illustrated in Fig. 11. We expect both mechanisms to become important at the same scale, Q2S{T,A), which is the critical gluon density at which the non-linear effects become of order one. This saturation scale is also the typical transverse momentum of the gluons in the hadron wavefunction at small x. As anticipated in the Introduction, and will be verified explicitly in what follows, this scale increases rapidly with r and A. The emergence of such a hard intrinsic momentum scale can also solve the 'infrared diffusion' problem of the BFKL approximation and therefore restore the applicability of perturbation theory to high energy processes in QCD. In the dipole frame, gluon recombination is seen as the merging of two gluon cascades, as illustrated in the r.h.s. of Fig. 11. It is interesting to see this process also from the boosted frame in which the quantum evolution proceeds via the dissociation of the incoming dipole into two dipoles. After the boost, the final gluon in the cascade in Fig. 11 — the one which couples to the dipole — gets incorporated within the dipole wavefunction, so the merging of two cascades now happens inside the dipole. Thus, from the boosted frame, the non-linear process is seen as the simultaneous scattering of the two final dipoles off the hadronic target. These considerations suggest the following simple equation which generalizes Eq. (3.4) by taking non-linear effects into account: d2z±
-^Nxy = as J •
(xj_ - 2/j_)2
2n (x± - zA_)2(y_L - z±)2
290
E. Iancu and R.
Venugopalan
X {Nxz + AfZy - Mxy ~ NxzKy]
•
(3.7)
53
This is the equation derived by Kovchegov within Mueller's 'Color Dipole approach' 4 8 . The main assumption used in its derivation was the fact the two final dipoles scatter independently off the target; it is correct only in the large-A^c limit. Deriving this equation, (together with its generalizations to finite Nc originally obtained by Balitsky 52 within a different formalism) within the framework of the CGC effective theory, will be a main objective of the forthcoming developments in this section. 3.2. Non-linear
evolution
for the
CGC
The CGC provides a natural framework for the description of the non-linear effects in the quantum evolution towards small x, and of the phenomenon of saturation. The main observation is that, to LLA, all the quantum corrections described previously — both the exponentially developing BFKL cascade, and gluon recombination which tames this rapid growth — can be incorporated into a change of the classical color charge and its correlations, namely, into a renormalization of the weight function Wj^+ [p] in Eq. (2.7). To see this at an intuitive level, let us reconsider the first radiative correction, the one-gluon emission in Fig. 10.b, and note that, to LLA, the typical contributions to the integral in Eq. (3.1) come from momenta pf such that p+ ^> pf 3> k+. That is, the condition of separation of scales is indeed satisfied for the intermediate gluon with momentum p* to be treated as a 'frozen' color source for the final gluon with momentum k+. The effect of this quantum correction is therefore simply to renormalize the effective color source at scale k+, as pictorially illustrated in Fig. 12.
P!
k+ Fig. 12.
( k+
Effective color source after including the lowest-order radiative correction.
By iterating this argument, it is quite clear that a whole BFKL cascade (see Fig. 10.c) can be included in the definition of the classical color source at the scale A + = x P + of interest. It is furthermore clear that the fusion between two gluon cascades, as illustrated in the l.h.s. of Fig. 13, can be
The Color Glass Condensate
and High Energy Scattering
in QCD
291
represented in the CGC theory as a non-linear effect in the classical dynamics of the color fields generated by this effective source (see the r.h.s. of Fig. 13).
Fig. 13.
The fusion of two gluon cascades and its interpretation in the CGC theory.
But non-linear effects are important also in the quantum evolution, and actually interfere with it, as illustrated in Fig. 14. Fig. 14.a is an immediate generalization of the one-gluon emission in Fig. lO.b. It is clear that what is renormalized by the scattering off the "semi-fast" (A + 3> p+ ~^> k+) quantum fluctuation is the classical field A% [p] at scale A + , which in turn is non-linear in p. (The Feynman rules for evaluating diagrams like those in Fig. 14, and also the present discussion, are adapted to the LC gauge A+ = 0, in which the quantum effective theory is written, cf. Eq. (2.7).) Fig. 14.b shows an additional source of non-linearity, arising from the propagation of the radiated gluon in the classical 'background' field Al[p\. If A + = x P + is small enough (x
292
E. Iancu and R.
Venugopalan
A[PL
,A[P]
A+: p+ K
^
al Fig. 14.
b)
c)
Some typical non-linear effects in the quantum evolution.
all-order inclusion of the classical field effects permits one to resum not only the large energy logarithms, namely the terms ~ ( a s l n l / x ) " , but also the dominant high density effects — the non-linear effects (like gluon recombination) which become of order one at saturation. The corresponding analysis is technically quite involved and has been described in detail in previous publications 23 ' 20 . Here we shall present only the final results and their consequences. As shown in Ref. 23, the condition that the new correlations induced by integrating out quantum fluctuations be reproduced by the CGC effective theory leads to a functional renormalization group equation (RGE) for the weight function WA+[P] = WT[/o], which is most succinctly written as 23 dWT[p]
Br
= \l 1 /*
S , 6 w , WT[p], Xab(x±,yx)[p] b 5paT{xL) Sp T(y±)
(3.8)
in notations that we shall shortly explain. Early versions of this equation can be found in the pioneering works of Refs. 22, 54. A formally similar, and physically equivalent 63 , functional evolution equation has been obtained by Weigert 5 5 , within a different formalism 52 . We shall discuss this latter approach in the next subsection. Let us now discuss the meaning and structure of the terms in Eq. (3.8). The rapidity variable r = ln(l/:r) = l n ( P + / A + ) indicates the dependence of the effective theory upon the separation scale. This is convenient since (as illustrated by the BFKL evolution discussed in Sect. 3.1), r is the natural "evolution time". The contribution of the quantum modes within a small layer in p+ (say A + > p+ > k+) is proportional to the rapidity extent A T = ln(A+/fc+) of that layer. The kernel x[p] is a positive definite non-linear functional of p (Eq. (3.8) is a diffusion equation) and is highly non-local in both longitudinal and transverse coordinates. The non-linearity in p and the non-locality in a; - are strongly correlated, since they have a common origin: \[p] depends upon p
The Color Glass Condensate
and High Energy Scattering
in QCD
293
via the Wilson lines (2.24). Physically, x[/°]^T i s the charge-charge correlator induced when integrating out quantum gluons within a rapidity interval dr and in the presence of a classical color charge distribution with density pa(x) 23>20. Technically, this is computed by evaluating the diagram in Fig. 14.c with the Feynman rules in Eq. (2.7). Note that, in addition to this real-gluon emission diagram, there are also virtual (self-energy and vertex) corrections which must be similarly computed 22>23. Such corrections are already included in the RGE (3.8), where they correspond to the functional derivative of the kernel x[p] 23The argument p^(x±) of the functional derivatives in Eq. (3.8) denotes the color charge density pa(x~,x±) at x~ = x~. The color source generated by the quantum evolution up to rapidity r has support within a limited interval in x~, namely at 0 < x~ < x~, with x~ oc e T . The reason this is so follows from the uncertainty principle: since the classical source at rapidity r is obtained by integrating out quantum modes with large longitudinal momenta p+ » A + = e _ r P + , it must be localized near x~ = 0, within a distance Ax~ ~ GTXQ (with XQ = 1/P+). However, Eq. (3.8) shows that the correlation between the quantum evolution in r and the longitudinal distribution of the resulting color source is even stronger. When the rapidity is further increased, say from r to r + d r , the additional contribution to the color source which is generated in this way has no overlap in x~ with the original source at rapidity r. Instead, this new contribution makes a new layer in a; - , which is located between x~ and x~+dT. This is why the functional derivatives in Eq. (3.8) involve just the color source pT = p{x~) in this outermost layer. This correlation is most simply formulated if one uses the space-time rapidity y, y = ln(aT/:co),
x^ = l/P+,
-oo
(3.9)
to indicate the longitudinal coordinate of a field. For example, Py(x±) = x~pa(x~,x±) fdyp°(x±)
for
x~ - x~ = XQ e y ,
= Jdx-pa(x-,xx),
(3.10)
and similarly for a, Eq. (2.18), or any other field. Eq. (2.24) can be rewritten as: V\x±)
= Pexp [ig fdya^{x1_)ta]
.
(3.11)
294
B. Iancu and R.
Venugopalan
The previous discussion shows that the space-time rapidity y of a given layer in p is identical to the usual (momentum) rapidity of the fast gluons that have produced that layer. In particular, the color source created by the quantum evolution up to r has support at space-time rapidities y < r (in agreement with the simple argument based on the uncertainty principle). Formally: WT\p]<xST\p],
(3.12)
where 6T[p] is a 6-functional enforcing that py = 0 for any y > r. As we shall see, this constraint is important because the r-dependence of the observables in the effective theory comes precisely from the upper limit on the longitudinal support of p. The Color Glass evolves by expanding in y. Since the Wilson lines (3.11) and many interesting quantities (like the S-matrix element (2.49), or the gluon distribution (2.32)) are more directly expressed in terms of the COV-gauge field ay(x±), rather than the color charge p, it is often preferable to use the 'a-representation', whose weight function WT[a] = WT[p — — V^a] satisfies the following RGE, obtained after a change of variables in Eq. (3.8): dWT[a\
1 =
-TT
6
abl
2fa££I)*
v
{x± y±)[a]
'
,
SWT
(3 13)
8a^7)-
-
We use compact notations in which repeated color indices (and coordinates) are understood to be summed (integrated) over. The relation of the kernel here to that in Eq. (3.8) is Va"(x±,y±)=
(
-V
Jz±,ux
V
_L
±
The analysis in Ref. 23 yields (see also 55):
rtx
y^) = -(^-
V [X±,y±)
^J
(*'-«W-*'>
( 2 ? r ) 2 {x±_
z±)2{y±_z±)2 x {1 + V}Vy - V}VZ - V}Vy}a\ ab
(3.14)
with V^ = V^(x±), etc. This is real and symmetric {ri (x±,y±) f]ba{y±Jx±))i a n d a l s o positive definite, as anticipated, since:
{1 + VjVy - VjVz - VXY" = (1 - VX)ca(l - VXU,
=
(3.15)
and the color matrix 1 — VjVx is hermitian. The transverse kernel in Eq. (3.14) is similar to the 'dipole kernel' in the BFKL equation (3.4). Their relation will be discussed in the next subsection.
The Color Glass Condensate
and High Energy Scattering
in QCD
295
The RGE in Eq. (3.13) [or Eq. (3.8)] has the structure of a FokkerPlanck equation. In the CGC formalism, the quantum evolution towards small x is a random walk in the space of Wilson lines 6 3 . The random process is one by which the Wilson lines are built. The physical random variable in the evolution is the elementary contribution a^(x±) to the classical field in the hadron arising from integrating out quantum fluctuations in the rapidity strip [T, T + dr}. Such a contribution changes the Wilson lines according to: Ut+dT{x±)
= ei°dTa'^Ta
Ul(x±),
(3.16)
whose iteration defines a path in the space of the U fields. (This path is unambiguously denned only after discretizing the rapidity variable; see Ref. 63 for details.) By exploiting this representation, an exact but formal solution to Eq. (3.13) has been constructed in the form of a path integral 63 . This random walk can be equivalently reformulated as a Langevin equation 63 , a formulation which is better suited for numerical simulations on a twodimensional lattice. 3.3. The Balitsky-Kovchegov
equation
In addition to the numerical simulations on a lattice, Eq. (3.13) can be made tractable via two strategies. Both involve some approximations. The first strategy consists in using this functional equation to deduce ordinary differential equations for quantities of interest. Because of the non-linearity of Eq. (3.13), the ensuing equations will generally not be closed, but rather form an infinite hierarchy of coupled equations. Nevertheless some progress can be made in various approximations, particularly in the large Nc limit, where we shall see that a closed equation emerges. The other strategy, to be developed in the next subsection, is to search directly for approximate solutions to the functional equation (3.13) for the weight function. If (0[a] )T is any observable which can be computed as an average over a (cf. Eq. (2.22)): {0[a] )r=
I VaO[a] WT[a],
(3.17)
then its evolution with r is governed by the following equation:
-(iSjfe'Sssb 0 '"!) •
(3 18)
'
296
E. Iancu and R.
Venugopalan
where, in writing the second line, we have used Eq. (3.13) and then integrated twice by parts within the functional integral over a. The last expression involves the functional derivative of r][a], which is easily computed by using [with Sxy = <5(2)(xx - y±)]^ = i ^ r - V t ( x x ) ,
^
= -i96xyV(X±)T«.
(3.19)
Here ay = 0 for y > r, cf. Eq. (3.12). The 2-point function ST(x±,y±) of the Wilson lines, Eq. (2.49) (which physically represents the 5-matrix element for dipole-hadron scattering) can straightforwardly be computed using repeatedly Eq. (3.19) (see 23 for details):
^
Z
-{x±-z±nyx-Zl)>
~ ti
(3-2°)
This equation was originally derived by Balitsky 5 2 , within a formalism based on the evolution of observables (in high-energy dipole-hadron scattering) which are built from Wilson lines. This is similar in spirit to the 'Color Dipole approach' by Mueller 4 8 , but it is not restricted to the largeNc limit. It is better suited for an asymmetric collision, like that between an "onium" (= a high-energy dipole) and a dense hadronic target, like a nucleus. In this respect, Balitsky's formalism is closer to the CGC formalism, where the focus is fully on the target wavefunction. As anticipated, the above equation is not closed: It relates the 2-point function to the 4-point function (tT(VjVz)tr(V^Vy))Physically, it is so since, except at large Nc, the qqg system formed after radiating one gluon from the original dipole is not exactly the same as a system of two dipoles (recall the discussion around Eq. (3.5)). One can similarly derive an evolution equation for the 4-point function 52 , but this will in turn couple the 4-point function to a 6-point function, and so on. Eq. (3.20) is merely the first in an infinite hierarchy of coupled equations 52 . In Ref. 55, Weigert managed to reformulate Balitsky's hierarchy as a single functional evolution equation for the generating functional of the n-point functions of the Wilson lines. As shown in Ref. 63, Weigert's equation is equivalent to the RGE (3.13) as far as the correlations of the Wilson lines are concerned. More recently, Mueller used a similar approach to give a simple derivation 56 for Eq. (3.13).
The Color Glass Condensate
and High Energy Scattering
in QCD
297
A closed equation can still be obtained in the large Nc limit in which the 4-point function in Eq. (3.20) factorizes:
— (trfltf Vi)) T
for Nc - oo.
Then Eq. (3.20) reduces to a closed equation for the 2-point function, which, when rewritten in terms of the scattering amplitude AfT(x±,y±.) = j^-(tr(l - V^Vy))T, is recognized as the Kovchegov equation (3.7) 53 . An early version of this equation has been proposed by Gribov, Levin and Ryskin n , and proven by Mueller and Qiu 12 in the 'double-logarithmic approximation'. More recently, Braun has rederived Eq. (3.7) by directly resumming 'fan' diagrams 61 . Following the recent literature, we shall refer to Eq. (3.7) as the "Balitsky-Kovchegov (BK) equation" Clearly, in the weak scattering approximation AfT(r±) "C 1 (which corresponds to a very small dipole, or, equivalently, to a relatively low gluon density in the hadronic target), the non-linear term can be neglected in the r.h.s. of Eq. (3.7), which then reduces to the BFKL equation (3.4). But in general, the feedback provided by this non-linear term ensures that the solution AfT(r±) to Eq. (3.7) respects the unitarity bound ftfr(r_i_) < 1Thus, Eq. (3.7) is a simple QCD-based non-linear equation consistent with unitarity. This explains the large interest in this equation in the recent literature, with important progress towards its resolution via both analytic 53,65,46 a n c j n u m erical methods 65>66>67.68. The conclusions reached in this approach are equivalent to those obtained from direct investigations of the RGE (3.13) 16 - 69 , and will be described in the next sections. 3.4. Saturation
momentum
and geometric
scaling
The solution to Eq. (3.7) is shown qualitatively in Fig. 15, which displays M-(rx) as a function of r±_ = x± — yx_ for two different rapidities. The scattering amplitude vanishes as r± —> 0, as it should from its definition AfT(r±) = ^ - ( t r ( l — V£Vy))T, and the fact that the Wilson lines are unitary matrices. For small r±, J^fT(f±) remains small ("color transparency"), and is well approximated by the BFKL solution (3.6). For large r±, it approaches the unitarity bound A/"T(rj_) = 1. The transition between "color transparency" at small r± and "blackness" at large r± takes place at a characteristic value of r± that we shall identify with the saturation length 1/QS(T). More precisely, we shall define QS{T) by the following convention: K(r±)
= 1/2
for
r± =
1/Q,(T)
.
As shown in Fig. 15, this saturation length decreases with T.
(3.21)
298
E. Iancu and R.
Venugopalan
1/QS(2)
1/QS(1)
Fig. 15. The solution to the BK equation as a function of r± for two values of r .
The result illustrated in Fig. 15 clearly shows that the non-linear BK equation (3.7) solves the unitarity problem of the BFKL equation (at least, at a fixed impact parameter; see also Sect. 4.4 below). Moreover, the emergence of an intrinsic "saturation scale" QS(T), which increases with r, also solves the problem of "infrared diffusion", as convincingly demonstrated by the numerical analysis in Ref.68. In addition to the numerical studies in Refs. 65, 66, 67, 68, the behaviour shown in Fig. 15 is supported also by analytic investigations focusing on qualitative features like the energy dependence of the saturation scale 15 . 46 > 94 . 44 j the "geometric scaling" behaviour 46 ' 94 - 44 ) or the approach towards the blackness with increasing rj_ 53,65,16,46,69 ^y e gjj^j describe here some of these analytical studies, whose results follow from general arguments, such as the validity of the BFKL dynamics at small r±. and the emergence (via non-linear effects) of an intrinsic momentum scale, the saturation momentum QS(T). Consider first the calculation of the saturation scale. Even though the BFKL solution (3.6) is valid only at small r±, well below the saturation length, it is nevertheless possible to compute the energy dependence of the saturation scale by extrapolating Eq. (3.6) up to r± ~ 1/QS(T) and then imposing the saturation condition (3.21) 15-46. More precisely, we shall see shortly that Q2S{T) is increasing exponentially with r. The BFKL computation alluded to above should then correctly reproduce the value of the exponent, but not necessarily the (slowly varying) prefactor as well (see however 9 4 ) . Towards this end, the solution of the BFKL equation in Eq. (3.6) is first
The Color Glass Condensate
and High Energy Scattering
in QCD
299
re-written as (p = \n(l/r2j_Ql)): A f
r
( ^ e x p { ^
s
T ^ - ^ y ,
(3.22)
where we have kept only the dominant terms in the regime p ^> 1 and asr 3> 1, with asr 3> p. Since the BFKL equation is now seen as just an approximation to more general non-linear equations like the BK equation (3.7), the solution (3.22) is acceptable only as long as r±_ -C 1/QS(T), or p > PS(T), with p.( T ) = HQ2S{T)/Q20). If nevertheless extrapolated down to p ~ PS(T), the saturation condition (3.21) amounts to the vanishing of the exponent in Eq. (3.22). This gives a second-order algebraic equation for PS(T) with the physical solution 15>46-. Q2s(r) = Q20ec&°T,
c = [-(3+^0(p
+
8u>)}/2 = 4.84... (3.23)
This estimate is consistent with the numerical solutions to the BK equation, which have found c ^ 4 65 . 66 - 68 ) but not with the phenomenology of DIS at HERA, which suggests rather a significantly lower value for the exponent 43 (see Sect. 4.3 below), namely, A « 0.3 instead of ca3 ~ 1. The factor in front of the exponential in Eq. (3.23) is not under control in present approximations. In Ref. 94, a more refined treatment was proposed, where the BFKL equation was solved with an absorbtive boundary condition at r± ~ 1/QS(T), and a weak dependence on T for this prefactor was obtained. It would be interesting to test their results against more accurate numerical solutions to Eq. (3.7). It is also interesting to study the behaviour of the scattering amplitude (3.22) for r± below but relatively close to 1/QS{T) — for p slightly above PS{T). Since: P ln
^ ^Ql=
Ps(T) + ln
^W)
^Ps+
5p
'
(3 24)
'
a simple calculation yields: A/- T (r x )~exp ( -
7
* p - M - \ ,
(3.25)
where 7 = l/2+c//3 w 0.64. Eq. (3.25) suggests a remarkable simplification: Assume that rj. is sufficiently close to 1/QS(T) (although still below it) for Sp/asT
K(r±.) « (rlQl(r))
(3.26)
300
E. Iancu and R.
Venugopalan
which shows geometric scaling 45 ' 46 ; it depends upon the two kinematical variables r± and r only via the combination T\Q2{T). Numerically, such a scaling behaviour has been seen in the solutions to the BK equation 65>68. Since Eq. (3.26) is the first term in an expansion in powers of 5p/ps, with ps = casr, this approximation is correct for 1 < In
J_
2n2(^ r2±Q2s(r)
< casr.
(3.27)
The condition on the left, r± < 1/QS(T), ensures we are still in a linear regime. For a dipole transverse resolution Q2 = 1/r^, this condition translates to the following scaling window 46 : Q2(r)
«
Q2 «
QUJ) ^ .
Ql
(3.28)
Since Q0 ~ AQCD and QS(T) » AQCD for sufficiently large r, the upper boundary of this scaling window is rather large. In particular, it is much larger than the saturation scale itself: Q2(T)/QO S> QS(T). Remarkably, as a consequence of saturation, knowledge of an intrinsic momentum scale is propagated through the linear evolution equations up to relatively large values of Q2 well outside the saturation regime. This is especially interesting since such values of Q2 are large enough for perturbation theory to be fully trustworthy. A property like (3.26) can be and has been tested against the experimental data 45>93>39. We shall return to phenomenological aspects of geometric scaling in Sect. 4.3. The previous results are obtained from the leading order BFKL equation. Recently, there has been some progress in including next-to-leadingorder a„ effects in the physics of saturation. In Refs. 46, 94, this was done heuristically, by simply replacing the fixed coupling as in the BFKL equation by the one-loop running coupling of QCD with the running scale set by the saturation momentum: as —• as(Q2(T)), with
^'-Ewte
-H3&-
(329)
'
The only modification due to the running coupling is in the functional form of the saturation scale, whose growth with T becomes somewhat milder (TO is an arbitrary constant, and c is the same number as in in Eq. (3.23)): Q2s(r) = A2QCD e^oc(T + r 0 ) ^
(3_30)
where the overall scale is now set by AQCD rather than the initial scale Qo- All the previous results on geometric scaling (the scaling law (3.26),
The Color Glass Condensate
and High Energy Scattering
in QCD
301
including the value of the "anomalous dimension" 7, and its range of validity (3.28)) remain unchanged, except for the expression for the saturation scale entering these results. For nuclei, Eq. (3.30) has an intriguing consequence for the dependence of the saturation scale upon the atomic number A 9 5 . Assume an initial condition of the MV type at r = 0, Q2S(T = 0,A) = Q2S(A) ~ A1'3 In A. For fixed coupling BFKL evolution, where Eq. (3.23) applies, this initial condition identifies the hitherto unspecified 'initial' scale QQ with the MV saturation scale, Q2S{T,A)
= Q2(A)ec&°T
(fixed coupling),
(3.31)
which preserves the A-dependence of the initial condition at any later 'time' r: Q2(T,A)~A1/3
In A.
For the running coupling BFKL evolution case where Eq. (3.30) applies (with ro fixed by the initial condition as 2boCTo = [III(Q2(A)/A'QCD)]2), one obtains a very different A-dependence at small and large r, respectively 95 : i) At relatively small energies, such that r -C ro ~ In2 A 1 / 3 , Q2(r, A) « Q2S (A) ec&°^^»T
,
(3.32)
which is the 'fixed-coupling'-like behaviour, with the as in the exponent being the running coupling (3.29) evaluated at the initial saturation scale, ii) At higher energies, T 3> In2 A1/3, one obtains Oj(r,A) - A ^ e ^ e x p j ^
^
^
}
}
-
^
which, for very large T, is nearly independent of A. In Ref. 44, which did not consider the A-dependence, Triantafyllopoulos presented a complete computation of the NLO effects on the energy dependence of the saturation scale. Recall that the NLO corrections to the BFKL equation 96 , turn out to be anomalously large and require resummation to obtain sensible results. Ref. 44 used the RG-improved resummation scheme of Ciafaloni, Colferai, and Salam 97 and found that, although Q2{T) is in general a more complicated function than the simple exponential (3.23), it can nevertheless be represented as such for a rather wide range of rapidities (including those of phenomenological interest). Specifically, if one defines A(T) = d\n(Q2/A2)/dr, then A(r) turns out to be a very slowly decreasing function, with A(r = 5 — 9) ~ 0.30 — 0.29. Remarkably, this value is also favoured by the current phenomenology at HERA 4 3 (and Sect. 4.3 below).
302
E. Iancu and R.
Venugopalan
T
/A2
4
Voii-liiic.ii
- Linear
Color Ciliiss / Aj!!e Condensate / ,. £•""
Parton Gas
*M ltl : KL A
-*• DGLAP •
InA Fig. 16.
InQ
lnkf
A map of the quantum evolution in the r — k±_ plane.
It would be very interesting to compute the yl-dependence within this fully NLO formalism. 3.5. Gluon saturation
and perturbative
color
neutrality
In this section, we return to the RGE (3.13), which describes the evolution of the hadron wavefunction as a whole, and construct approximate solutions to it. As usual, these approximations depend upon the transverse resolution scale Q2 at which correlations are measured. If Q2 is large enough (Q2 » Q2(T)), we probe color sources with small transverse size, which do not overlap with each other. In this dilute regime, a description in terms of uncorrelated sources, such as the MV model, may be a good approximation. With increasing T at fixed Q2 (or, equivalently, with decreasing Q2 at fixed T), spatial correlations start to develop, initially according to the linear BFKL evolution and then, once the density is high enough, according to the general non-linear RGE which predicts gluon saturation. A schematic map of the kinematical regimes for quantum evolution is shown in Fig. 16 (see the discussion below for details). i) High-momentum regime (Q2 » Q2S): Recovering BFKL In the dilute regime at Q2 > Q2, the color charge density is low, hence
The Color Glass Condensate
and High Energy Scattering
in QCD 303
the classical field is weak, g a < 1. One can therefore keep only the linear term in the expansion of the Wilson lines in Eq. (3.13) in powers of got: V^{x±) « 1 + ig dy ay(x±)
= 1 + iga(x±_),
1 - V}VX » ig{xj - xj)dja(x_L).
(3.34)
{aa(x±) is the effective color field in the transverse plane.) The kernel r)[a] in Eq. (3.14) then becomes quadratic in a(a;j_), and the RGE takes the generic form: dWT[a] 1 6 , .5 -B—H--(aKa)^Wr[a],
(3.35)
where the new kernel K. is non-local in the transverse coordinates. Its explicit form is easily extracted from Eq. (3.14) using Eq. (3.34). Even in this dilute regime, the RGE is non-linear and the corresponding weight function WT[a] is not a Gaussian. Nevertheless, compared to the general RGE (3.13), the evolution generated by Eq. (3.35) exhibits an important simplification: it does not mix correlations ( a ( l ) a ( 2 ) . . . a(n)) with different numbers n of fields 23 . Indeed, the quartic operator acting on W r [a] in the r.h.s. of Eq. (3.35) is formally the same as a the second-quantized Hamiltonian for a non-relativistic many-body system, and is diagonal in the number of "particles". Eq. (3.35) provides a closed evolution equation for the 2-point function, which is the BFKL equation. This equation is most commonly written for the charge-charge correlator fdf.(k±) oc (pa{k±)pa{—k±))T- In the linear regime, it is the same as the unintegrated gluon distribution (2.28). Specifically, if one defines (£{k±) as the Fourier transform of:
£(xuy±) = (pa( ^2 P _ a( r ))r > paM =Jdyp$M,
(3-36)
then Eqs. (2.29) and (2.28) imply, similar to Eq. (2.30), Vr(fcL) - ^P^
for
while from the RGE (3.35) one obtains d(J%(k±)
~d7~
_ [ cPp±.
=a
°j
22
k\
—pUki-p^bi.) 8
k± »
(3.37)
QS(T),
: ,
2
1
2
.
- -2 ^(fcx)) (3-38)
which is indeed the BFKL equation . Note that /^(fcj.) corresponds to p?A of the MV model, but, unlike the latter, it carries a non-trivial transverse momentum dependence, and a r
304
E. Iancu and R.
Venugopalan
dependence, both of which are obtained by solving Eq. (3.38). In fact, given the formal similitude between Eqs. (3.38) and (3.4), it is clear that the corresponding solution for p2(k±) can be obtained by replacing r^Qo —> ^i/*3o m Eq. (3.6). This shows the expected rapid exponential growth with T and infrared diffusion as well. ii) Low-momentum regime (Q2 -C Q2S): Saturation An external probe with low transverse resolution Q2 -C Q2 couples mostly to the saturated gluons, which have momenta k± £ Qs and occupation numbers ~ l/as. The corresponds classical fields are strong, ga(x±) ~ 1, so the Wilson lines (3.11) — which are complex exponentials built with these fields — oscillate around zero over a characteristic distance ~ 1/QS(T) in the transverse plane. This implies that Wilson lines which are separated by large distances 3> 1/QS(T) are necessarily uncorrelated (since their relative phases are random). Thus, when studying the dynamics over large transverse separations r± 3> 1/QS(T), it should be a good approximation to neglect the correlations of the Wilson lines (or, more generally, to treat them as small quantities). This is the "random phase approximation" (RPA) introduced in Refs. 16, 55. In this approximation, the RGE (3.8) simplifies drastically 16 . Neglecting the Wilson lines, the kernel 77 becomes independent of a, and the RGE reads in momentum space, dWT[a] 8T
=
1 [fk±
J _
2 J (2TT)
2
S2WT[p]
-Kk\ 6a«(k±)5a°(-k±)
Being quadratic, this equation can be immediately integrated r
WT
[p]«A/;exPj--y
{
'
'
16 20
'
Qs(y)
dy j
— ^ - j *
j , (3.40)
—00
which for convenience has been written as a functional of Py{k±) = k^a^kx). Eq. (3.40) is a low-momentum approximation: for a given rapidity y (with y < r ) . It is to be used only for modes k± < Qs(y). We see an interesting duality emerging at saturation: this strong field regime allows for a description in terms of a Gaussian weight function, like for a free theory. But even with such a Gaussian weight function, the CGC effective theory remains non-trivial, since the classical solution (2.16) and the observables for high energy scattering, like (2.49), are non-linear functionals of p.
The Color Glass Condensate
and High Energy Scattering
in QCD
305
Eq. (3.40) shows that the only non-trivial correlation of the color sources with small k± is the 2-point function, and reads: (pay(x±)pby,(y±))T
5ab6(y-y')e(T-y)\y(xj_,y±),
=
Ay(fcx) - - k\ ,
for fcx < Q,(y).
(3.41)
This distribution is local in (space-time) rapidity y, and homogeneous in all the (longitudinal and transverse) coordinates. In the transverse plane, it is only a function of the relative coordinate x± —y±, and in the longitudinal direction, it is independent of y. For a given k± <S QS{T), Eq. (3.41) applies only for y in the interval rs(k±) < y < r, with T3(k±) being the rapidity at which the saturation scale becomes equal to the momentum k± of interest: Q2S(T)
= kl
for
r = Ts(k±),
(3.42)
(see Fig. 16). It follows that the integrated quantity (cf. Eq. (3.36)): ^Msat =
/
* 7 = (
T
" T.(k±))&,
(k± «
QS(T)),
(3.43)
rs(k±)
which measures the density of saturated color sources (with given kx) in the transverse plane, grows only linearly with r. This should be contrasted with the exponential increase of the corresponding quantity at k± ^> Q s (r), obtained from the BFKL equation (3.38). (In Ref. 98, the result in Eq. (3.43) was obtained from a study of the BK equation.) We conclude that, at low momenta k±
Recall that Eq. (3.37) was obtained by using the linearized solution to the classical EOM, i.e., Eq. (2.29).
306
E. Iancu and R.
Venugopalan
(3.43) imply:
T-Ts(k±) •K
ncas
k^_
(3.44)
where in writing the second equality we have used Eq. (3.23) for the saturation scale 8 , together with the definition (3.42) of Ts(k±). A more careful calculation, based on the non-linear solution (2.16), shows that the correct answer for
Fig. 17. The gluon phase-density
In addition to T -saturation (the linear increase with T), Eq. (3.44) shows also k±-saturation — the k± spectrum is only logarithmic in l/k± at low momenta. Recall that, in the classical MV model, a similar spectrum emerged (Eq. (2.43)) only after fully taking into account the non-linear g
Recall that Eq. (3.23) corresponds to a fixed coupling as. In the case of a running coupling, where Eq. (3.30) applies, Eq. (3.44) remains formally the same, but as in the denominator must be understood as the running coupling (3.29) evaluated at Q 2 = k±Qs.
The Color Glass Condensate
and High Energy Scattering
in QCD
307
effects in the classical EOM (2.8). By contrast, in the quantum case, the non-linear effects responsible for saturation have already been encoded in the distribution of the color sources. This is natural since, as noticed at the end of Sect. 2, the difference between "sources" and "wee gluons" is a matter of convention and depends upon the resolution scale in k+. As anticipated, the gluon occupation factor at saturation, Eq. (3.44), is parametrically of order l / a s , as in the MV model. To conclude, the unintegrated gluon distribution predicted by the CGC effective theory is illustrated in Fig. 17, and looks qualitatively similar to that in the MV model, cf. Fig. 9. At large k± S> QS{T), the distribution in Fig. 17 is given by the solution to the BFKL equation, cf. Eqs. (3.37)(3.38), while at low k±_ -C Q s (r), it is given by Eq. (3.44). The saturation condition (3.21) can be also formulated in terms of the unintegrated gluon distribution as Mk±)
~ ^-
for
k±~Qs(T).
(3.45)
This condition, together with BFKL evolution at higher momenta, implies that, for k± within the scaling window (3.28), the unintegrated gluon distribution has a scaling form similar to Eq. (3.26) 69 :
where K is a yet undetermined numerical prefactor. We see that the scaling property characteristic of saturation (cf. Eq. (3.44)) is preserved by the linear BFKL evolution up to a relatively large moemtum k± ~ Q^(T)/AQCD (cf. Eq. (3.28)), which is well above the saturation scale. This "extended scaling" region 46 where the gluon density is relatively low, but takes the scaling form (3.46), is represented on the diagram in Fig. 16. iii) Color neutrality at saturation The vanishing of the charge-charge correlator (see (3.43)), as k\ when fc_L —* 0 has important consequences for the infrared behaviour of the CGC effective theory. Consider, the calculation of the dipole-hadron 5-matrix element. In the MV model, it was found to be logarithmically infrared divergent, cf. Eqs. (2.38) or (2.51). If in Eq. (2.37) one replaces the MV estimate for the charge correlator A^ by the corresponding quantum expression, the ensuing integral over k± becomes infrared finite, due to Eq. (3.41). When computed in the quantum effective theory, the dipole scattering amplitude is infrared safe, and therefore insensitive to the non-perturbative physics of
308
E. Iancu and R.
Venugopalan
confinement. A similar property holds for other gauge-invariant quantities like the gluon distribution. Furthermore, the smooth infrared behaviour in Eq. (3.41) is responsible for fcj_-saturation in the gluon distribution. In what follows, we shall argue that this behaviour has a simple physical interpretation-gluonic color sources are correlated over long distances to ensure that color neutrality is achieved over a transverse area of order 1 / Q 2 ( T ) 20-98>69. Consider the total color charge Qa enclosed within a surface E, as given by Eq. (2.4) with AS± —> E. This is a random quantity with zero average (since (pa(x)) = 0 at any point x), so we shall compute the average of the color charge squared Q 2 = QaQa. We have, (Q2)r = (N2 - 1) [d2x±
fd2y±tiT(x±,y±).
(3.47)
In the MV model, where the sources are uncorrelated (cf. Eq. (2.5)), we have (Q2)A
= (N2 - 1 ) E , £ ~ -
EQ 2 A ,
(3.48)
which increases rapidly with A, like A1/3. After including quantum evolution, the charge correlator acquires a nontrivial momentum dependence {p?A —• n2(k±)), and Eq. (3.47) can be estimated as (up to a color factor): (Q 2 ) T ~ £ / £ ( * £ ~ 1/E).
(3.49)
For a relatively small area, or moderately high energies, E _ 1 » Q^(T), and nl(k±) is given by the BFKL equation (3.38). Then Eqs. (3.37) and (3.46) imply: (Q2)T ~ -
(SQ2(r))7
for
1/E»Q2(r),
(3.50)
which shows incomplete color shielding. With increasing E, the total charge squared enclosed within this surface increases, but not as fast as the area itself. Thus, the density (Q 2 )/E ~ 1 / E 1 - 7 vanishes in the limit E —> oo, which is consistent with (global) charge conservation. Furthermore, for fixed E, the total charge (Q2)T increases exponentially with r. For larger surfaces, of the order of the saturation disk 1/Q 2 (r) or larger, one should rather us Eq. (3.43), which gives: (Q2)T ~ — In (EQ 2 (r))
for
1/E«Q2(r).
(3.51)
The Color Glass Condensate
and High Energy Scattering
in QCD
309
The total charge squared (3.51) is still non-zero, but unlike Eq. (3.50) it is only logarithmically increasing with both 1/x and E. This is complete shielding: When increasing E, the newly included color sources are completely screened by the other sources. If the total charge (squared) is still increasing with E (albeit only slowly), it is because the longitudinal width r — T S ( E ) ~ ln(EQg(r)) of the "saturated" piece of the hadron rises logarithmically with E. This complete shielding, together with the fact that the total charge (3.51) is much smaller than the total charge for a system of uncorrec t e d color sources with surface density Q2s/a3, cf. Eq. (3.48), enables us to speak about color neutrality already at the relatively short scale l/Qs( T ) ^ 1/AQCD- This interpretation is further confirmed by the fact that the color field created by gluon sources at large distances 3> 1/QS{T) cannot be distinguished from a dipolar field 69 . 3.6. A Gaussian
effective
theory
Consider the calculation of the dipole-hadron S'-matrix element (2.49) within the CGC effective theory. After expanding the Wilson lines in Eq. (2.49), one is led to evaluate n-point functions of the type gn(a(zi)a(z2) • • -a:(z„)) r , where the transverse arguments Zj_,t are either X.L or y±. Each such a n-point function receives contributions from either hard (k± > QS(T)) or semi-hard h (AQCD < fcx < QS(T)) momenta. The contributions of the modes with k±
310
E. Iancu and R.
Venugopalan
These considerations can be extended to any quantity which is not very sensitive to the momenta in the transition regime around QS(T). To systematically compute such quantities, construct a Gaussian approximation to the weight function which encodes the limiting behavior of the the 2point function at high and low momenta and which interpolates smoothly between these regimes. Such a Gaussian will be as simple to use as the original MV model, but will extend the latter by including the BFKL evolution at high momenta, and the physics of saturation and color neutrality at low momenta. Such a Gaussian weight function has been constructed in Ref. 69, and reads:
where the kernel Xy(x±,y±) — the 2-point function of the color sources (cf. Eq. (3.41)) — is such that its Fourier transform Ay(/cj_) satisfies the BFKL equation at momenta k± 3> QS{T) and reduces to Eq. (3.41) for momenta kj_
~ A?FKL(fc±=Qs(r)).
(3.53)
Then, the following function provides a smooth interpolation, that we shall use in Eq. (3.52):
W,
(3 54)
^W'
'
For momenta within the scaling window (3.28), the BFKL solution takes the scaling form (with 7 = 0.64): A ^
L
( ^ ) ^ i ( ^ ) \
which allows us to write a simple explicit expression for the kernel valid for all momenta k± & Q^(T)/AQCD-
(3.55) Xy(k±),
e (W In Sect. 4.3 , Eq. (3.56) will be used to compute the dipole-hadron scattering amplitude. Here, we shall present a different application, namely we
The Color Glass Condensate
•
1
and High Energy Scattering
1
•
,
in QCD 311
,
T TQ T = T0+ 2
1000
T=TQ + 4
T= ^LT:";"'*'-M'.-.L'. 'J '.•
100
TQ
+ 6
T = TQ + 8
—•——-
S ^ ^ f c ^ ^ — • • -
5 10
^V
\
\ ^ \
N.
\ N.
*
0.1 0.01
\ ^V
V
'••• N. •'••• \ . *** "**• ^
^\. >w *\ ^v *s
. . .
0.1
.
1
.
.
. \
10
.
\
100
. ':••
1000
kx/Qs(%> Fig. 18. Energy and momentum dependence of ipT(k±). We have plotted yv(fcx) as a function of k±/Qs(ro) (with TQ some value of reference) for six values of T. The lines, from the bottom to the top, correspond successively to r = TO, TO + 2, • • •, TO + 10. The increase with T is exponential at high momenta (giving equidistant curves in this log-log plot), but only logarithmic at low momenta. From Ref. 69.
shall deduce a simple analytic expression for the unintegrated gluon distribution. Specifically, Eqs. (3.36) and (3.41) imply ^T{k±_) = J dyXy(k±). When further combined with Eqs. (3.37) and (3.56), it leads to the following final result:
~ k\j dy\ {k±.)
y
TT~/cas
In 1 +
QM
(3.57)
(In performing the integral over y, we assumed Eq. (3.23) for the saturation scale.) Eq. (3.57) interpolates smoothly between Eq. (3.44) deeply at saturation (k± -C Qs) and Eq. (3.46) for momenta within the scaling window Qs & k± 5i Q S / A Q C D - In Fig. 18, we illustrate both the kj_-dependence and the r-dependence of the function (3.57), which is plotted as a function of k\_ for several values of T. 4. Deep Inelastic Scattering and the CGC In this section, we shall discuss some applications of the Color Glass Condensate picture to Deep Inelastic Scattering. We start by explicitly deriving a factorization formula introduced in Sect. 2.7, which allows one to compute
312
E. Iancu and R.
Venugopalan
the F2 structure function at high energy in terms of the dipole scattering. We shall then discuss a phenomenological "saturation model", proposed by Golec-Biernat and Wiisthoff, which is based on similar physical premises, and compares remarably well with the HERA data. This model will be further reconsidered from the CGC perspective, in which the dipole-hadron scattering is computed within QCD. This allows us to relate the "geometric scaling" observed in the HERA data to properties of the quantum evolution towards saturation. In addition, we discuss the Froissart bound can be realized conceptually in the Color Glass Condensate. We discuss next the relation of saturation to shadowing in DIS. Finally, we discuss inclusive and semi-inclusive signatures of the Color Glass Condensate in Deep Inelastic Scattering and specifically how they may be observed at a future Electron Ion Collider. 4 . 1 . Structure
Functions
in the Color Glass
Condensate
We shall show here how one computes structure functions for DIS in the presence of non-linear effects associated with saturation 59 . Towards this end, one needs the correlator of the electromagnetic current in the background of the strong classical gluon field representing the CGC. In DIS, the interaction between the hadron and the virtual photon is encoded in the following tensor expressed in terms of the forward Compton scattering amplitude TM„ 6 : W " V , P -q) = 2 Disc T^iq2,
P • q)
= -!-Im f dAxeiqx
(4.1)
J
27T
where "T" denotes a time-ordered product, J7* = {jj^ip is the hadron electromagnetic current and "Disc" denotes the discontinuity of T^ along its branch cuts in the variable P • q. Also, q2 < 0 is the transferred momentum squared (i.e., q^ is the momentum of the virtual photon, and Q2 = —q2) and P M is the momentum of the target. In the IMF, P+ —* 00 is the only large component of the momentum. Since < T{J»{x)r{y))
>=< T
ftx)7^(x)«!,)7^(!/))
> ,
(4.2)
the time ordered produced of currents can be expressed, in complete generality, as < T(J»(x)J»(y)) = Tr(rGA(x))Tr(YGA(y))
> = + Tv(^GA(x,
y)YGA{y,
x)),
(4.3)
The Color Glass Condensate
and High Energy Scattering
in QCD
313
where GA(X,V) = —i < ip{x)ip(y) >A is the quark Green's function in the external background gauge field A% of the hadron. The first term on the right hand side of Eq. (4.3) is a tadpole contribution without an imaginary part. It therefore does not contribute to H"" 1 . We are thus left with:
W^{q2,P-q) = ^ - ^ I m f d3X I (Pxj"* 27T M J (Tr(^GA(X + x/2,X-
J
x/2)jvGA{X
(4.4) - x/2, X + x/2))) .
The approximation one makes here is to replace the full background gauge field A^ by the classical background field -^1^.classical- ^ n other words, the Green's function which, in general, is computed in the full background field of the nucleus, is now computed in the saddle-point approximation where A^ —• ^ i C i a s s j c a [. Note that this expression makes no reference to the operator product expansion of DIS 6 . Thus, it is also valid at small values of x and moderate Q2, where the operator product expansion is not reliable 70 but where the classical approximation is sensible. Since the current-current correlator is gauge invariant, one can compute it by using background field propagator in an arbitrary gauge. For the reasons explained in Sect. 2.4, it is most convenient to use the covariant gauge, in which the propagator reads as follows 71>52'59; GA(X, y) = Go(x - y) - i / d4z G0{x - z)j~S(z~)G0(x {0(x-)9(-y-){V\zx)
- 1) - 9(-x-)e(y-)(V(z±)
- z) - 1)} , (4.5)
where Go is the free propagator and V and V^ are the Wilson lines of Eq. (2.24). This is obtained by assuming the color source to be a 5-function in a:~, which is appropriate since this source is due to relatively fast partons with longitudinal momentum fractions much larger than the Bjorken x = Q2/2P • q of the collision. Inserting the fermion propagator (4.5) in Eq. (4.4) and performing the integrations there, one obtains the final result for W^, which is conventionally expressed (for an unpolarized target and for Q2 « Mjy) in terms of two structure functions Fi and F2, denned by 6 :
W " =
-(9^v-^4-)Fi q2 '
H>»-£ip>){r-£!$<>)»
<«,
314
E. lancu and R.
Venugopalan
Consider first F 2 : the corresponding result takes the factorized structure anticipated in Eqs. (2.46)-(2.48), where |*(z,rj.)| 2 , which signifies the probability that the virtual photon splits into a qq-p&ii can be expressed as \®(z,r±)\2 = \^T(z,r±)\2 + \^L(z,r±)\2, and where each of the terms has the explicit form \*r(z,r±)\2
= f^
E
e
/ i ( * 2 + (! - z?)Q2fKl{Q}r)
\*L(z,r±)\2
= § f
X > / {4QV(l-z)2tf0(Q/r)} .
+ m2K2(Qfr)}
, (4.7)
Above, the sum runs over the quark flavours, Q2, = z(l — z)Q2 + raj, rrif is the quark mass, and K\ are modified Bessel functions. \^T(z,r±)\2 2 {\^L{Z,T±_)\ ) denotes the probability that a transversely (longitudinally) polarized photon splits into a qq-p&h. This decomposition implies a similar decomposition for F 2 , namely, F 2 = FT + F L . F I , this is proportional to FT — explicitly, FT = 2xF\ and the longitudinal structure function is F t = F2 — 2xF\. In the parton model, F t = 0 — the Callan-Gross relation — but is non-zero in QCD and is directly proportional to the gluon distribution. An independent measurement of FL is therefore of great phenomenological interest as will be discussed in Sect. 4.6. One can similarly derive expressions for the diffractive structure functions as discussed in Ref. 109. Some of the phenomenological implications of the CGC picture for measurements of structure functions were discussed in Sects. 2.6 and 2.7. 4.2. The Golec-Biernat-Wusthoff
model
In Ref. 43, Golec-Biernat and Wusthoff introduced a simple phenomenological model for the dipole-hadron cross-section, Eq. (2.48), which is generally referred to as the "Saturation Model": cTdiPoie(x,r±) = a 0 (1 - e - ^ W
4
)
(4.8)
with the parametrization Q2(x) = <5Q( X O/ X ) A - This shows color transparency at low r±
The Color Glass Condensate
and High Energy Scattering
in QCD
315
as a "saturation momentum". Note however that the approach towards the unitarity bound AfT(r±) = 1 for large r± is much faster then predicted by the CGC approach, or the solution to the BK equation (see Sect. 4.3 below). Besides, at small r±, Eq. (4.8) fails to reproduce, as it should, the leading-twist approximation — it misses the logarithmic factor In (1/r^A 2 ) in the exponent of Eqs. (2.51). Neither does it recover the BFKL prediction (3.6). By using the dipole cross-section (4.8) and the factorization formula (2.47) for DIS, Golec-Biernat and Wiisthoff were able to fit the HERA data for cr7«p for x < 10~ 2 and a wide range in Q2 with only three parameters, namely <7o = 23 millibarns, A = 0.288 and xo = 3.04 • 1 0 - 4 . (The reference scale Q2 n a s been fixed as Q2, = 1 GeV 2 .) These fits were performed for three light quark flavors; with the addition of charm quarks, the best fit was obtained with slightly changed parameters: CTQ = 29.1mb, A = 0.277, and xo = 0.4 • 1 0 - 4 . The fits are reasonable for Q2 up to Q2 ~ 20 GeV 2 , but are less successful beyond. This is related to the above observation that Eq. (4.8) does not have the right perturbative behaviour at small r±. By replacing Eq. (4.8) with a Glauber-type formula as shown in Eq. (2.53), the situation at high Q2 improves considerably and a wider range in Q2 can be fit within the framework of this model 73 . For other phenomenological models of the HERA data based on ideas of saturation see Refs. 74, 75. The diffractive structure function F^ (x-p, Q2, (3) (where x-p = (M2 + Q2)/(W2 + Q2) and /? = Q2/(M2 + Q2) where M is the diffractive mass and W is the total energy of the virtual photon-proton process) was measured at HERA and several striking properties of the diffractive structure function were observed. For instance, the ratio of adlff /a is large and is nearly independent of W, a feature that was not anticipated in pQCD based models. In Ref. 43 , the simple model that describes the inclusive scattering data also describes the diffractive structure function data. Interestingly, the form of the diffractive cross-section in this model is similar to the inclusive one except the dipole-hadron cross-section appearing in the latter is replaced by the square of this cross-section '. The parameters appearing in the fit of the inclusive cross-section to the data are therefore the same as those used in the diffractive fit. The agreement with data is quite impressive given these constraints. A version of the model has also been applied to study vector meson production with reasonable results 110>76. 'This feature of the model can be understood very simply in the CGC picture
316
E. Iancu and R.
Venugopalan
Whereas the behaviour at high Q2 can be improved rather easily, by replacing Eq. (4.8) by Eq. (2.53) as mentioned above, the impact parameter dependence is problematic. This is expected to be especially important in a proton. Data on the t-dependence of differential cross-sections, particularly for vector meson photoproduction, may be useful in understanding parton distributions in impact parameter space 76>77.
4.3. Geometric
Scaling in
DIS
The Golec-Biernat-Wusthoff dipole cross-section, Eq. (4.8), has the remarkable feature to depends upon the two kinematical variables x and r± only via the dimensionless combination (the "scaling variable") T = r^_Q2(x). Via the factorization formula (2.47), this scaling property transmits to the virtual photon total cross-section tT7.p which, in the limit where the quark masses are negligible, becomes a function of the ratio Q2/Q2(x) alone (a property usually referred to as "geometric scaling"). At a first sight, this may appear an artifact of the simple parametrization (4.8) specific to the saturation model. Indeed, while the scaling looks natural at saturation one would naively expect this scaling to be broken after generalizing Eq. (4.8) to reproduce the perturbative behaviour at high Q2 (e.g., the inclusion of the logarithm l n ( l / r ^ A 2 ) as in Eq. (2.51) would clearly violate scaling). It thus appeared as a surprise when Stasto, Golec-Biernat and Kwiecihski showed 4 5 that, to a rather good accuracy, the HERA data on cr7.p do show scaling for small enough x (x < 0.01) and all Q2 up to 450GeV2 (see Fig. 19). Such Q2 are significantly higher than the estimated value of the saturation scale at HERA, as extracted from the "saturation model" fits to F 2 43 : Q 2 ~ 1 • • • 2 GeV 2 . On the other hand, the data show no scaling for larger values of x. Subsequently, some indications of geometric scaling have been found also in the data for DIS off nuclei 9 3 , and even in the particle production at RHIC 38 , although, in these cases, the experimental evidence is more uncertain. In the light of the previous discussion in Sect. 3, such a scaling behaviour should not look surprising any more. As explained in Sects. 3.4 (for the scattering amplitude) and 3.5 (for the gluon distribution), the scaling property at saturation is preserved by the BFKL evolution up to relatively large Q2, of order <2f ( T ) / A Q C D (cf. Eq. (3.28)). If one uses the phenomenological values of Qs alluded to before, and AQCD = 0.2 • • • 0.3GeV, one finds that the maximum Q2 up to which scaling is expected is indeed of the order
The Color Glass Condensate
and High Energy Scattering
in QCD
317
ZEUS BPT 97 ZEUS BPC 95 HI low Q ! 95 ZEUS+H1 high Q2 94-95 E665 x<0.01 allQ 2
Fig. 19. HERA data on the cross section for 7*p DIS from the region x < 0.01 and Q2 < 400GeV 2 plotted versus the scaling variable T = Q 2 / Q o ( x ) (from Ref. 45).
of a few hundred GeV 2 , as observed at HERA 45 . Within the CGC formalism, the scaling properties of the dipole-hadron scattering amplitude can be explicitly studied by using the Gaussian approximation to the weight function introduced in Sect. 3.6. The kernel of the Gaussian shows explicit scaling for all momenta kj_ ^ Q^(T)/AQCD, cf. Eq. (3.56), and therefore so does also the ^-matrix element computed in this approximation. Specifically, a straightforward calculation yields 6 9 (in the case of a fixed coupling, cf. Eq. (3.23), for definiteness, and with K=
ATTCF/JCNC):
[ f d2kj_ 1 - eik-L'r± ST{r±) = exp < -K / —^TT r^ 2 (2TT)
*i
In 1 +
Q2s(r)
*i
(4.9)
Unlike the corresponding prediction of the MV model, Eq. (2.51), which was infrared sensitive, and thus dependent upon the non-perturbative scale AQCD , the integral in the equation above is well behaved both in the infrared
318
E. Iancu and R.
Venugopalan
and in the ultraviolet, so, clearly, the result is a scaling function: ST(r±) = S(r±Q3(r)). The limiting behaviour of Eq. (4.9) at either small, or large, transverse size r± as compared to the saturation length 1/QS(T) are 69 : i) For a small dipole, r±
*M*>£k
ln
g | ( ^ 2 - + (1 + ^ ( 2 ) + 2 1 n 2 ) . (4.10)
This is similar to the small-rj. behaviour in the MV model, Eq. (2.52), except that the infrared cutoff in the logarithm is now the saturation momentum, and not AQCD-
i) For a large dipole, r± ^> 1/QS(T) (but r± -C 1/AQCD), the dominant contribution comes rather from momenta k± in the range 1/rj. -C k± <§; QS(T), that is, from scattering off saturated gluon sources, and reads: 5T(rx)cxexp|-^-(lnriQ2(r))2| ,
(4.11)
in agreement with the results in Refs. 65, 16, 19. As anticipated, the approach towards the "black disk" limit ST = 0 predicted by the CGC is slower than that assumed in the saturation model, Eq. (4.8). 4.4. The Froissart
Bound for dipole
scattering
Let us now address the fundamental question of the asymptotic behaviour of the total cross-section at very high energies. Thus far, in the analysis of the dipole-hadron scattering, we have neglected the impact parameter dependence. We have assumed the hadron to be a homogeneous disk with radius R. Then all the theoretical descriptions which include saturation, from the phenomenological "saturation model" in Eq. (4.8) to the QCDbased formalisms like the CGC or the Balitsky-Kovchegov equation, lead a dipolar cross-section which approaches a constant value
The Color Glass Condensate
and High Energy Scattering
in QCD
319
be right. From experiments, we know indeed that hadronic cross-sections (e.g., for pp collisions) keep growing with s up to the highest energies that have been reached so far. At very high energy, the growth is rather slow, and can be fitted by either a small power of s (the "soft pomeron" ~ s 0 0 8 ) , or some power of In s, or a combination of them. A recent analysis 81 of the data for several high-energy processes appears to favour a dominant behaviour of the log-squared type: <7t0t(s) — coin s, with an universal prefactor GQ. If true, this would imply that the Froissart bound is actually saturated in nature. Such a steady growth of total cross-sections is also expected on physical grounds 78>79>80. As a quantum mechanical bound state, the hadron does not have a sharp edge but rather a tail. In a theory with a mass gap like QCD, this tail is an exponential whose width is fixed by the lowest mass in QCD, namely the pion mass. Impact parameters which, at some initial energy, are far away in the tail of the distribution will not contribute much to scattering, since the parton density is low there. But with increasing energy, the local gluon density will increase rather fast as a power of s, since gluon recombination is not effective when the density is low. Eventually, for sufficiently large s, the local gluon density will become high enough for these impact parameters to contribute significantly to scattering. That is, with increasing energy, the effective interaction radius of the hadron is expected to grow as well, which then results in an increase of the total cross-section. In fact, since the local scattering amplitude MT(r±, &x) cannot exceed the unitarity, or "black", limit AfT = 1, it is clear that, for sufficiently large energies, the increase of the cross-section with s will proceed via the expansion of the "black disk" (= the central area of the hadron where the unitarity limit has been reached already). This general discussion shows that a theoretical description of the dynamics of the black disk must combine two essential ingredients: i) a mechanism which ensures the unitarization of the scattering amplitude at fixed impact parameter, and ii) a description of the tail of the hadron wavefunction. In QCD, the second issue is certainly related to confinement, and is thus genuinely non-perturbative. But it has been unclear until recently whether the first issue, that of the unitarization, can be addressed in perturbation theory or not. Indeed, since gauge interactions are a priori long-ranged, it could well be that soft, non-perturbative, interactions are responsible for the approach towards "blackness" at a fixed impact parameter. The "infrared diffusion" of the BFKL equation may be seen as in argument in that sense.
320
E. Iancu and R.
Venugopalan
V Y*
f b
Q2= 1/r R(Q 2 ) Fig. 20. tions.
RH
b
Dipole-hadron scattering in longitudinal (left) and transverse right) projec-
However, our present understanding of the physics of high parton densities shows that the phenomenon of gluon saturation provides a perturbative framework for the study of unitarization. We have seen indeed that the perturbative, but non-linear, evolution equations yield a scattering amplitude which respects the unitarity bound (see Fig. 15 and Eq. (4.9)). To study the expansion of the black disk, these equations must be supplemented with some information about confinement and applied at impact parameters in the tail of hadron, namely, in the "grey area" at bj_ > R(T,Q2). Here, R(T,Q2) is the radius of the black disk for an incoming dipole with transverse resolution Q2 = \jr\ and relative rapidity r = In s/Q2. That is, AfT(r_i,b±) is of order one for b± < R(T, Q2), but it drops rapidly for 6j_ > R(T,Q2). Equivalently, R(T,Q2) is such that the local saturation scale QS(T, b±) — which is largest towards the center, and decreases with b±, so like the gluon density — becomes equal to Q2 at b\_ ~ R(T, Q2) (see Fig. 20). Perturbation theory is valid as long as Qs(T,b_i) 3> AQCD, that is, for b± < RH(T) in the plot in Fig. 20. This condition leaves us with a perturbative grey corona at R(T,Q2) < b± < RH(T), which is sufficient, as we shall see, to perform a controllable calculation of the expansion rate of the black disk. Specifically, for impact parameters within this corona, the dipole undergoes mostly hard scattering, with transferred momenta k\ > Q2S(T, & I ) (cf. the discussion after Eq. (4.9)). This implies that it interacts predominantly
The Color Glass Condensate and High Energy Scattering in QCD
321
with color sources which are relatively close in impact parameter space, within a saturation disk of radius 1/Q3(T, b±) centered at b±. Physically, it is so because of the screening phenomena associated with saturation (cf. Sect. 3.5): Color sources which lie further away create only dipolar fields at the impact parameter of the incoming dipole; such fields decrease rapidly with distance (more rapidly then the monopole fields due to the nearby color sources), and thus contribute less to scattering 80>69, Mathematically, this means that scattering in the grey corona is controlled by the linear BFKL equation (3.4), but with an infrared cutoff of order QS{T, b±). This cutoff simulates the non-linear terms, which, in the full equation (3.7), would limit the range of the dominant interactions to 1/QS(T, b±). Clearly, with such a cutoff, the BFKL equation is not afflicted by "infrared diffusion" any more. As in the previous calculation of the saturation scale in Sect. 3.4, the black disk radius can be computed by first solving the BFKL equation, and then imposing a saturation condition similar to Eq. (3.21): M-(r±, b±) ~ 1
for
b ~ R{T,Q2).
(4.12)
The bx-dependence of the initial condition is determined by the nonperturbative physics of the confinement, and thus requires a model. However, what we need to know about confinement is quite limited, and can be inferred from general principles. Firstly, what is the typical scale for transverse inhomogeneity in the hadron: this is clearly 1/AQCD- Secondly, what is the 6j_-dependence of the scattering amplitude in the hadron tail. This is an exponential fall-off oc e _ 2 m " 6 where twice the pion mass enters since the long range interactions between the dipole and the (isosinglet) gluons require the exchange of at least one pair of pions (recall that pions have isospin one). The first requirement tells us that the inhomogeneity occurs over transverse scales much larger than the typical range of the interactions, ^ 1 / Q S ( T , b±). The scattering of the small dipole proceeds quasilocally in the impact parameter space-1. As a consequence, the 6j_-dependence of the solution to the (effective) BFKL equation factorizes, and is fixed by the initial condition 8 0 : K(r±,&L)
« K(r±)S(b±),
(4.13)
JThis would not be true in the absence of the infrared cutoff generated by the non-linear effects.
322
E. Iancu and R.
Venugopalan
where AfT(rj_) is the solution to the homogeneous BFKL equation, as shown in Eq. (3.6). The second condition provides the transverse profile function in the hadron tail: S(bj_) oc e~2m"b (the proportionality coefficient is not important to the present accuracy). The previous arguments, together with Eq. (3.6) lead to the following estimate for the scattering amplitude valid at impact parameters in the grey corona k (with Q2 = 1/r^): K{Q2,
&x)| grey - exp 1 - 2 ^ 6 + w a , T - | I n ^ | .
(4.14)
This equation, together with the blackness condition (4.12), imply:
*'•«">-sb^-i
ta
&)-
(4 15)
-
We see that the black disk radius increases linearly with r. This behaviour ensures the compensation between the exponential increase with r (due to the perturbative BFKL evolution) and the exponential decrease with b± (due to confinement). Since MT is rapidly decreasing at b± » R(T, Q2), the total cross-section is dominated by the black disk:
2TTR2(T,Q2).
(4.16)
Together with Eq. (4.15), this yields the following dominant 1 behaviour of the cross-section at high energies 80 (see also Ref. 79 for an early study): 0-diPoie(s,<32) « 7T ( — - )
ln 2 s
as s —> oo.
(4.17)
This saturates the Froissart bound, with a universal coefficient for all hadrons and reflects the combined role of perturbative and non-perturbative physics in controlling the asymptotic behaviour at high energy. This behaviour is in qualitative agreement with the phenomenological analysis in Ref. 81, but the coefficient in Eq. (4.17) is too large to fit the data. The difference with respect to the data can be substantially reduced by using the RG-improved NLO estimate for the BFKL intercept, which decreases the leading-order value uas by roughly a factor of three 96 - 97 . In Refs. 82, the result in Eq. (4.17) has been generalized to 7* — 7* scattering. k
At very high energies, we can also neglect the diffusion term in Eq. (3.6). 'The subdominant behaviour is not completely described by Eq. (4.15), since it also receives contributions from the grey area 8 0 .
The Color Glass Condensate
and High Energy Scattering
in QCD 323
From Eq. (4.14), one can also estimate the radius RH(T) where the saturation scale decreases to AQCD (beyond which perturbation theory fails to apply). This is the same as the black disk radius for a large dipole with Q2 ~ A Q C D . Therefore
(418)
*»«"•£'•
and the radial extent of the perturbative corona can be estimated as: RH(r)
- R(T, Q2) « _ 1 _ l n | _
>
(419)
This is independent of r, and much larger than l / m x (because of the large logarithm ln(Q 2 /A 2 )), which demonstrates the self-consistency of the previous calculation: When T —» r + dr with aad,T ~ 1 (the typical rapidity increment at high energy), the black disk expands from R(T,Q2) to R(T,Q2) + w / 2 m T
The perturbative evolution equation Eq. (3.7) should not be used for very large impact parameters b± 3> RH(T), where Q 2 ( T , b) ^ ^QCD- Such equations lack confinement, so the long-range dipolar fields generated by the gluons within the black disk can propagate to arbitrarily large distances, thereby creating power-law contributions to the hadron tail. At impact parameters within the grey corona, these long-range contributions are relatively small, and the evolution is driven by the short-range interactions, with the conclusions outlined above. But for sufficiently large impact parameters, well beyond RH(T), the power-law contributions will eventually supersede the exponentially decreasing one, and the hadron will develop an unphysical power-law tail. (This is unphysical since, in the real world, it is removed by confinement.) If one pushes the perturbative expansion until the black disk enters this power-law tail, then its expansion rate speeds up, and violates the Froissart bound 83 . This is clearly an artifact of using perturbation theory outside its range of validity. The only way to circumvent this difficulty without introducing ad-hoc modifications in the evolution equations (to account for confinement) is to follow the perturbative evolution for only a limited interval of "time" A T , such that perturbation theory remains valid. As explained above, a limited evolution in time is indeed sufficient for the calculation of the expansion rate of the black disk. The result
324
E. Iancu and R.
Venugopalan
of this calculation can be then extrapolated to arbitrarily high energies, even though the perturbative evolution eventually becomes meaningless. To conclude, while perturbation theory alone appears to be sufficient to describe unitarization at fixed impact parameter, one still needs some information about the finite range of the strong interactions in order to compute total cross-sections. This is reminiscent of an old argument by Heisenberg 78 which combines unitarity and short-range interactions (as modelled by a Yukawa potential) to derive cross-sections which saturate the Proissart bound. Fifty years later, our progress in understanding strong interactions allows us to confirm Heisenberg's intuition, and identify shortrange interactions with confinement, and unitarization with saturation.
4.5. Saturation
and Shadowing
in Deep Inelastic 2
Scattering 2
Shadowing is the phenomenon where F^(x,Q )/AF2 (x,Q ) < 1 at small x (x < 0.1). It is a large effect in the region where the coherence length of the probe (the gq-pair in DIS) lcoh ~ 2m x ex ceeds the intra-nuclear longitudinal distance between any two nucleons in the nucleus. The nuclear parton distribution is not merely the sum of nucleon parton distributions but also contains the intereference between the parton distributions of the nucleons. When the coherence length is larger than the nuclear diameter (lcoh 3> 2A 1//3 , or x « l/(4mjv.<41/'3)), the qq-pa.ii interacts coherently with the entire nucleus, and the collective effects are expected to be imporatant. There are several questions about shadowing that have not been unambiguously resolved in the framework of QCD. • Is shadowing a "leadig twist" effect, or it is suppressed by powers of Q2? An empirical answer to this question would help settle whether shadowing is an intrinsically leading twist phenomenon 8 4 or whether it is due to weak coupling, higher twist/high parton density effects n>12>14. • What is the relation of shadowing to parton saturation? Does parton saturation provide a microscopic understanding of shadowing? • Does the shadowing ratio "saturate" at a minimum value for fixed Q2 and A with decreasing x? Does it saturate faster for quarks or gluons? • What is the relation of shadowing in nuclei to diffractive scattering of nucleons? The relation is well established at low parton densities 85 . In an interesting recent exercise, it has been shown that diffractive nucleon data at HERA could be used to predict the shadowing of quark distributions observed by NMC 86>108. Significant deviations from the simple relation be-
The Color Glass Condensate
and High Energy Scattering
in QCD
325
tween shadowing and diffraction, may again suggest the presence of strong non-linearities. • Is shadowing universal? For instance, are gluon structure functions extracted from p-A collisions at RHIC identical to those extracted from e-A in the same kinematic region? The naive assumption that this is true may be incorrect if higher twist effects are important. The answers to these questions will only be conclusively settled by the next generation of p-A collider experiments at RHIC and LHC 87 and by e-A collider experiments at RHIC (EIC/eRHIC) and/or DESY (HERA III)-see the following sub-section for a discussion of e-A collider plans. From the theoretical perspective, there are several shadowing models which consider only leading twist shadowing 108 ' 88 > 89 . Namely, shadowing effects are put in the non-perturbative initial conditions, which are then evolved in Q2 and x using the leading twist DGLAP equations 10 . Modifications of these leading twist models to include Mueller-Qiu type non-linear contributions have also been considered 90 . In the McLerran-Venugopalan saturation model, the gluon distribution (defined as the integral over the unintegrated gluon distribution) is additive in A. However, the structure functions computed as discussed in section 4.1 will exhibit shadowing for Q2 < Q2. There have been a few attempts to compute shadowing from the non-linear renormalization group equation. One of these calculations, which is based on the non-linear equation derived in Ref. 54, predicts that perturbative gluon shadowing will become large as one goes to smaller XBJ'S 9 1 . Other more recent computations of shadowing have been performed within the context of numerical solutions of the Balitsky-Kovchegov equation for both inclusive 6 7 , 9 3 and diffractive scattering 92 . In both cases, geometric scaling of the nuclear distributions is claimed but it is not clear that both groups obtain the same A-dependence for the saturation scale. A very interesting recent theoretical suggestion is that the scattering amplitude for high energy scattering for nuclei, at fixed impact parameter, is the same as for protons at asymptotic energies! 9 4 . Integrated over impact parameter, this would suggest that shadowing saturates at very small x. The understanding of shadowing in the saturation picture is still preliminary-more detailed global fits of the non-linear equations (as for instance performed for the linear DGLAP fits) are needed. In addition, different computations of the A-dependence of the saturation scale will likely converge as our theoretical understanding improves. Finally, as will be discussed in the following, the issue will likely not be resolved conclusively until DIS experiments off nuclei at small x are performed.
326
E. Iancu and R.
4.6. Probing
Venugopalan
the CGC with an Electron Ion
Collider
A high energy electron-nucleus collider, with a center of mass energy i / i = 60-100 GeV, presents a remarkable opportunity to explore fundamental and universal aspects of QCD. The nucleus, at these energies, acts as an amplifier of the novel physics of high parton densities — aspects of the theory that would otherwise only be explored in an electron-proton collider with energies at least an order of magnitude greater than that of HERA. An electon-nucleus collider will also make the study of QCD in a nuclear environment, to an extent far beyond that achieved previously, a quantitative science. In particular, it will help complement, clarify, and reinforce physics learnt at high energy nucleus-nucleus and proton-nucleus collisions at RHIC and LHC over the next decade. For both of these reasons, an eA collider facility represents an important future direction in high energy nuclear physics. We will briefly discuss here experimental observables in deep inelastic scattering (DIS) which are signatures of the novel physics of the Color Glass Condensate m . A more detailed discussion of the following can be found in Ref. 101. The regime of small xgj's (XBJ < 0.01) is easily accessed by an electron — heavy ion collider in the energy range y/s « 60-100 GeV. These energies would be most natural if the Electron Ion Collider (EIC) were constructed at BNL-this particular realization is called eRHIC. The kinematic coverage of EIC/eRHIC is shown in Fig. 21. What is novel about these energies is that for the first time one can study the physics of XBJ « 0.01 in a nucleus for Q2 » A Q C D , where AQCD ~ 200 MeV. Previous (fixed target) experiments such as NMC and E665 and current ones such as HERMES and COMPASS could only access small XBJ at small Q 2 's. The center of mass (cm) energy of the Electron-Ion Collider (EIC) is a factor of 10 smaller than that of the current ep-collider at HERA (the proposed HERA III plan-which includes e-A scattering would have a center-of-mass energy that's roughly 3 times greater than EIC). However, an eA collider has a tremendous advantage — the parton density in a nucleus, as experienced by a probe at a fixed energy, is much higher than what it would experience in a proton at the same energy. Since the parton density grows as A1/3, this effect is more pronounced for the largest nuclei-see Eq. (1.13). m W e will not cover here the interesting physics at intermediate and large XBJ that can be studied with an eA collider. A nice discussion of these issues can be found in Ref. 99, 100.
The Color Glass Condensate
and High Energy Scattering
in QCD
327
Fig. 21. The x - Q 2 range of the electron ion collider (EIC) compared to that of the HERA ep collider and fixed target experiments. The EIC's reach would encompass the fixed target regime as well as part of the HERA regime.
/
i\l/A
Eq. (1.13) suggests that xproton = ^nucleus/ ( ^ 3 )
• Since the nucleus
1 3
is dilute and conservatively taking the effective A ? = 4, then, for A ~ 0.3, one finds a;proton ~ ^nucleus/100. Thus the same parton density in a nucleus at XBJ ~ 10~ 4 and Q2 ~ a few GeV 2 is attained in a nucleon at XBJ ~ 1 0 - 6 and similar Q2\ Impact parameter tagging is feasible by counting knock-out neutrons 112 — if so, the gain in parton density in eA relative to ep may be even more spectacular. In the following, we will discuss both inclusive and semi-inclusive signatures of the CGC. The latter in particular are very difficult to measure in fixed target DIS Inclusive signatures of the CGC An obvious inclusive observable is the structure function F2(XBJ,Q2) 2 and its logarithmic derivatives with respect to XBJ and Q . The EIC should have sufficient statistical precision for one to extract the logarithmic derivative of F2 (and its logarithmic derivative!). Whether the systematic errors at small XBJ will affect the results is not clear at the moment. The logarithmic derivative dF2/d\n(Q2), at fixed XBJ, and large Q2, as a function of Q 2 , is the gluon distribution. QCD fits implementing the DGLAP evolution equations should describe its behavior at large Q2. At smaller Q2, one
328
E. Iancu and R.
Venugopalan
should see a significant deviation from linear QCD fits — in principle, if the Q2 range is wide enough, one should see a turnover in the distribution. The Q2 at which the turnover takes place should be systematically larger for smaller x's and for larger nuclei 74 . At eRHIC one can extract the longitudinal structure function 2 FL(XBJ, Q ) — F2 — 2 XBJFI at small XBJ independently since the energy of the colliding beams can be varied significantly. In the parton model, FL = 0 — thus FL is very sensitive to scaling violations. It provides an independent measure of the gluon distribution 102 and in particular of higher twist saturation effects which may be prominent in both FL and FT but may cancel in the sum 103 . The extended kinematic range of EIC may help determine whether shadowing is entirely a leading twist phenomenon, or if there are large higher twist perturbative corrections. As also discussed previously, there is a close relation between shadowing and diffraction. At EIC the validity of this relation can be explored directly — different nuclear targets are available, and the diffractive structure function may also be measured independently. Semi-inclusive signatures of the CGC A striking semi-inclusive measurement is hard diffraction wherein the virtual photon emitted by the electron fragments into a final state X, with an invariant mass M\ » A Q C £ ) , while the proton emerges unscathed in the interaction. A large rapidity gap — a region in rapidity essentially devoid of particles — is produced between the fragmentation region of the electron and that of the proton. In pQCD, the probability of a gap is exponentially suppressed as a function of the gap size. At HERA though, gaps of several units in rapidity are unsuppressed; one finds that roughly 10% of the crosssection corresponds to hard diffractive events with invariant masses M\ > 3 GeV. Hard diffraction probes the color singlet object (the "Pomeron") within the proton that interacts with the virtual photon. It addresses, in a novel fashion, the nature of confining interactions within hadrons. The mass of the final state is large and one can reasonably ask questions about the quark and gluon content of the Pomeron. A diffractive structure function F2 jj diffrs can be defined 104 ' 105 (analogous to F2) as
d4aeA^eXA dxBidQ2dxvdt
^
A
4™2m L xQ*
\
F?JC\0,Q2,xP,t),
,
^ 2[1+
I
R%(4)(/3,Q2,xv,t)}j
(4.20)
The Color Glass Condensate and High Energy Scattering in QCD
329
where y = Q2/sxBj and R°(i) = F°{4)/F°W. Also, Q2 = -q2 and xBj = 2 Q /(2P • q) have the usual DIS definitions and q •f
x-p
Here P is the initial nuclear momentum, P' the net momentum of the fragments Y in the proton fragmentation region and M\ the net momentum of the fragments X in the electron fragmentation region. An illustration of the hard diffractive event is shown in Fig. 22. Unlike F% however, F2 is not truly universal — it cannot be applied, for instance, to predict diffractive cross-sections in p-A scattering; it can be applied only in other leptonnucleus scattering studies 105 .
X (M x ) Largest Gap in Event
Y (MY) Fig. 22. The diagram of a process with a rapidity gap between the systems X and Y. The projectile nucleus is denoted here as p. Figure from Ref. 99.
In practice the structure function F2 y — J F2 ^ dt is measured, where \tmin\ < \t\ < \tmax\, where \tmin\ and \tmax\ and the limits of the empirically measurable momentum transfer to the nucleus. The ratio for two nuclei A\ and A2 RAIMP,
2
F2D$(P,Q2,xv)
Q , *V) = ^(D 3 V
1'
F2 $(P,Q2,xv)
;,
(4.22)
can be measured with high accuracy " . If RAI,A2 = 1, the structure of the Pomeron is universal, and one has an A-independent Pomeron flux. If - R A I , ^ = f(Al, A2), then albeit a universal Pomeron structure, the flux is ^-dependent. Finally, if Pomeron structure is yl-dependent, some models argue that RAI,AI = F2,AI/F2,A2-
330
E. Iancu and R.
Venugopalan
The ratio of RD = erdiffractive/ototai a t HERA is ~ 10% for Mx > 3 GeV. The systematics of hard diffraction at HERA can be understood in saturation models 4 3 . For eA collisions at EIC energies, saturation models predict that the ratio Rp can be much higher — on the order of 30% for the largest nuclei 106 . 108 . An important semi-inclusive observable in eA DIS at high energies is coherent (or diffractive) and inclusive vector meson production. For instance, the forward vector meson diffractive leptoproduction cross-section off nuclei is107 ^U=o(l*A
- VA) oc a2(Q2)
[GA(x,Q2)}2
,
(4.23)
for large Q2 and is therefore a sensitive probe of gluon saturation/shadowing. It is very important to measure inclusive and diffractive open charm and jets since they provide useful and complementary data to those of vector mesons. Due to large color fluctuations at small XBJ, on can expect: a) a broader rapidity distribution in larger nuclei relative to lighter nuclei and protons, b) Enhanced anomalous multiplicity distributions where anomalous multiplicity in one rapidity interval in an event would be accompanied by an anomalous multiplicity in rapidity intervals several units away m , and c) a correlation between the central multiplicity with the multiplicity of neutrons in a forward neutron detector 112 . In pA scattering at RHIC one also has the opportunity to study the high parton densitiy phenomena-see section 5.5 for a discussion. Some of the differences between pA and eA are as follows. In the pA Drell-Yan process, it is very hard to reliably extract distributions in the region below the \I>' tail — namely, one requires Q2 > 16 GeV 2 . In the x region of interest, one expects saturation effects to be important at lower Q2 of 1-10 GeV 2 . For Q2 = 16 GeV 2 , one might have to go to significantly smaller x's to see large saturation effects. Secondly, the survival probability of large rapidity gaps is smaller in pA relative to eA. This is because in pA (unlike eA) the gap is destroyed due to secondary interactions between "spectator" partons in the proton and the "Pomeron" from the nucleus. Thus one expects that diffractive vector meson and jet production in pA should be qualitatively different than what one will see in eA.
The Color Glass Condensate
and High Energy Scattering
in QCD
331
5. Melting the CGC in Nucleus-Nucleus and Proton-Nucleus Collisions In general, the problem of high energy hadron-hadron collisions is very difficult. As we discussed in the introductory section, a full understanding of multi-particle production in high energy collisions is one of the outstanding problems in high energy QCD. The approach developed in previous sections however suggests that classical methods may be useful in studying multi-particle production in high energy collisions. In this section, we will discuss applications of this approach in describing the initial stages of very high energy heavy ion collisions, proton (or deuteron)-nucleus collisions and briefly, peripheral nucleus-nucleus collisions. Understanding the initial conditions for heavy ion collisions is of the utmost importance in studying the high energy heavy ion experiments performed at the Relativistic Heavy Ion Collider (RHIC) at BNL and (to be performed) at the Large Hadron Collider (LHC) at CERN. The initial conditions are crucial in determining whether the quark-gluon matter produced immediately after the collision will equilibrate to briefly form a thermalized Quark Gluon Plasma (QGP). Classical methods are applicable because the occupation numbers of gluons in the nuclear wavefunctions (and immediately after the collision) are of order 1/as- One can therefore study the space-time evolution of partons which are "frozen" in the Color Glass Condensate through their release in the collision and their subsequent evolution. Unfortunately, it is not possible in this approach to follow their evolution all the way to equilibration. This is because the gluon occupation numbers become small well before equilibration thereby signalling the breakdown of the classical field approach. Since the classical field equations are fully non-linear, they cannot be solved analytically n . The solutions can however be determined numerically and the initial energy and number distributions can be obtained. A particular feature of our formulation of the scattering problem is that the classical Yang-Mills field equations are boost invariant-the equations are independent of the space-time rapidity. The boost invariance leads to a significant simplification-the classical Yang-Mills equations are only functions now of the two transverse spatial directions and the proper-time. In order to maintain gauge invariance, the numerical problem is formulated on the lattice and solved as a function of proper time. In the
n
For recent analytical work in this direction, see Ref. 114
332
E. Iancu and R.
Venugopalan
following, we will discuss briefly the numerical procedure and the results of the simulations. The non-perturbative results can be compared to the results of analytical computations at large transverse momenta and we will show that they agree with these computations. The results of these numerical computations can now be interpreted in light of the recent data from RHIC. We will discuss some of the phenomenological consequences of our results-in particular the implications for the importance of final state interactions of partons (beyond the initial stage of high occupation numbers) and their subsequent thermalization. Quantum effects, such as the non-trivial geometric scaling discussed earlier, may also play an important role in interpreting the very interesting and in some instances very puzzling RHIC data. As we will discuss, a contentious issue in deciphering the RHIC data is whether the phenomena observed are primarily due to initial state or final state effects. Experiments underway at RHIC on Deuteron-Gold collisions should help settle the issue since one expects final state effects in these collisions to be relatively unimportant ° Recently, there have been several papers studying various final states at RHIC energies in the CGC model. We will discuss these works and their predictions for the RHIC data. Finally, we will also discuss briefly the predictions of the CGC model for particle production in peripheral nucleus-nucleus collisions.
5.1. Classical
Picture
of Nuclear
Collisions
The classical picture of nuclear collisions was first formulated by Kovner, McLerran and Weigert 29 in the framework of the McLerran-Venugopalan model for a single nucleus. At very high energies, P+ —> oo of one of the nuclei (and P~ —> oo for the other), the hard valence quark (and gluon) modes are highly Lorentz contracted, static sources of color charge for the wee parton, Weizsacker-Williams, modes in the nuclei. For simplicity, we will first consider only central collisions of cylindrical nuclei with uniform matter distributions. The case of realistic nuclei and non-central collisions will be discussed later. The valence sources are then described by the current J"-°(r ± ) = V+pl{rL)5{x-)
+ 8"-fQ(r±)8(x+),
(5.1)
"Scattering effects that are often thought of as "final state" effects in proton-nucleus effects can, in a different gauge, be interpreted as initial state effects 2 7 . The final state effects emphasized here are those involving re-scattering of on-shell partons off each other.
The Color Glass Condensate
and High Energy Scattering
in QCD
333
where />i(2) correspond to the color charge densities of the hard modes in nucleus 1 (nucleus 2) respectively. The ^-function in x~ for the valence parton current of one nucleus (or 5-function in x+ for the other) implies that we are literally assuming the nuclei to move at the speed of light. In reality of course this condition should and can be relaxed. We will discuss this issue later on in this section. In the collision, these valence partons are assumed to be Eikonal sources-they continue on their straight line trajectories along the light cones. They may acquire a phase due to the rotation of the color charge in the scattering 115 . However, in the gauge we will use, no such phase will appear in our treatment of the classical scattering problem. For each color configuration in each of the nuclei, the classical field describing the small x modes in the Effective Field Theory (EFT) is obtained by solving the Yang-Mills equations in the presence of the two sources. We then have D ^ F " " = Jv .
(5.2)
To compute a physical quantity (O), the gauge field configurations have to be averaged over the respective Gaussian path integrals of the two nuclei, (0)p=
/ d / 5 i d p 2 0 ( p i , p 2 ) e x p l - / d rj_
^-^
I • (5.3)
For instance, the small x gluon distribution is simply related to the Fourier transform Af(k±) of the solution to Eq. ( 5.2) by < A?(k±)Aa(kx) >p, where the subscript denotes the average above. Here have assumed identical nuclei with equal Gaussian weights g4(J-\, where [i2A is the average color charge squared of a nucleus, denned in Eq. (2.5). We will henceforth use the variable A^ = g4^\. It is simply related to the saturation scale Qs by the relation
In practice, A s ~ Qs. In general, one can make the following ansatz for the gauge fields as a function of proper time r = V2x+x~ in the different light cone regions: Ai = ai(T,xT)0{x-)6(x+)
+
+
+ ai(T,xT)6(-x-)e(x ), A
±
±
a\(T,xT)e(x-)6(-x+) (5.5)
+
= ±x a(T,x1_)9(x-)6(x ),
(5.6)
334
E. Iancu and R.
Venugopalan
where a\ and a are respectively the transverse and longitudinal components of the gauge field the forward light cone while a\ 2(»~x) (* = 1> 2 are the two transverse Lorentz indices) are pure gauge fields defined through the gauge transformation parameters Aq(r],r±) U 6 o4 >2 (rx) = \ (Pe-ij^
*>'A lia (,',rxA
Vi
fPe
*i'^Mv'.r±)\
(57)
Here rj = ±ryproj ^log(a;=F/a;5:roj) is the rapidity of the nucleus moving along the positive (negative) light cone with the gluon field a\,2)- ^ n e ^1,2(77, r±) are determined by the color charge distributions Aj_Ag = pq (q=l,2) with Aj_ being the Laplacian in the perpendicular plane. We work in the Fock-Schwinger (or "radiation") gauge AT = x+A~ + x~A+ = 0, which is the interpolation between two light cone gauges. Fixing this gauge however does not fix the gauge completely. The residual gauge freedom can be fixed by imposing the Coulomb gauge condition Vj_ • A± = 0 in the two transverse dimensions. Substituting Eq. (5.5) in Eq. (5.2), and re-writing the equations in terms of the transverse coordinates x±, the proper time T and the space-time rapidity rj, one observes that the Yang-Mills are independent of 77, namely, they are boost invariant. The fields a\ and a are functions of x± and r only. The boost invariance of the solutions of the Yang-Mills equations is entirely due to our approximation-using J-function sources. Smearing the sources in rapidity would destroy the boost-invariance of the solutions. Nevertheless, one expects that the boost-invariance approximation is a reliable one especially at central rapidities in a nuclear collision. This point will be discussed further. The initial conditions for the solution of the Yang-Mills equations in the forward light cone are determined by matching the solutions in the space-like and time-like regions at r = 0. Requiring that the gauge fields be regular at r = 0, the coefficients of the singular pieces of the equations D^F^ = 0 and Z3 M ±F' t± = J ± (for x~,x+ -> 0) have to be set to zero. These give the boundary conditions at T = 0: ai(0,x±)=a\(0,x±) a(0,ZL) =
(5.8)
-[a{(0,x±),ai(0,x±)}.
(5.9)
l
+ aU0,x±),
These conditions, first formulated for infinitely large nuclei, are the same for finite nuclei as well. The boundary conditions remain the same even
The Color Glass Condensate
and High Energy Scattering
in QCD
335
when the fields a\ 2 before the collision are smeared out in rapidity properly account for singular contact terms in the equations of motion 116 . Further, since the equations are very singular at r = 0, the only condition on the derivatives of the fields that would lead to regular solutions are dTa\T=0,dTa\\T=o = 0. Perturbative solutions of the Yang-Mills equations to order p2 in the color charge density (or equivalently to second order in As/kj_) were found, and at late times, after averaging over the Gaussian sources, the number distribution of classical gluons was found to be 29 ' 116 > 115 d^kI
= nR
^
T
fcT
Hk±,X),
(5.10)
where L(k±,X) is an infrared divergent function at the scale A. This result agrees with the quantum bremsstrahlung formula of Gunion and Bertsch m . The distributions are very sensitive to L(k±, A). What is novel about the classical approach is that, at sufficiently high energies, the non-linearities in the Yang-Mills fields self-consistently regulate this infrared divergence. To confirm this claim, one needs to solve the Yang-Mills equations to all orders in As/k±. This is very difficult to do analytically. Fortunately, as we will discuss in section 5.2, the classical problem can be solved numerically. The discussion here has been extended to treat the collision of finite, ultrarelativistic nuclei with realistic nuclear matter distributions 119 . For the case of finite nuclei, the issue of color neutrality, as discussed in sections 2.4 and 3.10, becomes very important. A global neutrality constraint at the level of the nucleus is insufficient to ensure that there are no large field strengths outside the nuclear radius-color neutrality must be imposed at the nucleon level. The practical implementation of color neutrality will be discussed further in the following. As mentioned earlier, a major simplification occurs in the classical approach when boost invariance is assumed. It is likely a good approximation, especially in the central region, but the effects of relaxing this condition are unclear. A first step in answering this question (before performing a fully 3+1-dimensional simulation) is to study the stability of the 2+1dimensional results with respect to a perturbation in the ^-direction 120 . 5.2. Numerical
Gluodynamics
of Nuclear
Collisions
Classical real time numerical solutions of gauge theories were first discussed in the context of sphaleron transitions during the electroweak phase tran-
336 E. Iancu and R. Venugopalan
sition in the early universe 28 . Similar techniques can be applied to discuss the problem at hand. It is most convenient to follow the Hamiltonian approach, namely, to construct the appropriate lattice Hamiltonian and solve Hamilton's equations on the lattice with the lattice analog of the initial conditions discussed in section 5.1. These will be discussed below. We will subsequently discuss the results of our classical numerical simulations. i) Numerical Solution of the Yang-Mills Equations The QCD action for gauge field in the r, 77, f coordinates reads SQCD
= frdrjdTcPr
-^Tr^V^F^)
(5.11)
where F^ = d^Av — 3„AM — ig\A^Av\ and the metric is diagonal with = A^ta, and ta represent grr = _gxx = _gm = t a n d gr,r, = _ i / T 2 . ^ a gauge group matrices with the normalization of Tr(t a t b ) = 2#jj. The Lagrangian density in AT = 0 gauge is C = Tr ( £ ( a T A 0 2 + ^ A „ )
2
- IF^ - -^F2;) ,
(5.12)
where i,j runs over transverse coordinate x and y. Now let us assume 77 independence of the fields. As discussed previously, the Yang-Mills equations have this property if the sources are strictly 5function sources on the light cone. We have Ai(T,r1,f}=Ai{T,r),
i4,(T,J7,f)=*(T,f),
(5.13)
resulting in FVi = —Di§, where Di = di — ig[Ai, • ••] is the covariant derivative in the adjoint representation. Defining the conjugate momenta Ei = rdTAi and pn = ^dTAv, one finds that the boost invariant YangMills Hamiltonian is the QCD Hamiltonian in 2+1 dimensions coupled to an adjoint scalar 30 :
H = Jd\K [j-Ef + Jp2 + IE* + ^(Di*) 2 } .
(5.14)
In order to realize numerically the solutions to the equations of motion in the previous section, while maintaining the gauge symmetry, we introduce the link variables at the site i Uj,i = exp [iaAj(i)],
(j = x, y),
(5.15)
The Color Glass Condensate
and High Energy Scattering
in QCD
where, a is a lattice spacing. Defining the plaquette Ui,jUmj+iUi ,-+m£^„i,-, the Hamiltonian on the lattice is
+ h T,1*^ - u*j*i+*u&2 + i E ^ i j,n
C/Q
337
=
(5-16)
j
where the convention for the generators of the SU(3) color group is Tr(TaTb) = 2Sab. For g = 2, one obtains the correct normalization of the Hamiltonian in the continuum limit. Lattice equations of motion follow directly from Hi of Eq. (5.16). For any dynamical variable v, with no explicit time dependence, v = {HL,V}, where v is the derivative with respect to r, and {} denote Poisson brackets. We take Ei, Ui, Pj, and &j as independent dynamical variables, whose only nonvanishing Poisson brackets are {pf, $$} - SijSab; {E?, Um} = -tiimUiaa;
{£?,«, Ebm} = 2SlmeabcEf
(no summing of repeated indices). The initial conditions for the transverse gauge field and the adjoint scalar field on the lattice can be obtained in complete analogy to the procedure followed in the continuum. The details of this procedure and the expression for the initial conditions can be found it Ref. 30. We impose periodic boundary conditions on an N x iV transverse lattice, where N denotes the number of sites. The physical linear size of the system is L = a N, where a is the lattice spacing. It was shown in Ref. 30 that numerical computations on a transverse lattice agreed with lattice perturbation theory at large transverse momentum. ii) Numerical Method for Finite Nuclei In early studies 30 > 31 . 32 . 33 ; nuclear collisions, for simplicity, were idealized as central collisions of infinite, cylindrical nuclei. The color charge squared A^ was taken to be a constant for the uniform cylindrical nuclei. Furthermore, color neutrality was imposed only in a global sense 5 0 , namely, the color charge distribution over the entire nucleus was constrained to be zero. While very useful in obtaining first estimates of the space-time evolution of the produced gluonic matter, these studies did not make predictions for realistic nuclear collisions. In addition, studies of the distributions in peripheral collisions, in particular of the azimuthal anisotropy associated
338
E. Iancu and R.
Venugopalan
with elliptic flow, require finite nuclei and realistic nuclear matter distributions within each nucleus. These requirements were discussed in Refs. 118, 119. The problem with color neutrality for a finite nucleus can be stated simply as follows. If we impose the simple and obvious constraint that the color charge distribution must be zero outside the nucleus, the solution of Poisson's equation can still give a non-zero gluon distributions outside the nucleus. In two dimensions, the fall-off of the gluon field is rather slow as shown in Fig. 23. This slow fall-off is a problem for a finite nucleus since the gluon field is associated with a non-zero field strength. Clearly the simple prescription for color neutrality is not sufficiently stringent.
~
1
6
r" ' ! " ! • ,
<:i°'2 ir
1 i
t
.,
10* : • MV • • color neutral II 10 s ! • mc=1 fm" 1 it)"4
10*
I
* 1
I
•
10
12 r(fm)
Fig. 23. Gluon field as a function of radial distance. Original MV model is shown by circles, while squares correspond to the Color Neutral II prescription (see text). The triangles represent results from solution of Poisson's equation with a screening mass. The results are for AsoR = 37.
A more realistic prescription would be to apply the color neutrality constraint already at the nucleon level. Our numerical procedure to implement the constraint for finite nuclei is as follows. We first sample A nucleons on a discrete lattice requiring that they satisfy a Woods-Saxon nuclear density profile in the transverse plane. Note that this procedure generates the same distribution in the continuum as A^(x±) = A^0TA(X±) where TA(X±) = J_oodzn{r) is a thickness function, x± is the transverse coordinate vector (the reference frame here being the center of the nucleus), n(r) is the Woods-Saxon nuclear density profile, and A^0 is the color charge squared per unit area in the center of each nucleus. The only external dimensional variables in the model are Aso and the nuclear radius R.
The Color Glass Condensate
and High Energy Scattering
in QCD
339
Next, Gaussian color charge distributions are generated on the lattice. The probability distribution of color charge in a nucleon is expressed as P[p] = exp
(5.17)
where A^ • is the color charge distribution squared, per unit area, of a nucleon at a lattice site j and N is the number of lattice sites that comprise a nucleon. Anj is obtained from A^0 by assuming that the color charges of the nucleons add incoherently. There are two versions of the subsequent step. In the first (which we term Color Neutral I), we subtract from every pj the spatial average V . Pj/N in order to guarantee color neutrality (p) = 0 for each nucleon. In the second (termed Color Neutral II), the dipole moment d of each nucleon is eliminated in a similar manner. ~
1.4 MV color neutral I color neutral II
ij 1.2
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 PT^SO
Fig. 24. Color Charge Correlator in momentum space. Original MV model is shown by circles, while squares correspond to the result from color neutrality condition I and triangles correspond to color neutrality condition II (see text). The color charge correlator is plotted versus p j . in units of A3o •
In Fig. 24, we plot the Fourier transform of the charge correlator, which in the continuum is defined as P(P±) = fd2x±exp(ip±
• x±)(pa(x±)pa(0)),
(5.18)
for the MV model and for the two variants which impose color neutrality on the nucleon level. In the MV model, p(p±) is a constant everywhere except at p± = 0 where it is constrained to be zero from the global charge constraint. In the Color Neutral I (II) variant, we see that p(p±) ~ p\ (~ p\) for small momenta p\_ < Aso and is constant at larger momenta.
340
E. Iancu and R.
Venugopalan
The oscillatory behavior seen for Color Neutral I and II is due to the fact that the correlator in coordinate space is not strictly a delta-function. In the coordinate space the charge correlator for the two models, Color Neutral I and II, falls off rapidly, as ~ l/£j_ and ~ 1/zj. respectively, at larger distances. It is an interesting coincidence that the behavior of p(p±) in our model is similar to the behavior expected from the renormalization group (RG) evolution of color charges in the McLerran-Venugopalan model as discussed here in Section 3.10. In Refs. 16, 123, it is shown that the screening of color charges due to the RG evolution gives a behavior p(p±) ~ p\ for p±_ < A su (and p(pj_)=constant for p± > Aso). iii) Numerical Results for Distributions in Energy and Number in Central Collisions We will now discuss results for central collisions of very large cylindrical nuclei. We will see later that they are not very different from those for central collisions of realistic nuclei. There are only two free parameters for the problem of nuclear collisions as formulated int the EFT. One is the saturation scale As while the other is the nuclear radius R p (For an infinite cylindrical nucleus, one has irR2 = L2, where L is the lattice size.) Any dimensional quantity P, well defined within the EFT, can be written in terms of the physically relevant parameters A s and R as Af fp(AaR), where d is the dimension of P. The non-trivial physical information is therefore contained in the dimensionless function fp(A3R). On the lattice, P will generally depend on the lattice spacing a; this dependence can be removed by taking the continuum limit a —> 0. The broad range of physically relevant values of As for RHIC and LHC energies are ~ 1-2 GeV and 2-4 GeV respectively-corresponding to ASR w 30-120 approximately*1. Also, for central Au-Au collisions, we obtain L = 11.6 fm as the physical linear dimension of our square lattice. For the transverse energy of gluons we get, on purely dimensional
p
Strictly speaking, the saturation scale is not a free parameter since it can be determined from the gluon density as in Ref. 72, 27. However, for the momentum scales of interest, there is much uncertainity in the gluon density. Moreover, for nuclei, gluon shadowing contributions are not under control. Our results will therefore be non-perturbative formulae valid for a wide range of A s . q If we extrapolate from the Golec-Biernat HERA parametrization, we get A s = 1.4 GeV for RHIC and As = 2.2 GeV for LHC.
The Color Glass Condensate
and High Energy Scattering
in QCD
341
0.35 0.3 L °0.25 ^
(a)
* * * * .
0.2 -
* * * • » • •
0-15ti
i i i * < • SU3 • SU2*8/3
t
0.1, '0
2
4
6
8 10 12 14 16 18 20 A. T
Fig. 25. (a) er/Ag as a function of T A S for A s i t = 167.4. (b) £ T / A § as a function of Asa for AS.R = 167.4 (squares) and 50(circles), where a is the lattice spacing. Lines are fits of the form a — bx.
grounds, 1
dET.
\n=0 •KE? dr) '"'
1 5^
fE(AsR)A3s
(5.19)
The function fs is determined non-perturbatively as follows. In Figure 25 (a), we plot the Hamiltonian density, for a particular fixed value of ASR = 167.4 (on a 512 x 512 lattice) in dimensionless units as a function of the proper time in dimensionless units. We note that er converges very rapidly to a constant value. The form of er is well parametrized by the functional form ST = a + /3exp(—77-). Here dEr/dri/ivR2 = a has the proper interpretation of being the energy of produced gluons per unit area per unit rapidity, while TD = l / 7 / A , is the "formation time" of the produced glue. In Figure 25(b), the convergence of a to the continuum limit is shown as a function of the lattice spacing in dimensionless units for two values of ASR. In Ref. 31, this convergence to the continuum limit was studied extensively in an SU(2) gauge theory for very large lattices (up to 1024 x 1024 sites) and shown to be linear. The trend is the same for the SU(3) resultsthus, despite being further from the continuum limit for SU(3) (due to the significant increase in computer time) a linear extrapolation is justified. We can therefore extract the continuum value for a.
342
E. Iancu and R.
Venugopalan
Numerical results for the physically relevant RHIC and LHC initial conditions were discussed in several papers 31>32.33.i18>n9 Very recently, Lappi pointed out that the overall normalization in Eq. (5.19) was incorrect and the JE computed in Refs. 31, 33 is too large by a factor of two 121>122. As we will discuss below, the normalization for the number is correct. Wherever feasible, the corrected results will be discussed below. We find / B (50) = 0.269 and //j(167.4) = 0.248. The RHIC and LHC values likely lie in this wide range of ASR. The SU(2) value is approximately half the SU(3) value. Note that the variation of fs as a function of ASR is extremely weak.The formation time TD = l / 7 / A g is essentially the same for for both the SU(3) and SU(2) cases-for ASR = 167.4, 7 = 0.362 ± 0.023 and . As discussed in Ref. 31, it is ~ 0.3 fm for RHIC and ~ 0.13 fm for LHC (taking As = 2 GeV and 4 GeV respectively-for the values of A s extracted using the Golec-Biernat fit, the corresponding times are of course larger). We now combine our expression in Eq. (5.19) with our non-perturbative expression for the formation time to obtain a non-perturbative formula for the initial energy density 122 , £
~ A <
(5.20)
This formula gives a rough estimate of the initial energy density, at a formation time of TD = l/*//AaR where we have taken the average value of the slowly varying function 7 to be 7 ~ 0.3. For A s = 1.4 GeV, one obtains e ~ 10 GeV/fm 3 . We now report our results for the initial multiplicity of gluons produced at central rapidities. First consider a free field theory whose Hamiltonian in momentum space has the form
ff/ = | £ ( k ( * o i 2 + " 2 ( * ) w * o i 2 ) .
<5-21)
k
where
N(k) = w(k)Mk)\2) = v W F k W I 2 ) , •
(5-22)
In our case, the average () is over the initial conditions. We use two different generalizations of the particle number to an interacting theory. We have verified that the two definitions agree in the
The Color Glass Condensate
and High Energy Scattering
in QCD
343
weak-coupling regime at late times 31 . Our first definition is based on the behavior of a free-field theory under cooling. We obtain 32 AT=V-yo
_v(«), vW 0 v*
(5.23)
where t is the cooling time (not to be confused with real or proper time), and V(i) is the potential energy of the relaxed free field after cooling. The relaxed potential V(t) is gauge-invariant-hence so is this definition of the particle number. This is an attractive feature of the cooling method. Unfortunately, it presently only permits determination of the total particle number and cannot be used to find the number distribution N(k±). Our second definition of the multiplicity will enable us to compute N(k±). We impose the Coulomb gauge condition in the transverse plane, Vj_ • A± = 0, and substitute the momentum components of the resulting field configuration into Eq. (5.22). We can determine N(kj_) from the rightmost expression of Eq. (5.22); the middle expression of Eq. (5.22) can then be used to obtain w(k±). In Fig. 26(a), we plot the normalized gluon transverse momentum distributions versus k±/As with the value ASR = 167,4, for both the SU(3) and SU(2) gauge theories r . Clearly, we see that the normalized result for SU(3) is suppressed relative to the SU(2) result in the low momentum region. In Fig. 26 (b), we plot the same quantity over a wider range in k±/As for two values of ASR. At large transverse momentum, we see that the distributions scale exactly as N% — 1, the number of color degrees of freedom. This is as expected since, at large transverse momentum, the modes are nearly those of non-interacting harmonic oscillators. At smaller momenta, the suppression is due to non-linearities, whose effects, we have confirmed, are greater for larger values of the effective coupling ASR. The SU(3) gluon momentum distribution can be fitted by the following function,
^aWTT = 7UkT/A^ where fn(kT/As)
(5 24)
'
is ai exp (yJkl+m?/T^)
-1
(fcr/A. <1.5)
/n=<
(5-25) a 2 At log(47rfcT/As)fcX4
(kT/As
> 1.5)
r Our results for the number distribution agree with Ref. 121 if A s —» A 3 / 2 in the ensuing discussion of the p± distributions. See Ref. 122 for details.
344
E. Iancu and R.
Venugopalan
0.07 i 0.06 0.05 it ^•0.04 "£"0.03 0.02 fc, 0.01
ASR=167.4
(a)
0.2
%
°SU3 •SU2 -fit
0.4
0.6
0.8
1
MA. u .
10" 10"
fit
\
r
r
SU3: ASR=50 SU2: A.R=50 SU3: A S R-167.4
io"
^ ^ ^ ^
10" 2
)
4
6 8 kyA,
(b) 10
12
Fig. 26. Transverse momentum distribution of gluons, normalized to the color degrees of freedom, n(k±) = fn/(N* - 1) (see Eq. (5.24)) as a function of A s f i for SU(3) (squares) and SU(2) (diamonds). Solid lines correspond to the fit in Eq.(5.25).
with ai = 0.118, m = 0.034A„, Teff = 0.47AS, and a2 = 0.0087. At low momenta, the functional form is approximately that of a Bose-Einstein distribution in two dimensions even though the underlying dynamics is that of classical fields. The functional form at high momentum is motivated by the lowest order perturbative calculations 116 > 29 ' 115 . Integrating our results over all momenta, we obtain for the gluon number per unit rapidity, the non-perturbative result, 1
-KR2
dN
-
dr]
|rj=0
\fN(AsR)A2s.
(5.26)
We find that /AT(167.4) = 0.3. Our results for /JV are in reasonable agreement with those of Lappi 121 - 122 . We have checked for an SU(2) gauge theory that the results for a wide range of ASR vary on the order of 10%. The results from the Cooling and Coulomb methods also show very good agreement (less than 10%) especially for larger values of ASR. If we take the ratio of Eq. (5.19) to Eq. (5.26), we find that the initial transverse energy per gluon is dET/dr]/dN/r]\ri=o = *f- As = 0.88 A s . If we take A s = 1.4 GeV, we find that the energy per gluon is ~ 1.23 GeV-about a factor of 2 larger than the value for charged hadrons measured at RHIC. The topological charge generated in the initial stages of a heavy ion
The Color Glass Condensate
and High Energy Scattering
in QCD
345
collision can also be computed in the classical CGC framework. An interesting result is that if strict boost invariance is a good assumption sphaleron transitions are suppressed 124 . This is because the Chern-Simons number in this case is invariant under all rapidity independent gauge transformations. The primary mechanism for the generation of topological charge at the early stage is then by fluctuations of the color electric and magnetic fields. It was shown in Ref. 124 that the topological charge generated in this manner, at the early stage, are small. These results may be relevant for the formation of P and CP-odd metastable states in the late stages of heavy ion collisions 125 . iv) Numerical Results for Centrality Dependence of Energy, Multiplicity and Elliptic Flow For realistic nuclei, the non-perturbative relations discussed in section 5.2.3 are less simple. One can write Eq. (5.26) more generally as dN,g _ fN(b) A 2 0 drj Po
part
(b),
(5.27)
where po = p(0,0) = 4.321/m - 2 and Npart = f d2x_\_p(b,x±). For explicit expressions for /;v(&) for different values of Aso, see Ref. 119. One can similarly compute dEg/dr) for finite nuclei.
Fig. 27. The two components of the transverse pressure (Txx and Tvv) and the energy density e plotted as a function of T in dimensionless units. The results are for a impact parameter b = R and dimensionless coupling AsoR = 37. Also shown is the sum of the two transverse pressures.
346
E. lancu and R.
Venugopalan
The azimuthal anisotropy in the transverse momentum distribution is a sensitive probe of the hot and dense matter produced in ultra-relativistic heavy ion collisions 126>128. A measure of the azimuthal anisotropy is the second Fourier coefficient of the azimuthal distribution, the elliptic flow parameter V2- Its definition is 127
j:vd4>coS(2cP)fd'pTl^L_
V2 = (cos(2^)> = j-\: j;;^ \ z Jl«d
•
(5-28)
Td4
The classical Yang-Mills approach may be applied to compute elliptic flow in a nuclear collision. For peripheral nuclear collisions, the interaction region is a two-dimensional almond shaped region, with the x axis lying along the impact parameter axis and the y direction perpendicular to it and to the beam direction. Even though large electric and magnetic fields (and the corresponding transverse components of the pressure in the x and y directions) are generated over very short time scales r ~ 1/AS, the significant differences in the pressures, responsible for elliptic flow, are only built up over much longer time scales r ~ R. This can be seen in Fig. 27 where we plot the two transverse components of the pressure (Txx and Tyy) and the energy density as a function of proper time (in units of Aso) for a peripheral nuclear collision. Moreover, the elliptic flow is generated by soft modes pr ~ A s /8. Our result has important consequences for the theoretical interpretation of the RHIC data-these will be discussed later in the text. The elliptic flow, defined by Eq. (5.28), can be computed, as in the case of the gluon multiplicity, in two different ways; directly in Coulomb Gauge (CG) and by solving a system of relaxation (cooling) equations for the fields. It is easy to show that V2 N, N being the total gluon number, can be reconstructed from the cooling time history of Txx — Tyy, just as N can be reconstructed from that of the energy functional 32 : - / ^=(Txx(t)-Tvy{t)). (5.29) i" Jo Vt This expression for V2 N is manifestly gauge invariant. In Fig. 28 we plot V2 reconstructed from the cooling time history of only the potential terms in Txx — Tyy, along with the CG values (also including potential terms only) as a function of nch/ntot for different values of Aso-R as discussed in the figure. The systematic errors represented by the band (for Aso = 37 are primarily due to limited resources available to study the slow convergence of the cooling and CG computations. We have studied the
The Color Glass Condensate
and High Energy Scattering
in QCD
347
CG v.s.Cooling CM
"e 5 4 3 2 1
ft corrected values •
%
i ' • • • i • • '
• i * • • • i
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 n
ch'^max
Fig. 28. T h e centrality dependence of V2 at the earliest times is computed using cooling (open symbols) and CG (filled symbols). Results are for AsoR spanning t h e RHIC-LHC range, specifically, A3oR = 37 (squares), 74 (triangles), and 148 (stars). Full circles denote preliminary STAR data 1 2 9 . T h e band denotes the estimated value of V2 when extrapolated to very late times. "Corrected values" denotes the late time cooling and CG result for A s o ^ = 37 at one centrality value.
tr* 14 12 10 8 6 4 2 0
b/2R=0.75 A s0 R=148
ftr. i . 1 . . . . 1 . . . . 1 . . . . 1 . . . . 1 . . . . 1
0
. . . . I , . . . I .
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 PT'ASO
Fig. 29. V2(p±) As0R = 148.
as a function of transverse momentum in dimensionless units for
late time behavior of v^ for one impact parameter-the results are shown in the figure. The asymptotic values of V2, as predicted by the model, undershoot the data. For a fixed impact parameter, the model predicts that, as Asoi? —•> oo, the classical contribution to the elliptic flow goes to zero. This is because
348
E. Iancu and R.
Venugopalan
increasing AsoR is equivalent to increasing R for fixed A su and therefore reducing the initial anisotropy. In Fig 29, v2(p±) is plotted for b/2R = 0.75 for As0R = 148. Our calculations show that the elliptic flow rises rapidly and is peaked for p± ~ A s o/8 before falling rapidly. The theoretical prediction 130 is that for p± » A s0 , v 2(p±) ~ -^so/Pj.- The lattice numerical data appear to confirm this resultbetter statistics are required to determine the large momentum behavior accurately. The dominant contribution of very soft modes to v2 helps explain why the cooling and CG computations differ until very late times. The soft gluon modes have large magnitudes and therefore continue to interact strongly until very late proper times. Concomitantly, the occupation number of these modes is not small and the classical approach may be adequate to describe these modes even at the late times considered. 5.3. Melting
the Color Glass Condensate
4 3 8 % zs 3 8 ** 3.4 3.2 •o z 3 •o 2.8 2.6 2.4 2.2
at
RHIC
= •
* = •
2P
50
As(0)=1.18 As(0)=1.95 PHENIX(130) PHOBOS(130) I
I
.
I
I
I
I
I
I
I
I
I
L
100 150 200 250 300 350, 400 N
part
Fig. 30. Comparison of the centrality dependence of the gluon distribution from SU(3) lattice results to data from experiments 131>132 The strong coupling constant is fixed to the value g2 = 4. The lattice results for A s (0) = 1.18 GeV and A s (0) = 1.95 GeV are multiplied by a factor 2.4 and 1.1, respectively.
In Fig 30, we plot the computed centrality dependence of gluons together with the experimental data from PHOBOS 132 and PHENIX 131 . We assume here that the charged particle multiplicity is two thirds of the gluon number. The classical computation is performed for fixed as; the centrality dependence, as seen from Eq. (5.27), comes from the dependence of
The Color Glass Condensate
and High Energy Scattering
in QCD
349
on the impact parameter. In Ref. 32, it was shown that /AT = /AT(A,,.R) increases slowly with As.R-hence one expects it to vary with impact parameter. We see that the results agree reasonably well with the data. The centrality dependence of the transverse energy is studied in Fig. 31. As in the case of the multiplicity, even though the absolute normalization is strongly dependent on one's choice of A s , the centrality dependence is very similar for the two A s 's and shows reasonable agreement with the data. /AT
r
e.
& 1.8 %F P-
3. UJ •a
1.6 1.4 1.2 1
a
50
a
As(0)=1.18 scaled by 7/8
o
As(0)=1.95 scaled by 1 /4
•
PHENIX(130)
100 150 200 250 300 350 400 Npar,
Fig. 31. Comparison of the centrality dependence of the gluon transverse energy distribution from SU(3) lattice results to data from experiments 1 3 3 . The strong coupling constant is fixed to the value g2 = 4. The lattice results for A s (0) = 1.95 GeV are scaled by j while those for A 3 (0) = 1.18 are scaled by | .
Let us now compare our results with those derived previously by Kharzeev and Nardi in Ref. 36 and discussed further in Ref. 134. In these works, one obtains in terms of Qs, the average saturation scale, the result dNg
Af c 2 -1 [ ,2
Q29(b,X±)
^,
^2,t.^Npart
,- 0 „N
In the leading logarithmic approximation, if c;v is constant, one obtains a logarithmic dependence on the centrality entirely from xG(x, Q1(b)). One could thus attribute the logarithmic behavior at the classical level to fixed a3 and leading logarithmic behavior of the gluon distribution function or equivalently, at higher order, to the one loop running of as. In the physically interesting regime, it is difficult to distinguish between the two. A reasonable agreement with the data is also seen in this formulation of the problem.
350
E. Iancu and R.
Venugopalan
The relation between the two formulations is as follows. What we simulate numerically is the color charge squared per unit area, A2 32 . The saturation scale, on the other hand, is a scale determining the behaviour of the gluon number distribution 27 . The relation between the two is given in Eq. (5.4). The relation between c^r and /N is then CJV = 47r2/iv/(7V"c - l)/ln(<5 2 /A 2 5 C D ). Therefore, if cN is to be a constant, fN increases logarithmically with Qs. A weak rise in /JV is seen in our simulations. If the infrared scale, as argued in Ref. 46, is a number of order 0{1/Qs), we would have Aa « Qs. This is indeed the case in practice 119 . The initial transverse energy per particle is ET/N ~ 0.88AS, which for the Golec-Biernat value of A s = 1.4 GeV (also the value favored by Refs. 36, 37), gives ET/N = 1.23 GeV. This is a factor of 2 larger than the value for measured charged hadrons 133 . For this value of A s , the initial number of gluons is approximately half the multiplicity of hadrons at central rapidity. The excess ratio of ET/N in the initial state can be reduced by a) inelastic fragmentation of partons, which increases AT and b) thermalization and hydrodynamic expansion expansion which increases N and decreases ET respectively. Thermalization increases the particle number because it is driven primarily by inelastic partonic processes 34 . It is very conceivable that the multiplicity increases by a factor of two due to either a) or b) or both. Thermalization increases the particle number because it is driven primarily by inelastic partonic processes 34 . The p± distributions from the melting of the CGC were discussed previously. The distributions obtained numerically from purely classical considerations can be used as initial conditions in parton transport equations which describe the evolution of the system when the description in terms of classical fields is inadequate. The parton distributions exhibit geometric scaling-they are functions of p±/A s alone. (At large p± (pj_ » A s ), one expects this classical geometric scaling break down and distributions to be described by perturbative QCD.) It has been argued that the RHIC data demonstrate scaling behavior in a broad kinematic region in p± 38>135. A very similar scaling is seen in the string percolotion framework 136 . In Ref. 39, a novel mechanism for nuclear collisions was suggested, where, due to quantum evolution of saturation effects in the wavefunctions of the incoming nuclei, the p± window, where geometric scaling holds, may extend to p± > Q^/AQCD- This mechanism has precisely the same origin as the geometric scaling discussed at length in Sect. 3.4 and which was suggested in Sect. 4.3 to be an explanation the large scaling window observed in Deeply Inelastic Scattering at HERA. As a consequence of geometric scaling, the
The Color Glass Condensate
and High Energy Scattering
in QCD
351
p± distributions in AA-collisions at RHIC (and more strikingly at LHC) scale with centrality oc Npart. The boundary of this region is estimated to be ~ 4 GeV at RHIC and ~ 9 GeV at LHC. The CGC+geometric scaling picture therefore explains the qualitative behavior of the quenching seen in central collisions at RHIC 137 . For peripheral collisions, the saturation scale is small: the window for geometric scaling shrinks and the binary scaling of pQCD is restored just as observed in the data. This CGC+geometric scaling picture of Ref. 39 is in striking contrast to the QGP+energy loss description of the quenching of spectra seen in central AA-collisions 138 . The two scenarios, at face value, make very different predictions for quenching in d-Au collisions. The former predicts quenching in roughly the same p± range with the centrality dependence {Npart)1/2; the latter predicts a "Cronin-enhancement" just as seen in lower energy p-A collisions. However, multiple-scattering of the Cronin variety is also, in principle, present in the CGC picture. The magnitude of this effect must be quantified before we rush to any conclusions about the RHIC data. We will discuss p-A collisions further in Section 5.5. We now turn to the theoretical interpretation of the RHIC v?. data in the CGC approach. It is clear from Fig. 28 that our result for v-i contributes only about 50% of the measured V2 for various centralities. Our p± distributions also clearly disagree with experiment 139 ' 140 . Naively, one could argue that the classical Yang-Mills approach is only applicable at early times so additional contributions to V2 will arise from later stages of the collision. While there is merit in this statement, it is also problematic as we will discuss below. The reason the situation is complex is as follows. We observed that it takes a long time r ~ R to obtain a significant elliptic flow in the CGC. At these late times, one would expect that the classical approach would be inapplicable due to the rapid expansion of the system. On the other hand, we have seen that V2 in the classical approach is dominated by soft modes which are strongly interacting and don't linearize even at time scales T ~ R. Clearly, the soft modes cannot be treated as on-shell partons even at times r ~ -R! This conceptual problem will be discussed further in Section 5.4. The observation in Ref. 118 that the CGC alone is unable to explain the low p± behavior of the RHIC t>2 data is, thus far, the strongest argument for the importance of final state effects at RHIC. Finally, we note that v-i is extracted from a variety of techniques-in addition to a reaction plane method, two and four particle cumulant analyses are used 141 . As we noted previously, the CGC p±_ distribution at high p± decays as \jp\ while the data is flat up to p± ~ 10 GeV. Recently,
352
E. Iancu and R.
Venugopalan
non-flow two particle correlations in the CGC model were suggested as an explanation of the vi data 142 . It is unclear at present whether non-flow correlations can explain other features of the measured azimuthal anisotropy. In the approach reported previously, a procedure very similar to the experimental approach can be followed and two and four particle correlations can be determined 120 . It is not yet clear whether the results of this study would affect our above conclusion about the relevance of final state effects in heavy ion collisions.
5.4. Equilibration
and the Quark Gluon
Plasma
An outstanding issue in the physics of heavy ion collisions is whether the hot and dense matter formed equilibrates to form a Quark Gluon Plasma (QGP). Hydrodynamic models, which assume local equilibrium, have been used to study the evolution of the QGP and are successful in describing some of the data at RHIC 143 . It is however not easy to show from first principles that thermalization does occur in the extreme conditions of a high energy heavy ion collision. Even though the initial density is very high, the system is expanding and becoming dilute very rapidly; moreover, if the momenta of the gluons is large, asymptotic freedom dictates that the cross-section of the gluons is small. In the CGC approach, we have argued that the energy densities right after the collision are very high-e ~ 10 GeV/fm 3 at RHIC and e ~ 60 GeV/fm 3 at LHC energies. These energy densities are formed very rapidly after the collision-about Tforrnauon = 0.4 fm for RHIC and Tformation = 0.25 fm at LHC using the values of Qs from the Golec-Biernat parametrization. At the high energies of interest, a heavy ion collision in the classical picture is boost invariant s . The dynamics of particle production at early times is therefore purely transverse. Consider a narrow slice in space-time rapidity, around the central rapidity r\ = 0. Partons with any significant longitudinal momentum pz will not be found in this slice (over time scales of interest)-they have only transverse momentum p± ~ Q3. Our classical dynamics only describe the initial transverse dynamics of partons. In Fig. 27, we see that at a time of ~ 20A S OT (~ 2.8 fm), e = Txx + Tyy. This suggests that the partons that were strongly interacting in the transverse plane are free-streaming by this time.
"This assumption is modified by quantum effects-however, at central rapidities, boost invariance is a good approximation.
The Color Glass Condensate
and High Energy Scattering
in QCD
353
It is unlikely however that the classical approach is valid at these late times. Shortly after partons are formed ~ 1/QS, they begin to scatter off each other and off the transverse plane. The physics of this process is not contained in the classical picture. A first guess is to see if small angle elastic 2 —> 2 scattering will lead to thermalization. Analytical estimates 144 suggest and numerical solutions of the Boltzmann equation confirm 145 that the time scale for equilibration due to small angle scattering is ~ exp(l/y / o7)/Q s parametrically. In practice, this time scale is not much smaller than the lifetime of the system. Also, as pointed out in Ref. 146, the pressure computed from these simulations is smaller than that of an equilibrated QGP. It was suggested by Baier et al. 34 that inelastic 2 —» 3 scattering may provide the dominant mechanism driving the system towards equilibrium. Naive power counting suggests that this process is suppressed relative to elastic scattering by a power of a3-however, the 2 —> 3 process is more efficient in re-distributing momenta and changing particle number. This scenario for thermalization, termed "bottom-up" on account of our particular initial conditions, can be outlined briefly as follows. As we discussed previously, the classical fields can be linearized into partons for r > (Q s ) _ 1 The parton distribution can be expressed as in Eq. (5.24) with the functional form of the distribution given in Eq. (5.25). This description, in weak coupling, is valid (parametrically) up to a proper time r ~ (ois)~3^2/QsThis is the time at which, on account of the longitudinal expansion, the occupation number of partons / < 1. In a small slice in rapidity, pz ~ 1/r which is small for r > 1/QS- However, pz is built up gradually through random multiple scattering: p\ = m^Ncou, where mjj is the Debye mass and Ncou is the collision rate. Simple estimates of these give self-consistently Pz ~ Qs/iQsT)1^3- Thus even though pz decreases with T, it does so at a slower rate than given by free streaming. The occupation number of the hard gluons (with py ~ Qs) is then dNh 1
dzSAQ*pz
^
Q*
~ a3(QsT)
1
Q?
(QST)V* Qs
'
V"*1'
which gives / < 1 for r > a 7 ' /Qs. This estimate was made assuming one had only elastic 2 —> 2 scattering. There is also inelastic 2 —> 3 scattering going on but the number of these soft gluons Ns « Nh at these times. When / < 1, this inequality continues to hold but the soft gluons begin to dominate the contribution to the Debye screening mass. One obtains Ns ~ Nh at Q3T = ( a s ) - 5 / 2 .
354
E. Iancu and R.
Venugopalan
For QST > (as)~5'2, most of the gluons are soft. These collide frequently with each other and achieve an equilibrated distribution with a temperature T. However, there are still some hard gluons left which continue to transfer energy to the soft gluons thereby acting as a heat bath. Indeed, as a consequence, the temperature of the soft gluons increases linearly initially even though the system as a whole is expanding. The temperature finally stops increasing when the hard gluons have lost all their energy. This happens at a time Q3T = (as)~13/5 at a temperature of T = aj Qs. The temperature subsequently decreases as r - 1 / 3 as one would expect for a fluid undergoing one dimensional expansion. The bottom-up scenario is an attractive one and the authors in Ref. 134 have show that the centrality dependence of the RHIC experimental data can be fit successfully. Nevertheless, a fully successful application of the idea to phenomenology is still remote. For the various stages of this scenario to be realized, as must be very small-likely, much smaller than may be realized at RHIC and perhaps even LHC energies. In the bottom-up scenario itself, the final temperature and the equilibration time are not determined up to a constant-however this number can in principle be determined within the theoretical framework itself. A number that needs to be determined externally is the liberation coefficient CAT. AS we described previously in section 5.2.3 and 5.2.4, we obtain CN = 0.3 — 0.5 depending on what Qs is. From empirical considerations, a larger value CN ~ 1 is favored in the bottom-up picture 1 3 4 . Even for a favorable choice of parameters which fit the data, one finds Tequu w R/2 where R = 6.8 fm is the radius of a Gold nucleus. While still smaller than the size of the system, a necessary condition for thermalization, it is not smaller by a large enough margin to make it appear inevitable. The long equilibration times discussed here are problematic for understanding the RHIC data on vi- Hydrodynamical models that fit the data require early thermalization times of Tequa ~ 0.6 fm. The v-i generated by the CGC alone is not sufficient to explain the data. The correct way to treat the theoretical problem may be as follows. Hard modes with A;j_ > As linearize on very short time scales T ~ 1/QS. Their subsequent evolution is treated incorrectly in the classical approach, which has them free streaming in the transverse plane. In actuality, as discussed here, they are scattering off each other via elastic gg —> gg and inelastic gg <-» ggg collisions which drive them towards an isotropic distribution 34 . This dynamics would indeed provide an additional pre-equilibrium contribution to v-i and is calculable. An effect to consider here would be the possible screening of infrared diver-
The Color Glass Condensate
and High Energy Scattering
in QCD
355
gences in the hard scattering by the time dependent classical field. More complicated is the effect of these hard modes on the classical dynamics of the soft modes and on their possible modification of the contribution of the latter to V2- One has here a little explored dynamical analog to the interplay of hard particle and soft classical modes in the kinetic theory of Hard Thermal Loops 21>147. The overlap between the classical field and Boltzmann pictures in this context has been discussed recently in Ref. 35. A practical issue of interest is to use the results of the classical field simulations as the initial input to parton cascade models which simulate the later stages of heavy ion collisions 148 .
5.5. Proton-Nucleus & Peripheral collisions at RHIC and LHC
Nucleus-Nucleus
Proton-Nucleus collisions at RHIC and LHC will provide an excellent probe of the properties of the CGC. Deuteron-Gold collisions with the center of mass energy of ,/s = 200 GeV/nucleon at RHIC started in January 2003 and Proton-Nucleus collisions are an important part of the program at the upcoming LHC collider at CERN 87 . A big difference between p/D-A collisions and A-A collisions is that final state interactions are more important in the latter than in the former. Thus p/D-A collisions provide an important benchmark to disentangle novel phenomena such as the Quark Gluon Plasma, which arise as a consequence of strong final state interactions, from initial state phenomena which may result from the Color Gluon Condensate. A case in point is the remarkable observation of the suppression of the single particle spectra, as a function of p±, in Au-Au collisions relative to that in pp-collisions (per binary collision) at y's = 130 GeV/nucleon and 200 GeV/nucleon 137 . An interpretation is that highpj. partons traversing a QGP suffer significant energy loss resulting in fewer high p± partons 149 . If correct, the observed suppression is evidence that a QGP has been created in high energy heavy ion collisions. In contrast, in p/D-Au collsions, it is argued that one will see the Cronin effect-the ratio of d-Au relative to ppcollisions will show an enhancement at moderate p±, peaking at p± ~ 3 — 4 GeV before going down to unity at larger p± 138 . The argument is that energy loss in the "cold matter" of d-Au collisions is quite small and the p± broadening due to multiple scattering causes the Cronin effect. On the other hand, if no Cronin effect is seen and a suppression is seen instead it may be an indication that initial state effects are responsible for the phenomenon
356
E. lancu and R.
Venugopalan
in both d-Au and Au-Au collisions. In this section, we will briefly summarize recent work on proton-nucleus collisions in the CGC framework. Proton-Nucleus collisions in this framework were first considered by Kovchegov and Mueller 27 . The proton in this case was modelled by a gauge invariant gluonic current. They showed that one obtains a Glauber-type formula with a saturation scale Qs (of the nucleus) for the inclusive gluon cross-section. The problem was looked at in more detail in Ref. 40. The problem of proton-nucleus scattering was considered in a manner similar to the classical fields treatment of nuclear scattering 29 ' 30 , except now they introduced two saturation scales Qs\ and QS2 for the nucleon and nucleus respectively with Qs\ « QS2- They were able to obtain analytical solutions for classical gluon production in the regions k± > QS2 > Qsi as well as for QS2 > k± > Qs\. In the former, one obtains dN/d2k±d2b oc Q s i Q ^ A i while in the latter case they obtained dN/dkLd2b oc Q2sl\n{k\/Q2sl)/k\. In this latter regime, one is solving the classical equations to all orders in Q ^ / ^ l - No analytical solution is available for k± < Qsi < Qs2, even when k± > A-QCD- The classical problem discussed here can also be formulated numerically 119 . The k± dependence of the two kinematical regions is seen clearly in the numerical result. It was also argued in Ref. 40 that since the rapidity distributions, for fixed k±, in one or the other regime are so different, these rapidity distributions could be used to isolate the Renormalization Group (RG)-evolution of the partons in both the proton and the large nucleus. The proton fragmentation region provides an excellent probe of the CGC. Final states measured in this region are produced by the scattering of high x partons in the proton off very small x partons in the nucleus. The scattering can be described as the convolution of the probability to find a quark in the proton q{x\,Q2) times the probability for the quark to scatter off the classical field of the nucleus characterized by a saturation scale QS2{x2) 4 1 , 4 2 . As discussed in Ref. 40, the k± distribution for this scattering is modified from the usual tree level pQCD distribution-this modified distribution, when convolved with the appropriate fragmentation functions will be reflected in hadronic final states 41>150. Electromagnetic final states such as photons and di-leptons are a particularly sensitive probe of saturation dynamics in the proton fragmentation region 42 . Interestingly, the modified p± distributions lead to p± broadening of the final state and the Cronin effect 41>42. This effect is also seen in other (dipole) models of saturation 151 . Though the CGC picture gives rise to the Cronin effect (which can be
The Color Glass Condensate
and High Energy Scattering
in QCD
357
represented by multiple scattering tree level diagrams), quantum effects in the wavefunction can modify this picture significantly. In sections 4.3 and 5.3 respectively, we discussed the CGC+geometric scaling mechanism which was first applied to understand the HERA DIS data and subsequently the suppression of p± spectra of charged hadrons in Au-Au collisions at RHIC. In the geometric scaling regime, the anomalous dimensions for the evolution are very close to BFKL anomalous dimensions 46 ' 94 -this change in the anomalous dimensions (from the DGLAP one) is what causes the suppression of the p± spectra in the initial state CGC scenario. A similar suppression must then persist in p/D-A collisions. In Ref. 39, it is predicted that the dependence of semi-hard processes on the number of participating nucleons of the nucleus in D-A collisions will be ~ ( A ^ J 1 / 2 . The number of participants are determined a la the Glauber calculation of Kharzeev and Nardi 3 6 . Thus more quantitatively, the suppression in D-Au collisions (relative to pp ) at y/s = 200 GeV/nucleon due to the geometric scaling initial state effect translates into a 25-30% of moderately highpj, particles in the top 15% centrality range. This conclusion will be modified by the Cronin final state scatterings of Refs. 41, 42 — these will change the behavior and one may even recover an enhancement. A quantitative calculation including both geometric scaling and final state scattering is urgently required. Very recently, in Ref. 152, quantitative predictions (along the lines of the qualitative ones in Ref. 40) were made for hadron multiplicities in DeuteronGold scattering at RHIC-in particular for the rapidity and centrality dependence. The results of the current Deuteron-Gold run should therefore help determine whether a Color Glass Condensate is formed already at the moderately high energies at RHIC. An important part of the RHIC program is the study of peripheral nuclear collisions. Very intense electromagnetic fields are created at these energies and a variety of photon-photon and photon-Pomeron final states can be studied 153 . Of particular interest to us is inclusive and diffractive QQ-'m the color field of a nucleus 154 . The problem is analogous to the photo-production of heavy quark pairs in Deeply Inelastic Scattering. The transverse momentum and invariant mass distribution of quark pairs can be shown to depend sensitively on the saturation scale Q3. Thus ultra-peripheral nuclear collisions provide an independent method to extract properties of Color Glass Condensate.
358
E. Iancu and R.
Venugopalan
Acknowledgements We would like t o t h a n k our colleagues a n d friends, J. P . Blaizot, A. Dumitru, E. Ferreiro, F . Gelis, K. Itakura, J. Jalilian-Marian, D. Kharzeev, Y. Kovchegov, A. Krasnitz, A. H. Mueller, Y. Nara, D. Teaney a n d H. Weigert for their insights and collaborations over t h e years. We are especially grateful to L. McLerran whose approach to physics has h a d a large influence on our work in general and on this review in particular. R.V.'s research was supported by DE-AC02-98CH10886 and by t h e R I K E N - B N L Research Center at BNL.
References 1. P. A. Dirac, Rev. Mod. Phys. 21 (1949) 392. 2. L. Susskind, Phys. Rev. 165, 1535 (1968). 3. S. J. Brodsky, H. C. Pauli and S. S. Pinsky, Phys. Rept. 301, 299 (1998); R. Venugopalan, arXiv:nucl-th/9808023. 4. M. Froissart, Phys. Rev. 123 (1961) 1053. 5. A. Martin, Nuovo Cimento 42 (1966) 930; L. Lukaszuk and A. Martin, Nuovo Cimento 52 (1967) 122. 6. G. Sterman, An Introduction to Quantum Field Theory, Cambridge University Press, Cambridge, 1993; M.E. Peskin and D.V. Schroeder, An Introduction to Quantum Field Theory, Addison-Wesley, New York, 1995. 7. J. Breitweg et. al., Eur. Phys. J. 67, 609 (1999) and references therein. 8. L.N. Lipatov, Sov. J. Nucl. Phys.23 (1976), 338; E.A. Kuraev, L.N. Lipatov and V.S. Fadin, Sov. Phys. JETP 45 (1977), 199; YaYa. Balitsky and L.N. Lipatov, Sov. J. Nucl. Phys. 28 (1978), 822. 9. J.R. Forshaw and D.A. Ross, Quantum Chromodynamics and the Pomeron, Cambridge University Press, Cambridge, 1997. 10. V.N. Gribov and L.N. Lipatov, Sov. Journ. Nucl. Phys. 15 (1972), 438; G. Altarelli and G. Parisi, Nucl. Phys.B126 (1977), 298; Yu. L. Dokshitzer, Sov. Phys. JETP 46 (1977), 641. 11. L. V. Gribov, E. M. Levin, and M. G. Ryskin, Phys. Rept. 100 (1983) 1. 12. A. H. Mueller and Jian-wei Qiu, Nucl. Phys. B 2 6 8 (1986) 427. 13. J.-P. Blaizot and A. H. Mueller, Nucl. Phys. B289 (1987) 847. 14. L.McLerran and R. Venugopalan, Phys. Rev. D 4 9 (1994) 2233; ibid. 49 (1994) 3352; ibid. 50 (1994) 2225; A. Ayala, J. Jalilian-Marian, L. D. McLerran and R. Venugopalan, Phys. Rev. D 52, 2935 (1995); ibid. 53, 458 (1996). 15. A. H. Mueller, Nucl. Phys. B558 (1999), 285. 16. E. Iancu and L. McLerran, Phys. Lett. B510 (2001) 145. 17. L. McLerran, The color glass condensate and small x physics: 4 ' ec ~ tures, Lectures at 40th Internationale Universitatswochen fuer Theoretische Physik: Dense Matter (IUKT 40), Schladming, Austria, 3-10 Mar 2001, hep-ph/0104285.
The Color Glass Condensate and High Energy Scattering in QCD 359 18. E. Levin, Saturation at low x, hep-ph/0105205. 19. A. H. Mueller, Parton Saturation — An Overview, hep-ph/0111244, in QCD Perspectives on Hot and Dense Matter, NATO Science Series, Kluwer, 2002. 20. E. Iancu, A. Leonidov and L. McLerran, The Colour Glass Condensate: An Introduction, hep-ph/0202270, in QCD Perspectives on Hot and Dense Matter, NATO Science Series, Kluwer, 2002; E. Iancu, hep-ph/0210236, invited talk, Quark Matter 2002, Nantes, France, 18-24 July 2002. 21. J.-P. Blaizot and E. Iancu, Phys. Rept. 359, 355 (2002). 22. J. Jalilian-Marian, A. Kovner, A. Leonidov and H. Weigert, Nucl. Phys. B 5 0 4 (1997), 415; Phys. Rev. D 5 9 (1999), 014014. 23. E. Iancu, A. Leonidov and L. McLerran, Nucl. Phys. A 6 9 2 (2001) 583; Phys. Lett. B 5 1 0 (2001) 133; E. Ferreiro, E. Iancu, A. Leonidov and L. McLerran, Nucl. Phys. A 7 0 3 (2002) 489. 24. J. D. Bjorken, Phys. Rev. D27, 140 (1983). 25. Yu.V. Kovchegov, Phys. Rev. D 5 4 (1996), 5463; Phys. Rev. D 5 5 (1997), 5445. 26. J. Jalilian-Marian, A. Kovner, L. McLerran and H. Weigert, Phys. Rev. D 5 5 (1997), 5414. 27. Yu.V. Kovchegov and A.H. Mueller, Nucl. Phys. B529 (1998), 451. 28. D. Y. Grigoriev, V. A. Rubakov and M. E. Shaposhnikov, Nucl. Phys. B 326, 737 (1989); J. Ambjorn and A. Krasnitz, Phys. Lett. B 362, 97 (1995); G. D. Moore, Phys. Lett. B 412, 359 (1997). 29. A. Kovner, L. McLerran and H. Weigert, Phys. Rev D 5 2 3809 (1995); D 5 2 6231 (1995). 30. A. Krasnitz and R. Venugopalan, hep-ph/9706329, hep-ph/9808332; Nucl. Phys. B 5 5 7 237 (1999). 31. A. Krasnitz and R. Venugopalan, Phys. Rev. Lett. 84 (2000), 4309. 32. A. Krasnitz and R. Venugopalan, Phys. Rev. Lett. 86 (2001) 1717. 33. A. Krasnitz, Y. Nara, and R. Venugopalan, Phys. Rev. Lett. 87 (2001) 192302. 34. R. Baier, A.H. Mueller, D. Schiff and D.T. Son, Phys. Lett. B 5 0 2 (2001) 51. 35. A. H. Mueller and D. T. Son, arXiv:hep-ph/0212198. 36. D. Kharzeev, M. Nardi, Phys. Lett. B 5 0 7 (2001) 121. 37. D. E. Kharzeev, E. Levin, Phys. Lett. B 5 2 3 (2001) 79; D. Kharzeev, E. Levin and M. Nardi, arXiv.hep-ph/0111315. 38. L. McLerran, J. Schaffner-Bielich, Phys. Lett. B 5 1 4 (2001) 29; J. SchaffnerBielich, D. Kharzeev, L. McLerran and R. Venugopalan, Nucl. Phys. A 705, 494 (2002). 39. D. Kharzeev, E. Levin and L. McLerran, arXiv:hep-ph/0210332. 40. A. Dumitru and L. D. McLerran, Nucl. Phys. A 700, 492 (2002). 41. A. Dumitru and J. Jalilian-Marian, Phys. Rev. Lett. 89, 022301 (2002). 42. F. Gelis and J. Jalilian-Marian, Phys. Rev. D 66, 014021 (2002). 43. K. Golec-Biernat and M. Wusthoff, Phys. Rev. D 5 9 (1999), 014017; ibid. D 6 0 (1999), 114023; Eur. Phys. J. C20 (2001) 313. 44. D. N. Triantafyllopoulos, Nucl. Phys. B 648, 293 (2003).
360 E. Iancu and R. Venugopalan 45. A. M. Stasto, K. Golec-Biernat and J. Kwiecinski, Phys. Rev. Lett., 86, 596 (2001). 46. E. Iancu, K. Itakura and L. McLerran, Nucl. Phys. A 708, 327 (2002); arXiv:hep-ph/0205198. 47. J. D. Bjorken, J. B. Kogut and D. E. Soper, Phys. Rev. D 3, 1382 (1971). 48. A. H. Mueller, Nucl. Phys. B415 (1994), 373; ibid. B 4 3 7 (1995), 107; A. H. Mueller and B. Patel, Nucl. Phys. B 425, 471 (1994). 49. C. S. Lam and G. Mahlon, Phys. Rev. D 6 2 (2000) 114023; ibid. D 6 4 (2001) 016004. 50. R. V. Gavai and R. Venugopalan, Phys. Rev. D54, 5795 (1996). 51. J. Jalilian-Marian, S. Jeon and R. Venugopalan, Phys. Rev. D 63, 036004 (2001). 52. I. Balitsky, Nucl. Phys. B 4 6 3 (1996), 99; High-energy QCD and Wilson lines, hep-ph/0101042. 53. Yu. V. Kovchegov, Phys. Rev. D 6 0 (1999), 034008; ibid. D 6 1 (2000), 074018. 54. J. Jalilian-Marian, A. Kovner and H. Weigert, Phys. Rev. D 5 9 (1999), 014015; A. Kovner, J. G. Milhano and H. Weigert, Phys. Rev. D 6 2 (2000), 114005. 55. H. Weigert, Nucl. Phys. A 7 0 3 (2002) 823. 56. A. H. Mueller, Phys. Lett. B 5 2 3 (2001) 243. 57. A. H. Mueller, Nucl. Phys. B335 (1990) 115. 58. N.N. Nikolaev and B.G. Zakharov, Z. Phys. C49 (1991) 607, ibid. C53 (1992) 331. 59. L. D. McLerran and R. Venugopalan, Phys. Rev. D 59, 094002 (1999); R. Venugopalan, Acta Phys. Polon. B 3 0 (1999) 3731. 60. W. Buchmuller, M.F. McDermott and A. Hebecker, Nucl. Phys. B 4 8 7 (1997) 283, Erratum-ibid. B500 (1997) 621. 61. M. Braun, Eur. Phys. J. C16 (2000) 337. 62. I. Balitsky and A. V. Belitsky, Nucl. Phys. B 629, 290 (2002). 63. J. P. Blaizot, E. Iancu and H. Weigert, Nucl. Phys. A713, 441 (2003). 64. I. Balitsky, Phys. Lett. B518 (2001) 235. 65. E. Levin and K. Tuchin, Nucl. Phys. B 5 7 3 (2000), 833; Nucl.Phys. A 6 9 1 (2001) 779. 66. N. Armesto and M. Braun, Eur. Phys. J. C20 (2001) 517. 67. M. Lublinsky, E. Gotsman, E. Levin, and U. Maor, Nucl. Phys. A 696, 851 (2001); arXiv:hep-ph/0209074; E. Levin and M. Lublinsky, Nucl. Phys. A 696, 833 (2001); Phys. Lett. B 5 2 1 (2001) 233; M. Lublinsky, Eur. Phys. J. C21 (2001) 513. 68. K. Golec-Biernat, L. Motyka, and A. M. Stasto, Phys. Rev. D 65, 074037 (2002). 69. E. Iancu, K. Itakura and L. McLerran, arXiv:hep-ph/0212123. 70. A. H. Mueller, Phys. Lett. B 396, 251 (1997). 71. A. Hebecker and H. Weigert, Phys. Lett. B432 (1998) 215. 72. A. L. Ayala, M. B. Gay Ducati and E. M. Levin, Nucl. Phys. B 493, 305 (1997).
The Color Glass Condensate
and High Energy Scattering
in QCD
361
73. J. Baxtels, K. Golec-Biernat and H. Kowalski, Phys. Rev. D 66, 014001 (2002). 74. E. Gotsman, E. Levin, M. Lublinsky, U. Maor, E. Naftali and K. Tuchin, J. Phys. G 27, 2297 (2001);E. Gotsman, E. Levin, M. Lublinsky and U. Maor, arXiv:hep-ph/0209074. 75. L. Frankfurt, V. Guzey, M. McDermott and M. Strikman, Phys. Rev. Lett. 87, 192301 (2001). 76. S. Munier, A. M. Stasto, and A.H. Mueller, Nucl. Phys. B 6 0 3 (2001) 427. 77. J. Bartels, E. Gotsman, E. Levin, M. Lublinsky and U. Maor, arXiv:hepph/0212284. 78. W. Heisenberg, Z. Phys. 133 (1952) 65. 79. E.M. Levin and M.G. Ryskin, Phys. Rept. 189 (1990) 267. 80. E. Ferreiro, E. Iancu, K. Itakura and L. McLerran, Nucl. Phys. A 710, 373 (2002). 81. J.R. Cudell et al, Phys. Rev. D 6 5 (2002) 074024. 82. M. Kozlov and E. Levin, hep-ph/0211348; S.Bondarenko, M.Kozlov, and E.Levin, hep-ph/0303118. 83. A. Kovner and U.A. Wiedemann, Phys. Rev. D 6 6 (2002) 034031. 84. L. L. Frankfurt and M. I. Strikman, Phys. Rept. 160, 235 (1988). 85. V. N. Gribov, Sov. Phys. 29, 483 (1969); L. Frankfurt, V. Guzey and M. Strikman, J. Phys. G27, R23 (2001). 86. A. Capella, A. Kaidalov, C. Merino, D. Pertermann and J. Tran Thanh Van, Eur. Phys. J. C 5 , 111 (1998). 87. CERN Yellow Report on pA collisions at the LHC, in preparation. 88. K. J. Eskola, V. J. Kolhinen and C. A. Salgado, Eur. Phys. J. C 9 , 61 (1999). 89. M. Hirai, S. Kumano and M. Miyama, Phys. Rev. D 64, 034003 (2001). 90. K. J. Eskola, H. Honkanen, V. J. Kolhinen and C. A. Salgado, Phys. Lett. B 532, 222 (2002) 91. J. Jalilian-Marian and X. Wang, Phys. Rev. D 60, 054016 (1999). 92. E. Levin and M. Lublinsky, Nucl. Phys. A 712, 95 (2002). 93. A. Freund, K. Rummukainen, H. Weigert and A. Schafer, arXiv:hepph/0210139. 94. A. H. Mueller and D. N. Triantafyllopoulos, Nucl. Phys. B 640, 331 (2002). 95. A. H. Mueller, arXiv:hep-ph/0301109. 96. V.S. Fadin and L.N. Lipatov, Phys. Lett. B429 (1998) 127; G. Camici and M. Ciafaloni, Phys. Lett. B 4 3 0 (1998) 349. 97. G.P. Salam, J H E P 9807 (1998) 19; M. Ciafaloni, D. Colferai, Phys. Lett. B 4 5 2 (1999) 372; M. Ciafaloni, D. Colferai, and G.P. Salam, Phys. Rev. D 6 0 (1999) 114036. 98. A. H. Mueller, Nucl. Phys. B 6 4 3 (2002) 501. 99. M. Arneodo et al., in Proceedings of Future Physics at HERA, DESY, September 25th-26th, 1995, hep-ph/9610423. 100. See, for instance, proceedings of the 2nd eRHIC workshop, Yale University, April 2001, BNL report, BNL-52592. For more details on EIC, see http://www.bnl.gov/EIC. 101. R. Venugopalan, AIP Conf. Proc. 588, 121 (2001) [arXiv:hep-ph/0102087].
362
E. Iancu and R.
Venugopalan
102. A. Zee, F. Wilczek, and S. B. Treiman, Phys. Rev. D 10, 2881 (1974). 103. J. Bartels, K. Golec-Biernat and K. Peters, Eur. Phys. J. C17, 121 (2000). 104. A. Berera and D. E. Soper, Phys. Rev. D 53, 6162 (1996); L. Trentadue and G. Veneziano, Phys. Lett. B 3 2 3 , 201 (1994). 105. J. C. Collins, Phys. Rev. D 57, 3051 (1998). 106. E. Levin and U. Maor, hep-ph/0009217. 107. S. J. Brodsky, L. Frankfurt, J. F. Gunion, A. H. Mueller and M. Strikman, Phys. Rev. D 50, 3134 (1994); B. Z. Kopeliovich, J. Nemchick, N. N. Nikolaev and B. G. Zakharov, Phys. Lett. B324, 469 (1994). 108. L. Frankfurt and M. Strikman, Phys. Lett. B 3 8 2 (1996) 6. 109. Y. V. Kovchegov and L. D. McLerran, Phys. Rev. D 60, 054025 (1999), [Erratum-ibid. D 62, 019901 (2000)]. 110. A. C. Caldwell and M. S. Soares, hep-ph/0101085. 111. Y. V. Kovchegov, E. Levin and L. D. McLerran, Phys. Rev. C 63, 024903 (2001); L. Frankfurt and M. Strikman, Phys. Rev. Lett. 382, 6 (1996). 112. M. Strikman, M. G. Tverskoii and M. B. Zhalov, Phys. Lett. B459, 37 (1999). 113. B. Blattel, G. Baym, L.L. Frankfurt, and M. Strikman, Phys. Rev. Lett. 70, 896 (1993). 114. Y. V. Kovchegov, Nucl. Phys. A 692, 557 (2001). 115. Y. V. Kovchegov and D. H. Rischke, Phys. Rev. C56 (1997) 1084. 116. M. Gyulassy and L. McLerran, Phys. Rev. C56 (1997) 2219. 117. J. F. Gunion and G. Bertsch, Phys. Rev. D 25, 746 (1982). 118. A. Krasnitz, Y. Nara and R. Venugopalan, Phys. Lett. B 554, 21 (2003). 119. A. Krasnitz, Y. Nara and R. Venugopalan, Nucl. Phys. A 717, 268 (2003); arXiv:hep-ph/0209341. 120. A. Krasnitz, Y. Nara and R. Venugopalan, unpublished. 121. T. Lappi, arXiv:hep-ph/0303076. 122. A. Krasnitz, Y. Nara and R. Venugopalan, in preparation. 123. A. H. Mueller, arXiv:hep-ph/0206216. 124. D. Kharzeev, A. Krasnitz and R. Venugopalan, Phys. Lett. B 545, 298 (2002). 125. D. Kharzeev, R. D. Pisarski and M. H. Tytgat, Phys. Rev. Lett. 81, 512 (1998); D. Kharzeev and R. D. Pisarski, Phys. Rev. D 61, 111901 (2000). 126. J.-Y. Ollitrault, Phys. Rev. D46, 229 (1992). 127. S. Voloshin and Y. Zhang, Z. Phys. C70, 665 (1996); A.M. Poskanzer and S. Voloshin, Phys. Rev. C58, 1671 (1998). 128. H. Sorge, Phys. Rev. Lett. 78, 2309 (1997); Phys. Rev. Lett. 82, 2048 (1999). 129. Preliminary STAR data on centrality dependence of «2-we thank R. J. Snellings for providing us with the data. 130. D. Teaney and R. Venugopalan, Phys. Lett. B 539, 53 (2002). 131. K. Adcox et al. [PHENIX Collaboration], Phys. Rev. Lett. 86, 3500 (2001). 132. B. B. Back et al. [PHOBOS Collaboration], Phys. Rev. C 65, 031901 (2002). 133. K. Adcox et al. [PHENIX Collaboration], Phys. Rev. Lett. 87, 052301 (2001).
The Color Glass Condensate and High Energy Scattering in QCD 363 134. R. Baier, A. H. Mueller, D. Schiff and D. T. Son, Phys. Lett. B 539, 46 (2002). 135. R. C. Hwa and C. B. Yang, arXiv:nucl-th/0301004; arXiv:nucl-th/0302006. 136. M. A. Braun, F. Del Moral and C. Pajares, Phys. Rev. C 65, 024907 (2002); E. G. Ferreiro, F. del Moral and C. Pajares, arXiv:hep-ph/0303137; H. Satz, Lectures given at International School of Physics "Enrico Fermi": Course 153: From Nuclei and Their Constituents to Stars, Varenna, Lake Como, Italy, 6-16 Aug 2002, arXiv:hep-ph/0212046. 137. S. Mioduszewski [PHENIX Collaboration], arXiv:nucl-ex/0210021; G. J. Kunde [STAR Collaboration], arXiv:nucl-ex/0211018. 138. M. Gyulassy, I. Vitev, X. N. Wang and B. W. Zhang, arXiv:nucl-th/0302077. 139. STAR Collaboration, K.H. Ackermann et al., Phys. Rev. Lett. 86, 402, (2001). 140. S. A. Voloshin, arXiv.nucl-th/0202072. 141. N. Borghini, P. M. Dinh and J. Y. Ollitrault, Phys. Rev. C 64, 054901 (2001); Phys. Rev. C 63, 054906 (2001). 142. Y. V. Kovchegov and K. L. Tuchin, Nucl. Phys. A 708, 413 (2002). 143. P.F.Kolb, P. Huovinen, U. Heinz,and H. Heiselberg, Phys. Lett. B500, 232 (2001); D. Teaney, J. Lauret and E. V. Shuryak, Phys. Rev. Lett. 86, 4783 (2001). 144. A. H. Mueller, Nucl. Phys. B572 (2000) 227; Phys. Lett. B 4 7 5 220 (2000). 145. J. Bjoraker and R. Venugopalan, Phys. Rev. C63 024609 (2001); J. Serreau and D. Schiff, JHEP 0111, 039 (2001). 146. A. Dumitru and M. Gyulassy, Phys. Lett. B494, 215 (2000). 147. J. Jalilian-Marian, S. Jeon, R. Venugopalan and J. Wirstam, Phys. Rev. D 62, 045020 (2000). 148. S. A. Bass, B. Muller and D. K. Srivastava, arXiv:nucl-th/0207042; D. Molnar and M. Gyulassy, Phys. Rev. C 62, 054907 (2000); Y. Nara, S. E. Vance and P. Csizmadia, Phys. Lett. B 531, 209 (2002). 149. R. Baier, D. Schiff and B. G. Zakharov, Ann. Rev. Nucl. Part. Sci. 50, 37 (2000). 150. J. T. Lenaghan and K. Tuominen, arXiv:hep-ph/0208007. 151. J. Raufeisen, arXiv:hep-ph/0301052; B. Z. Kopeliovich, J. Raufeisen and A. V. Tarasov, Phys. Lett. B 503, 91 (2001); Phys. Rev. C 62, 035204 (2000). 152. D. Kharzeev, E. Levin and M. Nardi, arXiv.hep-ph/0212316. 153. G. Baur, K. Hencken, D. Trautmann, S. Sadovsky and Y. Kharlov, Phys. Rept. 364, 359 (2002). 154. F. Gelis and A. Peshier, Nucl. Phys. A 697, 879 (2002); Nucl. Phys. A 707, 175 (2002).
ELECTROMAGNETIC RADIATION FROM RELATIVISTIC N U C L E A R COLLISIONS
Charles Gale' and Kevin. L. Haglin§ 'Department of Physics, McGill University 3600 University St., Montreal, QC, H3A 2T8, Canada sDepartment of Physics, Astronomy and Engineering Science St. Cloud State University, St. Cloud, MN 56301, USA
We review some of the results obtained in the study of the production of electromagnetic radiation in relativistic nuclear collisions. We concentrate on the emission of real photons and dileptons from the hot and dense strongly interacting phases of the reaction. We consider the contributions from the partonic sector, as well as those from the nonperturbative hadronic sector. We examine the current data, some of the predictions for future measurements, and comment on what has been learnt so far.
364
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
365
Contents 1 2
Introduction Radiation from Hadronic Matter 2.1 The low dilepton invariant mass sector 2.1.1 A baseline calculation 2.1.2 Spectral density calculations 2.2 The dilepton intermediate invariant mass sector 2.3 Photons 2.3.1 General strategy 2.3.2 Establishing the rates 2.3.3 Refinements 2.3.4 Medium effects 2.3.5 Alternative approach: chiral reduction formulae 3 Radiation from Partons 3.1 Photons 3.1.1 Photon measurements 3.2 Dileptons 4 Predictions 4.1 Photons 4.2 Dileptons 4.3 Electromagnetic signatures of jets 4.4 Squeezing lepton pairs out of broken symmetries 5 Conclusions References
366 368 370 370 371 375 387 387 388 390 391 392 393 393 396 403 406 406 409 411 417 422 424
366
C. Gale and K. L. Haglin
1. Introduction The study of matter under extreme conditions constitutes a rich field of intellectual pursuit and is a vibrant research area of physics. It is popularised to nonspecialists by indicating that such studies reveal physics that governed the early Universe (microseconds after the Big Bang) and also continue today to govern physics of compact astrophysical objects (neutron stars and black holes). But it is indeed more than that. Practitioners concern themselves with a variety of very specific and technically challenging questions. When ordinary matter is heated to roughly a trillion degrees Kelvin, how does it respond? And what are the hallmark signatures of this response? What should one look for? When relatively cold matter is compressed to densities many times greater than that of normal nuclei, does it resemble something other than ordinary protons and neutrons? Quantum Chromodynamics (QCD) predicts under these extreme conditions, very far from the ground state, that matter will essentially change its properties to resemble a plasma of quarks and gluons (QGP) 1 , 2 . What is the QCD phase diagram? Indeed, the understanding of QCD under extreme conditions of high temperature or large baryon density has progressed considerably in recent years. Confinement, in the pure glue version of QCD, is a property that can be associated with a definite symmetry whose status is probed by the value of the Polyakov loop (L). This symmetry is valid at low temperature, but broken at high temperatures. In the limit of massless quarks, QCD is chirally symmetric, and that symmetry is valid at high temperatures and spontaneously broken at low temperatures. The order parameter there is the chiral condensate: (^ip). As a function of temperature, those order parameters are best studied on the lattice; although it is fair to say that the lattice has just started to venture into the finite baryon domain with any degree of quantitative assurance 3 . An idea on the status of finitetemperature lattice QCD can be had by consulting Ref. [1, 2]. The order of the transition actually depends on the details of the parameters of the theory, as shown in the "Columbia plot" (Fig. 1), but even in the absence of a robust prediction at finite temperature and baryon density, the lattice does provide tantalising clues of an eventually observable behaviour of the many-body nature of QCD. For example, it offers some support, together with effective models, to the discussions of a genuine tricritical point in the QCD phase diagram 4 . The experimental realisation of hot and dense strongly interacting mat-
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
367
Fig. 1. The nature of the QCD phase transition as a function of quark masses, along with theoretical expectations from effective models. From [2].
ter in terrestrial accelerators has been accomplished through the study of relativistic nuclear collisions. Producing quark gluon plasma (QGP) is one of the premier goals of the Relativistic Heavy Ion Collider (RHIC) and its broader experimental program and represents already enormous challenge; identifying the plasma and also studying its unique properties is yet something else, and has proven to be extremely challenging from both the theoretical and experimental points of view. Signatures of the QGP have been proposed: strangeness enhancement, suppression of the J/ip signal, effects of multiple collisions on the observed particle spectra, and electromagnetic radiation are but a few. Unfortunately, while most of the proposed signatures are plausible and evidently do occur at some level, it has been difficult to refine the mostly heuristic arguments into really precise predictions. It is now clear that certainty will be attained in this field through the simultaneous analysis of complementary observables. This being said, probes that do not interact strongly have the definite advantage of suffering little or no final state interaction: they open a privileged window to the hot and dense phases of the reaction. The price to pay is a small production rate. We will discuss in this article the status of theory and the current picture relative to the experimental data, surrounding electromagnetic radiation as probes of strongly-interacting many-body dynamics. This work is organised as follows: a brief review of the formalism germane to the emission of real
368
C. Gale and K. L. Haglin
and virtual photons from hot and dense systems is followed by a discussion of the radiation from the hadronic sector, then from the partonic sector. Some of predictions for current and future measurements are outlined, followed by a conclusion. The main thrust here pertains to models amenable to conventional experimental measurements: the high density low temperature phase of QCD 5 will not be discussed, even though its electromagnetic emissivity has been investigated 6 . A goal here is to provide the reader with a snapshot of this rapidly evolving field by discussing some of the recent theoretical and experimental developments. In doing so, it is unfortunately impossible to do justice to the wholeness of the body of work in this exciting area: we shall be brief on some topics and refer instead to the literature and in particular to previous reviews. 2. Radiation from Hadronic Matter The goal of this section of the text is to relate the spectrum of emitted radiation to some of the intrinsic properties of the strongly interacting matter. A thermalised medium is assumed, and the formalism below is developed in the one-photon approximation. Quite generally, the electromagnetic radiation from strongly interacting matter can be related to the imaginary part of the retarded in-medium photon self-energy, at finite temperature and density 7,8 ' 9 ' 10 ' 11 . We sketch a derivation 11 here, for the emission of real photons. Consider a transition i —> / 7 , i.e. from some initial hadronic state i to some final hadronic state / , plus a photon of momentum fcM = (w, k) and polarisation &. The transition rate is Rfi - —y~,
(1)
where r is the observation time, V is the volume of the system, and the S'-matrix element is
Sfi = (f\fd*xJll(x)A»(x)\i).
(2)
Jfi{x) is an electromagnetic current operator, and A"(x) = -j==
{eik-x + e~ik-x).
(3)
\I2UJV
Using translation invariance, summing over polarisations, and using the integral representation of the delta function, one can write
Rfi = -£y(27r)4[S\pi
+ k-pf)
+
S4(pi-k-Pf)]
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
x
369
(4)
where the parts having to do with emission and absorption are evident. The net thermal rate is obtained by summing the above over final states and taking a thermal average of the initial configurations. It is useful at this point to define some finite-temperature currentcurrent correlators 12 :
i
feW
=
fd'xe^Y.^M^Mxme-^/Z, J
i
/^eifc-^(i|[AM(x),^(0)]|i)e-^VZ.
fe(k)=
*'
(5)
i
The last correlation function is a retarded correlation function. The above all involve the electromagnetic current operator in the Heisenberg picture and are written in the grand canonical ensemble, where Z is the partition function. Assuming translational invariance, the first two can be rewritten together as fe(P°,P)
= £(27r) 4 <5 4 ( P / -Pi
±p)(i\All(0)\f)
(f\Av(0)\i)e-^/Z.
(6)
i,f
Clearly, /-* is involved with absorption of radiation, whereas f< deals with emission. Only, the latter case is treated here. Further defining a spectral density p^ = />„ - /<„, one may first show that f< =
PiiV
(7)
and also that R
i
_
^~lJ
f du
pnV{w,k)
2^k°-w
+ ie-
(8)
With the above elements, one can finally write pM„ = 2 I m / ^ . The emission probability is related to the imaginary part of the finite-temperature retarded current-current correlation function. In the one-photon approximation (i.e. to lowest order in e 2 ), the time-ordered current correlator is the one-particle irreducible photon self-energy, n M „ 1 3 . Putting all of this together, the differential rate for emitting real photons is
370
C. Gale and K. L. Haglin
The proof is easily generalised to the case of lepton pair emission (the lepton mass has been set to zero): E+E
-*WhZ
= (£)*F W- +W--ITP+ -P-] ™U*) 1
(io)
•
ePw - 1 '
Thus, the electromagnetic signal emitted during a nuclear reactions can be related to a quantity that is linked to properties of the medium itself. As it shall be seen, the retarded self-energy (or alternatively the currentcurrent correlator) will be modified in a strongly interacting environment. Also, the current-current correlator is calculable only perturbatively, unless a specific model is available: this again testifies to the importance and the value of electromagnetic measurement in nuclear collisions. They indeed open a window to the hot and dense phases of the reaction, and those regions can't be probed directly by other means. 2.1.
The Low Dilepton
Invariant
Mass
Sector
2.1.1. A Baseline Calculation As a calculation to set the scale of the physical processes under consideration, it is useful to consider first the following question: what is the magnitude of the radiation emitted by a hot gas of mesons? Specialising to the lepton pair sector 14 , this problem is briefly summarised here. Using relativistic kinetic theory, the lepton pair emission rates can be calculated with the help of effective interaction Lagrangians. The parameters of those effective Lagrangians are fitted to radiative decays measurements using vector meson dominance (VMD). Specifically, the "calibration" reactions are: p -> 7T7, K*± - • A- ± 7 , K *° (K*°) -> K°(K*°)j, u -» 7r°7, p° -+ 777, rf —* p°7, rf —> W7, <j) —> T77, <j> —> r],ry, and <j> —> ir°j. Note that the usage relativistic kinetic theory here is tantamount to evaluating the finite temperature photon self-energy at the one-loop level11. The rate for ab —> e+e~ is given by p nab_e+e-
d3 d d y f P° *Pb *P+ jv j 2 E a ^ 3 2Eb(27r)Z 2E+{2n)3 x\M\2(2n)45A(Pa+pb-P+-p-),
d3
Pf f 2E-{2^JaJb (11)
where the / ' s are appropriate distribution functions, and M is an overall degeneracy factor dependent upon the specific channel. The PP —> e+e~,
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
371
PV —» e+e~, and VV —> e + e _ reactions can be obtained from the radiative decay ones through crossing symmetry. The sum of them is shown in Fig. 2. Also shown is the contribution from the IT+IT~ —> e+e~~ reaction. This channel has often been considered as the sole source of hadronic dileptons in early calculations, owing mainly to multiplicity arguments. Even in this incoherent sum approach, one can see that this assumption is badly violated in the low mass region.
0.3 0.4 0.5 0.6 0.7 0.8 0.9
1.0
M (GeV) Fig. 2. The rates for dielectron emission from an incoherent sum of meson reactions 1 4 . The three sets of full curves are the net rates at a temperature of 100, 150, and 200 MeV, from bottom to top. The dashed curves represent the 7r+7r - —> e+e~ contribution only.
2.1.2. Spectral Density Calculations The full power of the formalism derived in section 2 reveals itself when it is made clear that the electromagnetic radiation rate is related to the in-medium vector meson spectral density: this quantity is of course not measurable directly. The spectral density is related to the imaginary part of the full propagator, by definition. It appears when the electromagnetic current operator in the retarded self-energy is expressed as the field operator, through VMD 1 8 ' 1 9 , 2 0 . We illustrate the spectral density approach here, by describing a calculation again done in the baryonless regime 21 . One may start with the model for the p-meson in free space employed
372
C. Gale and K. L.
Haglin
previously in Refs. [15, 16, 17]. Based on the standard pirn interaction vertex (isospin structure suppressed), •CpTTTT = ffpTTTT K P^^Pfl
.
(12)
M
(p : pion momentum) the bare p-meson of mass m°p is renormalised through the two-pion loop including a once-subtracted dispersion relation, giving rise to the vacuum self-energy Sj„(M) = S ° „ ( M ) - E ^ ( 0 ) ,
SJU(M) = / ^
WP)2
GUM,P)
,
(13)
with the vacuum two-pion propagator
and vertex functions
p2 V " r ( p ) = V o 9pn-K %P FpiTn(p) , V
(15)
*J
involving a hadronic (dipole) form factor Fp7r7r 17 . Resumming the two-pion loops in a Dyson equation gives the free p propagator D%M) = [M 2 - (m°) 2 - S ^ ( M ) ] " 1 , (16) which agrees well with the measured p-wave rnr phase shifts and the pion electromagnetic form factor obtained within VMD. To calculate medium corrections to the p self-energy in a hot meson gas, one can assume that the interactions are dominated by s-channel resonance formation. It is then possible to use the formal relationship that relates the self-energy to the forward scattering amplitude, integrated over phase space at finite temperature. A field-theoretic derivation of this connection can be found in [22]. At moderate temperatures relevant for the hadronic gas phase, the light pseudoscalar Goldstone bosons P = n, K are the most abundant species. The various resonances in pP collisions can be grouped into two major categories, namely vector mesons V and axialvector mesons A. The effective Lagrangians that regulate the interactions among all those species have their parameters chosen such that the measured hadronic phenomenology are reproduced 21 . This statement also holds true for hadronic form factors. In order to calculate the p spectral density at moderate temperatures, one includes the hadronic fields appearing in Table 1. The interaction vertices being completely determined, one may
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
373
M[GeV]
Fig. 3. The real and imaginary parts of the polarisation-averaged p self-energy (lower and upper panel, respectively). The different channels are labelled explicitly and are explained in the t e x t 2 1 . Note that the 7nr channel is absent for the sake of viewing clarity.
Table 1. Mesonic resonances R with masses ITIR < 1300 MeV and substantial branching ratios into final states involving direct p's (hadronic) or p-like photons (radiative). Taken from Ref. [21]. R u(782) h i (1170) ai(1260) Ki(1270) /i(1285)
IGJP
TT'(1300)
l-o-
o-io-i+ 1-1+
±1+ 2 + 0+1
Ttot [MeV] 8.43 ~360 ~400 ~90 25 ~400
ph Decay pn pit pit
pK PP pK
r°h [MeV]
F°h [MeV]
~5 seen dominant ~60 <8 seen
0.72 ?
0.64 ?
1.65 ?
first calculate the in-medium p self-energy, then the complete propagator in terms of its longitudinal and transverse parts 1 1 . The real and imaginary parts of the in-medium p self-energy are shown in Fig. 3. Each curve is labelled according to that species which interacts with the p. It is instructive to observe that the imaginary parts all add, while there is a significant amount of cancellation of the real parts. The first effect creates a sizeable
374
C. Gale and K. L. Haglin
-10
-8
O
°" -4 E
-2
"0.0
0.2
0.4
0.6 M [GeV]
0.8
1.0
1.2
Fig. 4. Imaginary part of the p propagator in the vacuum (solid curve) and in a thermal gas including the full in-medium self-energies for fixed three-momentum q = 0.3 GeV at temperatures T = 120 MeV (long-dashed curve), T = 150 MeV (dashed curve), and T = 180 MeV (dotted curve). The figure comes from Ref. [21].
width for the in-medium p, while the second determines the in-medium mass, which appears to be only slightly modified. Those different aspects are again seen in the representation of the imaginary part of the in-medium p propagator, shown in Fig. 4. The calculations of the in-medium vector meson spectral densities clearly show the richness of the many-body problem under scrutiny. The power of the techniques described above becomes evident when they are combined with dynamical models and confronted with experimental data. This story is well chronicled in [23, 24], and in references therein. See also [25]. The current situation can be summarised by writing that the low dilepton invariant mass data 2 6 ' 2 7 can be understood in terms of in-medium modifications of vector spectral densities. These data can not empirically exclude, however, other interpretations 28 ' 23 ' 29 . This unfortunate situation still prevails at a lower energy 30 , as shown in Fig. 5. The sources there are: free hadron decays without p decay (thin solid line), calculation with a vacuum p spectral density (thick dashed line), dropping in-medium p mass 28 (dash-dotted line), and with a medium-modified p spectral density 24 (thick solid line). Note that but the persistence of the dilepton excess at lower energies does support a baryon density-driven effect. Other suggestions to resolve the different models require high statistics 23 : the final analysis of
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
375
-4
10
\ Pb-Au 40 AGeV
o / o ^ . 30 %
> 2 5 5 10 o o A •g
Z*10 y
-«
A
$
E
V° z V •8
10
-9
10
•
0
0.2
0.4
0.6
0.8
1
1.2
mM(GeV/c2) Fig. 5. Dilepton spectrum from Pb-Au collisions at 40 AGeV/c. See the text for the meaning of the different curves. Data are from [30].
the CERES 2000 data with the TPC is eagerly anticipated. 2.2. The Dilepton
Intermediate
Invariant
Mass
Sector
Intermediate mass lepton pairs have traditionally been a focus of interest, as their spectrum has been suggested early on as a signature of the quarkgluon plasma 31 . In relativistic nuclear collisions, measurements have been carried out at SPS energies by the HELIOS-3 and the NA38/NA50 collaborations in the lepton pair invariant mass range m^ < M < mj/^. Both experimental collaborations have observed significant enhancement of dilepton yield in this region for central S + W and S + U collisions as compared to those in proton-induced reactions (normalised to the charged-particle multiplicity) 32 ' 33 . Chronologically, HELIOS-3 reported on the intermediatemass enhancement first. This experiment was designed to study virtual photons in the dimuon sector at low transverse mass. In this way, dimuon production was studied from threshold up the J/tp mass over a wide range in PT. A good summary of this experimental situation is shown in Fig. 6. Several explanations of the intermediate mass dimuon enhancement have been put forward. Those include additional production of cc pairs 34 , secondary Drell-Yan emission35, and charmed meson rescattering 36 . Note that in principle, all of those effects can coexist. However, such a global modelling has not been done, and we thus discuss them separately. It is fair
376
C. Gale and K. L. Haglin
io-»
10-'
I &
10-»
2
io-»
lO-io 1
1.5
3 M (GeV)
2.5
Fig. 6. Comparison of lepton-pair yield divided by the multiplicity of charged particles, in p + W and S + W collisions at 200 A GeV/c. The data are from [27],
to say that the first of those mentioned above is still admitted by current experimental data 3 7 , even though some calculations are slightly less optimistic 38 . The role of the last ingredient on the list has not been found to be large enough to account for the experimentally observed excess 39,37 . Another class of approaches consists of quantitative evaluation of thermal dilepton sources. Those may be from the hadronic, confined sector of QCD, and/or from the quark gluon plasma itself. One such model is described below. Recall that thermal hadronic sources have been shown to be crucial in the low mass sector. It is therefore legitimate to ask how high in invariant mass is the extent of the virtual photon radiation from those sources? Those concerns are carried to their logical conclusion in what follows. In the intermediate invariant mass region, relativistic theory estimates indicate that the following microscopic channels are relevant: TTTT —> l+l~, •np -> l+l~, TTCJ -> l+l~, Trai -> l+l~, KK -> l+l~, and KK*+ c.c. -> / + / [40]. Apart from sheer coupling constant values, the importance of those contributions stems simply from considering energy scales involved and
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
377
from phase space arguments. This combination of coupling constants and phase space is effective in maximising a particular contribution from the 7rai channel, for example 41 . Calculations of dilepton-emitting processes in the intermediate invariant mass region follow similar steps to those in the low mass sector. Effective Lagrangians are used, together with VMD, and the coupling constants and possible form factors are fitted to measured strong decays and electromagnetic radiative decays. Only, in the intermediate mass domain an extrapolation is required. The strong decay widths set a scale that is typically an order of magnitude below the mass region of interest: 1 GeV < M < 3 GeV. The radiative decays are even smaller, owing to the size of a e m [42]. The required extrapolation is then vulnerable to off-shell effects. Put another way, there is a risk of uncontrolled growth of form factors since the application region is far removed from the region where the empirical fitting was realized. Indeed, different Lagrangians known to agree in the low mass sector generating dileptons were found to differ significantly in their predictions of emission rates for intermediate mass lepton pairs 43 . Fortunately, there exists a wealth of data for e+e~ —-> hadrons, exactly in the invariant mass window relevant for this application. Those data can thus be used to extract an effective form factor for the inverse reactions. Alternatively, they may also be used to extract spectral densities: this point will be discussed later. As an example, consider e+e~ —> 7r+7r~, which has been measured with high accuracy 44 ' 45 , evidenced by the data shown in Fig. 7. The cross section for this reaction can be written as
(7(e+e -^^ 7r - ) =
^g!| i r 7 r ( M ) |2 )
(17)
where k is the three-momentum in the two-body rest frame, M is the leptonpair invariant mass, and Fn is the time-like pion electromagnetic form factor. With these data, one can extract Fn, and then use it in the calculation of the dilepton-producing reaction 7r+7r~ —> l+l~: , ±
,-i-, N +
87ra2/clTn .„,,,<>/' 2
m? \ /
2mf\
*{*+*- - l n = -^5-|F.(M)| ^1 - j±J (l + -^j
.„„.
,
(18)
where mi is the lepton mass. A similar procedure can be followed for other channels. Another example appears in Fig. 8, where the time-like electromagnetic form factors for the kaon systems have been extracted from electron-positron annihilation data. Both processes introduced above are of the pseudoscalar-pseudoscalar type. For the pseudoscalar-vector class, in the invariant mass region of interest, np —> l+l~, KK* + c.c. —> l+l~, and
378
C. Gale and K. L. Haglin
1000 ra
2
2.5
Fig. 7. The cross section for e + e - —» n+ir~. The solid curve is based on the model of Ref. [46]. The experimental data are from the OLYA collaboration 4 4 , the CMD collaboration 4 4 , and the DM2 collaboration 4 S .
•KW —> l+l~ are included. The first two processes effectively involve three pions, while the third one involves four pions. Note that in a transport approach, a process involving three or more pions in the initial state can only be described as a two-step process with an intermediate resonance. The first two channels above have been studied in Ref. [50]. The effective form factors one extracts are shown in [40]. Details about the TTU> channel are gotten from the study of four-pion final states. Using a Wess-Zumino VVP interaction Lagrangian, one finds a{1l"u-.l+l-)
=
^\FUM)\2
(19)
in the limit of vanishing lepton mass. The form factor may be parametrised in terms of three isovector p-like vector mesons, p(770), p(1450), and p(1700) 51 : Di
•F™(M) =
^ffy,
ml ' (rriy - M2) -
imvT\
(20)
Electromagnetic
Radiation from Relatimstic
1.2
1.4 1.6 M (GeV)
Nuclear Collisions
379
1.8
Fig. 8. The kaon electromagnetic form factor. The solid and dotted curves 4 6 are for the charged and neutral kaons, respectively. The symbols are for the charged kaon data from the CMD-2 collaboration 4 7 , the DM2 collaboration 4 8 , and the OLYA collaboration 4 9 .
This form is then used to fit the experimental data from the ND and ARGUS collaborations. The result is shown in Fig. 9. In the pseudoscalar axial-vector channel, we shall consider mainly •na\ —• l+l~, which is in effect a four-pion process. Considering here again the reaction where the lepton pair constitutes the initial state and the hadrons the final, one can attempt an extraction of an effective form factor. Some previous thermal rate calculations indicate that this specific channel is particularly important in the intermediate mass region 41 , even though it is difficult to calculate reliably a specific signal using effective Lagrangians. This fact owes mainly to off-shell effects43. One can pick a model that yields adequate hadronic phenomenology on-shell, and then extrapolate to the intermediate mass sector with the help of experimental data. Using a chiral Lagrangian where the vector mesons are introduced as massive Yang-Mills fields53 one may derive the following cross section cr(7rai —> ZZ) =
IWLH 72mll92pM5kn
M2;
1 +
2mf M2"
(21)
380
C. Gale and K. L. Haglin
C
1
1.2 1.4 M (GeV)
l.B
Fig. 9. The cross section for e+e —» 7r°oj. The solid curve is described in the text. The experimental data are from the N D S 1 and ARGUS 5 2 collaborations.
where TC is a nontrivial function of coupling constants, masses and momenta. kn is the magnitude of the pion momentum in the centre-of-mass. The issue of the electromagnetic form factor \F-Kai | 2 , can be settled, at least in principle, by analysing e+e~ —» n+ir~ir+Tr~ and e+e~ —> 7r+7r_7r07r° data. Although many such analyses have been carried out, an unambiguous result is still elusive, as many other intermediate states may contribute to the same four-pion final state. Several scenarios have been considered, and a discussion appears in [40]. What is probably a conservative estimate is highlighted here. The DM2 collaboration has determined the cross section ae+e-_>nai using a partial wave analysis (PWA) 54 . One may extract an effective form factor from these data, see Fig. 10, and then carry out an analysis for c„a ti, using detailed balance. Even in a careful analysis of the relevant intermediate invariant mass dilepton reactions, some concerns remain. These mainly stem from the need to account for all the sources of electromagnetic radiation. In kinetic theory approaches, risks exist of double-counting and possible omissions. In an attempt to bypass those, an approach which allows for a nonperturbative
Electromagnetic
20
Radiation from Relativistic
Nuclear Collisions
381
~\—'—I—'—I—'—I—'—I—r~
o
DM2 PWA our fit
15 -
Fig. 10. T h e cross section for e + e ~ —• 7rai. The open circles are the experimental data from the DM2 collaboration using a partial wave analysis 5 4 . The solid curve is a fit to the data.
treatment of the strong interaction and avoids a detailed enumeration of reactions was devised 55 . The dilepton emission rate is interpreted in terms of spectral functions of hadronic currents, tabulated from low energy e+e~~ annihilation reactions and from r lepton decays. A differential rate expression, obtained in the chiral (mw —> 0) limit, reads 55 dR dM2
4a 2 2TT
MTKX{M/T) pem(M)
\
) (PV(M)
PA(M))
(22)
where T is the temperature, e = T2/6F%, M is the dilepton invariant mass, and the superscripts on p denote the electromagnetic, vector, and axial spectral functions, respectively. These spectral distributions are displayed in Fig. 11. Using the spectral functions to generate the lepton pair emission rate, a comparison with the rates obtained via a summation of mesonic reaction channels is shown in Fig. 12. To summarise, the contributing channels producing lepton pairs in the invariant mass range 1 GeV < M < 3
382
C. Gale and K. L. Haglin
T~7
0.025 — 0.020 r 0.015 E0.010 ~—
Htt,
0.005 = 0.000 0.08
1
h—f-
0.06
1 h
T
ii—i—
-+-H
f
A
0.04 0.02 0.00 0.10
' ) " • » • H " !•
H
»»*(*»** u Lj
r M—i—h-ri-
0.08
^
A
0.06
H
1
h •
H rra.
1 h
MEA
/ s
0.04 0.02 0.00
I 500
1000
1500
,
2000
Vs (MeV) em
Fig. 11. The spectral functions p (s), nihilation and T decay d a t a 5 5 .
v
p (s),
and pA(s),
as compiled from e + e ~ an-
GeV have been found to correspond to the initial states rnr, np, iru, rjp, pp, irai, KK, KK* + c.c. [56]. The detailed channel-by-channel assessment clearly accounts for the net signal yielded by the "global" spectral function analysis. In order to compare with experimental data, the rates must be timeintegrated in a model that is also compatible with other measured observables, hadronic or otherwise. Furthermore, a precise simulation of the detector acceptance and resolution is necessary. An approach that incorporates both aspects is described presently. A class of models that produce time-evolution scenarios is that of hydrodynamic models. Specifically, the assumption is that, at SPS energies, a plasma is produced at proper time To. Assuming isentropic expansion, the temperature and proper formation time can be related to the measured differential multiplicity 58 2TT4
1
dN
45C(3) AT dy
4aT0Jr0
(23)
dN/dy is the measured particle rapidity density and a = 42.257r2/90 for a plasma of massless u, d, s, g partons. Once the transverse area AT is known
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
383
10
10
> 10
10
10
10
10
10
0.0
0.5
1.0
1.5
2.0
2.5
3.0
M (GeV) Fig. 12. Net dilepton production from a gas of mesons at a temperature of T = 150 MeV, as a function of dilepton invariant mass. The solid curve is the sum of the hadronic channels discussed in the text and in the references. The data points are from [57],
along with dN/dy, the above relation links To with To. Enumeration of the model premises is completed by the statement that the plasma is assumed to undergo a boost-invariant longitudinal expansion and an azimuthallysymmetric radial expansion, with a transition to a hot hadronic gas consisting of all hadrons having M < 2.5 GeV, in thermal and chemical equilibrium at temperature Tc. This makes for a rich equation of state. Once all parton matter is converted into hadronic matter, expansion continues until a kinetic freeze-out temperature Tp is reached. Those steps are generic in hydrodynamic calculations. Note that during the evolution, the speed of sound in matter is consistently calculated at every temperature that is input into the equation of state and needed to solve the hydrodynamic equations 59 . Additional details about setting up the initial conditions for the hydrodynamic evolution can be found in Ref. [60], The same reference also shows the result of hadronic spectra calculations with the hydrodynamic approach. It is vital to account for the finite acceptance of the detectors and for their resolution when comparing the results of theoretical calculations with
384
C. Gale and K. L. Haglin
measured experimental data. In the case at hand, those effects are indeed important in the NA50 experiment 37 . One approach to this problem in the past has been to model approximately and analytically the acceptance 62,63 . While this can be readily implemented, a legitimate doubt can subsist about the accuracy of the experimental representation, especially in regions where edge effects might be important. In order to circumvent this problem, a numerical subroutine developed to reproduce the NA50 acceptance cuts and finite resolution effects in the measurement of muon pairs in Pb + Pb collisions at the CERN SPS was used 60 . Thus, the invariant mass distribution of lepton pairs is computed in the hydrodynamic model, and then the pairs are run though the numerical detector simulation. The normalisation is determined by a fit to the Drell-Yan data using the MRSA parton distribution functions, as in the NA50 analysis. In order to get a px distribution, the dN/dM2 estimates for Drell-Yan were supplemented with a Gaussian distribution in PT62, this very closely reproduces estimates obtained by the NA50 collaboration. The resulting invariant mass distribution is shown in Fig. 13. The pr distribution is also computed. It is shown in Fig. 14. In
io3
10*
10° 2
3
4 M (GeV)
5
6
Fig. 13. The calculated dimuon invariant mass distribution, after correcting for the detector acceptance and resolution. The data are from the NA50 collaboration 6 4 . The Drell-Yan and thermal contributions are shown separately, as well as those coming from correlated charm decay and from the direct decays of the J/ip and ip'.
both cases, good agreement with the experimental data is clearly achieved.
Electromagnetic
:
Radiation from Relativistic
Nuclear Collisions
385
Central, Pb+PbOSPS
PT
(GeV)
Fig. 14. The dimuon transverse momentum spectrum, after accounting for the detector effects. The data and the different curves are from the same sources as in Fig. 13.
As this point it is appropriate to consider the following question: which initial temperature is demanded by the intermediate invariant mass dilepton data? A critical and quantitative assessment of this issue can be obtained by examining a linear plot of the lepton pair mass spectrum in the region under scrutiny. This is shown in Fig. 15. From this figure it is clear that the best fit is provided by To = 0.2 fm/c, and that the second best (less than two standard deviations away for most of the data points) belongs to To = 0.4 fm/c. In terms of initial temperatures, those correspond to To w 330 and 265 MeV respectively. A conservative and reasonable point of view is that it is probably not fair in such a challenging and complex environment as that of ultrarelativistic heavy ion collisions to ask for an agreement that is better than two standard deviations, considering all of the inherent uncertainties. The quark matter contribution (as modeled by qq annihilation) is ss 23% for TO = 0.2 fm/c, and « 19% for To = 0.4 fm/c, around a lepton pair invariant mass of 1.5 GeV. Focus so far has been placed on high multiplicity data only. However, to extend the hydrodynamic model to non-central events and to properly treat the azimuthal anisotropy is not a simple task. However, one can get an approximate estimate of the centrality dependence by ignoring the broken azimuthal symmetry and by approximating the region of nuclear overlap by a circle of radius R « 1.2(A^ part /2) 1/3 , where JVpart is the number of participants 65 ' 60 . A centrality-dependence is generated thusly and shown in
386
C. Gale and K. L. Haglin
. " ' ' ' 1 ' ' ' ' 1 ' ' ' ' 1 ' ' ''). 1500 —Central • Pb+PbOSPS
XI S5 •a
f-
Thermal+eharm+J/V+DY| •
5 NA50
1350
S
T„=0.2, 0.4, 0.6 f m / e
T
J-
1000
i
750 L
Thermal
'/
icharm+J/^+DY > 500 — ^^~-^'S—^ ,
,
.
1.0
.
1 // 1.5
,
.
^ ^ I ^ K / ^"^7 •> L . 2.0
, s,
.
1 . 2.5
.
~ . . 3.0
M (GeV) Fig. 15. A linear plot of the net dilepton spectrum in the intermediate mass region. The three solid curves correspond to formation time TQ= 0.2, 0.4, 0.6 fm/c, from top to bottom, respectively. The data are from [64]. The thermal contribution and that for hard processes are shown separately.
Fig. 16. It is seen that the agreement with the measured data is quite good 1
• • • 1 • •••I — I . . . . , . . . . , . . . . , . . . . ,
,
.
.
# , * ANch = Extra Charm; NA50 CO 4->
C 3
1000 500
Nth
= Thermal Dileptons Pb(15BA GeV)+Pb,
£> l-l
tO
& XI
100 50
S5
< 10 30
part
Fig. 16. Centrality dependence. The data represents the "extra charm" yield (as characterised by the NA50 collaboration 6 4 ) needed to describe the intermediate mass dimuon data. The solid curves are from the sources discussed in the text.
and that this approach gives a fair description of the centrality dependence
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
387
of the excess dilepton measurement. What has been learnt from this exercise is that many-body channels, important in the low mass region, also still play a vital role in the intermediate mass domain. This unity is satisfying. It also happens that the global dynamical behaviour of the electromagnetic radiation can be empirically modeled 61 . Perhaps more importantly, the case where the data shown in this subsection are interpreted as the signature of a charm excess no longer appears to be very compelling. The findings described here are in agreement with previous calculations of dilepton radiation at this mass scale 40 ' 62 . As mentioned previously, the portion of the signal that emanates from the deconfined sector is around 20%, a figure that is unfortunately too small to support convincing claims of a QGP presence, once all uncertainties are factored in. 2.3.
Photons
2.3.1. General Strategy Real photons differ from previously discussed dileptons in a couple of important ways. First, they are on-shell and thus cannot be accommodated kinematically with two-hadron annihilation processes. From the beginning then, the only one-loop contributions to the current-current correlator (or retarded self-energy) are the hadronic radiative decays ir° —> 77, r\ —> 77, and to —> 7r°7. Although the ui lifetime is ~ 23 fm/c and so that channel could be considered a thermal source (just barely), the others are clearly non-thermal sources with lifetimes much longer than the fireball. Such contributions must be considered in the overall yield, but are outside of the scope of the present discussion focusing on thermal emission. The oneloop contributions therefore play a smaller role in photon production as compared with dilepton production and consequently the discussion begins seriously here at two loops. Second, and something of a technical point, is that real photons have only two polarisation states over which to sum, rather than three as was the case for virtual photon propagation. In the same spirit then as the dilepton case, real photon emission from resonance hadronic matter can be most systematically studied by beginning with the photon self energy at two-loop order and working upward. The occupation numbers for the internal lines pay always a Boltzmann penalty and it is therefore natural to begin with the lightest species and with the minimum number of hadrons. As a aside remark, since the evaluation of two-loop topologies at finite temperature is technically challenging (although not
388
C. Gale and K. L. Haglin
impossible), many photon rate calculations rely on a kinetic theory approach. Since the energy regime is both relativistic and nonperturbative in terms of QCD degrees of freedom, it is commonplace to use effective theories for the composite hadron dynamics. Typically one starts with an effective Lagrangian with a large enough flavour symmetry to account for the lightest and relevant species. As a general rule, the pions are most important, followed by rho and so on, simply owing to increasing mass. Quantum numbers also play a role in terms of spin states and isospin states governing densities, and so one must be systematic. With interactions under some control relative to chiral symmetries, gauge invariance, conservation requirements of various sorts, one uses cutting rules on the two-loop self-energy diagrams in order to generate a list of reactions of the type ha + hb —> h\ 4- 7 and ha -> hi + h2 + 7. 2.3.2. Establishing the Rates The above mentioned strategy was first taken by Kapusta, Lichard and Seibert 66 where 7r-p and light meson dynamics were investigated. The dynamics were modeled with £=\D,$\2
- mlm2
- \P„vPr
+ \m2pP^
- \F^VF^.
(24)
Coupling of the rho and the photon to pions was accomplished with the covariant derivative D^ = d^ — ieAM — igpp,j,. The charged and neutral pions are embodied in the complex pseudoscalar field <£, the vector rho and photon field strength tensors are respectively p^v = d^p^ — dvp^ and F^v = dfj.Au — dvA^. Calibration is done by fitting the p —> 7r+7r_ decay rate with the choice gp — 2.9. The specific channels studied in Ref. [66] were dubbed annihilation + 7r 7r~ —> p7 and "Compton scattering" np —> 7T7, and finally, neutral rho decay p —> 7r + 7r _ 7 (essentially the finite temperature analog of the vacuum process studied by Singer 67 ). Since the T] meson mass is intermediate between pion and rho, its Boltzmann penalty is less than rho's. Its effects were also considered by including the channels TT+TT~ —» 777, n^ —> n^j. Owing mostly to coupling strengths (or weaknesses), these channels were found to be less important as compared to the purely n and p channels by more than an order of magnitude. Finally, in this initial study of photon production, the channel n+ir~ —> 77 was included, though it was seen to contribute very little.
Electromagnetic
Radiation from Relativistic Nuclear Collisions
389
The matrix elements for all the processes enumerated above are included in Ref. [66] and will therefore not be repeated here. The energy dependent invariant rate for producing photons is then obtained by folding in BoseEinstein (enhanced, if final) hadron distribution functions and Lorentz invariant phase space. For instance, for the channels pa +Pb —> Pi + p 7 , one has
JLJ (2ir)32Eb J0 (2n)32E1 v (2TT) 3 2' where M is the appropriate degeneracy factor counting the states. The resulting rates have been established numerically. However, analytical parametrisations valid for 100 MeV < T < 200 MeV and 0.2 GeV < Ey < 3 GeV have been proposed 68 a .
E7——(nn d3p7
-> py) = 0.0717T 1 8 6 6 exp(-0.7315/T+ 1 . 4 5 / V ^ - £ 7 / T ) ,
E
~* ni)
i-^{np
E7-^-(p
a p-f
= T2A
ex
P(-V(2T^)3/4 -
E7/T),
->7T7T7) = 0 . 1 1 0 5 T 4 - 2 8 3 £ - 3 - 0 7 6 + 0 - 0 7 7 7 / T exp(-1.18£ 7 /T).(26)
In these expressions, T is the temperature, E1 is the photon energy; both must be reported in GeV. The numerical constants have appropriate units in each case and the numerical values out in front have units f m - 4 G e V - 2 . The numerical results are shown in Fig. 17. The rate for the w —» 7r°7 is also included in Fig. 17. It can be written as b
x [1 + fBE (Eu - £ 7 )] where Em-m = mu(E% + EQ)/2E1EQ, rest frame of the w meson.
(27)
and EQ is the photon energy in the
a T h e process p —> irwy was slightly miscalculated in Ref. [66] owing to an omission of a Lorentz-boost factor. The parametrisation of the process published in Ref. [68] is therefore not optimal. A slightly different parametrisation is proposed here which correctly accounts for covariance effects. b Note that Eq. (54) of Ref. [66] has an incorrect Lorentz-boost factor. The formula has been corrected here and reported in Eq. (27).
390
C. Gale and K. L. Haglin
0
1
2
E r (GeV)
3
0
1
2
3
E7 (GeV)
Fig. 17. Photon emission rates from Ref. [68] plus an estimate including the a\ coherently. The u> —• 7T7 rate results from Eq. (27). Temperatures are fixed at T = 150 MeV and T = 200 MeV.
2.3.3.
Refinements
Soon after these initial rate calculations were done, Xiong, Shuryak and Brown 69 pointed out that the a\ meson would have an important effect on the 7rp —> 7T7 channel. It is a resonance in the np sector with pole mass roughly matching the average -y/s in the fireball and with a rather large width ~ 400 MeV. An interaction for the ainp vertex was proposed, stemming not from some symmetry argument but rather, a vertex having minimal momentum dependence and still respecting gauge invariance. The idea was to set the scale as simply as possible for the contribution from this process. Prom the strong-interaction vertex, vector dominance was employed to subsequently describe radiative decay. The a\ was indeed found to be important when studied this way—even dominating the other exchanges. However, by itself the s-channel a\ diagram does not carry complete information on the overall strength of the Compton process. Instead, a coherent sum of pion exchange and a\ exchange was necessitated. Song 70 later carried out a study of photon production using a Chiral Lagrangian
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
391
with vector and axial-vector fields introduced as dynamical gauge bosons of a hidden local symmetry. The role of the a-\_ once again seemed to be important. However, Song found that fitting the relevant masses and coupling strengths in the model allowed two solutions, i.e. two parameter sets, and was therefore not able to uniquely identify an emission rate from this channel. It was later found, while studying dilepton emission with the same model, that parameter set II was the more reliable one in terms of its ability to match the observable hadronic quantities 43 . In particular, the DjS ratio in the scattering amplitude for a\ —> 7rp was more closely respected with parameter set II. One thus observes a diminished presence of the a\ meson (as opposed to estimates cited above) in the final rate. A modest enhancement of 20%, attributable to the pseudo-vector, is illustrated in Fig. 17. An update on this is forthcoming71. A study with the hidden local symmetry approach also points to a reduced role of the ai meson 72 . Strange particles have been found to be only marginally important in the literature up to now. Specifically, the radiative decay K\ —> K'y was considered and was shown to be strong relative to the non-strange contributions only in a limited kinematic domain 73 . A final cautionary remark is in order here. When the rate spectra are studied at photon energies above 1 GeV, one must keep in mind that hadronic form factors have not been implemented in most of the rate calculations up to now (an exception is the Kapusta, Lichard and Seibert calculation 66 which estimated the form-factor effect on the TTTT —> pj channel), and could result in a suppression of a factor of 2 or more at higher photon energies. This is where the exchanged meson goes further off shell and brings forward possibly large form-factor effects. Advances in this direction will be important.
2.3.4. Medium Effects In terms of higher-order effects, corrections to these rates come in at least two forms. First, there are off-shell effects which can be conceptualised by dressing the propagators and vertices for the internal hadron species in the general photon self-energy structure. These are the so-called form factors mentioned earlier in cautionary remark. Second, and beyond this, there are bona fide medium effects (finite temperature and density effects, e.g. width smearing and pole mass adjustments) that could be quite important. The typical pursuit in studies of medium modifications is to investigate the effects of dramatic collision broadened vector meson spectral distributions 74
392
C. Gale and K. L. Haglin
and/or the dropping of the rho mass according to the so-called Brown-Rho scaling 75 or some other ansatz. Several authors have studied various pieces of the overall medium dependences 76 ' 77,78 ' 72 . The trends are the following. While the in-medium vector meson widths are expected to be rather large, the effect on photon production is not too significant. This makes sense since the vector spectral distributions contribute to photon production only as an integral over the specific distribution—and smearing the distribution does not affect the normalisation. Mass shifts, on the other hand, have been shown to affect the rates by anywhere from a factor of 3 up to an order of magnitude 76 ' 77 . The results are too model dependent to make specific concluding statements at present.
2.3.5. Alternative Approach: Chiral Reduction Formulae Instead of computing photon production rates using a channel-by-channel assessment, Steele, Yamagishi, and Zahed used chiral reduction formulae together with a virial expansion and they came forward with photon and dilepton emission rate estimates. The general idea is that the invariant production rate is proportional to the trace over a complete set of hadronic states of the hadronic (Boltzmann weighted) Hamiltonian convoluted with a current-current correlator. The hadronic part of the correlator is written as a virial type expansion truncated in a particular way. The expansion coefficients are constrained by various general arguments, e.g. broken chiral symmetry, unitarity, and gauge invariance and also, when available, constrained by observed spectra: electroproduction, T decay, radiative pion decay, and so on. The thermal photon emission estimates in this approach tend to be larger than those using an effective Lagrangian approach by a factor of 2-4 7 9 ' 8 0 . At present, this might be the honest theoretical error bar in the rate estimates even after a decade of model calculations. Progress continues especially with effective theories in the hadronic matter converging with results from models firmly rooted in QCD as the fundamental degrees of freedom as discussed in the next section. This is the so-called duality of hadronic matter and quark matter at the phase boundary that one expects. At this stage in the discussion it is somewhat premature to integrate the photon production rates over a space-time evolution, which would then facilitate a comparison with experiment, because radiation from partonic matter has not yet been discussed. So, before considering photon yields from nuclear collisions and making contact with data, the partonic contri-
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
393
butions to electromagnetic radiation will be presented, and then yields will be discussed. 3. Radiation from Partons It is of great theoretical importance to establish the production rates of electromagnetic radiation for many-body systems beyond the deconfinement phase boundary of nuclear matter. A model is employed whereby the matter is assumed to be fully in the partonic phase. Whilst experimental verification of an unequivocal identification of thermalised quark-gluon plasma is still forthcoming, it is the appropriate picture with which to work as a baseline. The general formalism established for photon production rates from hadronic matter in Sect. 2 is generic to all quantum field theories and is thus equally valid for partonic degrees of freedom. And yet, the massless nature of the up and down quarks requires special attention. Calculational tools known as hard-thermal-loop (HTL) methods have been applied to handle infrared singularities. Independent of the experimental advancements then, it would already be important to establish quark matter radiative emissivities. Since heavy-ion experiments at the CERN SPS and at RHIC have most likely probed into small areas of the deconfined region in the nuclear matter phase diagram, there is further motivation, and indeed some urgency, for theoretical investigations to converge and to report emission rates. Therefore, the status of theory for photon production from finite temperature quark matter is discussed below and a separate section is devoted to dilepton production. 3.1.
Photons
The imaginary parts of one-loop contributions to the photon self-energy, obtained with appropriate cuts, are identically zero due to vanishing phase space. Certain two-loop diagrams give nontrivial contributions. Cutting rules provide a bridge between kinetic theory and field theory where in fact, a mapping has been established 81 . The result of cutting two-loop diagrams gives QCD processes of the types qq —> 37 and qg —• q-y or qg —> 97. These processes, as well as bremsstrahlung processes, were studied using perturbative matrix elements two decades ago 82,83 ' 84 . The results were unfortunately infrared unstable (i.e. the rates diverged as the quark mass tended to zero). Significant improvement came when the "annihilation" and "Compton" processes were analysed by Kapusta et al.6e and Baier et al.85 using resummation techniques of Braaten and Pisarski 86 ' 87 . The basic
394
C. Gale and K. L. Haglin
idea behind the resummation technique, or the so-called hard-thermal-loop approximation, is that weak coupling at high temperatures allows a separation of scales, and a separation of the rate into soft (quark momentum ~ gT or smaller) and hard (quark momentum T or larger) contributions. The soft contribution can be computed with an appropriately dressed quark propagator in the one-loop photon self-energy, while the hard contribution can be computed using perturbative methods and kinetic theory. In each result, the separation scale appears as a sort of regulator. When the soft plus hard contributions are collected together and added, the result is independent of the separation scale, and of course also independent of quark mass since it was set to zero from the beginning. The exact result can be established only numerically. However, using an approximation which is valid for E^/T 3> 1, a simple pocket formula has been proposed. At the time, this result was thought to be complete to order aas. Specialising to two quark flavours the result is
A value of a3 = 0.4 (g2 = 5) is used and a " 1 " is added to the argument of the logarithm when plotting as suggested by Kapusta et al. to more closely match the exact numerical result for photon energies of the order of the temperature. This also ensures the rate is always positive. These results were subsequently generalised to finite quark chemical potential and also applied to chemical non-equilibrium systems. For a discussion of these effects see [88] and references therein. Having photon emission rates from QCD free from infrared instability ailments represented significant advancement and was at the time, thought to be the complete lowest order result. After all, the two loop contributions to the self-energy (that is, the dressed two-loop contributions, which actually contain arbitrarily many loops) seem naively to contribute to photon production at 0(aa2). They specifically correspond to bremsstrahlung processes and annihilation with scattering. The extra vertices would introduce an extra power of g2 as compared with the one-loop result. However, Aurenche et al. showed that the two-loop HTL contribution is curiously not of higher order, but instead contributes to order aas too 8 9 . This owes essentially to a collinear singularity when the exchanged gluon is soft. The resummed gluon propagator introduces a g2 in its denominator which cancels the "extra" g2 from the additional vertices. The overall contribution to the HTL for this category of two-loop diagrams is the same (lowest)
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
395
order in aas. In terms of the kinetic theory analog, these correspond to such processes as qq —> gq'f, gq —» 397, and qqq —» 57, or qqg —» 37 (and still others with antiquarks). Two- and even three-loop contributions were shown to contribute to lowest order, and the rates continued to rise! It is fair to say that after these features were pointed out by Aurenche et al, the situation appeared to signal a breakdown in perturbation theory for finite temperature QCD. However, it has been shown recently by Arnold, Moore and Yaffe that as long as E7 ~> gT, there is sufficient cancellation due to many-body effects so that the lowest-order rate is identifiable and fully under control 90 . This remarkable result represents a significant advancement in this field. There were unfinished details within the topics of bremsstrahlung 89 , magnetic mass 91 and coherence effects92 that Arnold et al. resolved by analysing multiple-loop ladder diagrams which introduce multiple scattering interference effects of Landau-PomeranchukMigdal (LPM) 93,94 . A digression will not be taken here to reproduce the lengthy and specialised argument, but the result is the following. The suppression is sufficient to regulate the rates because 1-loop, 2-loop and multiple-loop diagrams can be consistently resummed to give a finite rate! The efforts of many people over a decade of work have produced a complete photon production calculation from QCD to lowest order 0{aas). The simple expression below parametrises the exact numerical solution for two quark flavours. E
_dB_ = 5 aas_T2 3
'd p7
2
9 3TT
In (
1 e^/T
+ l
J + - In ( —
) + C2^2
+ Cbrem + Cannih
,(29)
where
r
C 2 -2 - 0.041(T/£ 7 ) - 0.3615 + 1 . 0 1 e - 1 3 5 ^ / T 0.633 ln(12.28 + ( T / £ 7 ) ) 0.154(JS7/T) r (£ 7 /T)3/2
V 1 + (£ 7 /16.27T)
(30)
Results are shown in Fig. 18, where a value for the strong coupling as = 0.4 (g2 = 5) is used as before, and superimposed onto the total hadron rate discussed previously. The striking feature is that after the dust has settled on the QCD calculations, with HTL to 1-, 2-, and even multiple-loop order, with LPM effects carefully included, the QCD rate at fixed temperature is once again the same as the hot hadronic gas rate. The QGP and the hadron gas seem to "shine just as brightly".
C. Gale and K. L. Haglin
10
10
-3
-4
£io-
5
=s 10
i inuiij i iiIIIII
396
1
'
1
"-
'
hodrons
^ %
^\
J
v\
"I
10
-10
li«"i' i 1
10-9
i iiiiml
o " 1 0 ~ 7 U=~
£ io- 8
J
— quarks—II q
\V
br-
g
T = 2 0 0 (MeV)
-
,
I
E7
i
J
I
i
1
(GeV)
Fig. 18. Photon emission rate for Q G P from Kapusta et al. in Ref. [66] (labelled Quarks-II) and Arnold et al. in Ref. [90] (labelled Quarks-I). Quarks-II includes Compton and annihilation, while Quarks-I includes in addition, bremsstrahlung and certain 3—»2 processes. Quarks-I is the "complete lowest order calculation". Temperature is fixed at T = 200 MeV. Total hadron contribution is also displayed for comparison purposes.
3.1.1. Photon Measurements Photon experiments using heavy-ion beams are notoriously difficult and signals of any kind are already a notable accomplishment. At high energies, there are at present two sets of data with which to compare the theory. The WA80 collaboration at CERN first reported and discussed their yields as absolute measurements, but were later forced to loosen the constraints somewhat and suggest upper limits only. Their direct photon limits came from 200.4 GeV 32 S + Au collisions95. Secondly, the WA98 collaboration measured direct photons in 158.4 GeV 2 0 8 Pb + 2 0 8 Pb collisions also at CERN 96 . The hope from the onset was to challenge the theory using production rates convoluted with a temperature profile evolving according to one of two possible scenarios: 1) the system first comes into equilibrium well above the phase boundary and therefore the quark rates contribute until such time as the system reaches the mixed phase. Overlapping four volumes mean that quarks and hadrons contribute until the latent heat is
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
397
absorbed fully into a hadronic state, and finally, the hadrons emit until freezeout; and scenario 2) where the system reaches a very hot and dense hadronic state and simply radiates photons while cooling and eventually freezes out. The burning question is which scenario is consistent with the measurements? Can either one be ruled out? There were several attempts to describe the WA80 results and do just that. Shuryak and Xiong 97 first used the hadron rates with their version (incoherent treatment) of a\ meson dynamics included, and their conclusion was that the excess photon signal could not be described with a conventional expansion scenario. They consequently suggested a long-lived mixed phase as a possible explanation. Since the data were later reported as upper limits only, the conclusion no longer rested on strong experimental support. Srivastava and Sinha applied the quark rates at the 1-loop HTL level and the hadron rates comparing scenarios (I) with and (II) without a phase transition to QGP 9 8 . They argued that the data (which later became upper limits) are well described by a scenario where QGP is formed initially. Bjorken hydrodynamics was employed with T, = 203 MeV, Tc = 160 MeV, and Tf = 100 MeV for scenario (I) and, % = 408 MeV for scenario (II). Other models came forward attempting to describe the experimental results. For example, Dumitru et al. used a three-fluid hydrodynamics without and with a phase transition". They came to similar conclusions, that without a phase transition to quark matter, the results were inconsistent with experiment. Improvements in rate calculations from quark matter brought advancement also in yield estimates. Two-loop HTL rates were coupled with hydrodynamics, and then later corrected due to numerical errors along the way 100 . The most recent and corrected comparison of the WA80 upper limits to hydrodynamic model estimates are displayed in Fig.19. The conclusion is that both scenarios, without and with a phase transition, seem to be consistent with the upper limits. A more complete hadronic equation of state (EOS) and up-to-date photon rates from quark matter lead to these new conclusions. With the WA80 results in hand and even anticipating the forthcoming WA98 data at that time, Cleymans, Redlich and Srivastava 101 used a hydrodynamical model, which arguably provided better description of the evolution as compared to previous model calculations and, in particular, could more completely describe the transverse flow likely to be generated at the SPS. The initial QCD rates of Kapusta et al. were used and the hadron rates, including the effects of the a\ meson, were implemented with
398
C. Gale and K. L. Haglin
108 P l i u e T r a n s i t i o n , ^ - l e i «teV T 0 -203 U«V,T0-1 f m / o S(200 AG«V)+Au; WA80 7.4%, moat cantral
No Phase Transition pt
(GeV/c)
p , (GeV/c)
Fig. 19. Upper limits at the 90% confidence level from WA80 on the invariant excess photon yield per event for the 7.4%
all hadrons up to 2.5 GeV mass contributing to the equation of state. They concluded that while the final yields were not significantly different in a QCD plus hadron matter scenario as compared with a fully hadronic picture, they argued that the physics seemed to favour the former since in the hadronic picture particle densities were beyond anything reasonable for hadronic language to be justified. Before moving to the WA98 data, one might make the remark that since the WA80 results are upper limits rather than measurements, and due to the uncertainties in the theoretical production rates and, mostly, with the uncertainties in the models for the evolutions of the nuclear systems, no definite conclusions can be reached. The eagerly anticipated direct photon measurement from 2 0 8 Pb + 2 0 8 Pb collisions at 158.4 GeV were published in 2000 by the WA98 collaboration 96 . The collaboration presented their data as compared with several protoninduced reactions at similar energies and scaled up to central 2 0 8 Pb + 2 0 8 Pb collisions. For pr > 1.5 GeV, where the signal is strongest, there is a clear excess beyond that which is expected from proton-induced reactions. In other words, the results are quite suggestive of thermal photon emission, or perhaps pre-equilibrium emission. One contribution that is non-negotiable in those data is that due to perturbative QCD. It owes its existence to collisions during the first instants of the reaction, and should appear in pp, pA, and AA measurements. The WA98 data is shown in Fig. 20, along with a pQCD estimate 102 . Even though the presence of pQCD effects at the energies under discussion here can't be argued against 103 , the appli-
Electromagnetic
Radiation from Relativistic
,208
, 20a
•
WABS This Result
Nuclear Collisions
399
D D 158AGeV™ , Pb + '°"Pb Central Collisions
pARaaultaata"*z 19.4 GeV
O
scaled to s' / 2 =17.3QeV
"3. 10
z "0
D
E628
V V V V
? 10
pOCO ( a ' ^ E l T . a G a V . y ^ J . W o n g e l a l . CTEQ4, <*?>=0.9 (GeV/e)2 CTEQ4, No Intrinsic kT
0.5
1
1.5
2
2.5
1
3.5
4
4.5
Transverse Momentum (GeV/c)
Fig. 20. The WA98 real photon measurements as a function of photon transverse momentum. The pQCD estimates are from [102].
cation to nucleus-nucleus data contains some ambiguities that need to be pointed out in order to make progress. Specifically, it is clear that some amount of parton intrinsic transverse momentum (neglected in strict NLO calculations) should manifest itself. Simple uncertainty principle arguments support this 29 , and soft gluon emission should increase the value further 104 . However, attempts to extract meaningful values from experiments have remained inconclusive; for example a recent survey found that fixed target data at ISR energies (y/s < 23 GeV) were inconsistent 105 . Furthermore, in nucleus-nucleus collisions, a part of the parton transverse momentum can be ascribed to multiple soft scattering of the nucleons prior to the hard scattering 106 , and this has to be modeled dynamically and independently. It is important to note that, at RHIC, several of those uncertainties will be lifted, as measurements of pp, pA, and AA reactions will be performed at the same energy with identical detector configurations. Bearing all those caveats in mind, a recent study 107 of E704 and WA98 data found that (A;2) ~ 1.3 GeV 2 could by extracted from pp reactions, leaving up to 1 GeV 2 for nuclear effects. This analysis is shown in Fig. 21. It is clear from this work that photon transverse momenta below 2.5 GeV are under-predicted by this pQCD estimate. Also, around this momentum, the exact value of
400
C. Gale and K. L. Haglin
Fig. 21.
The WA98 data with pQCD estimates of transverse partem momentum. From
[107].
the intrinsic transverse momentum ceases to be important. A softer component of the photon spectrum is called for, and this will be discussed shortly. Note that this value for a separation of scale between the "hard" and "soft" photon sources also appears if one fits the high momentum pQCD spectrum to the data, with a K factor 108 ' 61 . It is argued in these cited works that the soft component possesses thermal characteristics. Srivastava and Sinha 109 studied mechanisms for excess photon production using an hydrodynamic expansion applied to the 2 0 8 P b + 2 0 8 Pb system. The photon emission rate from quarks was input using the result from twoloop HTL calculations from Aurenche et al.S9. However, it is probably fair to say that those rates have been superseded by the calculations in Ref. [90], which incorporates higher loop topologies and thus LPM effects. Production rates from the hadronic phase were taken from the parametrisation of Kapusta et al. plus an incoherent a i -exchange contribution to the process np —> n-f. The results obtained there are shown in Fig. 22. The high initial temperature in this work is needed to generate a sufficient high transverse momentum component of the photon spectrum. In this respect, the WA98 data has been used to extract phenomenologically an initial radial velocity profile110. The result of that study is shown in Fig. 23. Both of those
Electromagnetic
Radiation from Relatimstic
401
dN/dy=750
\\ _
£-• 10-1
Nuclear Collisions
T 0 = 1 / 3 T 0 = 0 . 2 0 fm/o
>
-
T0=335 MeV, Tc=180 MeV: IWA98, Pb+Pb®SPS Single Photons
V O 10-2 XI M N
X(
\
S5 0
1
2
3
4
5
kT (GeV) Fig. 22. Single photon production in P b + P b collisions at relativistic energies. For specific model details, consult Ref. [109].
theoretical efforts concluded that the excess seemed to be consistent with a thermal source of photons at roughly ~ 200 MeV temperature, while detailed and quantitative conclusions on a partonic scenario versus hadronic with strong flow were not definitively reached.
158AGeV'*Pb + " , P b Central Collisions
158AGeV'"pb + " , Fb Central Collisions —
> WA9S data v,=0 VJ.0.1
PT«*V)
P,(GeV)
Fig. 23. Total photon yields from quark matter plus hadronic matter (left panel) and hadronic matter with medium-modified vector meson properties (right panel). The figure is reproduced here from Ref. [110].
Ruuskanen and collaborators also used hydrodynamics to compare theory with experiment 111,112,113 . These workers have challenged hydrodynam-
402
C. Gale and K. L. Haglin
ics to find consistency with not only photon spectra, but also hadron and dilepton spectra—all within the same model and simultaneously. They also insist on reproducing the longitudinal hadron characteristics 113 . Several equations of state and therefore several expansion scenarios seem to describe the photon spectra and hadron spectra equally well. The degeneracy between the different equations of state and initial conditions is not lifted empirically, even though the data do require a high density and temperature initial phase. 1=1 I | I I I I I I I I I | I I I I I I I I I | I I I I I I I I I | I I I b|
UrQMD
f 10 > w
10
WA98 L T I 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 f ffk+fi
Hydro
' I ' ' i i i i i i i I i i i i i M i i I i i i i i i i'
1
2
3 P T (GeV)
4
Fig. 24. Comparison of the WA98 photon spectrum to the predictions of the UrQMD model and the hydrodynamic model at several freezeout temperatures from Ref. [25].
A comparison of space-time models in reference to the WA98 data was recently carried out by Huovinen, Belkacem, Ellis and Kapusta 25 wherein hydrodynamics and a course-grained UrQMD were used to produce photon spectra (dilepton spectra were also calculated within theory and compared to experiment). Notably, the "complete" lowest order photon production rate [C(ao;s)] from the quark phase was used in this work. The basic con-
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
403
elusions were that UrQMD and hydrodynamics seem to give roughly the same qualitative features of expansion and cooling (although quantitatively, UrQMD cools more slowly due presumably to viscous and heat conduction effects); they give therefore, very similar results for photon production. The precise choice of freezeout temperature seemed to be irrelevant, indicating that the high temperature part of the evolution dominates photon production. Results are shown here in Fig. 24. The agreement between theory and experiment was described by these authors as "excellent", while they reminded the reader that the rates have uncertainties and the initial conditions which were fed into the models are responsible for further uncertainties propagating to the final spectra. A partial summary of the photon analyses is justified. It is an accurate statement that definite conclusions are elusive. Many physical ingredients have been invoked in the studies of heavy ion photon data, as seen above and in the quoted reviews, but uncertainties in many of those ingredients (if not all) preclude a clear interpretation of a signal that relies on a combination of their effects. But one example is the absence of the chemical potentials in hydrodynamics-based approaches. Another is the uncertainty in the basic photon rates. However, those uncertainties have narrowed down considerably in recent years, and this is true for rates in both the partonic and confined sectors. Also, as mentioned previously, the fact of being able to access data at the same energy in pp, pA, and AA events will make RHIC a fertile testing ground for theoretical models, and should allow the community to make more progress in differentiating between them.
3.2.
Dileptons
The yield of low mass dileptons (M < m^) from thermal quark-antiquark annihilation is not expected to be a great competitor of the two-pion annihilation simply owing to longevity effects in the two phases. The quark phase occupies a smaller four volume. Nevertheless, it is useful to assess the production rates as a benchmark and then to ask about higher-order corrections, especially in the medium. For high enough system temperatures, and for large enough invariant masses, qq —> 7* —» e+e~ is considered a more significant source and in terms of theory, can be reliably computed in a HTL approximation. The lowest order contribution [0(a°)] in field theory language corresponds to a one-loop graph with bare quarks occupying the internal lines. The imaginary part of the self-energy describes precisely the annihilation process mentioned above. The production rate is roughly
404
C. Gale and K. L. Haglin
the square of the density of quarks times the cross section times the relative velocity. These are now textbook formulae114 so one simply quotes the results
dR
dM2
="§5i^*<"/n
(31)
where the annihilation cross section is 47r a2 a{M) (32) YM2' and where TV" is an overall degeneracy factor (24 when using two quark flavours) and finally, K\ is the modified Bessel function of order 1. Here the quark and lepton masses have been set to zero. To compare the resulting rate with major hadronic contributors, Fig. 25 is presented. Already here, one sees that the Born term is not negligible and the natural next question is the role of higher order contributions. 10"
->
1
'
10"
r +
qq-»e e
qq-*e+e rtrs-*e*e~ KK->e+e"
1.5 M (GeV)
M (GeV)
Fig, 25. Thermal production rate for dileptons via lowest order quark-antiquark annihilation as compared to leading hadronic channels rnr —* e + e - and KK -* e + e - . The temperatures are set to T = 150 MeV (left) and 200 MeV (right).
Perturbative corrections to these annihilation rates were considered by Braaten, Pisarski and Yuan 118 , who found that for very soft dileptons at
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
405
rest in the medium (energies 4C 1 GeV) the corrections were orders of magnitude larger than the Born term. Also, unique structures emerged in these corrections owing to Van Hove singularities arising from significant softening of the quark dispersion relation in medium. There appears a minimum in the medium-modified quark dispersion relation (a plasmino) typically at dilepton energies less than that set by the quark mass. While these effects are quite intriguing, finite imaginary parts in the quark propagators and finite three-momentum effects for the dilepton 119 could dampen the peaks into undetectable artifacts. Also, the softer bremsstrahlung contributions 120 might overshine these total annihilation channels. For a review of these and other issues for the dilepton channels, see Ref. [24] by Rapp and Wambach. Quark-antiquark annihilation is of course not the only relevant parton process for dilepton production. For instance, the 2 —> 2 real photon production processes considered previously contribute also to lepton pair emission. In addition, there are annihilation processes where one of the incoming partons has already scattered and suffered an off-shell interaction. The resulting 3 —> 2 process comes from off-shell annihilation (also called annihilation with scattering). Such mechanisms have been shown to dominate at high enough photon energy. Since this is essentially a many-body initial state, formation time considerations and coherence effects for the virtual photon suggest once again that multiple scattering plays an important role. Aurenche, Gelis, and Zaraket 121 , and together with Moore 122 , have applied the HTL technique for lepton pairs with E/T 3> 1 (either low mass but high momentum, or high mass) and have shown that the two-loop contributions which include bremsstrahlung of a quark and annihilation with scattering, are free from infrared and collinear singularity effects. When added to to the Born term, the rescattering corrections plus the 2 —> 2 processes 123,124 result in a rate that is somewhat increased as compared with just the Born terms. Furthermore, threshold effects at M2 = 4m 2 (thermal quark mass) are smoothed out, all of which is illustrated in Fig. 26. The dilepton results from QGP discussed above assume equilibrium and and use asymptotic values for such quantities as thermal quark masses and screening masses. Strictly speaking, these are only valid at asymptotic values of temperature: the assumptions needed for the theoretical machinery to remain consistent might actually break down at terrestrial accelerators energies. Scenarios more realistic for RHIC and LHC could be studied if alternative schemes were used to compute masses in nonperturbative circumstances. First steps in the direction of lattice evaluations of thermal dileptons using maximum entropy methods 125 have recently been taken 126 .
406
C. Gale and K. L. Haglin
2M M = 2.3GeV 0.0001 r
1
1
le-08 '
' 0.5
' 1
1
1
1—|—
0" X
' ' 1.5 2 Photon mass (GeV)
]L
—i 2.5
^ 3
Born term rescattering corrections 2->2 processes Total Fig. 26. The dilepton production rate per unit four momentum at fixed energy as a function of photon mass (dilepton mass). Values are fixed at T = 1 GeV, go = 5 GeV, as = 0.3 and two quark flavours were considered. The figure is reproduced here from Ref. [122],
4. 4.1.
Predictions Photons
RHIC has been running, looking primarily at hadronic observables probing the later stages of ultrarelativistic nuclear reactions. Electromagnetic spectra will soon be available which, of course, probes deeper into the fireball and indeed cleanly into early stages of the reactions as well. Predictions for measurements of electromagnetic signals at RHIC are therefore very important. In addition, the Large Hadron Collider (LHC) is only five years away! While this number probably needs an appropriate error bar, it will soon become crucial to have formulated a set of model estimates for LHC experiments too. A section is devoted here to discussing these sorts of predictions. As one moves away from SPS systems and energies and goes to RHIC, and to LHC energies, there are increasing uncertainties in estimates for the initial energy densities. The initial state is very far from under control. But,
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
407
as in all cases, when theory is extrapolated to new territory, the simplest estimates are first used to set the scales and subsequent to this, refinements and various improvements are made. It is in this spirit that photon production (yields) were recently estimated at RHIC and LHC by several authors 1 1 0 , 1 2 7 ' 1 2 8 ' 1 2 9 . Simple 1+1 dimensional models show dominance of the QGP over the hadron gas for photon p? ^ 3 GeV (RHIC) and roughly 2 GeV (LHC) 129 . Transverse expansion, which builds up particularly later in the hadron phase, makes distinction less clear, but the QGP might still outshine the hadron gas. The results for photon production from Ref. [129] are displayed in Fig. 27. gh = 3
p T [GeV]
gh = 8
-
p T [GeV]
Fig. 27. Photon spectra (yields) from SPS, to RHIC, and LHC from P b + P b collisions. Two cases are shown, an ideal pion gas with g^ = 3 (left) and 8 (right) panels.
However, agreement is far from complete on this issue. For instance, Hammon et al.127 predict that QGP will not be visible at RHIC owing to a very strong contribution from prompt photons (a pre-equilibrium source
408
C. Gale and K. L. Haglin
which has not been discussed here, and one which is probably not entirely under control), while at LHC the situation is different where QGP will be visible for limited photon kinematics. They used To = 533 MeV (300 MeV) and r 0 = 0.12 fm/c (0.5 fm/c) for QGP (hadron gas) at RHIC, and they used T0 = 880 MeV (650 MeV) and r 0 = 0.1 fm/c (0.25 fm/c) for QGP (hadron gas) at LHC. One could however hope that the prompt photons due to pQCD could be measured separately (in pp collisions at the same energy, for example), and subtracted out. Alam et al.110 find, using less extreme initial condition parameters (lower initial temperatures), that thermal photons will be visible for px < 2 GeV. However, they also find that thermal photons from hot hadronic gas populate the high PT region even fairly strongly. Again, this is due to a strong flow built up later in the hadron phase. In the face of such lack of agreement, which owes essentially to large uncertainties in the initial conditions, in the nature of the expansion, and even in the quark and hadron rates themselves, one suggests that it is premature to make any definite statement at this point. In other words, the theoretical error bar is too large at present to formulate any physics conclusions from photons. And yet on the optimistic side, theory will progress when the newest QCD rates and hadronic rates are implemented into a dynamical model which attempts to describe the buildup of collective flow in some detail, on a species-by-species basis (viz. heavier species seem to flow differently from lighter ones). In almost all cases discussed in this work, the dynamical simulations used to model the dynamics of nuclear collisions assume some form of equilibrium. Many approaches assume both chemical and thermal equilibrium, while some only need the latter ingredient. There exists, however, a whole class of models that attempt an ab initio rendering of the heavy ion reactions. Those are currently the only window one has to the very early stages of the collisions, and thus they potentially offer precious insight on the importance of pre-equilibrium generation of electromagnetic radiation. At ultrarelativistic energies, the degrees of freedom that appropriately describe this phase are partonic. A recent prediction of the photon yields has been made 115 , using a version of the parton cascade model (PCM) 1 1 6 ' 1 1 7 . Along with gauging the importance of the above-mentioned pre-equilibrium effects, this calculation involves the application of perturbative QCD (pQCD) in a domain not necessarily restricted to large momentum transfers. The photons there are produced from Compton, annihilation, and bremsstrahlung processes at the parton level. All lowest-order
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
409
QCD scatterings between massless quarks and gluons are included in this model. The obtained photon spectrum for the collision of gold nuclei at RHIC is shown in Fig. 28. There, the contribution from interactions involv-
^ BMS: prim-prim & prim-2nd • BMS: full rescatterlng •^
10-
Au+Au; ECM=200 AGeV
(3 5 10*
Mo* PL
,!
.
10"
photons
! f i»
IYCMI < 0.5
Pt(GeV)
Fig. 28. Transverse momentum spectrum of photons from central collisions of gold nuclei at RHIC, calculated with the VNI/BMS parton cascade model 1 1 7 (see the text for an explanation of the symbols).
ing at least one primary parton (triangles) is compared with that obtained with full binary cascading (diamonds). Most of the photons between 2 and 4 GeV have their origin in the multiple semi-hard scattering of partons. This finding would support the claim that high energy quarks going through a quark gluon plasma would yield electromagnetic signatures (see later sections). 4.2.
Dileptons
The plasma signature in the lepton pair channels is expected to manifest itself mainly in the so-called intermediate mass sector 31 (see Section (2.2)). Owing to the large multiplicities germane to the collider conditions, a large background will render the extraction of any direct electromagnetic signal from the low mass region prohibitively difficult. However, there may be still hope to observe some distortions of the vector meson spectral densities. Using a dynamical simulation that accounts for a possible under-saturation of the parton chemical abundances, and estimates of the vector self-energies in a finite temperature meson gas, the yield on low invariant mass lepton
410
C. Gale and K. L. Haglin
pairs was calculated in Ref. [138] and is shown in Fig. 29. It can be seen
T — I — I — I — [ — I — I — I — I — | — I — I — I
.10-
-
—i—I—i—i—i—
Central Au+Au
. s1/2=200AGeV
IVj<0.35
,
> O
il°": free HG + QGP + cocktail \ in-med. HG + QGP + cocktail 0.5
0.6
0.7
0.8
0.9
1.0
1.1
M ee [GeV] Fig. 29. Net dilepton yield from an initial plasma phase evolving into a final hadronic phase. The full and dash-dotted lines respectively represent cases with and without inmedium modifications of the vector meson spectral densities (p, w, and
that the in-medium effects translate into a suppression of the p—u complex, and an enhancement below M = 0.65 GeV and above M = 0.85 GeV. The broadening of the a; is a candidate for experimental observation. Moving to the intermediate mass region, one obtains 138 the results displayed in Fig. 30. In this calculation, the sensitivity of the results on parton equilibrium has been examined, and the reader is invited to consult the relevant reference for the details. The bottom line, however, is that the quarkgluon plasma contribution (as approximated by Born-term qq annihilation) does not dip below the Drell-Yan. This being said, the differential invariant mass distribution is predicted to be dominated by the semileptonic decays of correlated cc pairs, in the intermediate mass region 148 . This source has not been shown in the figures above. However, if the heavy quarks that are progenitor for the semileptonic decay loose energy in the strongly interacting medium, this background will be suppressed 139 . A direct measurement would go a long way in lifting the ambiguities 140 .
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
411
M ee [GeV] Fig. 30. Intermediate-mass lepton pair spectra at RHIC energies [138]. Contributions from the hadron gas (HG), quark-gluon plasma (QGP), and Drell-Yan (DY) are shown individually, along with their sum.
4.3. Electromagnetic
Signatures
of
Jets
It is appropriate to discuss the electromagnetic signatures of jets in an environment devoted to predictions, as the physics necessary for those to exist necessitates high energy and intensity machines such as RHIC and the LHC. The fate of high energy jets traversing hot and dense matter is a fascinating study, and this whole subfield has become known as that of "jet quenching". The manner in which high energy jets loose energy in a strongly interacting medium has been shown to depend on the nature of the medium itself141. Thus, jet tomography is expected to be a sensitive probe of the quark-gluon plasma. However, if jets and plasma interact in such a way that the jet characteristics are modified, the jet-plasma interactions could by the same token lead to the emission of electromagnetic radiation. As discussed earlier, the microscopic processes leading to real photon emission at the parton level are quark-antiquark annihilation, Compton scattering, as well as bremsstrahlung. Therefore, a fast quark passing through the plasma will produce photons by Compton scattering with the thermal gluons and annihilation with the thermal antiquarks 142 . Those processes are higherorder in a s , when compared with photons from initial hard scatterings, but
412
C. Gale and K. L. Haglin
they will not form a sub-leading contribution as they correspond to multiple scattering (actually, double scattering) which grows with the system size. Working out the details, one can show that the rate of production of real photons due to annihilation and Compton scattering is 114 Nf
M d3pMP)[1
^^ = ^V/ l
+ f9iP) (33)
x * ( 0 ) ( « f e ^ + («-«).
(34)
The ft are parton distribution functions. In order to proceed one may assume that those may be decomposed as f(p)
= /thermal (p) + /jet(p)
(35)
where the thermal component is characterised by a temperature T: /thermal = exp(—E/T). This separation is kinematically reasonable as the jet spectra fall of as a power law and can thus easily be differentiated from their thermal counterpart. The phase space distribution for the quark jets propagating through the QGP is given by the perturbative QCD result for the jet yield 143 : (2TT)3
1
/jet(p) =
52 gq
TTR2±TP±
x5(v
dNiet
,2 , R(r) d2pxdy
- J / ) 0 ( T - Ti)e(T m a x - T)Q(R±
(36) - r) ,
(37)
where gq = 2 x 3 is the spin and colour degeneracy of the quarks, R± is the transverse dimension of the system, T, ~ l/p± is the formation time for the jet and r) is the space-time rapidity. R(r) is a transverse profile function. r m a x is the smaller of the life-time ry of the QGP and the time Td taken by the jet produced at position r to reach the surface of the plasma. The boost invariant correlation between the rapidity y and 77 is assumed 58 . Fig. 31 contains the results for thermal photons, direct photons due to primary processes, bremsstrahlung photons and the photons coming from jets passing though the QGP in central collision of gold nuclei at RHIC energies. The corresponding results for LHC energies are shown in Fig. 32.
Electromagnetic
1
Radiation from Relativistic
Nuclear Collisions
413
Photons for Au+Au at S 1/2 =Ax200 GeV
„io"5
•S io-7 P.
— 7fromjets through QGP • — Direct photons - - Bremsstrahlung Thermal photons
|io-8
IO" 9
,-io
10'
8
12
10 PrlGeV]
14
Fig. 31. Spectrum dN/d2p±dy of photons at y = 0 for central collision of gold nuclei at VSNN = 200 GeV at R.HIC. Plotted 1 4 2 is the yield for photons from jets interacting with the medium (solid line), direct hard photons (long dashed), bremsstrahlung photons (short dashed) and thermal photons (dotted).
10"
Photons for Pb+Pb at S 1/2 =Ax5.5 TeV
a io"5 H
-R
•jf-io 6 10" IO"8
— 7 from jets through QGP • — Direct photons ~- Bremsstrahlung Thermal photons
10
12 PT
14
16
18
20
[GeV]
Fig. 32. The same as Fig. 31 for central collision of lead nuclei at T/SNN LHC. Taken from the same reference.
= 5.5 TeV at
It is seen that the quark jets passing through the QGP give rise to a large yield of high energy photons. This contribution should be absent in pp collisions. For RHIC this contribution is the dominant source of photons up to px. K, 6 GeV. The jet-to-photon conversion falls more rapidly with p± than the direct photon yield, similar to a higher twist correction. It is clear that this new mechanism for the production of high energy photons contributes significantly. In fact, it is the leading source of directly produced
414
C. Gale and K. L. Haglin
photons at RHIC in the region py < 6 GeV/c. Very similar considerations should apply to the production of lepton pairs, even though the details will of course change. Phase space now allows the direct annihilation of a quark and an antiquark into a dilepton. The cross section for this process is a(M2)
(38)
M2
3
/
where the sum runs over the flavour of quarks, Nc — 3, and s and e/ stand for the spin and the charge of the quark. Using kinetic theory, the reaction rate for the above process can be written as
i?
= /0 / a ( P a ) /^ / 6 ( P b ) < T ( M 2 K
"rel ,
(39)
where / ; stands for the phase-space distribution of the quark or the antiquark, p a and p& are their momenta respectively and the relative velocity is (for massless quarks) Vvel =
EaEb
-PaPb EaEb
(40)
After some algebra this can be rewritten as dR dM2
M6 a(M2) 2 (2?r)6
Jxa
dxa d(t>a xb dxb dcj>b dya dybfa fb
x 5 [M2 - 2M2xaxb
cosh(ya - yb) + 2M2xaxb
cos 0 6 ]
(41)
where xa = p ^ / M , Xb = v\jM and ya and yb are the rapidities. The integrations over the azimuthal angles yield Mia{M2)
dR dM2 x
(2TT)5
j xa dxa xb dxb dya dyb fa fb -1/2
^x\x\
(42)
- {2xaxb cosh(z/a - yb) - 1}
such that _..
<
2xaxb cosh(ya -yb)-l
<
ZXaXb
(43)
0 < Xa
- o o < ya,b < oo . When fa and fb are given by a thermal distribution / t h ( p ) = exp(-E/T)
= exp(-pT
cosh
y/T),
(44)
Electromagnetic
10-5
=
1
1
Radiation from Relativistic
o
A
10- 7
M
io-
\
\ \
^
A
8
N
X
\
\
Quark Jota =^L__V^ through tJGP X 10oV~
+"*
T3
Pb+Pb@SPS
Drell--Yan
Thermal C
3
415
r^
x
^
Nuclear Collisions
\
10~ 9
^
\-
_L
10-10 0
2
4 M (GeV)
6
8
Fig. 33. Dilepton spectrum 1 4 4 for P b + P b at ^/SNN = 17.4 GeV at SPS. The solid line is the result with To = 0.2 fm/c. The long dashed curves give the results when the formation time TO is raised to 0.50 fm/c, thus lowering the temperature. See the quoted reference for details.
the above integral can be performed to obtain Eq. (31). Again assuming a Bjorken-scenario isentropic plasma evolution, one can plot results for thermal dileptons, dileptons from the Drell-Yan process, and the dileptons from the passage of quark jets through the plasma for SPS, RHIC, and LHC respectively. Those constitute figures 33, 34 and 35. At SPS energies, we recover (Fig. 33) the well known result that the high mass dileptons have their origin predominantly in the Drell-Yan process. Increasing the formation time from 0.20 fm/c to 0.50 fm/c — and thus lowering the initial temperature by 100 MeV — drastically alters the thermal production (from the dash-dotted curve to the long-dashed one) while the yield from the proposed jet-plasma interaction, even though essentially negligible, is reduced by a factor of w 2 (from the solid line to the long-dashed one). The jet-plasma interaction starts playing an interesting role at RHIC energies (Fig. 34), as now the corresponding yield is about only one third of the Drell-Yan contribution, and is much larger than the thermal contribution. Again lowering the initial temperature (now by about 150 MeV) by increasing the formation time to 0.50 fm/c further enhances the importance of the yield due to jet-plasma interaction. This production is of the same
416
C. Gale and K. L. Haglin
10 - 5
r^
i n - 6 =-
>
i,
io-'
T3 01
? \
10-8
I
r xl
io~9
r
10' -10
Fig. 34.
•
•
•
I
•
•
Same as Fig. 33 for central A u + A u at ,/sJvJv = 200 GeV at RHIC.
10 - 4
CM
> O
10-5
Drell-Yan\
&
10-6
U
10 - 7
_
CM
S
10"
Fig. 35.
Same as Fig. 33 for central P b + P b at y/sJ^N = 5.5 TeV at LHC.
order as that attributed to secondary-secondary quark-antiquark annihilation in a dilepton production calculation done using an earlier version of the parton cascade model 145 . The much higher initial temperatures likely to be attained at the LHC
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
417
and the much larger (mini)jet production lead to an excess of high mass dileptons from jet-plasma interactions which can be an order of magnitude greater (at M = 10 GeV) than that due to the Drell-Yan process. Again, reducing the initial temperature by raising the formation time to 0.50 fm/c reduces the jet-plasma yield by about a factor of 2 while the thermal yield is reduced far more. Recall that at LHC energies several calculations 146,147 ' 148 have reported a thermal yield larger than the Drell-Yan production. It is found that the jet-plasma interaction enhances the high mass dilepton production considerably. The calculations outlined above can be repeated for a plasma that is not in chemical equilibrium 144 . In this case, conclusions similar to the ones reached above are obtained. Summarising, it appears that a unique source of high mass dileptons is generated by the passage of quark jets through the quark-gluon plasma. The contribution is seen to be the largest at LHC energies, moderate at RHIC energies, and negligible at SPS energies. The measurement of this radiation could then be added to the list of QGP signatures, as well as providing a proof of existence for the conditions suitable for jet-quenching to occur. Finally, even though this is not a "direct" plasma signal, it is worth mentioning that electromagnetic radiation can also serve as a versatile jettag. This is especially useful in environments where the jet is expected to loose energy, or to be quenched out of existence. This statement holds true for real photons 149 , as well as for lepton pairs 150 .
4.4. Squeezing
Lepton Pairs
out of Broken
Symmetries
We have seen in the text above that the electromagnetic radiation measured in nuclear collisions is a precious measure of the in-medium photon self-energy. However, in a bath of finite temperature and density, new possibilities can manifest themselves. In some sense, the medium allows for the existence of correlators that vanish identically in the vacuum. This fact opens mixing channels that were previously closed. While several studies have sought to investigate the in-medium properties of hadrons, their mixing with other hadrons has up to now received little attention. An exception is the case of p — to 130 . This specific mixing may be omitted when dealing with isospin symmetric nuclear matter, as is done here. Also, one will concentrate on vector mesons, as they enjoy a privileged relationship with electromagnetic signals. First, one describes an exploratory calculation designed to highlight an eventual signal. The possibility of p — ao mixing is explored, via nucleon-nucleon excitations in strongly interacting systems.
418
C. Gale and K. L. Haglin
It is shown that this mixing opens up a new channel for the dilepton production and may thus induce an additional peak in the
+ JlB-Va ZTTIN
g^NN^l^^
T^Var & .
(45)
where \I/, <j>a, <j>ao, p, and u> correspond to nucleon, a, ao p, and ui fields, and TI, is a Pauli matrix. The values for the coupling parameters are from [131]. The existence of a preferred rest frame essentially creates a new vacuum state with quantum numbers different from those of the true vacuum. An immediate consequence of this fact is that one can now define a mixed correlator involving scalar and vector current operators: (jsj^)- This mixed correlator is identically zero in the true vacuum. The polarisation vector through which the ao couples to the p via the N — N loop is given by
/
d4k ^-^Tr[G(k)T^G(k
+ q)} ,
(46)
where 2 is an isospin factor and the vertex for p — N — N coupling is
In the above G(k) is the in-medium nucleon propagator 132 . For the sake of simplicity, one first uses the density-dependent and temperatureindependent propagator. This approximation will be relaxed later. With the evaluation of the trace and after a little algebra, Eq. 46 can be cast in a suggestive form: n , ,-h 9a0gPo2f0 * W2\ n M (ft, | 9 1) = - ^ 2q (2mN - _ j
fkF d3k ^ —
k^-pjk-q) qi
_
4{k
q?
.(48)
This leads immediately to two conclusions. First, q^IP = 0. Second, only two components of the polarisation vector survive after the angular integration. This will guarantee that only the longitudinal component of the p couples to the scalar meson, while the transverse mode remains unaltered. Further note that current conservation implies that out of the two nonzero components of IIM, only one is independent. The new mixing channel will affect the properties of the mesons in medium, i.e. affect their masses and
Electromagnetic
Radiation from Relativistic Nuclear Collisions
^0 S~\
\
P
419
/
7T
Fig. 36. mixing.
Feynman diagram for the process -K+T) —> e+e , which proceeds through p — ao
spectral function. Those aspects will not be discussed at length here, but details can be found in the literature 134 ' 137 . The p — ao mixing opens a new channel in dense nuclear matter, 77 + 77 —• + e e~, which may proceed through N-N excitations. The Feynman diagram for this process is shown in Fig. 36. One may then evaluate cross sections for the production of lepton pairs. Evaluating the polarisation in the zerotemperature limit, the cross section for this process is _ 4TTQ2 gl^r, 2
rrf
(M + m2pT2p{M) 1 1 2 2 2 2 2 " (M - mao) + m aoT o(M) JM - Am2 ' ^ 3q M
g
2
2
m2p)2
' '
(49)
where IIo is the zeroth component of the expression in Eq. (46). The numerical values for the couplings and the calculation details are to be found in Ref. [151]. The cross sections are shown in Fig. 37, for two different values of the nuclear density. The familiar process irir —» e+e~ is also plotted to set a scale. A noticeable feature of this plot is that the mixing process induces a peak at mao = 0.985 GeV. This constitutes a genuine in-medium effect which is mostly density-driven: this peak would be completely absent in vacuum. Furthermore, the contribution to the cross section at the ao mass is comparable in magnitude to that of the TTTT channel at its peak. Calculations of emission rates where the T = 0 simplification was not made also support this claim 136 . Those mixing effects will also affect the bulk features of in-medium mesonic behaviour 137 . A natural question to ask is whether those symmetry-breaking effects in the dilepton spectrum should have been observed in any of the past or present experiments. It turns out that the required baryonic density is too transient to influence significantly this signal at CERN energies 152 . However, the HADES 153 experiment at the GSI, in Darmstadt, Germany, has the resolution and the sensitivity to explore the appropriate invariant mass range. Those heavy ion reactions are
420
C. Gale and K. L. Haglin
(a)
J2
4
0.0
0.2
0.4
0.6
0.B
1.0
1.2
0.0
0.2
M (GeV)
0.4
0.6
0.8
1.0
1.2
M (GeV)
Fig. 37. Dilepton spectrum induced by -K + n —> e + + e~ and 7r + r\ —» e + + e~ considering matter-induced p — arj mixing, (a) p=1.5po (b) p=2.5po performed at lower energies, with respect to those of CERN, and thus lead to higher baryonic densities which persist longer. Evidencing continued theoretical interest at these lower energies, there are new advancements 154 ' 156 in theory pertinent for the dilepton measurements made by the DLS 155 , and to be made by HADES. There is finally another symmetry-breaking with an electromagnetic signature that will be mentioned, even though quantitative evaluations are still in their preliminary stage. At zero temperature, and at finite temperature and zero charge density, diagrams in QED that contain a fermion loop with an odd number of photon vertices are cancelled by an equal and opposite contribution coming from the same diagram with fermion lines running in the opposite direction, this is the basic content of Furry's theorem 157 (see also [158, 159]). In the language of operators we note that these diagrams are encountered in the perturbative evaluation of Green's functions with an odd number of gauge field operators: (0|A M iA M 2...A M 2n+l|0). In QED we know that CAflC~1 = — A^, where C is the charge conjugation operator. In the case of the vacuum |0), we note that C|0) = |0), as the vacuum is uncharged. As a result
<0|^1^a...^2n+1|0) =
(0\C-1CAtllC-1CA,2...A,2n+1C-1C\0)
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
421
Ho|A^ 2 ...^ 2 n + 1 |o>(-i) 2 " + 1 = -<0|A M l A M 2 ... J 4 M 2 n + 1 |0) = 0.
(50)
In an equilibrated medium at a temperature T, we not only have the expectation of the operator on the ground state but on all possible matter states weighted by a Boltzmann factor:
X)H^1AMa...AWB+1|n)e-^--"«-)I n
where /? = 1/T and /x is a chemical potential. We are thus calculating the expectation in the grand canonical ensemble. Here, C\n) = e l *| — n), where | — n) is a state in the ensemble with the same number of antiparticles as there are particles in \n) and vice-versa. If /x = 0 i.e., the ensemble average displays zero density then inserting the operator C~1C as before, we get ( n | A M l A^
. . . Afj,2ri+1
\n)e
= -(-n\AlilAfl2...A^n+1\-n)e-l3E" The sum all states will contain the mirror term {—r^A^A^ with the same thermal weight, hence
(51) ...
over Al^n+1\—n)e~l3En,
5 > l 4 , i Am ... A^+1 \n)e~^ = 0,
(52)
n
(the expectation over states which are excitations of the vacuum |0) will again be zero as in Eq. 50) and Furry's theorem still holds. However, if \i ^ 0 (=> unequal number of particles and antiparticles ) then
= -{-n\AlllAlta...All2n+1\-n)e-nE~->*~)t the mirror term this time is (—n|j4Mli4Ma ... AM2n+11 — n)e~^^En+llCin\ a different thermal weight, thus ^2(n\AlilA„2...A^n+1\n)e-0(E^Q")^O.
(53) with
(54)
n
The sum over all medium states which leads to the expectation value is no longer zero. This represents the breaking of Furry's theorem by the medium. Note that this occurs only in media with non-zero density or chemical potential. Summarising, if the medium contains a net charge, such that
422
C. Gale and K. L. Haglin
it breaks C spontaneously, Green's functions that vanish identically in the vacuum (or in a neutral medium) can survive. Making the simplest possible extension to QCD, one may replace two of the photons with gluons. This enables processes like gg —» £+£~, where the gluon fusion proceeds through a quark (antiquark) loop. This channel is exciting for the following reason: it offers a direct electromagnetic signature of early gluon populations. This represents pristine information on the state of the many-body system. Calculations are technically involved160, but results are finally forthcoming 161 .
5. Conclusions This ends our survey of the use of electromagnetic signals as probes of strongly-interacting relativistic many-body dynamics. The supporting framework has been relativistic quantum field theory, generalised to finite temperatures and densities, in order to formulate very general computational tools for estimating production rates from heated and compressed nuclear systems. The focus was on establishing within equilibrium circumstances, the number of radiated electromagnetic quanta per unit volume per unit time. For then, one can, and did, take the rate and evolve according to some "best guess" scenarios for the expansion dynamics in heavy ion collisions realized at facilities at SPS and RHIC. Dilepton experimental circumstances resembling those expected at the CERN SPS have been modeled in a variety of ways. Comparisons of theory and experiment have been fruitful in terms of suggesting a consistent picture of modified vector meson spectral properties. These modification are truly collective nuclear effects. Tremendous advancement in our understanding of the way in which nuclear matter responds when it is forced near the phase boundary between hadrons and quarks has come from these pursuits. To mention some specific achievements, one notes that the rho spectral function has been essentially measured at finite energy density and has been shown to be significantly modified from its vacuum structure. This is a significant achievement. The photon studies have been exceptionally fruitful too. Theory has advanced from a stage where rate estimates where plagued with infrared singularities, to first establishing regulated lowest-order results with new computational techniques and tools. The hard-thermal-loop approach is useful not only for photon production, but other studies in hot gauge theories and for a variety of observables. Next, one witnessed an impressive effort to understand the photon self-energy up to the many-loop order and includ-
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
423
ing multiple-scattering effects within the medium. The celebrated result is a complete lowest order photon production rate from finite temperature QCD that is stable and reliable. That too is a nontrivial accomplishment. When the QCD rates and hadronic rates are used to predict photon yields and then compared with experiments WA80 and WA98 from CERN, there is consistency, if not discriminatory features. One can say that the results strongly suggest thermal emission from a fireball at roughly 200 MeV temperature. Have we observed the QGP? The investigations reported on in this work only contain hints of an answer. As mentioned previously, the radiation from the partonic phase is present in the analyses, but does not constitute a large portion of the overall signal. Fortunately, this state of affairs is directly related to probed temperatures and to the space-time volume occupied by the plasma. Both those are expected to increase in the current and future generations of collider experiments. The only indirect proof is that, in many dynamical simulations, it is unavoidable for the initial phase to be elsewhere in the phase diagram than in the deconfined region. This is a consistency requirement brought about by our current knowledge of the equation of state. This however will change and evolve, especially with the experience gained with non-perturbative approaches which will in turn guarantee a better focused picture of the quasiparticle nature of the partonic sector. Have we observed chiral symmetry restoration? The approach to chiral symmetry is closely related to the properties of the inmedium spectral densities 162 . As the constraints of the Weinberg sum rules are extended to finite temperatures, they require a degeneracy of the vector and axial-vector correlators in the symmetric limit. This demand, however, can be satisfied in several ways 162 . Thus, a verdict on the status of chiral symmetry restoration is being hindered by the difficulty to access the axialvector correlator unambiguously. More theoretical work needs to be done in that respect as well, in order to provide a unified calculation with a credible degree of sophistication. It should finally be mentioned that RHIC and the LHC will have the intensity to make possible Hanbury-Brown-Twiss interferometry measurements of direct photons. The theory of this observable is well-developed163, and measurements of the correlation functions are expected to place constraints on the space-time extent of the photon-emitting sources 164 . If the estimates brought about by dynamical approaches and by analyses of hadrons abundances are reliable, we have just grazed the phase boundary of the deconfined sector at the SPS. This assertion is not in conflict with the evidence obtained from measurements of electromagnetic observables.
424
C. Gale and K. L. Haglin
Then, RHIC, and the LHC, should soon signal bold incursions into a new territory. Acknowledgements It is a pleasure to acknowledge helpful comments from, and discussions with, P. Aurenche, J.-e. Alam, A. Bourque, E. Bratkovskaya, A. Dutt-Mazumder, S. Gao, F. Gelis, P. Jaikumar, B. Kampfer, J. Kapusta, V. Koch, I. Kvasnikova, G. D. Moore, R. Rapp, D. Srivastava, O. Teodorescu, Y. Tserruya, and S. Turbide. C. G. thanks J. Bruce, J. Entwistle, J. Pastorius, Ch. Squire, and G. Willis for inspiration. This work has been supported in part by the Natural Sciences and Engineering Research Council of Canada, in part by the the Fonds Nature et Technologies of Quebec, and in part by the National Science Foundation under grant number PHY-0098760. References 1. F. Karsch and E. Laermann, this volume; F. Karsch, AIP Conf. Proc. 631, 112 (2003). 2. K. Kanaya, Nucl. Phys. A 715, 233c (2003). 3. Z. Fodor and F. Karsch, JHEP 03, 014 (2002). 4. M. Stephanov, K. Rajagopal, and E. Shuryak, Phys. Rev. Lett. 81, 4816 (1998). 5. See, for example, K. Rajagopal and F. Wilczek, At the Frontiers of Particle Physics/Handbook of QCD, M. Shifman ed., (World Scientific, Singapore, 2000), and references therein. 6. Prashanth Jaikumar, Ralf Rapp, and Ismail Zahed, Phys. Rev. C 65, 055205 (2002). 7. E. L. Feinberg, Nuovo Cimento 34A, 391 (1976). 8. L.D. McLerran and T. Toimela, Phys. Rev. D 31, 545 (1985). 9. C. M. Ko, L. H. Xia, and P. J. Siemens, Phys. Lett. B231, 16 (1989). 10. H. A. Weldon, Phys. Rev. D 42, 2384 (1990). 11. C. Gale and J. I. Kapusta, Nucl. Phys. B357, 65 (1991). 12. E. S. Fradkin, Proc. Lebedev Physics Institute 29, 6 (1965). 13. See, for example, Michael E. Peskin and Daniel V. Schroeder, An Introduction to Quantum Field Theory, (Addison-Wesley, Reading, 1995), and references therein. 14. Charles Gale and Peter Lichard, Phys. Rev. C 49, 3338 (1994). 15. G. Chanfray and P. Schuck, Nucl. Phys. A555, 329 (1993). 16. G. Chanfray, R. Rapp, and J. Wambach, Phys. Rev. Lett. 76, 368 (1996). 17. R. Rapp, G. Chanfray, and J. Wambach, Nucl. Phys. A617, 472 (1997). 18. J. J. Sakurai, Currents and Mesons, (University of Chicago Press, Chicago, 1969). 19. F. Klingl, N. Kaiser, and W. Weise, Z. Phys. A 356, 193 (1996).
Electromagnetic Radiation from Relativistic Nuclear Collisions 425 20. H. B. O'Connel, B. C. Pearse, A. W. Thomas, and A. G. Williams, Prog. Part. Nucl. Phys. 39, 201 (1997). 21. Ralf Rapp and Charles Gale, Phys. Rev. C 60, 024903 (1999). 22. Sangyong Jeon and Paul. J. Ellis, Phys. Rev. D 58, 045013 (1998). 23. W. Cassing and E. L. Bratkovskaya, Phys. Rept. 308, 65 (1999). 24. R. Rapp and J. Wambach, Adv. Nucl. Phys. 25, 1 (2000). 25. P. Huovinen, M. Belkacem, P.J. Ellis, and J.I. Kapusta, Phys. Rev. C 66, 014903 (2002). 26. G. Agakichiev et al., Phys. Rev. Lett. 75, 1272 (1995); Phys. Lett. B422, 405 (1998). 27. M. Masera et al, Nucl. Phys. A590, 93c (1995). 28. G. E. Brown and Mannque Rho, Phys. Rept. 269, 333 (1996). 29. See, for example, C. Gale, Nucl. Phys. A698, 143 (2002), and references therein. 30. Johannes P. Wessels et al. (CERES collaboration), Nucl. Phys A715, 262c (2003). 31. E. Shuryak, Phys. Lett. B79, 135 (1978). 32. A. L. S. Angelis et al, Eur. Phys. J. C 13, 433 (2000). 33. M. C. Abreu et al, Phys. Lett. B368, 230 (1996); Nucl. Phys. A610, 331 (1996). 34. C. Y. Wong and Z. Q. Wang, Phys. Lett. B367, 50 (1996). 35. C. Spieles et al, Eur. Phys. J. C 5, 349 (1998). 36. Z. Lin and X. N. Wang, Phys. Lett. B444, 245 (1998). 37. L. Capelli, PhD thesis, Universite Claude Bernard, 2001. 38. P. Levai, B. Miiller, and X. N. Wang, Phys. Rev. C 51, 3326 (1995). 39. P. Bordalo et al, Nucl. Phys. A661, 538c (1999). 40. G. Q. Li and C. Gale, Phys. Rev. Lett. 81, 1572 (1998); Phys. Rev. C 58, 2914 (1998). 41. L. Xiong, E. V. Shuryak, and G. E. Brown, Phys. Rev. D 46, 3798 (1992); C. Song, C. M. Ko, and C. Gale, Phys. Rev. D 50, R1827 (1994); J. K. Kim et al, Phys. Rev. D 53, 4787 (1996). 42. Particle Data Group, Phys. Rev. D 66, 010001 (2002). 43. S. Gao and C. Gale, Phys. Rev. C 57, 254 (1998). 44. L. M. Barkov al, (OLYA collaboration), Nucl. Phys. B256, 365 (1985). 45. D. Bisello et al, (DM2 collaboration), Phys. Lett. B220, 321 (1989). 46. M. E. Biagini, S. Dubnicka, E. Etim, snd P. Kolar, Nuovo Cimento A 104, 363 (1991). 47. R. R. Akhmetshin et al, (CMD-2 collaboration), Phys. Lett. B364, 199 (1995). 48. D. Bisello et al, (DM2 collaboration), Z. Phys. C 39, 13 (1988). 49. P. M. Ivanov et al, (OLYA collaboration), Phys. Lett. B107, 297 (1981). 50. K. Haglin and C. Gale, Phys. Rev. D 52, 6297 (1995). 51. S. I. Dolinsky et al, (ND collaboration), Phys. Rept. 202, 99 (1991). 52. N. Albrecht al, (ARGUS collaboration), Phys. Lett. B185, 223 (1987). 53. H. Gomm, 6 . Kaymakcalan, and J. Schechter, Phys. Rev. D 30, 2345 (1984). 54. D. Bisello et al, (DM2 Collaboration), Proceedings of Hadrons '91, (College
426
55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81.
82. 83. 84. 85. 86. 87.
C. Gale and K. L. Haglin Park, 1991). Z. Huang, Phys. Lett. B361, 131 (1995). I. Kvasnikova, PhD thesis, McGill University, 2001. Z. Huang, private communication. J. D. Bjorken, Phys. Rev. D 27, 140 (1983); R. C. Hwa and K. Kajantie, ibid. 32, 1109 (1985); M. Gyulassy and T. Matsui, Phys. Rev. D 29, 419 (1984). J. Cleymans, K. Redlich, and D. K. Srivastava, Phys. Rev. C 55, 1431 (1997); Phys. Lett. B420, 261 (1998). Ioulia Kvasnikova, Charles Gale, and Dinesh Srivastava, Phys. Rev. C 65, 064903 (2002). K. Gallmeister, B. Kampfer, O.P. Pavlenko, and C. Gale, Nucl. Phys. A688, 939 (2001). R. Rapp and E. Shuryak, Phys. Lett. B473, 13 (2000). D. K. Srivastava, B. Sinha, I. Kvasnikova, and C. Gale, Nucl. Phys. A698, 432c (2002). M. C. Abreu et al. (NA50 Collaboration), Eur. Phys. J. C 14, 443 (2000). D. K. Srivastava, Phys Rev. C 64, 064901 (2001); J.-Y. Ollitrault, Phys. Lett. B273, 31 (1991). J.I. Kapusta, P. Lichard, and D. Seibert, Phys. Rev. D 44 2774 (1991); Erratum-ibid. D47, 4171 (1993). P. Singer, Phys. Rev. 130 2441 (1963); 161 1694 (1967). H. Nadeau, J.I. Kapusta, and P. Lichard, Phys. Rev. C 45, 3034 (1992); Erratum Phys. Rev. C 47, 2426 (1993). L. Xiong, E. Shuryak and G.E. Brown, Phys. Rev. D 46 3798 (1992). C. Song, Phys. Rev. C 47 2861 (1993). Ralf Rapp, Simon Turbide, and Charles Gale, in preparation. M. Halasz, J.V. Steele, G. Li, G.E. Brown, Phys. Rev. C 58 365 (1998). K. Haglin, Phys. Rev. C 50 1688 (1994). K.L. Haglin, Nucl. Phys. A 584 719 (1995). G.E. Brown and M. Rho, Phys. Rev. Lett. 66 2720 (1991). S. Sarkar, J. Alam, P. Roy, A.K. Dutt-Mazumder, B. Dutta-Roy, and B. Sinha, Nucl. Phys. A 634 206 (1998). C. Song and G. Fai, Phys. Rev. C 58 1689 (1998). P. Roy, S. Sarkar, J. Alam, and B. Sinha, Nucl. Phys. A 653 277 (1999). J.V. Steele, H. Yamagishi, and I. Zahed, Phys. Lett. B 384 255 (1996). J.V. Steele, H. Yamagishi, and I. Zahed, Phys. Rev. D 56 5605 (1997). H A . Weldon, Phys. Rev. D 28, 2007 (1983); R. L. Kobes and G. W. Semenoff, Nucl. Phys. B260, 714 (1985); S. M. H. Wong, Phys. Rev. D 64, 025007 (2001); A. Majumder and C. Gale, Phys. Rev. C 65, 055203 (2002). E.V. Shuryak, Yad. Fiz. 28, 796 (1978). K. Kajantie and H.I. Miettinen, Z. Phys. C 9, 341 (1981). R.C. Hwa and K. Kajantie, Phys. Rev. D 32, 1109 (1985). R. Baier, H. Nakkagawa, A. Niegawa, and K. Redlich, Z. Phys. C 53, 433 (1992). R.D. Pisarski, Nucl. Phys. 309, 476 (1988); Phys. Rev. Lett. 63, 1129 (1989). E. Braaten and R.D. Pisarski, Nucl. Phys. B 337, 569 (1990).
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
427
88. T. Peitzmann and M. Thoma, Phys. Rept. 364, 175 (2002). 89. P. Aurenche, F. Gelis, H. Zaraket, and R. Kobes, Phys. Rev. D 58, 085003 (1998). 90. P. Arnold, G.D. Moore, and L.G. Yaffe, JHEP 0112, 009 (2001). 91. P. Aurenche, F. Gelis, H. Zaraket, Phys. Rev. D 61, 116001 (2000). 92. P. Aurenche, F. Gelis, H. Zaraket, Phys. Rev. D 62, 096012 (2000b). 93. L.D. Landau and I. Pomeranchuk, Dokl. Akad. Nauk Ser. Fiz. 92, 535 (1953). 94. A.B. Migdal, Phys. Rev. 103, 1811 (1956). 95. R. Albrecht et al., The WA80 Collaboration, Phys. Rev. Lett. 76, 3506 (1996). 96. M.M. Aggarwal et a l , The WA98 Collaboration, Phys. Rev. Lett. 85, 3595 (2000). 97. E.V. Shuryak and L. Xiong, Phys. Lett. B 333, 316 (1994); Erratum, ibid., B 339, 334 (1994). 98. D.K. Srivastava and B. Sinha, Phys. Rev. Lett., 73, 2421 (1994). 99. A. Dumitru, U. Katscher, J.A. Maruhn, H. Stocker, W. Greiner, and D.H. Rischke, Phys. Rev. C 51, 2166 (1995). 100. D.K. Srivastava and B. Sinha, Eur. Phys. J. C 12 109 (2000); Erratum, ibid, C 20 397 (2000). 101. J. Cleymans, K. Redlich, and D.K. Srivastava, Phys. Rev. C 55, 1431 (1997). 102. C. Y. Wong and H. Wang, Phys. Rev. C 58, 376 (1998). 103. P. Aurenche, R. Baier, M. Fontannaz, and D. Schiff, Nucl. Phys. B286, 509 (1987); Nucl. Phys. B297, 661 (1988). 104. J. C. Collins, D. E. Soper, and G. Sterman, Nucl. Phys. B250, 199 (1985). 105. P. Aurenche et al, Eur. Phys. J. C 9, 107 (1999). 106. Geoffrey T. Bodwin, Stanley J. Brodsky, and G. Peter Lepage, Phys. Rev. D 39, 3287 (1989). 107. A. Dumitru, L. Frakfurt, L. Gerland, H. Stocker, and M. Strikman, Phys. Rev. C 64, 054909 (2001). 108. K. Gallmeister, B. Kampfer, and O. P. Pavlenko, Phys. Rev. C 62, 057901 (2000). 109. D.K. Srivastava and B. Sinha, Phys. Rev. C 64 034902 (2001). 110. J.-e Alam, S. Sarkar, T. Hatsuda, T.K. Nayak, and B. Sinha, Phys. Rev. C 63, 021901R (2001). 111. J. Sollfrank, P. Huovinen, M. Kataja, P.V. Ruuskanen, M. Prakash, and R. Venugopalan, Phys. Rev. C55, 392 (1997). 112. P. Huovinen, P.V. Ruuskanen, and J. Sollfrank, Nucl. Phys. A 650, 227 (1999). 113. P. Huovinen, P. V. Ruuskanen, and S. S. Rasanen, Phys. Lett. B 535, 109 (2002). 114. C. Y. Wong, Introduction to Heavy Ion Collisions, (World Scientific, Singapore, 1994). 115. Steffen A. Bass, Berndt Miiller, and Dinesh K. Srivastava, Phys. Rev. Lett. 90, 082301 (2003). 116. K. Geiger and B. Miiller, Nucl. Phys. B369, 600 (1992). 117. S. A. Bass, B. Miiller, and D. K. Srivastava, Phys. Lett. B 551, 277 (2003).
428 118. 119. 120. 121. 122. 123. 124. 125.
C. Gale and K. L. Haglin
E. Braaten, R.D. Pisarski, and T.C. Yuan, Phys. Rev. Lett. 64, 2242 (1990). S. M. H. Wong, Z. Phys. C 53, 465 (1992). K. Haglin, C. Gale, and V. EmePyanov, Phys. Rev. D 47, 973 (1993). P. Aurenche, F. Gelis, and H. Zaraket, JHEP 0207 063 (2002). P. Aurenche, F. Gelis, G.D. Moore and H. Zaraket, JHEP 0212 006 (2002). T. Altherr, P.V. Ruuskanen, Nucl. Phys. 380, 377 (1992). M. H. Thoma and C.T. Traxler, Phys. Rev. D 56, 198 (1997). For a review of this technology, see M. Jarrel and J.E. Gubernatis, Phys. Rept. 269, 133 (1996). 126. I. Wetzorke, hep-lat/0305012; M. Asakawa, T. Hatsuda, and Y. Nakahara, Nucl. Phys. A 715, 863c (2003). 127. N. Hammon et al., Phys. Rev. C 57, 3292 (1998). 128. D.Y. Peressounko and Y.E. Pokrovsky, Nucl. Phys. A 669, 196 (2000). 129. F.D. Steffen and M. H. Thoma, Phys. Lett. B 510, 98 (2001) 130. A. K. Dutt-Mazumder, B. Dutta-Roy, and A. Kundu, Phys. Lett. B399, 196 (1997); W. Broniowski and W. Florkowski, Phys. Lett. B440, 7 (1998). 131. R. Machleidt, Adv. Nucl. Phys. 19, 198 (1989). 132. B. Serot and J. D. Walecka, Adv. Nucl. Phys. 16, 1 (1986). 133. S. A. Chin, Ann. Phys. (N.Y.), 108, 301 (1977). 134. K. Saito, K. Tsushima, A. W. Thomas, A. G. Williams, Phys. Lett. B433, 243 (1998). 135. H. A. Weldon, Phys. Lett. B274, 133 (1992); G. Wolf, B. Friman, and M. Soyeur, Nucl. Phys. A640, 129 (1998). 136. O. Teodorescu, A. K. Dutt-Mazumder, and Charles Gale, Phys. Rev. C 63, 034903 (2001). 137. O. Teodorescu, A. K. Dutt-Mazumder, and Charles Gale, Phys. Rev. C 66, 015209 (2002). 138. Ralf Rapp, Phys. Rev. C 63, 054907 (2001). 139. E. Shuryak, Phys. Rev. C 55, 961 (1997); Z. Lin, R. Vogt, and X.-N. Wang, ibid 57, 899 (1998); B. Kampfer and O. P. Pavlenko, Phys. Lett. B 391, 185 (1997). 140. http://na60.web.cern.ch/NA60/ 141. Miklos Gyulassy, Ivan Vitev, Xin-Nian Wang, and Ben-Wei Zhang, this volume. 142. Rainer J. Fries, Berndt Miiller, and Dinesh K. Srivastava, Phys. Rev. Lett. 90, 132301 (2003). 143. Z. Lin and M. Gyulassy, Phys. Rev C 51 217 (1995). 144. Dinesh K. Srivastava, Charles Gale, and Rainer J. Fries, Phys. Rev. C 67, 034903 (2003). 145. K. Geiger and J. I. Kapusta, Phys. Rev. Lett. 70, 1920 (1993). 146. J. Kapusta, L. D. McLerran and D. K. Srivastava, Phys. Lett. B 283, 145 (1992). 147. D. K. Srivastava, M. G. Mustafa and B. Miiller, Phys. Rev. C 56, 1064 (1997). 148. R. Vogt, B. V. Jacak, P. L. McGaughey and P. V. Ruuskanen, Phys. Rev. D 49, 3345 (1994); S. Gavin, P. L. McGaughey, P. V. Ruuskanen and R. Vogt,
Electromagnetic
Radiation from Relativistic
Nuclear Collisions
429
Phys. Rev. C 54, 2606 (1996); D. Fein, Z. P. Valerio and I. Sarcevic, Phys. Rev. C 56, 1637 (1997). 149. X. N. Wang, Z. Huang, and I. Sarcevic, Phys. Rev. Lett. 77, 231 (1996); X. N. Wang and Z. Huang, Phys. Rev. C 55, 3047 (1997). 150. Dinesh K. Srivastava, Charles Gale, and T. C. Awes, Phys. Rev. C, in press. 151. O. Teodorescu, A. K. Dutt-Mazumder, and C. Gale, Phys. Rev. C 61, 051901 (R) (2000). 152. O. Teodorescu, PhD thesis, McGill University 2002. 153. J. Friese et al., Prog. Part. Nucl. Phys. 42, 235 (1999). 154. E. L. Bratkovskaya, W. Cassing, M. Effenberger, and U. Mosel, Nucl. Phys. A 653, 301 (1999); E. L. Bratkovskaya, W. Cassing, and U. Mosel, ibid. 686, 568 (2001); E. L. Bratkovskaya, ibid. 696, 309 (2001). 155. R. J. Porter et al., Phys. Rev. Lett. 79, 1229 (1997); W. K. Wilson et al, Phys. Rev. C 57, 1865 (1998). 156. R. Shyam and U. Mosel, hep-ph/0303035; K. Schekhter et al., nuclth/0305015. 157. W. H. Furry, Phys. Rev. 51 135 (1937). 158. C. Itzykson and J. B. Zuber, Quantum Field Theory, McGraw Hill, New York, (1980). 159. S. Weinberg, The Quantum Theory of Fields, Vol. 1, Cambridge University Press, (1995). 160. A. Majumder, PhD thesis, McGill University 2002; A. Bourque, MSc thesis, McGill University 2002. 161. A. Bourque, A. Majumder, and C. Gale, in preparation. 162. J. I. Kapusta and Edward V. Shuryak, Phys. Rev. D 49, 4694 (1994). 163. R. Hanbury-Brown and R. Q. Twiss, Nature 178, 1046 (1956); G. Goldhaber, S. Goldhaber, W. Lee, and A. Pais, Phys. Rev. 120, 300 (1960); U. Heinz and B. Jacak, Ann. Rev. Nucl. Part. Sci. 49, 529 (1999); R. M. Weiner, Phys. Rept. 327, 249 (2000). 164. D. Neuhauser, Phys. Lett. B 182, 289 (1986); D. K. Srivastava and J. I. Kapusta, Phys. Lett. B 307, 1 (1993); D. K. Srivastava and C. Gale, Phys. Lett. B 319, 407 (1993); D. Peressounko, Phys. Rev. C 67, 014905 (2003); Jan-e Alam et al., Phys. Rev. C 67, 054902 (2003).
EVENT BY EVENT
FLUCTUATIONS
S. Jeon Physics Department, McGill University Montreal, QC H3A-2T8, Canada and Riken-Brookhaven Research Center Brookhaven National Laboratory Upton, NY 11973, USA E-mail: jeonhep.physics, mcgill. ca
V. Koch Lawrence Berkeley National Laboratory Berkeley, CA, 94720, USA E-mail: vkochlbl.gov
Contents 1 Introduction 432 2 Fluctuations in a Thermal System 435 2.1 Fluctuations in a grand canonical ensemble 435 2.1.1 Fluctuations of the energy and of the conserved charges . . 436 2.2 Fluctuations in a canonical ensemble 440 2.3 Phase transitions and fluctuations 442 3 Other Fluctuations 444 3.1 Volume fluctuations 444 3.2 Fluctuation from initial collisions 446 4 Fluctuations and Correlations 450 4.1 Correlations and charge fluctuations 453 4.2 Correlations and balance function 462 5 Observable Fluctuations 464
430
Event by Event Fluctuations 5.1 5.2
Fluctuations of ratios Fluctuations of the mean energy and mean transverse momentum 5.3 Fluctuations of particle ratios 5.4 Kaon fluctuations 6 Experimental Situation 6.1 Fluctuation in elementary collisions 6.2 Fluctuations in heavy ion reactions 7 Conclusions and Outlook References
431 469 471 474 477 480 480 482 485 487
432
S. Jeon and V. Koch
1. Introduction The study and analysis of fluctuations are an essential method to characterize a physical system. In general, one can distinguish between several classes of fluctuations. On the most fundamental level there are quantum fluctuations, which arise if the specific observable does not commute with the Hamiltonian of the system under consideration. These fluctuations probably play less a role for the physics of heavy ion collisions. Second, there are "dynamical" fluctuations reflecting the dynamics and responses of the system. They help to characterize the properties of the bulk (semi-classical) description of the system. Examples are density fluctuations, which are controlled by the compressibility of the system. Finally, there are "trivial" fluctuations induced by the measurement process itself, such as finite number statistics etc. These need to be understood, controlled and subtracted in order to access the dynamical fluctuations which tell as about the properties of the system. Fluctuations are also closely related to phase transitions. The well known phenomenon of critical opalescence is a result of fluctuations at all length scales due to a second order phase transition. First order transitions, on the other hand, give rise to bubble formation, i.e. density fluctuations at the extreme. Considering the richness of the QCD phase-diagram as sketched in Fig.l the study of fluctuations in heavy ions physics should
-Tri-Critical (End) point
Fig. 1.
Schematic picture of the QCD phase-diagram
lead to a rich set of phenomena. The most efficient way to address fluctuations of a system created in a heavy ion collision is via the study of event-by-event (E-by-E) fluctua-
Event by Event Fluctuations
433
tions, where a given observable is measured on an event-by-event basis and the fluctuations are studied over the ensemble of the events. In most cases (namely when the fluctuations are Gaussian) this analysis is equivalent to the measurement of two particle correlations over the same region of acceptance 12 . Consequently, fluctuations tell us about the 2-point functions of the system, which in turn determine the response of the system to external perturbations. In the framework of statistical physics, which appears to describe the bulk properties of heavy ion collisions up to RHIC energies, fluctuations measure the susceptibilities of the system. These susceptibilities also determine the response of the system to external forces. For example, by measuring fluctuations of the net electric charge in a given rapidity interval, one obtains information on how this (sub)system would respond to applying an external (static) electric field. In other words, by measuring fluctuations one gains access to the same fundamental properties of the system as "table top" experiments dealing with macroscopic probes. In the later case, of course, fluctuation measurements would be impossible. In addition, the study of fluctuations may reveal information beyond its thermodynamic properties. If the system expands, fluctuations may be frozen in early and thus tell us about the properties of the system prior to its thermal freeze out. A well known example is the fluctuations in the cosmic microwave background radiation, as first observed by COBE 1 . The field of event-by-event fluctuations is relatively new to heavy ion physics and ideas and approaches are just being developed. So far, most of the analysis has concentrated on transverse momentum and the net charge fluctuations. Transverse momentum fluctuations should be sensitive to temperature/energy fluctuations 57>53. These in turn provide a measure of the heat capacity of the system 4 4 . Since the QCD phase transition is associated with a maximum of the specific heat, the temperature fluctuations should exhibit a minimum in the excitation function. It has also been argued 55 ' 56 that these fluctuations may provide a signal for the long range fluctuations associated with the tri-critical point of the QCD phase diagram. In the vicinity of the critical point the transverse momentum fluctuations should increase, leading to a maximum of the fluctuations in the excitation function. Charge fluctuations 6 ' 3 5 , on the other hand, are sensitive to the fractional charges carried by the quarks. Therefore, if an equilibrated partonic
434
S. Jeon and V. Koch
phase has been reached in these collisions, the charge fluctuations per entropy would be about a factor of 2 to 3 smaller than in an hadronic scenario. In this review, we will systematically examine the principles and the practices of fluctuations such as the momentum and the charge fluctuations as applied to the heavy ion collisions. In doing so, various concepts of "average" need to be introduced. They are: (i) Thermal average (ii) Event-by-event average as a whole (iii) Averages over various parts of event-by-event average such as clusters and particles emitted by clusters. When necessary, we will use subscripts to distinguish these various averages. However whenever it is clear from the context, or if a relation is of a general nature such as the definition of a variance, we will simply use the symbol (...) to denote the average. The rest of this review is organized as follows. In section 2, we briefly review the basic concepts of thermal fluctuations and its relevance to heavy ion physics. In section 3, we examine possible sources of fluctuations other than the thermal ones and their effect on experimental measurements. In section 4, the relationships between the underlying correlation functions and charge fluctuations and also balance functions are established and their physical meaning made clear. In section 5, various ways of measuring fluctuations proposed so far in literature are briefly reviewed and their inter-relationships established. In section 6, past and current experimental results are examined in the light of the preceding discussions. We conclude in section 7. We regret that due to the lack of space and also expertise on the authors' side, such interesting topics as Disoriented Chiral Condensate or HanburyBrown-Twiss effects couldn't be discussed in detail in this review. Also for the same reason, it was impossible for us to cover all the relevant references although we tried our best. If there is any such glaring omission, we apologies to the authors.
Event by Event Fluctuations
435
2. Fluctuations in a thermal system To a good approximation the system produced in a high energy heavy ion collision can be considered to be close to thermal equilibrium. Therefore let us first review the properties of fluctuations in a thermal system. Most of this can be found in standard textbooks on statistical physics such as e.g. Ref.44 and we will only present the essential points here. Typically one considers a thermal system in the grand-canonical ensemble. This is the most relevant description for heavy ion collisions since usually only a part of the system - typically around mid-rapidity - is considered. Thus energy and conserved quantum numbers may be exchanged with the rest of the system, which serves as a heat-bath. There are, however, important exceptions when the number of conserved quanta is small. In this case an explicit treatment of these conserved charges is required, leading to a canonical description of the system 20 and to significant modifications of the fluctuations, as we shall discuss below. In the following, we first discuss fluctuations based on a grand canonical ensemble and then later point out the differences to a canonical treatment. 2.1. Fluctuations
in a grand canonical
ensemble
Assuming we are dealing with a system with one conserved quantum number (such as the electric charge, baryon number etc.) the grand canonical partition function is given by a Z = Y,
(* | e x P ( ~ / ? ( ^ - /xQ)) i) = Tr[exp(-/?(ff - /iQ))]
(1)
states i
where /3 = 1/T represents the inverse temperature of the system, Q is the conserved charge under consideration and fi is the corresponding chemical potential. Here the sum covers a complete set of (many particle) states. The relevant free energy F is related to the partition function via F=-TlogZ
(2)
We can further introduce the statistical operator PQ = \ exp [~P(H - nQ)\ = exp [-/?(# - nQ - F)] a
(3)
Here we restrict ourselves to one conserved charge. Of course the are several conserved quantum numbers for a heavy ion collision. The extension of the formalism to multiple conserved charges is straightforward.
436
S. Jeon and V. Koch
and the moments of the grand-canonical distribution: (Xn)G=Tr[XnpG}
(4)
For a thermal system, typical fluctuations are Gaussian 44 and are characterized by the variance defined by (6X2) = (X2) - {X)2
(5)
In the case of grand-canonical ensemble, fluctuations of quantities which characterize the thermal system, such as the energy or the conserved charges, can be expressed in terms of appropriate derivatives of the partition function. Of special interest in the context of heavy ion collisions are energy/temperature fluctuations, which are often related to the fluctuations of the transverse momentum as well as electric charge/baryon number fluctuations. 2.1.1. Fluctuations of the energy and of the conserved charges As already pointed out at the beginning of this section when analyzing the system created in a heavy ion collisions, one usually studies only a small subsystem around mid-rapidity. In a statistical framework, this situation is best represented by a grand canonical ensemble, where the exchange of conserved quantum number with the rest of the system is taken into account. The equilibrium state is then characterized by the appropriate conjugate variables, namely the temperature and the chemical potentials for the energy and the conserved quantities, respectively. As a consequence, energy as well as the conserved charges may fluctuate and the size of fluctuations reveals additional properties of the matter, the so called susceptibilities, which also characterize the response of the system to external forces. For example, the fluctuation of the conserved charge in the subsystem under consideration is given by W
)0=T^,o
g Z
=-r|lF
(6)
Similarly, fluctuations of the energy can be expressed in terms of derivatives of the partition function with respect to the temperature (^)a
= ^ ^ Z
= -
^
F
= T»Cy
(7)
Event by Event Fluctuations
437
Note that the energy fluctuations are proportional to the heat capacity of the system, CV- Thus one would expect that these fluctuations obtain a maximum as the system moves through the QCD phase-transition, which, among others, is characterized by a maximum of the heat capacity. An alternative way 57 ' 53 is to study temperature fluctuations in microcanonical ensemble which are inversely proportional to the heat capacity 44
(
f
f
2
)
M
C
=
%
•
( « )
However, for situation at hand, which is best described by a grand canonical ensemble, energy fluctuations appear to be the more appropriate observable. The second derivatives of the free energy, which characterize the fluctuations, are usually referred to as susceptibilities. Thus we have the charge susceptibility 1 d2
XQ =
~vwF
(9)
and the "energy susceptibility" XB =
-v&r*F = T
(10)
which is related to the specific heat. And just as the specific heat determines the response of the (sub)system to a change of temperature the charge susceptibility characterizes the response to a change of chemical potential. In case of electric charge, this would be the response to an external electric field. In thermal field theory these susceptibilities are given by correlation functions of the appropriate operators. For example the charge susceptibility is given by the space-like limit of the static time-time component of the electromagnetic current-current correlator 41.43>48>22 XQ = - n o o ( w = 0 , g - > 0 )
(11)
with I W * ) = i [ d*xexp(-ikx)
{pT*{ju(x)ju(0)))
(12)
It is interesting to note, that the charge susceptibility is directly proportional to the electric mass 4 1 , mlx = e2XQ
(13)
438 S. Jeon and V. Koch
Equation Eq. (11) allows to calculate the electric mass in any given model (see e.g. 23,22,48,14,13) a n d i n partidiaj. i n Lattice QCD 30>26. Since dilepton and photon production rates are given in terms of the imaginary part of the same current-current correlation function - taken at different values of (j and q - model calculations for these processes will also give predictions for the charge susceptibility, which then can be compared with lattice QCD results. As shown in 48 an extraction of \Q from Dilepton data via dispersion relations, however is not possible. For that one needs also information for the space-like part of IIoo which is not easily accessible by heavy ion experiments. Charge fluctuations are of particular interest to heavy ion collisions, since they provide a signature for the existence of a de-confined Quark Gluon Plasma phase 6 ' 35 . Let us, therefore, discuss charge fluctuations in more detail. Consider a classical ideal gas of positively and negatively charged particles of charge ±q. The fluctuations of the total charge contained in a subsystem of N particles is then given by (5Q2) =
t?((5N+-6N-)>) 2
= q [(5Nl) + (SNi)
- 2 (5NSN+)}
.
(14)
Since correlations are absent in an ideal gas, (5N-5N+) = 0. Furthermore, for a classical ideal gas, (6N2)UB
= WMB
(15)
where the subscript MB stands for Maxwell-Boltzmann. This implies (SQ2)MB
= I'
(N+
+ ^)MB
= Q2 (^ch)MB
(16)
where iVch = N++N- denotes the total number of charged particles. Taking quantum statistics into account modifies the results somewhat, since the number fluctuations are not Poisson anymore (see e.g. 52 ) (5N2)G
= (N)G (1 ±
(17)
fd3pn±(p)2 VJ:T,!, fd?pn±(p)
(is)
where __
<«±>G =
Here, (+) refers to Bosons and ( - ) to Fermions, and n±(p) represents the respective single particle distribution functions. For the temperatures and
Event by Event Fluctuations
439
densities reached in heavy ion collisions, however, the corrections due to quantum statistics are small. For a pion gas at temperature T = 170 MeV un = 1.13 10 . Obviously, charge fluctuations are sensitive to the square of the charges of the particles in the gas. This can be utilized to distinguish a Quark Gluon Plasma, which contains particles of fractional charge, from a hadron gas where the particles carry unit charge. Charge fluctuations per particle should be smaller in a Quark Gluon Plasma than in a hadron gas. The appropriate observable to study is then the charge fluctuations per entropy. To illustrate this point, let us consider a non-interacting pion gas and Quark-Gluon gas in the classical approximation. Corrections due to quantum statistics and due to the presence of resonances are discussed in detail in 6 > 35 ' 34 . In a neutral pion gas the charge fluctuations are due to the charged pions, which are equally abundant W>
M B
(19)
= Wr+>MB + Wr->MB
whereas in a Quark Gluon Plasma the quarks and anti-quarks are responsible for the charge fluctuations W M B
= G« W M B + Ql W)MB = I
(^)MB
(20)
where Nq = Nu = Nd denotes the number of quarks and anti-quarks. For a classical ideal gas of massless particles the entropy is given by 5MB=4(AT)MB
(21)
For a pion gas, this yields S„ = 4({NW+)MB
+ (N„-)UB
+ (Nno)MB)
(22)
and for a Quark Gluon Plasma SQGP = 4 ((JVU)MB + (Nd)MB = 4(2
+
+
<jVg)MB)
(Ng)UB) (23)
where Ng denotes the number of gluons. Therefore, the ratio of charge fluctuation per entropy in a pion gas is
Ml™ = i
(24)
440
S. Jeon and V. Koch
whereas for a 2-flavor Quark Gluon Plasma it is
M>MB_ 5QGP
24
(25)
Consequently, the charge fluctuations per degree of freedom in a Quark Gluon Plasma are a factor of four smaller than that in a pion gas. Hadronic resonances, which constitute a considerable fraction of a hadron gas reduce the result for the pion gas by about 30 % 35 ' 34 , leaving still a factor 3 signal for the existence of the Quark Gluon Plasma. The above ratio, (SQ2) /S can also be calculated using lattice QCD 3 0 . Above the critical temperature, the value obtained from lattice agree rather well with that obtained in our simple Quark Gluon Plasma model here 3 5 . More recent calculations for the charge susceptibility 26 give a somewhat smaller value, which would make the observable even more suitable. Unfortunately, at present lattice results are not available for this ratio below the critical temperature. Here one has to resort to hadronic model calculations. This has been done in 22 ' 23 using either a virial expansion, a chiral low energy expansion or an explicit diagrammatic calculation. In all cases, the ratio is slightly increased due to interactions, thus enhancing the signal for the Quark Gluon Plasma.
2.2. Fluctuations
in a canonical
ensemble
As pointed out in the beginning of this section, once the number of conserved quanta is small, i.e. of the order of one per event, the grand canonical treatment, where charges are conserved only on the average, is not adequate anymore. Instead the description needs to ensure that the quantum number is conserved explicitly in each event. Since the deposited energy is still large and is distributed over many degrees of freedom, the canonical ensemble is the ensemble of choice. Obviously the fluctuations of the energy is identical to the grand canonical ensemble, but fluctuations of particles which carry the conserved charge are affected. As an example, let us consider the fluctuations of Kaons in low energy heavy ion collisions. At 1-2 AGeV bombarding energy, only very few kaons are being produced (NK) ~ 0.1 58 , which makes an explicit treatment of strangeness conservation necessary. For simplicity, let us consider Kaons
Event by Event Fluctuations
441
and Lambdas/Sigmas b as the only particles carrying strangeness. In the canonical ensemble, where strangeness is conserved explicitly, the partition function is given by 20 Zc = Z?estJ2±(Z°KZl)n
(26)
n=0
where Z° est is the standard (grand canonical) partition for all the other non strange particles and Z^ , Z\ are the single particle partition functions for Kaons and Lambdas, respectively. In the classical limit, these are given by
ZK
° = 2V J (S^ ex P(-^*) = « ) M B
(27)
ZA
(28)
° = WI
ex
( ^
P ( - ^ ( ^ - MB)) - (K)MB
Here, the degeneracy factor of d = 2,4 for Kaons and Lambdas, respectively, take into account the presence of the K° and £ particles. Note that Z° is simply the number of particles in the grand canonical ensemble in the limit of vanishing strange chemical potential. Given the above partition function, the probability Pn to find n "Kaons" is given by 42>36>45 e" Pn
=
(29)
/0(2^)(n!)2
where Jo is the modified Bessel function and e = Z%Zl
= « )
M B
(ATO) M B
(30)
Given the probabilities Eq. (29), one can easily calculate the canonical ensemble average e2
rIi(2y/e)
(N2K)C
=
e3
e
(31)
so that the second factorial moment br
2
b
c
=
=
(N(N ~ l))c 2 7T772 {N) C
=
I + £+
=
I .
- O2 + 6 C + '-" '- - 02 +
(^)c 6e
+ ••'
(32)
I n the following we will denote both Lambdas and Sigmas as Lambdas, but include the appropriate degeneracy factor to take the sigmas into account
442
S. Jeon and V. Koch
This is to be contrasted with the grand canonical result, which follows in the limit of e 3> 1. In this case, (N2)G
- (N)G = (N)2G
(33)
F2G = 1
(34)
so that
Thus for (JV)
Transitions
and
Fluctuations
As already mentioned in the introduction, the QCD phase diagram is expected to be rich in structure. Besides the well known and much studied transition at zero chemical potential, which is most likely a cross over transition, a true first order transition is expected at finite quark number chemical potential. It has been argued 55 that the phase separation line ranges from zero temperature and large chemical potential to finite temperature and smaller chemical potential where it ends in a critical end-point (E) (see Fig.l). Here the transition is of second order. There is also a first attempt to determine the position of the critical point in Lattice QCD 25 . As discussed in 55 , the associated massless mode carries the quantum numbers of cr-meson, i.e. scalar iso-scalar, whereas the pions remain massive due to explicit chiral symmetry breaking as a result of finite current quark masses. Fluctuations are a well known phenomenon in the context of phase transitions. In particular, second order phase transitions are accompanied by fluctuations of the order parameter at all length scales, leading to phenomena such as critical opalescence 4 4 . However, since the system generated in a heavy ion collision expands rather rapidly, critical slowing down, another phenomenon associated with a second order phase transition, will prevent the long wavelength modes to fully develop. In Ref.9 these competing effects have been estimated and authors arrive at a maximum correlation length of about £ ~ 3 fm if the system passes through the critical (end) point of the QCD phase diagram.
Event by Event Fluctuations
443
In Ref.56 the authors argue that if the system freezes out close to the critical end-point, the long range correlations introduced by the massless <j-modes lead to large fluctuations in the pion number at small transverse momenta. In the thermodynamic limit, this fluctuations would diverge, but in a realistic scenario, where the long wavelength modes do not have time to fully develop, the fluctuations are limited by the correlation length. In Ref.9 it is estimated that a correlation length of £ ~ 3 fm will result in ~ 5 — 10% increase in fluctuations of the mean transverse momentum, which should be observable with present day large acceptance detectors such as STAR and NA49. Since the precise position of the critical point is not well known, what is needed is a measurement of the excitation function of these fluctuations. If the system undergoes a first order phase transition, bubble formation may occur. Since each bubble is expected to decay in many particles this leads to large multiplicity fluctuations in a given rapidity interval 8 ' 3 2 . Fluctuations of particle ratios, on the other hand, should be reduced due to the correlations induced by bubble formation 34 .
444
S. Jeon and V. Koch
3. Other
fluctuations
In heavy ion collisions, there are many sources of non-statistical fluctuations. To extract physically interesting information from the observed fluctuations, it is crucial that we know the sizes of these other fluctuations. In this section, we discuss the geometrical volume fluctuations and fluctuations coming purely from the initial reactions without further rescatterings within the context of wounded nucleon model 8 . 3.1. Volume
fluctuations
In heavy ion collisions, we have no real control over the impact parameter of each collision. This implies that geometric fluctuation is unavoidable in the fluctuations of any extensive quantities such as the particle multiplicity and energy. Furthermore, since the system created by a heavy ion collision is finite, one must also consider the thermal fluctuation of the reaction volume 44 . To be more specific, consider the multiplicity TV. Writing N = pV
(35)
where p is the density and V is the volume, the fluctuation of N can be written as ( ^ 2 ) e b e = < * A b e W e b e +
(36)
where the subscript 'ebe' indicates the event-by-event average measured in an experiment. Normally, what we are interested in are the fluctuations in the thermodynamic limit. In other words, we are only interested in the fluctuation of the density, {Sp2}. The second term containing (6V2) is therefore undesirable. To make a rough estimate of the volume fluctuation effect, we first decompose (^2)ebe = ( * n
h
+ (^2)geom
(37)
Assume the ideal gas law PV = NT and using
<*"*>«. = - * ( f ? ) T
<»>
(see 44 ) the thermodynamic volume fluctuation can be estimated as (W2)th
= (V)2/(N)th
= (V)/(p)th
(39)
Event by Event Fluctuations 445
or (p)l(6V)l
= (N)th
(40)
which is at least as big as the fluctuation due to the thermal density fluctuation (c.f. Eq.(17)). On top of this, we have purely geometrical fluctuation due to the impact parameter variation. To have an estimate, let us simplify the nucleus as a cylinder with a radius R and the length L. Assuming geometrically random choice of impact parameters, one can show that \
/geom _ 2 o m a x
ly? \
9^&
lOOmax +
27^R?
, ^^4
/p4\
+ 0(bmaJR
)
/41 \
(41)
'geom
if the impact parameter varies between 6 = 0 and b = 6 max - The expansion is of course valid when 6 m a x -C R. As an example, consider 6 % most central collision of two gold ions. For a gold ion, R w 7 fm and 6 % corresponds to 6 max « 3.0 fm. Plugging in these values to the above formula yields, (^2)geom«Wgeom/170
(42)
(43)
or
For (N) > 170 then, the observed multiplicity fluctuation can be overwhelmingly due to the impact parameter variation. This difficulty can be circumvented if the average of the fluctuating quantity, say the electric charge Q, is zero or close to it. This can be readily seen if one writes 8Q = Sq(V)ebe
+ (q)ehe5V
(44)
ignoring terms quadratic in 5q and SV. Squaring and averaging yield W)ebe = W e b e W
+ ^ O e b e <«>*.
(45)
assuming that the fluctuations in the charge density q and the volume V are independent. Usually, (Sq2) ~ (N) / {V) and from the above impact parameter consideration, {SV2) = (V) /C where C = 0(100). Therefore, the second term is negligible if ( ^ 2 ) e b e (?>ebe = «?>ebe /C «
(iV) e b e
(46)
446
S. Jeon and V. Koch
or
(Q)L/(N)ehe«C
(47)
The STAR detector at RHIC counts about 600 charged particles per unit rapidity in central collisions and (Q) e b e / {N)ehe « 0.1. In that case, (Q)ebe/Wabe«(0-l)2x600 = 6
(48)
which is much smaller than C = 0(100). If one is to talk about gross features, such error would not matter much. However, if a precision measurement is required, this has to be taken into account. 3.2. Fluctuation
from initial
collisions
One of the important issues in heavy ion collisions is that of thermalization. Seemingly 'thermal' behaviors of particle spectra are very common in particle/nucleus collision experiments including e+e~~ collisions and pp, pp collisions where one would not expect thermalization among secondary particles to occur. This seemingly 'thermal' feature results from the way an elementary collision distributes the collision energy among its secondary particles. Mathematically speaking, this is analogous to passing from the micro-canonical ensemble of free particles to the canonical ensemble of free particles. By doing so, one incurs an error. However as the size of the system grows the error becomes negligible. A way to distinguish simple equi-partition of energy from true thermalization through multiple scatterings is to consider not only the average values but also the fluctuations. For instance, as explained in section 2, the fluctuations of charged multiplicity in thermal system is Poisson or « )
M B
= <^h> M B
(49)
However, we do know that in elementary collisions such as pp or pp, the fluctuation is much stronger due to KNO scaling « >
e b e
c< (7Vch)e2be
(50)
Questions then arise: How would (5N^h) . behave if there is no scatterings among the secondaries? Will it resemble the thermal case or retain the features of KNO scaling? If it turned out to resemble thermal case, how can we distinguish?
Event by Event Fluctuations
447
To answer these questions, let us invoke the wounded nucleon model as explained in Ref.8. In this model, the final charged particle multiplicity is given by N„
Wch = ^ > i
(51)
j=i
where Np is the number of participating nucleons (participants) in the given event and n* is the number of charged particles from each of the participants. The assumption here is that the production of charged particles from each nucleon is independent and the produced particles do not interact further. In that case, it is not hard to show (iV ch ) ebe = (AT p ) ebe (n) N
(52)
and (^c h ) e b e - {Np)ehe
(n)N +
(n) N
(53)
Here (...) N denotes that this average is taken with respect to a single nucleon. It is independent of the system size. Therefore the second term is where KNO makes its appearance. It states that that ( J n 2 ) N = CKNO (n)-^ with CKNO ~ 0.35 for proton-proton collisions (for instance, see Ref. 33 ). In Ref.7, it is argued that since nucleons inside a nucleus are tightly correlated, the probability to have all nucleons in the reaction volume participate is very high. Therefore (SN^) / (Np)ebe should be small and i^^ch) I (-Wcii) = 0((n)N) which grows with the collision energy. To compare with a typical heavy ion experiment, however, two more effects need to be considered in addition. First, although the fluctuation of the number of participants due to the nuclear wavefunction is small, geometrical fluctuations due to the variation of impact parameter can introduce more substantial fluctuations in Np. Second, a heavy ion experiment usually can observe only a portion of the whole phase space. Therefore when calculating (n) and (n 2 ) we must fold in a binary distribution with p = (n) A / (n) full where Ar] is the detector window. This implies replacing {n)n^p{n)N, 2
(54) 2
2
(«5n )N -> (1 - p)p (n) N + p <«5n )N and
(55)
448 S. Jeon and V. Koch
(^P)ebe -> W ) g e o m (56) The fluctuation due to the impact parameter variation was estimated in the previous section. Putting everything together then yields WWNM =
^ ^
= (Nch)geom
( ^ 2 ^ 2 J + (1 - P ) +pc K NO (n) N
(57) Here we assumed that tight correlations renders intrinsic participant fluctuation very small and also used the fact that (NCh) m = P (^p)fie0m ( n )NA few conclusion can be drawn from Eq.(57). First, if p is sufficiently small or if the observational window is sufficiently small compared to the whole phase space, the fluctuation of the number of charged particle is Poisson. However, this has nothing to do with dynamics. Second, if p is sufficiently large, then WWNM becomes significantly larger than 1 since s CKNO (™) N i about 7 at RHIC energy. Third, the ratio W\VNM depends linearly on p. The first point is purely statistical in nature. The second and third points can be used to test the validity of this model where no secondary scattering occurs. What would change in this consideration if thermalization through multiple scatterings occurs? Suppose that the total number of produced particles is still governed by Eq.(51). In other words, only elastic scatterings happened to the secondary particles. Further suppose (<5n2)N is still given by the KNO scaling result. In this case, the fluctuations of iVCh in the full phase space must remain the same as before and Eq.(53) remains valid. Since binary process does not really care about clusters in this case, Eq.(57) also stays valid. Therefore elastic scatterings alone cannot make any difference in multiplicity fluctuations. Now let us relax the condition and allow inelastic collisions among the secondaries. The essential role of inelastic collisions is to convert energy into multiplicity and vice versa. Consider a set of events with the same number of participants and same amount of energy deposit. According to Eq.(53), the number of initially produced particles (we will call them 'initial particles') have a distribution that is much wider than the Poisson distribution if (n) N ^> 1. This wide distribution implies that the distribution of energy per initial particle also has a wide distribution. As argued above, elastic collisions cannot change this situation. If inelastic collisions are allowed, then an event with smaller number of initial particles would tend
Event by Event Fluctuations
449
to produce more particles since collisions in this event have more available energy. In this way, energy is re-distributed evenly and the the multiplicity distribution gets narrower. This is, of course, the process of equi-partition which lies at the heart of thermalization. Therefore if thermalization does occur starting from a certain energy or a certain centrality, <*JNch must show a corresponding behavior changing from approximately (n) N > 1 to lower values close to 1 as the energy goes higher or collisions get more central.
450
S. Jeon and V. Koch
4. Fluctuations and Correlations Fluctuations measure the width of the two particle densities 12 and therefore provides additional information than just the averages. To illustrate this point, let us consider a variable x(p) which depends on momentum of one particle but does not depend on the multiplicity in each event. Let us further consider a generic observable in a given event N
N
S(z) = 5 > ( P i ) E E £ > ( ; ) i=l
(58)
i=l
where N is the multiplicity of the event. Since x(p) does not depend on N, S is an extensive quantity. The event averaged moments of this quantity can be expressed as 1
M
Nm
Nm
(^ebe = M E E - E *»0l)-*m(i*) 771=1 1 1 = 1
(59)
«fc = l
where m labels the different events and M is their total number. Nm is the multiplicity of the event labeled by m. On the other hand, the moments of an extensive quantity x(p) calculated from n-particle inclusive distribution pn{pi, - , p n ) are defined as 1l-dpnPn(Pl,
-,Pn)[x(pi)}
' ...[x(pn)]kn
=
/ • M M
1
M M
N Nmm
N„ Nm
E E - Eb-fa)]* 1 -^'")]*" m=l ii=l
(eo)
in = l
where the sums over i\...in include only the terms for which all indices i\...in are different from each other. One sees immediately that (59) and (60) are related. <5>ebe = j dpPl{p)x{p) (S2)ebe
~
(S3)ebe
I dPldP2P2(Pl,P2)
=
/
x{pl)x(p2)
(61) + I dpPl{p)
d
[x{p)f
PldP2dP3P3(Pl,P2,P3)x(px)x{p2)x(p3)
+ 3 / dp1dp2p2(pi,P2)
x(pi)[x(p2)]2
(62)
Event by Event Fluctuations
+ JdPPl(p)[x(p)}3
451
(63)
Similar formulas can be derived for higher moments of S. By setting x(p) = 1, one can also find that the integral over pn gives the n-th factorial moments j dp1...dpnPn{p1,
...,Pn) = (N(N -i)...(N-n
+ l))ebe
(64)
We have thus established the relation between inclusive measurements and event-by-event fluctuations for single particle observables 27 , such as e.g. the total transverse momentum, particle abundances 28 etc. For observables concerning different species of particles, we define N
Sa{x) = ^2x{Pi)5ai
(65)
i=l
where the ^-function picks out the right species from N particles in the event. The event average (5 Q ) e b e is just as before with the replacement of Pi(p) —>• pa(p) and /dpp a (p) = (Na)ehe. The average number of pairs is then given by (Sa(x)Si3{x))ehe
= — ] T Y^ x(Pi)x(Pj) i=l
fyjSoci
(66)
j=l
In terms of the correlation function, this can be rewritten as (Sa{x)Sp(x))ebe
= / dpadpp pa0(pa,pp)x(pa)x(pp)
(67)
Again by setting x(p) = 1, one also obtains dpadp0 pa0(pa,pp)
= (NaNi3)ehe
(68)
Similar arguments can be constructed for variables which depend on two or more particle momenta, such as e.g. the fluctuations of Hanburry-Brown Twiss (HBT) two particle correlations, belong to this class. They are of practical interest and will be investigated in heavy ion experiments. Here we will restrict the argument to two particles but it can be readily extended to multiparticle correlations. The details are given in 12 . To summarize, E-by-E fluctuations of any (multiparticle) observable can be re-expressed in terms of inclusive multiparticle distribution. In case of Gaussian fluctuations the multiparticle distributions need to be known up to twice the order of the observable under consideration.
452
S. Jeon and V. Koch
Following these straightforward considerations, it is natural to discuss fluctuations in terms of two particle densities, or equivalently in terms of two particle correlations functions 37,49 Cap(Pl,P2) = Pap(Pl,P2) ~ Pcc{P2)P(}{Pl)
(69)
Here the labels a and ft denote general particle properties including different quantum numbers. Thus correlations between different particle types can also be discussed in the same framework. Instead of the correlation function C often the reduced correlation function , n v _ CQ0(pi,p2) _ pap(pi,P2) . p f7n. Raf3(Pl,P2)
=
-,—v / Pa{pi)p0{P2)
T=
/—v ;—r - 1 Pa{Pl)P0(P2)
(70)
is used. The advantage of using Ra0 over Cap is that the trivial scaling with the square of the number of particles is removed and, therefore, the true strength of the correlations is exposed more transparently. In particular, as shown in Ref.49, the observable Rap has the advantage that it is robust, i.e. it is independent of the detector efficiencies to leading order. Also, the analysis of elementary reactions such as proton-proton has traditionally been done based on the reduced correlation function Rap 61 (see also section 6.1). In some cases, such as the charge fluctuations (c.f. 2.1.1), the fluctuations of the number of particles in a given region of momentum space is considered. In this case one deals with the integrated quantities W*>A,
=
/
d
P«Pl(Pa)
(71)
dPadPl3p2(pa,p0)
(72)
J AT)
(NaN0)Arj
- 6ap (iVa)A„ = / J Art
where the notation (...)» is always to be understood as the event-by-event average in a given momentum space region A77
«=i
where 6(p £ AT?) = 1 if p falls inside A77 and zero otherwise. The correlations are then expressed in terms of the 'robust covariances' 49 _ (NaN0)Ar) - (Na)Ar]5a0 - (Na)Ar) (N0)Ar) s
Event by Event Fluctuations
fAridpidp2pa0(pi,P2)
453
i
/ A r ? dpipa(pi) JAr) dp2pp{p2) which shares the same virtue as Rap4.1. Correlations
and charge
fluctuations
We will now concentrate on charge fluctuations to illustrate the above general considerations. Here we will mainly work out the relationship between the charge fluctuations and the charged particle correlation functions. In full phase space, a conserved charge does not fluctuate, or (5Q2) = 0. What we are interested in is, however, fluctuations of the net charge in a small phase space window where the effect of this overall charge conservation is small. At the same time, this window should be big enough in order not to lose information on the widths of the correlation functions. Prom previous section 2.1.1, we know that if a QGP forms, the charge fluctuation per entropy becomes a factor of 2 to 3 smaller than that of the hadronic gas. What we are interested in this section is how that is related to the properties of the charge correlation functions. Since the net charge is Q = N+ — 7V_, the variance is = (SNl)An + < ^ - ) A , - 2 (M+ SN-)Ar, (75) Written this way, it is clear that to have a small charge fluctuation, we must have a positive correlation between the unlike-sign pairs and it must have enough strength to compensate the independent fluctuations of N±. The purpose of this section is to elaborate on these points. To write the charge fluctuations using the correlation functions, we first decompose W A ,
WA„
= W>A„ " <*+>!, + (N-)A, - (N-)ln - 2 [(N+N-)Av
- (N+)Ar] (iV_) A J
(76)
Using Eqs.(69) and (72), we can rewrite this as (SQ2)Av
= (N+)ATI + +
(N.)Ari
dpadp0C++{pa,Pi3)+ ./Arj
- 2 / JAn
dpadp0C—(pa,Pp) J Arj
dpa dpp C+ -(pa,pp)
(77)
454
S. Jeon
and
V.
Koch
The first two terms in the right hand side comes from the fact that integrating over like-particle correlations give (N(N - 1)) A „. If (N+)A « (N-)Av, this can be also rewritten in terms of the robust covariances
< ^ c h > Ari
~ 1H
;;
^dyn
(78)
where Pdyn = R++ + R— - 2R+-
(79)
and 7Vch = N+ + AL. This is the form advocated in Ref.49. In this paper, the authors argued that in Rap the detector efficiencies cancel out while (•Wch)A„ still depends on the efficiency which in modern detectors can introduce 10 ~ 20 % error. We know that the single particle distribution p±(j>) is proportional to the probability density function for single particle momentum. To give the two particle distributions similar meaning in terms of underlying correlations, let us consider a simple toy model. Consider a gas made up of only three species of resonances, one neutral, one positively charged and one negatively charged. No thermal pions are present. Furthermore let's further assume that the neutral resonance decays into a pair of 7r+7r~ and when a charged resonance decays it emits only one charged pions. Since our model does not contain thermal pions, all final state pions arise from resonance decay. We also assume that there is no correlation between the resonances themselves. The probability to have a given set of final state momentum {{pf}, {p^}} f° r charged pions is then given by the product of the 7r+ and 7r_ momentum distribution from each resonance decay:
V{{pt},{p-}\{Pi},M0)M+,M_) M0
M+
o=l
6=1
F
M_
= n Fo(pt,Pa\pa) n M\Pb) n F-^\PC)
<m
c=l
where {Pi} are the momenta of the resonances just before the decay and MQ, M+,M- are the number of the neutral and the charged resonances. The F's are normalized to 1. Since we are not so much interested in the distribution of resonance momenta, we integrate over it with a suitable weight (for instance thermal weight) before any other calculation. To simplify, we
Event by Event Fluctuations
455
make a further assumption that the resonance momenta are uncorrelated. In that case, c Mo
M+
M-
Hipth {pniM,, M+, M_)=n hbi,Pa) n Mpt) n f-fo) &) o=l
6=1
c=l
where MPT,P-)
f±(pf)
= jdPFo(pt,p-\P)Po(P)
(82)
= j dPF±(pf\P)P±(P)
(83)
Here Pi{P) is the momentum distribution for the resonance species i. The function V{{pf},{p^}|M0,M+,M_) contains full information about the momentum distribution of the final state charged pions given the number of underlying resonances. Prom this distribution all other correlation functions can be calculated. For instance, the single particle distribution for positively charged particles is M++M M + + M 00
P+(p)=
£
£ M0,M+M~
[dp] Y, S(p-Pi)V({pth{P7}\Mo,M+,M.) "
«=1
xV(M0,M+,M-) = <Mo>h+(p) + ( M + > / + ( P )
(84)
where [dp] is a short hand for integration over all p, and we defined h±{p±) =
dihff0{p+,p-)
(85)
Here V(Mo, M+, M_) is the probability to have the number configuration (Mo, M+, M_) in the whole event set and {M±to) is the event average of the resonance multiplicities in the full momentum space. Eq.(84) thus simply states that the number of positively charged particles is given by the number of resonances which decay in positively charged pions time the probability to for these pions to have the decay momentum p. c
To be quantum mechanically correct, we need to sum over all permutations of {p"*"} and { p - } and divide by (M+ + Mo)!(M_ + Mo)!. However, since this does not change the outcome of our consideration, we will omit that here.
456
S. Jeon and V. Koch
Likewise, two particle distributions are C±±{pi,P2) = (SM06M±) /i±(pi)/±(p 2 ) + (5M06M±)
h±(p2)f±(Pl)
+ {5Ml) f±(Pl)f±(P2) + (SMS) h±{Pl)h±(p2) - <M±> f±(pi)f±(p2)
- (M0)
h±(Pl)h±(P2) (86)
and C+_(pi,p2) = (SM06M-)h+(Pl)f_(p2)
+
(dM0SM+)h.(p2)f+(Pl)
+ (6M+8M-) / + ( P l ) / _ ( p 2 ) + (5M*) - (M0) h+(Pl)h_(p2)
h+(Pl)h_(P2)
+ (Mo) fo(pi,p2)
(87)
To simplify our consideration, let us regard the charged resonances to be iso-spin partners so that we have f+(p) = f-(p)
= f(p)
(88)
For the neutral resonances, f0 should satisfy /o(pi,P2) = /o(P2,Pi) which leads to h+(p) - M p ) = h(p)
(89)
The average multiplicities are then (N±)Av
= (M+) [
dpf(p) + (M0)f
dph(p)
(90)
J A?j
J At]
and the charge fluctuation is (SQ2)Av
=
(Nch)An + [((6M+ - SM-)2)
- (M+) - <M_>] ( f
dpf{p)
\JAn
-2 (M0)
dpadppfo(pa,p0)
J
(91)
J AT;
We can now consider two situations. First, consider the case where the underlying system is a thermal gas of free resonances in the grand canonical ensemble, where charge is only conserved on the average. In this case (8M?)UB = (Mt)MB and (5M+SM-)MB = 0 (using Boltzmann statistics for simplicity) and the above reduces to (6Q2)A°tm-
= (Nch)Ar)
- 2 (Mo) / J Arj
dpt dp2 / 0 (pi,p 2 )
(92)
Event by Event Fluctuations
457
If (SQ2)Aeim' is to be substantially different from (iVch)Ar;, we need to have (M 0 ) ~ (M±) and / A r ) dpi dp2 fo(Pi,P2) ~ JAlJdpf(p). In other words, the number of neutral resonances have to be large and the correlation sharp. In the full phase space, Eq.(92) leads to
w)Tjr (iVchC
rmal
(M+)+(M_)
=
(M+) + (M_) + 2(M0)
V ;
which corresponds to the result obtained in Ref.34 for the thermal resonance gas. To see how the finite phase space result (92) differs from the result (93), consider the extreme case of a flat distribution in the pair rapidity and a delta function in the relative rapidity f(y)
= % m a x + y ) % m a x ~ y)
,g^
^2/max
/o(2/i,y 2 ) = S(yi -y2)f((yi
+ya)/2)
(95)
Here we specified our momentum space variable to be the rapidity y and 2/max is the rapidity of the beam particles in the center of mass frame. In this extreme case, it is easy to see that (Nch)%;™- = {M+ + M_ + 2M 0 ) p
(96)
(6Q2)t£m=(M+ + M-)p
(97)
and
where p = Arj/2y max . Therefore in this case of infinitely sharp correlation, 2\therm.
/ » r \therm. (-Nch)therm. A7?
/r/n2\
th
erm.
/ » r \then /( j^. rc h \) therm. full
(98) v
'
for any values of ATJ. With more realistic / and /o, the relation (98) does not hold strictly. However as long as Ar? >
(99)
where OVT^ is the width of / 0 in the yrei = yi — yi direction and the single particle rapidity distribution is relatively flat within Arj, Eq.(98) should hold approximately as long as p is not too close to 1. In this way, one can say that the charge fluctuations per charged degree of freedom measured
458
S. Jeon and V. Koch
in a restricted rapidity window is a good approximation of the full thermal result. Prom proton-proton collision experiments, we can estimate ayrcl « 1
(100)
Therefore, our rapidity window must be at least that big. As noted previously the above results are for the grand-canonical ensemble. In real-life heavy ion collisions, the concept of grand-canonical ensemble cannot be applied to the full phase space since overall charge conservation strictly requires (SQ2) = 0. However, there is no dynamical information in this fact. In particular it has nothing to say about whether thermal equilibrium has been established within the system. To say something about the thermal equilibrium, we need to carve out a small enough sub-system and then use the concept of grand-canonical ensemble on it. It would be ideal if we could define a fluid cell within the evolving fireball that resulted from the collision of two heavy ions. Unfortunately, this is impossible for we have no information on the positions of the particles, produced or otherwise. Best we can do is to have (pseudo)rapidity slices which are supposed to be tightly correlated with the spatial coordinates in the beam direction. As shown above, if the system is grandcanonical, then this is not a big problem. However, since our underlying system is clearly not, we have to know what the effect of having a finite rapidity window is and what the measurements in that restricted window really signify. In our opinion, what one should try to get is the right handside of Eq.(93) which we know is the right grand-canonical limit for the resonance gas. To take the finite size effect and charge conservation effect into account, we make an restriction 6M+ = 5M_ or M+ - M_ = Qc is constant. The expression (91) now reduces to (SQ2)Ari
= {Nch)Ar)
- «M + > + (M_)) ( ^
- 2 (Mo) f
dp+dp-f0(p+,p-)
dpf(pfj (101)
JAri
which differs from the thermal result (92) by the second term. In the full phase space, this expression results in zero as it should. That also means that we need to find a way to extract the right-hand-side of Eq.(93) from the above expression. It will be ideal if we can just measure the second
Event by Event Fluctuations
459
term in Eq.(lOl) and subtract it. However, to do so we need to know (M±) which is not readily available. In the literature, two ways of compensating the charge conservation effect have been proposed. A multiplicative correction method was proposed in Ref.43 by the present authors and an additive correction method was advocated in Ref.49. To see how the multiplicative corrections work, let us assume for simplicity that / A / = / . h = p and rewrite Eq.(lOl) as
(SQ2)Ar) = (1 - P) (JVch)Ar, - 2 (Mo) Qf dP+ J dp- /o(p+,p-) - P 2 ) (102) The multiplicatively corrected charge fluctuation is given by ,o\ mult. A
"
(1-p)
(103) This is the origin of the so called "(1 — p) correction" 15 . With the extremely sharp correlation function (95) and the flat dN/dy (94), this yields
(*Q2)lT = {Nch)*» -2 {Mo)p ={M+ + M-)p
(104)
same as Eq.(97). In that case, muIt
-
mult.
(Nc^Tr,
/ r o 9 \ therm. / Tvr
\therm.
(^ch) full
^
'
holds for any At]. With a more realistic / and /o, Eq.(105) does not hold strictly. However, again as long the condition (99) holds and the single particle rapidity distribution is relatively flat within A77, Eq.(105) should hold approximately unless p is very close to 1. The authors of Ref.49 advocated an additive method. This method is best explained using the expression (78). Ref.49 proposes that this expression be corrected by replacing ^dyn -> ^dyn + 4 / (iVCh)full. The reason behind this correction is that - 4 / {Nch){uU is the value of Udyn when N+ — 7V_ is
460
S. Jeon and V. Koch
fixed and there is no correlation in the system. In our language, this correction corresponds to
= (iVch)A, - 2 (M0) (J dP+ J dp_ / 0 (p +> p_) - P2)(106) Again with the flat f(y) yields
(94) and the infinitely sharp /o (95), Eq.(106)
« _ w £ r .(*v\( (^ch) A „
h
(iVch)L ,r-
(Mo) \
V » » « ) \ ( M + + M- + 2M0) J
K
>
where we used p = At]/2ymsix. The second term is negligible only in the p —>• 0 limit. With a more realistic / and /b, this limit becomes 2/max
(108) which seems to be of more restrictive use for our purpose of extracting the right-hand-side of Eq.(93) (or the first term in Eq.(107)). In the full phase space limit,
{
^pSL =!
(•^eh/fun
(109)
independent of the specific choice for /o- Thus this method overcompensates the charge conservation effect slightly. One case where we do not need such corrections is when (M±) = 0 or (M±) <S (Mo). This is the case when all particles are produced in neutral clusters. Exactly this type of situation was analyzed in Ref.62. In this reference, however, "(1 — p)" correction 15 was made which actually overestimated the charge fluctuation. How then do we know the relative strength of (M±) and (M 0 )? From Eq.(lOl), it is clear that the unwanted second term scales like p2. The first term (JVch) of course scales like p. The last term in Eq.(lOl) however scales differently as the size of AJJ varies. If Arj is much smaller than the correlation length, this term varies like p 2 . However, if A77 exceeds the correlation length, then integration along p re i = p+ — p- ceases to change and hence it varies like p. This observation suggests the following method: Vary the observation window size A77 and plot (5Q2)^ as a function of
Event by Event Fluctuations
461
P — (-Nch)A,, / (•^ch)totai- According to Eq.(lOl), this plot can be described by a quadratic polynomial in p (5Q2)Av
= ap-bp2
(110)
If the single particle distribution is flat (as is the case for rapidity distribution), then p oc A77 and the parameterization (8Q2)A = a' A77 - b' (A77)2 can be used as well. The coefficients o and b are not truly constants but vary slowly as a function of p. For small p, a « W f u i i = (M+ + M_ + 2M 0 )
(111)
After A77 exceeds the correlation length in /o, the /o term becomes linear in p and the coefficient b becomes. b « (M+ + Af_)
(112)
Therefore by measuring the coefficients in these limits, one can estimate the relative strength of (M±) and (M0) and fully correct Eq.(lOl). At this point, one should ask how all this can be reconciled with the thermal fluctuations we considered in previous sections. The results are certainly similar. However, it seems that we have used no thermal features at all in deriving Eqs.(101-103). The expression (102) rather suggests that the these expressions result from binary process where a particle has a probability of p to end up in the detector. Of course, when N is large and p is small in such a way to have Np fixed, binary process becomes Poisson process but that does not mean that the underlying particle energies are thermally distributed. So how can the results of this section be reconciled with the QGP result from previous section? Recall that the arguments in section 2.1.1 that led to our conclusion of reduced charge fluctuation depended on two facts. One, the single particle distributions in particular the particle abundances are thermal. Two, the number fluctuations are all approximately Poisson. As Eq.(84) shows, the single particle distributions in this section do depend on the underlying momentum distribution. In particular, if one is to argue relationships such as Eq.(21) hold, one must have thermal momentum spectra. On the other hand, it doesn't really matter to the rest of the arguments whether the Poisson nature of the number fluctuation (Eqs. (19,20) is a result of having
462
S. Jeon and V. Koch
a thermal system or just a result of having a small detector window. Therefore, the fact that Eqs.(101, 102) are not the consequences of underlying thermal system does not negate our previous conclusion. Although the argument given above does reconcile the thermal consideration and the results from present section, it is still unsatisfactory in some aspects. The essential point here is our inability to have a sub-system in the sense of grand canonical ensemble. This is because we can't carve out a sub-system in the coordinate space to observe. Unfortunately this limitation is unavoidable. The best we can do is to define sub-systems in (pseudo-)rapidity space and argue that strong longitudinal flow strongly correlates the (pseudo-)rapidity and the coordinate in the beam direction. It would be desirable to work-out ensemble average taking into account this fact. That analysis, though, is still to be carried out. Another question to ask at this point is whether the underlying correlation fo(pi,P2) really corresponds to resonances or simply indicates that when the pions hadronize from a QGP most of the times they are pair produced. One way to distinguish the two scenario may be to actually measure the correlation length. Since the unstable neutral meson masses are all larger than twice the pion mass, thermal resonance gas must exhibit a certain characteristic momentum correlation length between the charged pions. For instance, if a p° at rest decays into two pions, they are 3.5 units of rapidity apart along the line of the decay. Pions directly coming out of hadronizing QGP on the other hand have no such strict constraint imposed on their correlation and should exhibit much sharper correlation. A lack of good hadronization scheme from a QGP makes this consideration difficult to substantiate. However, these points need to be further clarified. 4.2. Correlations
and Balance
Function
A balance function is a particular way of combining the correlation functions to articulate an aspect of the system. The balance function constructed by Pratt et.al. uses the number of like-sign pairs and unlike-sign pairs within a given phase space volume AT?: B(i?|Ai7) _ 1 (AT+_(r,|Ar?)) ~~ 2 (N.(Ar,))
+
(i\r_+(r?|Ar?)) (N+(Ar,))
(N++(r,\Ar,)) (N+(Ar,))
Event by Event Fluctuations
463
For instance, in this expression JV+_(77|A?y) is the number of unlike-sign pairs which are 77 apart from each other within the window AT;. TO relate (Nij) to to correlation functions, first we express the number of pairs within AT] and 77 apart as Nab(r,\Ar,) = £
m
E Arj)d(p) £ Aij)*(|p? - p)\ - r,)
(114)
ij
Integrating over the expression (81) give (Nab(ri\Ar))) =
dpxdp2S(\pi
- p2\ - r]) pab(p1,p2)
(115)
JAr,
With our model of resonance gas, the balance function can be expressed as 5(r7|Ar;) » — — — [2 (M 0 ) / (^ch)Af?
L
dPl dp2 5(\Pl - p2\ - r,) foipuPt)
J&V
+ «M+> + <M_» f
dPl dp2 6(\Pl -P2\-v)
/(Pi)/(P2)1 (116)
assuming that (Q)A„ <S (ATch)A Eq.(lOl) it is clear that
(Nc*)Av
From the expressions Eq.(116) and
1 - j dr) B{r)\Ar))
(117)
Balance function is originally devised to detect the change in the unlikesign correlation function. Just as the charge fluctuation, this is possible only if Mo ~ M± and /o is a sharply peaked function of 77. Pratt et.al. argued that if a QGP forms, the width of the balance function 5(r;|Ar7) will be reduced by a factor of 2 compared to the width of the resonance gas balance function. Recent measurement by STAR collaboration 24 indicates that going from peripheral collisions to the central collisions there is about 20 % reduction in the width of the balance function although it is yet not clear whether this is indeed the signal of QGP formation.
464
S. Jeon and V. Koch
5. Observable fluctuations Although the motivation for studying fluctuations is similar to that in solid state physics, the observables in heavy ion collisions are restricted to correlations in momenta and quantum numbers of the observed particles. Spatial correlation are only indirectly accessible via Fourier transforms of momentum space correlations, and thus limited. An example is the Bose Einstein correlations (see e.g. 3 1 ) . Consequently, the fluctuations accessible in heavy ion collisions are all combinations of (many) particle correlation functions in momentum space. In addition, as discussed in the previous section, even for tight centrality cuts there are fluctuations in the impact parameter which may mask the fluctuations of interest. In the thermal language, these impact parameter fluctuations correspond to volume fluctuations. Consequently, one should study so called intensive variables, i.e. variable which do not scale with the volume, such as temperature, energy density etc. Another issue is the presence of statistical fluctuations due to the finite number of particles observed. These need to be subtracted in order to access the dynamical fluctuations of the system. Finally, although this is outside the expertise of the authors, there are fluctuations induced by the measurement/detector, which also contribute to the signal. Those need to be understood and removed/subtracted as well. Let us return to the first two issues. As already discussed in some detail in section 4 the fluctuations can always be described in terms of the inclusive many-particle densities in momentum space pn(pi, • • • ,pn)- For the simplicity of the discussion let us concentrate on two particle correlations, which fully characterize fluctuations of Gaussian nature d . Let us start the discussion by defining the number of particles with quantum numbers a = a%,..., a n in the momentum space interval (p, p + dp) in a given evenf a _
d
"-"event
(118)
I n case of fluctuations of positive definite quantities, such as transverse momentum or energy,the appropriate distribution is a so called Gamma- distribution 5 9 e In this section, we will use the term 'event' and 'member of the given ensemble' interchangeably. We also use 'event average' and 'ensemble average' interchangeably.
Event by Event Fluctuations
465
and its fluctuations 6n% = n% - (nap)
(119)
In the case of ideal gas, (n£) takes the form of the Bose-Einstein distribution or the Fermi-Dirac distribution depending on the spin of the particles. The mean value of an observablef
x
= Y,x>p
( 12 °)
P,o>
is obtained by averaging over all the events in the ensemble
<x> = 5 > ? K >
(121)
Note that X defined in this way is an extensive observable. The goal of studying fluctuations is to see the effect of dynamics in terms of non-trivial correlation. If the system under study is totally devoid of correlation, then the single distribution function alone must be able to describe all the moments of X since higher order correlation functions would be just products of single particle distribution functions. Any deviation from this behavior is a sign of non-trivial correlation. To make a quantitative calculation, let us define the single particle probability density function
< " = jkpa{p)
(122)
where pa (p) is the single particle inclusive distribution function introduced in section 4. Note the factor 1/ (Na) in the definition of n £ m c which normalizes this distribution to unity. Note also that event average has already performed and the only variable left to average over is the single momentum. We define the average taken with this probability density as the 'inclusive average'.
JTC1 = 5 > ° ^ i n c l
(123)
For a single particle observable X
*incl = W)
(124)
'Here we use sums over momentum states, as appropriate for a finite box. The conversion to continuum states is a usual, J ^ —> V f F3 , where V is the volume of the system.
466
S. Jeon and V. Koch
For the discussion of fluctuations the relevant quantity to consider is A«f = (Sn^Snl)
(125)
where Sri" is defined in Eq.(119). For an ideal gas, only the identical particles are correlated and ((6rip)2) = up* (rip). Therefore,
^^n^K^'
( i26 )
Here o£ = 1
(127)
a £ = (1 ± « ) )
(128)
for a classical ideal gas, and
for a boson (+) and a fermion (—) gas. Before we continue, let us step back and try to understand eqs. (126) to (128) in simple terms. The term Yli fiaiPi ensures that only identical particles are correlated. For example, 7r+ and 7r~, are both pions but with different charge. Therefore they are not correlated in an ideal gas, i.e. {dTr+5ir~) = 0. Also there are no effect due to quantum statistics, as the particles are different. In the presence of dynamical correlations, the basic correlator (125) may contain off-diagonal elements in momentum space and/or in the space of the quantum numbers. For example the presence of a resonance, such as a po, correlates the number of TT+ and n~ in the final state due to resonance decay. Also their momenta are correlated. In addition, the occupation number np in a given momentum interval may be changed. This would be for instance an effect of hydrodynamical flow. Any two-particle observable can be expressed in terms of the basic correlator A°'^ 12,56 p o r t n e g e n e r j c observable X as given by eq. (120) the variance is given by
(5X2) = J2 KS,xv< p,q,a,0
(129)
Event by Event Fluctuations
467
To illustrate the formalism, let us consider the net charge 5
= J2
Q = N+-N-
( 13 °)
p,a
Here, qa is the charge of the particle whereas all other quantum numbers are still being summed over. Likewise the number of charged particle is given by p,a
The variances of these quantities are given by (SQ2) = (SNl) + (6N2_) - 2 {6N+6N-) = E(AM+
+ A
M"-2A?.i")
(132)
< ^ c h ) = (&N+) + (SNi) + 2 (5N+6N-) = £(
A
^
+
+
A
PT+
2 A P
V)
(133)
P,Q
Here the basic correlator A+;+ still contains a sum over all other quantum numbers, which we have suppressed. A+'+=
£
A<+->'{+^>
(134)
a j ^charge,/?; ^charge
and similar for the other combinations. This sum simply means that all charged particles are included independent of their flavor, spin etc. In the following, we will always use this short-hand notation with implicit summation over all the quantum numbers not specified. Obviously all the information revealed by the fluctuations is encoded in the basic correlator A ^ and the fluctuations of different observables expose different "moments" and elements of the basic correlator. Using the notation from the previous section, the basic correlator for a ^ f3 is nothing but the 2-particle correlation function defined in Eq.(69). For the identical particle correlator, Eq.(125) differs from Eq.(69) by a single g
Note the difference in notation: whereas Snp refers to the fluctuations in the momentum interval (p,p + dp), SN refers to the fluctuations of the total (integrated) number of particles.
468
S. Jeon and V. Koch
particle distribution so that A
P,f = Pap(P, q) - Pa{p)p0(q) + <W 6pq Pa(p) = Ca0 (P, q) + <W $pq Pa (?)
(135)
Therefore, an explicit measurement of the two-particle correlation function Cap would be extremely valuable, since all (Gaussian) fluctuation information can be extracted from it by properly weighted integrals. If the system has no genuine two particle correlations, then the basic correlator will be A £ f =
(136)
similar to the ideal gas, only that (n p ) does not have to follow a Boltzmann distribution. Using the inclusive single particle spectrum n^ mcl (122) as momentum distribution, the relation (136) may be utilized as an estimator of the fluctuations due to finite number statistics 27 , which follow Poisson statistics. A more detailed discussion on this aspect will be given below. Any correlations, on the other hand will lead to a non Poisson behavior of the basic correlator. The ideal quantum gases discussed above are examples where the symmetry of the wavefunction introduces 2-particle correlations which either reduce (Fermions) or enhance (Bosons) the fluctuations. Also global conservation laws such as charge, energy etc. will affect the fluctuations. However, if only a small subset of the system is discussed these constraints will be negligible11. As the fraction of accepted particles increases, these constraints may become more relevant, however, and thus need to be properly accounted for. This has been discussed in some detail in the previous section. As mentioned in the beginning of this section, in order to avoid contributions from volume / impact parameter fluctuations it is desirable to study so called intensive quantities, i.e. quantities which do not scale with the size of the system. Examples which are currently explored experimentally are the mean transverse momentum and mean energy fluctuations as well as fluctuations of particle ratios.
This is the same argument which leads to the canonical and grand-canonical description of a thermal system, as briefly discussed in section 2.
Event by Event Fluctuations
5.1. Fluctuations
of
469
Ratios
Let us begin with a general discussion of fluctuations of ratios, since mean energy or mean transverse momentum as well as particle ratios are all ratios of observables. Consider the ratio of of two observables A and B1
5
= E 6 X-
(l3?)
p,a
To find the average and the fluctuation of the ratio RAB
= |
(138)
we first write A = (A) + 8A and B = {B) + 5B and expand the numerator and the denominator to get _(A) RAB
(A)(5A
SB\
+
- W) W) { JA) " W))
2
(139)
( )
where 0(S2) indicates terms that are of quadratic and higher orders in 5A and 5B. Since A and B are extensive observables, the neglected terms are at most 0(1/ (N)) which we will neglect from now on. From Eq.(139), it is easy to see that (RAB)
= j|y
(140)
and the variance
(6R* \ -
{A)2
({5A2) + {SB2)
2{SA5BA
(Ul)
Using the basic correlator A ^ this can be rewritten as
{
B>
+
2
2
(6«)2
* -w-£h '" \w w- wm)
(142)
For an ideal gas, this simplifies to 2
(, 2AB)i d e a l _ ^ ) v{v , p), p/(a«)
~w\\
agfeg V{143)
[w w vm)
'Defined this way, both A and B scale with N, the number of particles in the final state.
470
S. Jeon and V. Koch
Obviously, deviations from the ideal gas value will give us insight into the dynamical correlations of the system. In order to expose these deviations, very often one compares to the fluctuations based on the 'inclusive' single particle distribution. The latter estimates the contribution of finite number fluctuations to the observed signal. In our formalism, this means that one evaluates equation (142) with the fully uncorrelated basic 'correlator' as defined in equ. (136). In addition, as it is common in the literature, one replaces the event averages (...) by inclusive averages T-r-rincl, which simply mean multiplying by the appropriate factors of (N). This let us define
( ri nVc l r2 y ^ cp , f Km) i + i ^in _ 2 ore \ mcl incl
(B
) V~
\(A y
(B y
(
A B )
In absence of any dynamical correlations
((8Rf) = j^WT*-
(145)
This observation has lead to the measure of dynamical fluctuations 0d y n a m i c 60 which is given by "dynamic = ( W )
~ <^) W
^
<146>
The first such measure to be proposed has been the so called $ variable ' which in terms of our variables here is given by
27 46
$ = y/(N) ((5R)2) - VW) 2 " 1 . The ratio has also been proposed i^=
(147)
56
_ i n / (148) (6R)2 which has the advantage of being dimensionless. Aside from simply subtracting the expected fluctuations from an uncorrelated system, the sub-event method has been developed in 60 . We note, that given the same momentum distribution the inclusive result is up to a trivial factor identical to that of an classical ideal gas, i.e ^dynamic = ^ = 0oxF = lm this case. However, in reality, the inclusive momentum distribution usually differs from a Boltzmann shape due to additional effects such as hydrodynamic flow. Notice also, that the results for
Event by Event Fluctuations 471
Bose and Fermi gases already differ from the inclusive estimator, reflecting the correlations induced by quantum statistics. For a Bose gas, $ > 0, F > 1, whereas for a Fermi gas $ < 0, F < 1. For the observable to be discussed below, the corrections to F are a few percent 5 6 . Finally, the above formalism allows us to discuss more general correlations between ratios of observables. Lets introduce two more observables C and D p,a
p,a
Then, the most general correlation of ratios can be written as 'A*
(§))"
(A)(C) ((5A6C) (B) (D) V {A) (C)
+
(8BSD) (B) (D)
(5A5D) {A) (D)
(SBSC)\ (B) (C) J
(
Uj
Using the basic correlator (125) we obtain
(*)'(§)>„a 0 aj^_
(A) (C) (B)(D) W J
^ J ^ (B) (D)
najp
a ^ (A) (D)
J.aj/3
\
bpd^ ^ (B) (C)
( i 5 i )
Note, that again this correlation function represents nothing but a specific moment of the basic correlator. Obviously, the replacements C — A and D = B leads to the ratio fluctuations discussed above. This general correlation will become useful below, when discussing charge dependent and charge independent transverse momentum fluctuations, as proposed by the STAR collaboration 51 . After having developed the general formalism for fluctuations of ratios it is now straight forward to discuss the actual observables. 5.2. Fluctuations momentum
of the mean energy and mean
transverse
The mean energy and transverse momentum are defined as =
E_=
Y,pnPE{p)
472
S. Jeon and V. Koch
Ep"PPt(p)
Pt N
(152) EP«P where E and (Pt) denote the Energy and transverse momentum of an event with N particles in the final state. Obviously, these observables are ratios and the above formalism can be readily applied with the substitutions P t =
A = E; or A = Pt B = N
(153)
So that for X being either the transverse momentum of the energy
((*)>(£.((»-$")') After rearranging terms one finds for an ideal gas
) )~
™
56
jy>)
(155)
" P ^
where X stands either for the energy E or the transverse momentum Pt • In addition to the transverse momentum fluctuations for all charged particles, one can investigate the pt fluctuations of the negative and positive charges independently as well as the cross correlation between them. Let us define Pt
Pt
_ EP"p(p)Pt(p)
(156)
where n^(p) is the momentum distribution of the positively (negatively) charged particles. Using Eq.(142), we then have
Pt(p)+Pt(q)
(N^iP?) , ±,± £ (MP) - {pt)) (Mi) - (pf)) AP,Q
(N±y
A±,±
(157)
p,q
where to get the last line we used (pf) = {Pt) I (N±)- For the unlike-sign pairs, we use Eq.(151) to get {SptSp
^
=
(N+)(N.)
£
(P*<«) -
~
Event
by Event
Fluctuations
473
Therefore it is natural to define the measure of dynamic fluctuation as
Aa2pf = (N±) ((Spf)2) - (5pfr
-incl
(159)
incl
For unlike-sign pairs, 5p^Spt
= 0 so that
A*J+_ = y/{N+) (N-) (Sp+Spt)
(160)
This then allows us to define the 'charge-independent' (CI) and 'chargedependent' (CD) combinations as used in the STAR and the PHENIX collaborations 51 (AT) Aa2ChCD(Pt)
= (N+) Aa\
+ {N.) A*2
± 2y/(N+)
(N-)Aa2+(161)
Here the (+)-sign in front of last term leads to the 'charge-independent' and the (—)-sign to the 'charge dependent' combinations, respectively. Under the quite reasonable assumption that (P+) _ (Pf)
_ (Pt++Pt~)
and (Spf)2
-incl
= (5pt ) 2
{Nch) Aa2ChCD(pt)
incl
_
{N++N-) ~ iPt)
JN^~WJ= (5pt)2
(162)
, this can be rewritten as
=
-2{Nch)(6pty
r-incl
(163)
where the assignment of plus and minus signs are as in Eq.(161). Notice the the combinations of the basic correlator A in the 'chargedependent' combination is the same as that of the net charge fluctuation Eq.(132) . Likewise, the 'charge-independent' combination resembles the fluctuation of the fluctuations of the number of charged particles, eq. (133). As we discussed in detail in section 4, charge conservation suppresses the fluctuations of the net-charge in a subsystem. Therefore, one would expect, that also the 'charge-dependent' transverse momentum fluctuations are suppressed, since they are nothing but a different moment of the same correlator. This is seen in the preliminary data of the STAR collaboration 51
474
S. Jeon and V. Koch
Other correlation, such as resonances and flow may also affect the transverse momentum fluctuations. In Ref.56 these have been estimated and found to be small, of the order of 1 — 2%. 5.3. Fluctuations
of particle
ratios
The fluctuation of particle ratios are also presently studied. Consider the ratio of particles " 1 " and "2" R12 = ^
(164)
then with A = Ni and B = N2 the dispersion is given by eq. (141) as ( ^ ) 2 \ (JVi)2
(AT2)2
(Ni)(N2)j
Notice that the last term introduces correlations between the particles which may reduce the fluctuations. One example for these correlations are hadronic resonances. Consider, for example the ratio of positively and negatively charged pions R = ?=. The the presence of neutral rho-mesons, which decay into ir+ and TT~ reduce the fluctuations. This can be easily understood by considering a gas made out of only neutral rho mesons. Independent, how large the fluctuations of these are, in every event on has as many TT+ as ir~, and hence the fluctuations of their ratio vanish. This effect can be utilized to estimate the number of resonance in the system at chemical freeze-out 3 4 . Next, consider the situation where one particle species is much more abundant then the other, i.e. (N2) 3> (Ni). This is the case of the fluctuations of the kaon to pion ratio, as investigated by the NA49 collaboration. Assuming that the number fluctuations are approximately Poisson, then the second term in eq. (165) dominates
&1!~J_»J_~»)! {Nlf
~ (Nl) >> (N2) ~
{N2f
(166)
(ibb)
so that <**?2> (R12)2
(1 + 0((N1)/(N2)))
(167)
~ (Ni)
In other words, if one particle species is much more abundant then the other, the correlations among the particles have to be very strong to be
Event by Event Fluctuations
475
visible in experiment. To give some numbers, in a standard resonance gas 18 the correlations due to resonance lead to 30% corrections in case of the 7r + /7r _ ratio and only to 4% in case of the K/TT ratio 3 4 . Thus, ratios of equally abundant particles are best suited to expose possible correlations. The simplest, and arguably most interesting is that of positively over negatively charged particles R+{XSRl +
= ^
(168)
) = R .f (M£ < ^ _2 M = A ' X{ ++ ' \{N+)2 + (N-) {N-)(N-)J
(169)
In the limit that the net charge (Q) = {N+ — N-) is much smaller than the number of charged particles {NCh) = (N+ + N-), or (Q) <S (iVCh) ( i ? + _ ) ~ l ; (N+) ~ (AT_) ~ < M
(170)
so that <*i^_> = j^-5
(SNl + 6N1 - 25N+ <57V_)
\-i>ch/
= 4<«!>
(171)
Since the number of charged particles is a measure for the entropy of the system, (Nch) oc S, the observable
is a measure for the charge fluctuations per entropy. And this, as discussed in detail in section 2.1.1, is an observable for the existence of the Quark Gluon Plasma. As discussed in detail in 35 for a pion gas, Dpion-gas = 4 whereas for a QGP, DQQP ~ 1 — 1.5, where the uncertainty arises from relating the entropy S with the number of charged particles (Nch). Hadronic resonances introduce additional correlations, which reduce the value of the pion gas to .Dhadron-gas — 3, but still a factor of 2 larger then the value for the QGP. (SO2)
In principle, one could directly measure the ratio \N I, without going through ratio fluctuations. However, since the net charge is an extensive quantity, this will introduce volume fluctuations into the measurement.
476
S. Jeon and V. Koch
Only in the limit that the total charge of the system is zero, volume fluctuations do not contribute to the order considered here 37 . The key question of course is, how can these reduced fluctuations be observed in the final state which consists of hadrons. Should one not expect that the fluctuations will be those of the hadron gas? The reason, why it should be possible to see the charge fluctuations of the initial QGP has to do with the fact that charge is a conserved quantity. Imagine one measures in each event the net charge in a given rapidity interval Ay such that AycM
< Ay < Ay max
(173) where Aymax is the width of the total charge distribution and Aycou is the typical rapidity shift due to hadronization and re-scattering in the hadronic phase. If, as it is expected, strong longitudinal flow develops already in the QGP-phase, the number of charges inside the rapidity window Ay for a given event is essentially frozen in. And if Ay S> Aycou neither hadronization nor the subsequent collisions in the hadronic phase will be very effective to transport charges in and out of this rapidity window. Thus, the E-by-E charge-fluctuations measured at the end reflect those of the initial state, when the longitudinal flow is developed. Ref. 54 arrives at the same conclusion on the basis of a Fokker-Planck type equation describing the relaxation of the charge fluctuation in a thermal environment. In Fig.2 we show the results of an URQMD calculation 15 (left figure), where the variable D is plotted versus the size of the rapidity window Ay. For large Ay the results have to be corrected for charge conservation effects; if all charges are accepted, global charge conservation leads to vanishing fluctuations (open symbols in Fig.2). This can be corrected for as explained in the previous section (for details see 1 5 ) . The resulting values for D are shown as full symbols in Fig.2. They agree nicely with the prediction for the resonance gas, as they should, since the URQMD model does not contain any partonic degrees of freedom. For small Ay the correlations imposed by the resonances are lost, because only one of the decay products is accepted. As a result we see an increase of D. The right figure shows a comparison of several transport models, including the Parton Cascade model 2 9 . This model starts with partons in the initial state, but has some model for hadronization included. If the general ideas about the reduced charge fluctuations are correct, this model should lead to smaller values of D, and it does.
Event by Event Fluctuations
• D
til "14
T'
• — • HIJING, cor • • HIJING/BBbar, cor • —-»VNIb, cor A—iUrQMD, cor T---TRQMD, cor Au+Au, 200A GeV
4
-1
\ I
Q3
++
•
X
1
-e-
" — 1
r
*
477
* ru m
Au+Au s,/2=200AGeV. b<2tm
1
*y
3
Ay
Fig. 2. Left: Charge fluctuations for different rapidity windows as obtained in the URQMD model. Open symbols: without correction for global charge conservation. Full symbols: With correction for global charge conservation. Right: Results for different transport model including the parton cascade 6 3 . These results are corrected for charge conservation.
For very small Ay, when (N) ~ 1, the ratio D is not well defined for events with iV_ = 0, and, therefore, cannot serve as a observable. Alternative observables, measuring the same quantity have been proposed and studied in 4 9 . 5.4. Kaon
Fluctuations
As discussed in section 2.2 at low energy the strangeness conservation has to be treated explicitly. This can be utilized to determine the degree of equilibration reached in these collisions 36 . By measuring the ratio F2 =
{NK(NK
M
- 1))
2
(174)
of the number of kaon pairs over the square of the inclusive number of kaons a factor two sensitivity on the degree of equilibrium can be obtained. From transport models (see e.g. 17 ) it is known, that most of the kaons are produced in secondary collisions, i.e. during the evolution of the fireball. Transport calculations also find that the equilibration time of kaons is of the order of 500fm/c, which is about ten time the lifetime of the system. Hence these models would not predict equilibration of the kaons. However, observed particle ratios including kaons are also consistent with a thermal
478
S. Jeon and V. Koch
model description (for a review see 1 8 ) . The measurement of F2 can resolve this issue and point to new physics such as multi-particle processes, in medium effects etc., if indeed it is found to be consistent with equilibrium. This is demonstrated in Fig.3 where the time evolution of F2 for several initial kaon numbers is shown. In all cases, F2 quickly rises close to F2 — 1 before it settles at the final equilibrium value of F | q ' ~ 1/2. Thus, by
i|
0.8
1
1
1
1
,
1
1
r
— N^N^-0.01 •••
"A.\
V
-
" ~ -
0.6
a in
.
L !
0.4
£ E = 0.1
- \
0.2
l
0
1
2
3
.
l
•
,
4
5
Fig. 3. Time evolution of Fi for various initial kaon numbers. The thin lines are analytical results for early times.
measuring F2 one can directly determine how close to chemical equilibrium the system has developed, before it freezes out. If the predictions of the transport models are correct, then a value of F2 — 1 should be found. If, on the other hand, equilibrium is indeed reached in these collisions, then F2 — \. In principle a similar measurement can also be done at higher energies for charmed mesons. To which extent this is technically feasible is another question. Let us conclude the section by mentioning other observables. In the context of so called DCC production 50 , the fluctuations of the fraction of neutral pions is considered a useful signal. Also the fluctuations of the elliptic flow has be proposed as an useful observable 47 , which may reveal
Event by Event Fluctuations
479
new configurations such as sphalerons, which could be created in heavy ion collisions.
480
S. Jeon and V. Koch
6. Experimental Situation 6.1. Fluctuation
in elementary
collisions
By now, it is clear that Quantum Chromodynamics is the right theory of strong interaction. However, essential non-perturbative nature of the strong interaction still remains a difficulty. In heavy ion collisions, this is more manifest in the sense that even the hard part of the spectrum is strongly influenced by the surrounding soft medium through multiple scatterings. Therefore, before we begin to consider the experimental results from heavy ion collisions, it is crucial that we understand elementary collision results such as proton-proton collision results since these elementary collision results can provide unbiased baseline. A large amount of data for pp collisions at various beam energies were taken from Fermi Lab in the 1970's 19 . 39 . 38 ' 40 (see also Ref. 61 ). Among them, data at the beam energy of 205 GeV were most thoroughly analyzed by Kafka et.al. 38 ' 39 . First, let us consider the 'charge fluctuation study'. In the case of pp collisions, the definition of 'charge fluctuation', however, differs from ours. Let us define the charge transfer u{y) = \[Qf{y)-Qb{y)\
(175)
where Q/(y) is the sum of the charges of the particles with rapidity larger than y and Qb(y) is the sum of the charges of the particles with rapidity smaller than than y. We then define the charge transfer fluctuation (6u(y)2)ebe
= (u2(y))ehe-(u(y))lhe
(176)
This quantity is usually referred to as 'charge fluctuation' in literature dealing with proton-proton collisions. This charge transfer fluctuation is certainly not the same as what we have been discussing so far which is the charge fluctuation within a given phase space window A77. However, as we will show shortly, (5u(y)2) is intimately related to the charge correlation functions. By considering the charge transfer fluctuation, we can then put a constraint on possible forms of the correlation functions. To do so, consider again our simple model defined by Eq.(81). As before, we impose the conditions M+ — M_ = Qc = 0, / + = / _ = / and h+ = h- = h. After straightforward but tedious algebra, we obtain (Su(y)2)ehe
= « M + ) + <M_» /
dy' f(y') /
dy" f(y")
Event by Event Fluctuations
/•oo
+ 2 (M0)/
py
dy+
Jy
481
dy-My+,y-)
(177)
J —oo
using the rapidity y as the phase space variable. In Ref.38, it is shown that the data satisfies (Su(y)\he
« 0.62^
(178)
at the beam energy of 205 GeV and the shape of dN^/dy is very well represented by a Gaussian with a2y « log(y/s/2mN) ~ 2.2. This result puts a condition on possible forms of / and /o- In our model, the rapidity distribution of the charged particle is ^
= «M+> + <M_» f(y) + 2 (M 0 ) h(y) (179) dy First, let us see if charge particles alone can satisfy Eq.(178). For this to be possible we must have f°° dy' f(y') f*^ dy" f(y") oc f(y) and / / = 1. These conditions are satisfied by f(y) = -^sech 2 (y/y)
(180)
2j/
where y is a constant specifying the width of the distribution.-* This form of f(y) is however, too sharply peaked to be consistent with the data. On the other hand a Gaussian with the above
- y2)g((yi
+ ya)/2)
(182)
where q(y) is a sharply peaked function at y = 0 with a small width A, and g(y) « ( 1 / (N))dN/dy. Then it can be also shown that h(y) PS g(y) and
/ dyi J
r
" dy"fo(v',v")*C*h(v)
(183)
" h ( x ) = 1 - t a n h 2 ( x ) = sech 2 (x) ax
(181)
This can be easily verified using dta
482
S. Jeon and V. Koch
with some constant C < 1. In fact the authors of Ref.38 argued that their proton-proton result is consistent with having only the neutral clusters. In our language, that corresponds to (M±) = 0. What the above consideration implies for our charge fluctuation is somewhat unclear. To the authors' knowledge, direct measurement of charge fluctuation in the sense defined in the previous sections has not been carried out in proton-proton experiments. What have been measured are the Cap functions defined in Eq.(69). In Ref39, the maximum height at t/i = j/2 = 0 are given as C++(0,0) « 0 . 3 6 C__(0,0) w0.25 C+_(0,0)«0.5 Using also the overall shape given in the same reference and the fact {Nch) ~ 6 at this energy one can then infer that
(184) 38
that
, , r , " « 0.5 ~ 0.6 (185) (^ch) A „ within — 1 < y < 1 without any charge conservation corrections. Correcting for charge conservation is a difficult task to perform. If as asserted in Ref.38 all particles are produced via neutral clusters, then there is no correction to perform. If however, some charged particles are emitted independently, then this is the lower bound. A more thorough analysis of the available data can undoubtedly yield more accurate estimate of the charge fluctuation. However, that is clearly beyond the scope of this review. We just make a remark here that our estimate of the charge fluctuation in QGP scenario (SQ2)A I (Nch)A « 0.25 ~ 0.3, seems to be still substantially smaller than the above proton-proton collision result. 6.2. Fluctuations
in Heavy Ion
Reactions
Unfortunately at the time of the writing of this review, very few published data on fluctuations in heavy ion collisions are available. Quite a few preliminary results are being discussed at conferences, which we will briefly mention. But we feel that an in depth discussion of these results prior to publication is not appropriate. The pioneering event-by-event studies have been carried out by the NA49 collaboration. They have analyzed the fluctuations of the mean transverse momentum 5 and the kaon to pion ratio 4 .
Event by Event Fluctuations
483
Both measurements have been carried out at at the CERN SPS at slightly forward rapidities. In Fig. 4 the resulting distributions are shown together with that from mixed events (histograms). In both cases the mixed event can essentially account for the observed signal, leaving little room for genuine dynamical fluctuations. Specifically, NA49 gives $ P t = 0.6 ± 1.0 MeVk for the transverse momentum fluctuations, which is compatible with zero. For the kaon to pion ratio they extract a width due to non-statistical fluctuations of anon3tat = 2.8% ± 0.5%, which would be compatible form the expectations of resonance decays 34 .
0.4 0.5 MdXj) (GeV/c)
0.1
0.2 0.3 0.4 Single Event K/JC ratio
Fig. 4. Results for the fluctuations of the mean transverse momentum (left) and kaon to pion ratio (right). Both results are from the NA49 collaboration.
The PHENIX collaboration recently also reported their results on mean transverse momentum and energy fluctuations at RHIC energies 2 . Similar to NA49, their result is compatible with the statistical fluctuations only, leaving no room for significant dynamical fluctuations. Their measurement was around mid-rapidity with a small azimuthal acceptance of A $ = 58° In contrast to that preliminary results by both the CERES collaboration 21 at the CERN SPS, as well as the STAR collaboration 51 at RHIC report significant dynamical fluctuations of the mean transverse momentum. For the definition of * p , see previous section.
484
S. Jeon and V. Koch
Both measure at mid-rapidity. It remains to be seen if these difference in measurements can be attributed simply to the different acceptance regions covered. As far as the charge fluctuations are concerned, first results have been reported. PHENIX 3 at RHIC which measures with a small rapidity acceptance, finds charge fluctuations consistent with a resonance gas. CERES 21 and NA49 16 , which both measure at SPS energies, report preliminary results on charge fluctuations, which are consistent with a pure pion gas. However, at the SPS the overall rapidity distribution is rather narrow, so that the correlation effect of the resonance gets lost when correcting for charge conservation 6 2 . But certainly, non of the measurements is even close to the prediction for the QGP. These findings have prompted ideas, that possibly a constituent quark plasma, without gluons, has been produced 11 . However, the measurement of additional observables would be needed in order to distinguish this from a hadronic gas. But maybe the present range of Ay is so small, that the charge fluctuations have time to assume the value of the resonance gas. As shown in Fig. 2, the results for the parton cascade arrive at the predicted value for the QGP only for Ay > 3. None of the present experiments has such a coverage yet and thus a detailed analysis of D as a function of Ay is needed, before any firm conclusions can be drawn. The Balance function measurement at ,/s = 130 GeV has been reported by STAR collaboration 24 . Going from peripheral to central collisions, the width of balance function steadily decreases. The trend is what one would expect if more of the system is filled with a QGP as the collision becomes more central. However, since the reduction is only about 20 % going from most perpheral to most central, it is not yet clear whether this signals the presence of a QGP or more mundane effect such as the strong flow.
Event by Event Fluctuations
485
7. Conclusions and Outlook In this review we discussed several aspects of the emerging field of event-byevent fluctuations in heavy ion collisions. We have briefly reviewed the basics concepts of fluctuations in a thermal system, and we have discussed the possible connection between a idealized thermal system and a "real world" heavy ion collision. Furthermore we have introduced the variables used in order to present fluctuations and discussed the relation between them. Finally we have present the presently published data, which are rather limited. But this is soon going to change. Crucial for the understanding of fluctuations is the connections to correlation functions. While the concept of fluctuations might be more intuitive and thus might lead us to new insights and approaches, the relevant information is encoded in the correlation functions. For example, consider the charge fluctuations, which we have discussed in some detail. It is true that like-sign and unlike-sign correlation functions encode all the necessary information on the charge fluctuations. However the physical meanings of those encoded information can be made clear only through considering charge fluctuations in and out of QGP and its connection to the relevant features of the correlation functions. But certainly from the observational point of view, the information to be extracted are the correlation functions between all possible quanta. Once they are available, the relevant observables can be readily obtained by folding the correlation functions with the appropriate variables On important aspect relevant to heavy ion collisions is the finiteness of the system. Thus approximations which are valid for a large system, such as the grand-canonical ensemble. Appropriate corrections have to be applied to either theory or data in order to extract the properties of the bulk matter. While this is not new to heavy ion physics, for the case of fluctuations these methods still require more refinement. Since the study of fluctuations is a new and developing area of heavy ion physics, this review can only be some kind of snapshot at a "hopefully" fast developing field. By no means can it be comprehensive. Most of the experimental data are available, if at all, only in preliminary form. Furthermore analysis methods and possible new observable based on fluctuations are being developed as we write the review. Therefore our idea was to present the basic concepts as well as some of the necessary formalism in a consistent
486
5. Jeon and V. Koch
and understandable way. To which extent we have succeeded is beyond our judgment.
Acknowledgments The authors thank S. Voloshin, L. Ray, C. Gale, C. Pruneau and S. Gavin for suggestions and discussions. V.K. was supported by the Director, Office of Science, Office of High Energy and Nuclear Physics, Division of Nuclear Physics, and by the Office of Basic Energy Sciences, Division of Nuclear Sciences, of the U.S. Department of Energy under Contract No. DE-AC03-76SF00098. S.J. thanks RIKEN BNL Center and U.S. Department of Energy [DE-AC02-98CH10886] for providing facilities essential for the completion of this work. S.J. is also supported in part by the Natural Sciences and Engineering Research Council of Canada.
Event by Event Fluctuations
487
References 1. E.g. COBE web page at http://space.gsfc.nasa.gov/astro/cobe/cobe_ home.html 2. K. Adcox et al. Event-by-event fluctuations in mean p(t) and mean e(t) in s(nn)**(l/2) = 130-gev au + au collisions. Phys. Rev., C66:024901, 2002. 3. K. Adcox et al. Net charge fluctuations in au + au interactions at s(nn)**(l/2) = 130-gev. Phys. Rev. Lett., 89:082301, 2002. 4. S. V. Afanasev et al. Event-by-event fluctuations of the kaon to pion ratio in central pb + pb collisions at 158-gev per nucleon. Phys. Rev. Lett, 86:19651969, 2001. 5. H. Appelshauser et al. Event-by-event fluctuations of average transverse momentum in central pb + pb collisions at 158-gev per nucleon. Phys. Lett, B459:679-686, 1999. 6. Masayuki Asakawa, Ulrich W. Heinz, and Berndt Muller. Fluctuation probes of quark deconfinement. Phys. Rev. Lett, 85:2072-2075, 2000. 7. Gordon Baym, B. Blattel, L. L. Frankfurt, H. Heiselberg, and M. Strikman. Correlations and fluctuations in high-energy nuclear collisions. Phys. Rev., C52:1604-1617, 1995. 8. Gordon Baym and Henning Heiselberg. Event-by-event fluctuations in ultrarelativistic heavy-ion collisions. Phys. Lett, B469:7-ll, 1999. 9. Boris Berdnikov and Krishna Rajagopal. Slowing out of equilibrium near the qcd critical point. Phys. Rev., D61:105017, 2000. 10. George F. Bertsch. Meson phase space density from interferometry. Phys. Rev. Lett, 72:2349-2350, 1994. 11. A. Bialas. Charge fluctuations in a quark antiquark system. Phys. Lett, B532:249-251, 2002. 12. A. Bialas and V. Koch. Event by event fluctuations and inclusive distributions. Phys. Lett, B456:l-4, 1999. 13. J. P. Blaizot, E. Iancu, and A. Rebhan. Comparing different hard-thermalloop approaches to quark number susceptibilities. 2002. 14. J. P. Blaizot, Edmond Iancu, and A. Rebhan. Quark number susceptibilities from htl-resummed thermodynamics. Phys. Lett, B523:143-150, 2001. 15. M. Bleicher, S. Jeon, and V. Koch. Event-by-event fluctuations of the charged particle ratio from non-equilibrium transport theory. Phys. Rev., C62:061902, 2000. 16. C. Blume et al. Results on correlations and fluctuations from na49. 2002. 17. E. L. Bratkovskaya, W. Cassing, and U. Mosel. Analysis of kaon production at sis energies. Nucl. Phys., A622:593-604, 1997. 18. P Braun-Munzinger, K Redlich, and J Stachel. this review volume. 19. C. M. Bromberg et al. Pion production in p p collisions at 102-gev/c. Phys. Rev., D9:1864-1871, 1974. 20. J. Cleymans, K. Redlich, and E. Suhonen. Canonical description of strangeness conservation and particle production. Z. Phys., C51:137-141,
488 S. Jeon and V. Koch 1991. 21. CERES collaboration, to be published, 2002. prelimnary results, private communication. 22. M. Doering and V. Koch. Event-by-event fluctuations in heavy ion collisions. Acta Phys. Polon., B33:1495-1504, 2002. 23. V. L. Eletsky, J. I. Kapusta, and R. Venugopalan. Screening mass from chiral perturbation theory, virial expansion and the lattice. Phys. Rev., D48:43984407, 1993. 24. J.Adams et.al. Narrowing of the balance function with centrality in au + au collisions s(nn)**(l/2) = 130-gev. 2003. 25. Z. Fodor and S. D. Katz. Lattice determination of the critical point of qcd at finite t and mu. JHEP, 03:014, 2002. 26. Rajiv V. Gavai and Sourendu Gupta. The continuum limit of quark number susceptibilities. Phys. Rev., D65:094515, 2002. 27. M. Gazdzicki and S. Mrowczynski. A method to study 'equilibration' in nucleus-nucleus collisions. Z. Phys., 054:127-132, 1992. 28. Marek Gazdzicki. A method to study chemical equilibration in nucleus nucleus collisions. Eur. Phys. J., 08:131-133, 1999. 29. Klaus Geiger and Berndt Muller. Dynamics of parton cascades in highly relativistic nuclear collisions. Nucl. Phys., B369:600-654, 1992. 30. S. Gottlieb et al. Thermodynamics of lattice qcd with two light quark flavours on a 16**3 x 8 lattice, ii. Phys. Rev., D55:6852-6860, 1997. 31. Ulrich W. Heinz and Barbara V. Jacak. Two-particle correlations in relativistic heavy-ion collisions. Ann. Rev. Nucl. Part. Sci., 49:529-579, 1999. 32. H. Heiselberg and A. D. Jackson. Anomalous multiplicity fluctuations from phase transitions in heavy ion collisions. Phys. Rev., 063:064904, 2001. 33. Henning Heiselberg. Event-by-event physics in relativistic heavy-ion collisions. Phys. Rept., 351:161-194, 2001. 34. S. Jeon and V. Koch. Fluctuations of particle ratios and the abundance of hadronic resonances. Phys. Rev. Lett, 83:5435-5438, 1999. 35. S. Jeon and V. Koch. Charged particle ratio fluctuation as a signal for qgp. Phys. Rev. Lett, 85:2076-2079, 2000. 36. S. Jeon, V. Koch, K. Redlich, and X. N. Wang. Fluctuations of rare particles as a measure of chemical equilibration. Nucl. Phys., A697:546-562, 2002. 37. Sang-yong Jeon and Scott Pratt. Balance functions, correlations, charge fluctuations and interferometry. Phys. Rev., 065:044902, 2002. 38. T. Kafka et al. Charge and multiplicity fluctuations in 205-gev/c p p interactions. Phys. Rev. Lett, 34:687-690, 1975. 39. T. Kafka et al. One, two, and three particle distributions in p p collisions at 205-gev/c. Phys. Rev., D16:1261, 1977. 40. T. Kafka et al. Correlations between neutral and charged pions produced in 300-gev/c p p collisions. Phys. Rev., D19:76, 1979. 41. J. Kapusta. Finite Temperature Field Theory. Cambridge University Press, 1989.
Event by Event Fluctuations
489
42. C. M. Ko et al. Kinetic equation with exact charge conservation. Phys. Rev. Lett, 86:5438-5441, 2001. 43. V. Koch, M. Bleicher, and S. Jeon. Event-by-event fluctuations and the qgp. Nucl. Phys., A698:261-268, 2002. 44. L Landau and L.M. Lifshitz. Statistical Physics. Pergamon Press, New York, 1980. 45. Zi-wei Lin and C. M. Ko. Baryon number fluctuation and the quark gluon plasma. Phys. Rev., C64:041901, 2001. 46. Stanislaw Mrowczynski. Transverse momentum and energy correlations in the equilibrium system from high-energy nuclear collisions. Phys. Lett., B439:611, 1998. 47. Stanislaw Mrowczynski and Edward V. Shuryak. Elliptic flow fluctuations. 2002. 48. Madappa Prakash, Ralf Rapp, Jochen Wambach, and Ismail Zahed. Isospin fluctuations in qcd and relativistic heavy-ion collisions. Phys. Rev., C65:034906, 2002. 49. C. Pruneau, S. Gavin, and S. Voloshin. Methods for the study of particle production fluctuations. 2002. 50. Krishna Rajagopal. The chiral phase transition in qcd: Critical phenomena and long wavelength pion oscillations. 1995. 51. R. L. Ray. Correlations, fluctuations, and flow measurements from the star experiment. 2002. Proceedings Quark Matter 2002, Nantes, France, July 2002. 52. P. Reif. Fundamentals of Statistical and Thermal Physics. McGraw-Hill, 1984. 53. Edward V. Shuryak. Event-per-event analysis of heavy ion collisions and thermodynamical fluctuations. Phys. Lett, B423:9-14, 1998. 54. Edward V. Shuryak and Misha A. Stephanov. When can long range charge fluctuations serve as a qgp signal? Phys. Rev., C63:064903, 2001. 55. Misha A. Stephanov, K. Rajagopal, and Edward V. Shuryak. Signatures of the tricritical point in QCD. Phys. Rev. Lett, 81:4816-4819, 1998. 56. Misha A. Stephanov, K. Rajagopal, and Edward V. Shuryak. Event-by-event fluctuations in heavy ion collisions and the QCD critical point. Phys. Rev., D60:114028, 1999. 57. L. Stodolsky. Temperature fluctuations in multiparticle production. Phys. Rev. Lett, 75:1044-1045, 1995. 58. C. Sturm et al. Kaon and antikaon production in dense nuclear matter. J. Phys., G28:1895-1902, 2002. 59. M. J. Tannenbaum. The distribution function of the event-by-event average p(t) for statistically independent emission. Phys. Lett, B498:29-34, 2001. 60. S Voloshin, V Koch, and H.G. Ritter. Event-by-event fluctuations in collective quantities. Phys. Rev., C60:0224901, 1999. 61. J. Whitmore. Multiparticle production in the fermilab bubble chambers. Phys. Rept, 27:187-273, 1976.
490 S. Jeon and V. Koch 62. Jacek Zaranek. Measures of charge fluctuations in nuclear collisions. Phys. Rev., C66:024905, 2002. 63. Q. H. Zhang, V. Topor Pop, S. Jeon, and C. Gale. Charged particle ratio fluctuations and microscopic models of nuclear collisions. Phys. Rev., C66:014909, 2002.
PARTICLE P R O D U C T I O N I N HEAVY ION COLLISIONS
Peter Braun-Munzinger°, Krzysztof Redlich 6 ' 0 , Johanna Stachel a
Gesellschaft fur Schwerionenforschung, GSI, D-64291 Darmstadt, Germany E-mail: [email protected] Fakultat fur Physik, Universitat Bielefeld, Postfach 100 131, D-33501 Bielefeld, Germany c Institute of Theoretical Physics, University of Wroclaw, PL-50204 Wroclaw, Poland E-mail: [email protected] Physikalisches Institut der Universitat Heidelberg, D 69120 Heidelberg, Germany E-mail: stachel@physi. uni-heidelberg. de The status of thermal model descriptions of particle production in heavy ion collisions is presented. We discuss the formulation of statistical models with different implementation of the conservation laws and indicate their applicability in heavy ion and elementary particle collisions. We analyze experimental data on hadronic abundances obtained in ultrarelativistic heavy ion collisions, in a very broad energy range starting from RHIC/BNL (v's = 200 A GeV), SPS/CERN ( ^ i ~ 20 A GeV) up to AGS/BNL (y/I ~ 5 A GeV) and SIS/GSI ( ^ s ~ 2 A GeV) to test equilibration of the fireball created in the collision. We argue that the statistical approach provides a very satisfactory description of experimental data covering this wide energy range. Any deviations of the model predictions from the data are indicated. We discuss the unified description of particle chemical freeze-out and the excitation functions of different particle species. At SPS and RHIC energy the relation of freezeout parameters with the QCD phase boundary is analyzed. Furthermore, the application of the extended statistical model to quantitative understanding of open and hidden charm hadron yields is considered.
491
492 P. Braun-Munzinger, K. Redlich and J. Stachel
Contents 1
Introduction 1.1 Initial conditions in A-A collisions and deconfinement 2 Statistical Approach — General Remarks 2.1 Statistical approach — grand canonical formalism 2.2 Thermal analysis of particle yields from AGS to RHIC energies . . 2.3 Comparison of measured particle densities with thermal model predictions 2.4 Statistical model and composite particles 3 Exact Implementation of the Conservation Laws in the Statistical Models 3.1 Kinetics of time evolution and equilibration of charged particles . 3.1.1 Kinetic master equation for probabilities 3.1.2 The equilibrium solution of the general rate equation . . . 3.1.3 The master equation in the presence of the net charge. . . 3.1.4 The kinetic equation for different particle species 3.2 The canonical description of an internal symmetry — projection method 3.2.1 Canonical models with a non-Abelian symmetry 3.2.2 The canonical partition function for Abelian charges . . . . 3.2.3 The equivalence of the canonical formalism in the grand canonical limit 4 The Canonical Statistical Model and Its Applications 4.1 Central heavy ion collisions at SIS energies 4.2 Particle production in high energy p-p collisions 4.2.1 Statistical hadronization and string dynamics in p-p collisions 4.3 Heavy quark production 4.3.1 Statistical recombination model 4.3.2 Results 4.3.3 Charmonium production from secondary collisions at LHC energy 4.3.4 Conclusions on heavy quark production 5 Unified Conditions of Particle Freeze-out in Heavy Ion Collisions . . . . 5.1 Chemical freeze-out and the QCD phase boundary 6 Particle Yields and Their Energy Dependence 7 Lifting of the Strangeness Suppression in Heavy Ion Collisions 8 Conclusions and Outlook Acknowledgements References
493 494 498 499 503 510 511 512 514 515 520 522 524 525 530 533 537 542 542 550 554 558 559 561 564 565 565 569 571 582 588 591 591
Particle Production
in Heavy Ion Collisions
493
1. Introduction The ultimate goal of the physics program with ultrarelativistic nucleusnucleus collisions is to study the properties of strongly interacting matter under extreme conditions of high energy density. Quantum Chromodynamics (QCD) predicts 1 - 5 that strongly interacting matter undergoes a phase transition from a state of hadronic constituents to a plasma of deconfined quarks and gluons (QGP). By colliding heavy ions at ultrarelativistic energies, one expects to create matter under conditions that are sufficient for deconfinement 1-10 . Thus, of particular relevance is finding experimental probes to check whether the produced medium in its early stage was indeed in the QGP phase. Different probes have been studied with the various SPS/CERN and RHIC/BNL experiments. The most promising signals of deconfinement are related to particular properties of the transverse momentum spectra of photons 11 ' 12 , dileptons 1 3 - 1 9 , and hadrons 9 ' 2 ° - 2 2 . The photon rate is studied to probe the temperature evolution from formation to decoupling of the fireball, implying sensitivity to a high temperature deconfined phase. The invariant mass distribution of dileptons is expected to be modified by in-medium effects related to chiral symmetry restoration 4 ' 5 ' 1 7 ' 1 9 ' 2 4 - 2 7 . The modification of charmonium production was argued to be a consequence of collective effects in the deconfined medium 1 ' 28 . Hadron multiplicities and their correlations are observables which can provide information on the nature, composition, and size of the medium from which they are originating. Of particular interest is the extent to which the measured particle yields are showing equilibration. The appearance of the QGP, that is a partonic medium being at (or close to) local thermal equilibrium and its subsequent hadronization during the phase transition should in general drive hadronic constituents towards chemical equilibrium 6 ' 7,9 ' 29 . Consequently, a high level of chemical saturation, particularly for strange particles 31 ' 33 , could be related to the deconfined phase created at the early stage of heavy ion collisions. The level of equilibrium of secondaries in heavy ion collisions was tested by analyzing the particle abundances 6 ' 9 ' 3 4 - 7 6 or their momentum spectra 9,20,21-23,37,46,47 j n fae grsf. c a s e o n e establishes the chemical composition of the system, while in the second case additional information on dynamical evolution and collective flow can be extracted. In this review we will discuss the formulation of statistical models and
494
P. Braun-Munzinger,
K. Redlich and J. Stachel
their applications to a phenomenological description of particle production in nucleus-nucleus collisions. We emphasize the importance of conservation laws and their different implementations in the statistical approach. We analyze experimental data on hadronic abundances obtained in ultrarelativistic heavy ion collisions, in a very broad energy range starting from RHIC/BNL (V5 = 130 A GeV), SPS/CERN (y/5 =± 20 A GeV) down to AGS/BNL (js~ ~ 5 A GeV) and SIS/GSI ( ^ 5 ~ 2 A GeV) to test equilibration. We argue that the statistical approach provides a very satisfactory description of experimental results covering this wide energy range. We further provide arguments for a unified description of chemical freeze-out of hadrons and discuss excitation functions of different particle species. An extension of the model for a quantitative understanding of open and hidden charm particle yields will be also discussed.
1.1. Initial
conditions
in A-A
collisions
and
deconfinement
In ultrarelativistic heavy ion collisions, the knowledge of the critical energy density ec required for deconfinement as well as the equation of state (EoS) of strongly interacting matter are of particular importance. The value of ec is needed to establish the necessary initial conditions in heavy ion collisions to possibly create the QGP, whereas the EoS is required as an input to describe the space-time evolution of the collision fireball*. Both of these pieces of information can be obtained today from first principle calculations by formulating QCD on the lattice and performing Monte-Carlo simulations. In Fig. (2) we show the most recent results 77 of lattice gauge theory (LGT) for the temperature dependence of energy density and pressure. These results have been obtained in LGT for different numbers of dynamical fermions. The energy density is seen in Fig. (2) to exhibit the typical behavior of a system with a phase transition b : an abrupt change in a very narrow temperature range. The corresponding pressure a
I n Fig. (1) we show a schematic view of the space-time evolution of heavy ion collisions in the Bjorken model 7 8 . "In a strictly statistical physics sense a phase transition in two flavour QCD can only appear in the limit of massless quarks where it is of second order. In three flavour QCD, with (u,d,s) quarks, the phase transition and its order depends on the value of the quark masses. In general it can be a first order, second order or cross-over transition. For physical quark masses, both the value of the transition temperature and the order of the deconfinement phase transition are still not well established.
Particle Production
in Heavy Ion Collisions
495
Ski
IT'
Fig. 1. Schematic space-time view of a heavy ion collision that indicates four basic stages in the evolution of the collision fireball: initial overlap region, pre-equilibrium partonic system, equilibrated quark-gluon plasma and its subsequent hadronization to a hadron gas.
curve shows a smooth change with temperature. In the region below Tc the basic constituents of QCD, quarks and gluons, are confined within hadrons and here the EoS is well parameterized 79 by a hadron resonance gas. Above Tc the system appears in the QGP phase where quarks and gluons can travel distances that substantially exceed the typical size of hadrons. The most recent results of improved perturbative expansion of the thermodynamical potential in continuum QCD indicate 81 ' 83 that, at some distance above Tc, the EoS of QGP can be well described by a gas of massive quasi-particles with a temperature dependent mass. In the vicinity of Tc the relevant degrees of freedom were argued 84 ' 85 to be described by Polyakov loops. Lattice Gauge Theory predicts, in two-flavour QCD, a critical temperature Tc = 173 ± 8MeV and corresponding critical energy density ec = 0.6 ± 0.3 GeV/fm 3 for the deconfinement phase transition 77 . The value of ec is surprisingly low and corresponds quantitatively to the energy density inside the nucleon. The initial energy density reached in heavy ion collisions can be estimated within the Bjorken model 78 . From the transverse energy ET measured in nucleus-nucleus collisions the initial energy
496
P. Braun-Munzinger,
1
K. Redlich and J. Stachel
!
=s
5.0
iW4 4.0
• P/T4
^
^
=
=
3.0
If/
2.0
•
II
1.0 0.0
3 flavour 2+1 flavour — 2 flavour
Br
J 1.0
17TC 1.5
2.0 2.5
3.0
3.5 4.0
.0
1.5
2.0
2.5
3.0
3.5
4.0
Fig. 2. The pressure P and energy density c normalized to the temperature to the fourth power, versus temperature normalized to its critical value. The calculations 7 7 were performed within LGT for different numbers of flavors. The values of the corresponding ideal gas results are indicated by the arrows.
density eo is determined as 1 eo(To)
1 dET
TTR2 TO
dy
(1)
where the initially produced collision fireball is considered as a cylinder of length dz = r^dy and transverse radius R ~ A1/3. Inserting for TTR2 the overlap area of colliding Pb nuclei together with an assumed initial time of To — 1 fm, and using an average transverse energy at midrapidity measured 86 at the SPS (y/s = 17.3 GeV) to be 400 GeV, one obtains eo
(TO ^ 1 fm) ~ 3.5 ± 0.5 GeV/fm J .
(2)
Increasing the collision energy to y/s = 130 A-GeV for Au-Au at RHIC and keeping the same initial thermalization time as at the SPS, would increase eo by only 50-60 %. However, at RHIC the thermalization time was argued in terms of different models 10 ' 90 to be shorter by a factor of 3-5. In the context of saturation models 1 0 ' 8 7 - 8 9 the thermalization time can be possibly related with the saturation scale 8 ' 10 . The basic concept of the
Particle Production
in Heavy Ion Collisions
497
saturation models is the conjecture that there is some transverse momentum scale p s a t where the gluon and quark phase space density saturates 1 0 ' 8 7 - 8 9 . For an isentropic expansion of the collision fireball, the transverse energy at Psat w a s related in Ref. (10) to that measured in nucleus-nucleus collisions in the final state. The saturation scale was also used to fix the thermalization time as req ~ l/ps&f Taking the value of ps&t predicted in Ref. (10) for RHIC energy, p s a t ~ 1.13 GeV, one gets req ~ 0.2 fm and a corresponding energy density eeq ~ 98 GeV/fm 3 . This is a larger value than expected for the initial energy density at RHIC in the McLerran-Venugopalan model 88 where e§HIC ~ 20 GeV/fm 3 , also in agreement with the prediction of Ref. (89). At SPS energy the saturation model described in Ref. (10) leads to e f
498
P. Brawn-Munzinger,
K. Redlich and J. Stachel
still not sufficient to create a QGP. The distribution of initially produced gluons is very far from being thermal, thus the system needs enough time to equilibrate. Recently it was shown8, in the framework of perturbative QCD and kinetic theory, that the equilibration of partons should definitely happen at LHC and most likely at RHIC energy. A previous microscopic study 93 within the Parton Cascade Model has led to the conclusion that thermalization can also be reached even at the lower SPS energy. Here, however, due to the relatively low collision energy, it is not clear whether a model inspired by perturbative QCD is indeed applicable. Assuming QGP formation in the initial state in heavy ion collisions one expects that the thermal nature of the partonic medium could be preserved during hadronization. c Consequently, the particle yields measured in the final state should resemble a thermal equilibrium population. 2. Statistical approach - general remarks In the approach of Gibbs (see, e.g., Ref. (94)) the equilibrium behavior of thermodynamical observables can be evaluated as an average over statistical ensembles (rather than as a time average for a particular state). The equilibrium distribution is thus obtained by an average over all accessible phase space. Furthermore, the ensemble corresponding to thermodynamic equilibrium is that for which the phase space density is uniform over the accessible phase space. In this sense, filling the accessible phase space uniformly is both a necessary and sufficient condition for equilibrium. Consequently, the agreement between observables and predictions using the statistical operator imply equilibrium (to the accuracy with which agreement is observed). "Filling phase space" is not a different statement, although it is often and erroneously used in the literature. In our further analysis we use in the statistical operator as Hamiltonian that leading to the full hadronic mass spectrum. In some sense this is synonymous with using the full QCD Hamiltonian. The only parameters in the statistical operator describing the grand-canonical ensemble are temperature T and baryon chemical potential HB- There is no room here for strangeness suppression (7S) factors. So the interpretation is that c
T h e fact that the phase transition is a driving force towards equilibration is found 29 ' e.g. in different kinetic models for Q G P evolution and hadronization.
Particle Production
in Heavy Ion Collisions
499
agreement between data and theoretical predictions implies statistical equilibrium at temperature T and chemical potential HB • If an additional factor 7S is needed to describe the data this implies a clear deviation from chemical equilibrium: a state in which e.g. strangeness is suppressed compared to the equilibrium value implies additional dynamics not contained in the statistical operator and not consistent with uniform phase space density. Similar arguments, of course, apply if one uses canonical phase space. If, in this regime, canonically calculated particle ratios agree with those measured, this implies equilibrium at temperature T and over the canonical volume V. To the extent that this describes data for e+e- or pp collisions, the same conclusions on thermodynamic equilibrium apply. However, we note that, in this approach e.g., particles ratios involving particles with hidden strangeness are generally not well predicted, again implying nonequilibrium behavior. 2.1. Statistical
approach
- grand canonical
formalism
The basic quantity required to compute the thermal composition of particle yields measured in heavy ion collisions is the partition function Z(T, V). In the Grand Canonical (GC) ensemble, ZGC{T,V^Q)
=Tr[e-/3("-£>'3-Qi)],
(3)
where H is the Hamiltonian of the system, Qi are the conserved charges and fiQ{ are the chemical potentials that guarantee that the charges Qi are conserved on the average in the whole system. Finally /? = 1/T is the inverse temperature. The Hamiltonian is usually taken such as to describe a hadron resonance gas. For practical reasons, the hadron mass spectrum contains contributions from all mesons with masses below ~1.5 GeV and baryons with masses below ~2 GeV. In this mass range the hadronic spectrum is well established and the decay properties of resonances are reasonably well known 91 . This mass cut in the contribution of resonances to the partition function limits, however, the maximal temperature to T m a x ~ 200 MeV, up to which the model predictions may be considered trustworthy 37 ' 38 ' 42,58 . For higher temperatures the contributions of ( in general poorly known) heavier resonances are not negligible. The interaction of hadrons and resonances are usually only included by implementing a hard core repulsions, i.e. a Van der Waals-type interaction. Details of such a implementation are discussed below. The main motivation of using the Hamiltonian of a hadron
500
P. Braun-Munzinger,
K. Redlich and J. Stachel
resonance gas in the partition function is that it contains all relevant degrees off freedom of the confined, strongly interacting medium and implicitly includes interactions that result in resonance formation. Secondly, this model is consistent with the equation of state obtained from the LGT below the critical temperature 79 ' 80 . In a strongly interacting medium, one includes the conservation of electric charge, baryon number and strangeness. The GC partition function (3) of a hadron resonance gas can then be written as a sum of partition functions In Zi of all hadrons and resonances
In Z(T, V,fi = Y,ln
Zi T y
( ' ' #'
(4)
where e, = yjp2 + m 2 and fi = ( / ^ B J ^ S I M Q ) with the chemical potentials Hi related to baryon number, strangeness and electric charge, respectively. For particle i of strangeness Si, baryon number S j , electric charge Q, and spin-isospin degeneracy factor g,,2 \nZi(T,V,fl
= ^ f y ° ° ±p2dpln[l ± Xiexp(-f3ei)},
(5)
with (+) for fermions, (-) for bosons and fugacity Xl(T, fi) = exp(
*)
(6)
Expanding the logarithm and performing the momentum integration in Eq. (5) we obtain
b.a.p-.V.ffl.^f^AW^).
(7)
fc=l
where K
m{T,n) = — = ^
1.—k—x*miK*(-f-)>
W
fc=i
The partition function (4) is the basic quantity that allows to describe all thermodynamical properties of a fireball composed of hadrons and resonances being in thermal and chemical equilibrium. In view of further application of this statistical operator to the description of particle production in heavy ion collisions we write explicitly the results for particle density
Particle Production
"I—
UB=550
in Heavy Ion Collisions
501
MeV MeV
*-
flB=250
/ T
/ /
/
/
/
^y —-"
. O.I
. O.IS
0.2
T
[GeVJ
Fig. 3. The ratio of the total density of positively charged pions that includes all resonance contributions to the density of thermal pions. The calculations are done in the hadron resonance gas model for /J,B =250, 550 MeV and for different temperatures.
obtained from Eq. (4). Of particular importance here is to account for resonances and their decay into lighter particles. The average number (AT,) of particles i in volume V and temperature T, that carries strangeness Si, baryon number Bi, and electric charge Qi, is obtained from Eq. (4) as
WXT./Z) = (Ni)th(T,P) + J2jrj^i(Nj)th'R(T,fl)
(9)
where the first term describes the thermal average number of particles of species i and second term describes overall resonance contributions to particle multiplicity of species i. This term is taken as a sum of all resonances that decay into particle i. The r.,_>j is the corresponding decay branching ratio of j —> i. The corresponding multiplicities in Eq. (9) are obtained from Eq. (8). The importance of the resonance contribution to the total particle yield in Eq. (9) is illustrated in Fig. (3) as the ratio of total to thermal number of 7r+. From this figure it is clear that at high temperature (or density) the overall multiplicity of light hadrons is indeed dominated by resonance decays. In the high-density regime, that is for large T and/or /i^, the repulsive interactions of hadrons should be included in the partition function
502
P. Braun-Munzinger,
K. Redlich and J. Stachel
(4). To incorporate the repulsion at short distances one usually uses a hard core description by implementing excluded volume corrections 58 . In a thermodynamically consistent approach 82 these corrections lead to a shift of the baryon-chemical potential. . We discuss below how this is implemented in our calculations. The repulsive interactions are important when discussing observables of density type. Particle density ratios, however, are only weakly affected38 by the repulsive corrections. The partition function (4) depends in general on five parameters. However, only three are independent, since the isospin asymmetry in the initial state fixes the charge chemical potential and the strangeness neutrality condition eliminates the strange chemical potential. Thus, on the level of particle multiplicity ratios we are only left with temperature T and baryon chemical potential (1B as independent parameters. In Fig. (4) we show the relation of (is = HS{T,(IB) obtained from the strangeness neutrality condition. For low temperature this relation is highly non-linear. For larger T, however, (is shows an almost linear dependence on (is- One sees by inspection of Fig. (4) that, at T ~ 200 MeV and (XB ~ 300 MeV, (is ~ \HB- This relation is obtained in a QGP from strangeness neutrality conditions. In the present context of a hadron resonance gas this is a pure accident with no dynamical information. At lower energies, in practise for T < 100 MeV, the widths of the resonances have to be included 49 ' 25 in Eq. (9). This is because the number of light particles coming from the decay of resonances is increased by the finite resonance width. In practice, the width of the A resonance is most important 25 ' 27 . Thus, the approximation of the resonance width by a 5 function is not justified. Assuming the validity of Boltzmann statistics one replaces the partition function in equation (7) by: \nZR = N ^ T f -
d
exp[(BR(iB + QRHQ + SR(is)/T} s s K ^ s , T
)
\
J
- - ^ -
m
(10)
where s m j n is chosen to be the threshold value for the resonance decay and y/smax ~ mR + 2TR. The normalization constant iV is adjusted such that the integral over the Breit-Wigner factor gives 1. The statistical model, outlined above, was applied 3 4 - 7 5 ' 7 6 to describe particle yields in heavy ion collisions. The model was compared with all available experimental data obtained in the energy range from AGS up to RHIC energy. Hadron multiplicities ranging from pions to omega baryons
Particle Production
in Heavy Ion Collisions
503
0.2
fB[0eV] Fig. 4. The strange chemical potential /is as a function of baryon-chemical potential for T=120,170 and 200 MeV. The results are obtained by imposing the strangeness neutrality condition in a hadron resonance gas.
and their ratios were used to verify that there is a set of thermal parameters (T,[1B) which simultaneously reproduces all measured yields. In the following Section we present the most recent analysis of particle production in A-A collisions at RHIC, SPS and AGS energies. 2.2. Thermal analysis RHIC energies
of particle
yields from AGS
to
For the analysis of data in the energy range of 40 GeV/nucleon and upwards'1 we use a grand canonical ensemble to describe the partition function and hence the density of the particles following Eqs. (4 -9). As discussed above the temperature T and the baryochemical potential HB are the two independent parameters of the model, while the volume of the fireball V, the strangeness chemical potential /is, and the charge chemical potential /J,Q are fixed by the following additional conditions. First, overall strangeness The results in this section were obtained in collaboration with D. Magestro and are published in part in Refs. (35, 59).
504
P. Braun-Munzinger,
K. Redlich and J. Stachel
conservation fixes [is- Note that this applies strictly for data integrated over 47r. For slices near mid-rapidity this condition is, however, also appropriate as the flow of strangeness in and out of the rapidity slice under consideration very nearly cancels. Charge conservation implies a condition on I3 according to:
i
Here, Z and N are the proton and neutron numbers of the colliding nuclei, I3 and If are the third component of the total isospin and that of particle i. This condition is appropriate (and relevant) at lower beam energies where there is full stopping and Air yields are used. For details see Refs. (38, 60). At higher energies and for data analyzed in rapidity slices the right hand side of Eq. (11) has to be replaced by the neutron excess of baryons transported into the rapidity slice under consideration. This number is clearly smaller than the full neutron excess entering Eq. (11) but in general not well known. However, its precise knowledge is less relevant for higher beam energies since the isospin balance is dominated by pions. For practical purposes isospin conservation is important for AGS energies and below but its effect is small (on the 10 % level) already at 40 GeV/nucleon beam energy (where we have used as an upper limit the full neutron excess of the colliding nuclei, leading to a slight overestimate of the pion charge asymmetry) and negligible at top SPS and RHIC energies. Finally, the volume (which drops out anyway for particle ratios) can be obtained from total baryon number conservation (for full stopping and quantities which are evaluated over the complete phase space) or is fixed by using the measured pion multiplicity in the rapidity slice under consideration. As discussed above the hadronic mass spectrum used in the calculations extends over all mesons with masses below 1.5 GeV and baryons with masses below 2 GeV. To take into account a more realistic equation of state we incorporate the repulsive interaction at short distances between hadrons by means of the excluded volume correction discussed above. A number of different corrections have been discussed in the literature. Here we choose that proposed in Refs. (60, 61, 82): pexcL(T,n)
= pid-9as(T,{i);
with fi = n - veigen pexcl(T,n).
(12)
This thermodynamically consistent approach to simulate interactions between particles by assigning an eigenvolume veigen to all particles modifies the pressure p within the fireball. Equation (12) is recursive, as it uses the
Particle Production
1
in Heavy Ion Collisions
\ I
c
I
c
*K
-
—m-
'g
+ -1
10)
505
IT)
f-
< ~#_
< II
T
t'l
S
±
•<•• CL
-2
10)
-
:
1
•
model dato
Fig. 5. Comparison between thermal model predictions and experimental particle ratios for P b - P b collisions at 40 GeV/nucleon. The thermal model calculations are obtained with T = 148 MeV and fiB = 400 MeV.
modified chemical potential ft to calculate the pressure, while this pressure is also used in the modified chemical potential, and the final value is found by iteration. Particle densities are calculated by substituting /x in Eq. (8) by the modified chemical potential p,. The eigenvolume has to be chosen appropriately to simulate the repulsive interactions between hadrons, and we have investigated the consequences for a wide range of parameters for this eigenvolume in Ref. (38, 60). Note that the eigenvolume is veigen = 4|-7r.R3 for a hadron with radius R. Assigning the same eigenvolume to all particles can reduce particle densities drastically but hardly influences particle ratios. Ratios may differ strongly, however, if different values for the eigenvolume are used for different particle species. Our approach here is, to determine, for nucleons, the eigenvolume according to the hard-core volume known from nucleon-nucleon scattering 62 . Consequently, we assigned 0.3 fm as radius for all baryons. For mesons we expect the eigenvolume not to exceed that of baryons. For lack of better theoretical guidance we chose also for the mesons a radius of 0.3 fm. For a discussion of the implications of varying these radius parameters see Ref. (38, 60). After thermal "production", res-
506
P. Braun-Munzinger,
-: 1
-
K. Redlich and J. Stachel
\* **
*B sz
's
••
— \ -*-
~
'B °B \
10
10
10
= " -
: -
<
Q.
^ ^ ^
-r +
*c
!.
"•"
I a.
<
T
< >
*•
•
I- ^
'B
+
f*
±
*•
\ SCNI
^ • "a
>
t
•
m o d e l , T = 170 MeV experiment
Fig. 6. Comparison between thermal model predictions and experimental particle ratios for P b - P b collisions at 158 GeV/nucleon. The thermal model calculations are obtained with T = 170 MeV and HB = 255 MeV.
onances and heavier particles are allowed to decay, therefore contributing to the final particle yield of lighter mesons and baryons, as indicated above. Decay cascades, where particles decay in several steps, are also included. Systematic parameters regulate the amount of decay products resulting from weak decays. This allows to simulate the different reconstruction efficiencies for particles from weak decays in different experiments. In the following we compare predictions of the model with results of measured particle ratios for central Pb-Pb collisions at SPS energies (40 and 158 GeV/nucleon) and for central Au-Au collisions at RHIC energies y/snn = 130 and 200 GeV. An important issue in this context is whether to use data at mid-rapidity or data integrated over the full phase space. While it is clear that full Am yields should be used at low beam energies, this is not appropriate any more as soon as fragmentation and central regions can be distinguished. In that case the aim is to identify a boost-invariant region near mid-rapidity and to choose a slice in rapidity within that region. For RHIC energies this implies that an appropriate choice, given the avail-
Particle Production
in Heavy Ion Collisions
507
able data, is a rapidity interval of width Ay = 1 centered at midrapidity. The anti-proton/proton ratio stays essentially constant within that interval, but drops rather strongly for larger rapidities and similar results are observed 63 for other ratios. Furthermore, the rapidity distribution exhibits a boost-invariant plateau near mid-rapidity 64 . As has been demonstrated in Ref. (65), effects of hydrodynamic flow cancel out in particle ratios under such conditions. At SPS energies a boost-invariant plateau is not fully developed but stopping is not complete, either. In addition, the proton and anti-proton rapidity distributions differ rather drastically, especially near the fragmentation regions, implying that particle ratios depend on rapidity (see. e.g., Ref. (66)). Under those circumstances we have decided to use, wherever available, data in a slice of ±1 unit of rapidity centered at midrapidity. This is slightly different from the analysis performed in Ref. (38), where both mid-rapidity and fully integrated data were used. We note, however, see below, that the fit parameters T and HB obtained at 158 A GeV are very close to those determined earlier. The criterion for the best fit of the model to data was a minimum in p del (13)
_<> y(^ -izr )\ .
of
In the above equations TZ™odel and 7£®xp- are the ith particle ratio as calculated from our model or measured in the experiment, and o^ represent the errors (including systematic errors where available) in the experimental data points as quoted in the experimental publications. For the data we used all information available including that presented at the QM2002 conference in July 2002. Details on the data selection, corrections for the weak-decay reconstruction efficiency, as well relevant references are found in Ref. (59). Under the conditions discussed above the data can all be well described, as is detailed below, by a thermal distribution with T and fis as independent parameters. There is no need to introduce additional parameters such as as strangeness suppression factors. The results of the fits for central Pb-Pb collisions at 40 and 158 GeV per nucleon are presented in Figs. (5,6). At 40 GeV/nucleon 11 particle ratios are included in the fit, while the number is 24 at 158 GeV/nucleon. We obtain values for (T,/xs) of (148±5, 400±10) and (170±5, 255±10), respectively, with reduced x2 values of 1.1 and 2.0. Obviously the fits are quite good. A possible exception is the (I>/(TT+ + 7r~) ratio at top SPS energy, where there are conflicting
508
P. Braun-Munzinger,
K. Redlich and J. Stachel
(0
o
p/p
<0
A/A
E/S
7C-/TC+
K"/K+
KVTT
p/re" K*°/h" K*°/h"
*T4*
-*10
r
* STAR o PHENIX a PHOBOS A BRAHMS
I i
10 Fig. 7. Comparison between thermal model predictions 3 5 and experimental particle ratios for P b - P b collisions at y/snn = 130 GeV. Calculations were performed for T = 174 MeV and / i S = 46 MeV.
data from NA49 and NA50. This is already discussed in detail in Ref. (38) and no new information on this problem has appeared since. Note that this ratio has not been used in the x 2 minimization. The somewhat larger values of x 2 at full SPS energy and the remaining uncertainty in fig are due to a systematic problem not yet sufficiently addressed by the experiments. The contribution of weak decays of strange baryons to final baryons has been discussed and cuts are applied to reduce this contribution in the data. However, there is also a contribution from weak decays to charged pions (or, more generally) charged hadrons which is up to now poorly quantified by the experiments. If, e.g., in the ratio k/h~, feeding of the A by decays of E's is suppressed due to cuts, but the 7r~ measurement has a 50 % efficiency for detection of pions from weak decays, the ratio would drop by 15-20 %, compared to the case with 0 % eficiency for weak decays. With this option the reduced x 2 value for the 158 GeV fit would drop from 2.0 to 1.5. This discussion indicates that there are sources of systematic uncertainties not
Particle Production
in Heavy Ion Collisions
509
included in the data. The corrections for weak decays are, consequently, of utmost importance when discussing the "precision" of fits.
p/p A/A 1 / 5 £J/njr/jt*K7K*K7?r p/irK °/h" ifj/h" A/h" =/hfi/n'*10
= -
•*• O • A
STAR PHENIX PHOBOS BRAHMS ^s^=130GeV
— ~
*
"#*"
Model re-fit with all data T = 176MeV, m, = 41MeV
p/p K"/K*K7ir p/icQ/h'*50
~*~
T
f +
4-
\Js^]=200 GeV Model prediction for T = 177MeV, |i,,=29MeV
Fig. 8. Comparison of the experimental d a t a on different particle multiplicity ratios obtained at RHIC at I/SNN = 130 and 200 GeV with thermal model calculations. The thermal model analysis is from Refs. (35, 36) and recent update by D. Magestro.
The results for RHIC energies are shown in Figs. (7, 8). In Fig. (7) we present the results as published in Ref. (35) in the summer of 2001. Since then, the data at ^snn — 130 GeV have been consolidated and extended and first (in some cases still preliminary) results have been provided for i/snn = 200 GeV. The current state of affairs in summarized in Fig. (7). The results demonstrate quantitatively the high degree of equilibration achieved for hadron production in central Au-Au collisions at RHIC energies. We obtain values for (T,/i B ) of (174±7, 46±5) and (177±7, 29±6), respectively, with reduced x 2 values of 0.8 and 1.1. We note that ratios involving multistrange baryons are well reproduced as is the 4>/h~ ratio. Even relatively wide resonances such as the K*'s fit well into the picture of chemical freezeout. This obviates the need for quark coalescence models as proposed in Ref. (67) and non-equilibrium models as proposed in Ref. (68). Very recently, the STAR collaboration has provided 69 first data, with about 30 - 50 % accuracy, on the p°/7r and /°(980)/7T ratios in semi-central Au-Au collisions. These mesons have been reconstructed in STAR via their decay channel in 2 charged pions. Comparing the preliminary results from STAR with our thermal model prediction reveals that the measured ratios exceed the calculated values by about a factor of 2. This is quite surprising,
510
P. Brawn-Munzinger,
K. Redlich and J. Stachel
especially considering that we use a chemical freeze-out temperature of 177 MeV for the calculation, while one might expect these wide resonances to be formed near to thermal freeze-out, i.e. at a temperature of about 120 MeV. At this temperature, the equilibrium value for the p°/n ratio is about 4 • 10~ 4 , while it is 0.11 at 177 MeV. Even with a chemical potential for pions of close to the pion mass and taking into account the apparent (downwards) mass shift of 60 - 70 MeV for the p° it seems difficult to explain the experimentally observed value of about 0.2. We finally note that the model discussed here was also applied to the AGS data collected in Ref. (37). The best fit, obtained for Rbaryon=Rmeson=0.3 fm, yields T = 125 (+3-6) MeV and /z B = 540 ± 7 MeV, well in line with the calculations reported in Ref. (37). In summary, hadron multiplicities produced in central nucleus-nucleus collisions in the range of AGS to full RHIC energy can be quantitatively described with a grand-canonical partition function based on the full hadron resonance spectrum, assuming complete chemical equilibrium. There is no need to introduce non-equilibrium parameters or strangeness suppression factors if data near mid-rapidity are considered. The physical relevance of the two model parameters T and HB is described in detail in our discussions below concerning the phase boundary between hadrons and the quark-gluon plasma. 2.3. Comparison of measured particle thermal model predictions
densities
with
As discussed below, the value for the energy density predicted by the presently used thermal model, including the excluded volume correction, agrees well with results from the lattice for temperatures below the critical temperature. It makes therefore sense to compare the densities for pions and nucleons predicted by the model with values determined from experiments. The CERES collaboration has recently performed an analysis of 2-pion correlation experiments for the energy range between AGS and RHIC, from which values for these densities have been determined 70 ' 71 from data taken at mid-rapidity. For the nucleon density (at thermal freeze-out) the experimental numbers are, at 40 and 158 GeV/nucleon e and at ^Jsnn = 130 e
We take here the d a t a published in Ref. (66); the data reported in Ref. (72) are about 20 % lower and would not fit the beam energy systematics.
Particle Production
in Heavy Ion Collisions
511
GeV, 0.077 ±0.005/fm 3 , 0.063 ±0.005/fm 3 and 0.06 ±0.009/fm 3 . From the model we deduce, at chemical freeze-out, values of 0.10/fm3, 0.10/fm3 and 0.08/fm3. This would imply a volume increase of about 40 % from chemical to thermal freeze-out. For pions the situation could be more complicated since yield ratios involving pions are (apparently) fixed at chemical freezeout, implying the build-up of a pion chemical potential between chemical and thermal freeze-out. From the data one deduces 70 ' 71 a pion density at (thermal) freeze-out of 0.28 ±0.03/fm 3 , 0.43 ±0.03/fm 3 , and 0.49 ±0.1/fm 3 , at 40 and 158 GeV/nucleon and at ^snn = 130 GeV. These values should be contrasted with the calculated (chemical) freeze-out values of 0.35/fm3, 0.59/fm3 and 0.62/fm3. From these numbers one would conclude a 30 % volume increase between chemical and thermal freeze-out, assuming that the pion chemical potential fixes the pion number to the value obtained at chemical freeze-out. This rather small volume increase indicates that the time between chemical and thermal freeze-out cannot be very long at SPS and RHIC energies. At AGS energy, the corresponding TT+ and proton densities of 0.051/fm3 and 0.053/fm3 agree well with those estimated 73 ' 74 from particle interferometry (0.058/fm3 and 0.063/fm3, respectively) implying that, at AGS energy, thermal and chemical freeze-out take place at nearly identical times and temperatures.
2.4. Statistical
model and composite
particles
An often overlooked aspect of the thermal model is the possibility to compute also the yields of composite particles. For example, the d/p and d/p ratios measured at SPS and AGS energies are well reproduced 97 with the same parameters which are used to describe 37 ' 38 baryon and meson ratios. Furthermore, the AGS E864 Collaboration has recently published 95 yields for composite particles (light nuclei up to mass number 7) produced in central Au-Au collisions at AGS energy near mid-rapidity and at small pt. In this investigation, an exponential decrease of composite particle yield with mass is observed over 7-8 order of magnitude, yielding a penalty factor Pp of about 48 for each additional nucleon. Extrapolation of the data to large transverse momentum values, considering the observed mass dependence of the slope constants, reduces this penalty factor to about 26, principally because of transverse flow. In the thermal model, this penalty factor can be related with thermal particle phase-space. In the relevant Boltzmann
512
P. Braun-Munzinger,
K. Redlich and J. Stachel
approximation, we obtain „
m ± Lit,
Rp~exp—^,
,„ ,.
(14)
where m is the nucleon mass and the negative sign applies for matter, the positive for anti-matter. Small corrections due to the spin degeneracy and the A 3//2 term in front of the exponential in the Boltzmann formula for particle density are neglected. Using the freeze-out parameters T=125 MeV and fib = 540 MeV appropriate 37 for AGS energy one gets 97 R p « 23, in close agreement with the data for the production of light nuclei. It was also noted that the anti-matter yields measured 96 by the E864 Collaboration yield penalty factors of about 2-105, again close to the predicted 97 value of 1.3-105 . This rather satisfactory quantitative agreement between measured relative yields for composite particles and thermal model predictions provides some confidence in the predictions for yields of exotic objects produced in central nuclear collisions. We briefly comment here on the results obtained in Ref. (97). In this investigation, the production probabilities for exotic strange objects and, in particular, for strangelets were computed in the thermal model. The results are reproduced in Table 1 for temperatures relevant for beam energies between 10 and 40 GeV/nucleon. We first note that predictions of the thermal model and, where available, the coalescence model of Ref. (98) agree (maybe surprisingly) well particularly for lighter clusters. Secondly, inspection of Table 1 also shows that, in future high statistics experiments which will be possible at the planned 180 new GSI facility, multi-strange objects such as 7?oA^He should be experimentally accessible with a planned sensitivity of about 10~ 13 per central collision in a years running, should they exist and be produced with thermal yields. Investigation of yields of even the lightest conceivable strangelets will be difficult, though. 3. Exact Implementation of the conservation laws in the statistical models The analysis of particle yields obtained in central heavy ion collisions from AGS up to LHC energy has shown that hadron multiplicities are very well described by assuming a complete thermalized state at fixed T and \IB • In this broad energy range, particle yields and their ratios are, within experimental error, well reproduced by the statistical hadron resonance gas model
Particle Production
in Heavy Ion Collisions
513
Table 1. Produced number of nonstrange and strange clusters and of strange quark matter per central Au+Au collision at AGS energy, calculated in a thermal model for two different temperatures, baryon chemical potential Hb= 0-54 GeV and strangeness chemical potential ns such that overall strangeness is conserved. Thermal Model Parameters Particles
d t+ 3 He a
Ho 5
AA Hn AA 6 He EOAAHe
J°st-8 fst- 91 1 ^St"
"st- 13
fst- 16
T=0.120 GeV 15 1.5 0.02 0.09 3.5 1 0 - 5 7.2 -lO" 7 4.0 l O - 1 0
1.6 1.6 6.2 2.4 9.6
10-14 10-17 10-21 10-24 10-31
T=0.140 GeV 19 3.0 0.067 0.15 2.3 1 0 - 4 7.6 1 0 - 6 9.6 1 0 - 9
7.3 1.7 1.4 1.2 2.3
Coalesce 11.7 0.8 0.018 0.07 4-10- 4 1.610-5 4 10-8
lO-13 10"15 10-18 10-21 10-27
Source: The Coalescence model predictions in the last column are from Table 2 of Ref. (98).
that accounts for the conservation laws of baryon number, strangeness and electric charge in the grand canonical ensemble. The natural question arising here is whether this statistical order is a unique feature of high energy central heavy ion collisions or is it also there at lower energies as well as in hadron-hadron and peripheral heavy ion collisions. To address this question one needs, however, to stress that when going beyond high energy central heavy ion collisions the grand canonical statistical operator (3) has to be modified. Within the statistical approach, particle production can only be described using the grand canonical ensemble with respect to conservation laws, if the number of produced particles that carry a conserved charge is sufficiently large. In view of the experimental data this also means that the event-averaged multiplicities are controlled by the chemical potentials. In this description the net value of a given charge (e.g. electric charge, baryon number, strangeness, charm, etc.) fluctuates from event to event. These fluctuations can be neglected (relative to the squared mean particle multiplicity) only if the particles carrying the charges in question are
514
P. Braun-Munzinger,
K. Redlich and J. Stachel
abundant. Here, the charge will be conserved on the average and the grand canonical description developed in the last section is adequate. In the opposite limit of low production yield the particle number fluctuation can be as large as its event averaged value. In this case charge conservation has to be implemented exactly in each event 99 ' 101 . The exact conservation of quantum numbers introduces a constraint on the thermodynamical system. Consequently, the time dependence and equilibrium distribution of particle multiplicity can differ from that expected in the grand canonical limit. To see these differences one needs to perform a detailed study of particle equilibration in a thermal environment. To discuss equilibration from the theoretical point of view one needs to formulate the kinetic equations for particle production and evolution. In a partonic medium this requires, in general, the formulation of a transport equation 102 ' 103 ' 104 involving colour degrees of freedom and a non-Abelian structure of QCD dynamics. In the hadronic medium, on the other hand, one needs 99 ' 100 ' 101 ' 105 ' 107 to account for the charge conservations related with the U(l) internal symmetry. 3.1. Kinetics particles
of time evolution
and equilibration
of
charged
In this section we will discuss and formulate the kinetic equations that include constraints imposed by the conservation laws of Abelian charges related with U(l) internal symmetry. We will indicate the importance of the conservation laws for the time evolution and chemical equilibration of produced particles and their probability distributions. In particular, we demonstrate that the constraints imposed by the charge conservation are of crucial importance for rarely produced particle species such as for particles with hidden quantum numbers like e.g. for J/ip. To study chemical equilibration in a hadronic medium we introduce first a kinetic model that takes into account the production and annihilation of particle—antiparticle pairs cc carrying U(l) quantum numbers like strangeness or charm. It is also assumed that particles c and c are produced according to a binary process ab —> cc and that all particle momentum distributions are thermal and described by the Boltzmann statistics. The charge neutral particles a and b are constituents of a thermal fireball with temperature T and volume V. We will consider the time evolution and equilibration of particles c and c inside this fireball, taking into account the
Particle Production
in Heavy Ion Collisions
515
constraints imposed by the U(l) symmetry. First, we formulate a general master equation for the probability distribution of particle multiplicity in a medium with vanishing net charge and consider its properties and solutions. Then we will discuss two limiting cases of abundant and rare particle production. Finally, the rate equation will be extended to a more interesting situation where there are different particle species carrying the conserved quantum numbers inside a thermal fireball that also has a non vanishing net charge.
3.1.1. Kinetic master equation for probabilities Consider P^c (r) as the probability to find Nc particles c, where 0 < Nc < oo. This probability will obviously change in time owing to the production ab —> cc and absorption cc —*• ab processes. The equation for the probability P/vc contains terms which increase in time, following the transition from Nc — 1 and Nc + 1 states to the iVc state, as well as terms which decrease since the state Nc can make transitions to iVc + 1 and Nc — 1 (see Fig. 9).
- > / \
• %.
:M-bl! )
s
/
Fig. 9. A schematic view of the master equation for the probability PN{T) due to ab <-> cc and the inverse process.
The rate equation is determined by the magnitude of the transition probability per unit time due to the production G/V and the absorption L/V of cc pairs through ab <-> cc process. The gain (G = < oab-*ccVab >) and the loss (L =< o-Cc^abVCc >) terms represent the momentum average of particle production and absorption cross sections. The transition probability per unit time from Nc + 1 —> iVc is given by the product of the probability L/V that the single reaction cc —> ab takes place multiplied by the number of possible reactions which is formally, (A^c + l)(JVg + 1). In the case when the charge carried by particles c and c is exactly and locally conserved, that is if (Nc + Ns = 0), this factor is just (Nc + l ) 2 . Similarly, the transition probability from iVc —> ./Vc + 1 is described by G(Na)(Nb)/V, where one assumes that particles a and b are
516
P. Bmun-Munzinger,
K. Redlich and J. Stachel
not correlated and their multiplicity is governed by the thermal averages. One also assumes that the multiplicity of a and b is not affected by the ab —> cc process. The master equation for the time evolution of the probability PNC(T) can be written" in the following form:
- ^(Na)(Nb)PNc
-~N2cPNc.
(15)
The first two terms in Eq. (15) describe the increase of PNC(T) due to the transition from Nc - 1 and Nc + 1 to the Nc state. The last two terms, on the other hand, represent the decrease of the probability function due to the transition from Nc to the iVc + 1 and Nc — 1 states, respectively. For a thermal particle momentum distribution and under the Boltzmann approximation the thermal averaged cross sections are obrained 32 ' 31 from
_ 0 C dtaab^cs(t)[t2 - (m+ )2}[t2 < °ab^c-cvab >-
8
mlmlK2(Pma)K2((3mb)
(m^)2]Ki{Pt) A b)
'
where K\, K2 are modified Bessel functions of the second kind, m^b = m a + mb and m~b = ma — mb, t = -^/s is the center-of-mass energy, /3 the inverse temperature, vab = ((kakb)2 — rri^ml)/EaEb is the relative velocity of incoming particles and the integration limit is taken to be to = max[(ma + mb), (m c + m s )]. The rate equation for probabilities (15) provides the basis to calculate the time evolution of the momentum averages of particle multiplicities and their arbitrary moments. Indeed, multiplying the above equation by 7VC and summing over iVc, one obtains the general kinetic equation for the time evolution of the average number (iVc) = Y^N =O •^c-fVc(r) of particles c in a system. This equation reads:
The above equation cannot be solved analytically as it connects particle multiplicity (Nc) with its second moment (N%). However, solutions can be obtained in two limiting situations: i) for an abundant production of c particles, that is when (Nc) » 1 or ii) in the opposite limit of rare particle
Particle Production
in Heavy Ion Collisions
517
production corresponding to (iVc)
(18)
where {SN2) represents the fluctuations of the number of particles c, one can make the following approximations: i) for (Nc) > 1 one has a (N2) « (Nc)2, and Eq. (17) obviously reduces to the well known 31 form:
^
« f W W ) - ^(Nc)2.
(19)
ii) however, for the rare production, particles c and c are strongly correlated and thus, for (Nc)
(20)
where the absorption term depends only linearly, instead of quadratically, on the particle multiplicity. From the above it is thus clear that, depending on the thermal conditions in the system (that is its volume and temperature), we are getting different results for the equilibrium solution and the time evolution of the number of produced particles c. This is very transparent when solving the rate equations (19) and (20). In the limit when (Nc) 3> 1, the standard Eq. (19) is valid and has the well known solution 99 ' 31 : (NC(T))=
(AQeqtanh
(T/T0),
(21)
where the equilibrium value {Nc)eq of the number of particles c and the relaxation time constant T 0 are given by: (Nc)eq = V~e , r 0 = - ^ = ,
(22)
respectively, with e = G{Na){Nb)/L. In the particular case when the particle momentum distribution is thermal, the ratio of the gain (G) to the loss (L) terms can be obtained " from Eq. (16) as G L
=
dcm2cK2(mc/T)d-cm2sK2(m-c/T) damlK2{ma/T)dbmlK2{mb/TY
[
>
518
P. Braun-Munzinger,
K. Redlich and J. Stachel
where we have employed the detailed balance relation between the cross sections for production oa\, and for absorbtion acB for ab <-> cc processes
with di being the spin-isospin degeneracy factor and m^ as in Eq. (16). In Boltzmann approximation, the equilibrium average number of particles c in Eq. (22) reads:
W e q = ^VTmlK2{mc/T).
(25)
This is a well known result for the average number of particles in the Grand Canonical (GC) ensemble with respect to the U(l) internal symmetry of the Hamiltonian. The chemical potential, which is usually present in the GC ensemble, vanishes in this case, because of the requirement of charge neutrality of the system. Thus, the solution of Eq. (19) results in the expected value for the equilibrium limit in the GC formalism where a charge is conserved on the average. In the opposite limit, where (Nc) -C 1, the time evolution of a particle abundance is described by Eq. (20), that has the following solution: (iV c (r)) c = ( 7 V c ) e c q ( l - e - ^ o c ) ,
(26)
with the equilibrium value and relaxation time given by (^c)eCq = e,T 0 C = ^ .
(27)
The above result, as will be shown in the next section, is the asymptotic limit of the particle multiplicity obtained in the canonical (C) formulation of the conservation laws 99 ' 100 . Here the charge related with the U(l) symmetry is exactly and locally conserved, contrary to the GC formulation where this conservation is only valid on the average. Comparing Eq. (22) with Eq. (27), we first find that, for (iVc) < 1, the equilibrium value is by far smaller than what is obtained in the grand canonical limit, i.e. (iVc)ecq = <JVc)e
2
« (Nc),
(28)
Particle Production
in Heavy Ion Collisions
519
Secondly, we can conclude that the relaxation time for a canonical system is shorter than the grand canonical value, i.e. ^
= T0(NC)^
«
r0,
(29)
since in the limit (ii) the equilibrium value (ATc)eq
^ T
= ^(Na){Nb)PNc^
+ ±(NC +
~ ^(Na)(Nb)PNc
- LNcmpNc.
l)(N5)PNc+1
(30)
Multiplying the above equation by Nc, summing over iVc and using the condition that (iVc) = (7Ve), one recovers Eq. (19), the rate equation for
520
P. Braun-Munzinger,
K. Redlich and J. Stachel
(Nc) in the GC ensemble. The above equation is thus indeed the general master equation for the probability function in the GC limit. Comparing this equation with the more general Eq. (15), one can see that the main difference is contained in the absorption terms that are linear in particle number instead of being quadratic . Eq. (30) can be solved exactly. Indeed, introducing the generating function g(x,r) for P/vc,
g(x,T)=
f>^Pjvc(r),
(31)
Nc=0
the iterative equation (30) for the probability can be converted into a differential equation for the generating function:
^
= £v^l-*)b'-v^],
(32)
with the general solution": g(x, T) = g0(l - xe~T) exp\y/e(l - x)(e~T - 1)],
(33)
where g' — dg/dx, f = {Ly/l/V)r and yfi = (Nc)eq given by Eq. (9). One can readily find out an equilibrium solution to the above equation. Taking the limit T = oo in the Eq. (33) leads to g eq (x) = exp[-\/e(l - x)\,
(34)
with the corresponding equilibrium multiplicity distribution: PNc,eq = ^ f f
1
^ .
(35)
This is the expected Poisson distribution with average multiplicity i/e. 3.1.2. The equilibrium solution of the general rate equation The master equation (30), that describes the evolution of the probability function in the GC limit, could be solved analytically. The general equation (15), however, because of the quadratic dependence of the absorption terms, requires a numerical solution. Nevertheless, the equilibrium result for the particle multiplicity can be given.
Particle Production in Heavy Ion Collisions
521
Converting Eq. (15) for P/vc into a partial differential equation for the generating function N
9{X,T)=YJX
(36)
Nc=0
one finds"
W?lll = ±(l-x){xg»
+ g>-e9).
(37)
The equilibrium solution ge
(38)
By a substitution of variables (x = y 2 e/4), this equation is reduced to the Bessel equation, with the following solution:
Seq(z) =
^I0{2y/S),
(39)
where the normalization is fixed by g(l) = 2Z-PJVC — 1The equilibrium value for the probability function PJVC is now written from Eqs. (36-39) as:
P/
^eq
=
/0(2^)(Ay)2-
(40)
We note that the equilibrium distribution of the particle multiplicity is not Poissonian. This fact was indicated first in equilibrium studies in Ref. (109). In our case this is a direct consequence of the quadratic dependence on the multiplicity in the loss terms of the master equation (15). The Poisson distribution is obtained from Eq. (40) if -y/e 3> 1, that is for large particle multiplicity where the C ensemble coincides with the GC asymptotic approximation. In Fig. (10) we compare the Poisson distribution from Eq. (35) with the distribution from Eq. (40) for two values of y/e. The result for the equilibrium average number of particles c can be obtained as:
<"•>--»'(" = ^ 5 ^ -
<41)
The above expression will be shown in the next section to coincide with the one expected for the particle multiplicity in the canonical ensemble with
522
P. Braun-Munzinger,
K. Redlich and J. Stachel
Fig. 10. T h e probability function from Eqs. (35,40) for two values of e = 4 and 16. T h e full lines represent Poisson distribution.
respect to U(l) charge conservation 105 ' 106 . The rate equation formulated in Eq. (15) is valid for arbitrary values of (Nc) and obviously reproduces (see Eqs. (101 - 104)) the standard grand canonical result for a large (Nc). Thus, within the approach developed above one can study the chemical equilibration of charged particles following Eq. (15), independent of thermal conditions inside the system. 3.1.3. The master equation in the presence of the net charge. So far, in constructing the evolution equation for probabilities, we have assumed that there is no net charge in the system under consideration. For the application of the statistical approach to particle production in heavy ion and hadron-hadron collisions, the above assumption has to be extended to the more general case of non-vanishing initial values of conserved charges. In the following we construct the evolution equation for P^ (t) in a thermal medium assuming that its net charge S is non-vanishing. The presence of a non-zero net charge requires modification of the absorption terms in Eq. (15). The transition probability per unit time from
Particle Production in Heavy Ion Collisions
523
the Nc to the Nc - 1 state was proportional to (L/V)NCNC. Admitting an overall net charge 5 ^ 0 the exact charge conservation implies that Nc - N£ = S. The transition probability from Nc to Nc - 1 due to pair annihilation is thus (L/V)NC(NC - S). Following the same procedure as in Eq. (15) one can formulate the following master equation for the probability Pjv (£) to find Nc particles c in a thermal medium with a net charge S: dPs C - ^ = yiNJiNJP^
T + -(Nc + 1)(NC + 1 -
-^(Na)(Nb)Pl
S)Pl+1
-LNc{Nc-S)Pl,
(42)
which obviously reduces to Eq. (15) for 5 = 0. To get the equilibrium solution for the probability and multiplicity, we again convert the above equation to the differential form for the generating
function
S 9
(X,T)
=
EN^O^PN^)-
£fel) = L{1 _ x) {xg>> + ^ ( i _ S) - e9s). In equilibrium, dgs(x,T)dr follows:
(43)
= 0 and the solution for g^ can be found as
rS/2
*W = 5575 , * (2 ^ 5) '
(44)
where the normalization is fixed by g(l) = ^ P „ = 1. The master equation for the probability to find TVg antiparticles c, its corresponding differential form and the equilibrium solution for the generating function can be obtained by replacing 5 with —S in Eqs. (42-44) The result for the equilibrium average number of particles (A r c ) eq and antiparticles (Ns)eci is obtained from the generating function using the relation: (iVc)eq = '(1)- The final expressions read:
(Nc)eq = V?
Is{2V-e)
,
I
s
M
•
(45)
The charge conservation is explicitly seen by taking the difference of these equations that results in the net value of the charge S.
524
P. Braun-Munzinger,
K. Redlich and J. Stachel
The thermal average values of the particle number given through Eq. (45) will be later derived from the equilibrium partition function by using the projection method 106 ' 107 . 3.1.4. The kinetic equation for different particle species The rate equations discussed until now were derived assuming that there is only one kind of particle c and its antiparticle c that carry conserved charge. To study equilibration of particles in a strongly interacting environment one also needs to include processes that involve different species. In low energy heavy ion collisions e.g. the K+ mesons are not only produced in pairs together with K~ but also with the strange hyperon A or E° due to the TTN —> KK+ process. The contribution of K° and K° has to be also included as these particles are produced with similar strength as charged kaons. To account for this situation one generalizes the rate equations described in the last sections. Consider PNK,NA(T) ^ t n e probability to find NK and NA number of K~ mesons and A baryons. Including the production and absorption processes such as: mm —> K+K~ and mN —> K+A this probability will obviously change in time. Here m(N) denotes a meson (nucleon). Following a similar procedure as was explained in Fig. (9) the master equation for the time evolution of the probability PNK,NA(T) can be written as.100,101
NK,NA
Gm
{Nm){Nrn)PNK^Nx
+ ^(NK
+ 1)(NK + 1 + NA)
XPNK+I,NA
- ^(Nm)(Nm)PNKtNA
- ^NK(NK
+ ^(Nm)(NN}PNK,NA^
+ ^(NA
+
NA)PNKtNA
+ 1)(NA + 1 + NK)
XPNK,NA+I
- ^r(Nm)(NN)PNK,NA
(46)
- ^-NA(NA
+
NK)PNK>NA.
with Gm and Lm being the production and absorption terms for the mm ^ K+K~ reaction and G^ and L^ denote equivalent terms for the mN ^ K+A process.
Particle Production
in Heavy Ion Collisions
The equilibrium solution for the probability function found as follows101: NK NA
'
+ iVA)!)2
IO(2^)((NK
525
can be
PNKNA
(47)
^K+NANkim
with etot = e m + CN and em(Ar) = G f m ( W )(A r mi ( m ))(./V m2 (;v))/L m ( N ). The equilibrium probability distribution is thus, according to Eq. (47), the product of the distribution of the number of pairs (NK + NA) and a binomial distribution that determines the relative weight of the individual particles, in our case the K~ and A. The probability Pitj is obviously normalized such that £V • Pitj = 1. The equilibrium value for the multiplicity (Ni) with i = K~ or i = A can be obtained as: eTO h{1y/tto~t)
("K" )eq = —
T frt
^
•,
V^tot h{1y/tto~t)
,N .
,
_
(JVA/eq -
'
e
£iv I\{^^/^to~t) - 7 ;
yfctot
T
,„
^
(48)
N
Io{1y/Uo~t)
with e m , CAT and etot defined as above. The average value of K+ can be obtained applying strangeness conservation leading to:
(NK+)eq = (NK-)eq
+ {NA)eq.
(49)
The results presented here can be extended 101 to an even more general case where there is an arbitrary number of different particle species carrying the quantum numbers related with U(l) symmetry of the Hamiltonian. 3.2. The canonical description projection method
of an internal
symmetry
-
Using the above kinetic analysis of charged particle production probabilities we have demonstrated that equilibrium distributions does not necessarily coincide with the GC value. It is thus natural to ask what is the corresponding partition function that can reproduce the kinetic results obtained in Eqs. (41,45,48). The main step in deriving these equations was an assumption of an exact conservation of quantum numbers in the kinetic master equations (15,42,47). Thus, one should account for this important constraint in constructing the partition function.
526
P. Braun-Munzinger,
K. Redlich and J. Stachel
The exact treatment of quantum numbers in statistical mechanics has been well established 105,106 for some time now. It is in general obtained 107 ' 108 by projecting the partition function onto the desired values of the conserved charge by using group theoretical methods. In this section we develop these methods and show how one gets the partition function that accounts for exact conservation of quantum numbers. The derivation will be not only restricted to the charge conservation related with an Abelian U(l) internal symmetries and their direct products, but it will include also symmetries that are imposed by any semi-simple compact Lie group. The usual way of treating the problem of quantum number conservation in statistical physics is by introducing the grand canonical partition function, as in Eq. (3). For only one conserved charge, e.g. strangeness 5, Z(jiS, T) = -nfc-WA-MsS)]
(50)
The chemical potential fis is then fixed by the condition that the average value of strangeness of a thermodynamical system is conserved and has the required value (S) such that: =
TdlnZ^s,T)
This method, as shown in the previous sections, is only adequate if the number of particles carrying strangeness is very large and their fluctuations can be neglected. In order to derive a partition function that is free from the above requirements let us first reorganize Eq. (50). Denoting the states under the trace as |s) such that H]s) = Es\s) and 5js) = s\s) one writes s=+oo
s=+oo
Z(»s,T)= Y, e-W-e*"" = £ s~~ oo
where we have introduced the fugacity As = e^s ZS =Trs[e-06]
ZSX%
(52)
s= — oo
and where (53)
is just the partition function that is restricted to a specific total value S of the conserved charge. This is the canonical partition function with respect to strangeness conservation. Thus, Zs is a coefficient in the Laurent series in the fugacity. Our goal is to calculate Zs- This is an easy task: starting from
Particle Production
in Heavy Ion Collisions
527
Eq. (52) we apply the Cauchy formula and take an inverse transformation to obtain
z T v
^ > )=hi§^z{Xs'T'v)
(54)
Choosing the integration path as the unit circle and parameterizing it as Xs = exp (i>) we can convert the contour integral into the angular one as ZS(T,V)
= |+*^Z(cj>,T,V)
(55)
where the generating function Z(
= Ti[e-^PS]
(56)
where Ps = P$ is the projection operator on the states with the exact value of S. For an Abelian symmetry, Ps is the 5-function Ps = Ss s. Introducing the Fourier decomposition of delta into Eq. (56) one can reproduce the projected result (55). The conservation of additive quantum numbers like baryon number, strangeness, electric charge or charm is related to the invariance of the Hamiltonian under the U(l) Lie group. In many applications it is important to generalize the projection method to symmetries that are related with a non-Abelian Lie group G. An example is the special unitary group SU(N) that plays an essential role in the theory of strong interactions. Generalization of the projection method would require to specify the projection operator or generating function. Consequently, the partition function obtained with the specific eigenvalues of the Casimir operators that fixes the multiplet of the irreducible representation of the symmetry group G could be determined. To find the generating function for the canonical partition function with respect to the symmetry group G, let us introduce the quantity Z(g) via
528
P. Braun-Munzinger,
K. Redlich and J. Stachel
Z(g)=Tr[U(g)e-eH].
(57)
This expression is a function on the group G with U(g) being the unitary representation of the group with g C G. The quantity U(g) can be decomposed into irreducible representations Ua(g)
0 U(g)=Y,Ua(9)
(58)
a
where a is labelling these representations. From Eq. (57) and (58) one has Z(g) =
^Tra[Ua(g)e-W} a
= E
E
<"<»& I U^9)e~0ii
| va,Za)
(59)
where ua labels the states within the representation a and £Q are degeneracy parameters. Introducing the unit operator 1 =|)(| into the above equation the expression factorizes %) = E
E
<"«»'&* I U°(9) I "cn<£a)K,£« I e~pk
= E E ( " « i u*(9) i "«>&. ie_/3" i w»
|
ua,ia)
(6°)
where we have used that, due to the exact symmetry, the only non-vanishing matrix elements of e_l3H are those diagonal in va. The matrix elements of Ua(g) are only non-zero if they are diagonal in £Q. Finally, the matrix elements of the Hamiltonian are independent of the states within representation (since due to symmetry they are dynamically equivalent) and those of U(g) of degeneracy factors (since U(g) does not distinguish dynamically different states that transform under the same representation). The last two sums in Eq. (60) can be further simplified as E > « I u°(9) I "«> = TraPaig)]
= xa(9).
(61)
Particle Production
in Heavy Ion Collisions
529
The quantity \a is by definition the character of the irreducible Ua(g) representation and
5> ' ^" ! *"> = ia)^e~Ptl = ia)Z^
V)
'
(62)
where Za is the canonical partition function with respect to the G symmetry of the Hamiltonian and d(a) is the dimension of the representation a. Calculating Za one considers under the trace only those states that transform with respect to a given irreducible representation of the symmetry group. We have thus connected, through Eq. (60) and (61-62), the canonical partition function with the generating functional on the group
2
(9) = ^^-Za(T,V)
(63)
The canonical partition function is the coefficient in the cluster decomposition of the generating function with respect to the characters of the representations. The character functions satisfy the orthogonality relation ^ y
/ dn(g)x*a(g)x~f{9) = <*a,7
(64)
where dfi(g) is an invariant Haar measure on the group. The orthogonality relation for characters allows to find the coefficients, the canonical partition function, in this cluster decomposition. Prom Eq. (63) and (64) one gets
Za(T, V) = d{a) J'dfi(g)x*a{g)Z{g)
(65)
This result is a generalization of Eq. (55) to an arbitrary symmetry group that is a compact Lie group. The formula holds for any dynamical system described by the Hamiltonian H. To find the canonical partition function we have to determine first the generating function Z(g) defined on the symmetry group G. If the symmetry group is of rank r, then the character of any irreducible representation are the functions of r variables {71,.... ,jr}- Denoting as Jk the commuting
530
P. Braun-Munzinger,
K. Redlich and J. Stachel
generators of G with k — 1, ...,r the character function Xa(7i».-,7r) = 5 Z K I e ' ^ i 7 " 7 ' | va)
(66)
is obtained. Here, va labels the state within the representation a. With the above form of the characters we can write Eq. (63) as Z( 7 i,..,7r) =Tr[e-f>A+iT^'«Jt]
(67)
Through the Wick rotation 7$ = — i/J/i, the generating function Z is just the GC partition function with respect to the conservation laws given by all commuting generators of the symmetry group G. The equations (65) and (67) are the basis that permits to obtain the canonical partition function for systems restricted to any symmetry. The simplicity of the projection formula (65) is that the operators that appear in the generating function are additive they are generators of the maximal Abelian subgroup of G. Thus, the problem of extracting the canonical partition function with respect to an arbitrary semi-simple compact Lie group G is reduced to the projection onto a maximal Abelian subgroup of G. The calculation of the generating function from the Eq. (67) can be done applying standard perturbative diagrammatic methods or a mean field approach. However, if interactions can be omitted or effectively described by a modification of the particle dispersion relations by implementing an effective mass, then the trace in Eq. (67) can be worked out 107 exactly, leading to
^)=-PE|f 2 a]
(68)
where 7 = (71,. .,7,-) and Z\ — J(gVdp/2n2)p2 exp (—\/P 2 + m\/T) is just the thermal particle phase-space in Boltzmann approximation belonging to a given irreducible multiplet of a symmetry group G. The sum is taken over all particle representations that are constituents of the thermodynamical system. 3.2.1. Canonical models with a non-Abelian
symmetry
To illustrate how the projection method described above works, we discuss a statistical model that accounts for the canonical conservation of nonAbelian charges related with the 5£/ c (N)xt/ B (l) symmetry with N = 3
Particle Production
in Heavy Ion Collisions
531
and B being the baryon number and c denoting the global gauge colour symmetry. Let us consider a thermal fireball that is composed of quarks and gluons at temperature T and volume V . We describe the canonical partition function that is projected on the global color singlet and exact value of the baryon number. The interactions between quarks and gluons are implemented effectively, resulting in dynamical particle masses that are temperature dependent, e.g. through mq,(g) ~ gT. Since the interactions are only trivially modifying the dispersion relations one can still use the free particle momentum phase-space. Thus, under this assumption, Eq. (68) provides the correct description of the generating function. The sum in the exponents in (68) gets contributions from quarks, antiquarks and gluons that transform under the fundamental (0,1), their conjugate (1,0) and adjoint (1,1) representation of the SUC(N) x £/s(l) symmetry group. Thus,
in z(T, K 7, 7B) = ^ ^ + ; r 4 + ; T Z G
(69)
where 7 = (71, ..,7;v-i) are the parameters of the SUC(N) and 75 of the f/jg(l) symmetry group. Through an explicit calculation of one-particle partition functions for massive quarks and gluons the corresponding generating function is obtained as Qo m20VT ~ lnZ Q (T, V , 7 , 7 B ) = £ ^ 2 ~ £ Q
n=0
( - l ) " + i[ n
-^-K2(mQ/T)
[ e^"/TXQ(n7) + e-^"/TXQ(n7)]
(70)
where the two terms in the bracket represent the contribution of quarks and antiquarks, respectively. The corresponding result for massive gluons reads 2 T/T 1 ° °
lnZ G (T,
V,7,7B)
= J - ~ ^
£
1
^K2(mG/T)[XG(n^
+xh(nj)}
(71)
n=0
where gc9Q and dQ = N, dG = N2 - 1, are respectively, the quark and gluon degeneracy factors and dimensions of the representations. Now we can apply this generating function in the projection formula (65) to get the canonical partition function. Of particular interest is the
532 P. Braun-Munzinger, K. Redlich and J. Stachel
color singlet partition function that represents global colour neutrality (phenomenological confinement) of a quark-gluon plasma droplet. The conjugate character for the SUC(N) singlet representation is particulary simple, ^(0,0) _ -j r j - ^ bgj-yojj n u m D e r D e treated grand canonically requiring a substitution 73 = — i / i s / T in Eq. (70). To find Z one still needs an explicit form of the fundamental and adjoint characters and the Haar measure on the SUC(N) group. Here we quote their structure for the SUC(3) group. The real LR and the imaginary Li parts of the character in the fundamental (quark) representation are LR = cos 71 + cos 72 + cos(7i + 72) Li = sin71 + sin72 - sin(71 + 72). For the adjoint (gluon) representation XG = 2[cos(7i - 72) + cos(27i + 72) + cos(272 + 71) + 1].
(72)
The invariant Haar measure on the SUC(3) internal symmetry group 2 / 7 l - 7 2 s . 2 / 2 7 l + 7 2 N . 2/ 2 72 + 7 l .
^(7i,72) = ^sm'(l^)Sin'(-!^)sm'C-^^).
(73)
From Eqs. (70)-(73) and (65) we write the final result for the SUC(3) color singlet partition function that for non-vanishing baryon chemical potential /is reads ZP(nB,T,V)
= Jd/x(7i,72)
exp{c l X G + c2[LR cosh(pfiB) + iLi sinh(/3)]} (74)
where the constants c\ and C2 can be extracted from Eqs. (70-71). The above partition function shows a complex structure of the integrand. However, due to its symmetry it is straightforward to show that the partition function is real. The thermodynamical properties of this color singlet canonical partition function and other thermodynamical observables can be studied 110 ' 111 ' 112 by a numerical analysis. In finite temperature gauge theory the zero component of the gauge field AQ takes on the role of the Lagrange multiplier guaranteeing that all states satisfy Gauss law. In Euclidean space one can choose a gauge in such a way that AQ(X,T)X1, is a constant in space-time. In such a gauge the Wilson loop defined as 1 L(x) = -TrPexp[igj
f0 A0(X,T)(IT}
(75)
Particle Production
in Heavy Ion Collisions
533
represents the character of the fundamental representation of the SUC(N) group 110 . The effective potential of the SU(N) spin model for the Wilson loop in the above gauge coincides 110,113 essentially with the generating function given in Eq. (74). In addition, this generating function could be also related to the strong coupling effective free energy of the lattice gauge theory with a finite chemical potential 110 . Thus, the effective model formulated above connects the colored quasi-particle degrees of freedom with the Wilson loop. 3.2.2. The canonical partition function for Abelian charges In this section we show how the projection method described above leads to a description of particle yields under the constraints imposed by the Abelian ^ B ( I ) symmetry. In this case the formalism is particularly transparent due to a simple structure of the symmetry group. The U(l) group is of rank one, thus the characters of the representations, numbered by the eigenvalues of the conserved charge B, depend only on one parameter <j>. They are of the exponential type:
xg fl (i)=e iB *-
(76)
For the conservation of a few Abelian charges inside the system like strangeness (S), baryon number (B) or electric charge (Q) and charm (C) one needs to account for the products of the U(l) symmetries: ?7s(l) x £/s(l) x UQ{1) x C/c(l). In this case the characters are numbered by the values of all conserved charges and they are expressed as the products of the corresponding characters of U(l) groups. For simultaneous conservation of baryon number and strangeness the characters read: •yS,B
_
S
B
_
i(Si>+B
(77)
The invariant measure on the E/s(l) x J7s(l) group is just the product of the differentials dfi(<j>s,
534
P. Braun-Munzinger,
K. Redlich and J. Stachel
others are treated using the GC formulation. The corresponding canonical partition functions can be obtained from Eqs. (65,68) as:
Zs
= h l * d
( 7 8)
and -1
ZB,S
= ^
/>Z7T
I
pZTX
dcf>e-iQ* I
diPe-iS^Z(T,V,
(79)
where Z is obtained from the grand canonical (GC) partition function replacing the fugacity parameter Ajg, As by the factors e%^ and e1"^ respectively, Z(T, V, 4>) = ZGC(T, V, XB - e**, Xs - e'*)
(80)
The particular form of the generating function Z in the above equation is model dependent. In applications of the above statistical partition function to the description of particle production in heavy ion and hadron-hadron collisions we calculate Z in the hadron resonance gas model. In our analysis we neglect interactions between a hadron and resonances as well as any medium effects on particle properties. In general, however, already in the low-density limit, the modifications of the resonance width or particle dispersion relation could be of importance 4,19,51,115 . For the sake of simplicity, we use a classical statistics, i.e. we assume a temperature and density regime such that all particles can be treated using Boltzmann statistics. Within the approximations described above and neglecting the contributions of multi-strange baryons, the generating function in Eq. (78), has the following form Z(T, V, tMQ,HB,
+ TV^-ie"^)
(81)
where N3-o,±i is defined as the sum over all particles and resonances having strangeness 0, ± 1 , Nt=0,±i=^2zl
(82) k
and Z\ is the one-particle partition function defined as Z\ = ^ m l T K 2 ( m k / T )
exp(Bk^B
+ QkfiQ)
(83)
Particle Production
in Heavy Ion Collisions
535
with the mass m^, spin-isospin degeneracy factor Qfc, particle baryon number jBfc and electric charge Qk- The volume of the system is V and the chemical potentials related to the charge and baryon number are determined by /XQ and HB, respectively. With the particular form of the generating function (81) the canonical partition function Zs is obtained from Eqs. (78-80) as Zs = Z0±- P dxtKr^W**3-*-"„
(84)
IK JO
where ZQ = exp (Ns=o) is the partition function of all particles having zero strangeness and where we introduce S±\ = Ns=±i with Na=±i defined as in Eq. (82). To calculate the canonical partition function (84) one can expand each term in the power series and then perform the <j> integration 116 . Rewriting the above equation as
and using the following relation for the modified Bessel functions eHt+-t) = Y,tsIs(x),
Is(x), (86)
—oo
one gets after the (^-integration the canonical partition function for a gas with the net strangeness S: ZS(T,V,HB,HQ)
= Z0(T,V^B,fiQ)(-p-)s/2Is(x)
(87)
where the argument of the Bessel function x = 2y/SiS-i.
(88)
The calculation of the particle density tik of species k in the canonical formulation is straightforward. It amounts to the replacement Z\ . - Afc Z\
(89)
of the corresponding one-particle partition function in equation (81) and taking the derivative of the canonical partition function (84) with respect to the particle fugacity Afc
536
P. Braun-Munzinger,
K. Redlich and J. Stachel
n
fc
(90) \k=i
As an example, we quote the canonical result for the density of kaons K+ and anti-kaons K~ in an environment with a net overall strangeness S, Zj,+ S-! IS-i(x) Z'K_ ft ls+1(x) c c HK+ nK ~ V ^ST, Is(x) ~ - V JSlS-i ls(x) ' ( 9 1 ) where x = y/SiS-i and Z1 are as in (83) and (84). For the particular case when S\ = 5_i the above equation coincide with (45). Thus, the master equation (42) represents the rate for the time evolution of the probabilities for which the equilibrium limit corresponds to the canonical ensemble. The partition function (85) and the corresponding results for particle densities (91) were derived neglecting the contribution of multistrange baryons to the generating functional (81). Multistrange baryons are, however, an important characteristics of the collision fireball created in heavy ion collisions. Thus, the canonical formalism described above should be extended to account for these particles. Under the constraints of the global strangeness neutrality condition 5 = 0 and including hadrons with strangeness content s = ± 1 , ±2, ± 3 the canonical partition function in Eq. (84) is replaced 48 ' 116 by
Z$=0 = ^fd
exp ( Yl S ^ i n A .
(92)
where Sn = ^2k Z\ and the sum is over all particles and resonances that carry strangeness n with Z\ defined as in Eq. (83). The integral representation of the partition function in Eq. (92) is not convenient for a numerical analysis as the integrant is a strongly oscillating function. The partition function, however, after <j> integration, can be obtained in a form that is free from oscillating terms. Indeed, rewriting Eq. (92) to
Zcs=0 = i - e s ° | * d4> n
exp [ ^ (ane^
+ o^e"**)]
(93)
Particle Production
in Heavy Ion Collisions
537
and using the relation (86) one finds, after integration 53 oo
oo
5
Zf =0 = e ° Y,
ap3a^2n-3pIn(x2)Ip(x3)I-2n-3p(xi),
E
(94)
n= — oo p— — oo
where ai = jS~fsIi
,
Xi
= 2^S~S~
(95)
and In are the modified Bessel functions. The expression for the particle density, n*, can be obtained from Eq. (90) and Eq. (92). For a particle i having strangeness s n
< = -%5-
E
E
a|a>^ 2 "- 3 p - s / n (a; 2 )/ p ( a ; 3)/_2n-3 P -3(x 1 ). (96)
5=0 n——oop= — oo
In the limit of x-i —» 0 and x3 —> 0 it is sufficient to take only terms with n = 0 and p = 0 in Eq. (94) and (96) 48 . In this case the density of particle n% and antiparticle nf with strangeness content s and s = —s respectively, reads
K i_i
io(^)
V
oi
io(a;)
with a; = -v/'S'i'f>-i and Zj as in (83). The above equation is an approximation and can be only used for a qualitative discussion. The quantitative description of multistrange particle production requires the exact result given in Eq. (94) and (96). 3.2.3. The equivalence of the canonical formalism in the grand canonical limit Discussing the strangeness kinetics in Section 4.1 we have already indicated that the canonical description of the conservation laws is valid over the whole parameter range. The grand canonical formulation, on the other hand, is the asymptotic realization of the exact canonical approach. This can be indeed verified when directly comparing particle densities obtained in the C and GC ensemble. Consider first a thermal system that contains
538
P. Braun-Munzinger,
K. Redlich and J. Stachel
only strangeness 1 particles and their antiparticles. In such an environment the GC result for the strangeness s = ±1 hadrons is obtained from (15) as n?f ± 1 = ^ A ?
1
-
(98)
with the fugacity Xs = exp (fis/T). Comparing the above GC and C result of Eq. (91) with 5 = 0 one sees that nf = ± 1 = nf£±1
(A.) .
(99)
where the effective fugacity parameter
xs = -^L=I-^l.
(loo)
In the limit of large x —» oo the canonical and the grand canonical formulations are equivalent. In the opposite limit, however, the differences between these two descriptions are large. This can be seen in the most transparent way, when directly comparing the two limiting situations of the large and small x in the Eq. (91). For x —> oo
lim £ g | _> i
(101)
and the ratio S^\/-\JS\S~l corresponds exactly to the fugacity Xs in the GC formulation (98). Indeed, a strangeness neutrality condition in the GC ensemble requires that (S) = 0, thus through Eqs. (80-81) one has: AS5X - A ; 1 ^ - ! = 0,
that is As =
(102)
S-i/y/SiS-i.
Thus, neglecting multistrange baryons in the generating functional (92) one gets
»f=±. = » ? A , g g .
(103)
Comparing Eq. (97) with the GC result (98) for the density ns of multistrange particles one finds that n
?*n?°7Wi-
(104)
Particle Production in Heavy Ion Collisions 539 However, one needs to remember t h a t t h e above relation is only valid if a t h e r m a l phase space of all multistrange hadrons is negligibly small. This assumption is, however, questionable particularly when approaching the thermodynamical limit.
I
,
-
Fs
•
1
Canonical
1 .o —
_ - — SPS - m AGS . • Ni-Ni — ay Au-Au — - •a Ni-Ni
0.5
1
suppression
factor
—
f^~~~ / 1 1.8A GaV
-
1
1 . OA GaV f 0.8A GaV J
_ -
O.O
10"
~a
I
.
I
10 1
10" 1
.
10 3
Fig. 11. The canonical strangeness suppression factor (see text for the explanation). The SPS and the AGS values are shown for Pb-Pb and Au-Au collisions, respectively.
Prom Eq. (103) one concludes t h a t the relevant parameter Fs t h a t describes deviations of particle multiplicities from their grand canonical value reads FS
(105)
Io(x)'
T h e largest differences appear in the limit of a small x where lim
h{x)
I-O/0(:E)
x/2
(106)
This limit reproduces t h e solution of our kinetic equation (27) for large particle number fluctuations. T h e argument of the Bessel functions in Eq. (106) describes the size of the thermal phase-space t h a t is available for strange particles. For a system free of multistrange hadrons t h e argument x can be also identified as being
540
P. Braun-Munzinger,
K. Redlich and J. Stachel
proportional to the total number of strange particle-antiparticle pairs in the GC limit. The canonical suppression factor Fs{x) is quantified in Fig. (11). Typical values of x expected for the SIS, AGS and SPS energies for central collisions of different nucleus are also indicated in this figure. Fig. (11) shows the importance of the canonical suppression of particle phase-space at SIS energies where it can even exceed an order of magnitude. In central heavy ion collisions at the AGS and particularly at higher energies (SPS, RHIC, LHC), the canonical suppression is seen in Fig. (11) to be negligible. Thus, here the GC formalism is adequate. In general, the canonical statistical interpretation of the particle production in central heavy ion collisions is important if the CMS collisions energy per nucleon pair becomes less than about (2 — 4) GeV. However, as we show later, the canonical suppression effect can also be important at high energy for non-central heavy ion collisions or for the description of heavy quark production. At the end of this section we also formulate thermodynamics of the canonical ensemble with strangeness and baryon number being exactly conserved. The corresponding partition function was already presented in Eq. (79). Neglecting the contribution from multistrange baryons and the particle-antiparticle charge asymmetry, the generating functional Z(T, V,
(107)
+ 2iV"s==0,6=i cos<^> + 2iV s=1)6= i cos((£ - ip) where NStb is defined as the sum over all particles and resonances having the strangeness s and baryon number b,
NSib = J2Zk
(108)
k
and Z\ = ^^- m\TK2{mk/T) is the thermal phase-space available for a particle that carries strangeness s and baryon number b. The above generating functional can be applied in Eq. (79) to get the canonical partition function of the fireball with net value of strangeness S and baryon number B
ZB,s(T> V) = Z0Z'*V) j * #e_iS0 &XP[ZN C ° S ^ ]
(1
°9)
Particle Production
p2n
/ Jo
in Heavy Ion Collisions 541
di/je~lS^ exp[zK cos tp + zy cos(0 — ip^dcpdip
where Z0 = exp(iVs=0,6=o), ZK = N«=i,&=o. ZN = Ns=0,b=i and zY = Ns=i^=i. This notation indicates the type of particles that carry corresponding quantum numbers: charge neutral hadrons, strange mesons, nonstrange and strange baryons. In the exponent of the ip integral we write Zfc COSIp + Zy COs((f> — tp) = z((j>) COs(tjj — a ( 0 ) )
(HO)
where a(—<j>) = —a{4>) and z
{
eiaW
=
^ K
z{4>)
+
j Y i
zY)1/2
t
(
m
)
z(<j>)
Since the ip integral goes over the whole period, we may shift the integration by a and perform the ip integral exactly to yield ZB,s(T, V) = Z ° ( ^ , y ) / * cos(B
(112)
Is(2z(cP))d(j> However, the <> / integration cannot be solved analytically. Starting from the above partition function one can find the mean multiplicity of particle species i. To get it one simply (i) separates for these species the particle and anti-particle term in Eq. (108), (ii) multiplies the relevant one by A, (iii) differentiates with respect to A following Eq. (90) and puts A = 1 afterwards. The result for the particle i with the strangeness sSi and baryon number Bj reads 106
{N^st
= Zl(mi,T,V)ZB~^s;"i{^V)
(113)
In view of further applications of these results in heavy ion collisions we restrict the discussion only to non-strange systems, that is these with overall strangeness S = 0. The results of Eq. (113) should coincide with the GC value in the limit of large B and V, however, with a fixed baryon density B/V. This can be shown 106 explicitly using a Chebyshev approximation of the corresponding integrals.
542
P. Bmun-Munzinger,
K. Redlich and J. Stachel
4. The canonical statistical model and its applications The results discussed in the last section indicate that the major difference between the C and GC treatment of the conservation laws appears through a strong suppression of thermal particle phase-space in the canonical approach 48 ' 53 ' 99 ' 100 ' 105,106 ' 107 ' 117 as well as in an explicit volume dependence of particle densities. Deviations from the asymptotic GC limit are thus expected to be large for low temperature 1 and/or small volume. In a thermal fireball created in heavy ion collisions these parameters are related to the CMS collision energy and the number of participating (wounded) nucleons, respectively. It is thus clear that the canonical formulation of quantum number conservation should be of importance in low energy central and high energy peripheral heavy ion collisions as well as in hadronhadron collisions. In the following section the applications of the canonical statistical model in the above collision scenarios and for different conserved quantum numbers will be discussed. A special case is the production of hadrons containing charm quarks. This will be dealt with in Section 5.3. 4.1. Central heavy ion collisions
at SIS
energies
The number of strange particles produced in heavy ion collisions depends on the energy and centrality of the collision. In low energy A-A collisions in the SIS/GSI energy range from 1 to 2 A-GeV, the average number of strange particles produced in an event is of the order of 1 0 - 3 . Thus, following the kinetic analysis presented in Section 4.1, a statistical description would require the canonical treatment of strangeness conservation. However, the conservation of baryon number and isospin can be treated grand canonically. Consequently, one expects a different centrality dependence of strange and non-strange particle yields. Fig. (12) shows experimental data on K+ and 7r+ yields divided by the number of participants ^4part as a function of Ap&vt measured 118 in Au-Au collisions at beam kinetic energy of 1 A'GeV. The data indeed exhibit the behavior expected in the canonical statistical model: a strong increase of the K+ yield per participant and an almost constant ir+ yield per participant with centrality. In the canonical model the particle densities depend on four parameters: the chemical potentials, ^Q and fig, related with the GC description f
T h e temperature T should be low relative to the lowest particle mass that carries the conserved charge.
Particle Production
in Heavy Ion Collisions
543
i — | — i — i — i — i — i — | — i — i — i — r
Au+Au 1 AGeV
0.12 0.10
K + x2000
A
\
. i r ^
0.08
< ^
0.06 0.040.02 J
0
50
100
150
L
200
250
300
<Apart> Fig. 12. The yield of kaons and pions measured in Au-Au collisions at beam kinetic energy of 1 A GeV from Ref. (118) versus centrality given by the number of participants Apart- The line represents the A^art fit 118 to experimental d a t a .
of the electric charge and baryon number conservation, the temperature T and the volume parameter appearing through the canonical treatment of the strangeness conservation. Constraints on these variables arise from the isospin asymmetry measured by the baryon number divided by twice the charge, B/2Q. For an isospin symmetric system this ratio is simply 1, for Ni+Ni it is 1.04 while for Au+Au this ratio is 1.25. When considering particle multiplicity ratios we are thus left with three independent parameters. The volume parameter V that is responsible for the canonical suppression, the freeze-out temperature T and freeze-out baryon-chemical potential /XB of the fireball. Figs. (13 - 16) show the location and sensitivity of the freeze-out parameters for different particle ratios in the (T — ^B) plane when varying these ratios in the range obtained at SIS energy. The deuteron to proton d/p and the ir+ /p ratios provide a good determination of the range of thermal parameters. The tr+ /p curve in the (T — /XB) plane shows temperature saturation for a large \X.Q that establishes the upper limit of the freeze-
544
P. Bmun-Munzinger,
K. Redlich and J. Stachel
T[CeV]
d/p =0.26, 0.28, 0.37 O.K. 0.43.
0.0 0.6
Fig. 13. Lines of a constant 7r+/proton ratio in the T-/J,B plane obtained 4 9 in the statistical model.
J 0.7
I
L 0.9
1.0
1.1
Fig. 14. Lines of constant deutron/proton ratio in the T-\XB plane obtained 4 9 in the statistical model.
out temperature T. On the other hand, the d/p ratio fixes the range of the freeze-out value for fj,g as it shows a steep dependence on the temperature. 8 The K+/K~ ratio in Fig. (15) exhibits a similar behavior as d/p and is also independent of the volume parameter as is evident from Eq. (45) when requiring that 5 = 0. The variation of thermal parameters with the system size is shown in Fig. (16) using as an example the K+/n+ ratio. For a given system size the K+ /TT+ ratio clearly determines a lower limit of the freeze-out temperature as it saturates for a large \±B • Changing the volume parameter V = 4/3nR3 implies a substantial modification of the line in the (T — (is) plane calculated for a fixed value of the K+/TT+ yields. Thus, in the canonical model, the strange to non-strange particle ratio requires an additional consideration of the range of correlation of strange particles that is quantified by the volume parameter V. This parameter is assumed to be related to the number of nucleons participating in A-A collisions. From a detailed anals
T h e deuteron, is the composite object as it is the proton-neutron bound state. It is most likely produced by nucleon coalescence at kinetic freeze-out. Thus, one could question if deuteron yield can be used to fix chemical freeze-out parameters (see Section 3.04). At SIS energies, however, chemical and thermal freeze-out coincide and the deuteron multiplicity seems1 to follow a statistical order with the same thermal parameters as its constituents.
Particle Production
n
1
i
1
r
TfGeVJ K/tC= 12
in Heavy Ion Collisions
1
1
T[Cl
=0.003
K'/pi 20
30
40
545
50
-
-
*"~--
"v%
R=6.0f«l
""--. "••
R-
infin
R=4 01m
R=12 f n
ty
C,f0eVJ O.O1 0.6
Fig. 15. Lines of a constant K+/K~ ratio in the T-fis plane obtained 4 9 in the statistical model.
i
1
1
1
Fig. 16. The freeze-out parameters in the T-fiB plane calculated 4 9 for a fixed value of the ratio K+/n+ = 0.003 and for different correlation volumes V = ATTR3 / 3 with R = 4,6,12 fm and R = oo.
ysis of experimental data from SIS up to AGS energies it was shown that V can be identified49 as the initial overlap volume of the system created in A-A collisions. Thus, it is obtained from the atomic number of the colliding nuclei and from the impact parameter by simple geometric arguments. In heavy ion collisions at SIS energies a good description of Ni-Ni and Au-Au data was obtained 49 ' 11 when choosing V ^ VoApart/2 with Vo ~ 7fm3 i.e. of the same order as the volume of the nucleon. The comparison of the thermal model with experimental data from AGS, SPS up to RHIC energy was discussed in Section 3 and it was shown that there are common freeze-out parameters which describe simultaneously all measured particle multiplicity ratios. In order to illustrate that this is also the case in low energy heavy ion collisions we show in Fig. (17) the lines in the (T — fis) plane corresponding to different particle multiplicity ratios measured 119 in Ni-Ni collisions at 1.8 A-GeV. The experimental errors are for simplicity not shown in the figure. All lines, except the one
h
T h i s volume parameter V can be in general y/s dependent. One way to include this dependence would be to replace the spherical symmetric V by cylinder with its longitudinal size being Lorentz contracted.
546
P. Braun-Munzinger,
• 1 . T[B*V]
r
K. Redlich and J. Stachel
i •—i Ni-Ni I.g A BtV
1 iC/A^plO*
1
/ •
/'mtm/tf- 033 - • • ' ' '
s
•
~
^_____
\/
—
R-4.0T1
// /1
\
«*/(>-• 17 _
t/K~*0.1 (inclusive;
-ft '
VjGiV] 1 0.7
1 0.8
1 0.9
1 1.0
/
1.1
Fig. 17. The lines in the T-/J.B plane calculated 4 9 in the statistical model for different particle ratios obtained in central Ni-Ni collisions at 1.8 A-GeV .
'o.O
\
Au 1.0 A G»V
//
'
\d/p»0.28 tC/tT'30
0.6
:
.
// i
.
100
200
*«•"
300
400
500
Fig. 18. A comparison of the statistical model results 4 9 for the K+/Apart ratio with the data from Fig. (12). The dashed and dashed-dotted lines represent predictions of the statistical model without and with a small Apart dependence of /ig and T. For more details see text.
for rj/n0, have a common crossing point around T ~ 70 MeV and HB ~ 760 MeV. A value of R ~ 4 fm is needed to describe the measured K+ /n+ ratio with the freeze-out parameters extracted from 7r + /p and d/p ratios. This radius is compatible with that expected for a central Ni-Ni collision and was found49 to be the same in the whole energy range from 0.8 up to 1.8 A-GeV. The corresponding results for a thermal description of Au-Au collisions at two different incident kinetic energies 1.0 and 1.5 A-GeV can be found in Ref. (49). As for Ni-Ni data, the particle ratios, n+ /p, K+/n+, TT+/-K~ , + K /K~ and d/p, with exception of r)/n°, could be described with the same value of the freeze-out parameters. The temperature T ~ 53 MeV and HB ~ 822 MeV were found in Au-Au collisions at 1.0 A-GeV.' Thus, the freeze-out temperature is obviously decreasing whereas the baryon chemical potential is an increasing function of the collision energy. A small variation
'With these responds to and leading to the total
thermal parameters the total density of particles at chemical freezeout corTIB ~ 0.05/fm 3 . Due to rather large value of the baryochemical potential contribution of baryons to total particle density this value also corresponds baryon density of the system.
Particle Production
in Heavy Ion Collisions
547
of freeze-out parameters with A in central A-A collisions for the same collisions energy was also extracted from the data 49 . The observed scaling of the volume parameter that determines the canonical suppression of the strange particle phase—space with the number of participants was also found to be valid in Au-Au collisions. Here, in the most central collisions a radius of ~ 6.2 fm is required to reproduce the measured K+ /Apart and K~ /Apart yields. The larger radius value obtained for Au compared to Ni data is compatible with the larger size of Au and corresponds to Apart ~ 330 in the most central Au-Au collisions. The importance of strangeness suppression due to canonical treatment of the conservation laws is particularly transparent in the comparison of the thermal model with the Au-Au data at 1 A-GeV . Using the grand canonical formulation of the strangeness conservation one would get a value of K+ /-K+ ~ 0.04 that overestimates the data by more than an order of magnitude. This shows that the thermal particle phase-space at SIS energies is far from the grand canonical limit and that the exact and local treatment of strangeness conservation is of crucial importance. The multiplicity of K+ per participant (K+ /Apart) was indicated in Fig. (12) to increase strongly with centrality while the corresponding ratio for pions 7r+ /Apart, is constant with Apart119. Consequently, the pion yield is proportional to the number of participants while the multiplicity of K+ scales with Apart as K+ ~ A.^art with a ~ 1.8. The canonical treatment of the strangeness conservation predicts the yield of strange particles to increase quadratically with the number of participants, see e.g. Eq. (28). In Fig. (18) the experimental data from Fig. (12) are compared with the thermal model. The parameters were chosen to reproduce the n/p and d/p ratios. The dashed-line in Fig. (18) describes the results of a thermal model under the assumption that both T and \XB are independent of Apart. One sees that already under this approximation the agreement of the model and the experimental data is very satisfactory as the model describes the magnitude and the centrality dependence of these data. Some differences between the model and the data seen in Fig. (18) can be accounted for when including a small variation of the freeze-out parameters with Apart. A smooth and almost linear increase of the temperature with centrality by a few MeV and a corresponding decrease of fis from very peripheral to central collisions is sufficient to get a very good description of the data. In Fig. (18) the dashed-dotted line shows thermal model results that include
548
P. Braun-Munzinger,
K. Redlich and J. Stachel
the variation of thermal parameters with centrality. Such a small dependence of the freeze-out temperature on impact parameter comes at first glance as a surprise since the experimental result on the apparent inverse slope parameter Tapp of particle yields as a function of pt in Au-Au collisions shows a strong dependence on A por -t 124 ' 123 - This difference, however, can be accounted for by including the concept of centrality dependent transverse collective flow of the collision fireball. A detailed analysis of particle spectra in Au-Au collisions at 1 A-GeV has shown that keeping the chemical freeze-out temperature at 53 MeV and including collective transverse flow reproduces the transverse momentum distributions of pions, protons and kaons as well as their centrality dependence 49 . This result indicates that chemical and thermal freeze-out coincide at SIS energy. The canonical thermal model provides a consistent description of the experimental data in the GSI/SIS energy range. The abundances of K+,K~, p, d, TT+ and ir~ hadrons (with the notable exception of 77 and possibly <j>) •> seem to come from a common hot source and with well defined temperature, T « 50,54,70 MeV, and baryon chemical potential us » 825,805,750 MeV for central Ni-Ni collisions at 0.8, 1.0, 1.8 A-GeV and correspondingly T « 53 MeV and /JLB ~ 822 MeV for central Au-Au collisions at 1.0 A-GeV. These temperatures are lower than the ones observed in the particle spectra but here a common explanation is possible in terms of hydrodynamic flow. The flow differentiates between particles of different mass since they acquire the same boost in velocity but very different boosts in momentum. The common freeze-out condition for almost all particle species is very strong evidence for chemical equilibrium in low energy heavy ion collisions. A satisfactory agreement of the model and the data could be only obtained when including a canonical description of the strangeness conservation. The relevant parameter that quantifies the canonical suppression of the particle phase-space was found to be the initial volume of the collision fireball that scales with the number of projectile participants. This volume parameter describes the range of strange particle correlations and is smaller than the
J
The observed deviations of the model from the measured yield of r] mesons require 1 2 2 further studies. It is conceivable that the canonical model described in Section 4 does not account correctly for hidden strange particle production. The recent result of
Particle Production
in Heavy Ion Collisions
549
radius of the fireball at chemical freeze-out. In central Au-Au collisions at 1 A-GeV the correlation radius was found to be 6.2 fm roughly corresponding to the size of Au whereas the radius required to reproduce measured particle yields is almost two times larger 49 . The appearance of two different space-like scales in the canonical description of particle production can be possibly understood from the kinetics of strangeness production. Introducing the locality of strangeness conservation in the kinetic equation (15) implies that the volume parameter in the loss terms can be different (smaller) than in the gain terms. Consequently, strange particle yields depend on two volume parameters: (i) the volume of the fireball which is also an overall normalization factor that determines the total strange and nonstrange particle yields originating from the collisions fireball and (ii) the volume that parameterizes the space-like correlations of strange particles that is required to satisfy exactly the strangeness conservation. The second parameter is also related with the initial number of nucleons participating in the collision. The apparent chemical equilibration of particle yields measured at SIS energies and the kinetic theory developed in Section 4 has recently inspired a more complete dynamical study of the problem in terms of microscopic transport models 1 2 5 - 1 2 8 . Recently the relativistic transport model was applied 129 ' 130 to describe the chemical equilibration of kaon and antikaon at energies that are below the N-N threshold. The results of these microscopic studies indicate that K+ can possibly appear k in chemical equilibrium during the lifetime of the collision fireball. The K~, on the other hand, approaches chemical equilibrium even at a earlier times However, it may eventually fall out of equilibrium at a later time due to the large annihilation cross sections in nuclear matter 130 . Thus, the results of transport models do not exclude chemical equilibration in low energy heavy ion collisions. The level of equilibration in these models is strongly related to the magnitude of production and absorption cross section of kaons inside the nuclear medium 131 . Although significant progress has been made in the theoretical description and understanding of in-medium kaon cross sections, the results are still far from complete. In particular, recently it was suggested that the coupling of kaons with the p-wave £(1385) res-
k However, this equilibration can be obtained if the K+ mass is substantially changed in the hot and dense medium.
550
P. Brawn-Munzinger,
K. Redlich and J. Stachel
onance can substantially increase the K production cross section that could influence132 the approach of kaons towards an equilibrium. 4.2.
Particle
production
in high energy p-p
collisions
The success of the statistical approach in the description of particle production in heavy ion collisions discussed in Sections 3 and 4 prompts the question if a statistical order of the secondaries is also observed in elementary collisions such as high energy p-p interactions, to which we restrict our discussion. The results of Section 4 make it clear that one should in this case apply a model that accounts for the canonical conservation of the quantum numbers. This is particularly the case for strangeness since even in very high energy p-p collisions the number of produced strange particles per event is of the order of unity. Thus, large, event by event, strange particle multiplicity fluctuations prevent the applicability of the GC approach. Whether or not the GC treatment of the baryon number and electric charge is adequate, would require a detailed study of the relative error between the canonical and the grand canonical results. The application of the statistical model to the elementary hadronhadron reactions was first proposed by Rolf Hagedorn 105 in order to describe the exponential shape of the m t -spectra of produced particles in p-p collisions. Hagedorn also pointed out phenomenologically the importance of the canonical treatment of the conservation laws for rarely produced particles. The first application of the canonical model to strangeness production in p-p collisions was done by Edward Shuryak 105 in the context of ISR data. Recently a complete analysis of hadron yields in p-p as well as in p-p, e+e~, 7r-p and in K-p collisions at several center-of-mass energies has been done in Refs. (45, 56, 133). This detailed analysis has shown that particle abundances in elementary collisions can be also described by a statistical ensemble with maximized entropy. In fact, measured yields are consistent with the model assuming the existence of equilibrated fireballs at a temperature T «160-180 MeV. The most general partition function ZB,Q,S(V, T) that is applied to test the chemical composition of the secondaries in elementary collisions should account for the canonical conservation of baryon number B, strangeness S, and electric charge Q. It can be constructed applying the projection method (see Section 4.2) for the UB(1)XUS(1)XUQ(1) symmetry. Following Eq. (79) one has
Particle Production in Heavy Ion Collisions 551
3 -3
! . 1
%
pp V s = 27.4 G e V
r
P* K* ">.••+
. K*oK lO
rr :
lO
-o
A*
s*
A+ *'>•"'
mberof SLDev. ^
*»#
?-* A
P
>
A°
• •
lO"1
r
/ ** + r
i
i
i
+
** i
i
P
P
1 Multiplicity (therm, model)
5 -5
.. '
,3^1* lO"2
O
. •
^»
•• • • • '
• • • • * •
• +
+
» ' •
*+ ' i
-J
i
• I*I
• i
•
•• i
+
i
i
-
7t° 71* n K"K*K°tl p ° p " p* to K"~K K**£° p p <j) A A £ £~ A A* A° A*% S "2:*+A(1520)
Fig. 19. A comparison133 of the p-p multiplicity data at y/s = 27A with the statistical model that accounts for the exact conservation of baryon number, electric charge, strangeness and includes the strangeness undersaturation factor 7$ ~ 0.51.
ZB,Q,S{V,T)=
r*pLe-*B+* Jo
27T
[* J0
d(t>Q 2TT
'
1-2-K
^ 5
Jo
-iS
2?r
(114)
where B, Q, and S are the initial values of the quantum numbers in hadronhadron collisions which are in p-p scattering B=Q=2 and S=0. The generating function Z in Eq. (114) can be obtained from the grand canonical partition function of a hadron resonance gas (3) by a Wick rotation of all appropriate chemical potentials Z(T,V,<j>B,
-i^Q^s-^-i^s).
(115)
Different particle multiplicities and their ratios are obtained from the partition function (114) following the procedure that was described in Section 4.2.3.
552
P. Braun-Munzinger,
K. Redlich and J. Stachel
Fig. (19) shows an example of the comparison of the canonical model prediction (114) with the experimental data for different particle yields obtained 133 in p-p collisions at yfs = 27.4-27.6 GeV. The agreement of the model with most of the experimental data is reasonable. However, the yields
• O
•
B.S.Q canonical V-77 fm' T*169±5..ram0.51, only S-canonicml, 0 S -19S MmV T*164±2.,r0mO.53, V'19 tm*y
10"
10"'
to" 8
y
jft •
-y
o Cmnenietl modml with 7,"0.S m Canonic* J modml with V9'Vpr^tmn ^
V"'
"
p
"
-<
r
•
- S *
s %%*
•
•p a t « " • •
27
*
OmV
10 - ' 10' Multiplicity-model
Fig. 20. A comparison of the p - p multiplicity obtained at v ^ = 27.4 with the canonical models. The open symbols represent the result of the model that accounts for 5, B and Q being exactly conserved 1 3 3 . T h e filled symbols are the statistical model results with only 5 being exact whereas B and Q are treated grand canonically.
10"' 10" Multiplicity-model
Fig. 21. As in Figure (20) but the filled circle are obtained in the canonical model that accounts for the strangeness undersaturation to be controlled by the correlation volume, Vp = 47rfl 3 /3 with Rp ~ 1.1 fm instead of -ys.
of the resonances like A 0 , p° and
Particle Production
in Heavy Ion Collisions
553
the comparison of p-p data with the canonical model that accounts for only strangeness being conserved exactly. The baryon number and electric charge conservation are treated in the GC ensemble, thus are controlled by chemical potentials. A satisfactory description of experimental data with He ~ 195 MeV, fiQ ~ 30 MeV, T ~ 165 MeV and -ys ~ 0.53 seen in Fig. (20) shows that canonical effects related with charge and baryon number conservation are indeed small 134 . This is, however, not the case in p-p and e + e~ collisions since the initial values of B = Q = 0 are there obviously too small to use a GC approximation. The experimental data shown in Fig. (19) can be also described in terms of the canonical model that was successfully applied in low energy heavy ion collisions (see Section 5.1). There, instead of the strangeness undersaturation factor 7,,, space-like correlations of strange particles were introduced. Consequently, there were two volume parameters that determined particle yields: (i) the volume of the fireball and, (ii) the correlation volume. Fig. (21) shows that choosing the correlation volume Vp = 47ri?3/3 and Rp ~ 1.1 — 1.2 fm the p-p data are well reproduced with the exception of <j>. The (/>-meson is not canonically suppressed because it only carries zero strangeness. Taking the correlation volume instead of 7S gives a deviation of the measured
We have to point out, however, that midrapidity d a t a in A-A collisions at SPS are consistent 3 9 with the value of fs = 1- See the more detailed discussion in Section 2.2. "Looked at from a different angle we conclude that in central nucleus-nucleus collisions at AGS energy and higher strangeness production reaches values consistent with complete chemical equilibrium. At lower energies, and in particular in elementary particle collisions, strangeness is strongly undersaturated. This will be discussed further in Section 6.
554
P. Braun-Munzinger,
K. Redlich and J. Stachel
the chemical freeze-out temperature extracted from heavy ion data 4 0 , 1 3 3 . At SPS and RHIC energies, T can be considered as a measure of thermal excitations of a non-perturbative QCD vacuum due to the particle scattering. Thus, T should be mostly correlated with the collision energy and not with the system size. The charge chemical potential, however, due to the isospin asymmetry, differers substantially in p-p and A-A collisions. In Pb-Pb collisions at the SPS the (XQ ~ —7 MeV 38 whereas in p-p at y/s ~ 27 the LIQ ~ 35 MeV is required. 4.2.1. Statistical hadronization and string dynamics in p-p collisions The apparent agreement of the canonical statistical model with experimental data on particle production in elementary collisions leads to the interpretation that hadronization in particle collisions is a statistical process. This result is difficult to reconcile with the popular picture that hadron production in hadron-hadron collisions is due to the decay of color flux tubes 135 , a model that has explained many dynamical features of these collisions. In the following we address the question, how one can possibly distinguish the string hadronization via the break up of a color flux tube from the statistical hadronization. We argue following Ref. (136) that the fl/Q, — Q+ /Q~ ratio in elementary proton-proton collisions is a sensitive probe to differentiate possibly these two scenarios. Color flux tubes, called strings, connect two SU(3) color charges [ 3 ] and [ 3 ] with a linear confining potential. If the excitation energy of the string is high enough it is allowed to decay via the Schwinger mechanism 137 , i.e. the rate of newly produced quarks is given by: -—- ~ exp [—7rmj_/«;] ,
(116)
where K is the string tension and m± = \/p\ + m2 is the transverse mass of the produced quark with mass m. However, specific string models may differ in their philosophy and the types of strings that are created: • In UrQMD 125 the projectile and target protons become excited objects due to the momentum transfer in the interaction. The resulting strings, with at most two strings being formed, are of the diquark-quark type.
Particle Production
in Heavy Ion Collisions
555
• In NeXuS 126 , the p-p interaction is described in terms of pomeron exchanges or ladder diagrams. Both hard and soft interactions take place in parallel. Energy is equally shared between all cut pomerons and the remnants. The endpoints of the cut pomerons (i.e. the endpoints of the strings) may be valence quarks, sea quarks or antiquarks. • In PYTHIA 127 , a scheme similar to that in UrQMD is employed. However, hard interactions may create additional strings from scattered gluons and sea quarks. Most strings are also of a diquarkquark form. Fig. (22 -left) depicts the antibaryon to baryon ratio at midrapidity in proton-proton interactions at 160 GeV. The results of the calculations by NeXuS, UrQMD and PYTHIA are included in this figure 136 . In all these models, the B/B ratio increases strongly with the strangeness content of the baryon. For strangeness |s| = 3 the ratio significantly exceeds the unity. In UrQMD and PYTHIA the hadronization of the diquark-quark strings leads directly to the overpopulation of Q. In NeXuS, however, the imbalance of quarks and anti quarks in the initial state leads to the formation of qvai—sSea strings (the sva\ — qsea string is not possible). These strings result then in the overpopulation of f2's. In order to understand the large fi/J7 values predicted by string models we include in Fig. (23), the color flux tube break-up mechanism. Fig. (23) shows the fragmentation of the color field into quark-antiquark pairs, which then coalesce into hadrons. While in large strings O's and fl's are produced in equal abundance (a), low-mass strings in UrQMD suppress fi production at the string ends (b), while in NeXuS Q's are enhanced (c). Thus, the microscopic method of hadronization leads to a strong imbalance in the Q,/Cl ratio in low-mass strings. In Fig. (22) the string model results are compared with the predictions of two statistical models (SM) 135 and the preliminary experimental data obtained 138 by the NA49 Collaboration. The main difference between these models is the implementation of baryon number and electric charge conservation and the way an additional strangeness suppression is introduced. In model (i) the calculation 56 is a full canonical one with fixed baryon number, strangeness and electric charge identical to those of the initial state. An extra strangeness suppression is introduced to reproduce the experimental
556
P. Braun-Munzinger,
m • A •
K. Redlich and J. Stachel
UrQMD NeXuS PYTHIA NA49
fi
.
• • *
Modell I Modell II NA49
. * g strangeness s
strangeness s
Fig. 22. The left hand figure: the anti baryon to baryon ratio at \y — yCm\ < 1 in p - p interactions at 160 GeV as given by PYTHIA, NeXuS and UrQMD. The right hand figure: the anti baryon to baryon ratio for the same reaction as given by the statistical models. Stars depict preliminary NA49 data for the B/B ratio at midrapidity.
multiplicities. This is done by considering the number of newly produced (ss) pairs as an additional charge to be found in the final hadrons. The ss pairs fluctuate according to a Poisson distribution and their mean number is considered as a free parameter to be fitted 56 . The parameters used for the prediction of the f2 + /Q~ ratio (T, the global volume V sum of single cluster volumes and (si)) have been obtained by a fit to preliminary NA49 p-p data 1 3 8 yielding T = 183.7 ± 6.7 MeV, VT3 = 6.49 ± 1.33 and (ss) = 0.405 ± 0.026 with a x 2 /dof = 11.7/9. It must be pointed out that the fi+/fi~ ratio is actually independent of the (ss) parameter and only depends on T and V. In model (ii) the conservation of the baryon number and electric charge is approximated by using the GC ensemble. The strangeness conservation is, however, implemented on the canonical level following the procedure that accounts for strong correlations of produced strange particles. In p-p collisions the strangeness is assumed not to be distributed in the whole volume of the fireball but to be locally strongly correlated. A correlation
Particle Production
_ 1=11
_ lf=^-i\
in Heavy Ion Collisions
557
1 __
Fig. 23. Fragmentation of a color field into quarks and hadrons. While in large strings Q's and fi's are produced in an equal abundance (a), small di'quark strings suppress H's at the string ends (b), sea-3 quarks enhance Q's (c).
volume parameter Vo = 4TTRQ/3 is introduced, where RQ ~ 1 fm is a typical scale of QCD interactions. The temperature T ~ 158 MeV and /is ~ 238 MeV were taken as obtained 40 from the SM analysis of a full phase-space Pb-Pb data of NA49 Collaboration. The volume of the fireball V ~ 17 fm3 and the charge chemical potential in p-p was then found to reproduce the average charge and baryon number in the initial state. The predictions of the statistical models are shown in Fig. (22 -right). In these approaches the B/B ratio is seen to exhibit a significantly weaker increase with the strangeness content of the baryon than that expected in the string fragmentation models. For comparison, both figures include preliminary data on the B/B ratios obtained 138 at midrapidity by the NA49 Collaboration. Note that the predictions of the statistical models in Fig. (22) refer to full phase-space particle yields whilst measurements of B/B ratios in p-p collisions have been performed at midrapidity, where they are expected to be the largest. Therefore, sizeable deviations of the model results from the data seen in Fig. (22) are to be expected. However, admitting the applicability of SM for midrapidity one reproduces 141 the experimental data quite well.
558
P. Braun-Munzinger,
K. Redlich and J. Stachel
In macroscopic string models the fi/f2 ratio depends in a strongly nonlinear fashion on the mass (energy) of the fragmenting string. All these models predict a strong enhancement of Q production at low energies, while for large string masses the ratio approaches the value of Q/fl = 1 (which should be reached in the limit of an infinitely long color flux tube). Statistical models, on the other hand, are not able to yield a ratio of Q/fi > 1. This can be easily understood in the GC formalism, where the B/B ratio is very sensitive to the baryon chemical potential HB- For finite baryon densities and including 100% feeding from resonances, the B/B ratio will always be < 1 and only in the limit of /is = 0 may fl/Q = 1 be approached. These features survive in the canonical framework, where the GC fugacities are replaced by the ratios of partition functions 105 ' 106 ' 107 . From the above discussion and from Fig. (22) it is thus clear that within the fragmenting color flux tube models the fi/fi ratio is significantly above the unity. This is in strong contrast to statistical model results, that always imply that B/B ratios are below or equal to a unity in proton-proton reactions. Since this observable is accessible to NA49 measurements at the SPS it can provide an excellent test to distinguish the statistical model hadronization scenario from that of a microscopic color-flux tube dynamics. We have to point out, however, that the classical string models considered above do not account for the so called string junction mechanism that allows for diffusion of baryon number towards midrapidity. This mechanism, recently included 139 in Dual Parton Model, was shown to be very important for (multi)strange baryon and antibaryon production. Thus, it would be of importance to study the energy dependence of B/B ratio in p-p collisions in terms of the model that includes this baryon number transport. 4.3. Heavy quark
production
Charm quarks are heavy (m c ~3> Tc) thus, thermal production of charm quarks and charmed hadrons is strongly suppressed in ultra-relativistic heavy ion collisions142. The situation has been recently discussed 75,76 with the conclusion that, compared to direct hard production, thermal production of charm quarks can be neglected at SPS energies and is small even at LHC energy. However, these investigations led to a new scenario for the production of hadrons which contain charm quarks in which production of heavy quarks through hard collisions is combined with a statistical procedure to produce open and hidden charm hadrons at hadronization. This idea
Particle Production
in Heavy Ion Collisions
559
of statistical hadronization of charm quarks ' 76 has sparked an intense activity in this field 143 ' 144 . Initial interest focussed on the available SPS data on J/ip production and their interpretation in terms of a conventional statistical model 145 . As we show below, these data can be well described, but only assuming a charm cross section which is enhanced compared to predictions within the framework of perturbative QCD. However, the largest differences between results from the statistical coalescence scenario (or a similar 146 model) and more conventional models are expected at collider energies. For example, in the Satz-Matsui approach 28 , one would expect very strong suppression compared to direct production of J/ip mesons (up to a factor 147 of 20) for central Au-Au collision at RHIC energy. In the present approach this suppression is overcome by statistical recombination of J/ip mesons from the same or different cc pairs, so that much larger yields are expected. We therefore focus in this section on predictions" for open and hidden charm mesons at RHIC and LHC energy, with emphasis on the centrality dependence of rapidity densities. 4.3.1. Statistical Recombination Model In this model it is assumed that all charm quarks are produced in primary hard collisions and (thermally) equilibrate in the quark-gluon plasma (QGP); in particular, no J/ip is preformed in QGP (complete screening) and there is no thermal production of charm quarks. For a description of the hadronization of the c and c quarks, i.e. for the determination of the relative yields of charmonia, and charmed mesons and baryons, we employ the statistical model, with parameters as determined by the analysis of all other hadron yields 175 . The picture we have in mind is that all hadrons form within a narrow time range at or close to the phase boundary. All charmed hadrons (open and hidden) are formed at freeze-out (at SPS and beyond, freeze-out is at the phase boundary 175 ) according to the statistical laws. Another interesting point concerns the ip /(J/ip) ratio. As is well known, this ratio is, in hadron-proton and p-nucleus collisions, close to 12 %, independent 152 of collision system, energy, transverse momentum etc.. In the thermal model, the ratio is 3.7 %, including feeding of the J/ip from "This section is based on work by the authors and A. Andronic and reported in Ref. (148, 149).
560
P. Braun-Munzinger,
K. Redlich and J. Stachel
heavier chaxmonium states. A temperature of about 280 MeV would be necessary to explain the ratio found in pp and p-nucleus collisions in a thermal approach. Clearly, J/ip and ip production in pp and p-nucleus collisions are manifestly non-thermal. This was previously realized in Refs. (154, 151). Similar considerations apply for the \ states. In fact, feeding from xi to J/ip is less than 3 % if the production ratios are thermal. The experimental situation concerning the evolution with participant number of the ip /(J/V0 ratio in nucleus-nucleus collisions, multiplied with the respective branching into muon pairs, is presented in Fig. (24). The data are from the NA38/50 collaboration 155 ' 156 ' 157 - 158 . With increasing N p a r t the ip /(3/tp) ratio drops first rapidly (away from the value in pp collisions) but seems to saturate for high N p a r ( values at a level very close to the thermal model prediction, both for S+U and Pb-Pb collisions • SUdalaNASO • PbPb95 D PbPb96
t? 0.02 i , pW 4 pd "pp"""
CD
average pp and pA
.2.0.015 m
u TTjIi
0.01
M
0.005
i i . . . .
0
50
i . , , .
i . . . .
i i
100 150 200 250 300 350 400 450 Npar,
Fig. 24. Comparison of the dependence of the measured ip /(J/i/') ratio on the number of participating nucleons with the prediction of the thermal model. T h e data of NA38 and NA50 Collaborations are from Refs. (155, 156, 157, 158). See text and Refs. (75, 76) for more details.
Taking this into account we note that predictions of the model should only be trusted from about N p a r t > 150 on, where also the rp'/(J/tp) ratio is close to the thermal value for Pb-Pb data. In our approach, ratios for all higher charmonia states including the \ c should approach the thermal value from N p a r t > 150 on, implying that for those N p a r t values feeding
Particle Production
in Heavy Ion Collisions
561
to J/ip should be small. In this picture, there should thus not be different "thresholds" for the disappearance of different charmonia. The total number of open charm hadrons expected in a purely thermal approach , N%£, is then readjusted to the number of directly produced cc pairs, N$r as (neglecting charmonia): N*r = %gcN£h(gcN£)/I0(gcN£), from which the charm enhancement factor gc is extracted. Here, /„ are modified Bessel functions. Note that we use here the canonical approach as, depending on beam energy, the number of charm quark pairs maybe smaller than 1. The grand-canonical limit will likely only be reached at LHC energy. For a detailed study of the transition from the canonical to the grand-canonical regime see Section 3.2 and Ref. (76). The yield of a given species X is then determined by Nx = gcN$h(geN%)/I0(gcN%) for open charm mesons and hyperons and Nx = g^N^ for charmonia (see Refs. (75, 76) for more details). The inputs for the above procedure are: i) the total charged particles yields (or rapidity densities), which are taken from experiments at P s p s i85,i59 a n d RHIC184 a n d extrapolated for LHC; and ii) JV* , which is taken from next-to-leading order (NLO) perturbative QCD (pQCD) calculations for pp 1 6 0 (the yield from MRST HO parton distributions was used here) and scaled to AA via the nuclear overlap function. A constant temperature of 170 MeV and the baryonic chemical potential //& according to the parameterization /i;,(MeV)=1270/(l + ^SATJV/4.3 have been used for the calculations 53 .
4.3.2. Results We first compare predictions 148 ' 149 of the model to 47r-integrated J/ip data 16 at the SPS from NA50 Collaboration replotted as outlined in Ref. (76). In Fig. (25) we present the model results for two values of N^r: from NLO calculations 160 (dashed-line) and scaled up by a factor of 2.8 (continuous line). The dashed-dotted line in Fig. (25) is obtained with the NLO cross section for charm production scaled-up by a factor 1.6, which is the ratio of the open charm cross section estimated 153 by NA50 for p-p collisions at 450 GeV/c and the NLO values from Ref. (149). For this case the Npart scaling is not the overlap function, but is taken according to measured 153 dimuon enhancement as a function of Npart. The results of Fig. (25) indicate that the observed centrality depen-
562
P. Braun-Munzinger,
K. Redlich and J. Stachel
dence of J/tp for 100 < Npart <350 is well reproduced by the statistical model using the NLO cross sections for charm production scaled by the nuclear overlap function. However, to explain the overall magnitude of the data a N^r increase by a factor of 2.8 compared to NLO calculations is needed. The drop of the J/tp yield per participant observed in the data for Npart >350 ( see Fig. (25)) is currently understood 150 in terms of energy density fluctuations for a given overlap geometry. We mention in this context that the observed enhancement of the dimuon yield at intermediate masses has been interpreted 163 by NA50 as possible indication for an anomalous increase in the charm cross section. We note, however, that other plausible explanations 13 ' 16 exist of the observed enhancement in terms of thermal radiation. *
0.3
"i
' i ' i—' i '
r
• NA50data
£0.25
2: "t>
0.2
0.15
0.05
50
J i I i L 100 150 200
250
300
350
400
N.part
Fig. 25. The centrality dependence of J/tp production at SPS. Model predictions are compared to 47r-integrated NA50 d a t a 1 6 1 ' 1 6 2 . Two curves for the model correspond to values of N*iT ~ 0.137 from NLO calculations 1 6 0 (dashed line) and scaled up by a factor of 2.8 (continuous line). The dashed-dotted curve is obtained when considering the possible NA50 Wpart~dependent charm enhancement 1 6 3 over their extracted 1 5 3 p - p cross section (see text).
We turn now to discuss model predictions for collider energies. For comparison we include in this study also results at SPS energy. The input parameters for these calculations for central collisions {Npart=350) are presented in Table 2. Notice that from now on we focus on rapidity densities, which are the relevant observables at the colliders. The results are compiled in Table 3 for a selection of hadrons with open and hidden charm. All predicted yields increase strongly with energy, reflecting the increasing charm cross section and the concomitant importance of statistical recombination.
Particle Production
in Heavy Ion Collisions
563
Also ratios of open charm hadrons evolve with increasing energy, reflecting the corresponding decrease in charm chemical potential. Table 2.
Input parameters for model calculations at SPS, RHIC and LHC. 17.3 170 253 430 861 0.064 1.86
v/sjviv (GeV) T (MeV) fib (MeV) dNch/dy dN^/dy 9c
200 170 27 730 1663 1.92 8.33
5500 170 1 2000 4564 16.8 23.2
Source: A. Andronic et al. from Ref. (148).
Table 3.
Results of model calculations at SPS, RHIC and LHC for iV p a r t=350.
v/Siviv ( G e V ) dND+/dy dND- /dy diV D o/dy dNAJdy ANj/^/dy dN^/dy
17.3 0.010 0.016 0.022 0.014 2.55-10" 4 0.95-10- 5
200 0.404 0.420 0.89 0.153 0.011 3.97-10- 4
5500 3.56 3.53 7.8 1.16 0.226 8.46-10-
Source: A. Andronic et al. from Ref. (148).
Predictions for the centrality dependence of J/tp production are presented in Fig. (26). In addition to the dramatic change in magnitude (note the scale-up by factors of 10 and 100 for RHIC and SPS energy, respectively) the results exhibit a striking change in centrality dependence, reflecting the transition from a canonical to a grand-canonical regime (see Ref. (76) for more details). The preliminary PHENIX results on J/ip production at RHIC 165 agree, within the still large error bars, with our predictions. A stringent test of the present model can only be made when high statistics J/tp data are available. Another important issue in this respect is the accuracy of the charm cross section, for which so far only indirect measurements are available 166 . In any case, very large suppression factors as predicted, e.g., by Ref. (147) seem not supported by the data. In Fig. (26) we present the predicted centrality dependence of charged D + -meson production for the three energies. The expected approximate scaling of the ratio D+/Npart like Np'art (dashed lines in lower Fig. (26)) is only roughly
564
P. Braun-Munzinger,
A IS •
LHC RHICxIO SPSxIOO
x °b *-
K. Redlich and J. Stachel
-m—•—•--••
_ » — • » — • " "
0.2
-+A • •
-+-
LHC RHICx5 SPSxIOO
' I
Fig. 26. Centrality dependence of J/tj) (upper figure) and of D+ (lower figure) rapidity density at SPS, RHIC and LHC energies.
' 1 ' I ' I
t x
N„.
fulfilled due to departures of the nuclear overlap function from the simple N pirt dependence. 4.3.3. Charmonium Production from Secondary Collisions at LHC Energy Another possibility to produce charmonium states is due to reactions among D mesons in the hadronic and mixed phase of the collision. This has been investigated in Refs. (142, 167). As demonstrated there, this mechanism does not lead to appreciable charmonium production at SPS and RHIC energies. However, the large number of cc pairs and consequently D,D mesons produced in Pb-Pb collisions at LHC energy can lead to an additional production of charmonium bound states due to reactions such as: DD* + D*D + D*D* -> J/ip + n and D*D* +£>£»-> J/ip + p. These processes were studied within a kinetic model taking into account the spacetime evolution of a longitudinally and transversely expanding medium. The results 142 demonstrate that secondary charmonium production appears almost entirely during the mixed phase and is very sensitive to the charmonium dissociation cross section with co-moving hadrons. Within the most
Particle Production, in Heavy Ion Collisions
565
likely scenario for the dissociation cross section of the J/t/j mesons their regeneration in the hadronic medium will be negligible, even at LHC energy. Secondary production of tp' mesons however, due to their large cross section above the threshold, can substantially exceed the primary yield. 4.3.4. Conclusions on Heavy Quark Production We have demonstrated that the statistical coalescence approach yields a good description of the measured centrality dependence of J/ip production at SPS energy, albeit with a charm cross section increased by a factor of 2.8 compared to current NLO calculation. Rapidity densities for open and hidden charm mesons are predicted to increase strongly with energy, with striking changes in centrality dependence. First RHIC data on J/ip production support the current predictions, although the errors are too large to make firm conclusions. For LHC energies we predict cross sections for charm production in central Pb-Pb collisions significantly exceeding the values predicted by scaling results for N-N collisions with the nuclear thickness function. The statistical coalescence implies travel of charm quarks over significant distances in QGP. If the model predictions will describe consistently precision data then 3/ip enhancement (rather than suppression) would be a clear signal for the presence of a deconfined phase. Regeneration of charmonia in the mixed and hadronic phase has also been studied. For J/ip mesons this will likely only be a small effect, even at LHC energy. However, secondary production of ip' mesons may be significant at LHC energy. 5. Unified conditions of particle freeze—out in heavy ion collisions A detailed analysis of experimental data in heavy ion collisions from SIS through AGS, SPS up to RHIC energy discussed in Section 3 and 5 makes it clear that the canonical or grand canonical statistical model reproduces most of the measured hadron yields. Figure (27) is a compilation of chemical freeze-out parameters that are required to reproduce the measured particle yields in central A-A collisions at SIS, AGS, SPS and RHIC energy. The GSI/SIS results have the lowest freeze-out temperature and the highest baryon chemical potential. As the beam energy increases a clear shift towards higher T and lower /xs occurs.
566
P. BrauTi-Munzinger,
K. Redlich and J. Stachel
200
O .4
O. 2
//'
O .6
[GeV]
Fig. 27. A compilation of chemical freeze-out parameters appropriate for A-A collisions at different energies: SIS results are from Ref. (49), AGS from Ref. (40), SPS at 40 A-GeV from Refs. (44, 168, 45, 175), SPS at 160 A-GeV from Refs. (38, 39, 40), and RHIC from Refs. (35, 54, 55). The full line represents the phenomenological condition of a chemical freeze-out at the fixed energy/particle ~ 1.0 GeV. 3 4
There is a common feature to all these points, namely that the average energy {E) per average number of hadrons (N) is approximately 1 GeV. A chemical freeze-out in A-A collisions is thus reached 34 when the energy per particle (E)/(N) drops below 1 GeV at all collision energies. In cold nuclear matter the (E)/(N) is approximately determined by the nucleon mass. For thermally excited nuclear matter, in the non-relativistic approximation
(El (N)
(m) + -T.
(117)
with (m) being the thermal average mass in the collisions fireball. This result makes it clear why at SIS the energy/particle at chemical freeze-out is of the order of 1 GeV since T ~ 53 MeV. At SPS and RHIC energy the leading particles in the final state (at thermal freeze-out) are pions.
Particle Production
in Heavy Ion Collisions
567
However, at chemical freeze-out most of the pions are still hidden in the mesonic and baryonic resonances. Thus, here the average thermal mass is larger than the pion mass and corresponds approximately to the p-meson mass. Consequently, since (m) > > T, Eq. (117) can be still used to justify approximately that (E)/(N) ~ 1 GeV at the SPS. Actually, this argument holds, to a large extent, in the whole energy range from SIS up to RHIC energy. The physical origin of the phenomenological freeze-out condition of fixed energy/particle would require further dynamical justification and interpretation. Recently, this question has been investigated in central Pb-Pb collisions at the SPS in terms of the Ultra-relativistic Quantum Molecular Dynamics model (UrQMD) 169 ' 170 . A detailed study has shown that there is a clear correlation between the chemical break-up in terms of inelastic scattering rates and the rapid decrease in energy per particle. If (E)/(N) approaches the value of 1 GeV the inelastic scattering rates drop substantially and further evolution is due to elastic and pseudo-elastic collisions that preserved the chemical composition of the collision fireball. Following the above UrQMD results and the previous suggestions 171 one could consider the phenomenological chemical freeze-out of (E)/(N) ~ 1 GeV as the condition of inelasticity in heavy ion collisions. Unified freeze-out conditions were also considered 172 in the context of hydrodynamical models for particle production and evolution in heavy ion collisions. There, it was suggested, that the condition for chemical freezeout, (E)/(N) ~ 1 GeV, selects the softest point of the equation of state, namely the point where the ratio of the thermodynamical pressure P to the energy density e has a minimum. The considerations were essentially based on the proposed 173 mixed phase model that seems to be consistent with the available QCD lattice data. The quantity P/e is closely related to the square of the velocity of sound and characterizes the expansion speed m of the reaction zone. Thus, the system lives for the longest time around the softest point that allows to reach the chemical equilibrium of its constituents. The above interpretation, however, crucially depends on the type of the equation of state used in the model. Chemical freeze-out in heavy ion collisions can also be determined 71 ' 187 by the condition of fixed density of the total number of baryons plus antibaryons. As it is seen in Fig. (28), within statistical uncertainties on the freeze-out parameters the above condition provides a good description of
568
P. Braun-Munzinger,
K. Redlich and J. Stachel 1
•
•
i
1
•
'. RHIC
SPS
• AGS
;* 100
^
—
n t - *-nB=cons t. GeV <E>/
—
• • O.O
,
i
0.2
m
SIS
0.4
0.6
^
0.8
/-GeV7
Fig. 28. The broken line describes the chemical freeze-out conditions of fixed total density of baryons plus antibaryons, ni, + n^ = 0.12/fm 3 from Ref. (49). The full line represents the condition of the fixed energy/particle ~ 1.0 GeV from Fig. (27). The freeze-out points are as in Fig. (27).
experimental data from the top AGS up to RHIC energy. However, in the energy range from SIS to AGS it slightly overestimates the freeze-out temperature for a given chemical potential. Consequently, e.g. the yield of the strange/non-strange particle ratios obtained at SIS turns out to be too large. The freeze-out conditions determined by the extensive thermodynamical observables are in addition very sensitive to the size and the model that describes repulsive interactions between hadronic constituents. The condition of fixed (E)/(N) ~ 1 GeV, is very insensitive to repulsive interactions. Independently on how the repulsive interactions are implemented, that is through a mean field potential 176 , an effective hard core 58 or a thermodynamically consistent implementation 38>82, the freezeout line in Fig. (27) is hardly modified. However, the energy per particle is being sensitive to the composition of the collision fireball. Considering heavy fragments like e.g. the He or Li as being the constituents of a thermal fireball would change the line shown in Fig, (27). In general these fragments
Particle Production
in Heavy Ion Collisions
569
would make the line steeper below the AGS energy. This is an open question whether at the SIS energy, the multiplicity and the spectra of such composite objects are of thermal origin and can be reproduced with the same parameters as all other hadrons. For higher energies beyond AGS, however, the above is not excluded as discussed in Section 3.04 and in Ref. (175). For the phenomenological determination of freeze-out parameters for different collision energies we use in the following the requirement (E)/(N) ~ 1 GeV. 5.1. Chemical
freeze-out
and the QCD phase
boundary
The chemical freeze-out temperature, found from a thermal analysis 38 ' 40 ' 35 of experimental data in Pb-Pb collisions at the SPS and in Au-Au collisions at RHIC energy is remarkably consistent, within errors, with the critical temperature Tc ~ 173±8 MeV obtained 77 from lattice Monte-Carlo simulations of QCD at a vanishing net baryon density. Thus, the observed hadrons seem to be originating from a deconfined medium and the chemical composition of the system is most likely to be established during hadronization of the quark-gluon plasma 6 ' 7 ' 9 . The observed coincidence of chemical and critical conditions in the QCD medium at the SPS and RHIC energy open the question if this property is also valid in heavy ion collisions at lower collision energies where the statistical order of the secondaries is phenomenologically well established. Recently, first attempts have been made to extent lattice calculations into the region of finite fis- This provided an estimate 177 ' 178 of the location of the phase boundary at finite baryon density. The generic problem of the Monte-Carlo simulation of QCD with the finite chemical potential, related with a complex structure of the fermionic determinant, was partly overcome. The reweighting method, in which the physical observables at finite HB are computed by simulating the theory at vanishing HB178 was successfully applied and first results on the phase boundary were obtained 178 . The region of applicability of this approach and uncertainties on the results due to a small lattice size and large strange quark mass are still, however, not well established. Another efficient method, at least for low baryon density, is based on the Taylor expansion in \IB of any physical observable 177 . The coefficients of the series are calculated at vanishing \IB , and thus could be obtained using a standard Monte-Carlo method. This procedure was recently applied to get a series expansion of the critical temperature in terms
570
P. Braun-Munzinger,
K. Redlich and J. Stachel
T [MeV]
200
150
!
100 •
50
const, energy density freeze—out curve 0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
Fig. 29. A Comparison of the chemical freeze-out curve from Fig. (27) with the phase boundary line. The upper thin line represents the LGT results obtained in Ref. (177) and the lower thin line describes the conditions of constant energy density that was fixed at n = 0. The upper point with crossed error bars denotes the end-point of the crossover transition from Ref. (178).
of the critical
HB177-
Fig. (29) shows the results on the position of the phase boundary that were obtained using the methods indicated above together with the freezeout curve from Fig. (27). The upper thin-line represents an extrapolation of the leading, (HB)1 order, term in the Taylor expansion of Tc to a larger values of the chemical potential 177 . It is interesting to note that, within statistical uncertainties, the energy density along this line is almost constant and corresponds to ec ~ 0.6GeV/fm3 (thin lower line in Fig. (29)), that is the same value as found on the lattice at HB = 0. It is thus, conceivable that the critical (fxcB — Tc) surface is determined by the condition of fixed energy density. This can be also argued phenomenologically. The transition from a confined to deconfined phase could appear if the particle (like in percolation models) or energy density is so large that hadrons start to overlap. It should not be important if this density is achieved by heating or compressing the nuclear matter. Thus, since the percolation type
Particle Production
in Heavy Ion Collisions
571
argument 179 is well describing critical conditions at /is = 0 it could be also valid at finite \IBFig. (29) shows that the chemical freeze-out points at SPS and RHIC energy are indeed lying on the phase boundary. The results of SPS at 40 A-GeV and top AGS are already below the boundary line. However, it is not excluded that also at these lower energies the collision fireball in the initial state appears in the deconfined phase. The initial energy density expected at AGS is of the order of 1 GeV/fm 3 (see Section 1.1) thus, it is larger than the critical energy density along the boundary line in Fig. (29). The canonical suppression effects for strangeness production, were shown to be negligible already at the top AGS energies. Here, strangeness was uncorrected and well described by the GC approach. It is quite possible that the asymptotic GC formulation and the maximal thermal phase space for strangeness is achieved if in the initial state the system was created in a deconfined, QGP phase. The abundant production of strangeness in the QQp3i t 0 g e ther with a long range correlations during a non-perturbative hadronization results in strangeness population that maximizes the entropy in the GC limit. In this context the energy range between SIS and 40 A-GeV is of particular interest and it is expected to be covered by the planned 180 for the future heavy ion experiments at GSI. 6. Particle yields and their energy dependence The hadronic composition in the final state obtained in heavy ion collisions is determined solely by an energy per hadron to be approximately 1 GeV per hadron in the rest frame of the system under consideration. This phenomenological freeze-out condition provides the relation between the temperature and the chemical potential at all collision energies. The above relation together with only one measured particle ratio, e.g. the ratio of pion/participant° as shown in Fig. (30) establishes 48 ' 53 the energy dependence of the two thermal parameters T and \IB • Consequently, predictions of particle excitation functions can be given in terms of the canonical statistical model. An alternative approach would be to interpolate and/or parameterize the energy dependence of the /xg and then using the unified freeze-out condition of (E)/(N) ~ 1 to get the energy dependence °The mean number of pion multiplicity is defined as: (it) = 1.5((7r+) + (7r - )) whereas the number of participant is calculated as the number of wounded nucleons
572
P. Braun-Munzinger,
K. Redlich and J. Stachel
•
A V
101
----^~-
:
*P-P
. . •
i.
10
•» RHIC (PHOBOS) • SPS (NA49)
c
:
/*
f
*
IZTAU
10"
E802/E859
m O+Ne JINR • Ni+Ni KAOS o Au-Au
i
i 101
s"a
',
".
10a [GeVJ
Fig. 30. T h e total number of pions per wounded nucleon ((ir)/ATw) versus the centerof-mass energy. The data at lower energies in A-A as well as in p - p collisions are from Refs. (181, 185). The RHIC results are from Ref. (182). T h e short-dashed and dashed lines are a fit to the data.
of T. The energy dependence of the chemical potential was shown 53 to be well parameterized as a (118) MBOO (1 + ^ / 6 ) where a cz 1.27 GeV and b cz 4.3 GeV. The result of this parameterization is shown by the full line in Fig. (31) together with the energy dependence of the freeze-out temperature. In the statistical approach, the knowledge of T(^/s) and fig(y/s) determines the energy dependence of different observables. Of particular interest are the ratios of strange to non-strange particle multiplicities as well as the relative strangeness content of the system that is measured by the Wroblewski factor 186 . We turn our attention first to a study of the energy dependence of the Wroblewski ratio defined as \s
2(ss) (uu) + (dd)'
(119)
where the quantities in angular brackets refer to the number of newly
Particle Production 1000
750
€" *=p
•
1
-
TAGS
—
-
V S P S
1i
200
1 1
«JT>S
%
ti
ISO
-
/ •
'1?AGS
50
RHIC T^~ ' R H I C J_
— _
100
1
573
'
^ SIS
500
2SO
1
i
in Heavy Ion Collisions
_ -
^SIS
1
Vs
(OcV)
Fig. 31. Behavior of the freeze-out baryon chemical potential \LB (upper curve) and the temperature T (lower curve) as a function of energy from Ref. (53). The temperature T as a function of beam energy is determined from the unified freeze-out conditions of fixed energy/particle.
formed quark-antiquark pairs, i.e., it excludes all the quarks that are present in the target and projectile. The quark content used in this ratio is determined at the moment of chemical freeze-out, i.e. from hadrons and especially, hadronic resonances, before they decay. This ratio is thus not easily measurable unless one can reconstruct all resonances from the final-state particles. The results are shown in Fig. (32) as a function of center of mass energy yfs. The values calculated from the experimental data at chemical freeze-out in central A-A collisions have been taken from reference (40). p All values of \ s were extracted from fully integrated data besides RHIC where the STAR collaboration results on particle ratios measured 190 at mid-pseudorapidity were used. The solid line in Fig. (32) describes the statistical model calculations 34 in complete equilibrium along the unified freeze-out curve and with the energy dependent thermal parameters shown p
Here t h e statistical model was fitted with an extra parameter "j3 to account for possible chemical undersaturation of strangeness. At the SPS, 7 S ~ 0.75 was required to get the best agreement with Ait data. See the discussion in chapter 2 concerning this issue.
574
P. Braun-Munzinger,
1.1 -• " 1 -
K. Redlich and J. Stachel
1
Thermal Model with y=\ ana radius = 7 fm Thermal Model with n B = 0.0 Thermal Model with u,„ = 0.0 and radius = 1.2 fm *
'
1 i i i | i |ii i | | i i n 1 | Mil Contributions to Xt from :
1
-
Strange mesons Hidden strangeness Sum
0.9 •
Xi=2<sS>/
0.8 •
1
AGS Si-Au
0.7 •
J
0.8 0.7 ~ r 0.6
<<" 0.6 0.5; AGS Au-Ai
Pb-PbiS^S^JRHlCAu-Au
-
,.is--Ag
0,4 •
0.5
0.3
0.2 •
0.2
0.1 -
0.1
10"
/
0.4 -
0.3 -
oC
/
10'
,
| 10'
|
Strange mesons
/
/••
\: lsjS-Au-AU|
s—^\ ^^~~"~--~
" " " " -
/
""'•..._
f/
Strange baryons.
Hidden strangeness
"
•
10'
Vs (GeV) Fig. 32. T h e Wroblewski ratio As as a function of y/s. For the description of the lines see text. The points are the statistical model results calculated with thermal parameters obtained from the fit to measured particle yields.
1
2
4
8 10
16
32
64 100
Vs (GeV)
Fig. 33. Contributions to the Wroblewski factor from strange baryons, strange mesons and hidden strange particles. Full line is a sum of all these contributions.
in Fig. (31). From Fig. (32) one sees that around 30 A-GeV lab energy the relative strangeness content in heavy ion collisions reaches a clear and well pronounced maximum. The Wroblewski factor decreases towards higher incident energies and reaches a limiting value of about 0.43. The appearance of the maximum can be related to the specific dependence of / i s on the beam energy. In Fig. (32) we also show Xs calculated under the assumption that only the temperature varies with collision energy but the baryon chemical potential is kept fixed at zero (dotted line). In this case the Wroblewski factor is indeed seen to be a smooth function of energy. The assumption of vanishing net baryon density is close to the prevailing situation in e.g. p-p and e + -e~ collisions. In Fig. (32) the results for Xs extracted from the data in p-p, p-p and e + — e - are also included 133 . The dashed line represents the results obtained with / i s = 0 and a canonical radius of 1.2 fm. There are two important differences in the behavior of
Particle Production
in Heavy Ion Collisions
K'/n*
v I
575
at An
1
!•
[GeV]
Fig. 34. The ratio of kaon to pion measured in heavy ion collision at different collisions energies. The left-hand figure describes midrapidity data, whereas the right hand figure represents the ratio of fully integrated yields. Data at SIS, AGS, SPS and RHIC energy are taken from Refs. (185, 187, 190). The short-dashed line describes the statistical model predictions along the unified freeze-out curve. The right hand figure also shows the parameterization of the p - p data (full-lines) from Ref. (191) and the canonical model results (dashed-line).
Xs in elementary compared to heavy ion collisions. Firstly, the strangeness content is smaller by a factor of two. This is mainly because in the elementary collisions particle multiplicities follow the values given by the canonical ensemble with radius 1.1-1.2 fm whereas in A-A collisions the grand canonical ensemble can be used, thus strangeness is uncorrelated and distributed in the whole fireball. Secondly, there is no evidence, at the moment, of a significant maximum in the behavior of As in elementary collisions. The importance of finite net baryon density on the behavior of \ s is demonstrated in Fig. (33) showing separately the contributions to (ss) coming from strange baryons, strange mesons and from hidden strangeness, i.e., from hadrons like 4> a n d r). As can be seen in Fig. (33), the origin of the maximum in the Wroblewski ratio can be traced to the contribution of strange baryons that is strongly enhanced in the energy range up to 30 A-GeV. The appearance of the maximum in the strangeness content of the collisions fireball can be also justified on the level of different ratios that includes strange particles. The measured 187 ' 190 K+/TT+ ratio (see also Fig. (34)) is
576
P. Braun-Munzinger,
K. Redlich and J. Stachel
a very abruptly increasing function of the collision energy between SIS up to the top AGS energy. At higher energies it reaches a broad maximum between 20 A-GeV - 40 A-GeV and gradually decreases up to RHIC energy. In microscopic transport models 192 the increase of the kaon yield with collision energy is qualitatively expected as being due to a change in the production mechanism from associated to direct kaon emission. However, hadronic cascade transport models do not, until now, provide quantitative explanation of the experimental data in the whole energy range. This is evident in Fig. (35) where the comparisons of RQMD, URQMD and BUU microscopic transport models with experimental data is presented. The RQMD
RHIC midrapitlity
0.2-
* 0.1
I
- / ft :j -
- -RQMD URQMD — Hadron Gas
•'/
10
10" \&GeV)
10
1110
1000
VA[GeV]
Fig. 35. The ratio of kaon to pion measured in heavy ion collision at different collisions energies in comparison with microscopic transport models. The left-hand figure represents the R Q M D 1 9 3 , U r Q M D 1 9 4 ' 1 2 8 and the statistical model predictions. The right hand figure shows the BUU results 1 9 2 . For the description of data see Refs. (57, 128).
provides a good description of the high energy data. The URQMD works quite well at the low energy, however, underestimates the yields at the SPS and RHIC energy. The BUU on the other hand is well suitable for SIS energy range, however, shows different energy dependence than that obtained in experiments. The statistical model in the canonical formulation 57 ,on the other hand, provides a good description of the K/n midrapidity ratio from SIS up to AGS as seen in Fig. (34). The abrupt increase from SIS to AGS and the broad maximum of this ratio are consequences of the specific dependence of thermal parameters on the collision energy and the canonical
Particle Production in Heavy Ion Collisions
577
strangeness suppression at SIS. A drop in the K+ /TT+ ratio for An yields reported 187 ' 190 by the NA49 Collaboration at 158 A-GeV (see Fig. (35) is, however, not reproduced by the statistical model without further modifications, e.g. by introducing an additional parameter 7S ~ 0.75 40 that accounts for additional suppression of strangeness. We also note that an abrupt drop in the K+/-K+ ratio is predicted 195 in a special model with particular conditions on the early stage of the collisions. This model, however, neglects completely the production of strange particles from the hadronization of gluons. 1 . 11 1 1
o.oa
1
&-/K-/3
I
1
1
0.1 0.09
i
0.07
0.04 0.O3
o.oa 0.01
1
°°°
1
yv
-££V= • \
1 «'S
* \
^-^ '•I
o.os
—i—n i i -r—"i
0.06
/ * * ^••^st-PBtiM)
to-pe
10
'
s " * [GeV]
I
1 2
°
Fig. 36. Particle ratios in A-A collisions versus the center-of-mass energy. Data at the SPS are fully integrated NA49 results. The corresponding ratio at the top AGS was obtained from E810 results on H~ measured 1 9 6 in Si-Pb collisions in the rapidity interval 1.4 < y < 2.9, normalized to the full phase-space values of 7r+ and K~ yield obtained 1 9 7 in Si-Au collisions by E802 Collaboration. The lines represent statistical model results 5 3 along the unified freeze-out curve.
The appearance of the maximum in the relative strange/non-strange particle multiplicity ratios already seen in K+ /TT+ is even more pronounced 53 for strange baryon/meson ratios. Fig. (36) shows the energy dependence of A/7r+ and S~ /n+. There is a very clearly pronounced maximum, especially in the A/n+ ratio. This maximum is related with a rather strong decrease of the chemical potential coupled with only a moderate increase in the as-
578
P. Braun-Munzinger,
K. Redlich and J. Stachel
sociated temperature as the energy increases 53 . The relative enhancement of A is stronger than that of E,". There is also a shift of the maximum to higher energies for particles with increasing number of strange quarks. This is because an enhanced strangeness content of the baryon suppresses the dependence of the corresponding ratio on / i s . This is also seen for H - /K~ ratio that shows a substantially narrower maximum since the strangeness dilution effect is compensated by the strangeness content of the K~. The actual experimental data for both A/TT+ and E~ /TV+ ratios shown in Fig. (36) are following the predictions of the statistical model. However, as in the case of kaons, midrapidity results are better reproduced by the model than Ait data. 11inij—i—i
11ini|—i
i 11 I I I I J —
,
S P S - - - +1 , - ' ' * RHIC:
Au+Au RHIC 70 + 70 AGeV • I - S - - i - l Pb+Pb 158 AGeV
^ \
A 6 S
0.1
#'' ?'' Au+Au
SIS
••••*
4 - - - V - T T * -
Au+Au 10.2 AGeV _
0.1 ,?
Ni+Ni Ni+Ni 1.93 AGeV - B - d - o - G J.-.A-.y..*...
0.01 i i mil
"'I
0.1
' ' ' ' ""I
1
M(ji°+JT)/Apart
10
1.5 AGeV 0.01 0.0
j
0.2
i
i
0.4
i
i
0.6
i
i,
0.8
i
1.0
centrality
Fig. 37. The left-hand figure, shows the ratio of K+/K~ as a function of (TT + — it0)/Apart- Points are the experimental results, a line is the statistical model result along the unified freeze-out curve. The right-hand figure shows K~/K+ ratio that appears to be constant as a function of centrality from SIS up to RHIC energy. Data are from the STAR, NA49, E866, and KaoS Collaborations. The broken lines are statistical model results.
The statistical model predicts that if at least two different ratios of non strange particles are constant with centrality then also strange particle/antiparticle ratio should be centrality independent. Dynamically this is a rather surprising result as strange particles and their antiparticles are generally produced and absorbed in the surrounding nuclear medium in a
Particle Production
in Heavy Ion Collisions
579
different way. This is particularly the case at lower energies (up to AGS) where e.g. K+ and K~ are predominantly produced due to -KN —> KK+ and 7rA —> K~N processes. In addition K+, mesons feel a repulsive potential whereas K~ mesons are attracted. Thus, the prediction of the thermal model that the K+ jK~ ratio is centrality independent was dynamically unexpected. Figure (37) represents the energy and centrality dependence of the K+ /K~ ratio from SIS to RHIC energy. The statistical model predictions are seen in Fig. (37 -right) to agree remarkably well with the data. The results of Fig. (37 -right) could be considered as the evidence of an apparent chemical equilibrium population of kaons in the final state. This behavior of data is also seen on a different level. The chemical equilibration of the associated production of K+ with a hyperon and strangeness exchange production of K~, indicated above, should result in a linear dependence of the K+ jK~ ratio on n/Apart198. Fig. (37 - left) shows that, indeed, in the energy range from SIS up to AGS, and almost independently from the colliding system, the above prediction is valid. It is also clear from this figure that between AGS and SPS the production mechanism of strange mesons is changing. The results for the K+ and K~ excitation function show an interesting behavior when expressed as a function of available energy y/s — ^/s^h (see Fig. (38 -left) 49 . The threshold energy s/sth corresponds to the production threshold in N-N collisions. For K+ mesons yfsth = 2.548 GeV, whereas for K~ the corresponding value is y/stii. = 2.87 GeV. It turned out that in this representation the measured yields of K+ and K~, close to threshold, in heavy ion collisions are about equal while they differ in p-p collisions by a factor of 10 - 100 199 . Fig. (38 -left) shows the experimental data together with the canonical statistical model results 49 . In the range where yfs — yfs^. is less than zero, the excitation functions for K+ and K~, obtained in the model, cross each other, leading to the observed equality of K+ and K~ at SIS energies. The yields differ at AGS energies by a factor of five. The difference in the rise of the two excitation functions can be understood in the statistical model as being due to the different dependence of K+ and K~ yields on fig. The K+ yield is strongly fig dependent through associated production with A whereas K~ yield is not directly effected by HB- Consequently, the excitation functions, i.e. the variation with T, exhibit a different rise for kaons and anti-kaons. In the statistical model all particle species are considered to be on shell.
580
P. Braun-Munzinger,
0
K. Redlich and J. Stachel
1
2
s v»
[GeV]
V s - V s ^ (GeV) Fig. 38. The left-hand figure: calculated K+/Apart and K~ /Apart ratios of yield/participant in the statistical model as a function of i/s — y/sth f° r Ni+Ni collisions. The points are the results for Ni+Ni collisions at S I S 1 1 9 ' 1 2 0 and for A u + A u collisions at 10.2 A-GeV (AGS) 1 2 1 energies. The right-hand figure: the yield of 4>/K~ ratio in Ni-Ni and Au-Au collisions calculated 2 0 8 in the statistical model along the unified freeze-out curve. The data points are from Ref. (203)
Thus, the statistical model reproduces the kaon yields and their excitation functions with current particle masses. A transport calculation 130 ' 192 ' 128 , on the other hand, required in-medium modifications (a reduction) of the K~ mass (as expected for kaon in the nuclear medium) in order to describe the measured yields. These differences are, however not necessarily in contradiction, as the transport model describes a time evolution of particle production, whereas the statistical models are only valid for particles that are measured in the final state, where the on-shell conditions are to be expected.*1 It is conceivable that the apparent chemical equilibration observed in the data at SIS energies could be a direct consequence of in-medium effects. One such possibility is an increase of the in-medium production cross section of strange particles. In the context of particle production in heavy ion collisions a particular role has been attributed to the vector meson resonances. The measureq
Nevertheless, in transport models the kaons and anti-kaons are produced in different time during the collision, thus also at different temperatures. In the statistical model there is a common temperature for kaons and their antiparticles.
Particle Production
in Heavy Ion Collisions
581
ments of these particles could possibly provide an information on the chiral symmetry restoration 4,19,24,25 ' 27 and in medium effects due to the collision broadening 115 of their decay widths. The production of
T h e results were obtained in the canonical model without strangeness suppression factor.
582
P. Brawn-Munzinger,
K. Redlich and J. Stachel
centrality dependence that increases with decreasing beam energy. Thus, the
7. Lifting of the strangeness suppression in heavy ion collisions An enhanced production of strange particles compared to the suppressed strangeness yield observed in collisions between elementary particles was long suggested 31,42 ' 43 as a possible signal of the QGP formation in heavy ion collisions.8 In the QGP the production and equilibration of strangeness is very efficient due to a large gluon density and a low energy threshold for dominant QCD processes of ss production 31,32 . In hadronic systems the higher threshold for strangeness production was argued in Ref. (31) to make the strangeness yield considerably smaller and the equilibration time much longer.4 Based on such arguments predictions have been developed for experimental signatures of deconfinement. Key predictions are 31,52 : i) the disappearance of the strangeness suppression observed in collisions s
Strangeness enhancement in heavy ion collisions was recently discussed and interpreted at the parton level in Ref. (211) without requirements of the Q G P formation in the initial state. 'Recently, it was argued in Refs. (128, 33), that multi-mesonic reactions could accelerate the equilibration time of strange antibaryons especially when the hadronic system is hot and very dense. This argument does not apply, however, for strange baryon, where also strong enhancements are seen.
Particle Production
in Heavy Ion Collisions
583
S
-
WA97
•
NA57
-*--»-•**
4
Yield/wound, nucl. re ativc to
I
i
A
Yield/wound, nucl. re ative to p+Bc
among elementary particles leading to e.g. an enhancement of multistrange baryons and anti-baryons in central A-A collisions, with respect to proton induced reactions. ii) chemical equilibration of secondaries: the appearance of the QGP being close to a chemical equilibrium and subsequent phase transition should in general drive hadronic constituents produced from hadronizing QGP towards chemical equilibrium. Heavy ion experiments at the CERN SPS reported 212 ' 213 actually a global increase of strangeness production from p-p, p-A to A-A collisions. This effect is seen e.g. in Fig. (34 -right) that shows the enhancement of the K+/TT+ ratio in Pb-Pb relative to p+p collisions. There is indeed an increase by a factor of about two in strangeness content when going from p+p to heavy ion collisions. This is also seen on the level of the Wroblewski factor in Fig. (32).
,
a
WA97
'•'
NA57
-hftf °'^
4411
i
\ pBe pPb
PbPb
PBe pPb
PbPb
Fig. 39. Particle yields per participant in P b - P b relative to p-Be and p - P b collisions centrality dependence. The d a t a are from WA97 2 1 2 and NA57 2 1 3 Collaborations.
A large strangeness content of the QGP plasma should, following (ii), be reflected in a very specific hierarchy of multistrange baryons 31 : implying an enhancement of Q > H > A. Fig. (39) shows the yield/participant in Pb-Pb relative to p-Be and p-Pb collisions measured 212 ' 213 by the WA97 and NA57 Collaborations. Indeed the enhancement pattern of the antihyperon yields is seen to increase with strangeness content of the particle and
584
P. Braun-Munzinger,
K. Redlich and J. Stachel
in WA97 data there is a saturation of this enhancement for Nwound > 100. The recent results of the NA57 collaboration are showing in addition an abrupt change of the anti-cascade enhancement for a lower centrality. Similar behavior was previously seen on the level of K+ yield measured 214 at Qlab = 0 by the NA52 experiment in Pb-Pb collisions. These results are very interesting as they might be interpreted as an indication of the onset of a new dynamics. However, a more detailed experimental study and theoretical understanding are still required here. Until now there is no quantitative understanding of this exceptional threshold behavior of S. A number of different mechanisms were considered to describe the magnitude of the enhancement and centrality dependence of (multi)strange baryons measured 4 8 , 1 8 9 ' 2 0 9 ' 2 1 5 - 2 2 0 by the WA97 Collaboration. Studies using microscopic transport models make it clear that data shown in Fig. (39) can not be explained by pure final state hadronic interactions. The combination of the former with additional pre-hadronic mechanisms like sting formation and their subsequent hadronization, baryon junction mechanism, color ropes or a color flux tubes overlap improves the agreement with the measured enhancement pattern and the magnitude for the most central collisions. However, the detailed centrality dependence is still not well reproduced within the microscopic models. The results of Section 2 have shown that the statistical model provides a satisfactory description of strange and multistrange particle yields in A-A collisions. The midrapidity data were argued in Section. 1 to be reproduced by the statistical model in a full equilibrium. The fully integrated results, on the other hand can be successfully reproduced when including a strangeness undersaturation parameter 7 S ~ 0.75 that accounts, at the SPS, for a 25% deviation from equilibrium. Strangeness production in p-p collisions was shown in Section 5 to be consistent with the canonical statistical model. The abundance of singlestrange particles could be described by this model by including, as in heavy ion collisions, the strangeness undersaturation factor j a ~ 0.51 or a correlation volume V ~ Vp that accounts for the locality of strangeness conservation. Consequently, the enhancement from p-p to A-A collisions of strangeness 1 particles could be well described in terms of the statistical model as the transition from the canonical to the asymptotic GC limit 48 . One of the consequences of the canonical model is a very particular volume dependence of multistrange particle densities. In heavy ion collisions
Particle Production
•inJ
10
1
u n i i l
100
I
i l l
in Heavy Ion Collisions
585
10" K
1000
10'
10'
10J
this volumes is usually assumed to scale with the number of participants. Prom Eq. (97) it is clear that the canonical suppression should increase with strangeness content of the particles. Indeed the approximate strangeness suppression factor Fs ~ Is(Vx)/Io(Vx) for a fixed Vx is a decreasing function of s. This is particularly evident in the limit of (Vx) << 1 where Is(Vx)/I0(Vx) ~ (Vx)s. The suppression factor is quantified in Fig. (40 -left) that indicates the expected suppression pattern. Fig. (41 -left) shows predictions 48 of the canonical model for the multiplicity/participant of fl, S, and A relative to their value in p-p collisions. Thermal parameters, T = 168 MeV and \IB = 266 MeV, that are appropriate 38 for a description of central Pb-Pb collisions at 160 A-GeV, were used here and assumed to be centrality-independent." The canonical volume parameter was taken to be proportional to the number of projectile "For a detailed discussion of the possible centrality dependence of thermal parameters see e.g. Ref. (209).
586
P. Braun-Munzinger,
K. Redlich and J. Stachel
participants, V ~ V0Npan/2 where V0 = 47ri?^/3 with Rp = 1.15 ± 0.5. The volume Vb is then, depending on the particular choice of Rp, of the order of the volume of a nucleon. Figure (40 -right) indicates that the statistical model in the canonical ensemble reproduces the basic features of WA97 data: the enhancement pattern and enhancement saturation for large ^4part- The appearance of the saturation of the enhancement indicates that the grand canonical limit was reached. It is also clear from Fig. (41-left) that this saturation is shifted towards a larger centrality with increasing strangeness content of the particle.
1 1 Mill,
1 1 MHM|
|
1000
i
i
T
:
; ' •
n 100
/ /
e a.
10
-
''
/
\l
!/
E9T7
• E866 • NA49
1^ ^
1
r
*'»/ : 1
\ ,,,
|
10
m|
100 <"•»*>
•
1000
2
4
6
B
10
12
14
16
IB
-
s /2
'
fGeVJ
Fig. 41. The left hand figure: the statistical model results on the centrality dependence of the relative enhancement of A, S and fi yields/participant in central P b - P b to p - p collisions at •/s = 8.73 GeV. The right hand figure: K+/n+ ratio in A + A relative to p + p collisions. For the compilation of data, see Refs. (191, 190). The dashed line represents the statistical model results.
The essential prediction 44 of the canonical statistical model is that the strangeness enhancement from p-p to A-A collisions should increase with decreasing collision energy. This is a direct consequence of e.g. Eq. (97) where the canonical suppression factor is seen to be a decreasing function of temperature and thus also collision energy. Fig. (41 -right) shows the compilation of data on the K+/n+ ratio in A-A relative to p-p collisions from Ref. (191, 190). This double ratio could be referred to as a
Particle Production in Heavy Ion Collisions
587
strangeness enhancement factor. The enhancement is seen in data to be the largest at the smallest beam energy and is decreasing towards higher energy. The line is a smooth interpolation between the canonical model results for ^/s = 17.3,12.3,8.73,5.56 GeV, calculated with the thermal parameters T and /xg that were extracted from Fig. (27). The canonical volume parameter was taken the same as used in Fig. (16- right). The enhancement seen in Fig. ( 41 -right) is due to the suppression of the K+/TT+ ratio in p-p collisions with decreasing energy and not due to a dilution of this ratio by excess pions in the A-A system. The K+ /ir+ ratio is known experimentally not to vary within 30% in the energy range from y/s ~ 5 GeV at AGS up to ^ = 130 GeV at RHIC 187 - 190 ' 191 . Recent data of NA49 187 and CERES 1 6 8 , 1 8 9 Collaborations on A yields exhibit a similar energy dependence of the enhancement factor as is seen in Fig. (41- right) for kaons. The canonical model (see Eq. (97)) also predicts that the multistrange baryon enhancement from p-p to A-A should be larger at lower collision energy. Fig. (41- left) shows the canonical model results for the strangeness enhancement and its centrality dependence for Au-Au collisions at 40 A-GeV. The qualitative behavior of the enhancement is like that at the SPS. However, the strength of the enhancement is seen to be substantially larger. For Q it can be as large as a factor of 100. The enhancement of S in A-A relative to p-p can be deduced from data. Indeed, combining the Si-Pb results for E~~ production obtained 196 by E810 Collaboration and the Si-Au results for pion or K~ yields obtained 197 by E802 Collaboration in collisions at the top AGS energy, one can estimate that E~ /ir+ ~ 0.0076. Within errors this agrees with the value of ET /ir+ ~ 0.0074 obtained 138 by NA49 in Pb-Pb at s/s = 17.3 GeV, as seen in Fig. (36). In p-p collisions the E~/TT+ ratio is obviously a strongly decreasing function of beam energy. From the above one would therefore expect that the relative enhancement of S~ from p p to A-A collisions should be larger at AGS than at SPS energy. This is seen in Fig. (41 -left). In addition ratios containing only newly produced particles, such as e.g. E~/K~, are also larger at AGS than at SPS energy, see Fig. (36). The above heuristic argument should be, however, verified by more complete experimental data. Recently, the production of E~ was study 218 in the relativistic transport model in the energy range from 2-11 A-GeV . The authors discussed the equilibration and the influence of a phase transition on E yields in A-A
588
P. Braun-Munzinger,
K. Redlich and J. Stachel
collisions. There, it was argued that, for beam energies above 4 A-GeV, there should be a sharp increase of the E~ yield if there is a deconfinement transition in the collisions fireball. This is an interesting conjecture. However, with the presently available experimental data this prediction cannot be tested. The canonical statistical model, predicts continuous increase of S yield with collisions energy as it is seen in Fig. (36). The prediction of the statistical model on the energy dependence of strange particle yields is also in contrast with the UrQMD results 219 . There, the production of strangeness is very sensitive to the initial conditions. In UrQMD the early stage multiple scattering may imply an increase of the colour electric field strength due to an overlap of produced strings 219 . Consequently, according to the Schwinger mechanism, this should increase the production of (multi)strange baryons. Under similar kinematical conditions as at the SPS, the UrQMD model predicts 219 at RIIIC an increase in relative strength of Q yields from p-p to A-A by a factor of 5. A recent analysis of multistrange baryon yields in Au-Au collisions at RHIC energy in the context of Dual Parton Model, leads 139 to a smaller increase of the enhancement at RHIC energies. The interpretation that the strangeness enhancement is explained by canonical effects is not necessarily in contradiction with the heuristic predictions 31 that strangeness enhancement and its pattern are due to a quarkgluon plasma formation in A-A collisions. This is particularly the case if one connects the asymptotic grand canonical description of strangeness production in A-A collisions with the formation of a QGP in the initial state. 8. Conclusions and outlook This article has focussed almost exclusively on the use of the statistical approach to understand yields of different particle species that have been measured in heavy ion collisions. We have discussed a statistical description of the conservation laws and described their kinetic implementations. We have argued, both on the qualitative and quantitative level, that exact conservation of quantum numbers is of crucial importance when applying the methods of statistical physics in the context of heavy ion and particle collisions. We have presented the systematics of particle production in heavy ion collisions from SIS up to LHC energy and discussed particular properties of strangeness production. One of the most intriguing results that comes from
Particle Production
in Heavy Ion Collisions
589
these investigations is the observation that particles seem to be produced according to the principle of the maximal entropy. In a very broad energy range, from yfs ~ 2.5 up to 200 GeV per nucleon pair, hadronic yields and their ratios observed in heavy ion collisions resemble a chemical equilibrium population along a unified freeze-out curve determined by the conditions of fixed energy/hadron ~ lGeV or complementary above SIS energy by fixed total density of baryons. Strangeness production follows this systematics from low to very high energy. However, there are some characteristic features of the system at chemical freeze-out in high energy central A+A collisions regarding strangeness production that are not present in low energy heavy ion and collisions among elementary particles. In nucleus-nucleus collisions, strangeness is un-correlated and redistributed in the macroscopic volume of a collision fireball and is conserved on the average. In hadronhadron collisions the thermal phase space available for strange particles is strongly suppressed since, with only few particles produced per event, strangeness is strongly correlated in a volume that approximately coincides with the size of a nucleon, i.e. a distance over which color is confined. Thus, following the statistical kinetics, strangeness has to be conserved exactly and locally. The associated production and locality of strangeness conservation is, in the context of the statistical model, the origin of the suppression of the thermal phase space for produced particles. Within this context the strangeness suppression observed in collisions among elementary particles finds its natural explanation. We also note that the suppression increases with the strangeness content of the particle as well as, for A-A collisions, with decreasing collision energy. A further consequence of the transition from the canonical to the grand-canonical regime is that strangeness production should be enhanced in A-A collisions compared to p-p collisions. At SPS and RHIC energies the freeze-out points approach the calculated QCD phase boundary. This fact lends strong support to the interpretation that the matter produced in nuclear collisions at SPS and RHIC energies was first thermalized in the deconfined quark-gluon plasma phase and subsequently expanded through the phase boundary into a thermal gas of mostly elastic and quasi-elastic interacting hadrons. The above connection between the QCD phase boundary and the observed chemical freezeout points is sometimes called into question 221 since hadron production in e+e~ and p-p or p-p can also be described in thermal models yielding a (apparently universal) temperature T e « 170 MeV. While the fact that
590
P. Braun-Munzinger,
K. Redlich and J. Stachel
T e is close to T values determined for heavy ion collisions at top SPS and RHIC energies might indeed reflect the fundamental hadronization scale of QCD, we have already noted above that there is an essential difference between thermal descriptions of central heavy ion collisions and elementary particle reactions. Strange particle densities and their ratios can be, for heavy ion collisions at full AGS energy and higher, well described in the grand-canonical ensemble. In contrast, for a description of particle production in elementary collisions, local quantum number conservation, that is the canonical description is need. Consequently, in central nucleus-nucleus collisions at ultra-relativistic energies, strangeness percolates freely over volumes of thousands of fm3! whereas in elementary processes is approximately restricted to the size of nucleon. At top SPS and RHIC energies it is natural to conclude that in nucleus-nucleus collisions the percolation has its origin in the quark-gluon phase, lending further strong support to the interpretation that the "coincidence" between experimentally determined chemical freeze-out points and the calculated phase boundary implies that a deconfined phase was produced in such collisions. In the context of statistical physics the fact that the measured particle yields coincides with a thermal multiplicities calculated with a given statistical operator is a necessary and sufficient conditions to maintained thermalization of the collisions fireball. In the sense of Gibbs interpretation of thermodynamics this implies that the parameters T and /ijg reflect the thermal properties of the fireball. Furthermore in heavy ion collisions there are experimental observables indicating the appearance of thermodynamical pressure and correlations that are expected in a thermalized medium. The build-up of pressure and collectivity is seen in heavy ion collisions on the level of particle transverse momentum distributions and elliptic flow parameters. An increase of average particle transverse momentum with particle mass, seen from SIS up to RHIC energy, is a typical property of a transversally expanding thermal medium. The appearance of strong elliptic flow is an indication of thermodynamical pressure that develops in the early stage in the collision. Finally, the measured particle number fluctuations are consistent with thermal expectations. Taken together, these observations lend strong support to the thermodynamic interpretation of T and \IB with the concomitant QGP interpretation.
Particle Production in Heavy Ion Collisions 591 Acknowledgments K.R acknowledges stimulating discussions with R. Baier, S. Bass, M. Bleicher, J. Cleymans, B. Friman, F . Karsch, V. Koch, W. Norenberg, H. Oeschler, H. Satz, R. Stock, A. Tounsi, Nu Xu and t h e support of the Alexander von Humboldt Foundation. References 1. For a review see eg. H. Satz, Rep. Prog. Phys. 63 (2000) 1511; S.A. Bass, M. Gyulassy, H. Stocker, and W. Greiner, J. Phys. G25 R l (1999). E.V. Shuryak, Phys. Rep. 115 (1984) 151; M.G. Alford, hep-ph/0209287, H. Satz, hep-ph/0209181. 2. J. Cleymans, R.V. Gavai and E. Suhonen, Phys. Rep. 130 (1986) 217. 3. B. Alessandro, et al., CERN-ALICE-INTERNAL-NOTE-2002-025. 4. G.E. Brown and M. Rho, Phys. Rep. 363 (2002) 85. 5. W. Cassing and H. Bratkovskaya, Phys. Rep. 308 (1999) 65. 6. P. Braun-Munzinger, and J. Stachel, Nucl. Phys. A606 (1996) 320; Nucl. Phys. A683 (1998) 3; J. Stachel, Nucl. Phys. A654 (1999) 119c. 7. R. Stock, Phys. Lett. B456 (1999) 277. 8. R. Baier, A.H. Mueller, D. Schiff, and D.T. Son, Phys. Lett. B 502 (2001) 51; Nucl. Phys. A698 (2002) 217. 9. U. Heinz, Nucl. Phys. A685 (2001) 414 and A661 (1999) 349. 10. K. J. Eskola, K. Kajantie, P. Ruuskanen and K. Tuominen, Nucl. Phys. B570 (2000) 379. 11. J. Alam, et al., hep-th/010802; Ann. Phys. 286 (2001) 159. 12. D. K. Srivastava, Eur. Phys. J. 10 (1999) 487. 13. R. Rapp, Phys. Rev. C63 (2001) 054907. 14. V. Ruuskanen, Acta. Phys. Polon. B18 (1987) 551; R. Rapp, and E. Shuryak, Phys. Lett. B473200013; I. Krasnikova, Ch. Gale, and D.K. Srivastava, Phys. Rev. C65 (2002) 064903, hep-ph/0112139; Ch. Gale, Nucl.Phys. A698 (2002) 143c, hep-ph/0104235. 15. R. Rapp and J. Wambach, Phys. Rev. C55 (1998) 916; Nucl. Phys. A638 (1998) 171c; Eur. Phys. J. A6 (1999) 415. 16. K. Gallmeister, B. Kampfer, and O.P. Pavlenko, Phys. Lett. B473 (2000) 20. 17. F. Karsch, K. Redlich, and L. Turko, Z. Phys. C60 (1993) 519. 18. K. Kajanie, J. Kapusta, L.D. McLerran, and A. Mekjian, Phys. Rev. D34 (1986) 2746 . 19. T. Hatsuda, Nucl. Phys. A698 (2002) 243; K. Yokokawa, et al., Phys. Rev. C66 (2002) 022201; D. Jido, et al., Phys. Rev. D63 (2001) 011901. 20. D. Teaney, J. Lauret, and E. Shuryak, Phys. Rev. Lett. 86 (2001) 4783. 21. U.A. Wiedemann and U. Heinz, Phys. Rep. 319 (1999) 145, and references therein.
592
P. Braun-Munzinger,
K. Redlich and J. Stachel
22. B. Tomasik, U.A. Wiedemann and U. Heinz, Nucl. Phys. A663 (2000) 753; R. Nix, Phys. Rev. C58 (1998) 2303. 23. Nu Xu, nucl-ex/0211012. 24. V. Metag, Nucl. Phys. A690 (2001) 140. 25. W. Weinhold, B.L. Friman and W. Norenberg, Acta. Phys. Polon. B27 (1996) 3249; Phys. Lett. B433 (1998) 236; M. Hermann, B.L. Friman and W. Weinhold, Nucl. Phys. A545 (1992) 2676. 26. G. Boyd, et al., Nucl. Phys. Proc. Suppl. 42 (1995) 469; Phys. Lett. B349 (1995) 170; E. Laermann, et al., Nucl. Phys. Proc. Suppl. 34 (1994) 292. 27. M.F.M. Lutz, G. Wolf and B. Friman, Nucl. Phys. A706 (2002) 431; Nucl. Phys. A661 (1999) 526. 28. T. Matsui and H. Satz, Phys. Lett. B178 (1986) 416. 29. H. W. Barz, H. Schulz, B.L. Friman and J. Knoll, Phys. Lett. B254 (1991) 315, Nucl. Phys. A545 (1992) 259. 30. T.S. Biro, hep-ph/0005067, hep-ph/0205049; T.S. Biro, Phys. Lett. B487 (2000) 133. 31. P. Koch, B. Miiller, and J. Rafelski, Phys. Rep. 142 (1986) 167; J. Rafelski, J. Letessier, and A. Tounsi, Acta Phys. Polon. B27 (1996) 1037; J. Rafelski Phys. Lett. B262 (1991) 333. 32. T. Matsui, B. Svetitsky and L.D. McLerran, Phys. Rev. D34 (1986) 783. 33. R. Rapp and E. Shuryak, Phys. Rev. Lett. 86 (2001) 2980; C. Greiner and S. Leupold, J. Phys. G27 (2001) L95. 34. J. Cleymans, and K. Redlich, Phys. Rev. Lett. 81 (1998) 5284. 35. P. Braun-Munzinger, D. Magestro, K. Redlich, and J. Stachel, Phys. Lett. B518 (2001) 41 and references therein. 36. D. Magestro, J. Phys. G28 (2002) 1745. 37. P. Braun-Munzinger, J. Stachel, J.P. Wessels and N. Xu, Phys. Lett. B344 (1995) 43 and B365 (1996) 1. 38. P. Braun-Munzinger, I. Heppe and J. Stachel, Phys. Lett. B465 (1999) 15. 39. S.V. Akkelin, P. Braun-Munzinger and Yu. M. Sinyukov, Nucl. Phys. A710 (2002) 439. 40. F. Becattini, J. Cleymans, A. Keranen, E. Suhonen and K. Redlich, Phys. Rev. C64 (2001) 024901. 41. A. Keranen, and F. Becattini, J. Phys. G28 (2002) 2041; Phys. Rev. C65 (2002) 044091. 42. J. Rafelski and J. Letessier, nucl-th/0209080. 43. M. Gazdzicki, et al., Z. Phys. C65 (1995) 215, Acta Phys. Polon. B30 (1999) 2705. 44. A. Tounsi and K. Redlich, Eur. Phys. J. C24 (2002) 529; J. Phys. G28 (2002) 2095. 45. F. Becattini, hep-ph/0202071. 46. J. Sollfrank, et al., Phys. Rev. C55 (1997) 392. 47. P. Huovinen, et al., Phys. Lett. B503 (2001) 58. 48. J. S. Hamieh, K. Redlich and A. Tounsi, Phys. Lett. B486 (2000) 61; J.
Particle Production
in Heavy Ion Collisions
593
Phys. G27 (2001) 413. 49. J. Cleymans and K. Redlich, Phys. Rev. C60 (1999) 054908; J. Cleymans, H. Oeschler and K. Redlich, Phys. Rev. C59 (1999) 1663; Phys. Lett. B485 (2001) 27. 50. J. Cleymans, D. Elliott, A. Keranen and E. Suhonen, Phys. Rev. C57 (1998) 3319. 51. D. Zschiesche et a l , Nucl. Phys. A681 (2001) 34. 52. J. Letessier, and J. Rafelski, Int. J. Mod. Phys. E9 (2000) 107. 53. P. Braun-Munzinger, J. Cleymans, H. Oeschler and K. Redlich, Nucl. Phys. A697 (2002) 902. 54. N. Xu, M. Kaneta, Nucl. Phys. A698 (2002)306, nucl-ex/0104021; M. Kaneta and N. Xu, J. Phys. G27 (2001) 589; Nucl. Phys. A698 (2002) 306. 55. W. Broniowski and W. Florkowski, Phys. Rev. Lett. 87 (2001) 272302. 56. F. Becattini and G. Passaleva, Eur. Phys. J. C23 (2002) 551, hepph/0110312. 57. K. Redlich, Nucl. Phys. A698 (2002) 94. 58. J. Cleymans and H. Satz, Z. Phys. C57 (1993) 135. 59. D. Magestro, P. Braun-Munzinger, and J. Stachel, in preparation. 60. I. Heppe, Diploma thesis, Heidelberg(1998). 61. G. Yen, M.I. Gorenstein, W. Greiner, S.N. Yang, Phys. Rev. C 56, (1997) 2210. 62. see, e.g., A. Bohr and B. Mottelson, Nucl. Structure (Benjamin, New York 1969), Vol. 1, p. 266. 63. D. Ouerdane et al., BRAHMS Coll., QM2002 contribution, Nucl. Phys. A (in print), nucl-ex/0212001. 64. T. Ullrich, QM2002 contribution, nucl-ex/0211004, Nucl. Phys. A (in print). 65. J. Cleymans in 3rd International Conference on Physics and Astrophysics of Quark Gluon Plasma (ICPAQGP 97), Jaipur, India, 17-21 Mar 1997. 66. H. Appelshauser et al., NA49 Coll., Phys.Rev.Lett. 82 (1999) 2471. 67. T.S. Biro, P. Levai, J. Zimanyi, J.Phys. G28 (2002) 1561, hep-ph/0112137; J. Zimanyi, et al., in S.A. Bass, et al., Nucl. Phys. A661 (1999) 205c. 68. J. Rafelski, J. Letessier, Qm2002 contribution, nucl-th/0209084, Nucl. Phys. A(in print); J. Rafelski, et al., in S.A. Bass, et al., Nucl. Phys. A661 (1999) 205c. 69. P. Fachini et al., STAR Coll., QM2002 contribution, nucl-ex/0211001, Nucl. Phys. A (in print). 70. D. Adamova et al., CERES coll., Nucl.Phys. A714 (2003) 124, nuclex/0207005. 71. D. Adamova et al., CERES coll., Phys.Rev.Lett. 90 (2003) 022301, nuclex/0207008. 72. M. van Leeuwen et a l , NA49 coll., QM2002 contribution, Nucl. Phys. A (in print), nucl-ex/0208014. 73. J. Barrette et al., E814 Coll., Phys. Lett. B333 (1994) 33. 74. N. Xu et a l , E814 Coll., Nucl. Phys. A566 (1994) 585c.
594 P. Braun-Munzinger, K. Redlich and J. Stachel 75. P. Braun-Munzinger and J. Stachel, Phys. Lett. B490 (2000)196 ; Nucl. Phys. A690 (2001) 119c. 76. P. Braun-Munzinger and J. Stachel, Nucl. Phys. A690 (2001) 119c. 77. F. Karsch, E. Laermann, and A. Peikert, Nucl. Phys. B605 (2001)579. 78. J.D. Bjorken, Phys. Rev. D27 (1983) 140. 79. F. Karsch, K. Redlich, and A. Tawfik, hep-ph/0303108. 80. P. Braun-Munzinger, J. Stachel, and C. Wetterich, in preparation. 81. J.P. Blaizot, E. Iancu, and A. Rebhan, Phys. Rev. D63 (2001) 065003. 82. E. Suhonen and S. Sohlo, J. Phys. G13 (1987) 1487; D. H. Rischke, M. I. Gorenstein, H. Stocker and W. Greiner, Z. Phys. C51 (1991) 485. 83. A. Peshier, B. Kampfer, and G. SofT, Phys. Rev. C61 (2000) 045203; Phys. Rev. D66 (2002) 094003; B. Kampfer, and O.P. Pavlenko, Phys. Lett. B391 (1997) 185. 84. R. Pisarski, Phys. Rev. D62 (2000) 111501. 85. A. Dumitru and R.D. Pisarski, Phys. Lett. B525 (2002) 95; A. Dumitru and R.D. Pisarski, Phys. Rev. D66 (2002) 096003, hep-ph/0204223. 86. M. Agarwal, et a l , WA98 Coll., Eur. Phys. J. C18 (2001) 651. 87. L. V. Gribov, et al., Phys. Rep. 100 (1983) 1; A.H. Mueller and J. Qiu, Nucl. Phys. B268 (1986) 427. 88. L.D. McLerran, and R. Venugopalan, Phys. Rev. D49 (1994) 2233; Phys. Rev. D49 (1994) 3352; Phys. Rev. D50 (1994) 2225. 89. D. Kharzeev and M. Nardi, Phys. Lett. B507 (2001) 121. 90. E.V. Shuryak, and L. Xiong, Phys. Rev. Lett. 70 (1993) 2241; J. Kapusta, L. McLerran, and D.K. Srivastava, Phys. Lett. B283 (1992) 145. 91. Particle Properties 92. D. K. Srivastava, M. G. Mustafa, and B. Miiller, Phys. Rev. C56 (1977) 1064; R.J. Fries, B. Miiller and D. K. Srivastava, nucl-th/0208001. 93. K. Geiger and D. K. Srivastava, Nucl. Phys A661 (1999) 592, nuclth/9808042. 94. K. Huang, Statistical Mechanics, 2nd Edition, Wiley, New York, 1987, sect. 3.4. 95. T. A. Armstrong, et al., E864 Coll., Phys. Rev. C61 (2000) 064908. 96. T. A. Armstrong, et al., E864 Coll., Phys. Rev. Lett. 85 (2000) 2685. 97. P. Braun-Munzinger, and J. Stachel, J. Phys. G: Nucl. Part. Phys. 21 (1995) L17. 98. A.J. Baltz et a l , Phys. Let. B325 (1994) 7. 99. C M . Ko, V. Koch, Z. Lin, K. Redlich, M. Stephanov and X.N. Wang, Phys. Rev. Lett. 86 (2001) 5438. 100. S. Jeon, V. Koch, K. Redlich and X.N. Wang, Nuc. Phys. A697 (2002) 546; K. Redlich, V. Koch and A. Tounsi, Nucl. Phys. A702 (2002) 326; K. Redlich, J. Cleymans, H. Oeschler, and A. Tounsi, Acta Phys. Polon. B33 (2002) 1609. 101. Z. Lin, and C M . Ko, Phys. Rev. C64 (2001) 041901. 102. H.-Th. Elze, M. Gyulassy, and D. Vasak, Nucl. Phys. B276 (1986) 706.
Particle Production
in Heavy Ion Collisions
595
103. J.P. Blaizot, and E. Iancu, Nucl. Phys. B 570 (2000) 326. 104. Q. Wang, K. Redlich, H. Stocker, and W. Greiner, Phys. Rev. Lett. 88 (2002) 132303; Nucl. Phys. A714 (2003) 293; J. Phys. G28 (2002) 2115. 105. R. Hagedorn, Thermodynamics of strong interactions, CERN Report 71-12 (1971); E.V. Shuryak, Phys. Lett. B42 (1972) 357; Rafelski and M. Danos, Phys. Lett. B97 (1980) 279; R. Hagedorn and K. Redlich, Z. Phys. C27 (1985) 541. 106. R. Hagedorn and K. Redlich, Z. Phys. C27 (1985) 541. 107. K. Redlich and L. Turko, Z. Phys. B97 (1980) 279; L. Turko, Phys. Lett. B104 (1981) 153; L. Turko and J. Rafelski, Eur. Phys. J. C18 (2001) 587. 108. K. Redlich, F. Karsch, and A. Tounsi, hep-ph/0302245. 109. J. Cleymans and P. Koch, Z. Phys. C52 (1991) 137. 110. D. Miller and K. Redlich, Phys. Rev. D37 (1988) 3716; Phys. Rev. D35 (1987) 2524; H. Th. Elze, D. Miller and K. Redlich, Phys. Rev. D35 (1987) 748. 111. H. Th. Elze, W. Greiner and J. Rafleski, Phys. Lett. B124 (1983) 515; Z. Phys. C24 (1984) 361; Ch. Derreth, W. Greiner, H. Th. Elze and J. Rafelski, Phys. Rev. C31 (1985) 1360; H. Th. Elze and W. Greiner, Phys. Rev. A33 (1986) 1879. 112. D.H. Rischke, M.I. Gorenstein, A. Schafer, H. Stocker and W. Greiner, Phys. Lett. B278 (1992) 19; D.H. Rischke, J. Schaffner, M.I. Gorenstein, A. Schaefer, H. Stocker and W. Greiner, Z. Phys. C56 (1992) 325. 113. A. Gocksch, R.D. Pisarski, Nucl. Phys. B402 (1993) 657. 114. J. Engels, et al., Z. Phys. C42 (1989) 341; A. Peshier, et a l , Phys. Rev. C61 (2000) 045203 and references therein. 115. R. Rapp and J. Wambach, Adv. Nucl. Phys. 25 (2000) 1. 116. J. Cleymans, K. Redlich and E. Suhonen, Z. Phys. C58 (1993) 347. 117. K. Zalewski, Acta. Phys. Polon. B62 (1965) 933. 118. A. Wagner et al., KaoS Coll., Phys. Lett. B420 (1998) 20, C. Muntz et al., KaoS Coll., Z. Phys. C357 39. 119. R. Barth et al., KaoS Coll., Phys. Rev. Lett. 78 (1997) 4007; F. Laue et al., KaoS Coll., Phys. Rev. Lett. 82 (1999) 1640. 120. M. Menzel, Ph.D.Thesis, Universitat Marburg, 2000. 121. L. Ahle et al., E-802 Coll., Phys. Rev. C58 (1998) 3523; 122. R. Averbeck, R. Holzmann, V. Metag and R.S. Simon, nucl-ex/0012007; A.R. Wolf et al.,Phys. Rev. Lett. 80 (1998) 5281; By TAPS Coll., Z. Phys. A359 (1997) 65; Phys. Rev. C56 (1997) 2920. 123. B. Hong et al., FOPI Coll., Phys. Rev. C57 1998 244. 124. C. Miintz et al., KaoS Coll., Z. Phys. C357 1997 399; J. Phys. G28 (2002) 1895; Euro. Phys. J. A9 (2000) 397. 125. M. Bleicher et a l , J. Phys. G25 (1999) 1859; S. A. Bass et al., Prog. Part. Nucl. Phys. 41 (1998) 225. 126. H.J. Drescher et al., Phys. Rep. 350 (2001) 93. 127. H.-U. Bengtsson and T. Sjostrand, Comput. Phys. Commun. 46 (1987) 43.
596
P. Braun-Munzinger,
K. Redlich and J. Stachel
128. C. Greiner, hep-th/0209021; J. Phys. G28 (2002) 1631. 129. C. Hartnack, H. Oeschler, J. Aichelin, Phys. Rev. Lett, (in print), nuclth/0109016. 130. S. Pal, C M . Ko, and Z. Lin, Phys. Rev. C64 (2001) 042201. 131. G.E. Brown, M. Rho, and Ch. Song, Nucl. Phys. A698 (2002) 483; Nucl. Phys. A690 (2001) 184. 132. W. Cassing and M. Lutz, privat communication. 133. F. Becattini, Z. Phys. C69 (1996) 485; F. Becattini and U. Heinz, Z. Phys. C76 (1997) 269. 134. M. Gazdzicki, and M. Gorenstein, Phys. Lett. B483 (2000) 60. 135. A. Casher, H. Neuberger and S. Nussinov, Phys. Rev. D20 (1979) 179. 136. M. Bleicher, et a l , Phys. Rev. Lett. 88 (2002) 202501; S.A. Bass, et al., nucl-th/0204049 137. J. S. Schwinger, Phys. Rev. 82 (1951) 664. 138. J. Bachler et al., NA49 Coll., Nucl. Phys. A661 (1999) 45; R.A. Barton et al., NA49 Coll., J. Phys. G27 (2001) 367; V. Afanasev et a l , NA49 Coll., Phys. Lett. B491 (2000) 59. 139. A. Capella, C. A. Salgado, and D. Sousa, nucl-th/0205014. 140. A. Capella and C. A. Salgado, Phys. Rev. C60 (1999) 054906. 141. J. Cleymans and K. Redlich, in preparation. 142. P. Braun-Munzinger, and K. Redlich, Eur. Phys. J. C16 (2000) 519; Nucl. Phys. A661 (1999) 546. 143. M. Gorenstein et al., Phys. Lett. B509 (2001) 277; K.A. Bugaev, M. Gazdzicki and M. Gorenstein, Phys. Lett. B544 (2002) 127; Phys. Rev. Lett. 88 (2000) 132301; Phys. Lett. B544 (2002) 127. 144. L. Grandchamp and R. Rapp, Phys. Lett. B523 (2001) 60; Nucl.Phys. A709 (2002) 415, hep-ph/0205305. 145. M. Gorenstein, and M. Gazdzicki, Phys. Rev. Lett. 83 (1999) 4003. 146. R.L. Thews, M. Schroedter and J. Rafelski, Phys. Rev. C63 (2001) 054905. 147. R. Vogt (p. 250c) in S.A. Bass et al., Nucl. Phys. A661 (1999) 205c. 148. A. Andronic, P. Braun-Munzinger, K. Redlich, J. Stachel, QM2002 contribution, Nucl. Phys. A (in print), nucl-th/0209035. 149. A. Andronic, P. Braun-Munzinger, K. Redlich, J. Stachel, nucl-th/0303036. 150. J.-P. Blaizot, P.M. Dinh, J.-Y. Ollitrault, Phys. Rev. Lett. 85 2000 4012 [nucl-th/0007020]. 151. H. Sorge, E. Shuryak, Phys. Rev. Lett. 79 (1997) 2775. 152. M.C. Abreu et al., NA50 Coll., Phys. Lett. B438 (1998) 35, and references therein. 153. M.C. Abreu et al., NA50 Coll., Nucl. Phys. A698 (2002) 539c. B438 (1998) 35, and references therein. 154. C. Gerschel, Acta Phys. Pol. B30 (1999) 3585. 155. M.C. Abreu et al., NA38 Coll., Phys. Lett. B449 (1999) 128. 156. M. Gonin et al., NA50 Coll., Proc. 3rd Conf. on Physics and Astrophysics of Quark-Gluon Plasma, Jaipur, India, March 1999, B. C. Sinha, D.K Sri-
Particle Production
157. 158. 159. 160. 161. 162. 163. 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. 175. 176.
177. 178. 179.
180. 181. 182. 183. 184. 185. 186.
in Heavy Ion Collisions
597
vastava, Y.P. Viyogi, editors, Narosa Publ. House 1998, p. 393. L. Ramello et al., NA50 Coll., Nucl. Phys. A638 (1998) 261c. M.C. Abreu et al., NA38 Coll., Phys. Lett. B466 (1999) 408. NA50 Coll., Phys. Lett. B530 (2002) 33. R. Vogt, hep-ph/0203151; hep-ph/0111271. NA50 Coll., Phys. Lett. B450 (1999) 456; Phys. Lett. B477 (2000) 28. J. Gosset et al., Eur. Phys. J. C13 (2000) 63. NA50 Coll., Nucl. Phys. A698 (2002) 539c. R. Rapp and E. Shuryak, Phys. Lett. B473 (2000) 13. J. Nagle, PHENIX Coll., Nucl. Phys. A698 (2002) 599. R. Averbeck, PHENIX Coll., Nucl. Phys. A698 (2002) 39. C M . Ko, X.N. Wang, B. Zhang and X.F. Zhang, Phys. Lett. B178 (1998) 237. W. Schmitz et al., CERES Coll., J. Phys. G28 (2002) 1861. M. Bleicher and J. Aichelin, Phys. Let. B530 (2002) 81. see also: L. Bravina, et al., Nucl. Phys. A698 (2002) 383; Phys. Rev. C66 (2002) 014906 M. Gyulassy, privat communication. V.D. Toneev, J. Cleymans, E.G. Nikonov, K. Redlich, and A.A. Shanenko, J. Phys. G27 (2001) 827; nucl-th/9904048. E.G. Nikonov, A.A. Shanenko and V.D. Toneev, Heavy Ion Physics 8 (1998) 89. C M . Hung and E.V. Shuryak, Phys. Rev. Lett. 75 (1995) 4003; Phys. Rev. C57 (1998) 1891. P. Braun-Munzinger and J. Stachel, J. Phys. G28 (2002) 1971. K.A. Olive, Nucl. Phys. B190 [FS3] (1981) 483; K.A. Olive, Nucl. Phys. B198 (1982) 461; J.I. Kapusta and K.A. Olive, Nucl. Phys. A408 (1983) 478; A.L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems, McGraw-Hill, 1971. S. Ejiri, et al., hep-lat/0209012; C. Schmidt, et al., hep-lat/0209009. Z. Fodor and S.D. Katz, Phys. Lett. B534 (2002) 87; JHEP 0203 (2002), hep-lat/0106002. S. Fortunato, F. Karsch, P. Petreczky, and H. Satz, Phys. Lett. B502 (2001) 321; D.S. Digal, S. Fortunato, P. Petreczky, and H. Satz, Phys. Lett. B541 (2002) 101. See e.g. P. Senger, J. Phys. G28 (2002) 1869, and GSI Report, Nov. 2001. M. Gazdzicki, and D. Roehrich, Z. Phys. C66 (1995) 77. The 47r results at RHIC are obtained from Ref. (183) and Ref. (184) using the extrapolation procedure described in Ref. (185). I.G. Bearden, et al., nucl-ex/0108016. B.B. Back, et al., PHOBOS Coll., Phys. Rev. Lett. 85, (2000) 3100; Phys. Rev. Lett. 87, (2001) 102303; Phys. Rev. Lett. 88, (2002) 022302. By Na49 Coll., nucl-ex/0205002. K. Wroblewski, Acta. Phys. Polon. B16 (1985) 379.
598 P. Braun-Munzinger, K. Redlich and J. Stachel 187. By STAR Coll., J. Phys. G28 (2002) 1535; Nucl. Phys. A698 (2002) 64. 188. V. Afanasev et al., NA49 Coll., nucl-ex/0208014; 0209002. 189. K. Redlich and A. Tounsi, hep-ph/0105201; hep-ph/0111159; A. Tounsi, A. Mischke, and K. Redlich, QM2002 contribution, Nucl. Phys. A (in print), hep-ph/0209284. 190. Ch. Blume et al., NA49 Coll., Nucl. Phys. A698 (2002) 104. 191. J.C. Dunlop, et al., Phys. Rev. C61 (2000) 031901. 192. W. Cassing, Nucl. Phys. A661 (1999) 468c. 193. F. Wang, H. Liu, H. Sorge, N. Xu and J. Yang, Phys. Rev. C61 (2000) 064904. 194. H. Weber, E.L. Bratkovskaya and H. Stocker, nucl-th/0205030. 195. M. Gazdzicki and M.I. Gorenstein, Acta. Phys. Polon. B30 (2000) 965; M. Gazdzicki and M. Gorenstein, Phys. Lett. B483 (2000) 60. 196. S. E. Eiseman et al., E810 Coll., Phys. Lett. B297 (1992) 44 and B325 (1994) 322. 197. T. Abbott et al., E802 Coll., Phys. Rev. C60 (1999) 044904; Y. Akiba et al., E802 Coll., Nucl. Phys. A590 (1995) 179c. 198. H. Oeschler, et al., in preparation. 199. F. Laue, C. Sturm et al., Phys. Rev. Lett. 82 (1999) 1640. 200. B. Holzman, E917 Coll., Nucl. Phys. A698 (2002) 643. 201. By NA49 Coll., Phys. Lett. B491 (2000) 59; H. Biaikowska., and W. Retyk, J. Phys. G27 (2001) 397. 202. By STAR Coll., Phys. Rev. C65 (2002) 041901; nucl-ex/0206008. 203. R. Kotte, FOPI Coll., Proceedings of Hirschegg XXVIII International Conference, Hadrons in Dense Matter (2000). 204. M.C. Abbreu et al., NA50 Coll., Nucl. Phys. A661 (1999) 534c. 205. S. Pal, C M . Ko, and Z. Lin, Nucl. Phys. A707 (2002) 525. 206. N. Herrmann, FOPI Coll., Nucl. Phys. A610 (1996) 49. 207. by FOPI Coll., ncl-ex/0209012. 208. N. Xu and K. Redlich, in preparation. 209. J. Cleymans, B. Kampfer, and S. Wheaton, QM2002 contribution, Nucl. Phys. A (in print), hep-ph/0208247; Contributed to 30th International Workshop on Gross Properties of Nuclei and Nuclear Excitation: Hirschegg 2002: Ultrarelativistic Heavy Ion Collisions, Hirschegg, Germany, 13-19 Jan 2002. hep-ph/0202134; S. Yacoob and J. Cleymans, hep-ph/0208246. 210. J. Cleymans, J. Phys. G28 (2002) 1575. 211. R. Hwa, and C.B. Yang, Phys. Rev. C66 (2002) 064903. 212. E. Andersen, et al., WA97 Coll, Phys. Lett. B449 (1999) 401. 213. N. Carrer, NA57 Coll, Nucl. Phys. A698 (2002) 29. 214. S. Kabana, et a l , NA52 Coll, J. Phys. G27 (2001) 495. 215. S. Soff, et a l , J. Phys. G27 (2001) 449. 216. L. Bravina, J. Phys. G27 (2001) 449. 217. S.E. Vance, et a l , Phys. Rev. Lett. 83 (1999) 1735; J. Phys. G27 (2001) 627.
Particle Production
in Heavy Ion Collisions
599
218. S. Pal, C M . Ko, J.M. Alexander, P. Chung, and R.A. Lacey, nuclth/0211020. 219. M. Bleicher, W. Greiner, H. Stocker and N. Xu, Phys. Rev. C62 (2000) 061901. 220. Z. Lin, et al., Phys. Rev. C64 (2001) 011902. 221. V. Koch, nucl-th/0210070.
HYDRODYNAMICAL DESCRIPTION OF COLLECTIVE FLOW
Pasi Huovinen School of Physics and Astronomy, University of Minnesota, Minneapolis, MN 55455, USA
I review how hydrodynamical flow is related to the observed flow in ultrarelativistic heavy ion collisions and how initial conditions, equation of state and freeze-out temperature affect flow in hydrodynamical models.
Contents 1 Introduction 2 Hydrodynamical Model 2.1 Basics 2.2 Initialization 2.3 Equation of state 2.4 Freeze-out 3 Transverse Flow and Its Anisotropies 3.1 Transverse 3.2 Flow anisotropies 3.2.1 Directed 3.2.2 Elliptic 4 Learning from RHIC Data 5 Summary and Outlook References
600
flow flow flow
601 602 603 604 608 608 610 610 615 616 617 623 629 631
Hydrodynamical Description of Collective Flow 601 1. Introduction One of the goals of the experimental heavy ion program at ultrarelativistic energies is to create a dense system of strongly interacting particles. It is hoped that the particles formed in the primary collisions would rescatter often enough to reach local thermal equilibrium and behave as a particle fluid, not as a cloud of free particles. If such a state is reached, the finally observed particles should depict signs of collective behaviour such as flow. Our intuitive understanding of flow, i.e. collective motion, is closely tied to a classical macroscopic description of flow using the language and tools of hydrodynamics. This means that it is often easiest to use hydrodynamical concepts like temperature, pressure and flow velocity to describe collective motion even if the applicability of such concepts is far from certain. Hydrodynamical models are thus particularly suitable to describe flow phenomena, but we have to be careful not to confuse what is actually observed with our way of describing observations. An example of the limits of hydrodynamical language is that there is no generally accepted definition of flow in the context of heavy-ion collisions, but the word is used in its intuitive meaning. A rigorous definition of flow is beyond the scope of this review. Instead I use a hydrodynamically practical definition of flow: collective flow is correlation of position and momentum during the dense, interacting stage of the collision regardless of the origin of these correlations. This means that I also call flow the correlation between the longitudinal momentum and the position of particles which has its origin in the initial particle producing processes. In a hydrodynamical model these correlations are manifested as an initial non-zero longitudinal velocity field. Unlike transverse and longitudinal flow, directed and elliptic flow do not directly refer to collective motion but to certain emission pattern where particle emission is not azimuthally isotropic (for definitions see sections 3.2.1 and 3.2.2). In principle elliptic anisotropy could be entirely due to the shape of the surface of the source and be finite even if flow the velocity is zero . Therefore to call elliptic anisotropy elliptic flow is unfortunate but firmly established in the literature. It has to be remembered, however, that even if elliptic anisotropy is not necessarily a sign of collective motion, it is a collective effect. Hydrodynamics connects the conservation laws to the equation of state, viscosity and heat conductivity of the fluid. Thus the properties of matter and flow are intimately connected and we hope to learn about the equation of state of nuclear matter by studying the flow in heavy ion collisions. In practice, however, this is a challenging task because of the nonlinear nature of the equations of hydrodynamics and the many unknowns in the hydrodynamical description of heavy-ion collision. In this review I describe briefly the basic concepts of a hydrodynamical model
602 P. Huovinen and the kind of collective flow generated in hydrodynamical simulation. My emphasis is on details which have particular significance in the description of elliptic flow and how initial shape of the system, equation of state and freeze-out temperature affect elliptic flow in Au + Au collisions at the Relativistic Heavy Ion Collider (RHIC) at Brookhaven National Laboratory (BNL) at
2 - 0± (zi)\
(69)
is an unambiguous sign of dynamically generated ^-dependent correlations in the source. 3. The second moments of the HBT radius parameters: The only non-vanishing second moments are
Ka2
= \ ((?)o ~ <*2>o) + \ mi
+ (i2)c2)
- \ (<*2>: - <^2>:) - \ <m\. Ka2
(70)
= - \ ((y2)o - (*2)o) + \ ((y2)c2 + (*2)c2)
~P±{(xt)l-(yt)t)+l3l(Pr2 - p± ((ii)l + (0)1) + \ ((*% - (?):) + \ (*y)l, (7i)
- % ««">s+mi) - \ 0 2 >: - ax)+\ <*»>:> (72) K22
= &)2 •
(73)
The equations (57) remain true in the presence of ^-dependent positionmomentum gradients in the source, while deviations from the Rcoa = -Rc22 = -Rsos 2 2 rule (58) are an unambiguous sign for an implicit $-
Central and Non-Central
HBT from AGS to RHIC 735
dependence, namely: Ro,22 + K22 = (i2)C2 +
(y2)2+Pl(P)2
-(3± {(xi)l - (yi){ + (xi)c3 + (yF)l) ,
(74)
K22 + Ksi = \ «*2>§ + if)!) + #>§ - % mi - (m + (£y)l
Ka - Ksi = \ «*2>2 + (y2)c2) - \p± mci - (yt)i - {^l - ml)
- \ 02>: - mi) - m: •
m
If the fourth order moments (x2). — (& 2 ) 4 and {xy)\ are negligible and if the transverse pair momentum K± is sufficiently small, then nc 2 r>s 2 ^ n c 2 . r>s 2 s,2 ~ nos,2 ~ no,2 "+" -""08,2 •
n
/'77^ V ' ' i
4. The n-th moments of the HBT radius parameters (n > 2): In general, the properties of the sin and cos functions imply that the n-th order harmonics -Rfj>n2, #ij,„ 2 in eqs. (49), (50) are built up of m-th order harmonics (x^y^)^, {x^y^)^ with n — 2 < m < n + 2. This limits the number of terms appearing in n-th order expressions. Since it is an open question whether the approximation (63) applies for realistic sources'5, we give these expressions here for completeness. For the side-long and out-long HBT radius parameters, only odd harmonic terms appear. For n = 3, 5, 7, . . . , we have K.n2
=
\{(*%en-l-m'n-l)-0±(&>en +\m
n +
^ / , „ 2 = \ m'n-1
l + (yS)n+l), - < ^ X - l ) + \ m'n+1 + <**>n+l) •
(78) (79)
The Fourier decomposition of the other HBT radius parameters contains b
We thank Mike Lisa for drawing our attention to this issue.
736
B. Tomdsik and U. A.
Wiedemann
only even harmonic terms. For n = 4, 6, 8, . . . , we have
+l{(*2)Xy%)
- \ (<*2>:+2 - <^2>:+2) - \ ay)n+2,
Kn2 = \ (<*2>:_2 - mu)
(so)
- \ (*y)n-2
+\((*2)>(y2):)+&(?): + \ (<*2>:+2 - (y%+2) + \ «
+\^
(<«">:.! -
- ^
(«+1+<^:+1)
- \ ((*%+2 - (y%+2) *?.»"
•
+ 2
(si)
(^:-0 + \ «
+ 2
= <^>: •
.
(82) (83)
Whether these higher harmonic coefficients are numerically important remains to be established experimentally. 2.3.6. Implicit azimuthal dependence at forward rapidity At non-central rapidity 7 ^ 0 , additional Fourier components of the spacetime variances can contribute to HBT radius parameters, as is seen from Table 1. The reason is that at Y ^ 0, the symmetries (46)-(48) do not hold. It is remarkable that the zeroth and second moments of the out, side and out-side radius parameters, given in eqs. (64), (65), (70), (71) and (72), do not receive extra contributions at forward rapidity. In particular, this implies that deviations from the R%^ = —RcSt22 = #L,2 2 r u ' e °f (58) remains an unambiguous test for the presence of angular dependent
Central and Non-Central
HBT from AGS to RHIC
737
position-momentum correlations at all rapidities. In addition, the out, side and out-side radius parameters acquire first harmonic oscillations away from mid-rapidity,
RcsA2 = \{{3?)l-2{xy)[+Z{?)l) -\((x%
+
2(xyy3-(?)l),
(84)
RcoA2 = \(Hx2)ci + ^mi + (y2)i) - fix (2(xi)0 + (xi)c2 + {yt)l) + (31^)1
+ \((i2)l + 2(xyr3-(f)l), Ks,2
= \{-{i2)l-^v)l
(85)
+ {y2)\)
+P±(xi)0 - \[3±. ((xi)c2 + (yi)c2)
+ 1-((x% + 2(xy)l-(f)l).
(86)
In the so-called blast wave model 153 ), there is no correlation between the transverse position and the time at which particles are emitted. Hence, the space-time variances linear in i vanish. Also, the emission duration (P) does not depend on the azimuthal angle. In addition, if the source shows a sufficiently smooth azimuthal dependence for the third order terms ( i 2 ) 3 , (xy)3, (y2)3 to be negligible, then R°s,l
- R°o,l
~
2
-^os,l
•
(87)
Moreover, it was observed in a class of model studies 98 that (J 2 )f should be much larger than (xy)l and (y2)f, since asymmetries with respect to the beam axis will occur predominantly in the direction of the impact parameter. This translates into the rule Ro,i
:
-^s.i
:
R0s,i
= 3 : 1: -1.
(88)
A test of (87), (88) allows to establish whether these additional modeldependent assumptions are satisfied. The three HBT radius parameters Rf, W"ol and R?sl involve longitudinal information and depend on the longitudinal velocity 0i. This leads to additional contributions away from mid-rapidity. For completeness, we list here
738
B. Tomdsik and U. A.
Wiedemann
the non-vanishing first and second moments: Rf,i2 = (~z2)i - 2A<**>? + 0f(P)i , 2
c
C
Rf/ = (z )l ~ Wi(zi) 2 + tf(P) 2 ,
(89)
(90)
Ki,i2 = (™)o + \ ((iz)c2 + (yz)'2) - /3±(zt)f + |/3« (2/?±
; - 2(£t)o - (2i}§ - <jft)5) . Rcoi,22 =
(91)
\((^)i-msi)-p±&)c2
+±pl(2(3x(Py2-(xi)c1 + (yi)l) + i «**>' + <^~>*) - | {{xt)l + (yi)l) , RSsi,i2 = -(^)o
+ \((xz)c2
(92)
+ (yzy2)
+ 1/?, (2(xt)0 - {xi)c2 - (yt)s2) ,
(93)
Kia = \ mi - @m) + \PI mci - (m + \ {{xz% + (yz)s3) - |
((xF)l + (yt)s3) .
(94)
2.3.7. Reconstruction of the reaction plane The above analysis of HBT radius parameters for non-central collisions requires the measurement of the angle $ and thus assumes knowledge about the event-wise orientation of the reaction plane. This orientation ^R is usually measured from the azimuthal dependence of single-particle transverse momentum spectra 63 ' 65
d3N
dN E
1T
=
A
dp
A
f
A AA. = /
4
pt dpt dy d
=
»2 AT
°°
1+2
2^^^[
g^ C ° S n ( ^ V ' f l ) ] -
(%)
However, the orientation of the true reaction plane can only be measured with limited accuracy. Since fluctuations in a finite multiplicity environment result in azimuthal anisotropics without geometrical origin, this limited accuracy arises largely as a consequence of the basic statistical properties
Central and Non-Central
HBT from AGS to RHIC
739
of a mesoscopic system, and cannot be reduced by larger event samples or refined measurements. The resulting uncertainty has to be corrected for if one aims for a geometrical interpretation of the ^-dependence of HBT radius parameters. Such corrections are discussed in literature 98,39 . 3. Two-particle correlations from model calculations The emission function S(x, K) is not determined uniquely by the correlator C(q, K). This is a consequence of the on-shell constraint (6) which implies that only a specific time-average over the emission function, the so-called relative distance distribution SK(X), is uniquely measurable 100
SK(x)=
J dtd(x + (3t,t;K),
d(x K)=(d*X { }
'
J
5 ( X +
t'*)
(96) S(X-hK)
fd*yS(y,K)fd*yS(y,K)-
(97)
The direct reconstruction of S/f (|aj|) from experimental data has been pursued successfully 15,16,115 . However, due to statistical uncertainties the numerical inversion of C(q, K) — \ = J d3x cos(q-x) SK(X) is complicated. In practice, it requires additional model assumptions to achieve convergence. Thus most data analyses proceed via model studies. Either they start from a model parametrisation of the emission function S(x, K), or they start from a dynamical calculation of S(x, K) based on a hydrodynamic or particlecascade based simulation. In this section, we review the main features of these approaches. 3.1. Model parametrisations
of the emission
function
The main features of the collision region at freeze-out can be characterised by its width in the different spatial and temporal extensions, its collective dynamical gradients (usually ascribed to a collective flow field u^{x) which determines the position-momentum correlations in the source) and its random dynamical component (usually ascribed to a local temperature T). Model parametrisations of S(x,K) implement these main features in a (minimal) analytical ansatz for S(x,K). The model parameters are then extracted from a fit to one- and two-particle spectra. Example of a model emission function: For illustration, consider a source in local thermal equilibrium at temperature T whose extension is given by the
740
B. Tomdsik and U. A.
Wiedemann
transverse width R, space-time rapidity width Arj and proper longitudinal emission time To smeared with a width A T . The transverse and longitudinal expansion of the collision system results in a longitudinally boost-invariant flow profile at freeze-out with a transverse component rjt(r) = Vfjs characterised by the transverse gradient rjf, Ufj,(x) = (cosh77 coshrft, fsiahrjt, ^sinhry^sinhrj cosh774) . 2
(98)
2
In longitudinal proper time r = y/t — z and rapidity I In [(t + z)/(t — z)], this source can be written as 2 , 1 9 ' 2 4 , 9 6 : Sr(x,p) V
= ;
T
—^=
(2 7 r) 3 v / 2^Ar 2 r
x exp
2R2
rn±_ cosh(y — ri) exp V " 2 r} (r - To)2 2(Ar?) 2
2(AT)2
p • u{x) — fir
'
(99)
Here, r labels the particle species which are produced in thermal abundances with chemical potentials \xr. The model emission function (99) is completely specified by the model parameters T,r)f ,R,Ar],AT,TO ,firThis basic model allows for a satisfactory fit to experimental data from the CERN SPS 89 . Overview of models and model extensions: In what follows, we review the physics arguments which motivated the study of modifications and extensions of the parameterisation (99): (1) Varying transverse density and flow profiles: The functional shape of (99) was varied by replacing the Gaussian transverse density distribution with a box profile89 or varying the functional dependence of the transverse flow profile95. This gives further support to the general statement that HBT radius parameters are mainly sensitive to the average r.m.s. width of S(x, K). However, details in the functional shape of S(x, K) can leave observable traces in the K± dependence of the HBT radii: in particular, experimental data from the SPS favour a transverse box profile over a Gaussian one. (2) Surface dominated versus bulk dominated emission: Model (99) implements bulk emission, i.e., particles decouple at the same average proper freeze-out time To from all spatial positions in the source with a probability proportional to the source density. However, if reabsorption of particles by the surrounding matter is significant, hadronic freeze-out may proceed via surface evaporation. In analytical
Central and Non-Central
HBT from AGS to RHIC
741
parametrisations of S(x, K), such "opaque sources" have been modelled via addition of absorption factors 40 . The main outcome of these studies is that surface-dominated emission can imply (x2)
flg(*x) » 1
L2
"T T
• If
(100)
742
B. Tomdsik and U. A.
Wiedemann
The size of this radius is proportional to the source size, but it is also sensitive to the transverse flow strength r]f of the source. This illustrates that HBT radii characterise only that part of a dynamically expanding source which can be viewed through a filter of wavelength K. This shrinking effect increases for increasing M± proportional to the ratio rft/T. The Makhlin-Sinyukov formula58 for the longitudinal radius,
*?""*£.
(WD
shows compared to (100) a stronger Mj_-dependence consistent with the stronger longitudinal expansion implemented in (99). Its dependence on ro is a direct consequence of the assumed longitudinal boost-invariance and receives corrections for sources of finite longitudinal extension. While quantitative corrections to these analytical expressions can be significant95, these pocket formulas illustrate qualitatively the interplay of geometry and dynamics in determining HBT radius parameters. This picture is supported by numerous numerical studies. (2) The difference R20-R2S: The main interest in this observable69 lies in its sensitivity to the emission duration [see also discussion of eq. (120)] Rl-Rl^Hl
).
(102)
Numerical calculations with a Gaussian density profile typically result in a small but positive signal for R^ — R?s. For steeper transverse profiles and particle emission at sufficiently large K±, or for opaque source 40,59 models with surface dominated emission, also negative values can be found for R% — R\. Equation (102) ignores the contribution from the (xt) correlation term which vanishes in the model (99) but is present in hydrodynamic models and Monte Carlo event generators, see below. (3) Influence of resonance decay contributions: Pions from resonance decays have a tendency to be emitted at later times and larger distances 28 ' 23 ' 97 . For models showing bulk emission, their effect on the size of HBT radius parameters is however small 97 . This is due to a combination of three effects: i) in models of the type (99), the emission region of the heavier resonances is smaller than that of direct pions, ii) the large decay widths of the most abundant resonances like p's and A's and their non-relativistic velocities imply that these decays occur within the emission region of the direct pions, iii)
Central and Non-Central
HBT from AGS to RHIC
743
resonances with large lifetime (rj, rj') decay so far outside that their decay pions interfere with the directly produced ones on a very small relative momentum scale (|q| < 1 MeV) only. This produces a peak of the correlation function at |q| < 1 MeV which is narrower than the experimental resolution and thus leads to an apparently reduced intercept A of the correlator C(q, K) without affecting its shape. Only pions from to decays stem from a resonance which is neither sufficiently short-lived nor sufficiently long-lived and thus can affect the shape of the correlator. This spoils a naive core-halo interpretation and contributes to non-Gaussian deviations of the two-particle correlator 96 . 3.2. Hydrodynamic
models
Hydrodynamic behaviour is an idealised but well-defined limiting case of the realistic dynamical evolution of the collision region in heavy ion collisions. It emerges as the zero mean free path limit of a particle cascade. In this limit, matter in the collision region is treated as an ideal, locally thermalised fluid whose dynamics is governed by the relativistic hydrodynamic equations. Input for simulations: A hydrodynamic model is fully specified by the equation of state and the initial conditions. Typically, the parametrisation of the latter models the outcome of an initial pre-equilibrium stage with initial energy density estimated from the Glauber approach to entropy and energy production in nucleon-nucleon collisions49. Freeze-out criterion: The freeze-out criterion, according to which the hydrodynamic simulation is terminated, is another important input in hydrodynamic model studies. Usually, the freeze-out criterion is set by a critical energy density or temperature. If the criterion is satisfied in a fluid cell, the cell is immediately assumed to freeze-out. Local properties of this cell are converted into a thermal ideal gas distribution of hadronic resonances with temperature and chemical potential set by the local energy and baryon density of the simulation. This leads to a sharp freeze-out along a threedimensional hypersurface and specifies the emission function S(x, K) entering the calculation of two-particle correlation functions. There are "hybrid models" in which the earlier hot stage of the collision is treated hydrodynamically but the hadronic phase is modelled with a Monte Carlo event generator code 80 . An event generator naturally leads to an emission function in a finite four-volume, see next subsection. Successes and problems at RHIC: At RHIC, hydrodynamic simulations
744
B. Tomdsik and U. A.
Wiedemann
Table 2. The main hydrodynamic model calculations with published results on HBT correlation functions. Codes follow either the full three-dimensional expansion or the two-dimensional expansion in the transverse plane (with assumed boost-invariance in the remaining longitudinal direction). Authors r e '-
74 13
HYLANDER '
energies
HBT data
studied
compared to
SPS
Dim.
NA44 7 6
(3+l)-dim
Rischke, Gyulassy 7 3
SPS, RHIC
RHIC prediction
(2+l)-dim
Zschiesche et a/. 105
SPS, RHIC
NA49, STAR
(2+l)-dim
Kolb, Heinz
38
Hirano, Morita et al.
RHIC SPS, RHIC
STAR, PHENIX 61 62
38
44 62
NA49 ' , STAR '
(2+l)-dim (3+l)-dim
compare in general well with the hadronic one-particle transverse momentum spectra up to « 2 GeV. The major success of this approach is the prediction of the size of the measured elliptic flow V2, as well as the correct description of its px-dependence for identified pion and proton spectra. This indicates that the main contribution to elliptic flow originates in the early stages of the collision where the system is very dense and the mean free path is close to the hydrodynamic limit zero. However, two-particle correlations are determined at freeze-out, where the mean free path (or rather the mean scattering time) is grown and a hydrodynamic picture becomes questionable. This may be one of the reasons why so far hydrodynamic simulations have significant problems in calculating two-particle correlators which are at least in qualitative agreement with experimental data 3 8 , see the following discussion. Generic properties of hydrodynamic simulations for HBT: An overview of hydrodynamic model calculations is given in Table 2. This list is limited to studies which include beyond the calculation of one-particle spectra also two-particle correlation functions. (1) Freeze-out hyper-surface shows strong outward-temporal correlations: The freeze-out criterion implemented in hydrodynamic simulations amounts to a sudden switch from a zero mean free path to an infinite mean free path approximation. This tends to favour sharp geometrical correlations along the freeze-out hyper-surface. In comparison to model
Central and Non-Central
HBT from AGS to RHIC
745
sources with emission from a finite four-volume, the (it) variance is significantly stronger 72 . Thus, the difference R2 — R2 does not measure a lifetime effect only. This consequence of a sharply localised freeze-out hyper-surface may be tamed in hybrid models in which hydrodynamic evolution is followed by a hadronic rescattering phase 80 . (2) Large "lifetime" effect and R20 » R2: Hydrodynamic simulations lead to sources with very large emission durations. The size of this lifetime signal depends on the equation of state (EOS): a softer EOS results in a more delayed pressure build-up and a longer lifetime 72,73 . Irrespective of model details, the resulting values for R0/Rs are generically much larger than the measured result 38 . This is the main problem of hydrodynamic simulations. (3) Resonance decays contribute significantly to HBT radii: In contrast to model sources of the type (99), resonance decay contributions added to the freeze-out of hydrodynamic simulations were reported to increase the size of HBT radii significantly13. This may be attributed to the different shapes of the freeze-out hyper-surfaces. The homogeneity regions for direct pion and resonance emission are the same in hydrodynamic simulations whereas the latter are smaller in the model (99). Thus even short-lived resonance decay contributions tend to increase the hydrodynamic pion source.
3.3. Monte
Carlo event
generators
Event generators are widely used to simulate particle production in ultrarelativistic heavy ion collisions. In particular, they allow to study how global collective dynamical properties emerge in a mesoscopic system from microscopic (2-to-2 or 2-to-3 body) interactions. In principle, each event generator output defines an emission function from which two-particle correlations can be calculated. However, since the event generator output is not a wavefunction with proper quantum-mechanical symmetrisation, an additional prescription is needed of how to relate it to the emission function. There is an extensive literature on the conceptual problem 100 . In practice, the afterburner program of Scott Pratt 7 0 is most frequently used. So far, there are only very few calculations of HBT correlation functions from event generators, see Table 3. The main conclusion from these calculations is that the late hadronic rescattering phase largely determines the size
746
B. Tomdsik and U. A.
Wiedemann
Table 3. Event generator model calculations with published results on HBT correlation functions. Code r e f -
RQMD83'84'0 hydro73+URQMD6'6 Humanic 4 5 , c AMTP
56,d
MPC60
energies
HBT data
studied
compared to
SPS
NA35 8 6
SPS, R H I C 8 0 ' 8 1
STAR 8 2
AGS, SPS, RHIC
E859/866, NA44 4 6 STAR 4 7
RHIC
STAR 5 6
RHIC
STAR, PHENIX 6 0
a
Code not maintained any more, no more recent studies available. Hadronic rescattering phase dominates HBT radii 8 2 . c This code models final state rescattering only. d This is a multi-phase transport model which includes initial partonic and final state hadronic interactions. b
of HBT radius parameters. Also, in contrast to hydrodynamic models, the generation of models with relatively small lifetime, satisfying R0/Rs ~ 1, does not appear to be a fundamental problem. While some simulations find a ratio R0/Rs which for large K± lies between 1.4 and 2.0, inconsistent with experimental data 8 2 , other simulations 47,56 are consistent with R0/Rs ~ 14. H B T m e a s u r e m e n t s Two-particle correlations have been measured at all energies from AGS to RHIC. Here we give an overview of the experimental situation. 4 . 1 . Coulomb final state
corrections
Data on two-particle momentum correlations between identical charged pions are usually corrected by the experimentalists for the pairwise Coulomb repulsion. These corrections are difficult since Coulomb interaction and Bose-Einstein interference effects are of similar size and affect the twoparticle correlator on similar relative momentum scales. Moreover, the used correction methods differ between experiments and sometimes even between different publications in one experiment. Differences between the used correction techniques can change the resulting size of the HBT radius parameter by more than 1 fm and they may affect the Xj_-slope of HBT radii.
Central and Non-Central
HBT from AGS to RHIC
747
Thus Coulomb final state corrections are a major source of systematic uncertainty in the space-time analysis of correlation measurements. The final state Coulomb interaction between two charged particles is described by the relative Coulomb wave-function of the particle pair, written in terms of the confluent hypergeometric function F 53 - 67 ! ^ ( r )
= r ( l + t»7)e-i"'ei«-'-F(-iT7;l;z_) ,
z± = \{qr±q-r)
= \qr(l±cos6).
(103) (104)
Here, r = \r\, q = \q\, and 9 denotes the angle between these vectors. The Sommerfeld parameter rj = a/(viei/c) depends on the particle mass m and the electro-magnetic coupling strength e. We write
where \i is the reduced mass and the plus (minus) sign is for pairs of unlikesign (like-sign) particles. If particle pairs are emitted from a static source at initial relative distance r with a probability Ssta.t(r; K), then the corresponding correlation is given by an average over the squared wave-function (103), C(q, K) = J
.
(106)
In the case of identical particles, the two-particle symmetrised version of (103) should enter equation (106). The if-dependence of the pair emission probability is often neglected when calculating Coulomb corrections. The following Coulomb correction methods are based on this starting point: (1) Point-like Gamow Correction For a point-like source 5stat('") = <^ 3 '(r), the correlator (106) is given by the Gamow factor G(rj) G(v) =
$ C q °/2(0)
2?a? e 27T7)
_
I
(107)
Early studies constructed the corrected like-sign two-particle correlation by dividing the measured correlator by this Gamow factor C£-Hq, K) = C^(q,
K)/G(r,.).
(108)
748
B. Tomdsik and U. A.
Wiedemann
(2) Static Finite Size Correction The point-like Gamow correction (107) largely overestimates the real effect of Coulomb corrections since particles are emitted in reality with finite separation r which leads to a weaker Coulomb interaction. An improvement advocated repeatedly 67 ' 14,7 is to calculate the correction factor for a finite size static source. Typically, a Gaussian ansatz Ssta.t(r) oc exp [—r2/4R2] is chosen in (106), « ( < / ) = / d \ S s t a t (r) | ^ 7 2 ' ( r ) | 2 , &~\q,
K) = Cl~^(q, K)/F*l?M) •
(109) (HO)
Instead of the analytical emission function in eq. (109), the particleemitting source can be characterized in terms of a discrete set of phasespace points obtained e.g. from a Monte Carlo simulation. In this case, the correlation due to particle symmetrization and final state interactions is usually calculated with a so-called afterburner routine. The most widely used afterburner is Scott Pratt's CRAB 70 (CoRrelation After-Burner). In practice, the value for the source width R in eq. (109) is determined iteratively from the extracted HBT radius parameter in the fitting procedure. A finite purity of the sample due to misidentified particle leads to an overall correlation strength A < 1. This can be taken into account by generalising78 eq. (110) to
C£J{q,K)
= (l-A) + AC(-7Hq,K)F?%(q)-
(HI)
(3) Correction of like-sign by unlike-sign correlations Rather than to calculate Coulomb corrections for finite size sources, one can make use of the fact that unlike-sign correlations receive no contribution from Bose-Einstein symmetrisation effects but depend on Coulomb correlations of the same magnitude (but opposite sign), Ci~-\q, K) = C^(q,
K) C^\q,
K).
(112)
Theoretical support for this procedure comes from the fact, that likesign and unlike-sign Coulomb correlations calculated from (106) compensate largely. For point-like sources, e.g., the product of the Gamow factors deviates from unity by less than five percent for relative mo-
Central and Non-Central
HBT from AGS to RHIC
749
menta q > 8 137
G
<"+) G <"->'i +( ,w+o(,,r
ma)
A further improvement over (112) is to take this deviation into account 78 , s-i( ) / K \ _ Cmeas (<7, K) Cmeas (q, ^corr.improvedvy'-"-^ G(n+) Girl-)
K) '
This was shown to work with excellent accuracy for a wide range of source parameters 78 . The effects of finite momentum resolution reduce both Cmeas {q,K) and Cmeas (q,K). As an unwanted consequence, these effects are amplified in the product defining the corrected like-sign correlation functions (112) and (114). An empirical parametrization which takes into account finite momentum resolution is discussed below; see eq. (117). (4) Experimental parametrisations of Coulomb corrections Unlike-sign correlations CmeaJ(q,K) function120 F(qim)
were parametrised by the
= 1 + (G(V+) - l) c-9-v/Qo
f
(115)
which depends on 9inv = Vq2 ~ (q0)2 •
(H6)
The parameter QQ is extracted from the fit. It quantifies a phenomenological finite-size correction for large relative momentum. The function F{.q\m) approaches the Gamow factor (107) for a point-like source, Qo -* oo. In order to take into account the imperfect purity of the sample and the finite experimental momentum resolution, CERES 132 parametrised the Coulomb correction by C^(q,K)
= (l-X)
+ XC^
[wK±(FCoui(<7inv) - 1) + 1] • (117)
In a Monte Carlo simulation of the final momentum resolution, the Coulomb correction function Fcou\(qmv) was obtained by evaluating eq. (109) and reducing F^(q) accordingly. The same Monte Carlo simulation determines WKX which accounts for the depletion of the parameter A due to finite momentum resolution effects. This is chosen
750
B. Tomdsik and U. A.
Wiedemann
such that XWK± gives the "true" corrected intercept parameter, which then multiplies the correction factor (.Fcoui^inv) ~ !)• F ° r perfect momentum resolution, WK± —• 1 and Fcou\(qinv) —> F^(qinv), and the prescription (117) agrees with (111). The above discussion is mainly for static sources and involves q w dependent correction factors only. In principle, Coulomb correction effects are different for the different relative momentum components, as seen from eq. (109). For dynamically expanding sources, a formalism for the calculation of Coulomb corrections exists 4 , but it has not been used so far in comparison to data. Only one of the correction methods listed above, eq. (112), contains some information about expansion effects since it uses the measured unlike-sign correlation as correction factor. 4.2. Experiments (AGS)
at the Alternating
Gradient
Synchrotron
At the AGS of the Brookhaven National Laboratory (BNL) three series of collaborations have measured and published results on Bose-Einstein correlations in fixed target experiments with beam energies varying between 2 AGeV and 11.6 AGeV. Due to lack of statistics, measured correlation functions were parametrised often by a 1-dimensional parametrisation C f a n v H l + Ae-tf-vtf-v,
(118)
where the invariant momentum difference qmv is defined in eq. (116). E802/E859/E866/E917 These experiments use a rotating spectrometer (the "Henry Higgins" Spectrometer) which in the E866 upgrade was supplemented by a Forward Spectrometer. The acceptance is at or close to mid-rapidity. The E802/E859/E866 Collaboration has published a systematic study of the dependence of HBT radius parameters on transverse mass Mx. = \/m2 + K\, system size and centrality dependences 107 . The last of the series, the E917 experiment, collected HBT data on beam energy dependence in Au+Au collisions from 6 to 10.6 AGeV. At the time of this writing, these data are still preliminary 117 . E895 This experiment uses the EOS time projection chamber inherited from Bevalac 112 . BE correlations were measured in 2, 4, 6, and 8 AGeV
"8£g o « _ S
^ CD O 3 3
- 3
cj
O .2 3
CD
°§e s .a •" ftJ
» 4? S
'5
c3 OT
>> 1> X I
-
is «
2 .£ " 3 - -
g B g ^
„
^ 8 "S 3
o
CD - 3 (3 fcj
"
CO
c
O
7,
O
4->
JM
S
ft
CB 3)
OT T 3
H ^ <
Cp < J S
in
«
co g
a -^
CO 3
XJ
.2 «s CO O H
"3 J J
a (S c
•9 "3
O
.3
3
"3 t.
a s
•«
O a3
•SI
ft.
: 3 i* 4^ -Z; co -u A ft
~
o ±1
< Cu a3
O
ft ¥
i-
faX !
ime
a _o 'm
"o u
3
+
< 3
<
CD
o c
-4-J
£ t-i +3 1 CO CO
0 U
B
a
bO 3
'3 3
13 •a 3 co
H
xi «3 'o
fcsO CD
CO
co CO
O
ct)
a t-i
to CO
O O
| a CO CD
XI B
>
OT
<^ 9> z" >.
CD
a CO
« ^
T-I
3 X
IM
O
V C5
00
751
!
CO
a o +" K + a 'ai S a
Si. §
CD
«° ><
CO
iO
CO O
CN
00
sa
O
X
<
^ i
<
3
00
a ^Sg a
OH
a a
00
-a1 T—1
30 3
^
3
3
1l
cB
W) 3
's 3
13 X)
5 ci3
ja «3
"3
-w
a
l" +
HBT from AGS to RHIC
6 B
•s O
+ ^
t
CO
c«
I-l
CO
S 0
'Jn 0
S
o
cO
£ a
> a of s.
=2
-4-9
(6 XI
18
0
Central and Non-Central
CO
2 «-
O >o > J .S i-3 .S II ft! II oi
s o s II
O CO
j
II
•J
CO
O
•s "°
"2 >>
>
+J
'5,
1
12 >>
X!
+J
>>
•*J
•3 ft!
3
^
c
1
ce
i-<
2 CD
G
+J
u S
'1 ft
m IH
i-<
» B c8
±1
^
• *
« ><
II
d
co~ O
CO II
+
X CO
o 3 ft!"
c3
cB
-M>
3
CO
>> •3 'ia,
a CO
> 0/
- ^
CD
^H
XI
o"0 O || iO ^ ^
t~
co
en
O
o T-I
CD
18
o
O
00
CM
r+
IV 00
H
"?l~00
H
CD
2^
> > o o > ^ > ^
"7H
d H 3 6 -3>^
>>
*3 CO
CU
< <^
I
Sr
t~
tO)
-~-> Z
XI
"
c3 ^ft -^
< ~-^ 2 o 2-^
CO
D
CO
r-
l-~ T-I O)
> o
i:
J) ^
ft
•3
8
>
13 t-i
co
O
CO
CO
d
•a
m
Oi 00
> 0)
O t;
2^
• ^
ft
^«
o
CO
s
O
o
3
CO
3
3
CO
13
3
3 ID
CO „ cj
0)
CO „
o
CO „ cj
CD
O
CO
CD
> o
• ^
>
> 3 3^ 2" > 00
0
XI X! ^ E 'a a 1 •§ S i
o
CO
O
CO
CO „
t^
di
d
^j.
in C31 00
CO
>
a a a a a
t^ i-< Ol
d
O
CD
CD
a> 0 0 H
lO
tr
d
d
CO
m
>
CTl 00
•*
d
d
O
0)
> o ^
:**
II •J} II o II
^
O
H
3 aO ^
corr.
752
B. Tomdsik and U. A.
Wiedemann
Au+Au collisions. Published data exist for the transverse mass dependences of correlation radii 114 , and the azimuthal dependences in non-central collisions for the three lower energies 113 . Results on average phase-space density as a function of K± and beam energy were shown at the Quark Matter 2001 conference116. All published data are at mid-rapidity. E814/E877 Particle correlations at projectile rapidity were measured with the Forward Magnetic Spectrometer of the E814/E877 Collaboration. The E814 is a Si-beam experiment 108 while E877 measured Au+Au at 10.8 AGeV 110 . E877 determined the pion phase-space density at freeze-out for the latter system 110 . At the Quark Matter 97 conference, they showed a direct fit of a Gaussian core-halo source function to K± and y-binned correlation function and extracted source radii for pions and for kaons 111 . A summary of AGS experiments is given in Table 4.
4.3. Experiments (SPS)
at the CERN Super Proton
Synchrotron
The CERN SPS was used first to accelerate 1 6 0 and 32 S nuclei to 200 AGeV. Then it was upgraded to accelerate a 158 AGeV 2 0 8 Pb beam. Recently, the CERN SPS delivered Pb-beams at lower energies: 40 and 80 AGeV. The step-by-step improvement of analysis tools during the CERN SPS heavy ion program is clearly seen in the available data. The correlation measurements for oxygen and sulphur beams were parametrised first in terms of qinv only. The later three-dimensional fits do not include the cross-term (18). Also, the Coulomb repulsion was corrected initially by multiplying with a Gamow factor which overestimates the repulsion significantly. Improved Coulomb corrections, based on averaging the squared Coulomb wave function over a finite source size, were only introduced approximately with the arrival of the Pb beam. NA35 Originally, this experiment used a large streamer chamber in a magnetic field to measure tracks of charged particles and their momenta in O+Au collisions. The large volume of the detector allowed for the study of three rapidity windows between -2.4 and 1.6 in the CMS of the nucleonnucleon collision118. Due to small statistics, the correlation function was
Central and Non-Central
HBT from AGS to RHIC
753
Table 5. BE correlations data for central oxygen-induced reactions at 200 AGeV/c. All correlation functions are constructed from hadron pairs and are corrected for final state interactions by the Gamow factor. The parametrisations used by both experiments are listed in the last column. They include an additional factor 1/2 in the exponent of the correlation function. target
collab. r e f -
rapidity
frame
C Cu Ag Au
WA80 1 4 2
- 1 < yiab < 1
lab
4 0 - - 200 MeV/c
9inv and 2d: q±,qi
Au
NA35 1 2 0
CMS
5 0 - - 600 MeV/c
BP no cross-term
0-5
K±
par am.
parametrised in qi = q\on% and q± = yjq20 + q] C{q) = 1 +Aexp
1 2p2
1 2 p2
-Tj^-ty ~ 2QjL J-
(119)
In contrast to other experiments, this parametrisation used by NA35 has an additional factor 1/2 in the exponent. The NA35 detector does not identify pions. Thus NA35 always studied hadron-hadron correlations. With the 0induced collisions a measurement of "single-event interferometry" 119 was attempted. For measurements with the sulphur beam, the detector was upgraded with a time projection chamber (TPC) which was crucial in gaining good statistics for correlation analysis. In a comprehensive study 120 of Sinduced reactions with C, S, Cu, Ag, and Au targets, NA35 measured the rapidity, K±, and multiplicity dependence of Bertsch-Pratt correlation radii. O+Au results were reanalysed in this work with better statistics. The K± dependence of correlation radii was also reported in a letter 121 . Although the existence of a sizeable cross-term was first confirmed by NA35 122 , no results with the cross-term were published. NA49 For the lead beam, this collaboration equipped the NA35 detector with four large TPCs which allow for precise tracking of the secondaries in the rapidity region 2 < y < 5.5 (values given in the laboratory system with J/CMS = 2.9 for the 158 ylGeV Pb beam). The NA49 detector is able to identify particles by a combination of energy loss and time-of-fiight measurements. So far, however, only unidentified hadron-hadron correlations
754
OH
03 CU
CD
4£
3 0
CD
-4J
3 0
CD
3
-t-»
3 O
CD
3
4J
3 O
CD
3
+J
Wiedemann
4J
3 O
3
CD
ft X (H
3
+J
3 O
3
CD
CD
3 3 O CD
15
3 CN
3 CN
3
3
+J J-i
o
+4>
B
CJ
a
1 (-
i
3
W t/1
&1|
3
T3 CO
13 2 CN CN
CO CO
J3 «3
J3 fl)
n ii
s s
4* CN H *
CO CN I f
- *
CD
+^
CD
O
M-l
cj"
O
4H
-H
+J
CJ
CN
H OH °" CQ g
c5
Sci C44 8
+4)
1-4
rf
CN CN
.S .5
X I J3 -4J
CD
ft ft X CD X
ft X CD
>.
CD
XI
ft X
ft CD
4J
CD X!
X
CD CD •4J
CD X!
a X CD -*J
CN
6
-3
3
<+H
o
m
0,
8
-»4>
^
(4-1
0.
-o d"
H *i
2
CO
• ^
V
rl
l o o ^ m
CO
3
2 Tt
in
^ -s
V S •8 V s
o
00
CN •»
v 7'
in
o O CN
in
CO
3
<
z z z>.
0,
< < < < < ^
ho
<
Z Z
CD
CN
o
CO CO
CJ t~.
t—<
s
C+H
t5 H o-
s oa
cj
1
in
I
o
CO
o Tf
CO
2 Tf
3
CM
i n " co
3
in~
'-< J3
^ is
V * x JS
,;
n* V
s -o ti
^ in
t*-
S V d
CN
s
in
o
1 CN
CN
CN
o
< Z
3
o
T« T*
CN
^
1 <
m
o 00
in CO
V
V
V
O
(H
i—1
H
3 XI
in
O
LH
1—1
8
C*H
O
0- OH 05 CQ
8
1
3 -O CO
in
d
O
LH
o
CM
i n " co
CO
w" V s>
CO R
V s> * » 3)
CN
r-t
CN
t-
r-t
N
in
o
CN CO
in
o CO
Tf
<
< < Z Z
"^
Z
OT
— S a V V
XI
• *
CO
lO
O
i-H
4->
J3
d?
t-t
O
%
•M
OH
u*
ce
ea
cj
S3
a
>>
t-.
t-t
t+-i 0>
3, 0
O
H *
i-H
.3 CO CD
+4)
3
T3
xi
CD
B cc5 i~i
ce a
6
01 m
B. Tomdsik and U. A.
CO
3
3 XS O +J
'-£ 3
CD
3 To-
C3
8
ce g
X5 +J
>CD X "co
3
+4>
Xi
CD
o ^ o o CN
cS T3
•4J
.2 cj
-U
•8
(-.
•&
CO
"3 144
>
CO
£ S
CO CD
•"J 1
T3 3 cS O in 1—i
n 0 4-3
J3
+3 CD
< z0
• ^
2
3
CO
1-4 O CJ
LH
CD
CO
3 •a O 3 0 0.
1
%
(H i-H
O O CO CD
CD
CD
X
u
~o
0
3
ce
C
CN
ll
CO CO
CJ CD
c 2 V ft +4> >, a J O •3 CJ 144
> • CD
^•1
v S§
a o '£
0
CD
0 in
SI "3 s. B
4"v T3
t-
'Si £
ce
O
4->
CD
^2 CD
£ ^0 O
0
CD
0 a CD
CD
T3
CD
B C a
2
k\ H ^ H fi -3
HHH H d
i
111
CO
0
1 CO
«3
CD t-H
2
O
t-
'1 CD
ce
444
ce
ft 144 CD
X! •4J
s 0 CO
CO
3 ^
ff^
1^
XI
CD
CQ
0H
ce O 0.
s
CD
4^
CO
3 -^ .2 S
CD
£
0
ns correspond t
CO
3 0
3
|H
3
X!
o
(4-4
_a xCD 3 CO 1
3
,
o
CD
3
<*3
8
-4J
o
8 Z ^ ki
,ou c ce
+^ cS T3
& o S ce
XI
CD
3O 13 4-4
t4
>>
o +^ cj
•8 +-> O
fH
8
PQ ,o <4H
CD
CO
$
3
CMS LCMS
4.5. ublish i was measured i for 2.5 < yiab ii were measure depe dene depe dene
LCMS LCMS
2di - 600 MeV/c < 400 MeV,
2di: 600 MeV/c CMS
600 MeV/c < 400 MeV, CMS LCMS
- 200 MeV/ lab
- 600 MeV/ CMS
600 MeV/c - 200 MeV/
noti frame apidit llab.
arget
are inter; term
nal eact ions ste ind tra ion
+•*
CD
si
CJ - u CU CD
CJ
X (-! £ O CO CD
ce
'8
s
3
CD
J3
T3
JS
s
LH
H c o rH • ^
c .2 cO
J3
t.
0 CO
3 §
6H
CO
6 ix
CT3
>>
CO
kH CD
O
6 CD
<*H
i o
CO
en
«
H Q;
oo "- 1
o c
«
2 O
w-
dies
V
CM
A
o c
t-H
>H
H
E S
o •h
Si
5?
g
CO
8
-a H a. J3 ho
!
J3
c -a CD
J3
S3
cO
-Q
m
cO
CO
J3
3 J2
cfl £3
!n O
3 cO
CO
CO
c
>H
3 o U
ho
4^
CD CO J5 X I
O
R o n. to
J3
$
I
!
S3 O
c
d
'
O) (31
.s
crt T3
a. Xi hn J3
ho
-a
cO
n, S CO
3 O
-a c
% u
J3
cil C
U
J3
1
•v
a a CO H
T5
CO
s-
HBT from AGS to RHIC
CQ
r-H
>
CD
CD 4J
T LH
CD
+3
OH ^ H
OH"
03 CJ
<3)
CN
C/D
d V
V
2 o V
d 1
<
£
00
'f
^1
CO
•J
00
d J3 1 re
s O
CO
^ < ^
CO
O J
« .S
s
CO
o
Cfi
d V
CO
W
OH
„
CQ
OH
CU CQ
>
o
CO
- I -0 6 •*
v 1
0
s
V
CN
^H
PH^ « >
^
CO
oo _ T ! 0 CD o JS -S •g! co .5
o
Central and Non-Central
CJ
^
3 ^"~
_1_
<£ r-t
o m
<*H
_
"g o" £ 00 3
S <8>
CD
2 o
o
^
OH
o^ I—I
OH
03
-S
ce
03
e "cfl
1 CD O
CO
CD CJ
CO
o
iff CO
CO
rCN CO
CO
2
•z
l<
<
j
s
CD
O
CO
< 2
00 CN
^
£
CO •J
S ±!
*H-;
cp
i-
o cj
dra pidit;
2 V.
+2
JS T3
a co -t a3 en
'+-»
c
S3
JH 73
T3
CD
cj
C O
a5 O H
+J
CO
CD
*CC
c H CO
£, S3
CD
( H
CD fc-i
CO
CD
s
CD t-H
3
J> 3> 3 CD £ S iH
LH
CD
1-1
ho C
O
CD -G
o o
kH lH
CD
C o •3 CI ft o p. -a CD cc
ho
"S
_c0
0
a o
CO
CD
cj
cu +3
h en fl O cci cp
-4-5
cO ft
S CO CD
-*J
T3
o
,— go s
'+2
a a .2 CO
K <
VH
o CO C O
, - J
to
**- _S0 en CO
1
CC
8 a In CD
s S
CO
c ft o d o
"o
3
s
c
o> & H 3
O CO OH
+J
CO CJ cu JS J3
co CO
C O 'co
^O
+
OH
OH
o CO X 3
Is -o a o CD
c_>
1 6 i CO CO
O cj
stead tion In det static WA98
a)
CD
3(ti
£
sligh NA44 has
PID rapidit [lab.
incti ed as
radii orr lati BE corn lat 7.
755
756
B. Tomdsik and U. A.
Table 8. collab. r e f E802 1 0 7
Wiedemann
Data on the centrality dependence of correlation radii. system and energy Si+Al @ 14.6 AGeV Si+Au O 14.6 AGeV Au+Au ® 11.6 AGeV
rapidity mid-rapidity
NA35 1 2 0
S+Ag ® 200 AGeV
- 2 . 5 < J/CMS < 1-5, 4 bins
NA44 1 2 7
S+S, S+Ag, S + P b @ 200 AGeV
mid-rapidity
NA45 1 3 2
P b + A u O 40, 80, 158 AGeV
mid-rapidity
NA49136,138
P b + P b @ 40, 80, 158 AGeV
y c M s = o 1 3 8 , 1.3 136
WA97 1 4 3
P b + P b @ 158 AGeV
- 0 . 3 < J/CMS < 0.9
STAR 1 5 0 ' 1 5 3
Au+Au @ V* = 130 AGeV 1 5 0 V« = 200 AGeV 1 3 3
mid-rapidity
PHENIX 1 4 8
Au+Au @ yfl = 200 AGeV
mid-rapidity
are available. For P b + P b collisions at 158 AGeV, the rapidity and K± dependence of YKP parameters was published 134 . Also, a rather comprehensive compilation of preliminary data at all rapidities and in both YKP and Bertsch-Pratt is available 135 . Preliminary results 137,138 exist for h~h~ correlations in 40, 80, and 158 AGeV P b + P b collisions with different centralities. Results for kaon-kaon correlations in central 158 AGeV P b + P b collisions were published very recently 139 . These latter studies are the only ones based on "global tracking" for which tracks from all TPCs are matched before constructing the correlation function. NA44 This experiment is based on a focusing spectrometer. Its narrow acceptance is around mid-rapidity and depends in detail on the detector setting which may be varied. The performance is optimised for particles with small momentum difference and good particle identification is achieved. For the S-beam, one-dimensional correlation functions for S+Pb collisions123, and a kaon interference study 124 were published first. Bertsch-Pratt radii without the cross-term were published in 1995 125 and a dedicated paper was written on their M± dependence 126 . NA44 studied collisions with other
Central and Non-Central
Table 9.
HBT from AGS to RHIC
Data on three-pion interferometry.
collab. r e f -
system and energy
rapidity
NA44 1 2 9
S + P b @ 200 AGeV
mid-rapidity
WA98145,146
STAR 1 5 4
757
P b + P b @ 158 AGeV A u + A u @ v ^ = 130AGeV
- 0 . 8 < ycMS < 0.2 (ycMs) = - 0 . 2 mid-rapidity
Table 10. Data on the azimuthally dependence of HBT radius parameters. All data are taken at central rapidity. collab. r e 1 -
(^-dependent radii
system and energy
E895 1 1 3
all six radii R^((j>)
A u + A u © 2, 4, 6 AGeV
STAR
153
2
2
Rl, R S, R oa, Rf
A u + A u @ y/l = 130, 200 AGeV
targets 127 : S+S, S+Ag, S+Pb. More recently an investigation of threeparticle correlations in S+Pb systems was published 129 . There are two papers with data from the 158 AGeV P b + P b collisions, one for pion 128 , and one for kaon 131 interferometry. These study the M± dependence of the BP correlation radii within the limited acceptance of the detector. NA45-CERES In 1998, the CERES collaboration upgraded their detector with a time projection chamber with radial drift field. This allows for interferometric studies. CERES measured correlations of nonidentified hadrons in Pb+Au collisions at 40, 80 and 158 AGeV with various centralities 132 . WA80 The correlation analysis of WA80 is based on the so-called plastic ball detector which has coverage in the target rapidity region and at low p_i_ < 220MeV/c. For the oxygen beam they used C, Cu, Ag, and Au targets. The sulphur beam was collided with Al and Au targets, the proton beam at 450 AGeV/c was collided with C and Au targets. An earlier analysis of target dependence of the observed HBT radii for O-induced
758
B. Tomdsik and U. A.
7 6 „ 5
40AGeV
•
Wiedemann
a,
80AGeV
•ea ®
*
*
'%. £ 4 h cc" 3 2 1
'•**•
NA44 i WA97 i WA98 i
CERES i • i NA49 prelim, i—9—'
8 7 6
.§ 5 ^ 4
'%.
#
•J
eft
•e_
3 2 -
0.2
0.4 0.6 K, [GeV/c]
0.8
0.2
0.4 0.6 K, [GeV/c]
0.8
0.2
0.4 0.6 K, [GeV/c]
Fig. 6. Summary of the Kj_-dependence of correlation radii measured at the CERN SPS in P b + P b collisions at midrapidity. Data are taken from NA44 1 2 8 , C E R E S 1 3 2 , preliminary NA49 1 3 7 , WA97 1 4 3 and WA98 1 4 6 .
reactions 140 was superseded by a study including also S and p as projectiles in which the detector performance correction was better understood 142 . The latter paper also shows a simple model fit to the correlation function with Coulomb correction (110) instead of Gamow factor multiplication (108). They used an additional factor of 1/2 in the correlation function like NA35 did. Another work 141 analyses these data in the context of intermittency.
WA98 For the lead beam runs, this collaboration made use of the plastic ball calorimeter. To measure charged particles the detector includes a two arm tracking spectrometer with "banana-shaped" 7r~ acceptance around mid-rapidity. WA98 published a study of the K± dependence of correlation radii for identified ir~ in Bertsch-Pratt and YKP parametrisations 144 ' 146 . Three pion interferometry 145 ' 146 and pion phasespace density 146 were also studied.
Central and Non-Central
7 6 „ 5
-i
1
1 1 130 AGeV
1
T _
r
:;
1
1 1 200 AGeV
1
r
1
1
1
h
1
I
h
759
fl
*,
1 4
HBT from AGS to RHIC
of 3 2 1
H
1
I
1
1
H
H
N.
& 4 a? 3
'**i
2 1
4-
J 45
£
H
PHENIX n V PHENIX J I V •—•—i STAR i V ' O i i STAR n i l "
8 7 6
1 -
::
7*.%.
3 2 1 0.2
0.4
_i 0.6 0.8 K, [GeV/c]
i 1
i_ 1.2
1
PHENIX prelim. I - - ^ H STAR prelim. •
_J
0.2
0.4
1
I
0.6 0.8 K, [GeV/c]
1
!_
1
1.2
Fig. 7. Summary of the K±-dependence of correlation radii at mid-rapidity measured at RHIC in Au+Au collisions. Data are taken from STAR 1 5 0 - 1 5 3 and P H E N I X 1 4 7 ' 1 4 8 . Data for v ^ = 200AGeV are taken from transparencies shown at the Quark Matter 2002 conference.
WA97 To measure BE correlations, WA97 used a silicon telescope which provides precise tracking of produced particles within the magnetic field. They do not identify particle species but sample the correlation function with h~h~ pairs. The data are mainly presented in the YKP parametrisation but consistency checks with the Bertsch-Pratt form were performed 143 . Transverse momentum and rapidity dependences of the correlation radii were investigated in the acceptance window —0.3 < y < 0.9 (in CMS) and 0.2GeV/c
4.4. Experiments (RHIC)
at the Relativistic Heavy Ion Collider
Three of the four collaborations at RHIC published results on BE interferometry: STAR, PHENIX, and PHOBOS. Data were taken from Au+Au collisions at CMS energies of 130 and 200 AGeV.
760
B. Tomdsik and U. A.
Wiedemann
STAR uses a time projection chamber in a solenoidal magnetic field. The analysis for y/s = 130 AGeV 150 covers in three bins the .^-dependence from 0.125 GeV/c to 0.45 GeV/c within a rapidity region |J/CMS| < 0.5. Preliminary data at %/s — 200v4GeV range in four bins up to K± ss 0.52 GeV/c 1 5 3 . The centrality dependence of the HBT radii at both energies was studied 150,153 . The azimuthal dependence of the /^.-integrated HBT radii R%, R?s and R^s was also analysed at both energies in minimum bias events 153 . So far, STAR determines the orientation of the reaction plane but it does not determine the direction of the impact parameter. Thus, the azimuthal dependence of R?sl and R^ cannot be measured. Finally, a first study of three pion correlations indicates that the source is fully chaotic 154 . PHENIX extends the results of STAR in a pseudo-rapidity window \r]\ < 0.35 to higher Kx: 0.2 < K± < 1.2 GeV/c at the lower and 0.2 < K± < 2 GeV/c at the higher CMS energy. The particle momentum is measured by a drift chamber and a pad chamber. At y/s = 130AGeV, correlation radii for both positive and negative identified pions are determined in three K± bins 147 . For collisions at -y/i = 200 AGeV, a much better statistics allowed to split the pairs into nine K± bins 148 . The centrality dependence of the correlation radii was also studied. Correlation radii from kaon-kaon correlations did not show 148 a simple M± scaling with the TTTT radii in contrast to expectations from certain hydrodynamically motivated parametrisations of the freeze-out state of the fireball24. PHOBOS presented so far two sets of BP correlation radii for the 15% most central -y/s = 200 AGeV Au+Au collisions. Data are for one K± bin from 0.15 to 0.35 GeV/c and for 0.2 < y < 1.5. One set was measured with 7r+ pairs, the other one with n~ pairs 149 . 4.5. Discussion
of the
data
4.5.1. Size and transverse momentum dependence of HBT radii The out-, side-, and longitudinal HBT radius parameters vary typically around 5 — 6 fm at small transverse pair momentum K± and decrease with increasing K±. Their absolute size shows no significant dependence on beam energy (see Figs. 8 and 9). For data from the CERN SPS, the K_i-s\ope of the longitudinal radius parameter is steeper (see Fig. 6) while all radius
Central and Non-Central HBT from AGS to RHIC
Fig. 8. The %/s and K± dependence of R0 and Ra- Data without error bars are summarised from E895 114 (Au+Au), NA45-CERES132 (Pb+Au), and STAR 7r+7r+150'153 (Au+Au). STAR results for y/s = 200AGeV are taken from transparencies shown at the Quark Matter 2002 conference.
762
B. Tomdsik and U. A.
Wiedemann
Fig. 9. The
parameters measured at RHIC show approximately the same slope (see Fig. 7). As explained in section 2.2.2, the HBT radius parameters of an expanding source correspond to the width of the Xj_-dependent region of homogeneity. This is smaller than the width of the entire collision region. The K± -slope of the HBT radii is a measure of the collective dynamical expansion. This picture can be illustrated by the pocket formulas (100), (101) and is supported by many model comparisons. If SPS data are fitted by a model with Gaussian transverse density distribution 85,89 , this leads to a radius of the entire collision region R » 7 fm. To put this number into perspective, we relate the two-dimensional rms width of the collision region, rrmUsrce = V(%2 + V2) = V%R ~ 10 fm, to the two-dimensional rms widths of a cold lead nucleus. The hard sphere radius Rfcs = 1.2 A1/3 fm is for lead R^ = 7.1 fm, and the corresponding two-dimensional transverse
Central and Non-Central
HBT from AGS to RHIC
763
rms width is r^Ph = y/{x2 + y2)Ph = ^/zjl R^ « 4.4 fin. SPS data favour a model with a transverse box density profile89 over a Gaussian density profile. The box radius is 10 — 12 fm. Irrespective of the transverse profile, one concludes that during the collision the system has expanded by a factor ~ 2 from the transverse size of the overlapping cold lead nuclei to the transverse extension at freeze-out. All available data are subject to significant systematic uncertainties. Additional uncertainties arise when comparing data from fixed target and collider experiments. In view of the rather mild changes of HBT radii between SPS and RHIC, this makes it difficult to assess to what extent the dynamical interpretation given above changes from SPS to RHIC. A first analysis of RHIC data 2 6 argues in favour of a more extended source with larger transverse flow, thus supporting the picture of a more vigorous transverse expansion at higher centre of mass energies. 4.5.2. Ro/Rs The main interest in the quotient or difference of the two transverse HBT radius parameters (15) and (16) lies in a model-dependent argument, that the emission duration (P) can be extracted from 9 ' 6 9 Rl{K) - R*.{K) * 01{P).
(120)
This statement is based on two model-dependent assumptions. First, the term -2f3±(xi) should be negligible compared to (120). This assumption, however, can be violated in models with strong expansion. Second, the difference {x2) — (y2) should be negligible compared to (120). This latter assumption can be violated at sufficiently high K±, in particular in models for which particle emission peaks close to the surface due to dynamical or opacity effects. Many model calculations predict R0/Rs ^> 1. In particular, Rischke and Gyulassy 73 emphasised that this would be an unambiguous signal of an equation of state which is sufficiently soft in the phase transition region to result in a significantly delayed build-up of transverse expansion. This would result in a large lifetime effect, R0/Rs ~ 1-5. In contrast, data indicate values R0/Rs < 1-1 even at RHIC, see Fig. 10. As mentioned in section 3.1, one can think of physics effects which result in R0/Rs < 1. This issue is presently under study 56 ' 59 .
764
B. Tomdsik and U. A.
Wiedemann
0.1
KT[GeV]
O.SNJ sqrt(s) [GeV]
Fig. 10. The y/s and K± dependence of R0/RsData are summarised from E895 1 1 4 , NA45-CERES 1 3 2 , and STAR ^+^+150,153^ S T A R r e s u l t s f r o m ^ = 200AGeV are taken from transparencies shown at the Quark Matter 2002 conference; their error bars may be underestimated.
4.5.3. Average phase-space density Idea and formalism: The spatial average of the phase-space density of pions at times later than freeze-out tf Jd3xf2(x,p,t>tf) (/>(P) = Jd3xf(x,p,t>t ) f
(121)
'
can be expressed in terms of the observable one- and two-particle spectra 11 ' 109 - 33 - 89 dN 1 1 VX E dYM±dM_Ld$
(f)(K±,Y) VeB(K±,Y)
=
Jd3q(C(q,K)-l)
Ve{{(K±,Y)
'
(122) (123)
Central and Non-Central
HBT from AGS to RHIC
765
E895 Negative Pion Phase Space Density
»-u.4
0.35 0.3
• 2AGeV o 4AGeV
T=i20MeV
1
A 6AGeV T 8AGeV
T=1b O M e V \ •\ • \ •
0.25
;
6
\ A
0.2 0.15 0.1
A
•\
:
-
•
o\
\»
T=80Mev\
•
0.05 : 0,0
0\
i\T
• i
i
i
i
I
1
0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 PT (GeV/c)
Fig. 11. Preliminary data on the average phase-space density as measured by the E895 Collaboration for different AGS projectile energies as a function of p_i_116. The lines show Bose-Einstein distributions of given temperature. This does not account for the effects of expansion.
The phase-space density is determined by the average number of pions with given momentum (the non-invariant spectrum) divided by the volume in which they are contained. The factor A~1//2 corrects for the "purity" of the sample: it ensures that only directly produced pions and not those coming from resonance decays are taken into account (see discussion at the end of section 2.1). As long as lifetime effects are small (which is consistent with all data measured so far, see Sec. 4.5.2), the volume is given by (123). For the Cartesian BP parametrisation, it takes the form
V*{K±,Y)
= ^R3(K)yjR*(K)R?(K)
- (R2ol(K))* .
(124)
Thermal model as a reference: For reference, it is customary to compare the measured (/) to that of a thermal distribution of a given temperature and chemical potential. Since (/) is a spatial average, it averages for an expanding source over homogeneity regions which move relatively to the point of maximum emissivity. This leads to deviations from a naive Bose-Einstein phase-space density which for the case of a boost-invariant longitudinally
B. Tomdsik and U. A.
0.35 0.3
Wiedemann
i
i
™r
""•'i
NA44 NA49 WA98
III
766
0.25 A
-
0.2
H—
V
0.15
o
.
o
0.1 0.05 0
0.1
0.2
0.3 0.4 pt [GeV/c]
0.5
0.6
Fig. 12. The average pion phase-space density measured by NA44 1 3 0 , NA49 3 3 , and WA98 1 4 6 in 158 AGeV P b + P b collisions at mid-rapidity.
expanding pion source with transverse expansion can be written as 90
(f)(p±) = f £ A.(P±)J / (f>-l),4 n (pj.)J ,
(125)
/•OO
An(p±) = 47rmXTf / rdre^WM Jo xl0(np±smh<^{r)/T) Ki(nmxcosh((r)/T).
(126)
Here, the transverse rapidity profile is characterised by £(r) and the transverse density profile by fi(r)/T. For illustration, the transverse geometry can be modelled e.g. by a box profile [box]
fi(r) = I
HB for —oo
r < Rbox otherwise'
(127)
or a Gaussian profile [Gauss]
(j,(r) = \IG — T
(128) ^•^Gauss
Central and Non-Central
0.1
0.2
0.3
HBT from AGS to RHIC
0.4
767
0.5
PT
Fig. 13. The average pion phase-space density as measured by STAR Collaboration 1 5 2 for A u + A u collisions at
The chemical potential /x(r) introduced here are position-dependent and result in a non-uniform density profile; HB or \XQ are the values in the centre of the fireball. A certain spatial average of this chemical potential determines the particle multiplicities 90 . For comparison of the present formalism to AGS data, the assumption of boost-invariance entering (125) has to be modified by a rapidity cut-off. Data: The average phase-space density was studied first by the E877 experiment for Au+Au collisions at the AGS 109 ' 110 in the projectile fragmentation region where it was found to decrease with increasing rapidity. The E895 collaboration measured the average phase-space density at mid-rapidity at 2, 4, 6, and 8 AGeV and observed its increase with the collision energy 116 (Fig. 11). At SPS energies there is an extensive compilation of phase-space densities for various collision systems based on data of the NA35/NA49 collaboration 33 , the NA44 measurements of S+S, S+Pb, and P b + P b
768
B. Tomdsik and U. A.
Wiedemann
E
+-
>~4000
*
2000
\
•
10
E895
•
CERES
•
STAR n'n
M
STAR TtV
10
Vs(GeV) Fig. 14. Dependence of the "freeze-out volume" Vf = {2n)3/2R^Ri on y/s. Data are measured by E895 1 1 4 , C E R E S 1 3 2 and STAR 1 5 0 for low-p ± pions at mid-rapidity.
collisions130, and the P b + P b data measured by WA98. For P b + P b collisions at 158 AGeV, the results of these experiments are mutually consistent, see Fig. 12. At RHIC energy, preliminary data indicate that the average pion phasespace density is significantly higher than at SPS. Moreover, the STAR collaboration reported a strong dependence of the phase-space density on centrality 152 (Fig. 13). The origin of this increase, as well as the apparently different p±-dependence of (/) at RHIC and SPS, is currently under study. 4.5.4. Energy and multiplicity dependence Despite the weak dependence of HBT radius parameters on centre of mass energy, the volume Vf from which particles decouple shows an interesting non-monotonous behaviour, see Fig. 14. Here the volume is estimated as Vf ex R^Ri. The radius R0 is not used in calculating Vf because it contains contributions from the temporal extent of the source, see eq. (16). The estimate Vf oc R^Ri does not take into account the effects of the expansion on the HBT radii. Figure 14 shows that the volume decreases gradually
Central and Non-Central
HBT from AGS to RHIC
769
within the AGS energy range, reaches a minimum around the highest AGS energies and then increases monotonously by almost a factor 2 up to the highest RHIC energy. At approximately fixed centre of mass energy at the SPS, the same freeze-out volume was found previously to grow linearly with the charged particle multiplicity per unit rapidity, see Fig. 15. Consistent with this finding is the centrality dependence of the freeze-out volume at SPS energies which grows linearly with the number of participants 132 . These two observations support the conjecture that freeze-out occurs at a fixed particle density. This is, however, contradicted by the non-monotonous energy dependence of Vf. A linear relation between freeze-out volume and particle multiplicity does not hold. Particle density, and thus particle multiplicity is certainly important in characterising the freeze-out condition, since it affects the hadronic escape probability from the medium. However, chemical composition, collective expansion and the momentum of the escaping particle are other factors which determine this escape probability 91 . To illustrate this, one can consider e.g. the mean free path of a pion at freeze-out133 A
mfP = 2np+pOV;v + 3nn-anir
= 2 - — a„N + 3-^-a^
.
(129)
While the freeze-out volume Vf and the numbers AT* of particles of species i contained in Vf both depend significantly on the centre of mass energy, the mean free path (129) is approximately ^-independent 1 3 3 . To understand in more detail how the interplay of different properties of the collision region determines the freeze-out volume, a realistic freeze-out criterion is required. A good starting point is the particle escape probability from the hot and dense but rapidly expanding collision region 29 ' 79 V{x,p,r) =exp - /
("/
dfK(x + vf,p)
.
(130)
Here, v is the velocity of the escaping particle and lZ(x,p) denotes the scattering rate which is defined as the inverse of the mean time between collisions for a particle at position x with momentum p. Freeze-out at different centre of mass energies is then assumed to occur when the probability V reaches a characteristic value. Recently, the scattering rate 1Z(x,p) was calculated for a full hadron resonance gas with chemical composition corresponding to SPS and RHIC
770
B. Tomdsik and U. A.
Wiedemann
CO
E 180
DC
CM ( 0
Go
160 140
S+Ag TPC
120
S+A
100-
P I
S+Au TBC
/ S /'
A (
r
<
80 60-
/
40-
S+Cuf
/
/
StS
S+Ag, SC (sub-samples)
20-
dN/dY -P*M—I-
10
20
30
40
50
60
70
80 90
Fig. 15. The "freeze-out volume" (V/ oc V8R%Ri) as a function of the number of produced negative particles per unit of rapidity. Data are by NA35 1 2 0 for different Sinduced reactions at 200 AGeV beam energy.
energies 91 . The main observation is that the scattering rate shows a significant momentum dependence suggesting that particles of different momenta are emitted at different times. This effect is neglected in hydrodynamic simulations which are based on the Cooper-Frye prescription 22 for freeze-out along a sharp three-dimensional hypersurface. Deviations from this Cooper-Frye prescription, i.e. freeze-out along finite four volumes may affect the transverse momentum slope of HBT radius parameters 30 , and the momentum dependence of the freeze-out volume. The role of a change in the chemical composition from SPS to RHIC was found to be relatively small in spite of the large increase of pion phase-space density (Fig. 13). This is a consequence of the small pion contribution to the total scattering rate, resulting from the comparatively small pion-pion cross-section. In contrast, collective transverse expansion gradients affect the freeze-out volume significantly. The reason is that an increase in the scattering rate at freeze-out can be compensated by stronger transverse flow gradients which lead to a faster density decrease in the collision region thus keeping the opacity integral in the exponent of (130) constant. The possible effect of
Central and Non-Central
HBT from AGS to RHIC
771
Au(2 AGeV)Au
+++, 4-i-f 1 4T +4-' i
0.6
b = 4-8 fm -0-3 < y™ < 0.9 J
cm
pT = 0-0.4 GeV/c ^0.2
out
-I;I
cc
I T
20 .16
long
side
-40
XA- out-long
_l
Tl
T
-
'
4- 4-
out-side
cr
-16
X. 200
... 0
200
.„. 0
200
400
Fig. 16. Azimuthal dependence of HBT radius parameters as published by the E895 Collaboration 1 1 3 for A u + A u collisions at 2 AGeV. Curves correspond to a static source according to the equations (51)-(56).
the flow gradients on the ^/I-dependence of the freeze-out volume (Fig. 14) remains to be studied. 4.5.5. Azimuthal dependence of HBT radius parameters Two years ago, the first measurements of the ^-dependence of HBT radius parameters were published by the E895 Collaboration 113 for beam energies of 2, 4 and 6 j4GeV in semi-peripheral Au+Au collisions at the AGS. Results at all three energies show a sizeable first order harmonics in the
772
B. Tomdsik and U. A.
Wiedemann
^-dependence of R^ and R%t, and comparatively small second order harmonics in the fully transverse HBT radii R%, R% and R%s. The longitudinal radius parameter and the intercept A are consistent with a ^-independent ansatz. According to (59), the first order harmonics allow to reconstruct the angle 0 by which the emission ellipsoid is tilted out of the beam axis, see Fig. 5. At lower AGS energies, this angle is with 0 « 30° surprisingly large. This value is consistent with RQMD transport model simulations 57 . Interestingly, the spatial tilt is found to point in the direction opposite to the directed flow in momentum space. This indicates that at lower AGS energies pion reflection from the bulk of the matter rather than pion absorption by this matter is at the root of the observed direct flow signal. With increasing centre of mass energy, a longitudinally approximately boost-invariant region develops around mid-rapidity. As a consequence, the tilt angle 0 of the emission ellipsoid is expected to decrease with increasing •y/I. However, there is so far no measurement of R%t and R*t at higher energies, which would be required to establish this effect experimentally. The ^-dependence of the fully transverse radius parameters R%, R^s and i?j is easier to measure than that of R?ol and R%t: while the former require the event-wise reconstruction of the orientation of the reaction plane, the latter require in addition the direction in which the impact parameter points. Due to this complication, at RHIC first preliminary data are available for the fully transverse radius parameters only 153 . These data are expected to contain information about whether the spatial orientation of the source at freeze-out is in-plane or out-of-plane. However, statistical and systematic uncertainties in these preliminary data are still too large to draw conclusions. A significant improvement in statistics is expected within the next run. Acknowledgements We are indebted to many colleagues for giving helpful informations. Thanks go in particular to Jan Pisiit for a critical reading of the manuscript. We thank Giuseppe Bruno, John Cramer, Mike Lisa, and Piotr Skowrohski for their comments to the manuscript. We profited from discussions with Harry Appelshauser, Ulrich Heinz, and Mike Lisa. We are also very grateful to Harry Appelshauser, Giuseppe Bruno, John Cramer, Peter Filip, Mike Lisa, Michael Murray, Laurent Rosselet, and Peter Seyboth who provided us with some of the data reviewed here.
Central and Non-Central HBT from AGS to RHIC
773
References Theoretical papers: J. Aichelin, Nucl. Phys. A 617 (1997) 510. 2. S. V. Akkelin and Y. M. Sinyukov, Z. Phys. C 72 (1996) 501. 3. S. V. Akkelin, P. Braun-Munzinger and Y. M. Sinyukov, Nucl. Phys. A 710 (2002) 439. 4. D. Anchishkin, U. W. Heinz and P. Renk, Phys. Rev. C 57 (1998) 1428. 5. H. W. Barz, Phys. Rev. C 53 (1996) 2536. 6. S. A. Bass et al., Prog. Part. Nucl. Phys. 41 (1998) 225. 7. G. Baym and P. Braun-Munzinger, Nucl. Phys. A 610 (1996) 286C. 8. G. F. Bertsch, M. Gong and M. Tohyama, Phys. Rev. C 37 (1988) 1896. 9. G. F. Bertsch, Nucl. Phys. A 498 (1989) 173C. 10. G. F. Bertsch, P. Danielewicz and M. Herrmann, Phys. Rev. C 49 (1994) 442. 11. G. F. Bertsch, Phys. Rev. Lett. 72 (1994) 2349 [Erratum-ibid. 77 (1994) 789]. 12. D. H. Boal, C. K. Gelbke and B. K. Jennings, Rev. Mod. Phys. 62 (1990) 553. 13. J. Bolz, U. Ornik, M. Pliimer, B. R. Schlei and R. M. Weiner, Phys. Rev. D 47 (1993) 3860. 14. M. G. Bowler, Phys. Lett. B 270 (1991) 69. 15. D. A. Brown and P. Danielewicz, Phys. Rev. C 57 (1998) 2474. 16. D. A. Brown and P. Danielewicz, Phys. Rev. C 64 (2001) 014902. 17. S. Chapman and U. W. Heinz, Phys. Lett. B 340 (1994) 250. 18. S. Chapman, P. Scotto and U. W. Heinz, Phys. Rev. Lett. 74 (1995) 4400. 19. S. Chapman, P. Scotto and U. W. Heinz, Heavy Ion Phys. 1 (1995) 1. 20. S. Chapman, J. R. Nix and U. W. Heinz, Phys. Rev. C 52 (1995) 2694. 21. G. Cocconi, Phys. Lett. B 49 (1974) 459. 22. F. Cooper and G. Frye, Phys. Rev. D 10 (1974) 186. 23. T. Csorgo, B. Lorstad and J. Zimany, Z. Phys. C 71 (1996) 491. 24. T. Csorgo and B. Lorstad, Phys. Rev. C 54 (1996) 1390. 25. T. Csorgo, arXiv:hep-ph/0001233. 26. T. Csorgo and A. Ster, arXiv:nucl-th/0207016. 27. G. Goldhaber, S. Goldhaber, W. Y. Lee and A. Pais, Phys. Rev. 120 (1960) 300. 28. P. Grassberger, Nucl. Phys. B 120 (1977) 231. 29. F. Grassi, Y. Hama and T. Kodama, Z. Phys. C 73, 153 (1996). 30. F. Grassi, Y. Hama, S. S. Padula and O. J. Socolowski, Phys. Rev. C 62 (2000) 044904. 31. M. Gyulassy, S. K. Kauffmann and L. W. Wilson, Phys. Rev. C 20 (1979) 2267. 32. M. Gyulassy and S. K. Kauffmann, Nucl. Phys. A 362 (1981) 503.
774 B. Tomasik and U. A. Wiedemann 33. D. Ferenc, U. W. Heinz, B. Tomasik, U. A. Wiedemann and J. G. Cramer, Phys. Lett. B 457 (1999) 347. 34. R. Hanbury Brown and R. Q. Twiss, Phil. Mag. 45 (1954) 663. 35. R. Hanbury Brown and R. Q. Twiss, Nature 178 (1956) 1046. 36. U. W. Heinz, B. Tomasik, U. A. Wiedemann and Y. F. Wu, Phys. Lett. B 382 (1996) 181. 37. U. W. Heinz and B. V. Jacak, Ann. Rev. Nucl. Part. Sci. 49 (1999) 529. 38. U. W. Heinz and P. F. Kolb, arXiv:hep-ph/0204061. 39. U. W. Heinz, A. Hummel, M. A. Lisa and U. A. Wiedemann, Phys. Rev. C 66 (2002) 044903. 40. H. Heiselberg and A. P. Vischer, Eur. Phys. J. C 1 (1998) 593. 41. H. Heiselberg, Phys. Rev. Lett. 82 (1999) 2052. 42. H. Heiselberg and A. M. Levy, Phys. Rev. C 59 (1999) 2716. 43. M. Herrmann and G. F. Bertsch, Phys. Rev. C 51 (1995) 328. 44. T. Hirano, K. Morita, S. Muroya and C. Nonaka, Phys. Rev. C 65 (2002) 061902. 45. T. J. Humanic, Phys. Rev. C 50 (1994) 2525. 46. T. J. Humanic, Phys. Rev. C 53 (1996) 901. 47. T. J. Humanic, arXiv:nucl-th/0205053. 48. W. Kittel, Acta Phys. Polon. B 32 (2001) 3927. 49. P. F. Kolb, J. Sollfrank and U. W. Heinz, Phys. Rev. C 62 (2000) 054909. 50. S. E. Koonin, Phys. Lett. B 70 (1977) 43. 51. G. I. Kopylov and M. I. Podgoretsky, Sov. J. Nucl. Phys. 15 (1972) 219 [Yad. Fiz. 15 (1972) 392]. 52. G. I. Kopylov and M. I. Podgoretsky, Sov. J. Nucl. Phys. 18 (1974) 336 [Yad. Fiz. 18 (1974) 656]. 53. R. Lednicky and V. L. Lyuboshits, Sov. J. Nucl. Phys. 35 (1982) 770 [Yad. Fiz. 35 (1982) 1316]. 54. R. Lednicky and V. L. Lyuboshits, Heavy Ion Phys. 3 (1996) 93. 55. R. Lednicky, V. L. Lyuboshits, B. Erazmus and D. Nouais, Phys. Lett. B 373 (1996) 30. 56. Z.W. Lin, C.M. Ko and S. Pal, arXiv:nucl-th/0204054. 57. M. A. Lisa, U. W. Heinz and U. A. Wiedemann, Phys. Lett. B 489 (2000) 287. 58. A. N. Makhlin and Y. M. Sinyukov, Z. Phys. C 39 (1988) 69. 59. L. McLerran and S. S. Padula, arXiv:nucl-th/0205028. 60. D. Molnar and M. Gyulassy, arXiv:nucl-th/0211017. 61. K. Morita, S. Muroya, H. Nakamura and C. Nonaka, Phys. Rev. C 61 (2000) 034904. 62. K. Morita, S. Muroya, C. Nonaka and T. Hirano, Phys. Rev. C 66 (2002) 054904. 63. J. Y. Ollitrault, Nucl. Phys. A 638 (1998) 195C. 64. M. I. Podgoretsky, Sov. J. Nucl. Phys. 37 (1983) 272 [Yad. Fiz. 37 (1983) 455].
Central and Non-Central HBT from AGS to RHIC 775 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100.
A. M. Poskanzer and S. A. Voloshin, Phys. Rev. C 58 (1998) 1671. S. Pratt, Phys. Rev. Lett. 53 (1984) 1219. S. Pratt, Phys. Rev. D 33 (1986) 72. S. Pratt, Phys. Rev. D 33 (1986) 1314. S. Pratt, T. Csorgo and J. Zimany, Phys. Rev. C 42 (1990) 2646. see S. Pratt's web page: http://www.nscl.msu.edu/~pratt/freecodes/crab/home.html. R. L. Ray, Phys. Rev. C 57 (1998) 2523. D. H. Rischke and M. Gyulassy, Nucl. Phys. A 597 (1996) 701. D. H. Rischke and M. Gyulassy, Nucl. Phys. A 608 (1996) 479. B. R. Schlei, U. Ornik, M. Plumer and R. M. Weiner, Phys. Lett. B 293 (1992) 275. B. R. Schlei, Phys. Rev. C 55 (1997) 954. B. R. Schlei, D. Strottman, J. P. Sullivan and H. W. van Hecke, Eur. Phys. J. C 10 (1999) 483. E. V. Shuryak, Phys. Lett. B 44, 387 (1973). Y. Sinyukov, R. Lednicky, S. V. Akkelin, J. Pluta and B. Erazmus, Phys. Lett. B 432 (1998) 248. Y. M. Sinyukov, S. V. Akkelin and Y. Hama, Phys. Rev. Lett. 89 (2002) 052301. S. Soff, S. A. Bass and A. Dumitru, Phys. Rev. Lett. 86 (2001) 3981. S. Soff, S. A. Bass, D. H. Hardtke and S. Y. Panitkin, Phys. Rev. Lett. 88 (2002) 072301 S. Soff, arXiv:hep-ph/0202240. H. Sorge, H. Stocker and W. Greiner, Nucl. Phys. A 498 (1989) 567C. H. Sorge, H. Stocker and W. Greiner, Annals Phys. 192 (1989) 266. A. Ster, T. Csorgo and B. Lorstad, Nucl. Phys. A 661 (1999) 419. J. P. Sullivan et al, Phys. Rev. Lett. 70 (1993) 3000. B. Tomasik and U. W. Heinz, Eur. Phys. J. C 4 (1998) 327. B. Tomasik and U. W. Heinz, Acta Phys. Slov. 49 (1999) 251. B. Tomasik, U. A. Wiedemann and U. W. Heinz, arXiv:nucl-th/9907096. B. Tomasik and U. W. Heinz, Phys. Rev. C 65 (2002) 031902. B. Tomasik and U. A. Wiedemann, arXiv:nucl-th/0207074. S. A. Voloshin and W. E. Cleland, Phys. Rev. C 53 (1996) 896. S. A. Voloshin and W. E. Cleland, Phys. Rev. C 54 (1996) 3212. S. Voloshin, R. Lednicky, S. Panitkin and N. Xu, Phys. Rev. Lett. 79 (1997) 4766. U. A. Wiedemann, P. Scotto and U. W. Heinz, Phys. Rev. C 53 (1996) 918. U. A. Wiedemann and U. W. Heinz, Phys. Rev. C 56 (1997) 3265. U. A. Wiedemann and U. W. Heinz, Phys. Rev. C 56 (1997) 610. U. A. Wiedemann, Phys. Rev. C 57 (1998) 266. U. A. Wiedemann, D. Ferenc and U. W. Heinz, Phys. Lett. B 449 (1999) 347. U. A. Wiedemann and U. W. Heinz, Phys. Rept. 319 (1999) 145.
776
B. Tomdsik and U. A.
Wiedemann
101. Y. F. Wu, U. W. Heinz, B. Tomasik and U. A. Wiedemann, Eur. Phys. J. C 1 (1998) 599. 102. F. B. Yano and S. E. Koonin, Phys. Lett. B 78 (1978) 556. 103. W. A. Zajc et al., Phys. Rev. C 29 (1984) 2173. 104. W. A. Zajc, Phys. Rev. D 35 (1987) 3396. 105. D. Zschiesche, S. Schramm, H. Stocker and W. Greiner, Phys. Rev. C 65 (2002) 064902. Experimental papers: 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. 133. 134. 135.
V. Cianciolo [E802 Collaboration], Nucl. Phys. A 590 (1995) 459C. L. Ahle [E802 Collaboration], Phys. Rev. C 66 (2002) 054906. J. Barrette et al. [E814 Collaboration.], Phys. Lett. B 333 (1994) 33. D. Miskowiec et al. [E877 Collaboration], Nucl. Phys. A 610 (1996) 227C. J. Barrette et al. [E877 Collaboration], Phys. Rev. Lett. 78 (1997) 2916. J.P. Wessels et al. [E877 Collaboration], Nucl. Phys. A 638 (1998) 69. G. Rai et al., IEEE Trans. Nucl. Sci. 37 (1990) 56. M. A. Lisa et al. [E895 Collaboration], Phys. Lett. B 496 (2000) 1. M.A. Lisa et al. [E895 Collaboration], Phys. Rev. Lett. 84 (2000) 2798. S. Y. Panitkin et al. [E895 Collaboration], Phys. Rev. Lett. 87 (2001) 112304. M. A. Lisa et al. [E895 Collaboration], Nucl. Phys. A 698 (2002) 185. B. Holzmann for the E917 Colaboration, Heavy Ion Physics from Bevalac to RHIC, ed. by R. Seto, (World Scientific, Singapore, 1999). A. Bamberger et al. [NA35 Collaboration], Phys. Lett. B 203 (1988) 320. T. J. Humanic et al. [NA35 Collaboration], Z. Phys. C 38 (1988) 79. T. Alber et al. [NA35 Collaboration], Z. Phys. C 66 (1995) 77. T. Alber et al. [NA35 Collaboration.], Phys. Rev. Lett. 74 (1995) 1303. T. Alber [NA35 and NA49 Collaborations], Nucl. Phys. A 590 (1995) 453C. H. Boggild et al. [NA44 Collaboration], Phys. Lett. B 302 (1993) 510 [Erratum-ibid. B 306 (1993) 418]. H. Beker et al. [NA44 Collaboration], Z. Phys. C 64 (1994) 209. H. Boggild et al. [NA44 Collaboration], Phys. Lett. B 349 (1995) 386. H. Beker et al. [NA44 Collaboration], Phys. Rev. Lett. 74 (1995) 3340. K. Kaimi et al. [NA44 Collaboration], Z. Phys. C 75 (1997) 619. I. G. Bearden et al. [NA44 Collaboration], Phys. Rev. C 58 (1998) 1656. H. Boggild et al. [NA44 Collaboration], Phys. Lett. B 455 (1999) 77. M. Murray and B. Holzer, Phys. Rev. C 63 (2001) 054901. I. G. Bearden et al. [NA44 Collaboration], Phys. Rev. Lett. 87 (2001) 112301. D. Adamova et al. [NA45 CERES Collaboration], arXiv:nucl-ex/0207005. D. Adamova et al. [NA45 CERES Collaboration], arXiv:nucl-ex/0207008. H. Appelshauser et al. [NA49 Collaboration], Eur. Phys. J. C 2 (1998) 661. R. Ganz et al. [NA49 Collaboration], arXiv:nucl-ex/9808006.
Central and Non-Central HBT from AGS to RHIC
777
136. R. Ganz [NA49 Collaboration], Nucl. Phys. A 661 (1999) 448. 137. P. Seyboth et al. [NA49 Collaboration], arXiv:hep-ex/0206046. 138. C. Blume for the NA49 Collaboration, arXiv:nucl-ex/0208020, (Proceedings of Quark Matter 2002, to be published in Nucl. Phys. A), see also transparencies at h t t p : / / a l i c e - f r a n c e . i n 2 p 3 . f r / q m 2 0 0 2 / T r a n s p a r e n c i e s / 19Plenary/Blume.pdf. 139. S. V. Afanasiev et al. [NA49 Collaboration], arXiv:nucl-ex/0210018. 140. R. Albrecht et al. [WA80 Collaboration], Z. Phys. C 53 (1992) 225. 141. R. Albrecht et al. [WA80 Collaboration], Phys. Rev. C 50 (1994) 1048. 142. T. C. Awes et al. [WA80 Collaboration], Z. Phys. C 69 (1995) 67. 143. F. Antinori et al. [WA97 Collaboration], J. Phys. G 27 (2001) 2325. 144. M. M. Aggarwal et al. [WA98 Collaboration], Eur. Phys. J. C 16 (2000) 445. 145. M. M. Aggarwal et al. [WA98 Collaboration], Phys. Rev. Lett. 85 (2000) 2895. 146. M. M. Aggarwal [WA98 Collaboration], arXiv:nucl-ex/0210002. 147. K. Adcox et al. [PHENIX Collaboration], Phys. Rev. Lett. 88 (2002) 192302. 148. A. Enokizono for the PHENIX Collaboration, arXiv:nucl-ex/0209026, (Proceedings of Quark Matter 2002, to be published in Nucl. Phys. A). 149. S. Manly for the PHOBOS Collaboration, arXiv:nucl-ex/0210036, (Proceedings of Quark Matter 2002, to be published in Nucl. Phys. A). 150. C. Adler et al. [STAR Collaboration], Phys. Rev. Lett. 87 (2001) 082301. 151. F. Retiere [STAR Collaboration], arXiv:nucl-ex/0111013. 152. L. Ray for the STAR Collaboration, arXiv:nucl-ex/0211030 (Proceedings of Quark Matter 2002, to be published in Nucl. Phys. A). 153. M. Lopez Noriega for the STAR Collaboration, arXiv:nucl-ex/0210031 (Proceedings of Quark Matter 2002, to be published in Nucl. Phys. A). 154. R. Willson for the STAR Collaboration, arXiv:nucl-ex/0211026 (Proceedings of Quark Matter 2002, to be published in Nucl. Phys. A).
Quark-Gluon Plasmai
World Scientific www.worldscientific.com 5029 he
ISBN 981-238-077-9