You are on page 1of 18

BNL—38530

BROOKHAVIN NATIONAL LABORATORY


DE87 001374
July 1986 BNL-

' > • • >

I9
QUARK DECONFINEMENT AND HIGH ENERGf
NUCLEAR COLLISIONS

H. Satz
Fakultat fur Physik
Universitat Bielefeld, D-48 Bielefeld, F.R. Germany
and
Physics Department
Brookltaven National Laboratory, Upton, NY 11973, USA

ABSTRACT

Statistical QCD predicts that with increasing density, strongly inter-


acting matter will undergo a transition to a plasma of deconfined quarks
and gluons. High energy heavy ion collisions are expected to permit exper-
imental studies of this transition and of the predicted new state of matter.

Talk given at the XXIII International Conference on High


Energy Physics, 1C-23 July 1980, Berkeley, California, USA.

This manuscript has been authored under contract number D10-ACO2-7OCIIOUU1C with the U.S. Depn.rt.-
ment of Energy. Accordingly, the U.S. Government retains a non-exclusive, royalty-free license to publish or
reproduce the published form of this contribution, or allow others to do so, for U.S. Government purposes.

DISTRIBUTION OF THIS DOCUMEKT !S UNLIMITED yfi


1. Introduction

High energy hadron collisions have provided the empirical basis for strong interaction
dynamics. It is our hope that high energy heavy ion collisions will do the same for strong
interaction thermodynamics, that they will become the tool with which we can carry out
the experimental analysis of strongly interacting matter. We recall that thermodynamics,
in studying the collective behaviour of many components, may well lead to new physics,
beyond the known dynamics. In solid state physics, ferromagnetism and superconductivity
are just two striking examples of this. We shall see shortly that the statistical mechanics
obtained with QCD as dynamical input in fact does predict new states of matter. Heavy
ion experiments, starting this fall at the AGS (Brookhaven)1 and the SPS (CERN)2 will
attempt to observe and study these states experimentally.
Already on a phenomenological level, we may expect strongly interacting matter to
undergo two phase transitions: colour deconfinement and chiral symmetry restoration.
Let us therefore begin by looking at these transitions as specific cases of rather general
phenomena in statistical physics.
The binding potential V (r) between two charges separated by a distance r is modified
by the presence of other charges. In atomic physics, we obtain the Debye-screened form
V (r) exp {—r/ru), where the screening radius rj) decreases with increasing charge density.
When TJJ ai rg, where rg denotes the binding radius, the system undergoes a transition
from insulator to conductor, since the screening now liberates the valence electron. Simi-
larly, we expect for strongly interacting matter a transition from colour insulator (hadronic
matter) to colour conductor (quark-gluon plasma), once the density of colour charges is
sufficient to make rjy^rjj, where rjj is the hadron radius.3 At sufficiently high density,
the long range, confining part of the potential thus becomes screened and we expect quark
deconfinement.
A conduction electron in a metal generally has a different effective mass than that of an
electron in vacuum or bound in an isolated hydrogen atom. This shift in mass is a collective
effect due to the lattice and the other electrons in the conductor. Similarly, we expect that
the quarks inside a hadron will have a different effective mass than they have in a plasma
at high density. For the latter, asymptotic freedom leads eventually to a vanishing quark
mass, while inside hadrons we have constituent quarks of mass m|ff a 300 MeV. A theory
with massless quarks is chirally symmetric. At low density, this symmetry must thus be
spontaneously broken; at high density, when m|ff —> 0, chiral symmetry is restored.
We thus expect two types of transition in strongly interacting matter: quark decon-
finement as the result of colour charge screening at high density, and chiral symmetry
restoration as the result of a shift in the quark mass when the system changes from colour
insulator to conductor. Are the two phenomena connected, and what are the transition
parameters? For this, we now turn to statistical QCD.

2. S t a t i s t i c a l Q C D

The basis for strong interaction dynamics is the QCD Lagrangian

t = -\{d»Aav -dvA%- gfe 4 Aiy -Y^$f(if tf

where A and ip are the gluon and quark fields, respectively; the corresponding colour
indices take on the values a, b, c = 1, • • •, 8 and a, (3 = 1,2,3. We shall in general consider
only two quark flavours (u, d); these we take as massless.
Strong interaction thermodynamics is then based on the partition function

Z (T, /x, V) = Tr{exp [- [H - n N) /T]}, (2)

where H (£,) is the Hamiltonian, N the net overall baryon number, and (j, the associated
baryonic chemical potential.
With Eq. (1) and (2), statistical QCD is in principle completely defined. Given Z, we
obtain by differentiation all thermodynamic observables of interest, such as

e=(T*/V)(dlnZ/dT)llfT>v (3)

for the energy density, or


nB = {T/V) (dinZ/dn)Ty (4)

for the baryon number density.


To evaluate Z, however, we need a regularizing, non-perturbative method of treating
the relativistic quantum field problem implied by Eq. (1). Since we want to study critical
behaviour, a perturbative approach as used in QED is not adequate. So far, the only
scheme which allows a solution of this problem is the lattice formulation proposed by K.
Wilson.4 It leads to a partition function similar in form to those encountered in the statis-
tical mechanics of spin systems and can therefore be solved by Monte Carlo simulation on
sufficiently large computers.5 We shall here go immediately to the results obtained through
this approach; for a recent survey giving more details of formulation and evaluation, see
Ref. 6.
The energy density of an ideal gas of massless quarks and gluons, for colour SU (3)
and two quark flavours, is given by the generalized Stefan-Boltzmann form

(5)

An ideal gas of massless pions leads to

(6)

The energy density of strongly interacting matter, as obtained by computer simulation of


the QCD expression (3) with Eq. (1) and (2), is shown in fig. 1. It is displayed there as
a function of the inverse coupling 6/g2, which is connected to the physical temperature
through the renormalization group relation lnT ~ 6/g2. At low temperature, the energy
density is seen to be close to the pion gas value (6); at a point corresponding to about
Tc ~ 200MeV, it changes abruptly and starts to approach the ideal plasma value (5).
To see if this transition indeed corresponds to deconfinement, we need an order pa-
rameter indicating which phase the system is in. It is given by the average thermal Wilson
loop L, 7 ' 8 which is related to the free energy F of an isolated quark,

L~exp{-F/T}. (7)

In the confinement regime, there can be no free quarks and F becomes infinite; once
colour screening sets in, it is possible to separate a qq pair, and hence F becomes finite.
Confinement is thus indicated by Z ~ 0, and when deconfinement occurs, L jumps abruptly
to much larger values. In fig. 2 we see that this is indeed the behaviour which follows from
statistical QCD, and the transition temperatures from L and e agree, as expected.
Next, we want to look at chiral symmetry restoration. Here the order parameter is
(•4>ip), which provides a measure of the effective quark mass. Hence for (ipip) a 0, we have
chiral symmetry, while for (iptp) >• 0 it is spontaneously broken. In fig. 3 we compare the
order parameters for deconfinement and chiral symmetry restoration; it is seen that the
two phenomena occur at the same transition temperature.
Let us note that the results of statistical QCD, which we have just summarized, have
been obtained by a number of different groups using quite different techniques to accom-
modate quarks in the lattice formulation (for a compilation, see Ref. 6). They all find
very similar behaviour, both qualitatively and quantitatively. We can therefore conclude
that statistical QCD predicts for vanishing baryon number density a meson gas at low
temperatures, colour deconfinement and chiral symmetry restoration at Tc OL 200 MeV,
and for higher temperatures, a plasma of deconfined, massless quarks and gluons. The
energy density necessary to attain this plasma is about 2.5GeV/fm3.
What happens when we now pass from "mesonic" matter with vanishing baryon num-
ber density to "baryonic" matter, for which fi ^ 0 in Eq. (2)? In this case the Monte
Carlo evaluation encounters technical difficulties (complex quark determinant), which are
presently being addressed by various groups; they are not yet resolved, however. There-
fore, only approximative results9 are available so far; nevertheless, they give us some
first indications of what to expect. In fig. 4 we show the temperature dependence of the
deconfinement order parameter L for different values of the baryonic chemical potential
(i. As fj. and hence the baryon number density increase, the transition point is shifted to
lower temperatures. Such an effect is expected: deconfinement can be caused either by
heating or by compression, and if compressed, the system requires less heating to become
critical. The use of pressure — here in the form of an increased baryon number density —
in inducing deconfinement is seen quite clearly in fig. 5, where the order parameter shows
a ^-dependence very similar to the T-dependence of fig. 2. The present, approximative
evaluation goes up to /j, a 800 MeV; so far, deconfinement and chiral symmetry restoration
continue to coincide, as seen in fig. 6. Whether this will hold down to T = 0 remains to
be seen; in principle, a three phase structure is not yet excluded.
In summary for this section, it can be said that we are slowly beginning to understand
the phase structure of strongly interacting matter as predicted by QCD. We hope that
adequate simulation techniques for \x ^ 0 will soon allow the completion of the T — fi
phase diagram and a precise determination of the transition parameters.
3. C o n d i t i o n s i n N u c l e a r Collisions

Can heavy ion collisions lead to systems of the type considered in statistical QCD?
And if that is the case, are the energy densities obtained sufficient for the formation of the
quark-gluon plasma? Let us begin by estimating the expected energy density.10
In proton-proton collisions, at incident energies high enough to be in the scaling regime,
we observe per unit rapidity interval in the central region (|yCms| 5: 1/2) about two to three
secondaries, each with an energy of about 0.5GeV. This corresponds to an energy deposit
of about 0.3GeV/ fm3. In the collision of two nuclei of mass number A, we have in the
same space-time region a superposition of interactions from those nucleons which "sit
behind each other." As a result, the energy deposit now becomes e^ ~ 0.3 A 1 / 3 ; for very
heavy ions, such as E/238, this brings us in the vicinity of the transition point. Two further
features lead to a considerable increase of e^. We can write more generally11

€A*fo)A(dN/dy)A/(*RA\A), (8)

where {po)A denotes the energy per secondary in the A — A collision, (dN/dy)A the number
of secondaries per central rapidity interval, R^ the nuclear radius, and A^ the formation
length — the longitudinal size of the interaction region leading to the production of the
secondaries considered. While A ~ 1 fm for p — p collisions, the observed enhanced nuclear
stopping12 corresponds to a A^ < 1 fm, which in turn increases eA.n In addition, a trigger
on high multiplicity events (larger than average [dN/dy)A) results in a further increase.
Energy densities of 20 — 50GeV/ fm3 thus become feasible, and we can conclude that the
eA values achievable in nuclear collisions do allow plasma formation.
The volume of strongly interacting matter produced in this way is evidently not
"macroscopic." Nevertheless, it is a factor A2/3 larger than the typical hadronic inter-
action volume, and for high initial values of eA, the system can expand considerably and
still remain in the plasma phase. Volumes of many hundred times the hadronic volume
thus become possible, so that there is hope for an application of statistical QCD.
Finally, we turn to the rather crucial issue of tests for a new state of matter. How do
we know if a quark-gluon plasma was formed, and how can we study its properties? While
probes for different plasma features have been considered for some time,13 an apparently
unambiguous signature for its formation was proposed only very recently.14 We shall first
sketch this signature, and then summarize briefly the main probes discussed so far.
The basic feature of the quark-gluon plasma is deconfinement: at sufficiently high
colour charge density, resonance formation is not possible in the plasma. On the other
hand, the production of heavy resonances, such as the J/ip, is possible only very early in
the collision; in the hadronization stage, the production is suppressed by exp (—M/Tc),
where M denotes the resonance mass and Tc ^ 200MeV. Hence the J/ip signal in the
lepton pair spectrum observed in nuclear collisions must be strongly suppressed if such a
collision indeed leads to plasma formation.14 Since the J/ip interacts only very weakly
with nuclear matter, it is moreover only the formation of a deconfined plasma which can
cause such a suppression.
To make these considerations more quantitative, we must know the J/rp radius as well
as the screening length ro at a given temperature; only for rp < rj^ can we expect J/tp
suppression. Charmonium models15 lead for T ~ Tc to a radius of about 0.5 fm; above Tc,
the J/ip can in fact exist only as cc state bound by a Debye-screened Coulombic potential,
and once rp < 0.3 fm, even that becomes impossible. Lattice calculations16 indicate that
at T/Tc ?i 1.5, rj) < 0.2 fm. Hence in a plasma of that temperature, we can indeed not
have any J/ip formation.
In the lepton pair spectrum from p — p collisions, the J/tp peak rises two orders of
magnitude above the Drell-Yan background. The suppression of this peak for dilepton
spectra from nuclear collisions should thus provide both a direct test for deconfinement
and an unambiguous signature for quark-gluon plasma formation.
The main features of the plasma, which are to be studied by the probes so far produced,
are its temperature, its expansion and subsequent hadronization, and its strangeness con-
tent.
The plasma temperature can be determined by measuring the thermal photon or dilep-
ton spectra from nuclear interactions.17 Thus the thermal lepton pair distribution will have
a Boltzmann form, exp (-M/T), which for not too small a pair mass M is dominated by
the initial temperature of the plasma.
If the transition from quark-gluon plasma to hadronic matter is of first order, then
there should be a coexistence regime, in which the entropy density changes at constant
8

Tc from its high plasma value to its low hadronic matter value, as a consequence of the
expansion of the system. Since the average transverse momentum px of the secondaries
give an indication of their emission temperature,18 and since the multiplicity measures the
entropy, this leads to a characteristic relation between py a n d the average particle number
per central rapidity interval.19 High multiplicity cosmic ray events may show a behaviour
of this form.20
If the quark plasma in its initial stage is also in equilibrium with respect to the produc-
tion of different quark species ("chemical" equilibrium), then the strange quark content is
determined at that point. It was first conjectured that this would lead to enhanced strange
particle production21; subsequent considerations,22 however, indicate that expansion ef-
fects must be taken into account.
For some further plasma probes — its stopping power, the spatial size of the plasma
bubbles formed, etc. — see Ref. 13.
In conclusion: we expect that high energy nuclear collisions will make possible an
experimental analysis of strongly interacting matter. On the one hand, this would allow a
test of statistical QCD; on the other hand, it provides an empirical simulation of the state
of matter of the early universe.

Acknowledgments

It is a pleasure to acknowledge the hospitality of the Aspen Center for Physics, where
this talk was prepared and written. Work supported in part by the Department of Energy
under contract DE-AC02-76CH00016.
REFERENCES

1. See the talk by R. Ledoux in these Proceedings.


2. See the talk by D. Lissauer in these Proceedings.
3. H. Satz, Nucl. Phys. A418 447c (1984).
4. K. Wilson, Phys. Rev. D10, 2445 (1974).
5. M. Creutz, Phys. Rev. Lett. 43, 553 (1979).
6. H. Satz, Ann. Rev. Nucl. Part. Sci. 35, 245 (1985).
7. L.D. McLerran an B. Svetitsky, Phys. Lett. 98B, 195 (1981).
8. J. Kuti, J. Poldnyi and K. Szlachanyi, Phys. Lett. 98B, 199 (1981).
9. B. Berg et al., Z. Phys. C31, 167 (1986).
10. J.D. Bjorken, Phys. Rev. D27, 140 (1983).
11. L.D. McLerran, in Quari Matter 1984, Lecture Notes in Physics 221. K. Kajantie
(Ed.), Springer Verlag, Berlin-Heidelberg-New York (1985).
12. W. Busza, Nucl. Phys, A418, 635c (1984).
13. For a survey, see M. Guylassy, Nucl. Phys. A418, 59c (1984).
14. T. Matsui and H. Satz, Phys. Lett. B (in press), Brookhaven Preprint BNL-38344,
June 1986.
15. See e.g., D.L. Scharre, in Physics in Collisionl, E.G. Bellini and P. Trower (Eds.),
Plenum Press, New York (1985).
16. K. Kanaya and H. Satz, Phys. Rev. D (in press), Bielefeld Preprint BI-TP 86/16,
May 1986;
T.A. DeGrand and C.E. DeTar, "Static Screening Lengths in the Gluon Plasma",
Colorado Preprint COLO-HEP-113 (1986).
17. E.L. Feinberg, Nuovo Cim. 34A, 391 (1976);
E.M. Shuryak, Sov. J. Nucl. Phys. 28, 408 (l978)Yad. Fiz. 28, 796 (1978);
G. Domokos and J. Goldman, Phys. Rev. D23. 203 (1981);
K. Kajantie and H. Miettinen, Z. Phys. C9, 241 (1981).
18. R. Hagedorn, Nuovo Cim. Supple. 3, 147 (1965).
10

19. L. Van Hove, Phys. Lett. 118B. 138 (1982).


20. See O. Miyamura, in Quark Matter 1984, Lecture Notes in Physics 221, K. Kajantie
(Ed.), Springer Verlag, Berlin-Heidelberg-New York (1985).
21. J. Rafelsky, Nucl. Phys. A418, 215C (1984).
22. T. Matsui, L.D. McLerran and B. Svetitsky, Phys. Rev. D (in press), MIT Preprint
CTP 1320, November 1985.
11

FIGURE CAPTIONS

Fig. 1: Energy density of strongly interacting matter at baryonic chemical potential //. = 0.
Fig. 2: Deconfinement order parameter at fi = 0.
Fig. 3: Deconfinement (•) and chiral symmetry restoration (o).
Fig. 4: Deconfinement for different values of the baryonic chemical potential, with fia = fi/3T.
Fig. 5: Deconfinement as function of the baryonic chemical potential.
Fig. 6: Critical parameters for deconfinement (x) and chiral symmetry restoration (•).
SB
12

10

FIG. 1
1.5

1.0

0.5

5.0 5.5 6.0 6.5


2
6/g -

FIG. 2
0.02 -

FIG. 3
1.5 = 0.600

488
400
330
0.200
0.100

1.0

0.5

0 Hh
5.0 5.2 5.4 5.6

FIG. 4
1.5

1.0

0.5

:
t I I I f I t I
0.0 0.2 • 0.4 0.6 0.8
fid = /x/3T

FIG. 5
1/Al

ISO1
x

«
X

125

100-
J 1 I I
0 100 200

FIG. 6
DISCLAIMER

This report was prepared as an account of work sponsored by an agency of the United States
Government. Neither the United States Government nor any agency thereof, nor any of their
employees, makes any warranty, express or implied, or assumes any legal liability or responsi-
bility for the accuracy, completeness, or usefulness of any information, apparatus, product, or
process disclosed, or represents that its use would not infringe privately owned rights. Refer-
ence herein to any specific commercial product, process, or service by trade name, trademark,
manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recom-
mendation, or favoring by the United States Government or any agency thereof. The v'ews
and opinions of authors expressed herein do not necessarily state or reflect those of the
United States Government or any agency thereof.

You might also like