1: \documentclass[12pt,epsf,epsfig,psfig]{article}
2: %\documentstyle[12pt]{article}
3: \usepackage{graphicx}
4: \usepackage{epsfig}
5: \oddsidemargin 15pt
6: \topmargin 0pt
7: \headheight 00pt
8: \headsep 00pt
9: \textheight 235mm
10: \textwidth 160mm
11: \voffset=-0.5cm
12: \hoffset=-0.5cm
13: \parindent=0pt
14:
15: % put your own definitions here:
16:
17: \def\J{$J/\psi$}
18: \def\j{J/\psi}
19: \def\X{$\chi_c$}
20: \def\x{\chi}
21: \def\P{$\psi'$}
22: \def\p{\psi'}
23: \def\U{$\Upsilon$}
24: \def\u{\Upsilon}
25: \def\C{c{\bar c}}
26: \def\B{b{\bar b}}
27: \def\cg{c{\bar c}\!-\!g}
28: \def\bg{b{\bar b}\!-\!g}
29: \def\b{b{\bar b}}
30: \def\q{q{\bar q}}
31: \def\Q{Q{\bar Q}}
32: \def\e{\epsilon}
33: \def\t{\tau}
34: \def\l{\Lambda_{\rm QCD}}
35: \def\A{$A_{\rm cl}$}
36: \def\a{\alpha}
37: \def\N{$n_{\rm cl}$}
38: \def\n{n_{\rm cl}}
39: \def\S{S_{\C}}
40: \def\s{s_{\rm cl}}
41: \def\bb{\bar \beta}
42: \def\chiral{\psi {\bar \psi}}
43: \def\CMP{{ Comm.\ Math.\ Phys.\ }}
44: \def\NP{{ Nucl.\ Phys.\ }}
45: \def\PL{{ Phys.\ Lett.\ }}
46: \def\PR{{ Phys.\ Rev.\ }}
47: \def\PRep{{ Phys.\ Rep.\ }}
48: \def\PRL{{ Phys.\ Rev.\ Lett.\ }}
49: \def\RMP{{ Rev.\ Mod.\ Phys.\ }}
50: \def\ZP{{ Z.\ Phys.\ }}
51: \def\EPJ{{Eur.\ Phys.\ J.\ }}
52:
53: \def\be{\begin{equation}}
54: \def\ee{\end{equation}}
55: \def\lsim{\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
56: \def\gsim{\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
57:
58:
59: \begin{document}
60:
61:
62: BI-TP 2006/06\hfill 28.\ 2.\ 06
63:
64: ~\vskip 1cm
65:
66: \centerline{\Large \bf Charm and Beauty in a Hot Environment$^*$}
67:
68: \vskip1cm
69:
70: \centerline{\large \bf Helmut Satz}
71:
72: \vskip 0.5cm
73:
74: \medskip
75:
76: \centerline{Universit\"at Bielefeld, Germany}
77:
78: %\centerline{Fakult\"at f\"ur Physik, Universit\"at Bielefeld}
79:
80: %\centerline{Postfach 100 131, D-33501 Bielefeld, Germany}
81:
82: \centerline{and}
83:
84: \centerline{Instituto Superior T\'ecnico, Lisboa, Portugal}
85:
86: %\centerline{Centro de F{\'i}sica Teorica de Particulas (CFTP)}
87:
88: %\centerline{Instituto Superior T\'ecnico, Av. Rovisco Pais, P-1049-001
89: %Lisboa, Portugal}
90:
91: \vskip1cm
92:
93: \centerline{\bf Abstract:}
94:
95: \vskip0.5cm
96:
97: We discuss the spectral analysis of quarkonium states in a hot medium of
98: deconfined quarks and gluons, and we show that such an analysis provides
99: a way to determine the thermal properties of the quark-gluon plasma.
100:
101: \vskip1cm
102:
103: \noindent{\large \bf 1.\ Introduction}
104:
105: \bigskip
106:
107: Since some thirty years it is known that besides the almost massless
108: {\it up} and $down$ quarks of the everyday world, and the still relatively
109: light {\it strange} quarks required to account for the strange mesons and
110: hyperons observed in hadron collisions, there are heavy quarks at the other
111: end of the scale, whose bare masses alone are larger than those of most of
112: the normal hadrons. These heavy quarks first showed up in the discovery of
113: the \J ~meson \cite{Ting}, of mass of 3.1 GeV; it is a bound state of a
114: {\it charm} quark ($c$) and its antiquark ($\bar c$), each having a mass
115: of some 1.2--1.5 GeV. On the next level there is the \U~meson \cite{Leder},
116: with a mass of about 9.5 GeV, made up of a {\it bottom} or {\it beauty}
117: quark-antiquark pair ($b \bar b$), with each quark here having a mass
118: around 4.5 - 4.8 GeV. Both charm and bottom quarks can of course also
119: bind with normal light quarks, giving rise to open charm ($D$) and open
120: beauty ($B$) mesons. The lightest of these `light-heavy' mesons have
121: masses of about 1.9 GeV and 5.3 GeV, respectively.
122:
123: \medskip
124:
125: The bound states of a heavy quark $Q$ and its antiquark $\bar Q$ are
126: generally referred to as quarkonia. Besides the initially discovered
127: vector ground states \J~and \U, both the $c \bar c$ and the $b \bar b$
128: systems give rise to a number of other {\it stable} bound states of
129: different quantum numbers. They are stable in the sense that their mass
130: is less than that of two light-heavy mesons, so that strong decays are
131: forbidden. The measured stable charmonium spectrum contains the $1S$
132: scalar $\eta_c$ and vector \J, three $1P$ states \X~(scalar, vector
133: and tensor), and the $2S$ vector state \P, whose mass is just below
134: the open charm threshold. There are further charmonium states above the
135: \P; these can decay into $D \bar D$ pairs, and we shall here restrict
136: our considerations only to quarkonia stable under strong interactions.
137:
138: \medskip
139:
140: \hrule width5cm
141:
142: \medskip
143:
144: $*$ Dedicated to Adriano Di Giacomo on the occasion of his 70th birthday
145:
146: \newpage
147:
148: The study of quarkonia has played and continues to play a major role
149: in many aspects of QCD: in spectroscopy, in production and decay, and
150: last but not least as probe of hot QCD media \cite{nora}. And in the
151: development of our understanding of and our feeling for the charm and
152: the beauty of strong interactions, Adriano Di Giacomo has played and
153: certainly will continue to play a major role. Neither my competence nor
154: the space and time available here allow me to give an overview of these
155: contributions; so let me just cite a few recent ones which are of
156: particular importance to what I want to discuss here \cite{DG,A-DG}.
157:
158: \medskip
159:
160: Quarkonia are rather unusual hadrons. The masses of the light hadrons,
161: in particular those of the non-strange mesons and baryons, arise almost
162: entirely from the interaction energy of their nearly massless quark
163: constituents. In contrast, the quarkonium masses are largely determined
164: by the bare charm and bottom quark masses. These large quark masses allow
165: a very straightforward calculation of many basic quarkonium properties,
166: using non-relativistic potential theory. It is found that, in particular,
167: the ground states and the lower excitation levels of quarkonia are very
168: much smaller than the normal hadrons, and that they are very tightly bound.
169: Now deconfinement is a matter of scales: when the separation between normal
170: hadrons becomes much less than the size of these hadrons, they melt to form
171: the quark-gluon plasma. What happens at this point to the much smaller
172: quarkonia? When do they become dissociated? That is the main question
173: we want to address here.
174: We shall show that the disappearance of specific quarkonia signals the
175: presence of a deconfined medium of a specific temperature \cite{M-S}.
176: Thus the study of the quarkonium spectrum in a given medium
177: is akin to the spectral analysis of stellar media, where the
178: presence or absence of specific excitation lines allows a determination
179: of the temperature of the stellar interior.
180:
181: \medskip
182:
183: We had defined quarkonia as bound states of heavy quarks which are
184: stable under strong decay, i.e., $M_{c\bar c} \leq 2 M_D$ for charmonia
185: and $M_{b\bar b} \leq 2 M_B$ for bottomonia. Since the quarks are heavy,
186: with $m_c \simeq 1.2 - 1.5 $ GeV for the charm and $m_b \simeq 4.5 - 4.8$
187: GeV for the bottom quark, quarkonium spectroscopy can be studied quite well
188: in non-relativistic potential theory. The simplest (``Cornell'') confining
189: potential \cite{Cornell} for a $\Q$ at separation distance $r$ has the form
190: \be
191: V(r) = \sigma ~r - {\alpha \over r}
192: \label{cornell}
193: \ee
194: with a string tension $\sigma \simeq 0.2$ GeV$^2$ and a Coulomb-like term
195: with a gauge coupling $\alpha \simeq \pi/12$. The corresponding
196: Schr\"odinger equation
197: \be
198: \left\{2m_c -{1\over m_c}\nabla^2 + V(r)\right\} \Phi_i(r) = M_i \Phi_i(r)
199: \label{schroedinger}
200: \ee
201: then determines the bound state masses $M_i$ and the wave functions
202: $\Phi_i(r)$, and with
203: \be
204: \langle r_i^2 \rangle = \int d^3r~ r^2 |\Phi_i(r)|^2 /
205: \int d^3r~|\Phi_i(r)|^2 .
206: \label{radii}
207: \ee
208: the latter in turn provide the (squared) average bound state ``radii'',
209: here defined as the $\Q$ separation for the state in question.
210:
211:
212: \medskip
213:
214: The solution of eq.\ (\ref{schroedinger}) gives in fact a very good
215: account of the full (spin-averaged) quarkonium spectroscopy, as seen in
216: Table 1 \cite{D-K-S}. The line labelled $\Delta M$ shows the differences
217: between the experimental and the calculated values; they are in all cases
218: less than 1 \%. The input parameters for these results were $m_c=1.25$ GeV,
219: $m_b=4.65$ GeV, $\sqrt \sigma = 0.445$ GeV, $\alpha=\pi/12$.
220: We see that in particular the \J~ and the lower-lying bottomonium
221: states are very tightly bound ($ 2M_{D,B} - M_0 \gg \l)$ and of very
222: small spatial size ($r_0 \ll 2 r_h \simeq 2$ fm). What happens to them
223: in a hot and dense medium?
224:
225: \vskip0.5cm
226:
227: \hskip1.5cm
228: \renewcommand{\arraystretch}{1.8}
229: \begin{tabular}{|c|c|c|c|c|c|c|c|c|}
230: \hline
231: {\rm state}& $J/\psi$ & $\chi_c$ & $\psi'$ & $\Upsilon$
232: & $\chi_b$ &
233: $\Upsilon'$ & $\chi_b'$ & $\Upsilon''$ \\
234: \hline
235: {\rm mass~[GeV]}&
236: 3.10&
237: 3.53&
238: 3.68&
239: 9.46&
240: 9.99&
241: 10.02&
242: 10.26&
243: 10.36 \\
244: \hline
245: $\Delta E$ {\rm[GeV]}&0.64&0.20&0.05 &1.10&
246: 0.67&0.54&0.31&0.20 \cr
247: \hline
248: $\Delta M$ {\rm[GeV]}&0.02&-0.03&0.03 & 0.06&
249: -0.06&-0.06&-0.08&-0.07 \cr
250: \hline
251: {$r_0$ \rm [fm]}&0.50&0.72&0.90&
252: 0.28&0.44& 0.56& 0.68 &0.78 \cr
253: \hline
254: \end{tabular}
255:
256: \bigskip
257:
258: \centerline{Table 1:
259: Quarkonium Spectroscopy from Non-Relativistic Potential Theory \cite{D-K-S}}
260:
261: \vskip0.8cm
262:
263: \noindent{\large \bf 2.\ Interaction Range and Colour Screening}
264:
265: \bigskip
266:
267: Consider a colour-singlet bound state of a heavy quark $Q$ and its
268: antiquark $\bar Q$, put into the medium in such a way that we can measure
269: the energy of the system as function of the $\Q$ separation $r$ (see Fig.\
270: \ref{string-b}). The quarks are assumed to be heavy so that they are static
271: and any energy changes indicate changes in the binding energy. We consider
272: first the case of vanishing baryon density; at $T=0$, the box is therefore
273: empty.
274:
275: \medskip
276:
277: \begin{figure}[htb]
278: \hspace*{0.1cm}
279: \centerline{\epsfig{file=box2.eps,width=8cm}}
280: \caption{String breaking for a $\Q$ system}
281: \label{string-b}
282: \end{figure}
283:
284: \medskip
285:
286: In vacuum, i.e., at $T=0$, the free energy of the $\Q$ pair is
287: assumed to have the string form \cite{Cornell}
288: \be
289: F(r) \sim \sigma r
290: \label{string}
291: \ee
292: where $\sigma \simeq 0.16$ GeV$^2$ is the string tension as
293: determined in the spectroscopy of heavy quark resonances (charmonium
294: and bottomonium states). Thus $F(r)$ increases with separation
295: distance; but when it reaches the value of a pair of dressed
296: light quarks (about the mass of a $\rho$ meson), it becomes energetically
297: favorable to produce a $\q$ pair from the vacuum, break the string
298: and form two light-heavy mesons ($Q\bar q$) and ($\bar Q q$). These
299: can now be separated arbitrarily far without changing the energy of the
300: system (Fig.\ \ref{string-b}).
301:
302: \medskip
303:
304:
305: The string breaking energy for charm quarks is found to be
306: \be
307: F_0 = 2(M_D - m_c) \simeq 1.2~{\rm GeV};
308: \label{charm-break}
309: \ee
310: for bottom quarks, one obtains the same value,
311: \be
312: F_0 = 2(M_B - m_b) \simeq 1.2~{\rm GeV},
313: \label{bottom-break}
314: \ee
315: using in both cases the quark mass values obtained in the solution
316: leading to Table 3. Hence the onset of string breaking is evidently
317: a property of the vacuum as a medium. It occurs when the two heavy
318: quarks are separated by a distance
319: \be
320: r_0 \simeq 1.2~{\rm GeV}/\sigma \simeq 1.5~{\rm fm},
321: \label{stringbreak}
322: \ee
323: independent of the mass of the (heavy) quarks connected by the string.
324:
325: \medskip
326:
327: If we heat the system to get $T>0$, the medium begins to contain light
328: mesons, and the large distance $\Q$ potential $F(\infty,T)$ decreases, since
329: we can use these light hadrons to achieve an earlier string breaking through
330: a kind of flip-flop recoupling of quark constituents \cite{miya}, resulting
331: in an
332: effective screening of the interquark force (see Fig.\ \ref{flip}).
333: Near the deconfinement point, the hadron density increases rapidly, and
334: hence the recoupling dissociation becomes much more effective, causing
335: a considerable decrease of $F(\infty,T)$.
336:
337: \begin{figure}[h]
338: \hspace*{0.1cm}
339: \centerline{\epsfig{file=flipflop.eps,width=9cm}}
340: \caption{In-medium string breaking through recoupling}
341: \label{flip}
342: \end{figure}
343:
344: \medskip
345:
346: A further increase of $T$ will eventually bring the medium to the
347: deconfinement point $T_c$, at which chiral symmetry restoration causes
348: a rather abrupt drop of the light quark dressing (equivalently, of the
349: constituent quark mass), increasing strongly the density of constituents.
350: As a consequence, $F(\infty,T)$ now continues to drop sharply.
351: Above $T_c$, light quarks and gluons become deconfined colour charges, and
352: this quark-gluon plasma leads to a colour screening, which limits the range
353: of the strong interaction. The colour screening radius $r_D$, which determines
354: this range, is inversely proportional to the density of charges, so that
355: it decreases with increasing temperature. As a result, the $\Q$ interaction
356: becomes more and more short-ranged.
357:
358: \medskip
359:
360: In summary, starting from $T=0$, the $\Q$ probe first tests vacuum string
361: breaking, then a screening-like
362: dissociation through recoupling of constituent quarks, and
363: finally genuine colour screening. In Fig.\ \ref{OKfree}, we show the
364: behaviour obtained in full two-flavour QCD for the colour-singlet
365: $\Q$ free energy as a function of $r$ for different $T$ \cite{K-Z}.
366:
367: \begin{figure}[htb]
368: \hspace*{-0.3cm}
369: \centerline{\epsfig{file=Ofree.eps,width=9cm}}
370: \caption{The colour singlet $\Q$ free energy $F(r,T)$ vs.\ $r$ at
371: different $T$ \cite{K-Z}}
372: \label{OKfree}
373: \end{figure}
374:
375: \medskip
376:
377: It is evident in Fig.\ \ref{OKfree} that the asymptotic value $F(\infty,T)$,
378: i.e., the energy needed to separate the $\Q$ pair, decreases with increasing
379: temperature, as does the separation distance at which ``the string breaks''.
380: For the moment we consider the latter to be defined by the point beyond which
381: the free energy remains constant within errors, returning in section 4.3 to a
382: more precise definition. The behaviour of both quantities is shown in Fig.\
383: \ref{free-T}. Deconfinement is thus reflected very clearly in the temperature
384: behaviour of the heavy quark potential: both the string breaking energy
385: and the interaction range drop sharply around $T_c$. The latter decreases
386: from hadronic size in the confinement region to much smaller values in the
387: deconfined medium, where colour screening is operative.
388:
389: \medskip
390:
391: \begin{figure}[htb]
392: \mbox{
393: \hskip1cm
394: \epsfig{file=Free-inf.eps,width=5cm}
395: \hskip3cm
396: \epsfig{file=debye.eps,width=5cm,height=3.6cm}}
397: %\vskip0.4cm
398: \caption{String breaking potential and interaction range at different
399: temperatures}
400: \label{free-T}
401: \end{figure}
402:
403: The in-medium behaviour of heavy quark bound states thus does serve
404: quite well as probe of the state of matter in QCD thermodynamics. We
405: had so far just considered $\Q$ bound states in general. Let us now
406: turn to a specific state such as the \J. What happens when the range
407: of the binding force becomes smaller than the radius of the state?
408: Since the $c$ and the $\bar c$ can now no longer see each other, the
409: \J~ must dissociate for temperatures above this point. Hence the
410: dissociation points of the different quarkonium states provide a
411: way to measure the temperature of the medium. The effect is illustrated
412: schematically in Fig. \ref{melt-t}, showing how with increasing
413: temperature the different charmonium states ``melt'' sequentially
414: as function of their binding strength; the most loosely bound state
415: disappears first, the ground state last.
416:
417: \medskip
418:
419: Moreover, since finite temperature lattice QCD also provides the
420: temperature dependence of the energy density \cite{schlad}, the
421: melting of the different charmonia or bottomonia can be specified
422: as well in terms of $\e$. In Fig\ \ref{melt-e}, we illustrate this,
423: combining $\e(T)$ with the force radii shown in Fig.\ \ref{free-T}.
424: It is evident that although \P~and \X~are both expected to melt
425: around $T_c$, the corresponding dissociation energy densities could
426: still differ.
427:
428:
429: \begin{figure}[htb]
430: \hskip1cm
431: {\epsfig{file=charm1.eps,width=6.5cm}}
432: \end{figure}
433: \hskip4cm
434: \begin{figure}[htb]
435: \vspace*{-6.4cm}
436: \hspace*{8.5cm}
437: {\epsfig{file=charm2.eps,width=7cm}}
438: %\end{figure}
439: \vspace*{0.5cm}
440: \caption{Charmonium spectra at different temperatures}
441: \label{melt-t}
442: \end{figure}
443:
444: \medskip
445:
446: \vspace*{-0.3cm}
447: \begin{figure}[htb]
448: \hspace*{-0.3cm}
449: \centerline{\epsfig{file=debye-L.eps,width=5.5cm}}
450: \caption{Charmonium dissociation vs.\ temperature and energy density}
451: \label{melt-e}
452: \end{figure}
453:
454: \medskip
455:
456: To make these considerations quantitative, we thus have to find a way
457: to determine the in-medium melting points of the different quarkonium
458: states. This problem has been addressed in three different approaches:
459: \vskip-0.2cm
460: \begin{itemize}
461: \vskip-0.2cm
462: \item{Model the heavy quark potential as function of the temperature,
463: $V(r,T)$, and solve the resulting Schr\"odinger equation
464: (\ref{schroedinger}).}
465: \vspace*{-0.2cm}
466: \item{Determine the internal energy $U(r,T)$ of a $\Q$ pair at separation
467: distance $r$ from lattice results for the corresponding free energy $F(r,T)$,
468: using the thermodynamic relation
469: \vskip-0.5cm
470: \be
471: U(r,T) = -T^2 \left( {\partial [F(r,T)/T] \over \partial T}\right)
472: = F(r,T) - T \left({\partial F(r,T) \over \partial T}\right),
473: \label{intern}
474: \ee
475: and solve the Schr\"odinger equation with $V(r,T)=U(r,T)$ as the binding
476: potential.}
477: \vskip-0.5cm
478: \item{Calculate the quarkonium spectrum directly in finite temperature
479: lattice QCD.}
480: \end{itemize}
481: Clearly the last is the only model-independent way, and it will
482: in the long run provide the decisive determination. However, the direct
483: lattice study of charmonium spectra has become possible only quite
484: recently, and so far most results are obtained in quenched QCD
485: (no dynamical light quarks);
486: corresponding studies for bottomonia are still more difficult.
487: Hence much of what is known so far is based on Schr\"odinger equation
488: studies with different model inputs. To illustrate the model-dependence
489: of the dissociation parameters, we first cite some early work using different
490: models for the temperature dependence of $V(r,T)$, then some recent
491: studies based on lattice results for $F(r,T)$, and finally summarize
492: the present state of direct lattice calculations of charmonia in finite
493: temperature media.
494:
495: \vskip0.5cm
496:
497: \noindent{\large \bf 3.\ Potential Model Studies}
498:
499: \bigskip
500:
501: The first quantitative studies of finite temperature charmonium dissociation
502: \cite{K-M-S} were based on screening in the form obtained in one-dimensional
503: QED, the so-called Schwinger model. The confining part of the Cornell
504: potential (\ref{cornell}), $V(r) \sim \sigma r$, is the solution of the
505: Laplace equation in one space dimension. In this case, Debye-screening leads
506: to \cite{Dixit}
507: \be
508: V(r,T) \sim \sigma r \left\{ {1-e^{-\mu r} \over \mu r} \right\},
509: \label{schwinger-e}
510: \ee
511: where $\mu(T)$ denotes the screening mass (inverse Debye radius) for the
512: medium at temperature $T$. This form reproduces at least qualitatively
513: the convergence to a finite large distance value $V(\infty,T) =
514: \sigma/\mu(T)$, and since $\mu(T)$ increases with $T$, it also gives
515: the expected decrease of the potential with increasing temperature.
516: Combining this with the usual Debye screening for the $1/r$ part of
517: eq.\ (\ref{cornell}) then leads to
518: \be
519: V(r,T) \sim \sigma r \left\{ {1-e^{-\mu r} \over \mu r} \right\}
520: - {\alpha \over r} e^{-\mu r} = {\sigma \over \mu}
521: \left\{ 1-e^{-\mu r} \right\} - {\alpha \over r} e^{-\mu r}
522: \label{schwinger}
523: \ee
524: for the screened Cornell potential. In \cite{K-M-S}, the screening mass
525: was assumed to have the form $\mu(T) \simeq 4~T$, as obtained in first
526: lattice estimates of screening in high temperature $SU(N)$ gauge theory.
527: Solving the Schr\"odinger equation with these inputs, one found that
528: both the \P~and the \X~become dissociated essentially at $T \simeq T_c$,
529: while the \J~persisted up to about $1.2~T_c$. Note that as function of
530: the energy density $\e \sim T^4$, this meant that the \J~really survives
531: up to much higher $\e$.
532:
533: \medskip
534:
535: This approach has two basic shortcomings:
536: \vspace*{-0.2cm}
537: \begin{itemize}
538: \item{The Schwinger form (\ref{schwinger-e}) corresponds to the screening
539: of $\sigma r$ in one space dimension; the correct result in three space
540: dimensions is different \cite{Dixit}.}
541: \vspace*{-0.2cm}
542: \item{The screening mass $\mu(T)$ is assumed in its high energy form;
543: lattice studies show today that its behaviour near $T_c$ is quite
544: different \cite{D-K-K-S}.}
545: \end{itemize}
546:
547: While the overall behaviour of this approach provides some first insight
548: into the problem, quantitative aspects require a more careful treatment.
549:
550: \medskip
551:
552: When lattice results for the heavy quark free energy as function of the
553: temperature first became available, an alternative description appeared
554: \cite{D-P-S1}. It assumed that in the thermodynamic relation (\ref{intern})
555: the entropy term $-T(\partial F /\partial T)$ could be neglected, thus
556: equating binding potential and free energy,
557: \be
558: V(r,T) = F(r,T) -T(\partial F /\partial T) \simeq F(r,T).
559: \label{DPS}
560: \ee
561: Using this potential in the Schr\"odinger equation (\ref{schroedinger})
562: specifies the temperature dependence of the different charmonium
563: masses. On the other hand, the large distance limit of $V(r,T)$
564: determines the temperature variation of the open charm meson $D$,
565: \be
566: 2 M_D(T) \simeq 2 m_c + V(\infty,T)
567: \label{D-mass}
568: \ee
569: In fig.\ \ref{masses}, we compare the resulting open and hidden charm
570: masses as function of temperature. It is seen that the \P~mass falls
571: below $2~M_D$ around 0.2 $T_c$, that of the \X~at about 0.8 $T_c$;
572: hence these states disappear by strong decay at the quoted temperatures.
573: Only the ground state \J~survives up to $T_c$ and perhaps
574: slightly above; the lattice data available at the time did not extend
575: above $T_c$, so further predictions were not possible.
576:
577: \begin{figure}[htb]
578: \centerline{\epsfig{file=free.eps,width=6cm}}
579: \caption{Temperature dependence of open and hidden charm masses \cite{D-P-S1}}
580: \label{masses}
581: \end{figure}
582:
583: The main shortcoming of this approach is quite evident.
584: The neglect of the entropy term in the potential reduces $V(r,T)$
585: and hence the binding. As a result, the $D$ mass drops faster with
586: temperature than that of the charmonium states, and it is this effect
587: which leads to the early charmonium dissociation. Moreover,
588: in the lattice studies used here, only the colour averaged free energy
589: was calculated, which leads to a further reduction of the binding
590: force.
591:
592: \medskip
593:
594: We conclude from these attempts that for a quantitative potential theory
595: study, the free energy has to be formulated in the correct
596: three-dimensional screened Cornell form, and it then has to be
597: checked against the space- and temperature-dependence of the corresponding
598: colour singlet quantity obtained in lattice QCD.
599:
600: \vskip0.5cm
601:
602: \noindent{\large \bf 4.\ Screening Theory}
603:
604: \bigskip
605:
606: The modification of the interaction between two charges immersed in a
607: dilute medium of charged constituents is provided by Debye-H\"uckel theory,
608: which for the Coulomb potential in three space dimensions leads to the
609: well-known Debye screening,
610: \be
611: {1\over r} ~\to~{1\over r} \e^{-\mu r},
612: \label{debye}
613: \ee
614: where $r_D=1/\mu$ defines the screening radius \cite{Landau}.
615: Screening can be evaluated more generally \cite{Dixit}
616: for a given free energy $F(r) \sim r^q$ in $d$ space dimensions, with
617: an arbitrary number $q$. We shall here apply this to the two terms
618: of the Cornell form, with $q=1$ for the string term, $q=-1$ for the gauge
619: term, in $d=3$ space dimensions \cite{D-K-K-S}.
620:
621: \medskip
622:
623: We thus assume that the screening effect can be calculated separately for
624: each term, so that the screened free energy becomes
625: \be
626: F(r,T) = F_s(r,T) + F_c(r,T) =
627: \sigma r f_s(r,T) - {\alpha \over r} f_c(r,T).
628: \label{screen-cornell}
629: \ee
630: The screening functions $f_s(r,T)$ and $f_c(r,T)$ must satisfy
631: \begin{eqnarray}
632: f_s(r,T)&=&f_c(r,T)= 1 ~~{\rm for}~ T \rightarrow 0,\nonumber\\
633: f_s(r,T)&=&f_c(r,T)= 1 ~~{\rm for}~ r \!\rightarrow 0,
634: \label{screen-f}
635: \end{eqnarray}
636:
637: since at $T=0$ there is no medium, while in the short-distance limit
638: $T^{-1} \gg r \to 0$, the medium has no effect. The resulting forms are
639: \cite{Dixit}
640: \be
641: F_c(r,T) = -{\alpha \over r}\left[e^{-\mu r} + \mu r\right]
642: \label{screen-gauge}
643: \ee
644: for the gauge term, and
645: \begin{eqnarray}
646: \hspace{-0.4cm}F_s(r,T) = {\sigma \over \mu}
647: \left[ {\Gamma\left(1/4\right) \over 2^{3/2}
648: \Gamma\left(3/4\right)}-{\sqrt{\mu r} \over 2^{3/4}\Gamma\left(3/4\right)}
649: K_{1/4}[{(\mu r)}^2]\right]
650: \label{screen-string}
651: \end{eqnarray}
652:
653: for the string term. The first term in eq.\ (\ref{screen-string})
654: gives the constant large distance limit due to colour screening; the
655: second provides a Gaussian cut-off in $x=\mu r$, since
656: $K_{1/4}(x^2) \sim \exp\{-x^2))$; this is in contrast to the exponential
657: cut-off given by the Schwinger form (\ref{schwinger}), but in accord
658: with string breaking arguments of Di Giacomo et al.\ \cite{A-DG}.
659:
660: \begin{figure}[htb]
661: \hspace*{-0.3cm}
662: {\epsfig{file=Oplcf_above.eps,width=8cm}}
663: {\epsfig{file=Oplcf_below.eps,width=8cm}}
664: \hspace*{4cm}(a)\hspace*{7.5cm}(b)
665: \caption{Screening fits to the $\Q$ free energy $F(r,T)$ for $T\geq T_c$
666: (left) and $T\leq T_c$ (right) \cite{D-K-K-S}}
667: \label{free}
668: \end{figure}
669:
670: At temperatures $T > T_c$, when the medium really consists of unbound
671: colour charges, we thus expect the free energy $F(r,T)$ to have the form
672: (\ref{screen-cornell}), with the two screened terms given by eqs.\
673: (\ref{screen-gauge}) and (\ref{screen-string}). In Fig.\ \ref{free}a,
674: it is seen that the results for the colour singlet free energy calculated
675: in two-flavour QCD \cite{OK} are indeed described very well by this form,
676: with $m_c=1.25$ GeV and $\sqrt \sigma = 0.445$ GeV, as above.
677: The only parameter to be determined, the screening mass
678: $\mu$, is shown in Fig.\ \ref{sc-mass}, and as expected, it first increases
679: rapidly in the transition region and then turns into the perturbative
680: form $\mu \sim T$.
681:
682: \medskip
683:
684: The behaviour of $\Q$ binding in a plasma of unbound quarks and gluons is
685: thus well described by colour screening. Such a description is in fact
686: found to work well also for $T<T_c$, when quark recombination leads to an
687: effective screening-like reduction of the interaction range, provided one
688: allows higher order contributions in $x=\mu r$ in the Bessel function
689: $K_{1/4}(x^2)$ governing string screening \cite{D-K-K-S}. The resulting fit
690: to the two-flavour colour singlet free energy below $T_c$ is shown in
691: Fig.\ \ref{free}b, using
692: $K_{1/4}(x^2+ (1/2) x^4)$ in eq.\ (\ref{screen-string}); the corresponding
693: values for the screening mass are included in Fig.\ \ref{sc-mass}.
694:
695: \begin{figure}[htb]
696: \hspace*{-0.3cm}
697: \centerline{\epsfig{file=mu_ttc.eps,width=7cm}}
698: \caption{The screening mass $\mu(T)$ vs.\ $T$ \protect\cite{D-K-K-S,D-K-S}}
699: \label{sc-mass}
700: \end{figure}
701:
702: \medskip
703:
704: With the free energy $F(r,T)$ of a heavy quark-antiquark pair given in
705: terms of the screening form obtained from Debye-H\"uckel theory, the
706: internal energy $U(r,T)$ can now be obtained through the thermodynamic
707: relation (\ref{intern}), and this then provides the binding potential
708: $V(r,T)$ for the temperature dependent version of the Schr\"odinger
709: equation (\ref{schroedinger}). The resulting solution then specifies
710: the temperature-dependence of charmonium binding as based on the
711: correct heavy quark potential \cite{D-K-S}.
712: Let us briefly comment on the thermodynamic
713: basis of this approach. The pressure $P$ of a thermodynamic system is
714: given by the free energy, $P=-F=-U+TS$; it is determined by the kinetic
715: energy $TS$ at temperature $T$ and entropy $S(T)$, reduced by the potential
716: energy $U(T)$ between the constituents. In our case, all quantities give
717: the difference between a thermodynamic system containing a $\Q$ pair
718: and the corresponding system without such a pair. The potential energy
719: of the $\Q$ pair, due both to the attraction of $Q$ and $\bar Q$ and
720: to the modification which the pair causes to the internal energy of
721: the other constituents of the medium, is therefore given by $U$.
722: To determine the dissociation points for the different quarkonium
723: states, we thus have to solve the Schr\"odinger equation (\ref{schroedinger})
724: with $V(r,T)=U(r,T)$.
725:
726: \medskip
727:
728: From eqs.\ (\ref{screen-gauge}) and (\ref{screen-string}) we
729: get for the $\Q$ potential
730: \be
731: V(r,T)= V(\infty,T) + \tilde V(r,T),
732: \ee
733: with
734: \be
735: V(\infty,T) = c_1{\sigma \over \mu} - \alpha \mu
736: + T{d\mu \over dT}[c_1 {\sigma \over \mu^2} + \alpha],
737: \ee
738: and $c_1=\Gamma(1/4)/2^{3/2} \Gamma(3/4)$, and where $\tilde V(r,T)$
739: contains the part of the potential which vanishes for $r\to \infty$.
740: The behaviour of $V(\infty,T)$ as function of the temperature is
741: shown in Fig.\ \ref{vinf}. It measures (twice) the energy of the cloud
742: of light quarks and gluons around an isolated heavy quark, of an extension
743: determined by the screening radius, relative to the energy contained in
744: such a cloud of the same size without a heavy quark. This energy difference
745: arises from the interaction of the heavy quark with light quarks and gluons
746: of the medium, and from the modification of the interaction between the light
747: constituents themselves, caused by the presence of the heavy charge. --
748: The behaviour of $\tilde V(r,T)$ is shown in Fig.\ \ref{vr} for three
749: different values of the temperature. It is seen that with increasing $T$,
750: screening reduces the range of the potential.
751:
752: \begin{figure}[h]
753: %\vskip0.5cm
754: \begin{minipage}[t]{7cm}
755: %\hskip0.1cm
756: \epsfig{file=vinf.eps,width=7.5cm}
757: %\vskip-0.1cm
758: \caption{The large distance limit of the quarkonium potential $V(r,T)$
759: \protect\cite{D-K-S}}
760: \label{vinf}
761: \end{minipage}
762: \hspace{1.3cm}
763: \begin{minipage}[t]{7cm}
764: \vspace*{-5.3cm}
765: \hskip-0.3cm
766: \epsfig{file=ovr.eps,width=7.5cm}
767: \caption{Variation of $\tilde V(r,T)$ with $r$ for different $T$
768: \protect\cite{D-K-S}}
769: \label{vr}
770: \end{minipage}
771: \end{figure}
772:
773: \medskip
774:
775: The relevant Schr\"odinger equation now becomes
776: \be
777: \left\{{1\over m_c}\nabla^2 - \tilde V(r,T)\right\} \Phi_i(r) =
778: \Delta E_i(T) \Phi_i(r)
779: \label{T-schroedinger}
780: \ee
781: where
782: \be
783: \Delta E_i(T) = V(\infty,T) - M_i - 2m_c
784: \ee
785: is the binding energy of charmonium state $i$ at temperature $T$.
786: When it vanishes, the bound state $i$ no longer exists, so that
787: $\Delta E_i(T)=0 $ determines the dissociation temperature $T_i$
788: for that state. The temperature enters only through the $T$-dependence
789: of the screening mass $\mu(T)$, as obtained from the analysis of the
790: lattice results for $F(r,T)$. -- Equivalently, the divergence of the
791: binding radii, $r_i(T) \to \infty$, can be used to define the
792: dissociation points $T_i$.
793:
794: \medskip
795:
796: The same formalism, with $m_b=4.65$ GeV replacing $m_c$, leads to the
797: bottomonium dissociation points. The combined quarkonium results are
798: listed in Table 2. They agree quite well with those obtained in a
799: very similar study based on corresponding free energies obtained
800: in quenched lattice QCD \cite{Wong}, indicating that above $T_c$ gluonic
801: effects dominate. Using a parametrically generalized screened Coulomb
802: potential obtained from lattice QCD results also leads to very compatible
803: results for the $N_f=2$ and quenched cases\cite{Alberico}. We recall
804: here that the main underlying change, which is responsible for the much
805: higher dissociation temperatures for the quarkonium ground states, is the
806: use of the full internal energy (\ref{intern}), including the entropy term,
807: as potential in the Schr\"odinger equation: this makes the binding much
808: stronger.
809:
810: \medskip
811:
812: We should note, however, that in all such potential studies
813: it is not so clear what binding energies of less than a few MeV
814: or bound state radii of several fermi can mean in a medium whose
815: temperature is above 200 MeV and which leads to screening radii
816: of less than 0.5 fm. In such a situation, thermal activation \cite{K-ML-S}
817: can easily dissciate the bound state.
818:
819: \medskip
820:
821: \begin{center}
822: \renewcommand{\arraystretch}{2.0}
823: \begin{tabular}{|c||c|c|c||c|c|c|c|c|}
824: \hline
825: state & J/$\psi(1S)$ & $\chi_c$(1P) & $\psi^\prime(2S)$&$\Upsilon(1S)$&
826: $\chi_b(1P)$&$\Upsilon(2S)$&$\chi_b(2P)$&$\Upsilon(3S)$\\
827: \hline
828: \hline
829: $T_d/T_c$ & 2.10 & 1.16 & 1.12 & $>4.0$ & 1.76 & 1.60& 1.19 & 1.17 \\
830: \hline
831: \end{tabular}\end{center}
832:
833: \medskip
834:
835: \centerline{Table 2: Quarkonium Dissociation Temperatures \cite{D-K-S}}
836:
837: \vskip0.5cm
838:
839: \noindent{\large \bf 5.\ Charmonium Correlators}
840:
841: \bigskip
842:
843: The direct spectral analysis of charmonia in finite temperature lattice
844: has come within reach only in very recent years \cite{lattice-charm}.
845: It is possible now to evaluate the correlation functions $G_H(\t,T)$ for
846: hadronic quantum number channels $H$ in terms of the Euclidean time
847: $\t$ and the temperature $T$. These correlation functions are directly
848: related to the corresponding spectral function $\sigma_H(M,T)$,
849: which describe the distribution in mass $M$ at temperature $T$ for
850: the channel in question. In Fig.\ \ref{spec}, schematic results at different
851: temperatures are shown for the \J~and the \X~channels. It is seen that
852: the spectrum for the ground state \J~ remains essentially unchanged
853: even at $1.5~T_c$. At $3~T_c$, however, it has disappeared; the
854: remaining spectrum is that of the $\C$ continuum of \J~quantum numbers
855: at that temperature. In contrast, the \X~is already absent at $1.1~T_c$,
856: with only the corresponding continuum present.
857:
858: \medskip
859:
860: \begin{figure}[htb]
861: \centerline{\epsfig{file=c-spectral.eps,width=9cm}}
862: \caption{\J~and \X~spectral functions at different temperatures}
863: \label{spec}
864: \end{figure}
865:
866: These results are clearly very promising: they show that in a foreseeable
867: future, the dissociation parameters of quarkonia can be determined
868: {\sl ab initio} in lattice QCD. For the moment, however, they remain
869: indicative only, since the underlying calculations were generally
870: performed in quenched QCD, i.e., without dynamical quark loops. Since
871: such loops are crucial in the break-up of quarkonia into light-heavy mesons,
872: final results require calculations in full QCD. Some first calculations
873: in two-flavour QCD have just appeared \cite{skullerud} and support
874: the late dissociation of the \J. The widths of the
875: observed spectral signals are at present determined by the precision of
876: the lattice calculations; to study the actual physical widths, much
877: higher precision is needed. Finally, one has so far only first signals
878: at a few selected points; a temperature scan also requires higher performance
879: computational facilities. Since the next generation of computers, in the
880: multi-Teraflops range, is presently going into operation, the next
881: years should bring the desired results. So far, in view of the mentioned
882: uncertainties in both approaches, the results from direct lattice studies
883: and those from the potential model calculations of the previous section
884: appear quite compatible.
885:
886: \vskip0.5cm
887:
888: \noindent{\large \bf 6.\ Conclusions}
889:
890: \bigskip
891:
892: The theoretical analysis of the in-medium behaviour of quarkonia has
893: greatly advanced in the past decade. Potential model studies based on
894: lattice results for the colour-singlet free energy appear to be converging
895: to results from direct lattice calculations, and within a few more
896: years, the dissociation temperatures for the different quarkonium
897: states should be known precisely. Through corresponding calculations
898: of the QCD equation of state, these temperatures provide the energy
899: density values at which the dissociation occurs. In statistical QCD,
900: quarkonia thus allow a spectral analysis of the quark-gluon plasma.
901: as many of the observed features as possible.
902:
903: \medskip
904:
905: Finally, it seems worthwhile to note that experimental measurements of
906: the relative dissociation points of the different quarkonium states could
907: be possible in the study of high energy nucleus-nucleus collisions. If
908: they succeed, this might in fact be a unique chance to
909: compare quantitative {\sl ab initio} QCD predictions directly to data.
910:
911: \vskip1cm
912:
913: \centerline{\large \bf Acknowledgements}
914:
915: \bigskip
916:
917: It is a pleasure to thank S.\ Digal, O.\ Kaczmarek,
918: F.\ Karsch, D.\ Kharzeev, P.\ Petreczky,
919: M.\ Nardi and R.\ Vogt for stimulating and helpful comments.
920:
921: \begin{thebibliography}{99}
922:
923: \bibitem{Ting} J.\ J.\ Aubert et al., \PRL 33 (1974) 1404;\\
924: J.\ E.\ Augustin et al., \PRL 33 (1974) 1406.
925:
926: \bibitem{Leder} S.\ W.\ Herb et al., \PRL 39 (1977) 252.
927:
928: \bibitem{nora} For a recent survey, see\\
929: N.\ Brambilla et al., {\sl Heavy Quarkonium Physics},
930: Yellow Report CERN-2005-005.
931:
932: \bibitem{DG} For a recent survey, see\\
933: A.\ Di Giacomo et al., Phys.\ Rep.\ 372 (2002) 319.
934:
935: \bibitem{A-DG} D.\ Antonov, L.\ Del Debbio and A.\ Di Giacomo, JHEP
936: 0308 (2003) 011;\\
937: D.\ Antonov and A.\ Di Giacomo, JHEP 0503 (2005) 017.
938:
939: \bibitem{M-S} T.\ Matsui and H.\ Satz, \PL B 178 (1986) 416.
940:
941: \bibitem{Cornell} E.\ Eichten et al., \PR D17 (1978) 3090; \PR D 21 (1980)
942: 203.
943:
944: \bibitem{D-K-S} S.\ Digal, F.\ Karsch and H.\ Satz, in preparation;\\
945: for earlier work at $T=0$ , see\\
946: S.\ Jacobs, M.\ G.\ Olsson and C.\ Suchyta,
947: \PR 33 (1986) 3338.
948:
949: \bibitem{miya} H.\ Miyazawa, \PR D 20 (1979) 2953;\\
950: V.\ Goloviznin and H.\ Satz, Yad.\ Fiz.\ 60N3 (1997) 523
951:
952: \bibitem{K-Z} O.\ Kaczmarek et al., Prog.\ Theor.\ Phys.\ Suppl.\ 153
953: (2004) 287;\\
954: O.\ Kaczmarek and F.\ Zantow, hep-lat/0503017.
955:
956: \bibitem{schlad} F.\ Karsch, Lect.\ Notes Phys.\ 583 (2002) 209.
957:
958: \bibitem{K-M-S} F.\ Karsch, M.-T.\ Mehr and H.\ Satz, \ZP C 37 (1988) 617
959:
960: \bibitem{Dixit} V.\ V.\ Dixit, Mod.\ Phys.\ Lett.\ A5 (1990) 227.
961:
962: \bibitem{D-K-K-S} S.\ Digal et al., \EPJ C 43 (2005) 71.
963:
964: \bibitem{D-P-S1} S.\ Digal, P.\ Petreczky and H.\ Satz, \PL B 514 (2001) 57.
965:
966: \bibitem{Landau} L.\ D.\ Landau and E.\ M.\ Lifshitz,
967: {\sl Statistical Physics}, Pergamon Press, London, 1958.
968:
969: \bibitem{OK} O.\ Kaczmarek and F.\ Zantow, \EPJ C 43 (2005) 63.
970:
971: \bibitem{Wong} C.-Y.\ Wong, hep-ph/0408020
972:
973: \bibitem{Alberico} W.\ Alberico et al., hep-ph/0507084;\\
974: C.-Y.\ Wong, hep-ph/0509088.
975:
976: \bibitem{K-ML-S} D.\ Kharzeev, L.\ D.\ McLerran and H.\ Satz, \PL
977: B 356 (1995) 349.
978:
979: \bibitem{lattice-charm} T.\ Umeda et al., Int.\ J.\ Mod.\ Phys.\ A16 (2001)
980: 2215;\\
981: M.\ Asakawa and T.\ Hatsuda, \PRL 92 (2004);\\
982: S.\ Datta et al., \PR D 69 (2004) 094507;\\
983: H.\ Iida et al., hep-lat/0509129
984:
985: \bibitem{skullerud} R.\ Morrin et al., hep-lat/0509115.
986:
987: \end{thebibliography}
988:
989: \end{document}
990:
991: