0708.1293/mu.tex
1: \section{Lattice QCD at finite density}
2: \label{sec:mu}
3: 
4: Since 2001, significant progress was made towards simulating QCD with realistic parameter values
5: at small baryon densities. This is achieved by a number of methods that circumvent the sign problem,
6: rather than solving it: i) Multi-parameter reweighting, 
7: ii) Taylor expansion around $\mu=0$ and iii) simulations at imaginary 
8: chemical potential, either followed by analytic continuation or Fourier transformed to the canonical ensemble. 
9: It is important to realize that all of these approaches introduce some degree of 
10: approximation. However, their systematic errors are rather different, thus allowing for powerful crosschecks.  
11: All methods are found to be reliable as long as $\mu/T\lsim 1$, or $\mu_B\lsim 550$ MeV, which includes the region of interest for heavy ion collisions. 
12: Reviews specialized on this subject can be found in \cite{oprev,csrev}. 
13: 
14: \subsection{Two parameter reweighting}
15: 
16: Significant progress enabling finite density simulations was made a few years ago, by a generalization of the Glasgow method \cite{gla} to reweighting in two parameters \cite{fk0}.  
17: The partition function is rewritten identically as
18: \be 
19: Z=\left \langle \frac{\ex^{-S_g(\beta)}\det(M(\mu))}
20: {\ex^{-S_g(\beta_0)}\det(M(\mu=0))}\right\rangle_{\mu=0,\beta_0},
21: \ee
22: where the ensemble average is now generated at $\mu=0$ and a lattice gauge coupling 
23: $\beta_0$, while a reweighting factor takes us to the values
24: $\mu,\beta$ of interest. 
25: The original Glasgow method \cite{gla} reweighted in $\mu$ only and was suffering from the
26: overlap problem: while the reweighting formula is exact, its Monte Carlo evaluation is not. The integral
27: gets approximated by a finite number of the most dominant configurations, which are different for the reweighted and the original ensemble, and this difference grows with $\mu$.
28: When searching for a phase transition at some $\mu$, one-parameter reweighting uses a non-critical ensemble at $\mu=0$, thus missing important dynamics.
29: By contrast, two-parameter reweighting uses an ensemble generated at the pseudo-critical coupling 
30: $\beta_c(\mu=0)$, which is then reweighted along the pseudo-critical line of the phase change.  Thus one is working with an ensemble that probes both phases, improving the overlap with
31: the physical ensemble.
32:  
33: A difficulty in this approach is that the determinant
34: needs to be evaluated exactly, which is costly.  Moreover, because of the sign problem the reweighting factor is exponentially suppressed with volume and chemical potential, thus limiting the applicability to presently moderate values of those parameters. 
35: Moreover, since the statistical fluctuations are those of the simulated ensemble instead of the physical one,  all reweighted measurements are correlated and it is difficult to obtain reliable error estimates. For a proposed procedure see \cite{fk3}.  A further problem arises with staggered quarks, where the
36: root of the determinant has to be taken. For $\mu\neq 0$ this may enhance cut-off effects
37: to $O(a)$ instead of  $O(a^2)$ for $\mu=0$ \cite{gss}.
38: 
39: \subsection{Finite density by Taylor expansion}
40: 
41: Another way to gain information about non-zero $\mu$ is to compute the coefficients of a Taylor series expansion of observables in powers of $\mu/T$. 
42: Early attempts have looked at susceptibilities and the response of screening 
43: masses to chemical potential \cite{milc0,taro,hlp,gg}. 
44: More recently it has also been used to gain information on the phase transition 
45: and its nature itself \cite{bisw1,bisw11,bisw2,bisw3,ggpd}.
46: This idea exploits the fact that on finite volumes the partition function 
47: $Z(m>0,\mu,T)$ is an analytic function of the parameters of the theory.
48: For small enough $\mu/T$ one may then hope to get away with only a few terms, whose coefficients are calculated at $\mu=0$. 
49: Moreover, because of CP symmetry of the QCD action the partition function is even in $\mu$, 
50: $Z(\mu)=Z(-\mu)$, such that physical observables have series expansions in $(\mu/T)^2$. Thus, 
51: {\it e.g.}~the pressure density can be expressed as an even power series, 
52: \be
53:  p(T,\mu)=-\,\frac{F}{V} 
54: = \left(\frac TV\right)\log Z(T,\mu),\quad
55: {p\over T^4}=
56: \sum_{n=0}^\infty c_{2n}(T) \left({\mu\over T}\right)^{2n}.
57: \label{press}
58: \ee
59: The coefficients are equivalent to generalized quark number susceptibilities at $\mu=0$ and measurable with standard simulation techniques.
60: Since all the $\mu$-dependence of the partition function is in the fermion determinant, its derivatives  need to be computed,
61: \be
62: \frac{\partial \ln \det M}{\partial \mu}  ={\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu} \right),\quad
63: \frac{\partial {\rm tr} M^{-1}}{\partial \mu} =
64: %\hspace{-4mm} = \hspace{-4mm} 
65: - {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
66:  M^{-1} \right), \quad \mbox{etc.},
67:  \ee
68: and iterations for higher orders. These expressions become increasingly complex and methods
69: to automatize their generation have been devised \cite{ggpd}.
70: Note that one now is dealing with traces of composite local operators, which greatly facilitates their numerical evaluation by statistical estimators compared to a computation of the full determinant as 
71: required for reweighting.
72: 
73: For high enough temperatures $T\gsim T_c$, the scale of the finite 
74: temperature problem is set by the 
75: Matsubara mode $\sim \pi T$, and one would expect coefficients of order one for an expansion in 
76: the `natural' parameter $\mu/(\pi T)$ \cite{fp2}. This is indeed borne out by numerical simulations
77: and explains the good convergence properties observed for $\mu/T\lsim 1$.
78: 
79: \subsection{QCD at imaginary $\mu$}
80: 
81: The hermiticity relation eq.~(\ref{gamma5}) tells us that the QCD fermion 
82: determinant with imaginary $\mu=i\mu_i$ is real positive, hence 
83: it can be simulated just as for $\mu=0$. 
84: It is then natural to ask whether such simulations can be exploited to learn 
85: something about physics at real $\mu$. 
86: The strategy to get back to real $\mu$ is to fit the full Monte Carlo results 
87: to a truncated Taylor series in $\mu_i/T$. In case of apparent convergence it is 
88: then easy to analytically continue the power series to real $\mu$. 
89: This idea was first used for observables like the chiral condensate and screening 
90: masses in the deconfined phase \cite{lom,hlp}. It was then shown to be 
91: applicable to the phase transition itself \cite{fp1}, which has recently been 
92: exploited in a growing number of works \cite{fp2,fp3,el1,el2,az1,cl,luo1}.
93: 
94: For complex $\mu=\mu_r+i\mu_i$, the QCD partition function eq.~(\ref{part})
95: is periodic in the imaginary direction, with period $2\pi/N_c$ for $N_c$ 
96: colours \cite{rw}. Hence, in addition to being even in $\mu$, 
97: the partition function has the additional symmetry $Z(\mu_r/T,\mu_i/T)=Z(\mu_r/T,\mu_i/T+2\pi/3)$. 
98: Because of the anti-periodic boundary conditions on fermions, this symmetry implies that a shift in $\mu_i$ by certain critical values is equivalent to a transformation by the $Z(3)$ centre of the gauge group. Thus, there are $Z(3)$ transitions between neighbouring centre sectors for all 
99: $(\mu_i/T)_c=\frac{2\pi}{3} \left(n+\frac{1}{2}\right), n=0,\pm1,\pm2,...$. It has been numerically verified that these transitions
100: are first order for high temperatures and a smooth crossover for low temperatures \cite{fp1,el1}. This limits the radius of convergence for analytic continuation, which is given by the first such transition, or $\mu/T=\pi/3$. Hence the approach is limited to $|\mu/T|\lsim 1$. 
101: 
102: Within this circle, the strategy then is
103: to measure expectation values of observables at imaginary $\mu$ and fit them by a Taylor series,
104: \be
105:  \langle O \rangle = \sum_n^N c_n \left(\frac{\mu_i}{\pi T}\right)^{2n}.
106:  \ee
107:  Working at imaginary $\mu$ has a couple of technical advantages. It is computationally simple and much cheaper than reweighting or computing coefficients of the Taylor expansion. Moreover, both parameters $\beta,\mu$ are varied and thus one obtains information from statistically independent ensembles. It also offers some control on the systematics by allowing a judgement on the convergence of the fits.
108: Furthermore, it is a good testing ground for effective QCD models: analytic results can always be continued to imaginary $\mu$ and be compared with the numerics there, as demonstrated for several examples in \cite{el2}.
109: 
110: A different approach making use of imaginary chemical potential is to employ 
111: the canonical ensemble at fixed quark number, which is related to the grand
112: canonical ensemble via the integral transform
113: \be
114: Z_C(T,B=3Q)=\frac{1}{2\pi}\int_{-\pi}^{\pi}d\left(\frac{\mu_i}{T}\right)\;\exp{(-i\mu_iQ/T)}\;
115: Z(\mu=i\mu_i,T,V).
116: \ee
117: One may then compute the grand canonical partition function at imaginary $\mu$
118: and perform the Fourier transform numerically \cite{kap}. This will only 
119: work for moderate $Q$ and small volumes, as the sign problem is reintroduced by the 
120: oscillations of the exponential. For small lattices $6^3\times 4,4^4$, 
121: first interesting results on the phase diagram have been obtained, 
122: both for staggered \cite{slavo} and Wilson
123: fermions \cite{andre}. However, this approach has no overlap problem, and one
124: might hope to push to larger chemical potentials once computational resources 
125: are available.
126: 
127: \subsection{Plasma properties at finite density}
128: 
129: Having developed computational tools for finite density, 
130: one can repeat the studies discussed in the
131: previous sections and see how finite baryon densities affect the screening 
132: masses \cite{milc,taro,hlp,gg}, the equation of state \cite{bisw2,bisw3,fk4} 
133: or the static potential \cite{fk5}. In all those cases the influence of the 
134: chemical potential is found to be rather weak, and for lack of space we will not further discuss 
135: those calculations here.
136: We likewise have to pass over the work done on certain modifications of QCD, 
137: for which the sign problem is manageable or absent, 
138: such as QCD in the static limit, two-colour QCD or QCD at finite isospin.
139: Such studies give important qualitative insights and some are covered in 
140: previous reviews \cite{latrevs,oprev}.   
141: Instead we concentrate here on calculations of the QCD phase diagram at finite 
142: density, where the order of the phase transition is expected to change  
143: as $\mu$ is increased, fig.~\ref{fig:1schem} (left).
144: 
145: \subsection{The critical temperature at finite density}
146: 
147: As in the case of zero density, let us first discuss the phase boundary, 
148: {\it i.e.}~the (pseudo-)critical 
149: temperature $T_c(\mu)$, before dealing with the order of the phase transition. 
150: The critical line has been calculated for a variety of flavours and 
151: quark masses using different methods. For a quantitative comparison one 
152: needs data at one fixed parameter set and also eliminate the uncertainties of 
153: setting the scale. Such a comparison is shown
154: for the critical coupling in fig.~\ref{fig:tccomp} (left), 
155: for $N_f=4$ staggered quarks with the same action and quark mass $m/T\approx 0.2$.
156: (For that quark mass the transition is first order
157: along the entire curve).
158: One observes quantitative 
159: agreement up to $\mu/T\approx 1.3$, after which the different results start 
160: to scatter. This vindicates our earlier statement that all methods appear 
161: to be reliable for $\mu/T\lsim 1$. 
162: Note that strong coupling results at $\beta=0$ 
163: predict $a\mu_c(\beta=0)=0.35-0.43$ \cite{kawa}, requiring the line to bend down
164: rapidly, and thus favour the data from the canonical approach.
165: 
166: 
167: \begin{figure}[t!]
168: \vspace*{-1cm}
169: \begin{minipage}{0.5\textwidth} 
170: \includegraphics[angle=-90,width=7cm]{pdplus.eps}
171: %{\rotatebox{0}{\scalebox{0.4}{\includegraphics{phase_diag.eps}}}}
172: %\caption{Left: Comparison of different methods to compute the critical couplings. From\cite{}.}
173: \end{minipage}
174: \begin{minipage}{0.5\textwidth} 
175: \includegraphics[width=7cm]{phase_diag.eps}
176: \end{minipage}
177: \vspace*{-1cm}
178: \caption[]{Left: Comparison of different methods to compute the critical couplings
179: \cite{slavo}.
180: Right: The phase diagram for physical quark masses as predicted by the two parameter reweighting method \cite{fk2}.}
181: \label{fig:tccomp}
182: \end{figure} 
183: 
184: The case of physical quark masses, after conversion to continuum units, 
185: is shown in fig.~\ref{fig:tccomp} (right) \cite{fk2}.
186: One observes that the critical temperature is decreasing only very 
187: slowly with $\mu$. This is consistent with a description by a 
188: Taylor expansion in powers of $(\mu/\pi T)^2$ with
189: coefficients of order one,
190: \begin{equation}
191: \frac{T_c(\mu)}{T_c(0)}=1-t_2(N_f,m_f)\left(\frac{\mu}{\pi T}\right)^2
192: +{\mathcal O}\left( \left( \frac{\mu}{\pi T}\right)^4 \right) \quad .
193: \label{texp}
194: \end{equation}   
195: The leading coefficients for various cases have been collected from the literature in \cite{csrev} and are
196: reproduced in table \ref{tab:tccomp}. The curvature gets stronger with increasing $N_f$, which is 
197: consistent with the observation at zero density that $T_c$ is lowered by light flavours, cf.~sec.~\ref{sec:tc}. Subleading coefficients are also beginning to emerge but at present not statistically significant yet. 
198: It has been noted that the convergence should speed up when constructing Pad{\'e}
199: approximants for the series, which also tends to increase the curvature 
200: towards larger $\mu$ \cite{el2,pade}.
201: Finally one should note that the continuum conversions relying on the two-loop beta function 
202: are certainly not reliable for these coarse lattices, while fits to 
203: non-perturbative beta functions tend to also increase the curvature.
204: %
205: \begin{table}
206: \begin{center}
207: \begin{tabular}{cccccccc}\hline \hline
208: $N_f$&$am$&$N_s$&$t_2$&Action&$\beta$-Function&Method&Reference    \\ \hline
209: 2  &0.1   & 16      &0.69(35) &p4   &non-pert.   &Taylor+Rew.& \cite{bisw1} \\
210:    &0.032 & 6,8     &0.500(54)&stag.&2-loop pert.&Imag.      & \cite{fp1}\\
211: 3  &0.1   & 16      &0.247(59)&p4   &non-pert.   &Taylor+Rew.& \cite{bisw11}\\
212:    &0.026 & 8,12,16 &0.667(6) &stag.&2-loop pert.&Imag.      & \cite{fp3}\\
213:    &0.005 & 16      &1.13(45) &p4   &non-pert.   &Taylor+Rew.& \cite{bisw11}\\
214: 4  &0.05  & 16      &1.86(2)  &stag.&2-loop pert.&Imag.      & \cite{el1}\\ \hline
215: 2+1&0.0092,0.25&6-12&0.284(9) &stag.&non-pert.   &Rew.       & \cite{fk2}\\
216: \hline \hline
217: \end{tabular}
218: \caption{Coefficient $t_2$ in the Taylor
219: expansion of the transition line, eq.~(\ref{texp}). 
220: All results have been obtained with
221: $N_t=4$. \label{tab:tccomp}}
222: \end{center}
223: \end{table}
224: 
225: \subsection{The QCD phase diagram for $\mu\neq 0$ and the critical point}
226: 
227: As in the case of $\mu=0$, a determination of the order of the transition, an hence the search for the critical endpoint, is much harder, and we begin by discussing the qualitative picture.
228: If a chemical potential is switched on for the light quarks, there is an additional parameter requiring
229: an additional axis for our phase diagram characterizing the order 
230: of the transition, fig.~\ref{fig:1schem} (right).
231: This is shown in fig.~\ref{fig:2schem}, where the horizontal plane is spanned by the $\mu=0$ phase diagram in $m_s, m_{u,d}$ and the vertical axis represents $\mu$. 
232: The critical line separating the first order section from the crossover will now extend to finite $\mu$ and
233: span a surface. A priori it is, of course, not known whether this surface might be leaning to the right or the left, or even have a more complicated behaviour as a function of the quark masses. 
234: %
235: \begin{figure}[t!]
236: \vspace*{-1cm}
237: \centerline{
238: \scalebox{0.65}{\includegraphics{3dphasediag_3.eps}}
239: \scalebox{0.65}{\includegraphics{3dphasediag_42.eps}}
240: }
241: \centerline{
242: \scalebox{0.58}{\includegraphics{2dphasediag5.eps}}
243: \scalebox{0.58}{\includegraphics{2dphasediagX5.eps}}
244: }
245: %\vspace*{-0.5cm}
246: %\vspace*{-0.5cm}
247: \caption{Upper panel: 
248: The chiral critical surface in the case of positive (left) and negative (right) curvature.
249: If the physical point is in the crossover region for $\mu=0$, a finite $\mu$
250: phase transition will only arise in the scenario (left) with positive curvature,
251: where the first-order region expands with $\mu$.
252: Note that for heavy quarks, the first-order region shrinks with $\mu$,
253: cf.~sec.~\ref{sec:potts}.
254: Lower panel: phase diagrams for fixed quark mass (here $N_f=3$) corresponding to 
255: the two scenarios depicted above.}
256: \label{fig:2schem}
257: \end{figure}
258: %\vspace*{-2cm}
259: %
260: However, the expected QCD
261: phase diagram is only obtained if the left situation of fig.~\ref{fig:2schem} 
262: is realized (unless there are additional critical surfaces yet unknown). 
263: The first order region expands as $\mu$ is turned on, so that the
264: physical point, initially in the crossover region, eventually belongs to the
265: critical surface. At that chemical potential $\mu_E$, the transition is second order:
266: that is the QCD critical point. Increasing $\mu$ further makes the transition
267: first order. 
268: A completely different scenario arises if instead the first-order
269: region shrinks as $\mu$ is turned on. In that case the physical
270: point remains in the crossover region for any $\mu$, fig.~\ref{fig:2schem} (right). 
271: 
272: There are then different strategies to determine the QCD phase diagram. 
273: One can fix a particular set of quark masses and for that theory switch on and increase the chemical potential to see whether a critical surface is crossed or not. Such calculations are covered in 
274: sec.~\ref{sec:cp}.
275: Alternatively, sec.~\ref{sec:crit} discusses how to start from the known critical line at $\mu=0$
276: and study its evolution with a finite $\mu$. That will give information on the whole phase diagram
277: in fig.~\ref{fig:2schem}, including, eventually, physical QCD. 
278: 
279: \subsection{Critical point for fixed masses from reweighting and Taylor expansion}
280: \label{sec:cp}
281: 
282: Reweighting methods produced the first finite density phase diagram 
283: from the lattice, 
284: obtained for light quarks corresponding to $m_\pi\sim 300$ MeV \cite{fk1}. 
285: A later simulation at physical quark masses puts the critical point 
286: at $\mu_B^E\sim 360$ MeV \cite{fk2}, fig.~\ref{fk} (left). 
287: In this work $L^3\times 4$ lattices with $L=6-12$
288: were used, working with the standard staggered fermion action. 
289: Quark masses were tuned to $m_{u,d}/T_c\approx 0.037, m_s/T_c\approx 1$, 
290: corresponding to the mass ratios 
291: $m_{\pi}/m_{\rho}\approx 0.19, m_{\pi}/m_K\approx 0.27$, 
292: which are close to their physical values.
293: A Lee-Yang zero analysis \cite{ly} was employed in order to find the 
294: change from crossover behaviour at $\mu=0$ to a first order 
295: transition for $\mu>\mu_E$.
296: This is shown in fig.~\ref{fk}. For a crossover the partition 
297: function has zeroes only off the real axis,
298: whereas for a phase transition the zero moves to the real axis 
299: when extrapolated to infinite volume.  
300: For a critical discussion of the use of Lee-Yang zeros in 
301: combination with reweighting, see \cite{shinji}.
302: %
303: \begin{figure}[t]    
304: \vspace*{-1cm}
305: %{\rotatebox{0}{\scalebox{0.3}{\includegraphics{conlyz244.eps}}}}\hspace*{1cm}
306: %\vspace*{1cm}
307: %{\rotatebox{0}{\scalebox{0.4}{\includegraphics{beta_im_chir.eps}}}}
308: %\includegraphics[height=4cm]{conlyz244.eps}\hspace*{1cm}
309: %\includegraphics[height=5cm]{beta_im_chir.eps}
310: \includegraphics[width=0.44\textwidth]{conlyz244.eps}\hspace*{0.5cm}
311: \includegraphics[width=0.5\textwidth]{beta_im_chir.eps}
312: 
313: %\vspace*{-1.5cm}
314: \caption[]{ Left: Lee-Yang zeroes in the complex $\beta$-plane for SU(3) 
315: pure gauge theory \cite{shinji}.
316: Right: Imaginary part of the Lee-Yang zero closest to the real axis as a 
317: function of chemical potential \cite{fk2}. }
318: \label{fk}
319: \end{figure}
320: %
321: Recently, reweighting has been combined with the density of states 
322: method \cite{dos}, in order to extend the applicability of reweighting 
323: to larger values of $\mu/T$. 
324: First interesting results have been obtained for $N_f=4$,  
325: indicating a new high density phase and a possible triple point, 
326: where the high density transition line meets the deconfinement line \cite{dos}. 
327: Unfortunately, the method is computationally very expensive and so far 
328: limited to coarse and small lattices, so that it is difficult to 
329: unambiguously establish those findings at present.
330: 
331:  
332: In principle the determination of a critical point is also possible 
333: via the Taylor expansion.
334: In this case true phase transitions will be signalled by an 
335: emerging non-analyticity, or a finite radius of convergence for the pressure series about $\mu=0$ as the volume is increased. 
336: The radius of convergence of a power series gives the distance between the expansion point and the nearest singularity, and may be extracted from the high order behaviour of the series. A possible 
337: definition is
338: \be
339: \rho = \lim_{n\rightarrow\infty}\rho_n \qquad \mbox{with}\quad
340:     \rho_n=\left|\frac{c_0}{c_{2n}}\right|^{1/2n}.
341: %     \qquad
342:  %  r_n=\left|\frac{c_{2n}}{c_{2n+2}}\right|^{1/2}.
343: \label{rad}
344: \ee
345: General theorems ensure that if the limit exists and asymptotically all coefficients of the series are positive, then there is a singularity on the real axis. 
346: More details as well as previous applications to strong coupling expansions 
347: in various spin models can be found in \cite{series}.
348: In the series for the pressure such a singularity would correspond to the critical point in the $(\mu,T)$-plane.
349:  
350: A critical endpoint for the $N_f=2$ theory, based on this approach, was reported  in \cite{ggpd}. The
351: authors perfomed simulations on $L^3\times 4$ lattices with $L=8-24$, using the standard staggered action. The quark mass was fixed in physical units to $m/T_c=0.1$. The aim of the simulations was to bracket the critical point by computing the Taylor coefficients of the quark number susceptibility up to sixth order ({\it i.e.}~8th order for the pressure) for various temperatures in the range $T/T_c=0.75-2.15$, and extrapolate to finite $\mu$.
352: This was done for different lattice volumes in order to get an estimate of finite voulme effects.
353: 
354: The results for the convergence radius eq.~(\ref{rad}) are shown in 
355: fig.~\ref{ggfig} (left). A rather strong volume dependence is apparent. 
356: While for the smaller $8^3$ lattice the estimator $\rho_n$ does not 
357: seem to converge to a finite radius of convergence, the results on the larger $24^3$ lattice are consistent with settling at a limiting value. 
358: %
359: \begin{figure}[t]
360: \includegraphics*[height=45mm]{ord1.eps}\hspace*{1cm}
361: \includegraphics*[height=45mm]{chiqO4O6vT.eps}
362: \caption[]{Left: Estimators of the radius of convergence $\rho_n$, 
363: eq.~(\ref{rad}), at $T/T_c=0.95$ on $N_t=4$ lattices. 
364: Circles represent $L=8$, squares $L=24$ \cite{ggpd}.
365: Right: Quark number susceptibility computed through $O(\mu^4)$ (dashed lines) and $O(\mu^6)$ (solid lines) \cite{bisw3}.}
366: \label{ggfig}
367: \end{figure}
368: %
369: Taking the large volume result at face value and extrapolating to all 
370: orders the estimate 
371: for the location of the critical point is $\mu^E_B/T_E=1.1\pm 0.2$ at 
372: $T_E/T_c(\mu=0)=0.95$ \cite{ggpd}.
373: 
374: Another investigation of the two-flavour theory,  also using the Taylor expansion of the pressure,
375: is reported in \cite{bisw2,bisw3}. 
376: This group works with a $16^3\times4$ lattice with p4-improved staggered fermions and a Symanzik-improved Wilson action,  
377: the quark mass is set to $m/T\approx0.4$.  The calculation to order 
378: $\mu^4$ was performed in \cite{bisw2} while results on $\mu^6$ are presented in 
379: \cite{bisw3}. The last work also contains detailed discussions of analytic 
380: calculations to compare with,
381: namely the pressure in high temperature perturbation theory \cite{vuo}, 
382: which is going to hold at very high temperatures, as well as the 
383: hadron resonance gas model, which gives a rather good description of 
384: the pressure in the confined phase \cite{krt}.
385: 
386: In agreement with \cite{ggpd} and qualitative expectations, 
387: their detailed results for the coefficients in the pressure series 
388: satisfy $c_6\ll c_4 \ll c_2$ for $T>T_c$, {\it i.e.}~one would have 
389: coefficients of order one for an expansion in $(\mu/\pi T)$. 
390: An impression of the convergence of the series can be obtained by 
391: looking at the quark number susceptibility calculated to consecutive orders, 
392: as shown in fig.~\ref{ggfig} (right).
393: For $T\lsim 1.2 T_c$, the series seems to converge rapidly and the $\mu^6$-result 
394: is compatible with the one through order $\mu^4$. Around the 
395: transition temperature $T_c$, the $\mu^4$-results show a peak emerging 
396: with growing $\mu/T_c$, which in \cite{bisw2}  was interpreted as evidence for
397: a critical point. However, the $\mu^6$ contribution suggests that in this region results do not yet converge,  and the structure is hence not a significant feature of the full pressure.
398: Furthermore, estimates for the radius of convergence through that order agreed with predictions
399: from the hadron gas model, which however has infinite radius of convergence. Hence the conclusion
400: in \cite{bisw3} that there is no signal for a critical point at that quark mass.
401: 
402: \subsection{The change of the critical line with $\mu$ \label{sec:crit}}
403: 
404: Rather than fixing one set of masses and considering the effects of $\mu$, 
405: one may map out the critical surface in fig.~\ref{fig:2schem} by measuring
406: how the $\mu=0$ critical boundary line changes under the influence of $\mu$.
407: This question was adressed using an imaginary 
408: chemical potential in 
409: \cite{fp2,fp3}. The methodology employed is the same 
410: as in sec.~\ref{sec:m1m2c}, {\it i.e.}~measurement
411: of the Binder cumulant of the chiral condensate. 
412: The results for $N_f=3$ are summarized in fig.~\ref{fig:immu} (left). 
413: The chemical potential is found to have almost no influence on $B_4$ as a function
414: of quark mass. 
415: A lowest-order fit, linear in $a m$ and $(a \mu)^2$, 
416: gives the error band fig.~\ref{fig:immu} (right), 
417: corresponding to
418: \be
419: a m^c(a \mu) = 0.0270(5) - 0.0024(160) (a \mu)^2 \; .
420: \label{c_prime1}
421: \ee
422: 
423: Care must be taken for the conversion to physical units. The crucial
424: point is that the gauge coupling $\beta$ is tuned with changing $\mu$ to 
425: stay on the critical line, so that $a(\beta)$ decreases.
426: Expressing the change of the critical quark mass with chemical potential in lattice
427: and continuum units as
428: \be
429: \label{c1prime}
430: %\frac{a m^c(\mu)}{a m^c(0)} = 1 + c'_1 (a \mu)^2 + ... \\
431: \frac{a m^c(\mu)}{a m^c(0)} = 1 + \frac{c'_1}{a m^c(0)} (a \mu)^2 + ..., \qquad
432: \frac{m^c(\mu)}{m^c(0)} = 1 + c_1 \left( \frac{\mu}{\pi T} \right)^2 + ...
433: \label{c1}
434: \ee
435: then $c_1$ and $c'_1$ are related by
436: \be
437: %c_1 = \frac{\pi^2}{N_t^2} c'_1 + \left( \frac{1}{T_c(m,\mu)} \frac{d T_c(m,\mu)}{d(\mu/\pi T)^2} \right)_{\mu=0}
438: c_1 = \frac{\pi^2}{N_t^2} \frac{c'_1}{a m^c(0)} + \left( \frac{1}{T_c(m,\mu)} \frac{d T_c(m^c(\mu),\mu)}{d(\mu/\pi T)^2} \right)_{\mu=0}.
439: \ee
440: Since $c_1'$ is observed to be nearly zero, it is the second term that dominates, leading
441: to an overall negative coefficient $c_1=-0.7(4)$ \cite{fp3}.
442: \begin{figure}[t!]
443: \centerline{
444: \scalebox{0.62}{\includegraphics{global_cons.eps}}
445: \scalebox{0.62}{\includegraphics{mc_cons.eps}}
446: }
447: \caption{Left: $B_4(am,a\mu_i)$ for $N_f=3$, $N_t=4$ and
448: imaginary chemical potentials.
449: Right: One-sigma error band for the critical mass $a m^c(a \mu_i)$ resulting
450: from a linear fit to the data on the left. From \cite{fp3}.}
451: \label{fig:immu}
452: \end{figure}
453: 
454: This is evidence that, in the $N_f=3$ theory on an $N_t=4$ lattice,
455: the region of first-order transitions {\em shrinks} as a baryon chemical potential
456: is turned on, and the ``exotic scenario'' of fig.~\ref{fig:2schem} (right) is realized.
457: Interestingly, similar qualitative conclusions are obtained from simulations of the three flavour theory
458: with an isospin chemical potential~\cite{DKS_mc}, as well as simulations at imaginary $\mu$
459: employing Wilson fermions \cite{luo1}.
460: 
461: This investigation has also been extended to the case of non-degenerate quarks \cite{fp3}.
462: Fig.~\ref{potts} (left) shows a comparison of the critical line for $\mu=0$  with some critical masses
463: calculated for a sizeable baryon chemical potential. The data show the same trend as for $N_f=3$: the critical mass is nearly unchanged 
464: or slightly increasing with $\mu_i$, in lattice units. 
465: The conversion to physical units
466: proceeds as in the three flavour case and gives
467: a dominant negative contribution to $c_1$.
468: Together with a very small value for $c'_1$, 
469: it implies again that the first-order region shrinks as the baryon chemical
470: potential is turned on, and the ``exotic scenario'' of fig.~\ref{fig:2schem} (right) is the correct one.
471: 
472: \subsection{The heavy quark limit: Potts model}
473: \label{sec:potts}
474: 
475: In light of these surprising results, it is also interesting to consider
476: the heavy quark corner of the schematic phase diagram of fig.~\ref{fig:2schem}.
477: Simulations of quenched QCD at finite baryon number have been done in \cite{quenb}. 
478: As the quark mass goes to infinity, quarks can be integrated out and QCD reduces to a gauge theory of Polyakov lines. First simulations of this theory with Wilson valence quarks can be found in \cite{nucu}.
479: At a second order phase transition, universality allows us to neglect the 
480: details of gauge degrees of freedom, so the theory reduces to the 
481: 3d three-state Potts model, which is in the 3d Ising universality class. 
482: Hence, studying the three-state Potts model should teach us about the behaviour of QCD in the neighbourhood of the critical line separating the quenched first order region from the crossover region.
483: For large $\mu$ the sign problem in this theory has actually been solved by means 
484: of cluster algorithms \cite{clust}. 
485: 
486: \begin{figure}
487: \centerline{
488: {\includegraphics*[width=0.5\textwidth]{m1m2c_mu0.2.eps}}
489: \includegraphics*[width=0.45\textwidth]{potts.eps}
490: }
491: \caption{Left: Comparison of the critical line at $\mu=0$ and $a \mu_I = 0.2$.
492: Right: The critical heavy quark mass separating first order from crossover as a function of $\mu^2$ from the Potts model \cite{Potts}.}
493: \label{potts}
494: \end{figure}
495: 
496: Here we are interested in simulations at small $\mu/T$ \cite{Potts}.  In this case, the sign problem is mild enough for brute force simulations at real $\mu$ to be feasible. 
497: In \cite{Potts}, the change of the critical heavy quark mass is 
498: determined as a function of real as well as imaginary $\mu$, 
499: as shown in fig.~\ref{potts} (right). 
500: Note that $M^c(\mu)$ rises with real chemical potential. {\it i.e.}~the first order region in fig.~\ref{fig:2schem} shrinks as finite baryon density is switched on. This system is thus an example of the non-standard scenario discussed in the previous section. 
501: The calculation also gives some insight in the problem of analytic continuation: while fig.~\ref{potts} clearly endorses the approach in principle, an $O(\mu^8)$-fit was required to reproduce the data
502: on both sides of $\mu^2=0$. Similar accuracy is much more difficult to achieve 
503: around the chiral line with present
504: resources.  
505: 
506: \subsection{Discussion of critical end point results}
507: 
508: The results about the critical surface from sec.~\ref{sec:crit} appear to be in qualitative contradiction 
509: with those of \cite{fk2}, \cite{ggpd}, which both conclude for the existence of a critical point
510: $(\mu_E,T_E)$ at small chemical potential $\mu_E/T_E \lsim 1$.  However,
511: in considering the reasons for such disagreement, one can see that the different data sets
512: are actually not inconsistent with each other, and the differing pictures can be explained
513: by standard systematic effects.
514: 
515: In \cite{ggpd} the critical point was inferred from an estimate of
516: the radius of convergence of the Taylor expansion of the free energy.
517: Regardless of the systematics when only 4 Taylor coefficients
518: are available, the estimate is made for $N_f=2$. 
519: The $(\mu,T)$ phase diagram of this theory might well be qualitatively different
520: from that of $N_f=2+1$ QCD, as illustrated in fig.~\ref{fig:syst} (right). 
521: Its critical endpoint point, obtained 
522: by intersecting a critical surface when going up vertically from the 
523: $N_f=2$ quark mass values,
524: is clearly a long way from the critical endpoint of physical QCD, and nothing follows quantitatively
525: from the value of one for the other.
526: 
527: In \cite{fk2} the double reweighting approach was followed.
528: By construction, this reweighting is performed at a quark mass fixed in lattice 
529: units:
530: $a m_{u,d}=\frac{m_{u,d}}{T_c} = const$. Since the critical temperature 
531: $T_c$ decreases 
532: with $\mu$, so does the quark mass. This decrease of the quark mass pushes
533: the transition towards first order, which might be the reason why a critical
534: point is found at small $\mu$. This effect is illustrated in the 
535: sketch fig.~\ref{fig:syst} (left), where the bent
536: trajectory intersects the critical surface, while the vertical line
537: of constant physics does not.
538: Put another way, \cite{fk2} measures the analogue of $c_1'$ instead of $c_1$
539: in eq.~(\ref{c1}).
540: From fig.~\ref{fk} (right) we see that the distance of the physical theory from
541: criticality stays initially constant, consistent with a coefficient $c'_1\approx 0$ 
542: as that in \cite{fp3}, sec.~\ref{sec:crit}. 
543: Taking the variation $a(\mu)$
544: into account could then make a dominant contribution, which might possibly change the results qualitatively.
545: 
546: Conversely, in \cite{fp3} only the leading order coefficient of $m^c(\mu)$ has been determined
547: from imaginary $\mu$. 
548: It cannot yet be excluded that there are cancelling terms
549: of higher order, alternating in sign. Such contributions would no 
550: longer cancel after continuation to real $\mu$, leading to a different picture.
551: Konwledge of the next term in the series will clarify this.
552: 
553: On the other hand, 
554: a robust finding is the high quark mass sensitivity
555: of the critical point: irrespective of the sign, if 
556: $c_1\sim O(1)$ in eq.~(\ref{c1}), 
557: $m^c(\mu)$ is a slowly varying function of $\mu$, just as the pressure, screening
558: masses or $T_c$. Hence $\mu_E(m)$ is rapidly
559: varying. A change of quark masses by a few percent will then imply
560: a change of $\mu_E$ by $O(100\%)$.
561: One should also remember that most investigations so far have used unimproved 
562: staggered quarks on coarse $N_t=4$ lattices only. This might be worrisome
563: given the exceedingly light quarks involved, cf.~sec.~\ref{latte} and \cite{sharpe,gss}.
564: Finally, a more complicated picture with additional
565: critical surfaces in fig.~\ref{fig:2schem} is yet another possibility.
566: In light of these circumstances even the qualitative features of the QCD phase 
567: diagram cannot be regarded as settled until they are probed with better
568: accuracy on finer lattices.
569: 
570: \begin{figure}[t]
571: \vspace*{-0.5cm}
572: \centerline{
573: \scalebox{0.62}{\includegraphics{fooFK.eps}}
574: \scalebox{0.62}{\includegraphics{fooGG.eps}}
575: }
576: \caption{Left: Effect of keeping the quark mass fixed in lattice units while changing $\mu$, as in \cite{fk2}.
577: Right: Location of critical points for the $N_f=2+1$ and the $N_f=2$ theory, as considered in \cite{ggpd}.
578: Knowing the latter does not predict the former.}
579: \label{fig:syst}
580: \end{figure}
581: 
582: \section{Conclusions}
583: 
584: We have summarized our current understanding of finite temperature and density QCD
585: from numerical studies on the lattice. 
586: While many results in the pure gauge theory are
587: available in the continuum limit, simulations with
588: dynamical fermions still suffer from systematic errors.
589: These are mainly due to the finite lattice spacing, and quark masses
590: which don't meet their physical values yet.
591: For the critical temperature and the equation
592: of state these problems are now being tackled and the first continuum extrapolations
593: become available.
594: 
595: Existing results
596: provide us with a detailed picture
597: of how equilibrium plasma properties change through the deconfinement
598: transition up to a few $T_c$.
599: Combinations of perturbative calculations and numerical
600: simulations have produced insight into the regime of very high
601: temperatures as well as the dynamics and mixing of hard and soft momentum modes.
602: Altogether this provides a quantitative understanding of the relevant
603: length scales in the plasma, as well as tests of the applicability
604: of thermal perturbation theory.
605: 
606: The naive picture of the deconfined phase as a weakly interacting
607: parton gas is not supported. For temperatures relevant to
608: heavy ion collisions, the plasma displays strong residual interactions
609: through soft gluonic modes, which cannot be treated
610: perturbatively, and which influence different
611: quantities in different ways. 
612: This gives a consistent explanation to the
613: various observed features: the equation of state,
614: susceptibilities and fermionic correlators
615: are dominated by hard modes and not far
616: from their ideal gas values, but the corrections are significant.
617: Gluonic correlators, on the other hand,
618: are dominated by soft modes and entirely off their leading perturbative
619: predictions. An ideal gas is established
620: only at asymptotically high temperatures.
621: 
622: Significant progress was made over the last five years regarding 
623: dynamical quantities as well as finite density simulations. 
624: In both fields methods have been developed, that
625: are currently being scrutinized for their reliability. 
626: Calculations of spectral functions provide
627: a picture of remnant binding forces in the plasma phase as well as first semi-quantitative
628: results for the transport coefficients. Calculations at finite density are 
629: now possible for $\mu/T\lsim 1$,
630: with good agreement between all methods whenever equal parameter sets are compared.
631: However, the way to a quantitative understanding of the QCD phase 
632: diagram is still long, mainly due to
633: the high quark mass and cut-off sensitivity of the critical endpoint. 
634: Current developments, also regarding computing speed-up for 
635: different fermion discretizations and improvement programmes,
636: give reason to believe that these questions can be significantly 
637: improved upon in the near future.
638: 
639: