1: \section{Calculations of $T_c$}
2: \label{tccc}
3: The critical temperature for an ideal Bose gas is given by
4: Eq.~(\ref{tc0}). A natural question to ask is: what is
5: the leading-order effect of a weak two-body interaction
6: on the critical temperature of a homogeneous Bose
7: gas? This question has been around for almost fifty years, but only
8: very recently has the issue been settled.
9: It has been discussed in detail by Baym {\it et al.}~\cite{baym2}
10: and we follow to some extent their paper.
11: Various approaches to the problem have also been discussed very recently
12: by Haque~\cite{hakk}.
13:
14: In the following, we assume that the interaction is repulsive, which
15: corresponds to a positive scattering length $a$. One might think
16: that effects of a repulsive interaction is to decrease the
17: critical temperature of Bose gas. For instance, the superfluid transition
18: in liquid $^4$He, takes place at a lower temperature than that of
19: an ideal gas of the same density. However, liquid $^4$He is not
20: weakly interacting and it turns out that the leading effects of
21: interactions in the dilute Bose gas is to increase $T_c$.
22:
23:
24: The first paper in which a
25: quantitative prediction appears, is from 1957 by Lee and Yang~\cite{leeyang}.
26: In that paper, the authors predict that the critical temperature increases
27: compared to that of an ideal Bose gas, and that the increase is proportional
28: to $\sqrt{a}$. Later the same authors predicted that the shift is linear
29: with $a$~\cite{leeyang2}.
30: This prediction was purely qualitative since neither sign nor
31: magnitude were given. A couple of years later,
32: Glassgold {\it et al}.~\cite{glass} also predicted an increase of $T_c$
33: which is proportional to $\sqrt{a}$. A couple of decades later, the
34: problem was revisited by Toyoda~\cite{toyota}.
35: He predicted a decrease of the critical temperature which is
36: proportional to $\sqrt{a}$. Since the sign agrees with the measurements
37: on $^4$He, there seemed to be at least qualitative agreement between theory
38: and experiments.
39: Long ago, Huang claimed an increase in $T_c$ that is proportional
40: $a^{3/2}$~\cite{huangbok}, and very recently he claims
41: in another paper that
42: $T_c$ increases proportional to $\sqrt{a}$~\cite{huang2}.
43: From this selection of papers, it is clear that there has been
44: considerable amount of confusion about how $T_c$ depends parametrically
45: on the scattering length. Common to these results as well
46: other attempts to calculate $T_c$~\cite{schakel,illu}, is that they
47: are based on perturbation theory. However, Bose condensation in
48: a dilute gas is governed by long-distance physics that is
49: inherently nonperturbative. These issues will be discussed later.
50:
51: There have also been other approaches to the calculation of $T_c$.
52: Stoof and Bijlsma have
53: carried out renormalization group calculations
54: of the critical temperature ~\cite{henk}.
55: This approach was discussed in the previous section and it predicts
56: that the leading shift is proportional to $a\log a$.
57: Since this approach is purely numerical, it is difficult to take the
58: limit $a\rightarrow0$ and thus obtain the correct dependence on $a$
59: in the dilute limit. Very recently,
60: a calculation of $T_c$
61: based on the exact renormalization group by
62: Ledowski, Hasselmann and
63: Kopietz~\cite{ledow}. They calculated the momentum-dependent two-point function
64: and showed that the leading behavior is proportional to the scattering length
65: $a$.
66:
67:
68: The first Monte Carlo simulations for hard-sphere bosons
69: in the low-density regime were done by Gr\"uter {\it et al.}~\cite{grut}.
70: In that paper, the authors predict a positive linear shift after extrapolating
71: to the limit $a\rightarrow 0$.
72: This result was somewhat surprising since
73: some early Monte Carlo simulations
74: as well as the experiments on $^4$He, show a decrease in $T_c$ due to
75: repulsive interactions.
76: Later, it has been shown rigorously
77: using effective field theory methods that the parametric dependence of $T_c$
78: indeed is linear in $a$~\cite{baym2}. Thus in the
79: dilute limit, we can write
80: \bqa
81: {\Delta T_c\over T_c^0}&=&cn^{1/3}a\;,
82: \label{proptcc}
83: \eqa
84: where $c$ is a constant that is to be determined.
85: The problem of determining the constant $c$ has been attacked by
86: analytical as well as numerical methods in recent years.
87: These methods include high-precision Monte Carlo simulations,
88: $1/N$ expansions, self-consistens calculations involving
89: summation of bubble and ladder diagrams, and variational perturbation theory.
90: We will discuss these in Secs.~\ref{1n}--\ref{oc}.
91:
92:
93:
94:
95: \subsection{Hartree-Fock Approximation and Breakdown of Perturbation Theory}
96: In this subsection, we will briefly discuss the Hartree-Fock approximation and
97: show that it predicts no shift in the critical temperature. We approach
98: the phase transition from above, so the condensate density $v$ is zero.
99:
100: In an ideal gas, the number
101: density of excited particles is given by Eq.~(\ref{nf}),
102: which can be written as
103: \bqa
104: n_{\rm ex}&=&{1\over\lambda_T^3}g_{3/2}(z),
105: \label{nonn}
106: \eqa
107: where the function $g(z)$ is the polylogarithmic function
108: \bqa
109: g_l(z)&=&\sum_{n=1}^{\infty}{z^n\over n^l}\;,
110: \label{non}
111: \eqa
112: and $z=e^{\beta\mu}$ is the fugacity.
113:
114: The simplest way to include the effects of interactions is to include
115: the self-energy
116: in the Hartree-Fock approximation.
117: The Feynman graph is the tadpole diagram shown in Fig.~\ref{hfe}.
118:
119: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
120: \begin{figure}[htb]
121: \epsfysize=0.95cm
122: \vspace{0.5cm}
123: \epsffile{hf.eps}
124: %\vspace{2mm}
125: \caption[a]{Self-energy diagram in the Hartree-Fock approximation.}
126: \label{hfe}
127: \end{figure}
128: The tadpole is independent of the external momentum and the
129: expression for the Hartree-Fock self-energy is
130: \bqa\nonumber
131: \Sigma_{\rm HF}&=&2g\sumint_{P}
132: {p^2-\mu\over\omega_n^2+(p^2-\mu)^2}
133: \\
134: &=&2gn\;,
135: \eqa
136: where we have used Eqs.~(\ref{nf}) and~(\ref{exs}).
137: Thus a particle with momentum ${\bf p}$ effectively has the energy
138: $\epsilon(p)=p^2+2gn$, where the term $2gn$ arises from the mean field
139: of the other particles. We can now generalize Eq.~(\ref{nonn})
140: by writing
141: \bqa
142: n&=&{1\over\lambda_T^3}g_{3/2}\left(e^{\beta(\mu-2gn)}\right)\;.
143: \label{sc}
144: \eqa
145: Eq.~(\ref{sc}) shows that we must increase the chemical potential
146: by an amount $\Delta\mu=\Sigma_{\rm HF}$ to
147: keep the same number density as that of an ideal gas at the same temperature.
148: It shows in particular that $\mu$ must approach $\Sigma_{\rm HF}$
149: from below to obtain the critical number density at a given temperature.
150: Thus the
151: critical temperature remains the same. The conclusion is that
152: including a constant
153: mean-field shift in the single-particle energies cannot
154: change the critical temperature of a Bose gas.
155: This is an example of the fact that mean-field theories
156: effectively treat interacting gases as ideal gases with
157: modified parameters and thus predict the same $T_c$.
158:
159:
160: Calculations using the Hartree-Fock approximation have been carried out
161: by Huang. He
162: applies a virial expansion to Eq.~(\ref{sc}) and obtains
163: a change in the critical temperature which is proportional to $a^{1/2}$.
164: However, this is an artifact of the approximation, as can be seen
165: by including more terms in the expansion.
166: This was discussed in some detail in~\cite{baym2}.
167: A correct treatment is given
168: in e.g.~\cite{pet}.
169: Similarly, the summation of the ring diagrams~\cite{haug}
170: in the effective potential does not change $T_c$.
171: The reason is that
172: these diagrams are evaluated at zero external momentum
173: and therefore merely corresponds to a
174: redefinition of the chemical potential.
175:
176:
177: We have seen that a leading perturbative calculation in the scattering
178: length $a$
179: gives no corrections
180: to the critical temperature of a dilute Bose gas. One might try to improve
181: on this result by going to higher orders in perturbation theory.
182: The Feynman diagrams contributing to the self-energy at second order in
183: perturbation theory are shown in Fig.~\ref{ptself}.
184:
185: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
186: \begin{figure}[htb]
187: \epsfysize=2.2cm
188: \vspace{0.5cm}
189: \epsffile{pt.eps}
190: %\vspace{2mm}
191: \caption[a]{Two-loop self-energy diagrams.}
192: \label{ptself}
193: \end{figure}
194:
195: If we focus on the
196: contribution from the $n=0$ Matsubara mode,
197: the diagrams read
198: \bqa\nonumber
199: \Sigma^{2a}(0,{\bf p})
200: &=&-4g^2T^2
201: \int{d^dk\over(2\pi)^d}\int{d^dq\over(2\pi)^d}
202: \\ &&\times
203: {1\over(k^2-\mu)
204: (q^2-\mu)^2}\;,\\ \nonumber
205: \Sigma^{2b}(0,{\bf p})
206: &=&-2g^2T^2
207: \int{d^dk\over(2\pi)^d}\int{d^dq\over(2\pi)^d}
208: \\ &&\hspace{-0.4cm}\times
209: {1\over(k^2-\mu)(q^2-\mu)(|{\bf p}+{\bf q}+{\bf k}|^2-\mu)}\;,
210: \eqa
211: where the chemical potential acts
212: as an infrared cutoff in the integral.
213: The left diagram is independent of the external momentum.
214: If we use the mean-field criterion that
215: $\mu\rightarrow0$ at the transition, the integral is linearly
216: divergent in the infrared.
217: The right diagram depends on the external momentum
218: ${\bf k}$. For $\mu=0$, it
219: is logarithmically divergent as the external momentum
220: goes to zero. As one goes to higher orders in the perturbation expansion,
221: the diagrams become increasingly infrared divergent for $\mu=0$.
222: If we denote a generic self-energy
223: diagram with $n$ loops by $\Sigma_n$, we have~\cite{baym2}
224: \bqa
225: \Sigma_n&\sim&T\left({a\over\lambda_T}\right)^2
226: \left({a^2\over\mu\lambda_T^4}\right)^{n-2\over2}\;.
227: \eqa
228: This shows that perturbation theory breaks
229: down in the critical region
230: due to the infrared
231: divergences.
232: Physically, these infrared divergences are screened and this can be
233: taken into account by summing certain classes of diagrams from all orders
234: of perturbation theory. Examples of this is summation of bubble or ladder
235: diagrams. We return to this issue at the end of this section.
236:
237:
238:
239:
240: \subsection{Dimensional Reduction}
241: In Sec.~\ref{rgapp},
242: we saw that the renormalization-group equations at high temperature
243: reduce to those of a three-dimensional $O(2)$-symmetric theory.
244: This is an example of dimensional reduction and for the dilute Bose
245: gas, it can be understood as follows.
246: In the imaginary time formalism, the fields are decomposed into
247: modes which are characterized by their Matsubara frequency $\omega_n=2\pi nT$.
248: At distances much larger than the thermal wavelength
249: and for temperatures sufficiently close to
250: the critical temperature, the time derivative term for $n\neq0$
251: is much larger than both the kinetic energy term and the effective chemical
252: potential term.
253: This implies that the nonstatic Matsubara modes
254: decouple and the long-distance physics can be described in terms of an
255: effective
256: three-dimensional field theory for the $n=0$ mode. %~\footnote
257: The fact that the long-distance physics associated with the
258: phase transition is well separated from the typical momentum scale $T$
259: associated with the nonzero Matsubara modes, makes effective field theory
260: methods ideal to study the phase transition.
261:
262:
263: %The condition that a gas is dilute is that the $s$-wave scattering length
264: %is small compared to the interparticle spacing $n^{-1/3}$. Close to
265: %the phase transition, the de Broglie wavelength
266: %$\lambda_T=2\sqrt{\pi/T}$ is of the order of the interparticle spacing.
267: %Thus for a dilute gas, we have $a\ll\lambda_T$.
268:
269: The effective three-dimensional theory
270: that describes the long-distance physics can be constructed
271: using the methods of effective field theory~\cite{eft}.
272: Once the symmetries of the theory have been identified, one writes down the
273: most general Lagrangian ${\cal L}_{\rm eff}$
274: that is consistent with these symmetries. In the present
275: case, we simply have a
276: complex scalar field with an $O(2)$-symmetry.
277: In addition, there a is three-dimensional rotational symmetry.
278: The effective three-dimensional theory is then described by the action
279: \bqa\nonumber
280: S_{\rm eff}&=&\int d^3x\bigg[
281: -{1\over2}\phi^*\nabla^2\phi+{1\over2}\mu_3\phi^*\phi
282: \\ &&
283: \hspace{3.5cm}
284: +{1\over24}u(\phi^*\phi)^2+...\bigg]\;,
285: \label{3eff}
286: \eqa
287: where we have used the conventional normalization of an $O(2)$-invariant
288: theory.
289: The dots indicate operators with more derivatives and more fields. Examples
290: are $\left[\nabla(\phi^*\phi)\right]^2$ and $(\phi^*\phi)^3$.
291: The relation between the parameters in the effective theory and in the
292: full theory can be determined by perturbative
293: matching; one requires that the effective
294: theory~(\ref{3eff}) reproduces static correlators at long distances
295: $R\gg1/T$ to a specified accuracy.
296: The reason why the parameters of the effective three-dimensional theory
297: can be determined in perturbation theory, stems from the fact
298: that the coefficients
299: of the effective theory encode the short-distance physics
300: at the scale $T$ which is perturbative and that the
301: matching procedure does not involve the nonperturbative long-distance
302: physics.
303:
304:
305:
306: At the tree level, the matching can be done simply by inspection.
307: By comparing the action~(\ref{s1})
308: that describes the full four-dimensional theory with the
309: action~(\ref{3eff}) that describes the effective three-dimensional theory,
310: we can read off the relation between the fields in the two theories.
311: This yields
312: \bqa
313: \psi&=&\sqrt{T\over2}\phi
314: \label{fmat}
315: \;,
316: \eqa
317: Similarly, by matching the other terms in the action in the two theories
318: at the tree level and using~(\ref{fmat}), one easily finds
319: \bqa
320: \mu_3&=&-\mu
321: \;,\\
322: u&=&3gT\;,
323: \eqa
324: Higher order operators as well as corrections to the coefficients of the
325: operators in Eq.~(\ref{3eff}) can be
326: ignored at the order of interest
327: in the diluteness expansion.
328: For instance, the coefficient of the operator
329: $\left[\nabla(\phi^*\phi)\right]^2$ is proportional to $a^2\lambda_T$.
330: From dimensional analysis, it follows that
331: the contribution to physical quantities
332: from this operator is then suppressed by a factor
333: of $n^{1/3}a$ compared to the operator $(\phi^*\phi)^2$.
334: The contribution to physical quantities from other operators are analyzed
335: in a similar manner.
336:
337: For an ideal Bose gas, the critical
338: temperature is given by Eq.~(\ref{tc0}). Equivalently, the
339: critical number density $n_c^0$ at fixed temperature
340: satisfies
341: $n_c^0\lambda_T^3=\zeta\left({3\over2}\right)$.
342: Due to the repulsive interactions in the
343: dilute Bose gas, the critical number densitiy changes.
344: The first-order change in the critical temperature
345: $\Delta T_c=T_c-T_c^0$
346: is related to the first-order change $\Delta n_c=n_c-n_c^0$
347: in the critical number density
348: at fixed $T_c$
349: by~\cite{baym2}:
350: \bqa\nonumber
351: {\Delta T_c\over T_c^0}&=&-{2\over3}{\left[n_c(T_c)-n_c^0(T_c)\right]\over n}\\
352: &=&
353: -{1\over3}{T^0\Delta\langle\phi^*\phi\rangle\over n_c^0}\;.
354: \label{delten}
355: \eqa
356: The factor $2/3$ in the first line
357: comes from the relation $T^0\propto (n^0)^{2/3}$.
358: The last equality follows from the fact that $n=\langle\psi^*\psi\rangle$
359: and that the contribution from the the zeroth Matsubara
360: mode is ${1\over2}T\langle\phi^*\phi\rangle$, which
361: follows from Eq.~(\ref{fmat}).
362:
363: In order to calculate the critical temperature, we must evaluate the
364: quantity $\Delta\langle\phi^*\phi\rangle$. We discuss this next.
365:
366: \subsubsection{$1/N$ expansion}
367: \label{1n}
368: The $1/N$ expansion is a nonperturbative method thas has been widely used
369: in high-energy and condensed matter physics~\cite{largen,moshe}.
370: In condensed matter physics, it has been used to
371: study the critical behavior of $O(N)$ spin models and calculate
372: their
373: critical exponents~\cite{zinn}.
374: The idea is to generalize a Lagrangian with a fixed number of fields to
375: $N$ fields and then let $N$ be a variable. The expansion is defined
376: as an expansion in powers of $1/N$ while $gN$ is held fixed ($g$ is
377: the coupling constant). The method is nonperturbative in the sense that
378: calculations at every order in $1/N$, there are Feynman diagrams
379: contributing from all orders of perturbation theory.
380: In this way, one sums up graphs from all orders of
381: perturbation theory. One hopes that this expansion captures some of the
382: essential physics that cannot be captured by e.g. perturbative methods.
383:
384: The critical temperature was recently calculated by Baym, Blaizot, and
385: Zinn-Justin~\cite{baym3} in the large-$N$ limit. The next-to-leading-order
386: result was obtained by
387: Arnold and Tom\'a\^sik~\cite{at}.
388:
389:
390: In the present case, the
391: Lagrangian~(\ref{3eff}) is generalized to a
392: scalar field theory with $N$ real
393: components. The Lagranigan is now $O(N)$ invariant
394: and reads
395: \bqa
396: {\cal L }_{\rm eff}&=&-{1\over2}\phi_i\nabla^2\phi_i
397: +{1\over2}\mu_3\phi_i^2
398: +{1\over24}u(\phi_i\phi_i)^2
399: \;,
400: \eqa
401: where $i$ runs from 1 to $N$. Summation over $i$ is implicitly
402: understood.
403: The large-$N$ limit is obtained by taking $N\rightarrow\infty$, while
404: keeping $uN$ constant.
405:
406: Generally, the diagrams that contribute to $\Delta\langle\phi^2\rangle$
407: can be obtained from vacuum diagrams by
408: inserting an operator $\phi^2$.
409: For example, at the three-loop level, there are two
410: diagram that contribute and they are shown in Fig
411: diagrams in Fig.~\ref{ex}.
412: It can be shown that the first diagram is suppressed by a factor of $1/N$
413: relative to the second.
414:
415: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
416: \begin{figure}[htb]
417: \epsfysize=2.2cm
418: \vspace{0.2cm}
419: \epsffile{ex.eps}
420: %\vspace{2mm}
421: \caption[a]{Three-loop
422: Feynman diagrams contributing to $\Delta\langle\phi^2\rangle$.}
423: \label{ex}
424: \end{figure}
425:
426:
427:
428: The Feynman diagrams that contribute to $\Delta\langle\phi^2\rangle$
429: at leading order in $1/N$ are shown in Fig.~\ref{nden}.
430: The dot denotes an insertion of the operator $\phi^2$.
431: These diagrams are called bubble diagrams and the bubble summation
432: is thus exact in the large-$N$ limit.
433: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
434: \begin{figure}[htb]
435: \epsfysize=1.4cm
436: \vspace{0.2cm}
437: \epsffile{nden.eps}
438: %\vspace{2mm}
439: \caption[a]{Feynman diagrams contributing to $\Delta\langle\phi^2\rangle$
440: to leading order in $1/N$.}
441: \label{nden}
442: \end{figure}
443: The expression for the expectation value $\Delta\langle\phi_i^2\rangle$
444: then becomes
445: \bqa
446: \Delta\langle\phi_i^2
447: \rangle&=&%\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
448: -N\int{d^dp\over(2\pi)^d}{1\over p^4}\Sigma(p)\;,
449: \label{sigmaapp}
450: \eqa
451: where $\Sigma(p)$ is the self-energy.
452: The Feynman diagrams for the self-energy
453: are obtained from those in Fig.~\ref{nden} by
454: cutting the propagator line that goes through the dots.
455: The self-energy diagrams are shown in Fig.~\ref{nself}.%\vspace{1.2cm}
456: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
457: \begin{figure}[htb]
458: \epsfysize=0.95cm
459: \vspace{0.5cm}
460: \epsffile{nself.eps}
461: %\vspace{2mm}
462: \caption[a]{Feynman graphs contributing to the self-energy to leading order
463: in $1/N$.}
464: \label{nself}
465: \end{figure}
466:
467: After mass renormalization, so that $\Sigma(0)=0$,
468: the expression for the self-energy is~\cite{zinn}
469: \bqa\nonumber
470: \Sigma(p)&=&{2\over N}
471: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
472: \int{d^dk\over(2\pi)^d}
473: {1\over6/Nu+{\cal T}_1(k)}
474: %\\ &&\times
475: \left[
476: {1\over|{\bf p}+{\bf k}|^2}-{1\over k^2}
477: \right]\;,
478: \\ &&
479: \label{lnself}
480: \eqa
481: where ${\cal T}_1(k)$ is the one-loop contribution to the
482: four-point function:
483: \bqa
484: {\cal T}_1(k)&=&%\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
485: \int{d^dq\over(2\pi)^d}{1\over q^2|{\bf q}+{\bf k}|^2}\;.
486: \label{tdeff}
487: \eqa
488: In the appendix, we show how to calculate
489: the function ${\cal T}_1(k)$ in dimensional regularization. One finds
490: \bqa
491: {\cal T}_1(k)&=&
492: M^{2\epsilon}
493: %\left(4e^{\gamma}M^2\right)^{\epsilon}
494: {\Gamma\left(2-{d\over2}\right)\Gamma\left({d\over2}-1\right)
495: \over2^{2d-3}\pi^{{d-1\over2}}\Gamma\left({d-1\over2}\right)}
496: k^{d-4}
497: \;.
498: \label{tdr}
499: \eqa
500: The next step is to evaluate the integral over $p$ in Eq.~(\ref{sigmaapp}).
501: Using Eq.~(\ref{pint}), we obtain
502: \bqa \nonumber
503: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
504: \int{d^dp\over(2\pi)^d}{1\over p^4}
505: \left[
506: {1\over|{\bf p}+{\bf k}|^2}-{1\over k^2}
507: \right]&=&
508: \\ &&\hspace{-2cm}
509: M^{2\epsilon}
510: {\Gamma\left(3-{d\over2}\right)\Gamma\left({d\over2}-2\right)
511: \over2^{2d-4}\pi^{d-1\over2}\Gamma\left({d-3\over2}\right)}k^{d-6}
512: \;.
513: \label{pdr}
514: \eqa
515: Inserting Eqs.~(\ref{tdr}) and~(\ref{pdr}) into Eq.~(\ref{lnself}),
516: we obtain
517: \bqa\nonumber
518: \Delta\langle\phi_i^2\rangle&=&-M^{2\epsilon}
519: {\Gamma\left(3-{d\over2}\right)\Gamma\left({d\over2}-2\right)
520: \over2^{2d-5}\pi^{d-1\over2}\Gamma\left({d-3\over2}\right)}
521: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
522: \\ &&\times
523: \int{d^dk\over(2\pi)^d}{k^{d-6}\over a+bk^{d-4}}
524: \;,
525: \label{above2}
526: \eqa
527: where
528: \bqa
529: a&=&6/Nu\;,\\
530: b&=&
531: M^{2\epsilon}
532: {\Gamma\left(2-{d\over2}\right)\Gamma\left({d\over2}-1\right)
533: \over2^{2d-3}\pi^{{d-1\over2}}\Gamma\left({d-1\over2}\right)}
534: \;.
535: \eqa
536: The final step consists of integrating over $k$.
537: Using Eq.~(\ref{neeed}) in the Appendix
538: we obtain
539: \bqa\nonumber
540: \Delta\langle\phi^2_i\rangle&=&M^{4\epsilon}
541: %\left(4e^{\gamma}M^2\right)^{\epsilon}
542: a^{d-2\over d-4}b^{3-d\over d-4}
543: \\ &&\hspace{-1cm}\times
544: {\Gamma\left(3-{d\over2}\right)\Gamma\left({d\over2}-2\right)
545: \Gamma\left({2-d\over4-d}\right)
546: \Gamma\left({6-d\over4-d}\right)
547: \over2^{3d-6}\pi^{d-{1\over2}}\Gamma\left({d-3\over2}\right)}\;.
548: \eqa
549: The limit $d\rightarrow3$ is regular and one obtains
550: \bqa
551: \Delta\langle\phi^2_i\rangle
552: &=&-{Nu\over96\pi^2}
553: \label{1/n}
554: \;.
555: \eqa
556: Inserting the result~(\ref{1/n}) into~(\ref{delten}),
557: we obtain the critical temperature to leading order in $1/N$:
558: \bqa\nonumber
559: {\Delta T_c\over T_c^0}&=&{8\pi\over3
560: \left[\zeta\left({3\over2}\right)\right]^{4/3}}an^{1/3}
561: \\ &\approx&2.33n^{1/3}a
562: \;,
563: \label{lotc}
564: \eqa
565: where we have used that $u=3gT$ and $g=8\pi a$ and set $N=2$.
566: Thus $T_c$ {\it increases} linearly with $a$. As noted
567: before~\cite{baym3}, the result~(\ref{lotc}) is independent
568: of $N$. However, it is only valid in the limit $N\rightarrow\infty$.
569: %The reason for the independence of $N$ is simply that
570:
571: The $1/N$ correction to the above result has recently been calculated by
572: Arnold and
573: Tom\'a\^sik~\cite{at}.
574: This is a very lengthy and technically complicated calculation.
575: For instance the integrals that one encounters
576: at order $1/N$ are difficult to evaluate in $3-2\epsilon$ dimensions.
577: Instead, Arnold and Tom\'a\^sik always reduced their diagrams
578: to unambigious integrals in three dimensions that are simpler to evaluate.
579: We shall not review the calculation, but merely state the result.
580: Through next-to-leading order in $1/N$, the shift in $T_c$ is
581: \bqa
582: {\Delta T_c\over T_c^0}&=&
583: {8\pi\over3\left[\zeta\left({3\over2}\right)\right]^{4/3}}\left[
584: 1-{0.5272\over N}\right]n^{1/3}a
585: \;.
586: \eqa
587: For $N=2$, it gives a correction of only $26\%$:
588: \bqa
589: {\Delta T_c\over T_c^0}&=&1.71n^{1/3}a
590: \;.
591: \eqa
592:
593:
594: \subsubsection{Monte Carlo simulations}
595: The action~(\ref{3eff}) can also be used as the starting point for numerical
596: calculations of the critical temperature and other
597: nonuniversal effects of a dilute Bose gas.
598: This is done by putting the theory on a lattice and using Monte Carlo methods
599: to solve the theory
600: nonperturbatively. Such numerical simulations have been carried out by
601: several groups~\cite{grut,kraut,arnold1,arnlat,svis}.
602: This approach has been discussed in considerable detail by Arnold and
603: Moore~\cite{arnold1}.
604: The value of the coefficient in~(\ref{proptcc}) reported by Gru\"ter
605: {\it et al.}~\cite{grut} is $c\simeq0.34$, while Holzmann and
606: Krauth~\cite{kraut} obtained $c\simeq2.3$.
607: The most recent values reported by
608: Arnold and Moore~\cite{arnlat} and by Kashurnikov {\it et al.}~\cite{svis}
609: are $c\simeq1.32$ and $c\simeq1.29$, respectively.
610: One source of discrepancy lies in the nonlinear corrections
611: to $T_c$ as a function of $a$ at the densities where the
612: simulations of~\cite{grut} were carried performed.
613: In~\cite{kraut}, the authors expand the integrand in the path
614: integral in powers of the interaction and keep only the first
615: term in that expansion. This perturbative treatment of the
616: interaction is incorrect since the physics close the phase transition
617: is inherently nonperturbative. This is in contrast to the
618: Monte Carlo simulations based on the action~(\ref{3eff})
619: carried out~\cite{arnlat,svis}. These calculations agree
620: within error bars. We will regard these lattice results as the
621: correct result for the coefficient $c$ in Eq.~(\ref{proptcc}).
622:
623:
624:
625: \subsubsection{Other calculations}
626: \label{oc}
627: In this subsection, we will briefly discuss other calculations
628: of the shift in $T_c$ in the dilute Bose gas.
629: Generally, the expectation value $\langle\phi^2\rangle$ can be written as
630: \bqa
631: \langle\phi^2\rangle&=&%\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
632: \int{d^dk\over(2\pi)^d}
633: {1\over k^2+\mu_3+\Sigma(k)}
634: \;,
635: \eqa
636: where $\Sigma(k)$ is the self-energy function. The critical point
637: is determined by the condition that
638: the correlation length becomes infinite, or equivalently that
639: the effective chemical potential $\mu_3+\Sigma(0)$ vanishes:
640: \bqa
641: \mu_3+\Sigma(0)&=&0\;.
642: \label{epcon}
643: \eqa
644: In absence of interactions, the condition~(\ref{epcon})
645: reduces to the well known $\mu_3=0$. Using Eq.~(\ref{epcon}) to eliminate
646: the chemical potential $\mu_3$, we can write Eq.~(\ref{delten}) as
647: \bqa%\nonumber
648: \Delta\langle\phi^2\rangle
649: &=&
650: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
651: \int{d^dk\over(2\pi)^d}
652: \left[{1\over k^2+\Sigma(k)-\Sigma(0)}-{1\over k^2}\right]\;.
653: %\\ &&
654: \label{selfeq}
655: \eqa
656: The next step is to make an approximation for the
657: self-energy function appearing in Eq.~(\ref{selfeq}). In the previous
658: section, we used the large-$N$ expression for the self-energy~(\ref{lnself}).
659: Baym {\it et al.}~\cite{baym2} considered three different
660: equations for the self-energy.
661: \begin{itemize}
662: \item{One-bubble approximation:}
663: \end{itemize}
664: The self-energy is approximated by the second diagram in Fig.~\ref{ptself}:
665: \bqa\nonumber
666: \Sigma(p)-\Sigma(0)&=&-2g^2T
667: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
668: \int{d^dk\over(2\pi)^d}{\cal T}_1(k)
669: \\ && \nonumber
670: \hspace{-2.4cm}
671: \times
672: \left[
673: {1\over|{\bf p}+{\bf k}|^2+\Sigma(|{\bf p}+{\bf k}|)
674: -\Sigma(0)}-{1\over k^2+\Sigma(k)-\Sigma(0)}
675: \right]\;,
676: \\ &&
677: \label{selfcon1}
678: \eqa
679:
680:
681: \begin{itemize}
682: \item{Bubble-summation approximation:}
683: \end{itemize}
684: The self-energy is approximated by the bubble sum in Fig.~\ref{nself}:
685: \bqa\nonumber
686: \Sigma(p)-\Sigma(0)&=&-2g^2T
687: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
688: \int{d^dk\over(2\pi)^d}{{\cal T}_1(k)\over1+2g{\cal T}_1(k)}
689: \\ && \nonumber
690: \hspace{-2.4cm}
691: \times
692: \left[
693: {1\over|{\bf p}+{\bf k}|^2
694: +\Sigma(|{\bf p}+{\bf k}|)-\Sigma(0)}-{1\over k^2+\Sigma(k)-\Sigma(0)}
695: \right]\;,
696: \\ &&
697: \label{selfcon2}
698: \eqa
699:
700: \begin{itemize}
701: \item{Ladder-summation approximation:}
702: \end{itemize}
703: The self-energy is approximated by the ladder sum similar to the bubble sum.
704: \bqa\nonumber
705: \Sigma(p)-\Sigma(0)&=&
706: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
707: \int{d^dk\over(2\pi)^d}{{\cal T}_1(k)\over1+g{\cal T}_1(k)}
708: \\ && \nonumber
709: \hspace{-2.4cm}
710: \times
711: \left[
712: {1\over|{\bf p}+{\bf k}|^2
713: +\Sigma(|{\bf p}+{\bf k}|)-\Sigma(0)}-{1\over k^2+\Sigma(k)-\Sigma(0)}
714: \right]\;,
715: \\ &&
716: \label{selfcon3}
717: \eqa
718: Note that Eqs.~(\ref{selfcon1})--(\ref{selfcon3}) have been made
719: self-consistent
720: by replacing the free propagators on the right hand side by the
721: interacting propagators. Note also that the only difference between the
722: bubble sum and the ladder sum is a factor of two.
723:
724: Eqs.~(\ref{selfcon1})--(\ref{selfcon3}) have been solved numerically and the
725: results were used to evaluate Eq.~(\ref{selfeq}) to obtain the
726: corresponding shifts in $T_c$. The results for the coefficient $c$ are
727: $3.8$, $2.5$, and $1.6$, respectively and thus
728: within a factor three. Note also that the prediction for the
729: shift in $T_c$ from the non self-consistent bubble summation
730: (the leading $1/N$ result with $N=2$) is very close to the result from
731: self-consistent bubble summation.
732:
733:
734: We can gain more insight into the mechanism behind the increase in $T_c$
735: by looking at the modification of the spectrum for small momenta.
736: Let us consider the non self-consistent bubble sum:
737: \bqa
738: \epsilon(k)&=&k^2+\Sigma(k)-\Sigma(0)\;,
739: \eqa
740: where
741: \bqa
742: \Sigma(k)&=&
743: \int{d^dk\over(2\pi)^d}{{\cal T}_1(k)\over1+g{\cal T}_1(k)}
744: {1\over|{\bf p}+{\bf k}|^2}\;.
745: \eqa
746: For small momenta $k$,
747: the difference $\Sigma(k)-\Sigma(0)$ can be approximated by~\cite{baym2}
748: \bqa
749: \Sigma(k)-\Sigma(0)&=&-{2\over3\pi^2}
750: k^2\left(\log{k\over k_c}-{1\over3}\right)\;,\hspace{0.2cm}k\ll k_c\;,
751: \label{sigmahard}
752: \eqa
753: where $k_c=8\pi^2a/\lambda_T^2$ is the screening wave number.
754: The logarithmic term in~(\ref{sigmahard}) indicates a modified
755: spectrum for small wave numbers $\sim k^{2-\eta}$
756: and thus a hardening (The spectrum is of the form $k^{\alpha}$,
757: where $\alpha<2$)
758: of the spectrum
759: compared to the noninteracting case.
760: The hardening results from correlations among particles with low momentum,
761: which leads to a
762: decrease in the critical density and thus an increase in the
763: critical temperature. Other approximations show a different functional
764: dependence of the difference $\Sigma(k)-\Sigma(0)$ at small $k$, but
765: the basic mechanism remains the same, namely a hardening of the spectrum
766: at the critical temperature. It has been pointed out~\cite{gordon}
767: that the hardening of the spectrum for the $n=0$ Matsubara mode
768: only takes place exactly at $T_c$. The Bogoliubov operator
769: inequality
770: guarantees that the spectrum remains quadratic away from $T_c$.
771:
772:
773:
774: Variational methods have also been used
775: recently~\cite{ram1,ram2,tceric}
776: to calculate $T_c$. The basic idea is to compute~(\ref{selfeq}) using
777: an effective three-dimensional Lagrangian that has been reorganized
778: according to the discussion in Sec.~\ref{subopt}.
779: The Lagrangian is written as ${\cal L}={\cal L}_0+{\cal L}_{\rm int}$, where
780: \bqa
781: {\cal L}_0&=&-{1\over2}\phi_i\nabla^2\phi_i+{1\over2}m^2\phi^2_i\;,
782: \label{vpt1def}\\
783: {\cal L}_{\rm int}&=&
784: {1\over2}\delta(\mu_3-m^2)\phi_i^2+{1\over24}\delta u(\phi_i\phi_i)^2\;,
785: \label{vpt2def}
786: \eqa
787: and $i$ runs from 1 to $N$.
788: Calculations are carried out by using $\delta$ as a formal expansion parameter,
789: expanding to a given order in $\delta$ and setting $\delta=1$ at the end
790: of the calculation. Finally, we need to give a prescription for the mass
791: parameter $m$. In the calculations below, we will use the PMS criterion.
792: In this context, it reads
793: \bqa
794: {\partial \Delta\langle\phi^2\rangle\over\partial m}&=&0\;.
795: \label{pmsm}
796: \eqa
797: If we wish to apply variational methods,
798: to calculating the shift in $T_c$, we need to generalize
799: the quantity $\Delta\langle\phi^2\rangle$ appearing in
800: Eq.~(\ref{selfeq})
801: to the field theory defined by Eqs.~(\ref{vpt1def}) and~(\ref{vpt2def}).
802: One must be able to expand this quantity in powers of $\delta$ and it must
803: reduce to $\Delta\langle\phi^2\rangle$ when $\delta=1$.
804: Several prescriptions for generalizing $\Delta\langle\phi^2\rangle$
805: have been proposed in the literature\cite{ram1,ram2,tceric}.
806: Some of these prescriptions are well behaved in the limit $N\rightarrow\infty$
807: and some of them are not. Since the result for the shift in $T_c$ is
808: known analytically in this limit, this a desireable property of a
809: prescription. Two generalizations that have this property have been
810: considered~\cite{tceric}:
811: \bqa
812: \Delta_a\langle\phi^2\rangle&=&
813: N\int{d^dk\over(2\pi)^d}\left[
814: {1\over k^2+\Sigma(k)-\Sigma(0)}-{1\over k^2}
815: \right]\;,
816: \label{propa1}
817: \\\nonumber
818: \Delta_b\langle\phi^2\rangle&=&
819: N\int{d^dk\over(2\pi)^d}\left[
820: {1\over k^2+m^2(1-\delta)+\Sigma(k)-\Sigma(0)}\right.
821: \\ &&
822: \left.
823: -{1\over k^2+m^2(1-\delta)}
824: \right]\;.
825: \eqa
826: In the following, we will consider $\Delta_a\langle\phi^2\rangle$.
827: The strategy is to calculate the difference
828: $\Sigma(k)-\Sigma(0)$ in a powers series in $\delta$, substitute
829: the result into Eq.~(\ref{propa1}), and finally expand the resulting
830: integral in powers of $\delta$. The Feynman diagram that contributes
831: to the self-energy to first order in $\delta$ is the leftmost diagram
832: in Fig.~\ref{ptself}. It is independent of the external momentum and
833: so the difference $\Sigma(k)-\Sigma(0)$ vanishes.
834: The first nonzero contribution to the
835: quantity $\Sigma(k)-\Sigma(0)$ is then given by the two-loop diagram
836: in Fig.~\ref{ptself}. The expression is
837: \bqa\nonumber
838: \Sigma_2(k)-\Sigma_2(0)&=&\delta^2{N(N+2)\over6}u^2
839: \int{d^3p\over(2\pi)^3}\int{d^3q\over(2\pi)^3}
840: \\ && \nonumber
841: \hspace{-1cm}
842: \times
843: \left[
844: {1\over p^2+m^2}{1\over q^2+m^2}{1\over(p+q+k)^2+m^2}
845: \right.
846: \\ &&
847: \hspace{-1cm}
848: \left.
849: -{1\over p^2+m^2}{1\over q^2+m^2}{1\over(p+q)^2+m^2}
850: \right]\;,
851: \eqa
852: where the subscript $n$ indicates the order in the loop expansion.
853: We have set $d=3$ since the integral finite in three dimensions.
854: The integral is calculated in the appendix. Using Eq.~(\ref{selfdiff}),
855: we obtain
856: \bqa\nonumber
857: \Sigma_2(k)-\Sigma_2(0)&=&\delta^2{N(N+2)\over6}u^2
858: {1\over(4\pi)^2}\Bigg[
859: 1
860: \\ &&
861: \hspace{-1cm}
862: -{3m\over k}\arctan{k\over3m}
863: -{1\over2}\log{k^2+9m^2\over9m^2}
864: \Bigg]\;.
865: \label{selfdel}
866: \eqa
867: The self-energy~(\ref{selfdel}) is itself second order in $\delta$.
868: The second-order result for $\Delta_a^{(2)}\langle\phi^2\rangle$
869: is then obtained by expanding~(\ref{propa1}) in powers of the subtracted
870: self-energy and keeping only the first term:
871: \bqa
872: \Delta_a^{(2)}\langle\phi^2\rangle&=&
873: \int{d^3k\over(2\pi)^3}{1\over k^4}\left[
874: \Sigma_2(k)-\Sigma_2(0)
875: \right]\;,
876: \eqa
877: where the superscript indicates the order of $\delta$.
878: Using Eq.~(\ref{delta2c})
879: \bqa
880: \Delta_a^{(2)}\langle\phi^2\rangle&=&
881: -{1\over108(4\pi)^3}{1\over m}
882: \delta^2u^2N(N+3)\;.
883: \label{pms2}
884: \eqa
885: The PMS criterion~(\ref{pmsm}) has no solution at second order in $\delta$
886: because~(\ref{pms2}) is a monotonic function of $m$.
887: Thus one has
888: to go the third order
889: in order to obtain a value for $m$. The result
890: for
891: $\Delta_a^{(3)}\langle\phi^2\rangle$
892: reads~\cite{tceric}
893: \bqa\nonumber
894: \Delta_a^{(3)}\langle\phi_i^2\rangle&=&
895: -{1\over108(4\pi)^3}{1\over m}
896: \delta^2u^2N(N+3)\left(1+{1\over2}\delta\right)
897: \\ &&
898: -\delta^3{N(N+2)(N+8)}u^3I_3\;,
899: \eqa
900: where
901: \bqa\nonumber
902: I_3&=&-{1\over24(4\pi)^4}\left[
903: \pi^2+16\log{3\over4}+12{\rm Li}_2(-1/3)
904: \right]{1\over m^2}\;.
905: \\ &&
906: \eqa
907: Here
908: \bqa
909: {\rm Li}_n(x)&=&\sum_{i=1}^{\infty}{x^i\over i^n}
910: \eqa
911: is the polylogarithmic function.
912: At this order, the PMS criterion has a single real solution:
913: \bqa
914: m&=&1.04u{N+8\over24\pi}\;.
915: \eqa
916: The resulting value for
917: $\Delta_a^{(2)}\langle\phi_i^2\rangle$ is
918: \bqa
919: \Delta_a^{(2)}\langle\phi_i^2\rangle&=&
920: 0.4813{N+2\over N+8}\left(-{Nu\over96\pi^2}\right)\;.
921: \eqa
922: Setting $N=2$, the prefactor becomes $0.19$.
923: The lattice results~\cite{arnlat} is
924: \bqa
925: \Delta\langle\phi_i^2\rangle&=&
926: \left(0.284\pm0.004\right)
927: \left(-{Nu\over96\pi^2}\right)\;.
928: \eqa
929: Thus the third-order approximation differs from the lattice Monte Carlo
930: result by $66$\%. Similarly, it was found~\cite{tceric} that the fourth
931: order approximation differs from the numerical simulations by 61\%
932: In the paper by Braaten and
933: Radescu~\cite{tceric}, the convergence of the linear $\delta$
934: expansion to the exact result in the large--$N$ limit was studied as well.
935: It was shown that it converges to the lattice Monte-Carlo result~(\ref{lotc}),
936: but that the convergence is rather slow.
937:
938: A straightforward
939: application of the linear delta-expansion to field theoretic
940: problems have recently been criticized by Kleinert~\cite{kleinert,kleinertb}.
941: One must take into account the correct Wegner
942: exponent~\cite{we} that governs
943: the approach to the
944: strong-coupling limit. The method is then called variational perturbation
945: theory (VPT)~\cite{vptdef1}. The failure to take into account the
946: correct Wegner exponent is the reason for why one finds complex-valued
947: solutions in the linear delta-expansion.
948: These matters are highly technical and
949: beyond the scope of this paper. Interested
950: readers are referred to the textbook by Kleinert and
951: Schulte-Frolinde~\cite{kleinert2}.
952: VPT has very
953: reently been applied in a seven-loop calculation by Kastening~\cite{boris2}
954: (See also~\cite{kleinert,boris} for five and and six-loop calculations.).
955: The result for the coefficient is $c=1.28\pm 0.1$ in excellent agreement
956: with lattice field theory results. We also note that seven-loop
957: calculations have been carried out for the case $N=1$ and $N=4$ as well.
958: The values for the coefficient $c$ are $1.07\pm0.10$ and
959: $1.54\pm0.11$, respectively. These results are also in good agreement
960: with the lattice simulation of Sun~\cite{sun} who obtained the
961: values $1.09$ and $1.59$, respectively.
962:
963:
964:
965: We close this section by listing in Table.~\ref{tctable}
966: the predictions for the shift
967: in the critical temperature that has been obtained by the various methods
968: discussed in this section.
969:
970: \begin{table}[htb]
971: \begin{center}
972: \begin{tabular}{llllll}\hline
973: \\
974: ${\Delta T_c\over T_0}$&=&$2.33n^{1/3}a$,%\hspace{0.3cm}
975: & Leading order $1/N$
976: \\
977: &&&~\cite{baym3}.
978: \\
979: &&&\\
980: ${\Delta T_c\over T_0}$&=&$1.71n^{1/3}a$,\hspace{0.3cm}&
981: Next-to-leading order $1/N$
982: \\
983: &&&~\cite{at}.\\
984: &&&\\
985: ${\Delta T_c\over T_0}$&=&$\left(1.32\pm0.02\right)n^{1/3}a$,%\hspace{0.3cm}
986: &
987: Lattice~\cite{arnlat}.
988: \\
989: &&&\\
990: ${\Delta T_c\over T_0}$&=&$\left(1.29\pm0.05\right)n^{1/3}a$,%\hspace{0.3cm}
991: &
992: Lattice~\cite{svis}.
993: \\
994: &&&\\
995: ${\Delta T_c\over T_0}$&=&$0.7n^{1/3}a$,%\hspace{0.3cm}
996: &
997: One-bubble approximation
998: \\
999: &&&~\cite{laloe}.\\
1000: &&&\\
1001: ${\Delta T_c\over T_0}$&=&$3.8n^{1/3}a$,%\hspace{0.3cm}
1002: &
1003: One-bubble self-consistent
1004: \\
1005: &&&approximation
1006: \\
1007: &&&~\cite{baym2}.\\
1008: &&&\\
1009: ${\Delta T_c\over T_0}$&=&$2.5n^{1/3}a$,%\hspace{0.3cm}
1010: &
1011: Ladder-summation approximation
1012: \\
1013: &&&~\cite{baym2}.\\
1014: &&&\\
1015: ${\Delta T_c\over T_0}$&=&$1.6n^{1/3}a$,%\hspace{0.3cm}
1016: &
1017: Bubble summation approximation
1018: \\
1019: &&&~\cite{baym2}.\\
1020: &&&\\${\Delta T_c\over T_0}$&=&$(1.27\pm0.11)n^{1/3}a$,
1021: &
1022: 7-loop VPT~\cite{boris2}.
1023: \\
1024: &&&\\${\Delta T_c\over T_0}$&=&$(1.23\pm)n^{1/3}a$,
1025: &
1026: RG in three dimensions\\
1027: &&&~\cite{ledow}.
1028: \\
1029: &&&\\${\Delta T_c\over T_0}$&=&$(1.3\pm0.4)n^{1/3}a$,
1030: &
1031: Simulations of classical field theory
1032: \\
1033: &&&~\cite{tcdavis}.
1034: \\
1035: &&&\\ \hline
1036: \end{tabular}
1037: \end{center}
1038: \caption{The critical temperature for a dilute Bose
1039: gas obtained by various analytic and numerical methods.}
1040: \label{tctable}
1041: \end{table}
1042: