cond-mat0511437/RG.tex
1: \documentstyle[aps,preprint,pra,epsfig]{revtex}
2: %\documentstyle[aps,multicol,pra,epsfig]{revtex}
3: %\documentclass{article}                  
4: 
5: \begin{document}
6: 
7: \draft
8: 
9: \title{Finite-Temperature Renormalization Group Analysis of 
10: Interaction Effects in $2D$ Lattices of Bose-Einstein Condensates}
11: 
12: \author{A. Smerzi$^{1,2}$, P. Sodano$^{3}$, and A. Trombettoni$^{4}$}
13: \address{$^1$ Istituto Nazionale per la Fisica della Materia BEC-CRS and
14: Dipartimento di Fisica, Universita' di Trento, I-38050 Povo, Italy\\
15: $^2$ Theoretical Division, Los Alamos
16: National Laboratory, Los Alamos, NM 87545, USA\\
17: $^3$ Dipartimento di Fisica and Sezione I.N.F.N., Universit\`a di 
18: Perugia, Via A. Pascoli, I-06123 Perugia, Italy\\
19: $^4$ Istituto Nazionale per la Fisica della Materia and
20: Dipartimento di Fisica, Universita' di Parma,
21: parco Area delle Scienze 7A, I-43100 Parma, Italy\\
22: }
23: 
24: \date{\today}
25: \maketitle
26: 
27: \begin{abstract}
28: By using a renormalization group analysis, 
29: we study the effect of interparticle interactions on the 
30: critical temperature $T_{BKT}$ at which the Berezinskii-Kosterlitz-Thouless 
31: (BKT) transition occurs for Bose-Einstein condensates loaded at 
32: finite temperature in a $2D$ optical lattice. 
33: We find that $T_{BKT}$ decreases as the 
34: interaction energy decreases; when $U/J=36/\pi$ one has $T_{BKT}=0$, 
35: signaling the possibility of a quantum phase transition of BKT type.
36: \end{abstract}
37: 
38: %\begin{multicols}{2}
39: 
40: \section{Introduction}
41: 
42: It has been recently suggested \cite{trombettoni03} 
43: that a $2D$ optical lattice of Bose-Einstein 
44: condensates \cite{greiner01} may allow for the observation 
45: of a finite-temperature phase transition to a superfluid regime 
46: where the phases of the single-well condensates are coherently aligned. 
47: In fact, in an appropriate range of parameters, 
48: the thermodynamical properties of the bosonic lattice at finite temperature 
49: may be well described by the Hamiltonian of the $XY$ model 
50: \cite{trombettoni03}, and, as it is well known, the $2D$ $XY$ model 
51: exhibits the Berezinskii-Kosterlitz-Thouless (BKT) 
52: transition \cite{BKT,minnaghen87,kadanoff00}. 
53: The BKT phase transition occurs via an 
54: unbinding of vortex defects: the low-temperature phase, 
55: for $T$ below the BKT temperature $T_{BKT}$,  
56: is characterized by the presence of bound vortex-antivortex pairs 
57: and the spatial correlations exhibit a power-law 
58: decay. For $T \sim T_{BKT}$ the pairs starts to unbind 
59: (see e.g. \cite{tobochnik79,leoncini98}): in the high-temperature phase, 
60: only free vortices are present, leading to an
61: effective randomization of the phases and to an exponential decay 
62: of the correlation functions.
63: 
64: The Hamiltonian describing the properties of bosons 
65: in deep optical lattices 
66: is the so called Bose-Hubbard Hamiltonian \cite{jaksch98}, 
67: in which two terms are present: a kinetic term 
68: describing the hopping of the bosons with 
69: tunneling rate $t$, and a potential term describing 
70: the interaction between bosons in the wells of the array 
71: with interaction energy $U$ (which is proportional to the 
72: $s$-wave scattering length). The description of bosons in 
73: deep lattices by means of the Bose-Hubbard Hamiltonian holds 
74: also at finite temperature, provided that the temperature effects 
75: do not induce the occupation of higher bands: for a $2D$ optical 
76: lattice confined to a plane by a magnetic potential having frequency 
77: $\omega_z$, this implies that the Bose-Hubbard Hamiltonian 
78: description is 
79: valid up temperatures $T$ such that $\hbar \omega_z \gtrsim k_B T$.   
80: When the average number of bosons $N_0$ per site is high enough 
81: and the interaction energy is larger than $t/N_0$, one may map - 
82: {\em also at finite temperature} - the Bose-Hubbard Hamiltonian 
83: in the quantum phase Hamiltonian \cite{trombettoni03}: 
84: the conjugate variables are the phases and the particle numbers 
85: of the condensates in the different sites of the lattice. 
86: Thus, also at finite temperature, the phase diagram is 
87: determined by the competition of two energies: 
88: the Josephson energy $J \approx 2 t N_0$, 
89: proportional to the tunneling rate between neighbouring sites of the $2D$ 
90: square optical lattice, and the interaction energy $U$. For $U=0$, the 
91: quantum phase Hamiltonian reduces to the $XY$ Hamiltonian, exhibiting then 
92: a BKT transition at $T_{BKT} \sim J/k_B$: for $T < T_{BKT}$ the system 
93: as a whole behaves as a superfluid with phase coherence across the array.  
94: Very accurate Monte Carlo simulations yield, in the thermodynamic limit, 
95: $T_{BKT} = 0.898 J/k_B$ \cite{gupta88}. 
96: 
97: In this paper we determine the effect of the 
98: interaction energy $U$ on the BKT transition. 
99: When $U \ll J$, a BKT transition still occurs at a critical 
100: temperature $T_{BKT}(U)$: we find that  
101: $T_{BKT}(U)$ is smaller than $T_{BKT}$ and decreases with $U$ 
102: (see Fig. 1). Intuitively speaking, the quantum phase Hamiltonian, 
103: as well as the Bose-Hubbard Hamiltonian, in the limit $U \gg J$ 
104: describes a Mott insulator, while in 
105: the opposite limit, $J \gg U$, the array behaves as a 
106: phase-coherent superfluid 
107: (e.g., see \cite{sondhi97,fazio01}): thus, when $U$ increases, 
108: the superfluid region in the phase diagram decreases. 
109: Although in two dimensions there is not long-range order at finite temperature 
110: due to the Mermin-Wagner theorem, 
111: still the system exhibits superfluid behavior 
112: for $T < T_{BKT}$ \cite{fazio01}: thus one expects  
113: that $T_{BKT}(U)$ should decrease when $U$ increases. 
114: We also find (see Section III) that, when $U/J$ is equal to the 
115: critical value $(U/J)_{cr}=36/\pi$, 
116: one has that $T_{BKT}(U) = 0$, signaling the possibility of a ($T=0$) 
117: quantum phase transition. 
118: 
119: The quantum phase Hamiltonian describes also the behavior 
120: of superconducting Josephson networks below the temperature 
121: $T_{BCS}$ at which each junction becomes 
122: superconducting \cite{fazio01,simanek94,martinoli00}. 
123: The momenta conjugate to the phases of the macroscopic 
124: wavefunctions of the superconducting grains are 
125: the number of Cooper pairs, $J$ is the Josephson energy of the 
126: superconducting Josephson junctions and $U$ is the charging energy 
127: due to the Coulomb repulsion between Cooper pairs. 
128: We observe that, in superconducting 
129: Josephson arrays, the interaction term is written 
130: in general as $\sum_{ij} U_{ij} N_i N_j$ where $N_i$ is the Cooper pairs 
131: number at the site $i$: $U_{ij}$ is proportional to the inverse 
132: of the capacitance matrix and may be also non-diagonal (i.e., $U_{ij} 
133: \neq 0$ with $i \neq j$). At variance, for bosons in deep optical lattices, 
134: the interaction term is $\sum_{i} U_{ii} N_i^2$, which corresponds, 
135: in a suitable range of parameters, to a $diagonal$ quantum phase model. 
136: 
137: The analogy between superconducting Josephson networks and 
138: atomic gas in deep optical lattices is then clear: 
139: in each well of the periodic potential there is a condensate grain, 
140: appearing at the Bose-Einstein condensation temperature 
141: $T_{BEC}$. When $T_{BEC}$ is larger than all other energy scales, 
142: the atoms in the well $i$ of the $2D$ optical lattice may be 
143: described by a macroscopic wavefunction. 
144: It becomes apparent, then, that an optical network 
145: can be regarded as a network of {\em bosonic} Josephson junctions. 
146: Furthermore, in both systems, when the interaction term is neglected 
147: respect to the energy associated to the particle hopping, one has 
148: that the whole $2D$ array becomes superfluid at the temperature 
149: $T_{BKT}$ at which the BKT transition occurs, with $T_{BKT}$ smaller 
150: than $T_{BCS}$ for superconducting Josephson 
151: networks and smaller than $T_{BEC}$ for bosonic Josephson networks.
152: 
153: The plan of the paper is the following: in Section II we introduce the 
154: effective Hamiltonian describing bosons hopping on a deep optical lattice 
155: and, by using a semiclassical approximation 
156: \cite{jose84}, we compute the effective Josephson 
157: energy in presence of the interaction energy $U$.
158: In Section III we evaluate the effect of the quantum fluctuations 
159: on $T_{BKT}$ by putting in the renormalization group equations 
160: the effective Josephson energy obtained in Section II; 
161: a comparison with previous results is then carried out. 
162: Details of the computations skipped in the text are in the 
163: Appendix A. Section IV is devoted to some concluding remarks. 
164: 
165: \section{The renormalized Josephson energy}
166: 
167: $2D$ optical lattices are created using two standing waves \cite{greiner01}: 
168: when the polarization vectors of the two laser fields are orthogonal, 
169: the periodic potential is 
170: \begin{equation}
171: V_{opt}=V_0 [ \sin^2{(kx)} + \sin^2{(ky)} ],
172: \label{V_opt}
173: \end{equation}
174: where $k=2 \pi/\lambda$ is the wavevector of the lattice beams. 
175: $V_0$ is usually expressed in units of $E_R=\hbar^2 k^2/2m$ 
176: (where $m$ is the atomic mass): 
177: in \cite{greiner01} it is $\lambda=852 \, nm$ and 
178: $E_R=h \cdot 3.14 kHz$. 
179: Around the minima of the potential (\ref{V_opt}) one has 
180: $V_{opt} \approx m\tilde{\omega}_r^2 (x^2 + y^2)/2$ 
181: with 
182: \begin{equation}
183: \tilde{\omega}_r=\sqrt{\frac{2 V_0 k^2} {m}}.
184: \label{omega_tilde}
185: \end{equation} 
186: When $\tilde{\omega}_r \gg \omega_z$ (with $\omega_z$ 
187: the frequency of the confining magnetic potential superimposed 
188: to the optical potential and acting along $z$), 
189: the system realizes a square array of tubes, 
190: i.e. an array of harmonic traps elongated along the $z$-axis 
191: \cite{greiner01}. 
192: 
193: When all the relevant energy scales are small compared to the excitation 
194: energies, one can expand the field operator \cite{jaksch98} as 
195: \begin{equation}
196: \hat{\Psi}(\vec{r},t)= \sum_i \hat{\psi}_i(t) 
197: \Phi_i(\vec{r})
198: \label{expansion} 
199: \end{equation}
200: with $\Phi_i(\vec{r})$ the Wannier wavefunction localized 
201: in the $i$-th well 
202: (normalized to $1$) and $\hat{N}_i=\hat{\psi}^{\dag}_i 
203: \hat{\psi}_i$ the bosonic number operator. 
204: Substituting the expansion of $\hat{\Psi}(\vec{r},t)$ 
205: in the full quantum Hamiltonian, one gets  
206: the effective Hamiltonian describing the bosons hopping on 
207: the deep optical lattice \cite{jaksch98,fisher89} 
208: \begin{equation}
209: H=-t \sum_{<i, j>} 
210: (\hat{\psi}^{\dag}_i \hat{\psi}_j+ h.c.)
211: + \frac{U}{2} \sum_i \hat{N}_i (\hat{N}_i-1).
212: \label{B-H}
213: \end{equation}
214: In Eq.(\ref{B-H}) $\sum_{<i, j>}$ denotes a sum over all the 
215: distinct pairs of nearest neighbours; 
216: $t$ and $U$ are respectively the tunneling rate and 
217: the interaction energy, and are given by 
218: \begin{equation}
219: t \simeq - \int d\vec{r} \, \Biggl[ \frac{\hbar^2}{2m} 
220: \vec{\nabla} \Phi_i \cdot \vec{\nabla} \Phi_j + 
221: \Phi_i  V_{ext} \Phi_j \Biggr] 
222: \label{t}
223: \end{equation}  
224: and 
225: \begin{equation}
226: U = \frac{4 \pi \hbar^2 a}{m} \int d\vec{r} \, 
227: \Phi_i^4
228: \label{U} 
229: \end{equation}
230: ($a$ is the $s$-wave scattering length).
231: 
232: As discussed in the Introduction, one can show that, 
233: when $V_0$ and $\omega_z$ are large enough, 
234: the system is described by the Hamiltonian (\ref{B-H}) up to temperatures 
235: $T \sim \hbar \omega_z / k_B$ \cite{trombettoni03}. 
236: Upon defining $J \approx 2 t N_0$, the Hamiltonian (\ref{B-H}),  
237: for $N_0 \gg 1 $ and $J/N_0^2 \ll U$ \cite{anglin01}, reduces to 
238: \begin{equation}
239: \hat{H}=- J \sum_{<i,j>} \cos{(\theta_i-\theta_j)}
240: - {U \over 2} \sum_i \frac{\partial^2}{\partial \theta_i^2}.  
241: \label{Q-P-M}
242: \end{equation}
243: For a $2D$ lattice with $V_0$ between $20E_R$ and $25E_R$ and 
244: $N_0 \approx 170$ as in \cite{greiner01}, one sees 
245: that the condition $\gg J/N_0^2$ is rather well 
246: satisfied and that $J/k_B$ is of order of $20nK$. 
247: For $U=0$, (\ref{Q-P-M}) is the $XY$ Hamiltonian:
248: \begin{equation}
249: \hat{H}=- J \sum_{<i,j>} 
250: \cos{(\theta_i-\theta_j)}.
251: \label{X-Y}
252: \end{equation}
253: 
254: The Hamiltonian (\ref{Q-P-M}) describes the so-called 
255: quantum phase model \cite{fazio01,simanek94,martinoli00}. 
256: There is an huge amount of literature on the properties of $2D$ 
257: superconducting Josephson arrays studied by means of the quantum phase model 
258: \cite{fazio01,simanek94,martinoli00}. A (not exhaustive) list of 
259: relevant papers includes mean-field and coarse-graining 
260: approaches \cite{vanotterlo93,kopec00,grignani00}, 
261: Monte Carlo results 
262: \cite{jacobs84,rojas96,capriotti03,alet03,alsaidi04}, 
263: renormalization group calculations \cite{jose84,rojas96} 
264: and self-consistent harmonic approximations \cite{rojas96,kim97,cuccoli00} 
265: (more references are in \cite{fazio01}). In the following we shall study 
266: the renormalization-group equations in which it is used 
267: an effective value of the Josephson energy computed within the 
268: harmonic approximation \cite{jose84}. 
269: 
270: The starting point is the partition function $Z$ of the quantum phase model: 
271: using the path integral formalism, 
272: from (\ref{Q-P-M}), one has 
273: \begin{equation}
274: Z=\int {\cal D} \theta e^{-\frac{1}{\hbar} S[\theta]},
275: \label{part_funct}
276: \end{equation}  
277: where the action $S$ is given by
278: \begin{equation}
279: S[\theta]=
280: \int\limits_{0}^{\hbar \beta} {d\tau \Biggl[\frac{\hbar^2}{2U} 
281: \sum\limits_j { \Biggl(\frac{\partial \theta_j}{\partial
282: \tau}\Biggr)^2}+ J \sum\limits_{<i, j>} {(1-
283: \cos\theta_{ij})} \Biggr]}
284: \label{azione}
285: \end{equation}
286: with $\beta=1/k_B T$ and 
287: $\theta_{ij} \equiv \theta_{i}-\theta_{j}$.
288: Separating the phases as $\theta_i(\tau)=\varphi_i+
289: \delta_i(\tau)$, where $\varphi_i$ is a
290: static vortex configuration and $\delta_i(\tau)$ is a fluctuation
291: about $\varphi_i$, the path-integral partition function 
292: (\ref{part_funct}) can be written as 
293: \begin{equation}
294: Z=\int {\cal D} \varphi {\cal D} \delta e^{-\frac{1}{\hbar} 
295: S[\varphi,\delta]}
296: \label{ZETA} 
297: \end{equation}
298: where 
299: \begin{equation}
300: S[\varphi,\delta]=\int\limits_{0}^{\hbar \beta} 
301: {d\tau \Biggl[ \frac{\hbar^2}{2U} \sum\limits_{i} 
302: {\dot{\delta}_i^2 (\tau)}+J \sum\limits_{<i,j>} 
303: {(1- \cos\varphi_{ij} \, \cos\delta_{ij})} \Biggr]},
304: \label{azione_phi_delta}
305: \end{equation}
306: with $\varphi_{ij} \equiv \varphi_{i}-\varphi_{j}$ e 
307: $\delta_{ij} \equiv \delta_{i}-\delta_{j}$. 
308: Assuming that $\varphi_i$ and $\delta_i$ are slowly varying over the 
309: size of the array \cite{jose84,simanek94}, i.e. 
310: \begin{equation}
311: \cos\varphi_{ij} \approx 1-\varphi_{ij}^2/2
312: \label{APPR1}
313: \end{equation}
314: and 
315: \begin{equation}
316: \cos\delta_{ij} \approx 1-\delta_{ij}^2/2,
317: \label{APPR2}
318: \end{equation}
319: one gets 
320: \begin{equation}
321: Z=Z_0 \int{{\cal D}\varphi \exp\Bigl\{-{1 \over 2} 
322: \beta \bar{J} \sum\limits_{<i,j>} {\varphi_{ij}^2}\Bigr\}},
323: \label{part_funct_phi_delta}
324: \end{equation} 
325: where $Z_0 \equiv \int{{\cal D}\delta \, e^{-{1 \over \hbar} S_0[\delta]}}$ 
326: and 
327: \begin{equation}
328: S_0[\delta]=\int\limits_{0}^{\hbar \beta} 
329: {d\tau \Biggl[\frac{\hbar^2}{2U} \sum\limits_{i} {\dot{\delta}_i^2}+
330: {J \over 2} \sum\limits_{<i,j>} {\delta_{ij}^2}\Biggr]}.
331: \end{equation} 
332: In Eq.(\ref{part_funct_phi_delta}) 
333: $\bar{J}$ is the renormalized Josephson energy, which is given by
334: \begin{equation}
335: \bar{J} \simeq J \Bigl(1-{1 \over 2} <\delta_{ij}^2>_0\Bigr)
336: \label{J_rin}
337: \end{equation} 
338: with 
339: \begin{equation}
340: <\delta_{ij}^2>_0 \equiv \frac{1}{Z_0} 
341: \int{{\cal D}\delta \, e^{-{1 \over \hbar} S_0[\delta]}\delta_{ij}^2}.
342: \label{delta_ij}
343: \end{equation} 
344: The evaluation of $<\delta_{ij}^2>_0$ can be carried out in a standard way  
345: \cite{simanek94} and one has
346: \begin{equation}
347: <\delta_{ij}^2>_0=\sqrt{\frac{\pi U}{J}} 
348: \frac{1}{\eta^3} \int\limits_{0}^{\eta} x^2 \coth{x} \, dx
349: \label{delta}
350: \end{equation}
351: where 
352: \begin{equation}
353: \eta=\beta \sqrt{\pi U J}.
354: \label{eta}
355: \end{equation} 
356: 
357: \section{Renormalization group equations}
358: 
359: In the renormalization group equations for the $2D$ $XY$ model 
360: \cite{BKT,minnaghen87,kadanoff00} the scale-dependent screened charge 
361: ${\cal K}$ depends on the dimensionless scaling parameter 
362: $l=\log{(r/a)}$ (where $r$ is the vortex
363: distance and $a$ is the lattice spacing) and it is given by 
364: ${\cal K}(l)=\beta J/\epsilon(l)$, where the 
365: dielectric constant $\epsilon(l)$ expresses the screening of the 
366: vortex-antivortex interaction due to the presence of other vortices 
367: \cite{kadanoff00}. The renormalization group recursion equations 
368: read \cite{kadanoff00}
369: \begin{equation}
370: \frac{d {\cal K}^{-1}(l)}{dl}=4 \pi^3 y^2(l)
371: \label{prima_rec}
372: \end{equation}
373: and 
374: \begin{equation}
375: \frac{d y(l)}{dl}=[2- \pi {\cal K}(l)] y(l)
376: \label{seconda_rec}
377: \end{equation}
378: where $y \propto r^2 e^{-\beta V(r)/2} e^{-\beta \mu}$, 
379: with $e^{-\beta \mu}$ is the fugacity for creating a vortex pair 
380: and $V(r)$ correspond to the screened vortex-antivortex potential. 
381: The study of the scaling equations (\ref{prima_rec})-(\ref{seconda_rec}) 
382: about the fixed point $y_f=0$, ${\cal K}_f=2/\pi$ shows that 
383: the BKT transition occurs at $2-\pi {\cal K}(l=0) \approx 0$, where 
384: ${\cal K}(l=0)=\beta J/\epsilon(l=0)=\beta J$ \cite{kadanoff00}.
385: 
386: In presence of the quantum fluctuations, one has to replace the 
387: scale-dependent charge ${\cal K}(l)$ by 
388: $\bar{\cal K}(l)$ with $\bar{\cal K}(l=0)= \beta \bar{J}$ and  
389: $\bar{J}$ given by Eqs.(\ref{J_rin}) and (\ref{delta}). 
390: Denoting with $T_{BKT} (U)$ 
391: the BKT transition temperature for a given $U$, one finds  
392: the following equation for $K \equiv J/ k_B T_{BKT}(U)$:
393: \begin{equation}
394: F(K)=2-\pi K  \Biggl(1-\frac{1}{2 \pi^2 X_u K^3} \int_{0}^{\pi \sqrt{X_u} K}
395: x^2 \coth{x} \, dx \Biggr)=0
396: \label{radice}
397: \end{equation}
398: where $X_u=U/\pi J$.
399: 
400: The root of Eq.(\ref{radice}) indicates the critical point 
401: at which a BKT transition occurs. The eventual occurrence 
402: of a double root for Eq.(\ref{radice}) might correspond to what is called 
403: in literature a {\em reentrant} behavior. Indeed it has been 
404: often argued (see Refs. in the review 
405: \cite{fazio01}) that the quantum phase model 
406: may undergo at low temperatures a reentrant transition 
407: induced by the quantum fluctuations: namely, fixing $U/J$ and 
408: lowering the temperature, one could switch from
409: an insulating phase to a superconducting one at $T_{BKT}(U)$ and then, 
410: lowering further the temperature, one finds another critical temperature 
411: $T^{(1)}(U)$ at which the system comes back ({\em reenters}) in the 
412: insulating phase. Consistent with the reentrant scenario is the dramatic 
413: decrease of the specific heat at very low temperatures  
414: \cite{jacobs84} and, as discussed, 
415: the presence of double roots for $T_{BKT}$ in the 
416: renormalization group equations. 
417: 
418: The phenomenon of the reentrance - although not universal - 
419: is a non-perturbative effect: in fact, opposite 
420: results may be obtained by means of 
421: different truncations for the series expansion of the function $F$. 
422: For instance, if in Eq.(\ref{radice}) one expands the hyperbolic cotangent 
423: as $\coth x \approx 1/x+x/3$, one gets an expression for $F(K)$ 
424: which gives two roots for $T_{BKT}$. If instead, for $\eta<\pi$, 
425: one uses the expansion $\coth x=\frac{1}{x}+\sum\limits_{n=1}^{\infty} 
426: \frac{2^{2n} B_{2n}}{(2n)!} x^{2n-1}$ where the $B_n$'s are the 
427: Bernoulli numbers \cite{abramowitz72}, one obtains the equation: 
428: \begin{equation}
429: F(K)=2-\pi K \Biggl(1-\frac{1}{4 K}-\sum\limits_{n=1}^{\infty} 
430: \frac{6^n \,  2^{4n} \, B_{2n} \, x_u^n \, K^{2n-1}}
431: {2 \, (2n+2) \, (2n)!} \Biggr)=0
432: \label{bernoulli}
433: \end{equation}  
434: with $x_u=\pi U/24 J$. Since $B_{4n}<0$ for $n=1,2, \ldots$, and 
435: $B_{4n+2}>0$ for $n=0,1,2, \ldots$, if one truncates the sum 
436: in Eq.(\ref{bernoulli}) to the $(2n+1)$-th order with 
437: $n=0,1,2,\ldots$, one finds 
438: $F(K) \to \infty$ for $K \to \infty$ and 
439: then Eq.(\ref{bernoulli}) has two roots; 
440: at variance, if one truncates the sum 
441: to the $(2n)$-th order with 
442: $n=1,2,\ldots$, one finds 
443: $F(K) \to -\infty$ for $K \to \infty$ and 
444: Eq.(\ref{bernoulli}) admits only one root. Thus one cannot truncate 
445: the sum in Eq.(\ref{bernoulli}) to any finite order, even for $U/J \ll 1$; 
446: the origin of this problem is that the expansion used 
447: to get (\ref{bernoulli}) is a series with terms having alternating 
448: signs. In addition, to find the roots 
449: of the equation $F(K)=0$ one has to evaluate the integral 
450: in (\ref{delta}) up to $\sqrt{UJ}/k_B T_{BKT}$, and, also 
451: with $U$ small, $\eta$ may become large requiring to use all the orders 
452: in Eq.(\ref{bernoulli}). 
453: 
454: A way to overcome the above mentioned 
455: difficulties is to use in Eq.(\ref{radice}) 
456: the Mittag-Leffler expansion of 
457: the hyperbolic cotangent, i.e. 
458: $\pi \coth \pi x=1/x+2x\sum\limits_{n=1}^{\infty} (x^2+n^2)^{-1}$: 
459: in this way Eq.(\ref{radice}) can be written as 
460: \begin{equation}
461: F(K)=2-\pi K  \Biggl[1-\frac{1}{4 K}-\frac{1}{X_u 
462: K^3} \, g(K,X_u) \Biggr]=0
463: \label{radice_impr}
464: \end{equation}
465: where
466: \begin{equation}
467: g(K,X_u) \equiv \sum\limits_{n=1}^{\infty} 
468: \Bigl[\frac{X_u K^2}{2}-\frac{n^2}{2} 
469: \log \Bigl(1+ \frac{X_u K^2}{n^2}\Bigr)\Bigr].
470: \label{g}
471: \end{equation}
472: An analytic expression for $F(K)$ valid for $U/J \ll 1$ is given 
473: by Eq.(\ref{approxx}) in the Appendix A.
474: 
475: A detailed study of the function $F(K)$ is 
476: in the Appendix A. Here we observe that, in the limit $U/J \to 0$, 
477: one finds $2+\frac{\pi}{4}-\pi K=0$, from which
478: $K_0 \equiv K(U/J \to 0)=\frac{8+\pi}{4 \pi}$ and 
479: $k_B T_{BKT} (0) \simeq 1.128 J$. Furthermore, 
480: for $U/J<36/\pi$, one can show (see the Appendix A) that 
481: $F(K) \to - \infty$ for $K \to \infty$ and that $F^\prime (K)<0$: 
482: since $F$ tends to the positive value $2+\pi/4$ for $K \to 0$, 
483: one can safely conclude 
484: that, for $U/J<36/\pi$, Eq.(\ref{radice_impr}) has an {\it unique} solution. 
485: At variance, one can show that, for $U/J>36/\pi$, Eq.(\ref{radice_impr}) 
486: does not admit any solution, while for 
487: $U/J=36/\pi$ one has $K \to \infty$, i.e. $T_{BKT} (U) \to 0$. 
488: A plot of $F(K)$ and $F^\prime(K)$ for three values of $U/J$ 
489: respectively smaller than, equal to and larger than $36/\pi$ is 
490: given in Figs. 2 and 3.
491: One may infer that, at $T=0$, a phase transition is expected to occur 
492: at the critical value $(U/J)_{cr}=36/\pi$: this 
493: value is in reasonable agreement with the 
494: mean-field estimates for the $T=0$ transition. In fact, 
495: for the diagonal quantum phase model, one has that 
496: the mean-field prediction 
497: \cite{vanotterlo93,grignani00} 
498: is $(U/J)_{cr} \approx 2z =8$ where $z=4$ is the number 
499: of nearest neighbours. 
500: 
501: In Fig. 1 we plot as empty circles the values of $T_{BKT}(U)/T_{BKT}(0)$ 
502: as a function of $U/J$ from the numerical solution 
503: of Eq.(\ref{radice_impr}). 
504: If one uses the analytic expression (\ref{approxx}) 
505: for the function $F(K)$ one gets, for $U/J \lesssim 1/2$, 
506: an error lesser than $1 \%$. A much better estimate may be obtained 
507: by expanding the function 
508: $F(K)$ near $K_0$: in this way Eq.(\ref{radice_impr}) reads 
509: \begin{displaymath}
510: F(K) \approx F(K_0)+ (K-K_0) \cdot 
511: F^{\prime}(K_0)=0,
512: \end{displaymath}
513: and then
514: \begin{equation}
515: K \approx K_0 - \frac{F(K_0)}
516: {F^{\prime}(K_0)}.
517: \label{sviluppo}
518: \end{equation}
519: In Fig. 1 Eq.(\ref{sviluppo}) is plotted as a solid line.
520: 
521: We may conclude that Eq.(\ref{radice_impr}) admits, for small $U/J$, 
522: a unique solution and this excludes the possibility of reentrant 
523: behavior. Of course, our conclusion 
524: relies on the approximations (\ref{APPR1})-(\ref{APPR2}) 
525: made in order to estimate $\bar{J}$ and $<\delta_{ij}^2>_0$. 
526: A more careful treatment accounting for the 
527: periodicity of the phases in Eq.(\ref{azione_phi_delta}) 
528: seems to be highly desirable to put on a more solid base 
529: all the contrasting issues related to the reentrant 
530: behavior of the systems described by the Hamiltonians (\ref{B-H}) and 
531: (\ref{Q-P-M}). It is comforting to see that the experimental study  
532: of weakly interacting bosons on $2D$ optical networks and 
533: the investigations of their superfluidity properties at very low temperatures 
534: could provide new insights also on this 
535: intriguing and yet poorly understood problem 
536: (see also the recent papers \cite{kleinert03}).
537: 
538: \section{Concluding remarks}
539:   
540: In this paper we studied, by means of a renormalization group analysis, 
541: the effect of interparticle interactions on the 
542: critical temperature at which the Berezinskii-Kosterlitz-Thouless 
543: transition occurs for Bose-Einstein condensates loaded in a $2D$ 
544: optical lattice at finite temperature. We determined the shift 
545: of the Berezinskii-Kosterlitz-Thouless transition temperature 
546: induced by the interaction term and we compared our findings with 
547: previously known results.  
548: 
549: \appendix 
550: 
551: \section{Properties of the renormalization group equation}
552: 
553: In this Appendix we study the properties of the function 
554: $F(K)$ defined in Eq.(\ref{radice_impr}). 
555: We show in (1) that $F(0)>0$, in (2) that 
556: $F(K) \to -\infty$ for $K \to \infty$ when $U/J<36/\pi$ and 
557: in (3) that for $U/J<36/\pi$ one has $F^\prime(K)<0$: 
558: one may then conclude that the equation $F(K)=0$ has only 
559: one root for $U/J<36/\pi$. Finally, we derive an analytic expression 
560: for $F(K)$ holding for small $U/J$. 
561: 
562: From Eq.(\ref{radice_impr}) one can see that:\\
563: (1) for $U/J \to 0$ one has $2+\frac{\pi}{4}-\pi K=0$, 
564: from which $K=\frac{8+\pi}{4 \pi}$ and $K_B T_{BKT} \simeq 1.128 J$: 
565: indeed, for small $U/J$, one has that
566: \begin{displaymath}
567: \frac{X_u K^2}{2}- \frac{n^2}{2} \log 
568: \Bigl(1+\frac{X_u K^2}{n^2}\Bigr)=\frac{X_u K^2}{2}- 
569: \frac{n^2}{2} \Bigl(\frac{X_u K^2}{n^2}-\frac{X_u^2 K^4}{2 n^4}+
570:  \ldots\Bigr)=\frac{X_u^2 K^4}{4 n^2}-\ldots
571: \end{displaymath}
572: and thus 
573: \begin{displaymath}
574: \frac{1}{X_u K^2} g(K,X_u) \approx \frac{1}{X_u K^2} 
575: \sum\limits_{n=1}^{\infty} \frac{X_u^2 K^4}{4 n^2} \to 0.
576: \end{displaymath}
577: In the same way, $F(K) \to 2+\frac{\pi}{4}$ for $K \to 0$.\\
578: 
579: (2) $F(K) \to -\infty$ per $K \to \infty$ for $x_u<3/2$ 
580: (where $x_u=\pi^2 X_u/24=\pi U/24J$): indeed for large $K$ 
581: \begin{displaymath}
582: F(K) \approx 2-\pi K \Bigl(1-\frac{1}{48 x_u K^3} \int\limits_{0}^{2 \sqrt{6} 
583: \sqrt{x_u} K} dx x^2 \Bigr)=2-\pi K \Bigl(1-\frac{\sqrt{6}}{3} 
584: \sqrt{x_u}\Bigr):
585: \end{displaymath}
586: then, for $K \to \infty$, $F(K) \to -\infty$ with $x_u<3/2$ 
587: and $F(K) \to \infty$ with $x_u>3/2$.\\ 
588: 
589: (3) For $x_u<3/2$, it is $F^\prime (K)<0$ for $K>0$: 
590: indeed 
591: \begin{equation}
592: F^\prime(K)=-\pi-\frac{2 \pi}{X_u K^3} \, g(K,X_u)+\frac{\pi^2 X_u^{1/2}}
593: {2} \Bigl[\coth(\pi X_u^{1/2} K)- \frac{1}{\pi X_u^{1/2} K}\Bigr].
594: \label{21}
595: \end{equation}
596: Since $\lim_{x \to 0} \Bigl(\coth x-\frac{1}{x}\Bigr)=0$, 
597: one gets
598: \begin{displaymath}
599: \lim_{K \to 0} F^\prime(K)=-\pi.
600: \end{displaymath}
601: For large $K$ one finds
602: \begin{displaymath}
603: F^\prime(K) \approx -\pi+\frac{\pi^2}{6} X_u^{1/2}.
604: \end{displaymath}
605: Therefore one has that $\lim_{K \to \infty} F^\prime(K)<0$ for 
606: $X_u<36/\pi^2$, and $\lim_{K \to \infty} F^\prime(K)>0$ for 
607: $X_u>36/\pi^2$. 
608: Using similar arguments, one can show that, for $K>0$, it is 
609: \begin{equation}
610: F^\prime(K)<-\pi+\frac{\pi^2}{6} X_u^{1/2}.
611: \label{21d}
612: \end{equation}
613: which implies that, for $X_u<36/\pi^2$, one has $F^\prime(K)<0$. 
614: The behavior of $F(K)$ and $F^\prime(K)$ for three values of $U/J$ 
615: respectively smaller than, equal to and larger than $36/\pi$ is 
616: plotted in Figs. 2 and 3.
617: 
618: One may also obtain an analytic approximation for $F(K)$ holding 
619: for $U/J \ll 1$ by putting 
620: $z=\sqrt{X_u} K$ and using $\log (1+z)=\sum\limits_{j=1}^{\infty} 
621: \frac{(-1)^{j+1} z^j}{j}$ for $\mid z \mid<1$: one has 
622: \begin{equation}
623: g(K,X_u)=\sum\limits_{j=2}^{\infty} \frac{(-1)^j z^{2j}}{2j} \zeta(2j-2)
624: \label{appross}
625: \end{equation}
626: where $\zeta (j)=\sum\limits_{n=1}^{\infty} \frac{1}{n^j}$ is the
627: Riemann zeta-function. One finds
628: \begin{equation}
629: \sum\limits_{j=2}^{\infty} \frac{(-1)^j z^{2j}}{2j} \zeta(2j)=
630: \frac{1}{2} \Bigl\{\frac{\pi^2 z^2}{6}+\log \frac{\pi z}{\sinh \pi z}\Bigr\}.
631: \label{appross1}
632: \end{equation} 
633: For large $n$ one has the recurrence relation \cite{abramowitz72}
634: \begin{displaymath}
635: \zeta(n+1) \simeq \frac{1}{2} [1+ \zeta(n)]. 
636: \end{displaymath}
637: Substituting back in Eq.(\ref{appross}) and using (\ref{appross1}), 
638: one has
639: \begin{displaymath}
640: g(K,X_u) \approx X_u K^2 \Bigl(\frac{\pi^2}{3}-{3 \over 2}\Bigr)+
641: 2 \log \frac{\pi X_u^{1/2} K}{\sinh \pi X_u^{1/2} K}+
642: {3 \over 2} \log (1+X_u K^2)
643: \end{displaymath}
644: and finally 
645: \begin{equation}
646: F \approx 2-\pi K 
647: \Bigg[1-\frac{1}{4 K}-\frac{1}{X_u 
648: K^3} \Bigg( X_u K^2 \bigg(\frac{\pi^2}{3}-{3 \over 2}\bigg)+
649: 2 \log \frac{\pi X_u^{1/2} K}{\sinh \pi X_u^{1/2} K}+
650: {3 \over 2} \log (1+X_u K^2) \Bigg) \Bigg].
651: \label{approxx}
652: \end{equation}
653: 
654: \begin{thebibliography}{10}
655: 
656: \bibitem{trombettoni03} A. Trombettoni, A. Smerzi, and P. Sodano, 
657: New J. Phys. {\bf 7}, 57 (2005).
658: 
659: \bibitem{greiner01} 
660: M. Greiner {\em et al.}, Phys. Rev. Lett. {\bf 87}, 160405 (2001); 
661: Appl. Phys. B {\bf 73}, 769 (2001).
662: 
663: \bibitem{BKT} V. L. Berezinskii, Sov. Phys. JETP {\bf 32}, 
664: 493 (1971); J. M. Kosterlitz and D. J. Thouless, 
665: J. Phys. C {\bf 6}, 1181 (1973); J. M. Kosterlitz, 
666: {\em ibid.} {\bf 7}, 1046 (1974).
667: 
668: \bibitem{minnaghen87} P. Minnaghen, Rev. Mod. Phys. {\bf 59}, 
669: 1001 (1987).
670: 
671: \bibitem{kadanoff00} L. P. Kadanoff, {\em Statistical Physics: Statics, 
672: Dynamics and Renormalization}, (Singapore, World Scientific, 2000), 
673: Chapt.s 16-17 and reprints therein.
674: 
675: \bibitem{tobochnik79} J. Tobochnik and G. V. Chester, 
676: Phys. Rev. B {\bf 20}, 3761 (1979). 
677: 
678: \bibitem{leoncini98} X. Leoncini, A. D. Verga, and S. Ruffo, 
679: Phys. Rev. E {\bf 57}, 6377 (1998).
680: 
681: \bibitem{jaksch98} D. Jaksch, C. Bruder, 
682: J. I. Cirac, C. W. Gardiner, and P. Zoller,  
683: Phys. Rev. Lett. {\bf 81}, 3108 (1998).
684: 
685: \bibitem{gupta88} R. Gupta, J. DeLapp, 
686: G. G. Batrouni, G. C. Fox, C. F. Baillie, and J. Apostolakis, 
687: Phys. Rev. Lett. {\bf 61}, 1996 (1988).  
688: 
689: \bibitem{sondhi97} S. L. Sondhi, S. M. Girvin, 
690: J. P. Carini, and D. Shahar, 
691: Rev. Mod. Phys. {\bf 69}, 315 (1997).
692: 
693: \bibitem{fazio01} R. Fazio and H. van der Zant, 
694: Phys. Rep. {\bf 355}, 235 (2001).
695: 
696: \bibitem{simanek94} E. Sim\`anek, 
697: {\em Inhomogeneous Superconductors},
698: Oxford University Press, New York, 1994.
699: 
700: \bibitem{martinoli00} P. Martinoli and C. Leeman, 
701: J. Low Temp. Phys. {\bf 118}, 699 (2000). 
702: 
703: \bibitem{jose84} J. V. Jos\'e, 
704: Phys. Rev. B {\bf 29}, 2836 (1984).
705: 
706: \bibitem{fisher89} M. P. A. Fisher, 
707: P. B. Weichman, G. Grinstein, and D. S. Fisher, 
708: Phys. Rev. B {\bf 40}, 546 (1989). 
709: 
710: \bibitem{anglin01} J. R. Anglin, P. Drummond, and A. Smerzi, 
711: Phys. Rev. A {\bf 64}, 063605 (2001).
712: 
713: \bibitem{vanotterlo93}
714: A. van Otterlo, K.-H. Wagenblast, R. Fazio and G. Sch\"{o}n,
715: Phys. Rev. B {\bf 48}, 3316 (1993).
716: 
717: \bibitem{kopec00}
718: T. K. Kope\'c and J.V. Jos\'e,
719: Phys. Rev. Lett. {\bf 84}, 749 (2000).
720: 
721: \bibitem{grignani00}
722: G. Grignani, A. Mattoni, P. Sodano, and A. Trombettoni,
723: Phys. Rev. B {\bf 61}, 11676 (2000).
724: 
725: \bibitem{jacobs84} L. Jacobs, J. V. Jos\'e, and M. A. Novotny, 
726: Phys. Rev. Lett. {\bf 53}, 2177 (1984).
727: 
728: \bibitem{rojas96} C. Rojas and J. V. Jos\'e, 
729: Phys. Rev. B {\bf 54}, 12361 (1996).
730: 
731: \bibitem{capriotti03} L. Capriotti, A. Cuccoli, A. Fubini, 
732: V. Tognetti, and R. Vaia, Phys. Rev. Lett. {\bf 91}, 247004 (2003). 
733: 
734: \bibitem{alet03} F. Alet and 
735: E. S. S$\o$rensen, Phys. Rev. E {\bf 67}, 015701(R) (2003). 
736: 
737: \bibitem{alsaidi04} W. A. Al-Saidi and D. Stroud, 
738: Physica C {\bf 402}, 216-222 (2004). 
739: 
740: \bibitem{kim97} B. J. Kim, J. Kim, S. Y. Park, and M. Y. Choi, 
741: Phys. Rev. B {\bf 56}, 395 (1997).
742: 
743: \bibitem{cuccoli00} A. Cuccoli, A. Fubini, V. Tognetti, and 
744: R. Vaia,  
745: Phys. Rev. B {\bf 61}, 11289 (2000).
746: 
747: \bibitem{abramowitz72} M. Abramowitz and I. Stegun, 
748: {\em Handbook of Mathematical Functions}, U.S. Government Printing Office, 
749: Washington, DC, 1972.
750: 
751: \bibitem{kleinert03} H. Kleinert, S. Schmidt, and A. Pelster, 
752: Phys. Rev. Lett. {\bf 93}, 160402 (2004). 
753: 
754: \end{thebibliography}
755: 
756: %\end{multicols}
757: 
758: \begin{figure}
759: \centerline{\psfig{figure=fig1.ps,width=60mm,angle=270}}
760: \caption{The BKT critical temperature $T_{BKT}(U)$ 
761: (in units of $T_{BKT}(0)$) as a function  
762: of $U/J$. Empty circles: numerical solution of Eq.(\ref{radice_impr}); 
763: solid line: Eq.(\ref{sviluppo}).}
764: \label{fig1}
765: \end{figure}
766: 
767: \begin{figure}
768: \centerline{\psfig{figure=fig2.ps,width=60mm,angle=270}}
769: \caption{$F(K)$ for $U/J=1$ (solid line), $36/\pi$ (dashed line) 
770: and $15$ (dotted line). The dot-dashed line is the analytic approximation 
771: (\ref{approxx}) for $F$ holding for $U/J \ll 1$. For 
772: $U/J=36/\pi$ $F$ asymptotically tends to $0$.}
773: \label{fig2}
774: \end{figure}
775: 
776: \begin{figure}
777: \centerline{\psfig{figure=fig3.ps,width=60mm,angle=270}}
778: \caption{$F^\prime(K)$ 
779: for $U/J=1$ (solid line), $36/\pi$ (dashed line) 
780: and $15$ (dotted line).}
781: \label{fig3}
782: \end{figure}
783: 
784: \end{document}
785: 
786: 
787: 
788: 
789: 
790: 
791: 
792: 
793: 
794: 
795: 
796: 
797: