1: \documentclass[aps,twocolumn,superscriptaddress,showpacs,floats]{revtex4}
2: %\documentclass[aps,twocolumn,superscriptaddress,floats]{revtex4}
3: %\documentclass[aps,preprint,showpacs,floatfix,superscriptaddress]{revtex4}
4: %\documentclass[aps,twocolumn,groupedaddress,superscriptaddress,showpacs,floats,amssymb]{revtex4}
5:
6: \usepackage{amsmath}
7: \usepackage{amsfonts}
8: \usepackage{amssymb}
9: \usepackage{euscript}
10: \usepackage{bm}
11: \usepackage[dvips]{graphicx}
12:
13: %\topmargin=5pt
14:
15: \begin{document}
16:
17: \title{On-site number statistics of ultracold lattice bosons}
18:
19: \author{Barbara Capogrosso-Sansone}
20: \affiliation{Department of Physics, University of
21: Massachusetts, Amherst, MA 01003}
22: \author{Evgeny Kozik}
23: \affiliation{Department of Physics, University of
24: Massachusetts, Amherst, MA 01003}
25: \author{Nikolay Prokof'ev}
26: \affiliation{Department of Physics, University of Massachusetts,
27: Amherst, MA 01003}
28: \affiliation{Russian Research Center
29: ``Kurchatov Institute'', 123182 Moscow, Russia}
30: \author{Boris Svistunov}
31: \affiliation{Department of Physics, University of Massachusetts,
32: Amherst, MA 01003} \affiliation{Russian Research Center
33: ``Kurchatov Institute'', 123182 Moscow, Russia}
34:
35: \begin{abstract}
36: We study on-site occupation number fluctuations in a system of
37: interacting bosons in an optical lattice. The ground-state
38: distribution is obtained analytically in the limiting cases of
39: strong and weak interaction, and by means of exact Monte Carlo
40: simulations in the strongly correlated regime. As the interaction
41: is increased, the distribution evolves from Poissonian in the
42: non-interacting gas to a sharply peaked distribution in the
43: Mott-insulator (MI) regime. In the special case of large
44: occupation numbers, we demonstrate analytically and check
45: numerically that there exists a wide interval of interaction
46: strength, in which the on-site number fluctuations remain Gaussian
47: and are gradually squeezed until they are of order unity near the
48: superfluid (SF)-MI transition. Recently, the on-site number
49: statistics were studied experimentally in a wide range of lattice
50: potential depths [Phys. Rev. Lett. \textbf{96}, 090401 (2006)]. In
51: our simulations, we are able to directly reproduce experimental
52: conditions using temperature as the only free parameter.
53: Pronounced temperature dependence suggests that measurements of
54: on-site atom number fluctuations can be employed as a reliable
55: method of thermometry in both SF and MI regimes.
56: \end{abstract}
57:
58: \pacs{32.80.Pj, 39.90.+d, 67.40.Db}
59:
60: \maketitle
61:
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: \section{Introduction}
64: \label{sec:intro}
65: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66:
67: Experiments on ultracold atoms trapped by an optical potential
68: \cite{Bloch_SFMI,Ospelkaus,Bloch_spin_dynamics,Ketterle_spatial,Bloch_SF3D},
69: offer a unique possibility to explore fundamental properties of
70: strongly correlated quantum many-body systems allowing virtually
71: unlimited control over the microscopic Hamiltonian parameters
72: (see, e.g., \cite{Zoller} and references therein). System
73: flexibility along with relatively long decoherence times puts it
74: among top candidates for implementation of quantum information
75: algorithms \cite{Zoller}. In atomic interferometry \cite{Berman},
76: ultracold gases in the strongly correlated regime allow to achieve
77: accuracies below the standard shot noise limit
78: \cite{Burnett,Rodriquez_interferometry}. A new exciting
79: application, ``atomtronics'', is suggested by a remarkable analogy
80: between the physics of ultracold atoms in optical lattices and
81: that of electrons in crystals \cite{Seaman}. With the current
82: experimental technique, it seems plausible to produce such basic
83: atomtronic devises as diodes and bipolar junction transistors,
84: which serve as building blocks for amplifiers and logic gates.
85:
86: On the fundamental physics side, strongly interacting lattice
87: bosons provide insight into the nature of quantum phase
88: transitions. In particular, these systems are an accurate
89: experimental realization of the Bose-Hubbard model
90: \cite{Zoller98},
91: \begin{equation}
92: H=-t \sum_{<i,j>} a^{\dagger}_i a^{\:}_j + \frac{1}{2}U \sum_i
93: \hat{n}_i(\hat{n}_i-1) + \sum_i \epsilon_i \hat{n}_i,
94: \label{Bose-Hubbard-Hum}
95: \end{equation}
96: where $\hat{n}_i=a^{\dagger}_i a_i$ is the number operator on site
97: $i$; $a^{\dagger}_i$ and $a_i$ respectively create and
98: annihilate bosons on lattice sites, and $<i,j>$ denotes the sum
99: over nearest neighbors. The first term describes tunneling between
100: neighboring potential wells of the optical lattice, the second
101: term is the effective repulsion within a well, while the last term
102: is due to an additional smooth space-varying potential, such as,
103: e.g., a magnetic trap. This system exhibits a transition between
104: superfluid (SF) and Mott-insulator (MI) groundstates governed by
105: the competition of atom mobility and interatomic interaction
106: \cite{Fisher}. The SF-MI phase transition is a topic of intense
107: current research both theoretically
108: \cite{Batrouni,Bloch_theor,Isacsson} and experimentally
109: \cite{Bloch_spatial,Gerbier,Clark}. In typical experiments, a
110: prepared Bose-Einstein condensate of ultracold atoms is driven to
111: the strongly correlated regime by means of its adiabatic loading
112: into a periodic optical potential (optical lattice) induced by the
113: a.c. Stark effect of interfering laser beams. The mobility of
114: atoms, namely the hopping $t$, and their interaction $U$ are
115: controlled by the depth of the optical potential, i.e. by the
116: laser intensity. In relatively shallow potentials, atoms are
117: delocalized over the entire lattice giving rise to long-range
118: coherence and thus \cite{Penrose_Onsager} to superfluid behavior.
119:
120: If the lattice filling is commensurate, i.e. there is on average
121: an integer number of atoms per lattice site, increasing the
122: lattice depth brings the system across the phase transition to the
123: MI state, which is characterized by zero compressibility and a gap
124: in the spectrum of elementary excitations. Here, the key
125: observable is the atom interference pattern obtained upon
126: releasing the atoms and letting the atom cloud expand for a
127: transient time of flight \cite{Bloch_SFMI}. Phase correlations
128: between lattice sites result in pronounced interference peaks
129: smeared by the finite correlation length in the MI regime. At the
130: same time phase coherence is fundamentally connected with the
131: statistics of atom number fluctuations on lattice sites. In
132: particular, one expects the on-site atom number to behave as a
133: canonically conjugate variable with respect to the phase field
134: and, therefore, experience suppressed fluctuations in the MI
135: regime, analogous to the number squeezed states with
136: sub-Poissonian number fluctuations \cite{Orzel} widely studied in
137: quantum optics (see, e.g., \cite{quant_opt}). In practice, number
138: squeezed states are important for high-precision atomic
139: interferometry \cite{Berman}, where their use can potentially lead
140: to sensitivities limited only by the Heisenberg uncertainty
141: principle \cite{Caves,Burnett,Rodriquez_interferometry}, and for
142: atom-based quantum computing techniques \cite{Zoller}, where
143: unwanted number fluctuations necessitate correction procedures in
144: operation of quantum gates.
145:
146:
147: Until recently, the number distribution was not measured directly.
148: The situation changed with the development of the spin-oscillation
149: technique \cite{Bloch_spin_dynamics}, which is sensitive to the
150: number of atom pairs and works at arbitrary lattice depths
151: \cite{Gerbier}, and, most recently, the microwave spectroscopy
152: using atomic clock shifts \cite{Ketterle_spatial}. In
153: Ref.~\cite{Gerbier}, Gerbier \textit{et al.} observed a drastic
154: change of atom number statistics as the system of $^{87}$Rb atoms
155: was driven through the SF-MI transition. On the theoretical side,
156: apart from the recent mean-field calculation \cite{Lu-Yu}, a
157: comprehensive study of atom number fluctuations in the strongly
158: correlated regime is still missing.
159:
160:
161: In the present work, we attempt to close this gap by tackling the
162: problem both analytically and numerically. In
163: section~\ref{sec:homog}, we focus on the academic case of a
164: homogeneous square lattice in the thermodynamic limit and in the
165: limit of zero temperature in one-(1D), two-(2D) and three-(3D)
166: dimensions. At $U=0$, the ground state of an ideal Bose gas is a
167: Bose-Einstein condensate with characteristic Poisson distribution
168: of number fluctuations. In subsection~\ref{subsec:weak_coupling},
169: we consider the limit of weak interaction $\nu U/t \ll 1$, where
170: $\nu$ is the filling factor. This parameter region corresponds to
171: perturbative squeezing of the Poisson distribution, which, as a
172: function of $\nu U/t$, qualitatively depends on the space
173: dimension $d$. We consider the case of large filling factors $\nu
174: \gg 1$ separately in subsection~\ref{subsec:large_nu} because it
175: is qualitatively different from that of $\nu \sim 1$. Indeed, the
176: quantum nature of the SF-MI transition implies that the number
177: fluctuations in the vicinity of the transition must be of order
178: unity. At the same time, at $U=0$ the variance of the on-site
179: number distribution is $\sigma^2=\nu \gg 1$. Therefore, there
180: exists an extensive range of interactions (defined by the
181: condition $U/t \lesssim \nu$), in which the system remains
182: superfluid, but its on-site number distribution is drastically
183: squeezed before the SF-MI can take place. We show that, at $U/t
184: \ll \nu$, the on-site number statistics are Gaussian and derive
185: the variance $\sigma^2$ of the distribution, which scales as
186: $\sigma^2 \propto \sqrt{\nu t / U}$ at $1/\nu \ll U/t \ll \nu $ in
187: all dimensions. In subsection~\ref{subsec:inter_formula}, we
188: suggest an expression that interpolates $P(n)$ between the
189: limiting cases of small interaction and large occupation numbers,
190: which is found to properly describe $P(n)$ up to $U/t$ of the
191: order of the critical value $(U/t)_c$. The strong coupling limit,
192: $\nu t/U \ll 1$, at integer filling is considered in
193: subsection~\ref{subsec:strong_coupling}. In this limit, the system
194: is in the MI regime and the on-site number distribution is
195: governed by rare particle-hole fluctuations.
196:
197: We study the distribution in the strongly correlated regime
198: connecting the limiting cases by means of a direct numeric
199: simulation of the model~(\ref{Bose-Hubbard-Hum}) at $\nu=1$ in
200: subsection~\ref{subsec:numerics}. The distribution of the on-site
201: occupation number is a local property, revealing no critical
202: features at the transition. However, the strongly correlated
203: region is responsible for the crossover that changes the
204: statistics qualitatively. As the interaction strength is
205: increased, we observe a gradual squeezing of the on-site number
206: distribution and the emergence of the symmetry between particle-
207: and hole-like fluctuations, characteristic of a MI. In this
208: section, we also present numerical data for the case of large
209: filling factors and demonstrate that the analysis of
210: subsection~\ref{subsec:large_nu} is applicable already at $\nu=5$.
211:
212: The worm algorithm quantum Monte Carlo (MC) technique
213: \cite{Worm_Algorithm} allows us to simulate system sizes that are
214: currently realized in experiments without any approximations,
215: including the particle number. The results of a direct numeric
216: simulation of the experimental setup of Ref.~\cite{Gerbier} are
217: presented in section \ref{sec:exp}. With the lattice parameters
218: fixed by the experiment, we are left with temperature as the only
219: free parameter. Due to a smooth confining potential present on top
220: of the optical lattice, the number distribution is not
221: homogeneous. We focus on an integral characteristic of the number
222: distribution, namely, the fraction of atoms found on lattice sites
223: with occupation $n$, which can be systematically measured
224: experimentally. This quantity has a pronounced temperature
225: dependence in both SF and MI regimes.
226:
227: The problem of thermometry in optical lattices, especially in the
228: MI regime, is a long standing one. The ability to control the
229: temperature is of crucial importance for applications that rely on
230: the peculiar properties of a MI state. At $T=0$ fluctuations are
231: of quantum nature and can be efficiently controlled externally
232: through the lattice parameters, whereas temperatures comparable to
233: the Mott excitation gap destroy the insulating state by activating
234: particle-hole excitations. At the moment, there are no
235: experimental techniques to measure the temperature of a strongly
236: interacting system. Unlike in the weakly interacting regime, where
237: the temperature can be straightforwardly extracted from the
238: momentum distribution (e.g., from the interference pattern of
239: matter waves or from the condensate fraction observed after the
240: trap is released and the gas expands freely), in the strongly
241: correlated regime, both temperature and interatomic interaction
242: are responsible for filling the higher momentum states making
243: standard absorption imaging techniques inapplicable.
244:
245: The idea of using occupation number distributions to estimate the
246: temperature was explored in Ref.~\cite{Prokofev}, where in was
247: argued that the temperature dependence of the total number of
248: pairs and their spatial distribution (in traps) provides a
249: sensitive method of thermometry deeply in the MI phase at energies
250: smaller than the interatomic interaction, but much larger than the
251: effective hopping between the sites. [Recently, it has become
252: possible to directly sample spatially-resolved number
253: distributions by spin changing collisions \cite{Bloch_spatial},
254: microwave spectroscopy \cite{Ketterle_spatial}, and the scanning
255: electron microscope \cite{Gericke_scan_microscp} promising a
256: complete practical realization of this method.] In this paper, we
257: perform thermometry of the system in all strongly correlated
258: regimes by comparing experimental data and numerical results for
259: the statistics of occupation numbers. More specifically, we
260: compare numerically calculated fraction of pairs ($n=2$) with that
261: measured in Ref.~\cite{Gerbier} across the SF-MI transition
262: estimating the range of experimental temperatures. The accuracy of
263: this method is mainly limited by the error bars of the
264: experimental data and by the range of applicability of the
265: Bose-Hubbard model.
266:
267: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
268: \section{ Homogeneous Lattice} \label{sec:homog}
269: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
270:
271: In this section, we assume that there is no space-varying
272: potential on top of the optical lattice, and set $\epsilon_i
273: \equiv \epsilon_0$. Let $N_s$ be the number of lattice sites and
274: $N$ be the total number of particles. The goal of this section is
275: to obtain the ground state probability $P_n$ to detect $n$
276: particles on a given lattice site in the limit of $N_s,N
277: \rightarrow \infty$, at a fixed filling factor $\nu=N/N_s$.
278: Mathematically, $P_n$ can be defined as
279: \begin{equation}
280: P_n (U/t)= \sum_{\{n_{i \ne 1}\} }\big|\; \langle n_1=n, \{n_{i
281: \ne 1}\} | \:|\Psi_{U/t} \rangle \big| ^2 \; , \label{Pn}
282: \end{equation}
283: where $| \Psi_{U/t} \rangle$ is the many-body ground state
284: wavefunction, and $| \{n_{i}\} \rangle$ are Fock states.
285:
286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
287: \subsection{Weak coupling limit} \label{subsec:weak_coupling}
288: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
289:
290: At sufficiently small $U/t$, the relevant representation of
291: Eq.~(\ref{Bose-Hubbard-Hum}) is obtained by the diagonalization of
292: the kinetic energy term with the canonical transformation
293: $a_i=N_s^{-1/2}\sum_\mathbf{k} b_\mathbf{k} \exp(i 2\pi\:
294: \mathbf{k} \, \mathbf{r}_i/L)$ (periodic boundary conditions are
295: assumed), where $L=N_s^{1/d}$ is the linear system size,
296: $\mathbf{r}_i$ and $\mathbf{k}$ are respectively the position of
297: the site $i$ and a quasi-momentum in $d$ dimensions with
298: integer-valued components, $-(L-1)/2 \leq r_{i\: \mu},\: k_{\mu}
299: \leq (L-1)/2$, $\mu=1,\, ... \, ,d$. The result is
300: \begin{equation}
301: H=\sum_{k} \varepsilon_{\mathbf{k}} b^{\dagger}_{\mathbf{k}}
302: b^{\;}_{\mathbf{k}} + \frac{U}{2N_s} \sum_{\mathbf{k}_1 +
303: \mathbf{k}_2= \mathbf{k}_3 + \mathbf{k}_4}
304: b^{\dagger}_{\mathbf{k}_1}b^{\dagger}_{\mathbf{k}_2}
305: b^{\:}_{\mathbf{k}_3}b^{\:}_{\mathbf{k}_4}
306: \label{Bose-Hubbard-Hum_momentum_space}
307: \end{equation}
308: with $\varepsilon_{\mathbf{k}}\, =\,
309: 2t\sum_{\mu=1}^{d}\,[1-\cos(2\pi k_{\mu}/L)]$. At $U=0$ the ground
310: state of the Hamiltonian (\ref{Bose-Hubbard-Hum_momentum_space})
311: is a pure Bose-Einstein condensate, which can be expressed as a
312: coherent state, $ |\Psi_{(U/t)=0}\rangle = \exp(\sqrt{N}
313: b_0^{\dagger} - N/2) | 0 \rangle$. Then, transforming the
314: wavefunction back to the on-site representation in terms of $\{
315: a_i \}$ yields the Poisson distribution for the probability to
316: find $n$ particles on a given site:
317: \begin{equation}
318: P^{(0)}_n=e^{-\nu} \, \frac{{\nu}^{n}}{n!} \, .
319: \label{Pn_Ideal_gas}
320: \end{equation}
321: At a finite, but small, $U$ we employ the standard Bogoliubov
322: method \cite{LLStatMech2} of separating the system into the
323: classical-field condensate part and non-condensate particles
324: interacting with it, omitting the terms of the third and forth
325: order with respect to the non-condensate operators. In this
326: approximation, the Hamiltonian
327: (\ref{Bose-Hubbard-Hum_momentum_space}) is reduced to a bilinear
328: in $b_{\mathbf{k}}$ and $b_{\mathbf{k}}^{\dagger}$ form and
329: diagonalized by the canonical transformation
330: $c_{\mathbf{k}}=u_\mathbf{k} b_{\mathbf{k}} + v_\mathbf{k}
331: b_{-\mathbf{k}}^{\dagger}$, where
332: \begin{gather}
333: u_\mathbf{k} = \left[(\varepsilon_\mathbf{k} + \nu U)/2\omega_\mathbf{k} + 1/2 \right]^{1/2}, \notag \\
334: v_\mathbf{k} = \left[(\varepsilon_\mathbf{k} + \nu U)/2\omega_\mathbf{k} - 1/2 \right]^{1/2}, \notag \\
335: \omega_\mathbf{k}=\left[\varepsilon_\mathbf{k}^2 + 2 \nu
336: \varepsilon_\mathbf{k} U\right]^{1/2}. \label{UkVk}
337: \end{gather}
338: The ground-state wavefunction is then obtained from the equation
339: $c_{\mathbf{k}}\:|\Psi_{U/t} \rangle = 0$ for all ${\mathbf{k}}
340: \neq 0$ and has the form
341: \begin{equation}
342: |\Psi_{U/t}\rangle = C \exp\Biggl[\sqrt{N_0} b_0^{\dagger} -
343: \frac{N_0}{2} - \frac{1}{2}
344: \sum_{\mathbf{k}\neq0}\frac{v_\mathbf{k}}{u_\mathbf{k}}b^{\dagger}_{\mathbf{k}}b^{\dagger}_{-\mathbf{k}}\Biggr]
345: | 0 \rangle, \label{Psi_Bogolubov}
346: \end{equation}
347: where $C$ is the normalization factor and $N_0=N -
348: \sum_{\mathbf{k}\neq0} \langle
349: b^{\dagger}_{\mathbf{k}}b_{\mathbf{k}}\rangle$ is the number of
350: condensate particles.
351:
352:
353: Now we can express $|\Psi_{U/t}\rangle$ in terms of the on-site
354: operators,
355: \begin{gather}
356: |\Psi_{U/t}\rangle = C \exp\Biggl[ \sum_i \left(\sqrt{\nu_0}
357: a_i^{\dagger} - \frac{\nu_0}{2} \right) - \sum_{i,j} S_{ij}
358: a^{\dagger}_i a^{\dagger}_j \Biggr] | 0 \rangle \;,
359: \label{Psi_onsite}
360: \end{gather}
361: with $\nu_0=N_0/N_s$ and
362: \begin{equation}
363: S_{i j} = \frac{1}{2N_s}
364: \sum_{\mathbf{k}\neq0}\frac{v_\mathbf{k}}{u_\mathbf{k}} e^{i 2 \pi
365: \mathbf{k} (\mathbf{r}_i - \mathbf{r}_j)/L}. \label{Sij}
366: \end{equation}
367: In this form, the wavefunction can be used to obtain the on-site
368: number distribution in the whole range of $U/t \ll \nu$ by a
369: straightforward application of Eq.~(\ref{Pn}).
370:
371: We derive a closed-form expression for $P(n)$ in the limiting case
372: of
373: \begin{equation}
374: \alpha \, = \, \nu \, U/t \, \ll 1, \label{alpha}
375: \end{equation}
376: which corresponds to the range of sufficiently weak squeezing of
377: $P(n)$ allowing us to consider only the first correction to the
378: Poisson distribution in the leading power of $\alpha$.
379: Mathematically, Eq.~(\ref{alpha}) implies that $S_{ij} \ll 1$. If,
380: in addition, $S_{ij}$ is short-range, i.e. it decays at distances
381: $|\mathbf{r}_i - \mathbf{r}_j| \sim 1$, which implies that the
382: leading correction is insensitive not only to the system size, but
383: also to the value of the healing length $ \propto \alpha^{-1/2}$,
384: then we can expand the exponential in Eq.~(\ref{Psi_onsite}) in
385: powers of $S_{ij}$. Rather straightforward but lengthy algebra
386: yields the distribution in the form
387: \begin{equation}
388: P_n=P^{(0)}_n- \frac{\nu \,\lambda(\alpha)}{2} \,
389: [P^{(0)}_n-2P^{(0)}_{n-1}+P^{(0)}_{n-2}] \:,
390: \label{Pn_weak_coupling}
391: \end{equation}
392: where $\lambda$ is an interaction-dependent squeezing parameter
393: and $P^{(0)}_n$ is given by Eq.~(\ref{Pn_Ideal_gas}) assuming
394: $P^{(0)}_n=0$ for $n < 0$. Identically,
395: \begin{equation}
396: P_n = P^{(0)}_n \left(1 + \frac{\lambda(\alpha)}{2} \, \left[
397: \frac{n - (n-\nu)^2}{\nu}\right] \right) . \label{Pn_relative}
398: \end{equation}
399:
400: Let us postpone writing an explicit expression for $\lambda$ and
401: discuss the underlying assumptions leading to
402: Eqs.~(\ref{Pn_weak_coupling}),(\ref{Pn_relative}). It turns out
403: that $S_{ij}$ is local only in 3D, where the main contribution to
404: the integral in Eq.~(\ref{Sij}) comes from large momenta close to
405: the edge of the Brillouin zone, $|\mathbf{k}| \sim L/2$. In 2D,
406: the integral has a logarithmic divergency at low momenta in the
407: limit $\alpha \to 0$, while, in 1D, the dominant contribution to
408: $S_{ij}$ comes from small momenta, meaning that in both 1D and 2D
409: cases we can not rely on the perturbative expansion of the wave
410: function. Nevertheless,
411: Eqs.~(\ref{Pn_weak_coupling}),(\ref{Pn_relative}) are still valid
412: in 1D and 2D, since the functional form of $P(n)$ must be the same
413: in all dimensions. To prove this statement, we note that $P(n)$ is
414: unambiguously determined by its characteristic function
415: $\chi(t)=\langle \exp(i \, t \, a_i^{\dagger} a_i) \rangle$, which
416: can be used to generate all moments of $P(n)$. The function
417: $\chi(t)$ can be calculated explicitly as a series expansion in
418: powers of $(i t a_i^{\dagger}a_i)$. Since averages of bosonic
419: quasiparticle operators in the Bogoliubov theory obey Wick's
420: theorem, each term in the series is a function of $\zeta \equiv
421: \langle a_i^{\dagger} a_i \rangle$ and $\zeta' \equiv \langle a_i
422: a_i \rangle$. Therefore, $\chi(t) = \chi(t, \zeta, \zeta')$,
423: meaning that all physical parameters, including the space
424: dimension, enter $P(n)$ only through $\zeta$ and $\zeta'$. In
425: Eqs.~(\ref{Pn_weak_coupling}),(\ref{Pn_relative}), this
426: combination determines $\lambda(\alpha)$.
427:
428: To determine $\lambda(\alpha)$ in all dimensions were note that it
429: is directly related to the dispersion of $P(n)$ in
430: Eq.~(\ref{Pn_relative})
431: \begin{equation}
432: \sigma^2 \equiv \langle (n-\nu)^2 \rangle = \nu \, (1-\lambda).
433: \label{sigma-lambda}
434: \end{equation}
435: On the other hand, $\sigma^2 = \langle n_i^2 \rangle - \nu^2 =
436: \langle a_i^{\dagger}a^{\;}_i a_i^{\dagger}a^{\;}_i \rangle -
437: \nu^2$ can be calculated in a standard way by replacing $a_i$ with
438: their expressions in terms of the Bogoliubov modes and the
439: classical-field condensate part, i.e.
440: $a_i=\sqrt{\nu_0}+N_s^{-1/2}\sum_{\mathbf{k} \neq 0} [u_k
441: c_\mathbf{k}-v_k c^{\dagger}_{-\mathbf{k}}] \exp(i 2\pi\:
442: \mathbf{k} \, \mathbf{r}_i/L)$. Then, an application of the Wick's
443: theorem along with $\langle c_{\mathbf{k}}
444: c^{\dagger}_{\mathbf{k}'} \rangle =
445: \delta_{\mathbf{k},\mathbf{k}'}$ and $\langle c_{\mathbf{k}}
446: c_{\mathbf{k}'} \rangle = 0$ gives
447: \begin{equation}
448: \sigma^2=\frac{\nu}{N_s} \sum_{\mathbf{k} \neq 0}
449: \frac{\varepsilon_{\mathbf{k}}}{\omega_{\mathbf{k}}} \; .
450: \label{sigma-general-zeroT}
451: \end{equation}
452:
453: Strictly speaking, Eq.~(\ref{sigma-general-zeroT}) is valid as
454: long as the number of non-condensed particles is small (the 1D
455: case is special in this regard and is further discussed below),
456: $(N-N_0)/N \ll 1$, which implies $U/t \ll \nu^{(2-d)/d}$ at small
457: $\nu$ and $U/t \ll \nu$ at large $\nu$. In the latter case, the
458: distribution change can be non-perturbative since $\lambda \sim 1$
459: is typical in this parameter regime, which will be discussed in
460: more detail in the next subsection.
461:
462: Thus, we arrive at the following result
463: \begin{equation}
464: \lambda=\frac{1}{N_s} \sum_{\mathbf{k} \neq 0} \big[ 1-
465: \frac{\varepsilon_{\mathbf{k}}}{\omega_{\mathbf{k}}} \big] \; .
466: \label{lambda-general}
467: \end{equation}
468: The asymptotic behavior of $\lambda(\alpha \to 0)$ qualitatively
469: depends on the dimensionality. In 1D, the main contribution to the
470: integral (\ref{lambda-general}) comes from low momenta resulting
471: in
472: \begin{equation}
473: \lambda_{(d=1)} (\alpha) \, \to\, \frac{\sqrt{2}}{\pi}\,
474: \sqrt{\alpha}. \label{lambda_1D}
475: \end{equation}
476: In 3D, there is no low-momentum singularity at $\alpha \to 0$, and
477: the asymptotic expression is linear in $\alpha$:
478: \begin{equation}
479: \lambda_{(d=3)} (\alpha) \, \to \, \frac{B}{2 \pi^3}\; \alpha,
480: \label{lambda_3D}
481: \end{equation}
482: \begin{equation}
483: B\,=\, \int_0^{\pi}\!\!\! \int_0^{\pi}\!\!\!\int_0^{\pi}
484: \frac{dz_1\, dz_2\, dz_3}{\sum_{\mu=1}^3 (1-\cos z_{\mu})}\,
485: \approx \, 15.673. \label{B}
486: \end{equation}
487: In 2D, the low-momentum singularity is logarithmic
488: \begin{equation}
489: \lambda_{(d=2)} (\alpha) \; \to \; {\ln (C/\alpha)\over\ 4 \pi}\;
490: \alpha, \;\;\;\;\; C\, \approx\, 23.54 \label{lambda_2D}
491: \end{equation}
492:
493:
494: A comment is in order here concerning the derivation procedure for
495: the 1D case. Formally, the condensate fraction is zero in the
496: macroscopic limit even at $T=0$, while the derivation is based on
497: the assumption that almost all the particles are Bose condensed.
498: Nevertheless, our final results for the probabilities are valid
499: even in 1D, and the generic reasoning---based on the notion of
500: quasicondensate---leading to this conclusion is as follows
501: \cite{KSS}. The quasicondensate correlation properties
502: characteristic of a weekly interacting 1D gas at $T=0$ imply two
503: different correlation radii, $r_c$ and $R_c$, $R_c \gg r_c$. Here
504: $r_c$ defines the length scale upon which the system can be
505: considered as macroscopic, while $R_c$ is the length at which
506: (quantum) fluctuations of phase are of order unity. If the system
507: size $L$ is such that
508: \begin{equation}
509: r_c \ll L \ll R_c \; , \label{L}
510: \end{equation}
511: then the system is macroscopic with respect to all {\it local}
512: properties, while still featuring a genuine condensate. (In a 1D
513: weakly interacting system at $T=0$ the density of this condensate
514: is close to the total density of particles.) Hence, for all local
515: properties, including the ones discussed in the present paper, one
516: can assume, without loss of generality, that the system size is
517: finite and satisfies the condition (\ref{L}). It should be
518: emphasized that this assumption is {\it implicit} and is used
519: exclusively to justify the derivation procedure. Otherwise, it
520: does not lead to any explicit dependence of final answers on $L$.
521: Indeed, the first inequality in Eq.~(\ref{L}) guarantees that all
522: discrete sums over momenta can be replaced with integrals, and,
523: since the integrals are convergent, the answer is independent of
524: $L$.
525:
526: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
527: \subsection{Large occupation number limit} \label{subsec:large_nu}
528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529:
530: At $\nu \gg 1$ there exists a wide ($U/t \ll \nu$) superfluid
531: region, in which large number fluctuations given by
532: Eq.~(\ref{Pn_Ideal_gas}) are gradually suppressed by the
533: interaction until they become of order unity at the SF-MI
534: transition. This physically appealing regime is not captured by
535: Eqs.~(\ref{Pn_weak_coupling}),(\ref{Pn_relative}), since they are
536: applicable only for $ U/t \ll 1/\nu$.
537:
538: At $\nu \gg 1$ and $U/\nu t \ll 1$, the number distribution is
539: easily obtained due to the fact that the typical values of the
540: occupation number fluctuations, $\eta_i = a_i^{\dagger}a_i - \nu$,
541: are large $1 \ll |\eta_i| \ll \nu$. Thus, the transformation
542: $a_i=\sqrt{\nu + \eta_i}\exp(i \Phi_i)$, where $\eta_i$ and
543: $\Phi_i$ are canonically conjugate Hermitian operators, along with
544: $\eta_i/\nu \ll 1$ reduces the Bose-Hubbard
545: Hamiltonian~(\ref{Bose-Hubbard-Hum}) to
546: \begin{multline}
547: H=-t \nu \sum_{<i,j>} \left[ 1 +
548: \frac{1}{4\nu^2}(\eta_i\eta_j-\eta_i^2)\right]\cos(\Phi_i-\Phi_j)
549: \\ + \frac{U}{2}\sum_i \eta_i^2, \label{Extended-Rotor-cos-Hum}
550: \end{multline}
551: where $0 \leq \Phi_i \leq 2 \pi$. In fact, in this form,
552: Eq.~(\ref{Extended-Rotor-cos-Hum}) is applicable at {\it any}
553: $\eta_i \ll \nu$ including the description of the SF-MI transition
554: and the MI phase. At sufficiently large interaction, namely $U/t
555: \gg 1/\nu$, the $\propto \eta_i^2/\nu^2$ term in
556: Eq.~(\ref{Extended-Rotor-cos-Hum}) can be neglected, and the
557: Hamiltonian coincides with the well-studied quantum rotor model
558: (QRM).
559:
560: In this work, we are interested in the $U/ t \ll \nu$ limit of
561: model Eq.~(\ref{Extended-Rotor-cos-Hum}), i.e. when phase
562: fluctuations between the nearest-neighbor sites are small $|\Phi_i
563: -\Phi_j| \ll 1$ and the number fluctuations are large. This allows
564: us to consider $\eta_i$ as a continuous variable and formally
565: redefine the domain of $\eta_i,\Phi_i$ as $-\infty < \eta_i,
566: \Phi_i < \infty$. The result is a harmonic approximation of the
567: Hamiltonian~(\ref{Extended-Rotor-cos-Hum}),
568: \begin{multline}
569: H=t \nu \sum_{<i,j>} \left[ \frac{1}{2} (\Phi_i-\Phi_j)^2 +
570: \frac{1}{4\nu^2}(\eta_i^2-\eta_i\eta_j)\right] \\ +
571: \frac{U}{2}\sum_i \eta_i^2.~~~~~~~~~ \label{Extended-Rotor-Hum}
572: \end{multline}
573:
574: From its functional (quadratic) form we immediately conclude that
575: the distribution of $\eta_i$, which is Gaussian at $U=0$ ($\nu \to
576: \infty$ limit of Eq.~(\ref{Pn_Ideal_gas})), remains Gaussian in a
577: wide range of coupling strength---all the way to the vicinity of
578: the SF-MI transition where $\sigma^2 \sim 1$ and the
579: model~(\ref{Extended-Rotor-Hum}) breaks down. The proof is as
580: follows. Since the Hamiltonian~(\ref{Extended-Rotor-Hum}) is
581: bilinear in $\{\eta_i, \Phi_i\}$, all averages are subject to the
582: Wick's theorem. Therefore, the characteristic function of the
583: distribution $W(\eta_i)$, which is given by the integral
584: $\int_{-\infty}^{\infty} \exp(ik\eta_i) W(\eta_i) \: d \eta_i$, is
585: a Gaussian $\exp(-k^2 \sigma^2/2)$.
586:
587: The only parameter of the distribution, $\sigma^2$, is given by
588: \begin{equation}
589: \sigma^2=\frac{\nu}{N_s} \sum_{\mathbf{k}}
590: \frac{\varepsilon_{\mathbf{k}}}{\omega_{\mathbf{k}}} \; \left[\,
591: 1+ 2\langle c_{\mathbf{k}}^{\dagger}c_{\mathbf{k}} \rangle \,
592: \right], \label{sigma-general}
593: \end{equation}
594: where $c_{\mathbf{k}}$ and $c_{\mathbf{k}}^{\dagger}$ are the
595: creation and annihilation operators of the normal modes of the
596: Hamiltonian~(\ref{Extended-Rotor-Hum}) with frequencies
597: $\omega_\mathbf{k}$ given by Eq.~(\ref{UkVk}). The Hamiltonian is
598: diagonalized by the transformation
599: \begin{gather}
600: \eta_i = \sum_\mathbf{k}
601: \sqrt{\frac{\omega_\mathbf{k}}{2U+\varepsilon_{\mathbf{k}}/\nu}}
602: \; \left[ \; \chi_{\mathbf{k} \, i} \, c_\mathbf{k} +
603: \chi^*_{\mathbf{k} \, i}\, c^{\dagger}_\mathbf{k} \; \right],
604: \notag
605: \\ \Phi_i = - i \sum_\mathbf{k} \sqrt{\frac{\omega_\mathbf{k}}{2\nu\varepsilon_{\mathbf{k}}}} \; \left[ \; \chi_{\mathbf{k} \, i} \, c_\mathbf{k}
606: - \chi^*_{\mathbf{k} \, i}\, c^{\dagger}_\mathbf{k} \; \right],
607: \label{phonon_fields} \\
608: \chi_{\mathbf{k} \, i}=\frac{1}{\sqrt{N_s}}\exp(i 2\pi\:
609: \mathbf{k} \, \mathbf{r}_i/L) \notag
610: \end{gather}
611:
612: Here, we restrict ourselves to ground state properties only and
613: thus set $\langle c_{\mathbf{k}}^{\dagger}c_{\mathbf{k}} \rangle
614: \equiv 0$, which leads exactly to Eq.~(\ref{sigma-general-zeroT}).
615: This is hardly surprising, since the
616: model~(\ref{Extended-Rotor-Hum}) is equivalent to the Bogoliubov
617: approximation of the
618: Hamiltonian~(\ref{Bose-Hubbard-Hum_momentum_space}) in the limit
619: of large $\nu$, and thus, we could formally demonstrate that
620: $P(n)$ is Gaussian at $\nu \gg 1$ already in the framework of
621: section~\ref{subsec:weak_coupling}. However, the hydrodynamic
622: approach chosen in this section seems more natural and physically
623: transparent when dealing with a dense system.
624:
625: Let us examine properties of the distribution variance in more
626: detail. An explicit expression for $\sigma^2$ reads
627: \begin{equation}
628: \sigma^2=\frac{\nu}{\pi^d} \int_0^\pi \cdots \int_0^\pi
629: \sqrt{\frac{\sum_\mu (1-\cos x_\mu)}{\nu U/t + \sum_\mu (1-\cos
630: x_\mu)}} \: d x_1 \cdots d x_d, \label{sigma-zeroT}
631: \end{equation}
632: In the limit of $\alpha=\nu U/t \rightarrow 0$, this expression
633: reduces to Eqs.~(\ref{lambda_1D})-(\ref{lambda_2D}). In the
634: opposite limit, we find
635: \begin{equation}
636: \sigma^2= \frac{I_d}{\pi^d} \sqrt{\frac{\nu t}{U}},~~~~~~~ U/t \gg
637: 1/\nu,\label{sigma-largeU}
638: \end{equation}
639: where $I_1 = 2 \sqrt{2}$, $I_2 \approx 13.373$, and $I_3 \approx
640: 52.348$. The latter formula allows to estimate the range of $U/t$
641: at which Eq.~(\ref{sigma-zeroT}) is applicable. The condition
642: $\sigma^2 \gg 1$ gives $U/t \ll \nu$, consistent with the
643: applicability range of Eq.~(\ref{sigma-general-zeroT}).
644:
645: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
646: \subsection{Interpolation formula in the SF regime} \label{subsec:inter_formula}
647: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
648: In sections~\ref{subsec:weak_coupling} and \ref{subsec:large_nu},
649: we have derived asymptotically exact solutions describing week
650: squeezing ($\lambda \ll 1$) and strong squeezing ($0 \leq \lambda
651: \lesssim 1$) at $\nu \gg 1$ respectively. The two limits overlap,
652: which suggests that one can write a single interpolation formula
653: to capture both limits correctly. This formula is also expected to
654: predict $P(n)$ correctly in a broader region of $U/t$ even at $\nu
655: \sim 1$. Formally, we have to find a function $\tilde{P}(n, \nu,
656: \lambda)$ such that (i) it coincides with Eq.(\ref{Pn_relative})
657: in the limit of $\lambda \ll 1$, (ii) it becomes Gaussian as a
658: function of $n$ at $\nu \gg 1$ and $0 \leq \lambda <1$ with the
659: average $\nu$ and variance $\tilde{\sigma}^2=\nu(1-\lambda)$, and
660: (iii) it is analytic with respect to $\nu$ and $\lambda$. The
661: solution is not unique, and we simply suggest one which satisfies
662: the above mentioned criteria
663: \begin{multline}
664: \tilde{P}\,(n,\nu,\lambda) = c \, P^{(0)}(n) \\
665: \exp\left[\frac{\lambda (n-\nu) + (n- \nu^2)}{2\nu} -
666: \frac{(n-\nu)^2}{2 \nu (1-\lambda)}\right],
667: \label{Pn-tilde-general}
668: \end{multline}
669: where $c$ is the normalization factor and $P^{(0)}(n)$ is the
670: Poisson distribution. By comparing Eq.~(\ref{Pn-tilde-general})
671: with numerical simulations at $\nu=1$ (described below) we find
672: that this formula accurately describes the actual form of $P(n)$
673: in a much broader range of $U/t$ than the
674: solution~(\ref{Pn_relative}).
675:
676:
677: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
678: \subsection{Strong coupling limit} \label{subsec:strong_coupling}
679: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
680:
681:
682: Let us now turn to the strong coupling limit, $\nu t / U \ll 1$,
683: at integer $\nu$, when atoms are well in the Mott insulator
684: regime, and the zeroth order approximation to the wavefunction is
685: a direct product of local Fock states, $|\Psi_{(t/U)=0}\rangle =
686: |\{ n_i=\nu \} \rangle$. The effect of finite hopping $t$ can be
687: taken into account as a perturbation of the on-site interaction
688: term in the Hamiltonian~(\ref{Bose-Hubbard-Hum}). To the first
689: approximation, this results in an admixture of the particle-hole
690: pairs to $|\Psi_{(t/U)=0}\rangle$,
691: \begin{equation}
692: |\Psi_{t/U}\rangle \propto \Bigl(
693: 1+\frac{t}{U}\sum_{<i,j>}a_i^{\dagger}a_j
694: \Bigr)|\Psi_{(t/U)=0}\rangle, \;\;\; \nu t/ U \ll 1.
695: \label{Psi_Mott}
696: \end{equation}
697: With this wavefunction Eq.~(\ref{Pn}) leads to the following
698: distribution
699: \begin{gather}
700: P_{\nu}=1-2P_{\nu-1}, \notag \\
701: P_{\nu-1}=P_{\nu+1}=2d \, \nu(\nu+1) \; t^2/U^2, \notag \\
702: P_n= 0, \; \mathrm{if} \;|n-\nu| > 1 . \label{Pn_Mott}
703: \end{gather}
704:
705:
706: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
707: \subsection{Numerics} \label{subsec:numerics}
708: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
709:
710: \begin{figure*}[htb]
711: \includegraphics[width=0.95\textwidth, keepaspectratio=true]{homog.eps}
712: %
713: %
714: \caption{(Color online.) The probability $P_n$ to detect $n$
715: particles on a single site of a homogenous square lattice as a
716: function of $U/t$ at zero temperature and unitary filling. The
717: results of MC simulation are shown for $n=0$ (squares), $n=1$
718: (circles) and $n=2$ (triangles). The uncertainties in $P_n$ are
719: smaller than the symbol size. The predictions of
720: Eq.~(\ref{Pn-tilde-general}) in the SF regime and that of
721: Eq.~(\ref{Pn_Mott}) in the strong coupling limit are plotted by
722: thick and thin solid lines respectively. The critical points of
723: the SF-MI transition, marked by the arrows on the graph, are
724: $(U/t)_c = 29.34(2)$ in three dimensions \cite{Barbara} (a),
725: $(U/t)_c=16.4(8)$ in two dimensions \cite{Krauth} (b), and
726: $(U/t)_c=3.30(2)$ in one dimension \cite{Svistunov_1DSFMI} (c).}
727: %
728: \label{fig:homog_case}
729: \end{figure*}
730:
731:
732: Clearly, as $U$ is changed from $0$ to $\infty$ the number
733: distribution must change qualitatively. Having an essentially long
734: tail at $n > \nu$ in the weakly interacting limit, $P_n$ becomes
735: sharply peaked at large $U$ with equal probabilities for $\nu-1$
736: and $\nu+1$ particles on a site. To obtain $P_n(U/t)$ in the
737: crossover regime, $t \sim U$, we perform MC simulations using the
738: continuous-time Worm Algorithm scheme \cite{Worm_Algorithm}. We
739: set $\nu=1$ and take the limit $N_s \rightarrow \infty$,
740: $\beta=1/T \rightarrow \infty$, where $T$ is the temperature. For
741: the linear system size $L=N_s^{1/d}=24$ and $\beta \gg L/2\pi c$,
742: where $c \sim 6t$ is the typical value of sound velocity in the
743: superfluid phase (higher temperatures can be used in the MI phase
744: with gaps $\sim U$), the shape of the distribution is already well
745: saturated, within a fraction of one percent in 3D and 2D, and
746: within a few percent in 1D, consistent with the fact that $P_n$ is
747: an essentially local characteristic.
748:
749: The simulation results for 3D, 2D and 1D are shown in
750: Fig.~\ref{fig:homog_case}, where $P_0$, $P_1$ and $P_2$ are
751: plotted as functions of $U/t$ along with the predictions of
752: Eq.~(\ref{Pn-tilde-general}) and Eq.~(\ref{Pn_Mott}). The main
753: observation is that, although $P_n(U/t)$ fundamentally does not
754: reveal any critical behavior, in 3D and 2D the SF-MI transition is
755: marked by a substantial change in $P_n$ curves---they rapidly
756: plateau in the Mott regime for $U/t \gtrsim (U/t)_c$. As expected,
757: in 1D the curves are much more smooth and the saturation at high
758: $U$ is not pronounced. Remarkably, Eq.~(\ref{Pn-tilde-general}),
759: deviates notably (by a few percent) from the numerical data only
760: at $U/t$ as high as $\sim 2$ in 1D, $\sim 4$ in 2D, and $\sim 8$
761: in 3D.
762:
763: We compare the theoretical predictions of
764: section~\ref{subsec:large_nu} for the asymptotics in the $\nu \gg
765: 1$ limit with the results of Monte-Carlo simulations. From
766: Fig.~\ref{fig:variance}, where the simulated probability
767: distribution at $U/t=2$ is plotted along with Gaussian curves with
768: variances calculated by Eq.~(\ref{sigma-zeroT}) at corresponding
769: $\nu$, we conclude that the shape of the distribution is perfectly
770: well (within the error bars) described by the Gaussian
771: distribution already at $\nu=5$. In Fig.~\ref{fig:variance_of_U},
772: the numerically simulated $\sigma^2$ and the prediction of
773: Eq.~(\ref{sigma-zeroT}) are plotted together as a function of
774: $U/t$.
775:
776:
777:
778: \begin{figure}[tbh]
779: \includegraphics[width=0.95\columnwidth,keepaspectratio=true]{var.eps}
780: %
781: %
782: %
783: \caption{(Color online.) Probability $P_n$ of the on-site
784: occupation number $n=\eta_i+\nu$ at $U/t=2$ from the Monte-Carlo
785: simulation (the error bars are smaller then the symbol size) with
786: the filling factor $5$ (squares), $10$ (circles), and $15$
787: (triangles). The solid lines are the Gaussian curves with the
788: variances calculated from Eq.~(\ref{sigma-zeroT}) with the
789: parameters of simulations.}
790: %
791: \label{fig:variance}
792: \end{figure}
793:
794: \begin{figure}[tbh]
795: \includegraphics[width=0.95\columnwidth,keepaspectratio=true]{varU.eps}
796: %
797: %
798: %
799: \caption{(Color online.) The variance $\sigma^2$ as a function of
800: interaction strength $U/t$ from the Monte-Carlo simulation (the
801: error bars are smaller then the symbol size) with the filling
802: factor $5$ (squares), $10$ (circles), and $15$ (triangles). The
803: solid lines represent the prediction of Eq.~(\ref{sigma-zeroT})
804: with the parameters of simulations.}
805: %
806: \label{fig:variance_of_U}
807: \end{figure}
808:
809:
810: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
811: \section{Comparison with experiment: effects of finite temperature}
812: \label{sec:exp}
813: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
814: In actual experiments, atoms are confined in the lattice by a
815: trapping potential, typically of a parabolic form. This results in
816: an inhomogeneous density profile in the SF regime and in the
817: formation of Mott-plateaus \cite{Gerbier, Bloch_spatial,
818: Ketterle_spatial}---spherical shells of integer filling---in deep
819: lattices. Correspondingly, the distribution of the number
820: fluctuations is also inhomogeneous, i.e. $P_n=P_n(i)$, where $i$
821: is the site index. The development of experimental techniques
822: allowing detection of atoms with a single-site spatial resolution
823: \cite{Gericke_scan_microscp} should open an exciting possibility
824: to directly measure atom correlation functions and $P_n(i)$ in
825: particular. Current experiments \cite{Gerbier, Bloch_spatial,
826: Ketterle_spatial} typically deal with integral characteristics of
827: the number distribution, such as the fraction of the total number
828: of atoms found on lattice sites with occupation $n$,
829: \begin{equation}
830: f_n = \sum_i n P_n(i). \label{fn}
831: \end{equation}
832:
833: The fraction of pairs $f_2$ in both SF and MI regimes can be
834: accurately probed by the spin-changing collisions
835: \cite{Bloch_spin_dynamics}. The measurement is set up in the
836: following way \cite{Gerbier}. After the system is allowed to
837: equilibrate at a fixed value of the lattice potential depth $V_0$,
838: the configuration of atoms is frozen by a rapid increase of $V_0$.
839: Then, coherent spin dynamics in the two-particle channel can be
840: induced with a near-unitary efficiency, and the spin oscillation
841: amplitude is measured to yield the fraction of pairs. With this
842: technique, Gerbier \textit{et al.} \cite{Gerbier} observed $f_2$
843: as the system was driven from a SF regime across the transition
844: deep into MI regime, corresponding to the values of the initial
845: lattice potential $V_0$ ranging from $4 E_r$ to $40 E_r$, where
846: $E_r=h^2/2m \lambda^2$ is the single photon recoil energy, $m$ is
847: the mass of a $^{87}$Rb atom, and $\lambda$ is the lattice laser
848: wavelength.
849:
850: In our simulations, the system of $^{87}$Rb atoms of
851: Ref.~\cite{Gerbier} is described by the Bose-Hubbard
852: model~(\ref{Bose-Hubbard-Hum}). The external potential is
853: harmonic, $\epsilon_i = m \omega_0^2 \, \mathbf{r}_i^2/2$, and all
854: the parameters of the Hamiltonian~(\ref{Bose-Hubbard-Hum}) are
855: fixed by the experimental setup. We study the number fluctuations
856: at two values of the lattice depth, $V_0 = 8 E_r$ and $V_0 = 13
857: E_r$ (corresponding to $U/t \approx 7.4$ and $U/t \approx 35.6$
858: respectively), which serve as examples of typical SF and MI phases
859: in the strongly correlated regime. The trapping frequency
860: $\omega_0$ is equal to $2\pi \times 37$Hz and $2\pi \times 46$Hz
861: for $V_0 = 8 E_r$ and $V_0 = 13 E_r$ correspondingly
862: \cite{Gerbier,Gerbier_private}. We perform simulations in a
863: sensible range of temperatures. Direct comparison between
864: experimental and numerical data enables us to estimate the final
865: temperature of the system. The results of the simulations are sown
866: in Fig.~\ref{fig:exp}, where the curves for $f_1$,$f_2$, and $f_3$
867: as functions of the total atom number in the trap $N$ are
868: parameterized by the temperature.
869:
870: \begin{figure}[tbh]
871: %\includegraphics[width=0.95\textwidth, keepaspectratio=true]{v8v13.eps}
872: \includegraphics[width=1\columnwidth,keepaspectratio=true]{v8v13.eps}
873:
874: %\includegraphics[width=0.9\columnwidth,keepaspectratio=true]{v13.eps}
875: %
876: %
877: \caption{(Color online.) Fractions of atoms $f_1$, $f_2$, $f_3$
878: occupying lattice sites with one, two and three particles on them
879: respectively versus the total number of particles in the trap at
880: the lattice depths $V_0=8E_r$, (a), (b), (c), and $V_0=13E_r$,
881: (d), (e), (f). The numerical results at different temperatures are
882: plotted with the experimental data for $f_2$ of
883: Ref.~\cite{Gerbier}.}
884: %
885: \label{fig:exp}
886: \end{figure}
887: Thermal effects vanish and a system can be considered to be in its
888: ground state, if the temperature is substantially smaller than the
889: energy of the low-lying excitations. In the SF regime, these modes
890: have typical frequencies of order $\sqrt{m/m_*} \, \omega_{0}$
891: ($m^*$ is the effective mass in the lattice), which for the
892: parameters of Fig.~\ref{fig:exp}(a)-(c), gives a rough estimate $T
893: \ll 3t$. Clearly, the curves for $T=t/3$ and $T=t$ in
894: Fig.~\ref{fig:exp}(a)-(c) are in the regime of effective zero
895: temperature. Already at $T=2t$ the thermal effects become
896: significant resulting in a larger atom cloud and, consequently,
897: reduced average density. At $T < 5t$ the size of the cloud for the
898: maximum number of atoms in the trap is $\sim 100$ lattice sites in
899: each dimension. To obtain $f_n$ at $T=10t$, we have to resort to
900: finite-size scaling, being limited by computer memory at the
901: linear system size of 150.
902:
903: In the MI phase, the zero temperature regime is reached for
904: temperatures smaller than the excitation gap, which is of order
905: $U/2$ for the parameters of Fig.~\ref{fig:exp}(d)-(f). This leads
906: to the condition $T \ll 20t$, consistent with the saturation of
907: $f_n$ below $T=5t$ (see Fig.~\ref{fig:exp}(d)-(f)). For $N
908: \lesssim 5 \times 10^4$, in the zero temperature limit, the curves
909: $f_n$ are flat, corresponding to the filling of the $\nu=1$ Mott
910: shell, and the number distribution is essentially squeezed. Simple
911: estimates \cite{Gerbier} show that the decrease of $f_1$
912: accompanied by the increase in $f_2$ at $N \sim 5 \times 10^4$,
913: and the saturation of $f_2$ with a peak in $f_3$ at $N \sim 2
914: \times 10^5$ are consistent with the formation of Mott plateaus
915: with $\nu=2$ and $\nu=3$, respectively. As seen from
916: Fig.~\ref{fig:exp}(d),(e), final temperature effects degrade the
917: degree of number squeezing in the MI by favoring particle-hole
918: excitations.
919:
920: The comparison of the calculated fraction of atom pairs $f_2$ with
921: the measurements of Ref.~\cite{Gerbier} (open circles in
922: Fig.~\ref{fig:exp}(b),(e)) gives the typical experimental
923: temperatures of order of $5t \approx 1.5\times 10^{-1} E_r$ in the
924: SF regime and $10t \approx 10^{-1}E_r$ in the MI regime.
925: Temperatures on the order of a few $t$ have been also observed in
926: a (one-dimensional) Tonks-Girardeau gas \cite{Bloch_Shlyapnikov},
927: where the effective fermionization due to strong interactions
928: allows to deduce the experimental temperature from the momentum
929: distribution. Note that the system acquires a finite temperature
930: as a result of its loading into the optical lattice, and therefore
931: the final temperature is supposed to depend on the number of atoms
932: in the trap. However, the fact that the experimental data lie
933: above the $T=0$ curve in Fig.~\ref{fig:exp}(b) at high $N$ is
934: unlikely to be explained by the the effects of heating. When a
935: large $f_3$ fraction is present, a change in the spin resonance
936: condition can result in a considerable contribution of spin
937: collisions on triply occupied sites to the observed spin
938: oscillation amplitude \cite{Gerbier}, which could explain the
939: anomalously high apparent $f_2$ \cite{Gerbier_private}. Such
940: drifts of the resonance parameters are carefully checked for, but
941: can not be ruled out completely \cite{Gerbier_private}.
942:
943:
944:
945: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
946: \section{Conclusions}
947: We studied the ground-state on-site number statistics of
948: interacting lattice bosons. We considered the limits of weak
949: interatomic interactions, the limit of large filling in the SF
950: regime, and the limit of strong interatomic interactions. At $\nu
951: U/t \ll 1$, the correction to the Poisson distribution is
952: described by Eq.~(\ref{Pn_relative}), with an essentially
953: dimension-dependent scaling---Eq.~(\ref{lambda_1D}) in 1D,
954: Eq.~(\ref{lambda_2D}) in 2D, and Eq.~(\ref{lambda_3D}) in 3D. In
955: the case of large occupation numbers, $\nu \gg 1$, we show that,
956: in the region of interactions $U/t \ll \nu$, the on-site
957: occupation number distribution is Gaussian and its variance, given
958: by Eq.~(\ref{sigma-zeroT}), gradually decreases with the
959: asymptotic form $\sigma^2 \propto \sqrt{\nu t/U}$ at $1/\nu \ll
960: U/t \ll \nu$ in all dimensions. An excellent agreement with
961: numeric simulations is found already at $\nu=5$. At $\nu t/ U \ll
962: 1$ and integer filling $\nu$, the distribution,
963: Eq.~(\ref{Pn_Mott}), is dominated by particle-hole fluctuations on
964: top of an ideal MI. At $\nu=1$, we performed Monte Carlo
965: simulations of the on-site occupation number distribution in 1D,
966: 2D, and 3D in a wide range of $U/t$ covering the crossover region
967: between the limiting cases.
968:
969: We simulated a parabolically confined system in the realistic case
970: of Ref.~\cite{Gerbier} with the final temperature of the system
971: being the only free parameter. By direct comparison between
972: experimental data and numerical results at different temperatures,
973: we were able to estimate the experimental temperature, which we
974: found to be of the order of a few $t$ near the transition. The
975: error bars are small enough to determine $T$ with accuracy of
976: order $t$. Our results suggest that measurements of the on-site
977: atom number fluctuations can serve as a reliable method of
978: thermometry in both superfluid and Mott-insulator regimes. We
979: believe that with more elaborate techniques, such as that giving
980: access to the spatial number distribution $P_n(i)$, the
981: temperature resolution can be further improved. \label{sec:concl}
982: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
983:
984: We are grateful to Fabrice Gerbier for providing us with the
985: experimental parameters and for fruitful discussions of the
986: results of this work. This research was supported by the National
987: Science Foundation under Grant No. PHY-0426881.
988:
989:
990: \begin{thebibliography}{99}
991:
992: \bibitem{Bloch_SFMI} M. Greiner, O. Mandel, T. Esslinger, T.W. H\"{a}nsch, and I. Bloch, Nature
993: \textbf{415}, 39 (2002).
994:
995: \bibitem{Ospelkaus} S. Ospelkaus, C. Ospelkaus, O. Wille, M. Succo, P. Ernst, K. Sengstock, and K.
996: Bongs, Phys. Rev. Lett. \textbf{96}, 180403 (2006).
997:
998: \bibitem{Bloch_spin_dynamics} A. Widera, F. Gerbier, S. Foelling, T. Gericke, O. Mandel, and I.
999: Bloch, Phys. Rev. Lett. \textbf{95}, 190405 (2005).
1000:
1001: \bibitem{Ketterle_spatial} G. K. Campbell, J. Mun, M. Boyd, P. Medley, A. E. Leanhardt, L. Marcassa, D. E. Pritchard, and W.
1002: Ketterle, \textit{cond-mat}/0606642 (2006).
1003:
1004: \bibitem{Bloch_SF3D} T. Gericke, F. Gerbier, A. Widera, S. Foelling, O. Mandel, and I.
1005: Bloch, \textit{cond-mat}/0603590 (2006).
1006:
1007: \bibitem{Zoller} D. Jaksch and P. Zoller, Ann. Phys. \textbf{315}, 52 (2005).
1008:
1009: \bibitem{Berman} P. Berman Ed., \textit{Atom Interferometry} (Academic Press, New York,
1010: 1997).
1011:
1012: \bibitem{Burnett} J. A. Dunningham, K. Burnett and S. M. Barnett,
1013: Phys. Rev. Lett \textbf{89}, 150401 (2002); J. A. Dunningham and
1014: K. Burnett, Phys. Rev. A, \textbf{70}, 033601 (2004).
1015:
1016: \bibitem{Rodriquez_interferometry} M. Rodr\'{i}guez, S. R. Clark, and D. Jaksch, \textit{cond-mat}/0607397 (2006).
1017:
1018:
1019: \bibitem{Seaman} B. T. Seaman, M. Kr\"{a}mer, D. Z. Anderson, and
1020: M. J. Holland, \textit{cond-mat}/0606625 (2006).
1021:
1022: \bibitem{Zoller98} D. Jaksch, C. Bruder, J.I. Cirac, C.W. Gardiner, and P.
1023: Zoller, Phys. Rev. Lett. \textbf{81}, 3108 (1998).
1024:
1025:
1026: \bibitem{Fisher} M.P.A. Fisher, P.B. Weichman, G. Grinstein, and D.S.
1027: Fisher, Phys. Rev. B, \textbf{40}, 546 (1989).
1028:
1029:
1030: \bibitem{Batrouni} G.G. Batrouni, V. Rousseau, R.T. Scalettar, M. Rigol, A. Muramatsu,
1031: P.J.H. Denteneer, and M. Troyer, Phys. Rev. Lett. \textbf{89},
1032: 117203 (2002).
1033:
1034: \bibitem{Bloch_theor}F. Gerbier, A. Widera, S. Foelling, O. Mandel,
1035: T. Gericke and I. Bloch, Phys. Rev. A, \textbf{72}, 053606 (2005)
1036:
1037: \bibitem{Isacsson} A. Isacsson, Min-Chul Cha, K. Sengupta, and S. M.
1038: Girvin, \textit{cond-mat}/0508068 (2005).
1039:
1040: \bibitem{Bloch_spatial} S. F\"{o}lling, A. Widera, T. M\"{u}ller, F. Gerbier, and I.
1041: Bloch, \textit{cond-mat}/0606592 (2006).
1042:
1043: \bibitem{Gerbier} F. Gerbier, S. F\"{o}lling, A. Widera, O.
1044: Mandel, and I. Bloch, Phys. Rev. Lett. \textbf{96}, 090401 (2006).
1045:
1046: \bibitem{Clark} S.R. Clark and D. Jaksch, \textit{cond-mat}/0604625 (2006).
1047:
1048: \bibitem{Penrose_Onsager} O. Penrose and L. Onsager, Phys. Rev. \textbf{104}, 576
1049: (1956).
1050:
1051: \bibitem{Orzel} C. Orzel, A. K. Tuchman, M. L. Fenselau, M.
1052: Yasuda, and M. A. Kasevich, Science, \textbf{291}, 2386 (2001).
1053:
1054: \bibitem{quant_opt} D. F. Walls and G. J. Milburn, \textit{Quantum Optics} (Springer-Verlag, Berlin,
1055: 1994).
1056:
1057:
1058: \bibitem{Caves} C. M. Caves, Phys. Rev. D, \textbf{23}, 1693
1059: (1981).
1060:
1061: \bibitem{Lu-Yu} X. Lu and Y. Yu, \textit{cond-mat}/0609230 (2006).
1062:
1063: \bibitem{Worm_Algorithm} N. V. Prokof'ev, B. V. Svistunov, and I.
1064: S. Tupitsyn, Phys. Lett. A, \textbf{238}, 253 (1998); JETP
1065: \textbf{87}, 310 (1998).
1066:
1067: \bibitem{Prokofev} G. Pupillo, C. J. Williams, and N. V. Prokof'ev,
1068: Phys. Rev. A, \textbf{73}, 013408 (2006).
1069:
1070: \bibitem{Gericke_scan_microscp}T. Gericke, C. Utfeld, N. Hommerstad, and H.
1071: Ott, Laser Phys. Lett. \textbf{3}, 415 (2006).
1072:
1073: \bibitem{LLStatMech2} see, e.g., E.M. Lifshitz and L.P. Pitaevskii,
1074: \textit{Statistical Mechanics}, Part 2, Pergamon Press, New York,
1075: 1980.
1076:
1077: \bibitem{KSS} Yu. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Sov. Phys. - JETP {\bf 66}, 314 (1987).
1078:
1079: \bibitem{Barbara} B. Capagrosso-Sansone, N. V. Prokof'ev, and B.
1080: V. Svistunov, in preparation.
1081:
1082: \bibitem{Krauth} W. Krauth and N. Triverdi, Europhys. Lett.
1083: \textbf{14}, 627 (1991).
1084:
1085: \bibitem{Svistunov_1DSFMI} V. A. Kashurnikov and B. V. Svistunov,
1086: Phys. Rev. B, \textbf{53}, 11776 (1996).
1087:
1088: \bibitem{Bloch_Shlyapnikov} B. Paredes, A. Widera, V. Murg, O. Mandel, S. F\"olling, I. Cirac, G. V. Shlyapnikov, T. W. H\"ansch,
1089: and I. Bloch, Nature \textbf{429}, 277 (2004).
1090:
1091: \bibitem{Gerbier_private} Fabrice Gerbier, private communication.
1092:
1093:
1094:
1095:
1096:
1097: %\bibitem{Freericks} J. K. Freericks, H. Monien, Phys. Rev. B
1098: %\textbf{53}, 2691 (1996).
1099:
1100:
1101:
1102:
1103: \end{thebibliography}
1104:
1105:
1106:
1107: \end{document}
1108: