0803.0059/JJ.tex
1: 
2: \documentclass[aps,prl,showpacs,twocolumn]{revtex4}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \usepackage{amsmath}
7: \usepackage{graphicx}
8: \usepackage{epsfig}
9: 
10: \setcounter{MaxMatrixCols}{10}
11: %TCIDATA{OutputFilter=Latex.dll}
12: %TCIDATA{Version=5.00.0.2552}
13: %TCIDATA{<META NAME="SaveForMode" CONTENT="1">}
14: %TCIDATA{LastRevised=Monday, January 28, 2008 10:25:27}
15: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
16: 
17: 
18: \begin{document}
19: 
20: \title{Probing Mott lobes via the AC Josephson effect}
21: \author{M.X. Huo$^{1}$}
22: \author{Ying Li$^{1}$}
23: \author{Z. Song$^{1}$}
24: \email{songtc@nankai.edu.cn}
25: \author{C.P. Sun$^{2}$}
26: \email{suncp@itp.ac.cn}
27: \homepage{http://www.itp.ac.cn/~suncp}
28: \affiliation{$^{1}$Department of Physics, Nankai University, Tianjin 300071, China}
29: \affiliation{$^{2}$Institute of Theoretical Physics, Chinese Academy of Sciences,
30: Beijing, 100080, China}
31: 
32: \begin{abstract}
33: The alternating-current (AC) Josephson effect is studied in a system
34: consisting of two weakly coupled Bose Hubbard models. In the framework of
35: the mean field theory, Gross-Pitaevskii equations show that the amplitude of
36: the Josephson current is proportional to the product of superfluid order
37: parameters. In addition, the chemical potential--current relation for a
38: small size system is obtained via the exact numerical computation. This
39: allows us to propose a feasible experimental scheme to measure the Mott
40: lobes of the quantum phase transition.
41: \end{abstract}
42: 
43: \pacs{03.65.Ud, 03.67.MN, 71.10.FD}
44: \maketitle
45: 
46: \emph{Introduction.} Recent development of experiments allows the detection
47: of the quantum phase transition in strongly correlated many-body systems
48: \cite{Probe}. The relevant physics is captured by the Bose-Hubbard model,
49: which describes the competition between the kinetic-energy and
50: potential-energy effects. The Mott insulator (MI) to superfluid (SF) phase
51: transition of the Bose-Hubbard model was described qualitatively using a
52: mean field theory \cite{Fisher} and realized in a gas of ultracold atoms
53: \cite{QBEC}. Generally, for the MI to SF transition, the superfluid order
54: parameter predicted by the mean field theory can not be measured directly.
55: However, in this letter, we propose a feasible realization to detect the
56: phase transition and obtain the Mott lobes of the order parameter. The main
57: idea relates to the well known alternating-current Josephson effect, which
58: has been well investigated in the Bose-Einstein condensates (BEC) \cite{Jc,
59: Jc2, Jc3, Jc4}.
60: 
61: The Josephson junction is composed of two superfluid parts separated by a
62: tunnelling barrier. When a constant chemical potential difference is
63: applied, the AC Josephson effect occurs, correponding an oscillating
64: particle flow through the barrier. If one part is fixed in surperfluid
65: state, the phase of other part should determine the amplitude of the AC
66: current. Inversely, it is possible to detect the state of the target part
67: via measuring the Josephson current across the barrier.
68: 
69: In this letter, through the well established mean field theory and
70: spontaneous symmetry breaking mechanism for quantum phase transition, the
71: analytical analysis reveals that the amplitude of the Josephson current is
72: proportional to the product of superfluid order parameters of the two parts.
73: Through the measurement of the AC Josephson current, the Mott lobes can be
74: obtained indirectly. Moreover, numerical simulations show that the profile
75: of the Mott lobes for small size system is consistent with that in
76: thermodynamic limit. This opens a possibility to detect the BEC in small
77: quantum device. As an application we discuss an experimental realization in
78: a cavity-atom hybrid system based on the recent work in Ref. \cite%
79: {Photonnonlinearity}.
80: 
81: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
82: 
83: \begin{figure}[tbp]
84: \includegraphics[bb=77 569 527 778, width=7 cm]{fig1.eps}
85: \caption{\textit{(Color online) Schematic plot of the lattice model used in
86: this work. (A) is a Bose-Hubbard model with the on site repulsion }$U$%
87: \textit{\ triggering the Mott transition; (B) is a free boson lattice model
88: with a chemical potential }$\protect\mu $\textit{\ which controls the
89: density of bosons in lattice A; (C) is a thin contact surface between A and
90: B which can be represented by a direct weak tunnelling across A and B or a
91: free boson lattice model with an on site repulsion larger than }$\protect\mu
92: $\textit{.}}
93: \label{Model}
94: \end{figure}
95: 
96: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97: 
98: \emph{Hamiltonian and the mean field theory. }The setup is depicted as the
99: following Hamiltonian
100: \begin{equation}
101: H=H_{A}+H_{B}+H_{C},  \label{H1}
102: \end{equation}%
103: where
104: \begin{eqnarray}
105: H_{A} &=&-\kappa \sum_{\left\langle \mathbf{i},\mathbf{j}\right\rangle \in
106: A}\left( a_{\mathbf{i}}^{\dag }a_{\mathbf{j}}+h.c.\right) +\frac{U}{2}\sum_{%
107: \mathbf{i}\in A}a_{\mathbf{i}}^{\dag }a_{\mathbf{i}}^{\dag }a_{\mathbf{i}}a_{%
108: \mathbf{i}},  \notag \\
109: H_{B} &=&-\kappa \sum_{\left\langle \mathbf{i},\mathbf{j}\right\rangle \in
110: B}\left( b_{\mathbf{i}}^{\dag }b_{\mathbf{j}}+h.c.\right) +\mu \sum_{\mathbf{%
111: i}\in B}b_{\mathbf{i}}^{\dag }b_{\mathbf{i}},  \label{Habc} \\
112: H_{C} &=&-g\sum_{\mathbf{i}\in C}\left( a_{\mathbf{i}}^{\dag }b_{\mathbf{i}%
113: }+h.c.\right) .  \notag
114: \end{eqnarray}%
115: Here $a_{\mathbf{i}}$ and $b_{\mathbf{i}}$ are the boson operators. In this
116: paper, we consider lattices A and B having an identical structure with size $%
117: N$ for simplicity. Each lattice site $\mathbf{i}$ $(=1,2,...,N)$ corresponds
118: to two positions in lattices A and B, respectively. Among them, $\mathbf{i}%
119: \in C $\ lies on the contact surface. We consider the weakly coupled case $%
120: g\ll \kappa $. Fig. 1 is a schematic representation of the setup.
121: 
122: We start our investigation from the mean field approximation. The order
123: parameters can be defined as the expectation values of boson operators $a_{%
124: \mathbf{i}}(t)$ and $b_{\mathbf{i}}(t)$ in the Heisenberg picture by
125: ignoring the fluctuations, i.e.,
126: \begin{equation}
127: \psi _{a,\mathbf{i}}(t)\equiv \left\langle a_{\mathbf{i}}(t)\right\rangle ,%
128: \text{ }\psi _{b,\mathbf{i}}(t)\equiv \left\langle b_{\mathbf{i}%
129: }(t)\right\rangle ,  \label{OP}
130: \end{equation}%
131: where the average is taken with respect to the ground state. In this paper,
132: we use $a_{\mathbf{i}}$ and $b_{\mathbf{i}}$ to denote the boson operators
133: in the Schr$\ddot{o}$dinger picture. In order to study the time evolution of
134: the order parameter, we make use of the Gross-Pitaevskii (GP) equations in
135: the lattice \cite{GPE, GPE2}. The GP equations corresponding to the
136: Hamiltonian (\ref{H1}) read
137: \begin{subequations}
138: \begin{eqnarray}
139: i\frac{\partial }{\partial t}\psi _{a,\mathbf{i}} &=&-\kappa
140: \sum_{\left\langle \mathbf{i},\mathbf{j}\right\rangle }\psi _{a,\mathbf{j}%
141: }+U\left\vert \psi _{a,\mathbf{i}}\right\vert ^{2}\psi _{a,\mathbf{i}}-g_{%
142: \mathbf{i}}\psi _{b,\mathbf{i}},  \label{GPa} \\
143: i\frac{\partial }{\partial t}\psi _{b,\mathbf{i}} &=&-\kappa
144: \sum_{\left\langle \mathbf{i},\mathbf{j}\right\rangle }\psi _{b,\mathbf{j}%
145: }+\mu \psi _{b,\mathbf{i}}-g_{\mathbf{i}}\psi _{a,\mathbf{i}},  \label{GPb}
146: \end{eqnarray}%
147: where $g_{\mathbf{i}}=g$ for $\mathbf{i}\in C$ and $g_{\mathbf{i}}=0$ for $%
148: \mathbf{i}\notin C$.
149: 
150: When $g$ is small enough, we can define solutions of Eqs. (\ref{GPa}) and (%
151: \ref{GPb}) as
152: \end{subequations}
153: \begin{equation}
154: \psi _{a,\mathbf{i}}=\phi _{a}\left( t\right) \Phi _{a,\mathbf{i}},\text{ }%
155: \psi _{b,\mathbf{i}}=\phi _{b}\left( t\right) \Phi _{b,\mathbf{i}},
156: \label{Psi_ai}
157: \end{equation}%
158: where $\Phi _{a,\mathbf{i}}$ and $\Phi _{b,\mathbf{i}}$ are solutions for
159: the ground states of Eqs. (\ref{GPa}) and (\ref{GPb}) when $g=0$. $\Phi _{a,%
160: \mathbf{i}}$ and $\Phi _{b,\mathbf{i}}$ are set to be real and normalized as
161: $\sum_{\mathbf{i}}\Phi _{a,\mathbf{i}}^{2}=\sum_{\mathbf{i}}\Phi _{b,\mathbf{%
162: i}}^{2}=1.$ Taking the periodic boundary condition, $\Phi _{a,\mathbf{i}}=$ $%
163: \Phi _{b,\mathbf{i}}$ $=1/\sqrt{N}$. After replacing $\psi_{a,\mathbf{i}}$
164: and $\psi_{b,\mathbf{i}}$, GP equations (\ref{GPa}) and (\ref{GPb}) become
165: the two-state model \cite{Jcformula}
166: \begin{subequations}
167: \begin{eqnarray}
168: i\frac{\partial }{\partial t}\phi _{a}\left( t\right) &=&\left[
169: E_{a}+U_{a}\left\vert \phi _{a}\left( t\right) \right\vert ^{2}\right] \phi
170: _{a}\left( t\right) -K\phi _{b}\left( t\right),  \label{TSM_a} \\
171: i\frac{\partial }{\partial t}\phi _{b}\left( t\right) &=&E_{b}\phi
172: _{b}\left( t\right) -K\phi _{a}\left( t\right) ,  \label{TSM_b}
173: \end{eqnarray}%
174: where
175: \end{subequations}
176: \begin{subequations}
177: \begin{eqnarray}
178: E_{a} &=&-2\kappa \sum_{\left\langle \mathbf{i},\mathbf{j}\right\rangle
179: }\Phi _{a,\mathbf{i}}\Phi _{a,\mathbf{j}}=-2d\kappa ,  \label{Ea} \\
180: E_{b} &=&\mu -2\kappa \sum_{\left\langle \mathbf{i},\mathbf{j}\right\rangle
181: }\Phi _{b,\mathbf{i}}\Phi _{b,\mathbf{j}}=\mu -2d\kappa ,  \label{Eb} \\
182: U_{a} &=&U\sum_{\mathbf{i}}\Phi _{a,\mathbf{i}}^{4}=\frac{U}{N},  \label{Ua}
183: \\
184: K &=&g\sum_{\mathbf{i}\in A}\Phi _{a,\mathbf{i}}\Phi _{b,\mathbf{i}}=\frac{g%
185: }{N^{\frac{1}{d}}},  \label{K}
186: \end{eqnarray}%
187: for two $d$-dimensional lattices with the periodic boundary condition and
188: contact surface with size $N^{\frac{d-1}{d}}$. $\phi _{a,b}\left( t\right)=%
189: \sqrt{N_{a,b}}e^{i\theta _{a,b}}$ where $N_{a,b}$ and $\theta _{a,b}$ are
190: the particle number and phase of the superfluid component in lattices A and
191: B.
192: 
193: Substituting $\phi _{a,b}\left( t\right) $ into the above two-state model,
194: we get
195: \end{subequations}
196: \begin{subequations}
197: \begin{eqnarray}
198: \frac{\partial z}{\partial t} &=&-2K\sqrt{1-z^{2}}\sin \Theta ,  \label{Zt}
199: \\
200: \frac{\partial \Theta }{\partial t} &=&\Delta E+\Lambda z+\frac{2Kz}{\sqrt{%
201: 1-z^{2}}}\cos \Theta ,  \label{Theta_t}
202: \end{eqnarray}%
203: where $\Theta =\theta _{a}-\theta _{b}$, $z=(N_{b}-N_{a})/(N_{a}+N_{b})$, $%
204: \Delta E=E_{b}-E_{a}-\frac{U_{a}}{2}\left( N_{a}+N_{b}\right) $, and $%
205: \Lambda =U_{a}\left( N_{a}+N_{b}\right) /2$. Eqs. (\ref{TSM_a}) and (\ref%
206: {TSM_b}) show that $N_{a}+N_{b}$ is conserved. Then the Josephson current
207: (if the boson is neutral, the current corresponds to the flow of bosons) is
208: \end{subequations}
209: \begin{equation}
210: J(t)=\frac{\partial N_{b}}{\partial t}=-2K\sqrt{N_{a}N_{b}}\sin \Theta .
211: \label{J(t)}
212: \end{equation}%
213: When $\mu \gg U,\kappa $, it becomes
214: \begin{equation}
215: J(t)=-\frac{2g\sqrt{N_{a}N_{b}}}{N^{\frac{1}{d}}}\sin \mu t.  \label{Jt}
216: \end{equation}%
217: Here $N_{a,b}$ is time-dependent. On the other hand, the upper bound of the
218: particle immigration across the junction during a half period of
219: oscillation, which corresponds to the case $N_{a,b}\sim N$, is of the order $%
220: \Delta N_{b}\sim $ $g\sqrt{N_{a}N_{b}}/(\mu N^{\frac{1}{d}})$ $\sim gN^{%
221: \frac{d-1}{d}}/\mu $. For $N^{\frac{d-1}{d}}\ll N$, $N_{a,b}$ can be
222: regarded as a constant, which corresponds to the order parameter defined in
223: the framework of the mean field theory \cite{Fisher} as $\psi _{\gamma
224: }\equiv \left\langle \gamma _{\mathbf{i}}\right\rangle $ $=\left\langle
225: \gamma _{\mathbf{i}}^{\dag }\right\rangle $ $=\sqrt{N_{\gamma }/N}$, ($%
226: \gamma =a,b$). Then the Josephson current is obtained as
227: \begin{equation}
228: J(t)=-2gN^{\frac{d-1}{d}}\psi _{a}\psi _{b}\sin \mu t.  \label{J_ab}
229: \end{equation}
230: 
231: \emph{Measurement of the order parameter. }In this section, we focus on the
232: experimental scheme to obtain the Mott lobes of the Bose-Hubbard model. It
233: is based on measuring the magnitude and frequency component of the Josephson
234: current. Our proposal for the measurement of the order parameter
235: experimentally is as follows: (a) Parameters $U$, $\kappa $ and $\mu $ are
236: set to be values corresponding to a point ($\mu /U$, $\kappa /U$) on the
237: phase diagram, then cool the system to the ground state. (b) Take the ground
238: state as an initial state, and shift the chemical potential to $\mu +\Delta
239: \mu $ ($\Delta \mu \gg U,\kappa $). (c) Obtain the current $J(t)$ between A
240: and B. (d) Numerically analyze the curve $J(t)$ to obtain
241: \begin{equation}
242: J_{m}\left( \mu /U,\kappa /U\right) =\max \left\{ J\left( \mu /U,\kappa
243: /U,t\right) \right\}  \label{Jm}
244: \end{equation}%
245: and the frequency component of the current from the formula
246: \begin{equation}
247: J\left( \mu /U,\kappa /U,\omega \right) =\sqrt{\frac{2}{\pi }}\int_{0}^{\tau
248: }dtJ\left( \mu /U,\kappa /U,t\right) \sin \omega t  \label{Jomiga}
249: \end{equation}%
250: as $\tau \rightarrow \infty $. (e) In the weak coupling limit $g\ll \kappa $%
251: , all the bosons in lattice B are in the superfluid state, i.e., $\psi
252: _{b}\simeq \langle b_{\mathbf{i}}^{\dag }b_{\mathbf{i}}\rangle ^{1/2}$. Then
253: $\psi _{b}$\ can be measured via the measurement of the average particle
254: density in lattice B. (f) Construct the phase diagram $\psi _{a}\left( \mu
255: /U,\kappa /U\right) $ according to Eq. (\ref{J_ab}).
256: 
257: Theoretically, the Josephson current arises from the time evolution of the
258: superfluid state of the Hamiltonian (\ref{H1}). In this Hamiltonian, we add
259: the chemical potential $\mu $ only in system B, but this is equivalent to
260: adding a chemical potential $-\mu $ in system A, since adding a term $-\mu
261: \sum_{\mathbf{i}}\left( a_{\mathbf{i}}^{\dag }a_{\mathbf{i}}+b_{\mathbf{i}%
262: }^{\dag }b_{\mathbf{i}}\right) $ to the whole system does not bring any
263: physical change. Then the order parameter of lattice A corresponds to the
264: point ($\mu /U$, $\kappa /U$) of the obtained phase diagram. At $t=0$, the
265: ground state of the Hamiltonian (\ref{H1}) is $\left\vert \varphi _{g}\left(
266: \mu /U,\kappa /U\right) \right\rangle =\left\vert \varphi \left( 0\right)
267: \right\rangle $, which is the initial state for the time evolution driven by
268: the Hamiltonian $H+\Delta \mu \sum_{\mathbf{i}}b_{\mathbf{i}}^{\dag }b_{%
269: \mathbf{i}}$, i.e., $\left\vert \varphi \left( t\right) \right\rangle
270: =e^{-i(H+\Delta \mu \sum_{\mathbf{i}}b_{\mathbf{i}}^{\dag }b_{\mathbf{i}%
271: })}\left\vert \varphi \left( 0\right) \right\rangle $. Then the current
272: across A and B is
273: \begin{eqnarray}
274: &&J\left( \mu /U,\kappa /U,t\right)  \label{J(t)_Phi(t)} \\
275: &=&-ig\sum_{\mathbf{i}\in C}\left\langle \varphi \left( 0\right) \right\vert
276: \left( a_{\mathbf{i}}^{\dagger }\left( t\right) b_{\mathbf{i}}\left(
277: t\right) -h.c.\right) \left\vert \varphi \left( 0\right) \right\rangle ,
278: \notag
279: \end{eqnarray}%
280: where $a_{\mathbf{i}}^{\dagger }\left( t\right) $ and $b_{\mathbf{i}}\left(
281: t\right) $ are boson operators in the Heisenberg picture. To demonstrate
282: this scheme, the numerical simulation is performed for $N=2$ system with $%
283: \kappa =1$, $\Delta \mu =100$, $g=0.1$, and $\mu /U=0.5$, which corresponds
284: to a parallel line in the phase diagram of lattice A. Fig. 2(a) is the 3D
285: plot of the Josephson current $J\left( U,t\right) $, which is a good
286: sinusoidal curve vs time $t$ when $U$ is not very large, in which region
287: systems A and B are both in the superfluid phase. When $U$ is large enough,
288: the amplitude of the current drops, which indicates that system A enters
289: into the Mott insulating phase. Fig. 2(b) is the 3D plot of the Fourier
290: component $J\left( U,\omega \right) $ obtained according to Eq. (\ref{Jomiga}%
291: ). The upper limit of the time integration (\ref{Jomiga}) is taken as $\tau
292: \left( 1/\Delta \mu \right) =20$. The larger values of $\tau $ will make the
293: peak with frequency $100$ higher and sharper. The current drops rapidly as $%
294: U $ increases, which indicates the quantum phase transition of system A.
295: These results show that the current is a Josephson alternating current with
296: frequency $\Delta \mu $ and can be employed to witness the quantum phase
297: transition.
298: 
299: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
300: 
301: \begin{figure}[tbp]
302: \includegraphics[bb=9 135 594 691, width=4 cm]{fig2a.eps} %
303: \includegraphics[bb=6 143 595 710, width=4 cm]{fig2b.eps}
304: \caption{\textit{(Color online) 3D plots of the Josephson current (a) }$%
305: J\left( U,t\right) $\textit{\ and its Fourier component (b) }$J\left( U,%
306: \protect\omega \right) $\textit{\ for a small size system obtained by the
307: exact diagonalization with }$\protect\kappa =1$\textit{, }$\Delta \protect%
308: \mu =100$\textit{, }$g=0.1$\textit{\ and }$\protect\mu /U=0.5$\textit{. In
309: (b), the upper limit of the time integration is taken as }$\protect\tau %
310: (1/\Delta \protect\mu )=20$\textit{. As }$\protect\tau $\textit{\ increases,
311: the peak with frequency }$100$\textit{\ will become higher and sharper. The
312: current drops rapidly as }$U$\textit{\ increases. These results show that
313: the current is a Josephson alternating current with frequency }$\Delta
314: \protect\mu $\textit{\ and can be employed to witness the quantum phase
315: transition.}}
316: \end{figure}
317: 
318: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
319: 
320: According to the mean field results, when $g$ is small enough, the Josephson
321: current is proportional to the order parameter of an isolated system with
322: the Hamiltonian $H_{A}\rightarrow H_{A}-\mu \sum_{\mathbf{i}}a_{\mathbf{i}%
323: }^{\dag }a_{\mathbf{i}}$. However, the Mott lobes obtained in a finite size
324: system from the above procedures (a-f) cannot be compared with the phase
325: diagram obtained from the mean field method in the thermodynamic limit,
326: which is usually obtained via the Gutzwiller trial wave function \cite%
327: {Fisher}. To compare the Josephson current with the mean field phase
328: diagram, we need to develop another way to obtain the order parameter for a
329: finite system in the framework of the spontaneous symmetry breaking. It is
330: well known that the quantum phase transition is a consequence of U(1)
331: symmetry breaking for an interacting boson system \cite{Sachdev}. Then
332: introducing an auxiliary field as
333: \begin{equation}
334: H_{A}-\mu \sum_{\mathbf{i}}a_{\mathbf{i}}^{\dag }a_{\mathbf{i}}\rightarrow
335: H_{A}-\mu \sum_{\mathbf{i}}a_{\mathbf{i}}^{\dag }a_{\mathbf{i}}+\lambda
336: \sum_{\mathbf{i}}\left( a_{\mathbf{i}}^{\dagger }+a_{\mathbf{i}}\right)
337: \label{auxiliary}
338: \end{equation}%
339: breaks the U(1) symmetry breaking and induces the order parameter.
340: Accordingly, the order parameter is determined by $\psi _{a}=\lim_{\lambda
341: \rightarrow 0}\lim_{N\rightarrow \infty }\left\langle a_{\mathbf{i}%
342: }\right\rangle ,$ where the average is taken for the ground state of the
343: Hamiltonian (\ref{auxiliary}). In the thermodynamic limit $N\rightarrow
344: \infty $, the ground state $\left\vert \varphi _{g}^{a}\right\rangle $ is a
345: coherent state and the vanishing $\psi _{a}$ discriminates two phases. In
346: this case, the order parameter is independent of $\lambda $. On the other
347: hand, when lattice A is coupled to B, the order parameter obtained from Eqs.
348: (\ref{J_ab}) and (\ref{J(t)_Phi(t)}) should also be independent of $g$ in
349: the limit $N\rightarrow \infty $ and $g\rightarrow 0$.
350: 
351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352: 
353: \begin{figure}[tbp]
354: \includegraphics[bb=8 347 589 784, width=7 cm]{fig3.eps}
355: \caption{\textit{(Color online) Mott lobes calculated by the exact
356: diagonalization with an auxiliary field for a small size system (color
357: contour map), and by the mean field method via the Gutzwiller trial wave
358: function. The contour line denotes the boundary of two phases which
359: corresponds to a vanishing order parameter. It is shown that although the
360: result of the small system can not give the boundary of two phases, its
361: contour lines are consistent with that from the mean field method in the
362: thermodynamic limit very well.}}
363: \end{figure}
364: 
365: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
366: 
367: In a finite system, although $\psi _{a}$ obtained from the above two ways
368: are not independent of $\lambda $ and $g$, it is believed that their
369: consistency should be revealed from the contour maps. To demonstrate this,
370: the numerical simulation for a small size system is performed and compared
371: with the mean field method. Even for a small size system, the dimension of
372: the Hilbert space of the Hamiltonian (\ref{auxiliary}) is infinite. So the
373: truncation approximation is taken for the exact diagonalization of the
374: matrix for $N=2$ and the average density $\left\langle a_{\mathbf{i}}^{\dag
375: }a_{\mathbf{i}}\right\rangle \in \lbrack 0,2]$. In Fig. 3, the Mott lobes
376: calculated by the exact diagonalization with an auxiliary field for $N=2$
377: system and by the mean field method via the Gutzwiller trial wave function
378: are plotted. The contour line denotes the boundary of two phases, which
379: corresponds to a vanishing order parameter. It is shown that although the
380: result of the small size system can not give the boundary of two phases, its
381: contour lines are consistent with that of the mean field method in the
382: thermodynamic limit very well.
383: 
384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385: 
386: \begin{figure}[tbp]
387: \includegraphics[bb=17 369 523 789, width=7 cm]{fig4.eps}
388: \caption{\textit{(Color online) Mott lobes of the order parameter for a
389: small size system calculated by the exact diagonalization with }$\protect%
390: \kappa =1$\textit{, }$\Delta \protect\mu =100$, and $g=0.1$. \textit{The
391: color contour map is obtained from the Josephson current, while the contour
392: lines are obtained via an auxiliary field with }$\protect\lambda =0.1$%
393: \textit{. It is shown that two results are consistent very well.}}
394: \end{figure}
395: 
396: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
397: 
398: In Fig. 4, Mott lobes of the order parameter for a small size system are
399: plotted through two different mechanisms with $\kappa =1$ and $g=0.1$. The
400: color contour map is obtained from the Josephson current via the above
401: procedures (a-f), while the contour lines are obtained from the Hamiltonian (%
402: \ref{auxiliary}) by the exact diagonalization. It is shown that two results
403: are consistent very well, which indicates that the property of the small
404: size system can shed light on the profile of Mott lobes in the
405: thermodynamics limit.
406: 
407: \emph{Discussion. }In order to detect the Mott lobes of the quantum phase
408: transition, we investigate the AC Josephson effect in a system consisting of
409: two weakly coupled Bose-Hubbard models. The mean field theory in the
410: thermodynamic limit and the numerical simulation for a small size system
411: show that, through measuring the magnitude and frequency component of the
412: Josephson current, the Mott lobes can be measured.
413: 
414: To realize this scheme experimentally, as mentioned before, a good candidate
415: is the coupled cavity system with each cavity interacting with 4-level atoms
416: driven by an external laser \cite{Photonnonlinearity}. In this system, the
417: repulsion $U$ can indeed be strong enough to observe the Mott insulator
418: state for photons. Moreover, according to Refs. \cite{Photonnonlinearity}
419: and \cite{parameters}, the controllable range of the chemical potential
420: required by our scheme is in experimentally accessible parameter regimes. In
421: fact, in such an effective Bose-Hubbard system, the on-site interaction $U$
422: and chemical potential $\mu $ for photons are determined by%
423: \begin{equation}
424: U=S\left( \frac{g_{13}}{\Omega }\right) ^{2}\frac{g_{24}^{2}}{\Delta },\text{
425: }\mu =S\left( \frac{g_{13}}{\Omega }\right) ^{2}\epsilon ,  \label{U}
426: \end{equation}%
427: under the conditions $g,\Delta ,g_{24},\epsilon \ll \Omega $; $g_{24}g\ll
428: \left\vert \Delta \Omega \right\vert $. Here $\Omega $ is the Rabi frequency
429: of the driving laser; $S$ is the atom number in each cavity; $g_{13}$, $%
430: g_{24}$ are the couplings between cavity mode to the atomic levels; $\delta $%
431: , $\Delta $ and $\epsilon $ are detunings of atomic transitions with respect
432: to the cavity and laser fields. All the notations are identical with those
433: used in Fig. 1 of Ref. \cite{Photonnonlinearity}.
434: 
435: By fixing $S\left( g_{13}/\Omega \right) ^{2}$\ and adjusting $%
436: g_{24}^{2}/\Delta $ and $\epsilon $, $U$ and $\mu $ are tunable
437: independently. This allows us to simulate the Hamiltonian (\ref{Habc}) by
438: setting $\epsilon =0$ in lattice A to get $\mu =0$, and $g_{24}=0$ in
439: lattice B to get the vanishing $U$. Subsequently, to drive the quantum phase
440: transition and probe the Mott lobes of lattice A, the parameters in two
441: systems A and B are tuned according to the procedures (a-f). In this scheme,
442: the Josephson current $J(t)$ is the flow of photons in photonic crystal
443: waveguides, which can be imaged via a high-resolution imaging technique,\
444: the collection scanning near-field optical microscope \cite{SNOM}. This
445: predicts that the Mott lobes for a small quantum device can be detected
446: experimentally.
447: 
448: This work is supported by the NSFC with grant Nos. 90203018, 10474104 and
449: 60433050, and NFRPC with Nos. 2006CB921206 and 2005CB724508.
450: 
451: \begin{thebibliography}{99}
452: \bibitem{Probe} I.B. Mekhov \textit{et al.}, Nature \textbf{3}, 319 (2007)
453: 
454: \bibitem{Fisher} M.P.A. Fisher \textit{et al.}, Phys. Rev. B \textbf{40},
455: 546 (1989).
456: 
457: \bibitem{QBEC} M. Greiner \textit{et al.}, Nature \textbf{415}, 39 (2002).
458: 
459: \bibitem{Jc} F.S. Cataliotti \textit{et al.}, Science \textbf{293}, 843
460: (2001).
461: 
462: \bibitem{Jc2} S. Giovanazzi, A. Smerzi and S. Fantoni, Phys. Rev. Lett.
463: \textbf{84}, 4521 (2000).
464: 
465: \bibitem{Jc3} M. Albiez \textit{et al.}, Phys. Rev. Lett. \textbf{95},
466: 010402 (2005).
467: 
468: \bibitem{Jc4} S. Levy \textit{et al.}, Nature \textbf{449}, 579 (2007).
469: 
470: \bibitem{Photonnonlinearity} M.J. Hartmann and M.B. Plenio, Phys. Rev. Lett.
471: \textbf{99}, 103601 (2007).
472: 
473: \bibitem{GPE} F. Dalfovo \textit{et al.}, Rev. Mod. Phys. \textbf{71}, 463
474: (1999).
475: 
476: \bibitem{GPE2} A. Polkovnikov \textit{et al.}, Phys. Rev. A \textbf{66},
477: 053607 (2002).
478: 
479: \bibitem{Jcformula} A. Smerzi \textit{et al.}, Phys. Rev. Lett. \textbf{79},
480: 4950 (1997).
481: 
482: \bibitem{Sachdev} S. Sachdev, \textit{Quantum Phase Transition}, (Cambridge
483: University Press, Cambridge, 1999).
484: 
485: \bibitem{parameters} D.K. Armani \textit{et al.},Nature (London) \textbf{421}%
486: , 925 (2003); T. Aoki \textit{et al.}, Nature (London) \textbf{443}, 671
487: (2006); S.M. Spillane \textit{et al.}, Phys. Rev. A \textbf{71}, 013817
488: (2005).
489: 
490: \bibitem{SNOM} S.I. Bozhevolnyi \textit{et al.}, Phys. Rev. B \textbf{66},
491: 235204 (2002).
492: \end{thebibliography}
493: 
494: \end{document}
495: