0807.3475/T2.tex
1: \documentclass[prb,twocolumn,showpacs,groupedaddress]{revtex4}
2: \usepackage{amsmath}
3: \usepackage{graphicx}
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \begin{document}
7: \title{Coherent-incoherent transition in a Cooper-pair-box coupled to a quantum oscillator: an equilibrium approach}
8: %\title{Effect of a nanomechanical resonator on quantum tunneling in a Cooper-pair-box : an equilibrium approach }
9: \author{Ying-Hua Huang}
10: \author{Hang Wong}
11: \author{Zhi-De Chen}\email[Author to whom correspondence should be addressed: ]{tzhidech@jnu.edu.cn}
12: \affiliation{Department of Physics, Jinan University, Guangzhou
13: 510632, China}
14: %\date{July 24, 2002}
15: 
16: \begin{abstract}
17: Temperature effect on quantum tunneling in a Cooper-pair-box coupled
18: to a quantum oscillator is studied by both numerical and analytical
19: calculations. It is found that, in strong coupling regions, coherent
20: tunneling of a Cooper-pair-box can be destroyed by its coupling to a
21: quantum oscillator and the tunneling becomes thermally activated as
22: the temperature rises, leading to failure of the Cooper-pair-box.
23: The transition temperature between the coherent tunneling and
24: thermo-activated hopping is determined and physics analysis based on
25: small polaron theory is also provided.
26: 
27: \end{abstract}
28: \pacs{85.85.+j, 85.25.Cp, 71.38.-k} \maketitle
29: 
30: \section{Introduction}
31: Since the experimental realization of the single-electron transistor
32: (SET) in 1987,\cite{al,fd} the SET has become an important element
33: in both extremely precise measurement\cite{bw,zb,kc,la} and quantum
34: information processing.\cite{aa,mss,dv} The development of the
35: so-called radio-frequency SET electrometer is considered as the
36: electrostatic ``dual" of the well known superconducting quantum
37: interference devices (SQUIDs) and has shown its power of fast and
38: ultrasensitive measurement.\cite{sch,sch1} On the other hand, the
39: direct observation of the coherent oscillation in superconducting
40: SET (SSET, or the so-called Cooper-pair-box (CPB)) shows the SSET
41: can be served as a physical realization of a coherent two-state
42: system,\cite{na,na1} and since then the SSET or the SSET-based
43: circuit has been considered as a candidate of a qubit in a solid
44: state device and a read-out device of a qubit.\cite{aa,mss,dv}
45: 
46: The high frequency (up to 1 GHz) and high quality factor ($Q\sim
47: 10^3$) nano-mechanical resonator (NR) was firstly fabricated from
48: bulk silicon crystal in 1996,\cite{cr} and therefore it brought a
49: lot of interest in demonstrating the quantum nature of this small
50: mechanical device.\cite{is,ir,lt,ts,ar} A nano-mechanical resonator
51: capacitively coupled to a SSET forms the so-called
52: nano-electromechanical system (NEMS), a system shows promises for
53: fast and ultrasensitive force microscopy.\cite{la,pb} Especially,
54: displacement-detection approaching the quantum limit by this system
55: has been demonstrated.\cite{la} From a view point of theoretical
56: study, NEMS can be interesting in their own right. In fact, NEMS
57: provides a physical realization of a simple quantum system: a
58: two-level system (TLS) coupled to a single phonon mode, which has
59: basic relation to some important models in solid state physics and
60: quantum optics. First of all, it represents an oversimplified
61: spin-boson model,\cite{leg,weiss} i.e., the spin-boson model in a
62: single-phonon mode case, also it can be considered as the Einstein
63: model in a two-site problem in the exciton-phonon (or
64: polaron-phonon) system.\cite{ss} By taking the analog of the
65: resonator as the single mode cavity, it is the Jaynes-Cummings model
66: in quantum optics.\cite{lt,jm,jc,sun}
67: 
68: To be a physical realization of a TLS, the SSET needs to work in
69: severe constrains. It is well known that both thermal and quantum
70: fluctuations from the environment can destroy the Coulomb blockade
71: of tunneling and lead to the failure of the SET.\cite{al1,hd}
72: Additional conditions are needed for a SSET,\cite{al1} and one
73: important factor is to maintain the coherent tunneling. The study on
74: polaron-phonon system shown that the coherent motion of the electron
75: can be destroyed by its coupling to the phonon and a transition from
76: band motion to hopping motion happens as the temperature
77: rises.\cite{mah} By analogy, it is important to know if the coherent
78: tunneling of the SSET in NEMS can be destroyed by its coupling to a
79: NR, leading to failure of the SSET. We shall state such a
80: phonon-induced transition as coherent-incoherent transition and this
81: is the main interest of the present paper.
82: 
83: The coherent-incoherent transition in NEMS will be studied by both
84: numerical and analytical analyses. It is shown that the
85: coherent-incoherent transition exists for some frequency ranges with
86: strong coupling parameters. The rest of the present paper is
87: organized as follows. In Sec.\ II, we explain the model and the way
88: to trace out the environment, the results from adiabatic
89: approximation is also discussed. The coherent-incoherent transition
90: is studied in Sec.\ III and conclusion and discussion are presented
91: in Sec.\ IV.
92: 
93: \section{The model and explanation of the approximation}
94: The Hamiltonian of the NEMS, i.e., a TLS $+$ a NR, is given by
95: (setting $\hbar=1$)\cite{mss,ar}
96: \begin{equation}
97: \hat{H}_0=\frac{1}{2}\epsilon_0\hat{\sigma}_z-\frac{1}{2}\Delta_0\hat{\sigma}_x+\Omega\hat{a}^{\dagger}\hat{a}+\lambda\hat{\sigma}_z(\hat{a}^{\dagger}+\hat{a}),
98: \end{equation}
99: where $\epsilon_0=E_C(1-2n_g)$ and $E_C$ is the charging energy of a
100: Cooper pair, $n_g=C_gV_g/(2e)$ with $C_g$ and $V_g$ are the gate
101: capacity and gate voltage, respectively. $\Delta_0=E_J=I_c/(2e)$ and
102: $I_c$ is the critical current of the Josephson junction.
103: $\hat{\sigma}_i$ ($i=x,y,z$) are the Pauli matrices and
104: $\hat{a}^{\dagger}$  $(\hat{a})$ is the creation (annihilation)
105: operator of the phonon mode with energy $\Omega$, while $\lambda$ is
106: the coupling parameter. The above Hamiltonian indicates that our
107: main interest is the case when the SSET is closed to the degeneracy
108: point.\cite{mss,ar} It should be noted that the system described by
109: the above Hamiltonian is not a thermodynamical system, a true
110: thermodynamical system should include the environment, which is
111: modeled as a collection of harmonic oscillators.\cite{leg,weiss} The
112: whole Hamiltonian can be written as
113: \begin{equation}
114: \hat{H}=\hat{H}_0+\hat{H}_e,\quad\hat{H}_e=\sum_k\omega_k\hat{b}_k^{\dagger}\hat{b}_k+\hat{H}_{int}, %\hat{\sigma}_z\sum_k g_k(\hat{b}_k^{\dagger}+\hat{b}_k),
115: \end{equation}
116: where $\hat{H}_{int}$ represents the coupling between the
117: environment and NEMS. In the present paper, we shall assume that the
118: effect of the environment is just to keep the NEMS in equilibrium.
119: Here we apply the concept of reduced density matrix to explain the
120: approximation. The density matrix of the whole system is
121: \begin{equation}
122: \rho=\rho_0\otimes\rho_e,
123: \end{equation}
124: where $\rho_0$ and $\rho_e$ are the density matrix of the NEMS and
125: environment, respectively. The expectation value of any operator
126: $\hat{Q}$ that acts only on the variables of the NEMS can be found
127: by\cite{bl}
128: \begin{equation}
129: \langle\hat{Q}\rangle={\rm
130: Tr}\rho_r(\phi)\langle\phi|\hat{Q}|\phi\rangle,\quad \rho_r={\rm
131: Tr}_{bath}\rho,
132: \end{equation}
133: where $\rho_r$ is the reduced density matrix and ``${\rm
134: Tr}_{bath}$" means the trace operation over the environment. Our
135: approximation means that the above expression can be rewritten as
136: \begin{equation}
137: \langle\hat{Q}\rangle\simeq \frac{1}{Z}\sum_ie^{-\beta E_i}\langle\phi_i|\hat{Q}|\phi_i\rangle,~~~~~~~~
138: \hat{H}_0|\phi_i\rangle=E_i|\phi_i\rangle,
139: \end{equation}
140: with $Z=\sum_ie^{-\beta E_i}$ is the partition function of the
141: NEMS and $\beta=1/T$ (setting $k_B=1$). In previous
142: treatments,\cite{ir,lt} the effect of the environment is just to
143: keep the NR in equilibrium. The present treatment, taking a small
144: step further, suggests that the NEMS is in equilibrium with the
145: environmental bath.
146: 
147: By taking this approximation, thermodynamical properties of the NEMS
148: can be known, provided that the eigenvalue problem of $\hat{H}_0$ in
149: Eq.\ (1) can be solved. The eigenvalues of $\hat{H}_0$ can be obtain
150: by the standard numerical diagonalization technique. In a practical
151: treatment, the number of eigenvalues treated must be finite, we
152: therefore truncate the eigenspectrum by just taking the first
153: smallest $N$ eigenvalues, i.e.,
154: \begin{equation}
155: \langle\hat{Q}\rangle\simeq \frac{1}{Z}\sum_ie^{-\beta E_i}\langle\phi_i|\hat{Q}|\phi_i\rangle
156: \simeq \frac{1}{Z}\sum_{i=1}^Ne^{-\beta E_i}\langle\phi_i|\hat{Q}|\phi_i\rangle,
157: \end{equation}
158: with $Z\simeq \sum_{i=1}^Ne^{-\beta E_i}$, while $N$ is determined by the following condition
159: \begin{equation}
160: \beta(E_N-E_1)\ge L,
161: \end{equation}
162: where $L$ is a fixed number to control the calculation error and in
163: the present paper we have $L\ge 20$. Accordingly, one can find the
164: interested physical quantities from the numerical results, for
165: example, the free energy of the NEMS is
166: \begin{equation}
167: F(T)\simeq  -T\ln\left(\sum_{i=1}^Ne^{-\beta E_i}\right),
168: \end{equation}
169: from which all the thermodynamical quantities of the NEMS can be
170: found. In our numerical calculation, we have randomly checked the
171: dependence of the results on the value of $L$. We found no
172: observable dependence on $L$ for $L\ge 20$ by extending $L$ to 50 or
173: even 100. It should be noted that, for calculating the temperature
174: dependence of tunneling splitting, Eq.\ (5) should be modified as
175: \begin{equation}
176: \Delta(T)\simeq \frac{1}{Z}\sum_{i=1}^Ne^{-\beta E_i}|\langle\phi_i|\hat{\sigma}_x|\phi_i\rangle|,
177: \end{equation}
178: this is because the sign of $\langle\phi_i|\hat{\sigma}_x|\phi_i\rangle$ is not relevant in the calculation of $\Delta(T)$.
179: 
180: Before going to numerical results, we shall first discuss the
181: adiabatic approximation in the case of $\epsilon_0=0$ which was
182: shown to be a good approximation for $\Delta_0/\Omega \le
183: 1$.\cite{ir} In the adiabatic approximation, the eigenvalues of
184: $\hat{H}_0$ are characterized by pairs of energies that are given
185: by\cite{ir}
186: \begin{equation}
187: E_a^{\pm}(n)=n\Omega\pm\Delta_a(n)-\frac{\lambda^2}{\Omega},\quad\Delta_a(n)=\frac{\Delta_0}{2}\langle
188: n_+|n_-\rangle,
189: \end{equation}
190: where $|n_{\pm}\rangle=\exp\{\mp(\lambda/\Omega)(a^+-a)\}|n\rangle$
191: are the displaced-oscillator states and $n=0,1,2,\cdots$. Taking
192: these eigenvalues, one can find the temperature dependence of the
193: tunneling splitting according to Eq.\ (9) and the result is
194: \begin{equation}
195: \Delta_a(T)\simeq \frac{1}{Z_a}\sum_{n=0}^{\infty}\Delta_a(n)
196: e^{-\beta n\Omega}~2\cosh(\beta\Delta_a(n)),
197: \end{equation}
198: where $Z_a=\sum_{n=0}^{\infty}e^{-\beta
199: n\Omega}~2\cosh(\beta\Delta_a(n))$. In the limit of
200: $\Delta_a(n)\ll\Omega$, it can be found that
201: \begin{equation}
202: \Delta_a(T)\simeq \sum_{n=0}^{\infty}\Delta_a(n) p_{{\rm th}}(n),~~~~~~p_{{\rm th}}(n)=e^{-\beta n\Omega}(1-e^{-\beta \Omega}),
203: \end{equation}
204: a result that was used in the previous treatments.\cite{ir,lt}
205: \section{Coherent-incoherent transition}
206: %To perform numerical diagonalization, one needs to find out the
207: %matrix representation of $\hat{H}_0$ in the basis of
208: %$|\uparrow,\downarrow\rangle \otimes |n\rangle$, where
209: %$|\uparrow\rangle(|\downarrow\rangle)$ and $|n\rangle$ are the
210: %eigen-states of $\hat{\sigma}_z$ and $\hat{a}^{\dagger}\hat{a}$
211: %respectively. One can also do diagonalization in the basis of
212: %$|\uparrow,\downarrow\rangle \otimes |n_{\pm}\rangle$ as down in
213: %ref.\cite{ir}. We find that both schemes give the same result and
214: %the speed of calculation is almost the same.
215: To perform numerical diagonalization, one needs to represent the
216: Hamiltonian $\hat{H}_0$ with a suitable basis. In the treatment of
217: Ref.\ \onlinecite{ir}, the basis of $|\uparrow,\downarrow\rangle
218: \otimes |n_{\pm}\rangle$ is applied, where $|\uparrow\rangle$ and
219: $|\downarrow\rangle$ are the eigenstates of $\sigma_z$. However,
220: with this basis, the elements of the Hamiltonian matrix are flooded
221: with the overlap terms between different states $|n_\pm\rangle$ as
222: shown in the Eq.\ (8) of Ref.\ \onlinecite{ir}. Such overlap terms
223: are rather complicated to compute and therefore the generation of
224: the Hamiltonian matrix will be slow. In our treatment, for
225: convenient, we use a very simple basis that consist of the
226: eigenstates of $\sigma_z$ and $\hat{a}^\dagger\hat{a}$ to represent
227: the Hamiltonian, i.e., a basis of $|\uparrow,\downarrow\rangle
228: \otimes |n\rangle$. Hence, we represent the TLS terms with
229: $|\uparrow,\downarrow\rangle$ and the NR terms with $|n\rangle$ and
230: do a tensor product, the matrix of $\hat{H}_0$ is generated. Of
231: course, both schemes give the same results. The diagonalization can
232: give the eigenvalues of $\hat{H}_0$, i.e., $\{E_i\}$, and the
233: corresponding eigenvectors $\{\phi_i\}$, from which the tunneling
234: splitting $\langle\phi_i|\hat{\sigma}_x|\phi_i\rangle$ can be found.
235: Noted that, to reach the accuracy condition (7), the matrix size is
236: $\Omega$-dependent, the lower the frequency, the larger the matrix
237: size. In the following, we shall firstly concentrate our interest on
238: the unbiased case of $\epsilon_0=0$ and the biased case will be
239: discussed in the end of this section. The eigenvalues in the case of
240: $\epsilon_0=0$ we found are in good agreement with the results in
241: Ref.\ \onlinecite{ir}, i.e., the results of the adiabatic
242: approximation given in Eq.\ (10) work well for $\Omega/\Delta_0\ge
243: 1$. The corresponding tunneling splitting of the eigenstates for
244: $\lambda/\Omega=0.5$ and some typical values of $\Delta_0/\Omega$
245: are shown in Fig.\ 1, where the results of the adiabatic
246: approximation are also given for comparison.
247: 
248: \begin{figure*}
249: \centering
250: \begin{minipage}[c]{0.7\textwidth}
251: \centering
252: \includegraphics[width=\textwidth]{f1_1.eps}
253: \end{minipage}
254: \begin{minipage}[c]{0.25\textwidth}
255: \centering \caption{Tunneling splitting for eigenstates of
256: $\hat{H}_0$ in the case of $\lambda/\Omega=0.5$ and $\epsilon_0=0$
257: for some typical values of $\Delta_0/\Omega$ by numerical
258: calculation (shown as triangles). The results of the adiabatic
259: approximation (shown as ``$\ast$") are also shown for comparison.
260: The sketches of both curves show good agreement for
261: $\Delta_0/\Omega\le 1$ and the higher the excited states the better
262: the agreement. However, obvious discrepancy appears for both the
263: ground state and the first excited state even when $\Delta_0/\Omega=
264: 1/3$. The results of the adiabatic approximation show large
265: discrepancy with the numerical results in low frequency region with
266: $\Delta_0/\Omega> 1$.}
267: \end{minipage}%
268: \end{figure*}
269: 
270: \begin{figure}[h]
271: \centering
272: \includegraphics[width=0.5\textwidth]{f2_1.eps}
273: \caption{Comparison of the tunneling splitting of both the ground
274: state and the first excited state by numerical calculation with
275: adiabatic approximation in the case of $\epsilon_0=0$ for some
276: typical values of $\Delta_0/\Omega$. It should be noted that the
277: result by adiabatic approximation is
278: $\Delta_0/\Omega$-independent, which is shown as solid and dash
279: lines (dash line is the ground state). Numerical results are shown
280: in different patterns for different values of $\Delta_0/\Omega$
281: (the hollow patterns are for the ground state).}
282: \end{figure}
283: 
284: It can be found that the sketches of both curves show good agreement
285: for $\Delta_0/\Omega\le 1$ and the higher the excited states the
286: better the agreement. However, a closed look at the curves shows
287: that obvious discrepancy appears for both the ground state and the
288: first excited state even when $\Delta_0/\Omega= 1/3$. Figure 2 shows
289: the details of the discrepancy. We found that the discrepancy begins
290: to appear even when $\Delta_0/\Omega=0.1$ and becomes obvious when
291: $\Delta_0/\Omega = 1/3$. The present results show that, if one
292: compares the eigenvalues of the adiabatic approximation with the
293: numerical diagonalization, the adiabatic approximation seems to work
294: pretty well for $\Delta_0/\Omega\le 1$; however, if one measures the
295: tunneling splitting, the adiabatic approximation works well only for
296: $\Delta_0/\Omega\le 0.1$.
297: 
298: After finding out the eigenvalues and the corresponding tunneling
299: splitting, one can calculate temperature dependence of the tunneling
300: splitting according to Eq.\ (8). For all the frequencies we studied
301: (down to $\Delta_0/\Omega=50$), it is found that the tunneling
302: splitting decreases with temperature at the beginning in low
303: temperature regions. However, when the coupling strength
304: $\lambda/\Omega$ increases to some critical value, $\Delta(T)$ has a
305: ``upturn" at some temperature $T_t$, which is
306: $\lambda/\Omega$-dependent. Typical results are shown in Fig.\ 3.
307: 
308: \begin{figure*}
309: \centering
310: \begin{minipage}[c]{0.74\textwidth}
311: \centering
312: \includegraphics[width=\textwidth]{f3_1.eps}
313: \end{minipage}
314: \begin{minipage}[c]{0.25\textwidth}
315: \centering \caption{Temperature dependence on the tunneling
316: splitting in the case of $\epsilon_0=0$ for some typical values of
317: $\Delta_0/\Omega$. At the beginning, $\Delta(T)$ decreases with
318: temperature, but as the coupling parameter  $\lambda/\Omega$
319: increases, $\Delta(T)$ shows a ``up-turn" at some temperature $T_t$,
320: which is $\lambda/\Omega$-dependent.}
321: \end{minipage}%
322: \end{figure*}
323: 
324: The curves shown in Fig.\ 3 indicate that, for a given value of
325: $\Delta_0/\Omega$ and in the low temperature region, the main effect
326: of the NR on quantum tunneling is to decreases the tunneling
327: splitting in the weak coupling regions. The situation is more or
328: less the same as that in the polaron-phonon (or exciton-phonon)
329: system, say, coupling to phonons makes the electron become a
330: ``dressed" one and lowers the hopping rate of the electron. However,
331: as the coupling strength increases, the effect of the NR on quantum
332: tunneling changes as temperature increases to $T_t$, e.g., the
333: tunneling splitting is enhanced by a NR as temperature increases
334: further. The result is reminiscent of the transition from Bloch-type
335: band motion to phonon-activated hopping motion in the polaron-phonon
336: (or exciton-phonon) system.\cite{mah} Moreover, $\Delta_a(T)$
337: obtained from Eq.\ (11) does not show such ``upturn" for the
338: frequency regions we studied. This implies that the ``up-turn" is a
339: non-adiabatic effect.
340: 
341: As we have mentioned in Sec. I, the present model can be considered
342: as an analog to the polaron-phonon system of
343: Einstein model.\cite{mah} We therefore employ the small polaron theory to analyze the temperature dependence of the tunneling splitting. %\cite{mah}
344: At the beginning of the low temperature regions, the expectation
345: value of phonon number is small, the main contribution to the
346: tunneling splitting is the so-called diagonal transitions. The
347: diagonal transition rate can be found by following the way given in
348: Ref.\ \onlinecite{mah}. Firstly, a canonical transformation is
349: applied to the Hamiltonian in Eq.(2), i.e.,
350: \begin{equation}
351: \hat{H}'=e^S\hat{H}e^{-S}=\hat{{\cal H}}_0+\hat{V}+\sum_k\omega_k\hat{b}_k^{\dagger}\hat{b}_k+\hat{H}_{int}',%\hat{H}_e,
352: \end{equation}
353: where $S=(\lambda/\Omega)\hat{\sigma}_z(\hat{a}^{\dagger}-\hat{a})$,
354: \begin{equation}
355: \hat{{\cal H}}_0=\frac{\epsilon_0}{2}\hat{\sigma}_z+\Omega\hat{a}^{\dagger}\hat{a}-\frac{\lambda^2}{\Omega},
356: \end{equation}
357: \begin{equation}
358: \hat{V}=\frac{\Delta_0}{4}(\hat{\sigma}_+e^{(\lambda/\Omega)(\hat{a}-\hat{a}^{\dagger})}+h.c.),
359: \end{equation}
360: $\hat{H}_{int}'=e^S\hat{H}_{int}e^{-S}$ and $\hat{\sigma}_{\pm}=\hat{\sigma}_x\pm i\hat{\sigma}_y$. The diagonal transition rate is given by
361: \begin{equation}
362: w_d(T)={\rm Tr}_{\hat{{\cal H}}_0}\langle i|\hat{V}|i\rangle=\frac{\Delta_0}{2}e^{-S_T},%~~~~~n=(e^{\beta\omega}-1)^{-1},
363: \end{equation}
364: where $S_T=(2\lambda/\Omega)^2(n+1/2)$ and
365: $n=(e^{\beta\Omega}-1)^{-1}$ is the expectation value of phonon
366: number. Here, we have employed the approximation elucidated in Sec.
367: II, i.e., the effect of
368: $\sum_k\omega_k\hat{b}_k^{\dagger}\hat{b}_k+\hat{H}_{int}'$ is just
369: to keep the NEMS in equilibrium. $w_d$ and hence the tunneling
370: splitting will decrease with increasing temperature. On the other
371: hand, the main contribution to tunneling splitting, at higher
372: temperature, is the so-called non-diagonal transitions. One can
373: calculate the correlation function in the way given in Ref.\
374: \onlinecite{mah}, e.g.,
375: \begin{equation}
376: W(t)={\rm Tr}_{\hat{{\cal H}}_0}[\langle
377: i|\hat{V}(t)\hat{V}(0)|i\rangle-|\langle i|\hat{V}|i\rangle|^2],
378: \end{equation}
379: with $\hat{V}(t)=e^{i\hat{{\cal H}}_0t}\hat{V}e^{-i\hat{{\cal H}}_0t}$ and the result is
380: \begin{equation}
381: W(t)=(\Delta_0/2)^2e^{-2S_T}[e^{\varphi(t)}-I_0(\epsilon)],
382: \end{equation}
383: where $I_0(x)$ is a Bessel function and
384: \begin{equation}
385: \varphi(t)=\epsilon \cos[\Omega(t+i\beta\hbar/2)],\quad
386: \epsilon=2(2\lambda/\Omega)^2[n(n+1)]^{1/2}.
387: \end{equation}
388: It is interesting to note that, in the present case, the factor
389: $I_0(\epsilon)$ still survives after taking thermodynamical limit, a
390: situation that is different from the small polaron theory of the
391: Einstein model.\cite{mah} This is because only one phonon mode
392: coupled to the TLS in the present case and all the other modes just
393: serve as the bath. This result can also help to get rid of the delta
394: function problem shown in Ref.\ \onlinecite{mah}. The non-diagonal
395: transition rate can be found by
396: using the saddle-point integration%(setting $\hbar=1$)
397: \begin{equation}
398: w_n(T)=\int_{-\infty}^{\infty}W(t)dt\simeq (\Delta_0/2)^2(\pi/\gamma)^{1/2}e^{-2S_T+\epsilon},
399: \end{equation}
400: where $\gamma=\lambda^2[n(n+1)]^{1/2}$. It can be easily checked that $w_n$ shows opposite temperature dependence to $w_d$ and the transition
401: temperature $T_t$ is determined by $w_d=w_n$, which leads to
402: \begin{equation}
403: \frac{\Delta_0}{2}[\pi/\gamma(T_t)]^{1/2}e^{\epsilon(T_t)-S_T(T_t)}=1.%I_0(\epsilon(T_t))
404: \end{equation}
405: The above equation can be solved numerically and the transition
406: temperatures obtained are shown in Fig.\ 4, where the transition
407: temperatures determined from $\Delta(T)$ by numerical calculation
408: (i.e., curves shown in Fig.\ 3) are also presented for comparison.
409: 
410: \begin{figure*}
411: \centering
412: \begin{minipage}[c]{0.74\textwidth}
413: \centering
414: \includegraphics[width=\textwidth]{f4_1.eps}
415: \end{minipage}
416: \begin{minipage}[c]{0.25\textwidth}
417: \centering \caption{Transition temperatures obtained by analytical
418: (from Eq.(21)) and numerical calculations in the case of
419: $\epsilon_0=0$ for some typical values of $\Delta_0/\Omega$. The
420: analytical results are always higher than the numerical ones. For
421: $\Delta_0/\Omega<1$, both transition temperatures show similar
422: $\lambda/\Omega$ dependence, i.e., $T_t$ decreases with increasing
423: $\lambda/\Omega$. However, opposite $\lambda/\Omega$ dependence is
424: found for $\lambda/\Omega\ge 1$, showing that the analytical results
425: are invalid in the low frequency regions. The inset of the lefttop
426: figure shows the whole transition temperature curve obtained by the
427: analytical calculation.}
428: \end{minipage}%
429: \end{figure*}
430: 
431: It can be seen from Fig.\ 4 that the transition temperatures from
432: numerical calculations are always lower than the analytical ones. In
433: high frequency region with $\Delta_0/\Omega<1$, both obtained
434: transition temperatures show similar $\lambda/\Omega$ dependence,
435: i.e., $T_t$ increases as $\lambda/\Omega$ decreases. However, the
436: transition temperature obtained from analytical calculation shows
437: opposite coupling parameter dependence in low frequency region with
438: $\Delta_0/\Omega<1$, a result that is in confliction with an
439: intuitive picture. We notice that, as shown in the inset of Fig.\ 4,
440: analytical calculation predicts a transition for coupling strength
441: as low as $\lambda/\Omega\sim 0.2$ for $\Delta_0/\Omega\le 1/3$;
442: however, numerical result shows that there is no transitions for
443: coupling strength $\lambda/\Omega$ lower than 0.7. Figure 5 draws
444: the the transition boundary as a function of $\lambda/\Omega$ and
445: $\Delta_0/\Omega$ obtained by the numerical calculation.
446: 
447: \begin{figure}
448: \centering
449: \includegraphics[width=0.5\textwidth]{f5_1.eps}
450: \caption{Transition boundary obtained by the numerical calculation
451: in the case of $\epsilon_0=0$. The larger the $\Delta_0/\Omega$
452: (i.e., the lower frequency for a given $\Delta_0$), the larger the
453: coupling strength $\lambda/\Omega$ is needed for the transition to
454: happens, indicating the SSET is more stable when coupling to a lower
455: frequency NR.}
456: \end{figure}
457: 
458: We believe that the discrepancy mainly comes from the approximation
459: made in the analytical calculation of the diagonal transition rate.
460: It can be shown that
461: \begin{equation}
462: w_d(T)=\sum_{n=0}^{\infty}\Delta_a(n) p_{{\rm th}}(n)\simeq \Delta_a(T),
463: \end{equation}
464: which indicates that effectively the diagonal transition rate is
465: found by adiabatic approximation, an approximation is good for
466: finding the tunneling splitting only when $\Delta_0/\Omega\le 0.1$.
467: Figure 6 shows the details of how the discrepancy comes from the
468: adiabatic approximation.
469: 
470: \begin{figure}
471: \centering
472: \includegraphics[width=0.5\textwidth]{f6_1.eps}
473: \caption{Illustration of how the discrepancy comes from $w_d(T)$
474: obtained by the adiabatic approximation in the case of
475: $\epsilon_0=0$, $\Delta_0/\Omega=1/3$, and $\lambda/\Omega=1$. As
476: the temperature increases, the descending rate of $w_d(T)/w_d(0)$ is
477: much lower than $\Delta(T)/\Delta(0)$ from the numerical
478: calculation, leading to a higher transition temperature. The
479: ``up-turn" of $w_n(T)$ in the low temperature regions is due to the
480: factor $\gamma^{-1/2}$ which diverges as $T\rightarrow 0$. }
481: \end{figure}
482: 
483: It turns out that, as the temperature increases, the descending rate
484: of $w_d(T)/w_d(0)$ is much lower than $\Delta(T)/\Delta(0)$ from the
485: numerical calculation. Such a result has two consequences. The first
486: one is the transition temperature by analytical calculation is
487: higher than the numerical result since a slower descending
488: $w_d(T)/w_d(0)$ will lead to a higher crosspoint with a given
489: $w_n(T)$; Still another is the extremely high transition temperature
490: in the weak coupling regions shown in the inset of Fig.\ 4. It is
491: found that, in the weak coupling regions, the slow descending
492: $w_d(T)/w_d(0)$ can still survive to some high temperature, at where
493: $\Delta(T)$ from the numerical calculation has died out, bringing
494: about an artifact high transition temperature. On the other hand, As
495: one can see from Fig.\ 1 and Fig.\ 2, tunneling splitting from the
496: adiabatic approximation shows large discrepancy with the numerical
497: result when $\Delta_0/\Omega\ge 1$. This implies that $w_d(T)$ is a
498: bad approximation to calculate the diagonal transition rate in low
499: frequency regions with $\Delta_0/\Omega\ge 1$. Numerical analysis
500: indicates that $w_d(T)$ shows different
501: $(\lambda/\Omega)$-dependence from $\Delta(T)$, leading to different
502: $\lambda/\Omega$-dependence of $T_t$. In other words, the break-down
503: of the adiabatic approximation in the low frequency region with
504: $\Delta_0/\Omega\ge 1$ leads to an abnormal
505: $\lambda/\Omega$-dependence of the transition temperature in the low
506: frequency regions.
507: 
508: Now we turn to see the effect of bias on the coherent-incoherent
509: transition. In the case of $\epsilon_0\not=0$, intuitively, the
510: hopping between the TLS needs the assistance of the phonon unless
511: the bias is very small comparing with $\Delta_0$, i.e., the bias
512: can be overcome by the quantum fluctuation. Accordingly, the
513: diagonal transition is still possible and hence the
514: coherent-incoherent transition is expected to survive only when
515: $\epsilon_0/\Delta_0\ll 1$. Nevertheless, the appearance of the
516: non-zero bias will weaken the diagonal transition rate. As
517: $\epsilon_0$ increases to some value, at where tunneling becomes
518: impossible without the assistance of phonon, then the diagonal
519: transition makes no contribution to tunneling and
520: coherent-incoherent transition disappears. Numerical calculation
521: in the case of $\epsilon_0\not=0$ is the same as $\epsilon_0=0$
522: and some typical results obtained by the numerical calculation are
523: shown in Fig.\ 7. As $\epsilon_0/\Delta_0$ increases, the
524: transition temperature decreases and the transition disappears
525: when $\epsilon_0/\Delta_0$ reaches some critical value which is a
526: function of $\lambda/\Omega$ and $\epsilon_0/\Omega$. Obviously,
527: the numerical result is in accord with the analysis presented
528: above.
529: 
530: \begin{figure*}
531: \centering
532: \begin{minipage}[c]{0.74\textwidth}
533: \centering
534: \includegraphics[width=\textwidth]{f7_1.eps}
535: \end{minipage}
536: \begin{minipage}[c]{0.25\textwidth}
537: \centering \caption{The effect of bias on the coherent-incoherent
538: transition for some typical values of $\lambda/\Omega$ and
539: $\Delta_0/\Omega$. The transition temperature decreases with
540: increasing $\epsilon_0/\Omega$ and the transition disappears when
541: $\epsilon_0/\Omega$ reaches some critical value which depends on
542: both $\lambda/\Omega$ and $\Delta_0/\Omega$.}
543: \end{minipage}%
544: \end{figure*}
545: 
546: \section{Conclusion and discussion}
547: In conclusion, we have presented study on the effect of a
548: nanomechanical resonator on quantum tunneling in a Cooper-pair-box
549: at $T\not=0$. We found that the coherent tunneling of a
550: Cooper-pair-box can be destroyed by the coupling to a NR at a
551: temperature much lower than the coupling energy $E_J$ of the
552: Josephson junction. The present analysis shows that, for a NEMS to
553: work well, one additional condition, e.g., $T\ll T_t$, is needed.
554: The coherent-incoherent transition boundary as a function of
555: $\Delta_0/\Omega$ and $\lambda/\Omega$ is calculated. It turns out
556: that the transition happens only for $\lambda/\Omega>0.7$, and the
557: lower the $\Delta_0/\Omega$, the larger the $\lambda/\Omega$ is
558: needed for the transition. We also found that the transition
559: temperature is a monotonic descending function of
560: $\lambda/\Omega$. The present result shows that experimental
561: observation on the coherent-incoherent transition is still
562: impossible since the corresponding coupling parameter cannot be
563: achieved in the present stage. Taking some typical experimental
564: parameters:\cite{is} $\Delta_0=4\mu$eV and $\hbar\Omega=1.2\mu$eV,
565: the coupling parameter needed for the transition is
566: $\lambda/\Omega\simeq 1.4$, which is larger than the strong
567: coupling limit possible to achieve in experiment
568: $(\lambda/\hbar\omega_0\sim 1)$.\cite{ir} To see the transition in
569: the coupling parameters can be achieved in experiment, the result
570: shown in Fig.\ 5 tells $\Delta_0/\Omega$ should lower than 1,
571: i.e., the frequency of the NR should be as high as 1GHz, which is
572: almost the limitation in the present stage. Nevertheless, it is
573: believed that, in a real system, the transition can be seen in a
574: lower coupling parameter regions since the coupling of the NR to
575: the environment can help the transition to happen. It is also
576: found that other thermodynamical quantities, like specific heat,
577: vary smoothly over the transition point, showing that the
578: transition is not a thermodynamical transition.
579: 
580: Coherent-incoherent transition is an important issue in the
581: spin-boson model and analysis on the transition at $T\not=0$ was
582: also provided in Ref.\ \onlinecite{leg}. However, the starting
583: point in Ref.\ \onlinecite{leg} is different from the present
584: analysis. In the present analysis, the key element is the
585: temperature dependence on the tunneling splitting, while in the
586: previous analysis, it is the relaxation behavior, i.e., the time
587: dependence on transition rate $P(t)$.\cite{leg} As a matter of
588: fact, the present analytic calculation is similar to the
589: calculation provided in Sec. III D of Ref.\ \onlinecite{leg}, and
590: accordingly the transition rate we found here is corresponding to
591: $\Gamma$ (or $1/(2\tau)$), which was predicted to have a monotonic
592: temperature dependence in their analysis. It seems that this
593: result is suitable for the case with large bias while the
594: ``up-turn" of $\Delta(T)$ for small (or zero) bias cannot be
595: explained. Nevertheless, the present result can help to understand
596: the coherent-incoherent transition at $T\not=0$ in the spin-boson
597: model with small bias. To the first order approximation (i.e.,
598: omitting the cooperative effect between different phonon modes),
599: the combined contribution of all the phonon modes with a
600: frequency-dependent weight for $0 <\omega<\omega_c$ is
601: approximately the contribution of the whole bath, accordingly a
602: coherent-incoherent transition is expected to exist in the
603: spin-boson model since the coupling of the TLS to all the phonon
604: modes can lead to the transition when the coupling strength
605: exceeds some value.
606: 
607: \begin{acknowledgements}
608: This work was supported by a grant from the Natural Science
609: Foundation of China  under Grant No. 10575045.
610: \end{acknowledgements}
611: 
612: \begin{thebibliography}{99}
613: \bibitem {al}D. V. Averin and K. K. Likharev, J. Low Temp. Phys. {\bf 62}, 345 (1986).
614: \bibitem {fd}T. A. Fulton and G. J. Dolan, Phys. Rev. Lett. {\bf 59}, 109 (1987).
615: \bibitem {bw}M.P. Blencowe, M.N. Wybourne, Appl. Phys. Lett. {\bf 77}, 3845 (2000).
616: \bibitem {zb}Y. Zhang and M.P. Blencowe, J. Appl. Phys.  {\bf 91}, 4249 (2002).
617: \bibitem {kc}R. G. Knobel, A. N. Cleland, Nature (London) {\bf 424}, 291 (2003).
618: \bibitem {la}M. D. LaHaye, O. Buu, B. Camarota, and K. C. Schwab, Science {\bf 304}, 74 (2004).
619: \bibitem {aa}A. Aassime, G. Johansson, G. Wendin, R. J. Schoelkopf, and P. Delsing, Phys. Rev. Lett. {\bf 86}, 3376 (2001).
620: \bibitem {mss} Y. Makhlin, G. Sch\"on, and A. Shnirman, Rev. Mod. Phys. {\bf 73}, 357 (2001).
621: \bibitem {dv}D. Vion, A. Aassime, A. Cottet, P. Joyez, H. Pothier, C. Urbina, D. Esteve, and M. H. Devoret, Science
622: {\bf 296}, 886 (2002).
623: \bibitem {sch}R. J. Schoelkopf, P. Wahlgren, A. A. Kozhevnikov, P. Delsing, and D. E. Prober, Science {\bf 280}, 1238 (1998).
624: \bibitem {sch1}M. H. Devoret and R. J. Schoelkopf,  Nature (London) {\bf 406}, 1039 (2000).
625: \bibitem {na} Y. Nakamura, C. D. Chen, and J. S.
626: Tsai, Phys. Rev. Lett. {\bf 79}, 2328 (1997).
627: \bibitem {na1}Y. Nakamura, Yu. A. Pashkin, and  J. S. Tsai,
628: Nature (London), {\bf 398}, 786 (1999).
629: \bibitem {cr}A. N. Cleland and M. L. Roukes, Appl. Phys. Lett. {\bf 69}, 2653(1996); Nature (London), {\bf 392}, 160 (1998).
630: \bibitem {is}E. K. Irish and K. Schwab, Phys. Rev. B {\bf 68}, 155311 (2003).
631: \bibitem {ir} E. K. Irish, J. Gea-Banacloche, I. Martin, and K. C. Schwab, Phys.
632: Rev. B {\bf 72}, 195410 (2005).
633: \bibitem {lt}L. Tian, Phys. Rev. B {\bf 72}, 195411 (2005).
634: \bibitem {ts} T. Sandu, Phys. Rev. B {\bf 74}, 113405 (2006).
635: \bibitem {ar}A. D. Armour, M. P. Blencowe, and K. C. Schwab, Phys. Rev. Lett. {\bf 88}, 148301 (2002).
636: \bibitem {pb}M. P. Blencowe, Contemp. Phys., {\bf 46}, 249 (2005).
637: \bibitem {leg} A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg, and W. Zwerger, Rev. Mod. Phys.
638: {\bf 59}, 1 (1987) and references there in.
639: \bibitem {weiss} U. Weiss, {\it Quantum Dissipative Systems}, (World Scientific, Singapore, 1999).
640: \bibitem {ss} H. B. Shore and L. M. Sander, Phys. Rev. B {\bf 7}, 4537 (1973).
641: \bibitem {jm}J. M. Raimond, M. Brune, and S. Haroche, Rev. Mod. Phys. {\bf 73}, 565 (2001).
642: \bibitem {jc}E. T. Jaynes and F. W. Cummings, Proc. IEEE {\bf 51}, 89 (1963).
643: \bibitem {sun}F. Xue, L. Zhong, Y. Li, and C. P. Sun, Phys. Rev. B {\bf 75}, 033407 (2007).
644: \bibitem {al1}D. V. Averin and K. K. Likharev, in {\it Mesoscopic Phenomena in Solids}, ed. by B. L. Altshuler, P. A. Lee, and R. A. Webb,
645: (North-Holland, Amsterdam, 1991).
646: \bibitem {hd}M. H. Devoret, D. Esteve, H. Grabert, G.-L. Ingold, H. Pothier, and C. Urbina, Phys. Rev. Lett. {\bf 64}, 1824 (1990).
647: \bibitem {mah} G. D. Mahan, {\it Many-Particle Physics}, (Plenum Press, New York, 1990).
648: \bibitem {bl}K. Blum, {\it Density Matrix and Applications}, (Plenum Press, New York, 1996).
649: \end{thebibliography}
650: 
651: \end{document}
652: 
653: Dear Editor,
654: 
655:      The present paper is our recent work on coherent-incoherent
656:      transition in a Cooper-pair-box coupled to a nano-mechanical
657:      oscillator. It is found that the coherent tunneling of a
658:      Cooper-pair-box can be destroyed by its coupling to a
659: quantum oscillator. The result also has relation to the transition
660: from band motion to hopping motion in polaron-phonon system and
661: coherent-incoherent transition in spin-boson model. we hope it
662: could be published in your Phys. Rev. B.
663: 
664: Recommended referees: E. K. Irish; K. C. Schwab; T. Sandu; L.
665: Tian.
666: