cond-mat0511281/T1.tex
1: %%\documentclass[english]{article}
2: %\documentclass [aps,showpacs,showkeys,preprint,amsmath,amssymb]{revtex4}
3: %\documentclass [aps,showpacs]{revtex4}
4: %\usepackage[T1]{fontenc}
5: %\usepackage[latin1]{inputenc}
6: %\usepackage{setspace}
7: %\doublespacing
8: %\usepackage{amssymb}
9: %\documentclass[showpacs,preprintnumbers,amsmath]{revtex4}
10: 
11: 
12: 
13: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath]{revtex4}
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15: \usepackage{graphicx}
16: \usepackage{amsmath}
17: 
18: \begin{document}
19: 
20: \title{\textbf{Decoherence induced by electron accumulation \\
21: in quantum measurement of charge qubits}}
22: \author{\textbf{Ming-Tsung Lee$^{a}$ and Wei-Min Zhang$^{a,b}$}}
23: \date{\today}
24: \affiliation{$^{a}$Physics Division, National Center for Theoretical
25: Sciences,
26: Tainan, Taiwan 70101, ROC \\
27: $^{b}$Department of Physics and Center for Quantum Information
28: Science, National Cheng Kung University, Tainan, Taiwan 70101, ROC}
29: 
30: \begin{abstract}
31: In this paper, we study the quantum decoherence induced by
32: accumulation of electron tunnellings during the quantum measurement
33: of a charge qubit. The charge qubit is a single electron confined in
34: coupled quantum dots. The measurement of the qubit states is
35: performed using a quantum point contact. A set of master equations
36: for qubit states is derived within a non-equilibrium perturbation to
37: the equilibrium reservoir due to the electron accumulation between
38: the source and drain of the quantum point contact. The quantum
39: decoherence of the qubit states arose from the electron accumulation
40: during the measurement is explored in this framework, and several
41: interesting results on charge qubit decoherence are obtained.
42: \end{abstract}
43: 
44: \pacs{03.65.Yz,85.35.Be,03.65.Ta,03.67.Lx,73.63.Kv} \maketitle
45: 
46: 
47: \section{INTRODUCTION}
48: 
49: Quantum decoherence \cite{decoh1,decoh2,decoh3} is mainly induced by the
50: interaction of a microscopic system coupled with its environment. The
51: system-environment coupling results in a non-unitary evolution of the
52: system, which destroys the purity of quantum states and leads to information
53: loss toward the environment. These issues have attracted much attention in
54: the study of quantum computation and quantum communication in recent years
55: \cite{qis}. Some investigations of quantum decoherence have been focused on
56: the measurement induced decoherence \cite
57: {gurvitz,korotkov,goan,mozyrsky,clerk,stace,li}, and the dynamical controls
58: of the decoherence \cite{free decoh, dy decoup}. For the implementation of
59: realistic quantum information processors, these investigations become the
60: most champion works in the field. In the solid-state quantum computer with
61: charge qubits, quantum measurement of the qubit states can be performed by
62: coupling the charge qubit with a sensitive electrometer such as quantum
63: point contacts (QPCs) \cite{qpc} or a single-electron transistor \cite
64: {set,cooper}. In this paper, we shall focus on the charge qubit measurement
65: using a QPC, and study the non-equilibrium dynamical effect of the QPC on
66: the decoherence of the qubit states, here the charge qubit is a single
67: electron confined in coupled quantum dots (CQD).
68: 
69: Literaturely, to study the quantum decoherence induced by quantum
70: measurements, an equilibrium approximation is applied to the reservoir state
71: \cite{goan,stace,li}. For example, in the study of a charge qubit measured
72: by a QPC (Fig.~\ref{fig1}), the electronic reservoir, the source and the
73: drain of the QPC, is assumed to be macroscopic enough in comparison with the
74: CQD. Then the thermal equilibrium of the reservoir is kept continuously
75: through rapid relaxation processes. This condition of a perfect heat bath
76: causes electrons tunneling such that the extra electrons arriving at the
77: drain will flow back rapidly into the source through a close loop of the
78: transport circuit. No extra electron accumulates in the steady reservoir.
79: Practically, however, the condition of the perfect heat bath is not
80: guaranteed at mesoscopic scale of the QPC. The electron accumulation (EA) in
81: the source and the drain of the QPC may destroy the equilibrium of reservoir
82: and influence the outcome of the electrometer. As a result, the effects of
83: EA may become significant to the decoherence of the measured qubit. In this
84: paper, we shall treat approximately this intermediate non-equilibrium effect
85: as a perturbation to the equilibrium reservoir. The quantum decoherence of
86: the charge qubit due to the EA is then studied within this framework.
87: 
88: The paper is organized as follows: The model of the charge qubit measurement
89: and the perturbation scheme of taking the EA effect into account are
90: presented in Sec. II. A set of master equations for the reduced density
91: operator of the qubit is also derived in this section. In Sec. III., the
92: quantum decoherence of the qubit in the thermal equilibrium limit is
93: illuminated according to the resultant master equations. An abnormal
94: dependence of the qubit decay mode to the temperature and bias voltage is
95: obtained. The EA effect on the quantum decoherence of the qubit states is
96: explored in Sec. IV. The dynamics of the EA is analyzed, and the EA induced
97: decoherence of the qubit is extensively discussed. Remarkably, we also found
98: a decoherenceless EA effect in the low temperature regime. Finally, a
99: summary is given in Sec. V.
100: 
101: \section{CHARGE QUBIT MEASUREMENT USING QPC}
102: 
103: The model we consider here is a single electron confined in a CQD (as a
104: charge qubit) that is measured by the QPC detector, see Fig.~\ref{fig1}a.
105: The Hamiltonian of the whole system is given by \cite
106: {gurvitz,korotkov,goan,stace,li}
107: \begin{eqnarray}
108: \hat{H} &=&\hat{H}_{S}+\hat{H}_{B}+\hat{H}^{\prime },  \label{eq:h1} \\
109: \hat{H}_{S} &=&E_{L}\hat{c}_{L}^{+}\hat{c}_{L}+E_{R}\hat{c}_{R}^{+}\hat{c}%
110: _{R}+\Omega _{0}\left( \hat{c}_{L}^{+}\hat{c}_{R}+\hat{c}_{R}^{+}\hat{c}%
111: _{L}\right) ,  \label{eq:hd} \\
112: \hat{H}_{B} &=&\sum_{l}\varepsilon _{l}\hat{a}_{l}^{+}\hat{a}%
113: _{l}+\sum_{r}\varepsilon _{r}\hat{a}_{r}^{+}\hat{a}_{r},  \label{eq:hpc} \\
114: \hat{H}^{\prime } &=&\left( \Omega -\delta \Omega \hat{c}_{R}^{+}\hat{c}%
115: _{R}\right) \sum_{lr}\left( \hat{a}_{l}^{+}\hat{a}_{r}+\hat{a}_{r}^{+}\hat{a}%
116: _{l}\right) .  \label{eq:hi}
117: \end{eqnarray}
118: Here, $\hat{H}_{S}$ is the Hamiltonian of the charge qubit in which $\hat{c}%
119: _{L}^{+}\left( \hat{c}_{L}\right) $ and $\hat{c}_{R}^{+}\left( \hat{c}%
120: _{R}\right) $ are the electron creation (annihilation) operators of the
121: electron sited in the two dots labelled by $L$ and $R$ as shown in Fig. 1a.
122: They satisfy the condition $\hat{c}_{L}^{+}\hat{c}_{L}+\hat{c}_{R}^{+}\hat{c}%
123: _{R}=1$. The parameter $\Omega _{0}$ is the electron tunneling amplitude
124: between the two dots. $\hat{H}_{B}$ denotes the Hamiltonian of the QPC
125: reservoir, which is decomposed into two terms for the source and the drain,
126: respectively, and $\hat{a}_{l}^{+}\left( \hat{a}_{l}\right) $ and $\hat{a}%
127: _{r}^{+}\left( \hat{a}_{r}\right) $ are the corresponding electron creation
128: (annihilation) operators. Since the presence of an electron in a dot close
129: proximity to the QPC will cause a variation in barrier of the QPC, $\Omega
130: \rightarrow \Omega -\delta \Omega $, the interaction Hamiltonian $\hat{H}%
131: ^{\prime }$ describes the coupling of the quantum dots and the QPC in close
132: proximity to the dot $R$. The information of the qubit-states can then be
133: read out through the QPC current \cite{gurvitz}.
134: \begin{figure}[tbph]
135: \includegraphics*[angle=0,scale=.65]{fig1}
136: \caption{(color online). (a) Schematic plot of a single electron in a
137: coupled quantum dots measured by the QPC. (b) Schematic diagram of the
138: electron accumulation in reservoir, the first figure shows the forward
139: tunneling processes (FTP) of electrons from the source to the drain which
140: causes electrons accumulating in the drain, and induces a variation of the
141: Fermi energy. The second one shows the backward tunneling processes (BTP) of
142: electrons from the drain to the source and a corresponding variation of the
143: Fermi energy.}
144: \label{fig1}
145: \end{figure}
146: 
147: There are two channels for electron tunneling in the QPC. One is the forward
148: tunneling processes from the source to the drain, and the other is the
149: backward tunneling processes from the drain to the source, see Fig.~\ref
150: {fig1}b. In the high-bias regime, magnitude of the electron tunneling rate
151: for the forward tunneling processes is much larger than the one in the
152: backward tunneling processes. The contribution of the backward tunneling
153: processes could be ignored as treated in \cite{gurvitz}. However, in a
154: non-equilibrium description, both contributions of the forward tunneling
155: processes and backward tunneling processes should be included for an
156: arbitrary bias voltage. The Hilbert space is then spanned by the following
157: basis
158: \begin{equation}
159: \left\{ \left| \phi _{i}\right\rangle \otimes \left| D\left( \bar{L}%
160: ^{n},R^{n}\right) \right\rangle ,\left| \phi _{i}\right\rangle \otimes
161: \left| \bar{D}\left( L^{m},\bar{R}^{m}\right) \right\rangle \right\} ,
162: \label{eq:basis}
163: \end{equation}
164: where $|\phi _{i}\rangle $ is an electron eigen-state in the CQD (we define $%
165: |\phi _{0}\rangle =|g\rangle $ and $|\phi _{1}\rangle =|e\rangle $ as the
166: ground and excited states in the eigen-state representation, respectively), $%
167: |D(\bar{L}^{n},R^{n})\rangle $$=\hat{a}_{r_{1}}^{+}\cdots \hat{a}_{r_{n}}^{+}%
168: \hat{a}_{\bar{l}_{1}}\cdots \hat{a}_{\bar{l}_{n}}\left| B_{0}\right\rangle $
169: is a state with $n$ electrons accumulated in the drain through the forward
170: tunneling processes, and $\left| \bar{D}\left( L^{m},\bar{R}^{m}\right)
171: \right\rangle =\hat{a}_{l_{1}}^{+}\cdots \hat{a}_{l_{m}}^{+}\hat{a}_{\bar{r}%
172: _{1}}\cdots \hat{a}_{\bar{r}_{m}}\left| B_{0}\right\rangle $ a state with $m$
173: electrons accumulated in the source through the backward tunneling
174: processes. The state $|B_{0}\rangle $ is the reservoir (equilibrium) vacuum,
175: and the operator $\hat{a}_{\bar{l}(\bar{r})}$ denotes the annihilation of an
176: electron from $|B_{0}\rangle $ [or the creation of a hole below the Fermi
177: energy of the source (drain)], while the operator $\hat{a}_{l(r)}^{+}$
178: denotes the creation of an electron above the Fermi energy of the source
179: (drain). Therefore, an arbitrary state of the whole system can be expressed
180: as
181: \begin{eqnarray}
182: |\psi (t)\rangle &=&\sum_{i}|\phi _{i}\rangle \otimes \Big(\sum_{n}\sum_{%
183: \bar{L}^{n}R^{n}}b_{i,\bar{L}^{n}R^{n}}(t)|D(\bar{L}^{n},R^{n})\rangle
184: \nonumber \\
185: &&+\sum_{m}\sum_{L^{m}\bar{R}^{m}}\tilde{b}_{i,L^{m}\bar{R}^{m}}(t)|\bar{D}%
186: (L^{m},\bar{R}^{m})\rangle \Big).  \label{eq:tstat}
187: \end{eqnarray}
188: 
189: To study the charge qubit measured by the QPC, the electronic reservoir is
190: usually assumed to be in the reservoir vacuum state (a thermal equilibrium
191: state). The reservoir vacuum state is uniquely determined for a given
192: temperature and chemical potential. The reservoir time correlation function
193: is usually calculated in this state. However, tunneling electrons from the
194: forward tunneling processes will occupy the levels above the Fermi energy of
195: the drain for a short period and disturbs the reservoir vacuum state, see
196: Fig.~\ref{fig1}b. This phenomenon, called the electron accumulation (EA) in
197: QPC, should be taken into account, vice versa for the backward tunneling
198: processes. Meanwhile, the EA may also be induced by the impurity of the
199: reservoir or a circuit with a worse transport mobility (the imperfect
200: condition of the reservoir). Literaturely, the effect of the EA is ignored
201: \cite{gurvitz,goan,mozyrsky,stace,li}. One assumes that electrons created in
202: drain (for the forward tunneling processes) will be forced into the circuit
203: instantly, no electrons are accumulated in the drain temporarily. Therefore,
204: there is no any variation to the Fermi energy. The read-out current is
205: completely contributed by the created electron of drain through the circuit,
206: as a result in the macroscopic limit. However, for a mesoscopic reservoir, a
207: variation of the Fermi energy can be generated by the EA. It is certainly
208: interesting to study what the influence of the EA effect is on the
209: decoherence of the measured qubit.
210: 
211: In general, the EA should be considered by treating the reservoir as a fully
212: non-equilibrium system. However, the reservoir is indeed an asymptotic
213: stationary state at mesoscopic scale. This is because the external bias can
214: be regarded as a bigger macroscopic reservoir compared with the QPC (see
215: Fig.~\ref{fig1}). The external bias forces the QPC reservoir to evolve into
216: an asymptotic stationary state in a short period. For such a case, we can
217: treat approximately the intermediate non-equilibrium effect as a
218: perturbation to the equilibrium reservoir. Thus, the EA can be described
219: simply by a variation of the chemical potentials in the source and the
220: drain. In the forward tunneling processes, the tunneling of $n$ electrons
221: moved from the source to the drain results in an effective increase of the
222: chemical potential in the drain $\mu _{R}\rightarrow \mu _{R}^{+}=\mu
223: _{R}+\delta \mu _{R}\left( n,\beta \right) $ and an effective decrease of
224: the chemical potential in the source $\mu _{L}\rightarrow \mu _{L}^{-}=\mu
225: _{L}-\delta \mu _{L}\left( n,\beta \right) $, see Fig.~\ref{fig1}b. Here, $%
226: \beta =1/k_{B}T$ is the inverse temperature. In thermal equilibrium state,
227: the electron distributions obey Fermi-Dirac function $F_{l}=\frac{1}{1+\exp
228: \beta \left( \varepsilon _{l}-\mu _{L}\right) }$ for the source and $F_{r}=%
229: \frac{1}{1+\exp \beta \left( \varepsilon _{r}-\mu _{R}\right) }$ for the
230: drain. Obviously, at a given temperature, the chemical potential determines
231: the number density of electrons $\bar{N}$ in reservoir. The relation of its
232: inverse function can be solved. Furthermore, because the QPC is a
233: 2-dimensional electron gas, the density of states, $g_{L}$, for the source ($%
234: g_{R}$ for the drain) can be assumed to be energy independent \cite{gurvitz}%
235: . Then the chemical potential can be expressed as $\mu _{L,R}(\bar{N},\beta
236: )=\frac{1}{\mathbf{\beta }}\ln \left[ \exp (\beta \bar{N}/g_{L,R})-1\right] $%
237: . The variation due to the EA is given by $\mu _{R,L}(\bar{N}\pm n/A,\beta
238: )-\mu _{R,L}(\bar{N},\beta )=\pm \delta \tilde{\mu}_{R,L}(n,\beta )$, where $%
239: A$ is the area of the QPC. One can find that up to the first order of $n/A$,
240: \begin{equation}
241: \delta \tilde{\mu}_{R,L}(n,\beta )=\frac{n}{A}\left( \frac{1+e^{-\beta \mu
242: _{R,L}}}{g_{R,L}}\right) .  \label{eq:variatn}
243: \end{equation}
244: 
245: In addition, the bias effect on the EA should be taken into account. The
246: external bias and the QPC-dots interaction determine electron tunnelings in
247: the QPC. The current partially induced by the QPC-dots interaction records
248: the qubit information. The bias effect simply forces the reservoir into a
249: local equilibrium state. That is, the fluctuation of the chemical potential
250: arisen from the QPC-dots interaction will be suppressed by the bias effect
251: asymptotically, $\mu _{L}\left( t\right) -\mu _{R}\left( t\right)
252: \rightarrow V_{d}$. The EA should decay to vanish (a electron-release
253: process, ER for short) by the bias effect. This ER process is also
254: associated with the imperfect condition of the reservoir. To simplify the
255: problem, the stochastic EA process (also including the effects of the bias
256: and QPC-dots interaction) is assumed to be described by
257: \begin{equation}
258: \delta \mu _{R,L}\left( n,\beta ,t\right) =\delta \tilde{\mu}_{R,L}\left(
259: n,\beta \right) e^{-\frac{\Gamma _{R,L}\left( n\right) t}{\hbar }}\cos
260: \left( \omega _{d}t\right) .  \label{eq:verls}
261: \end{equation}
262: The exponential decay factor in Eq. (\ref{eq:verls}) comes from the bias
263: effect. The detailed expression of the ER decay rate $\Gamma _{R,L}\left(
264: n\right) $ depends on the imperfect condition of the QPC associated with the
265: bias effect. In general, $\Gamma _{R,L}$ are functions of the tunneling
266: electron number $n$, and are different for the forward tunneling processes
267: and backward tunneling processes. $1/\Gamma _{R,L}\left( n\right) $ denotes
268: the release time of the ER. The processes repeat continually through the
269: transport circuit which is given by the oscillation function in Eq. (\ref
270: {eq:verls}), where $\omega _{d}$ is the frequency of the repeated processes.
271: It can be estimated that $\omega _{d}\sim v_{t}/l$ with $l$ the length of
272: the QPC in the tunneling direction and $v_{t}$ the electron drift velocity
273: in the transport circuit. Thus, Eq. (\ref{eq:verls}) describes properly the
274: stochastic nature of the tunneling processes embedded in the ER process and
275: the cyclic motion of the external transport circuit. Obviously, the
276: magnitude of the chemical potential variation decays to zero as the
277: reservoir approaches to the local equilibrium state, namely, $\mu _{L}\left(
278: t\right) -\mu _{R}\left( t\right) =V_{d}$. Similar to the forward tunneling
279: processes, the tunneling of $m$ electrons moved from the drain to the source
280: for the backward tunneling processes leads to the chemical potential
281: decreasing $\mu _{R}\rightarrow \mu _{R}^{-}=\mu _{R}-\delta \mu
282: _{R}(m,\beta ,t)$ in the drain and increasing $\mu _{L}\rightarrow \mu
283: _{L}^{+}=\mu _{L}+\delta \mu _{L}(m,\beta ,t)$ in the source. Up to the
284: first order of $\delta \mu _{R,L}\left( n,\beta ,t\right) $, the perturbed
285: Fermi-Dirac function is given by
286: \begin{equation}
287: F_{r,l}^{\pm }(n,\beta ,t)\approx F_{r,l}\pm \delta \mu _{R,L}(n,\beta ,t)%
288: \frac{\partial F_{r,l}}{\partial \mu _{R,L}}.  \label{eq:pf}
289: \end{equation}
290: 
291: As a result, even if the EA disturbs slightly the equilibrium of the
292: reservoir, it may induce decoherence to the measured qubit. This effect can
293: be explored from the master equation of the measured qubit which has the
294: form as
295: \begin{equation}
296: \frac{d}{dt}\hat{\rho}\left( t\right) =\frac{1}{i\hbar }\left[ \hat{H}_{S},%
297: \hat{\rho}\left( t\right) \right] -\hat{R}\hat{\rho}\left( t\right) ,
298: \label{eq:mr}
299: \end{equation}
300: where $\hat{\rho}(t)$ is the reduced density operator of the qubit
301: \begin{equation}
302: \hat{\rho}(t)=\sum_{n=0}^{\infty }\left( \mathrm{Tr}_{D^{(n)}}[\hat{\rho}%
303: _{tot}(t)]+\mathrm{Tr}_{\bar{D}^{(n)}}[\hat{\rho}_{tot}(t)]\right) .
304: \label{eq:dentot}
305: \end{equation}
306: $\hat{\rho}_{tot}(t)$ in Eq. (\ref{eq:dentot}) denotes the total density
307: operator of the whole system, and the conditional partial traces are defined
308: by $\mathrm{Tr}_{D^{(n)}}[\hat{O}]={\sum_{\bar{L}^{n},R^{n}}\langle D(\bar{L}%
309: ^{n},R^{n})|\hat{O}|D(\bar{L}^{n},R^{n})\rangle }$ and $\mathrm{Tr}_{\bar{D}%
310: ^{(m)}}[\hat{O}]={\sum_{L^{m},\bar{R}^{m}}\langle \bar{D}(L^{m},\bar{R}^{m})|%
311: \hat{O}|\bar{D}(L^{m},\bar{R}^{m})\rangle .}$ Here, $\mathrm{Tr}_{D^{\left(
312: n\right) }}$ ($\mathrm{Tr}_{\bar{D}^{(m)}}$) means to sum over all the
313: allowed energy levels of the reservoir with the condition of $n$ electrons
314: tunneling the QPC barrier through the forward (backward) tunneling processes
315: and then accumulating in the drain (source). $\hat{R}$ in Eq. (\ref{eq:mr})
316: denotes the dissipation operator, the detailed form of which can be found in
317: the Appendix. The qubit decoherence due to the electrical measurement is
318: governed by the dissipation term $-\hat{R}\hat{\rho}\left( t\right) $ in Eq. (\ref{eq:mr}%
319: ). This term is composed of the reservoir spectrum functions, which are the
320: Fourier transformation of the reservoir time correlation functions defined
321: by \textrm{Tr}$_{D^{(n)}(\bar{D}^{(m)})}\big[\hat{f}^{+}(t)\hat{f}(t^{\prime
322: })\hat{\rho}_{tot}\big]$ and $\mathrm{Tr}_{D^{(n)}(\bar{D}^{(m)})}\big[\hat{f%
323: }(t)\hat{f}^{+}(t^{\prime })\hat{\rho}_{tot}\big]$. The reservoir
324: fluctuation due to the QPC-dots interaction is encoded in the reservoir
325: operators $\hat{f}^{+}(t)\hat{f}(t^{\prime })$ and $\hat{f}(t)\hat{f}%
326: ^{+}(t^{\prime })$, where $\hat{f_{t}}={\sum_{lr}e^{i(\varepsilon
327: _{r}-\varepsilon _{l})t/\hbar }\hat{a}_{r}^{+}\hat{a}_{l}}$ is the electron
328: tunneling operator in the interaction picture of the reservoir. The
329: tunneling processes of electrons relative to the EA have also been taken
330: into account in these functions by treating the EA-fluctuated reservoir
331: state as the perturbed Fermi-Dirac function through Eq. (\ref{eq:pf}). As a
332: result, the decoherence of the measured qubit due to the EA effect is
333: manifested in the master equation through the first order perturbation of
334: the reservoir spectrum functions, which is associated with the perturbed
335: term $\pm \delta \mu _{R,L}(n,\beta ,t)\frac{\partial F_{r,l}}{\partial \mu
336: _{R,L}}$ of the perturbed Fermi-Dirac function in Eq. (\ref{eq:pf}). A
337: perturbation scheme taking the EA effect into account is then given as
338: follows
339: 
340: \begin{equation}
341: \hat{\rho}_{tot}(t)=\hat{\rho}_{0,tot}(t)+\xi \hat{\rho}_{1,tot}(t)+\xi ^{2}%
342: \hat{\rho}_{2,tot}(t)+\cdots ,  \label{eq:pertot}
343: \end{equation}
344: where $\xi =\mathcal{U}/\bar{\mu}$ is a perturbation parameter
345: characterizing the EA effect with $\mathcal{U}=\left( (\delta \tilde{\mu}%
346: _{L}(n,\beta )+\delta \tilde{\mu}_{R}(n,\beta )\right) /n$ and $\bar{\mu}%
347: =(\mu _{L}+\mu _{R})/2$. The zero order term $\hat{\rho}_{0,tot}(t)$ in Eq.~(%
348: \ref{eq:pertot}) represents the usual states without the EA. The higher
349: order terms describe the contribution of the EA. By taking partial trace to
350: Eq.~(\ref{eq:pertot}), we have the perturbation scheme for the reduced
351: density operator of the qubit
352: \begin{equation}
353: \hat{\rho}\left( t\right) =\hat{\rho}_{0}(t)+\xi \hat{\rho}_{1}(t)+\cdots .
354: \end{equation}
355: 
356: The QPC, as a sensitive charge detector, may be required a large electron
357: transparency and a relatively strong coupling strength to maximize the
358: detector sensitivity in the current experiments \cite{clerk,elzerman,averin}%
359: . However, we shall focus in this paper on the tunneling
360: transparency regime (the weak dot-QPC coupling). This is because,
361: as Gurvitz has pointed out, the occupation probability of a single
362: electron in the coupled dot (as a qubit) can be traced through the
363: QPC readout current in this regime \cite {gurvitz}. To take the
364: weak dot-QPC coupling into account, the second order cummulant
365: expansion technique \cite{mori,li,goan,stace} is used. Here, the
366: Markovian approximation has been introduced in the technique. It
367: should be noted that since the electron tunneling in the QPC is
368: determined by the external bias and the QPC-dots interaction, it
369: results in the qubit dynamics with three time scales: the qubit
370: decoherence time, the ER decay time due to the electron release by
371: the bias, and the correction time due to the QPC-dots interaction
372: associated with the bias, temperature and the electron
373: accumulation. In literature, the validity of the Markovian
374: approximation could depend on the time scale of the correlation
375: time \cite{gardiner}. When the correlation time is much smaller
376: than other time scales (in the Markovian regime), the
377: non-Markovian memory effect in the qubit dynamics can be ignored.
378: The qubit decoherence under the environment effect through the
379: QPC-dots interaction is simply a decay process. Taking the
380: electron accumulation effect into account, an extra qubit
381: relaxation may be induced but the enhancement should not be so
382: significant under the perturbative treatment. Therefore, the
383: Markovian approximation \cite{gardiner} is still applicable for
384: the derivation of the master equations here.
385: 
386: As a result, the master equations for the measured qubit in our perturbation
387: scheme are obtained as follows
388: \begin{eqnarray}
389: \xi ^{0} &:&\frac{d}{dt}\hat{\rho}_{0}(t)=-i\hat{L}_{D}\hat{\rho}%
390: _{0}(t)-\lambda \left[ \hat{q},\Big[\left[ \hat{G}^{(0)},\hat{\rho}%
391: _{0}(t)\right] \Big]\right] , \nonumber \\
392: \label{eq:ml0} \\
393: \xi ^{1} &:&\frac{d}{dt}\hat{\rho}_{1}(t)=-i\hat{L}_{D}\hat{\rho}%
394: _{1}(t)-\lambda \Big[\hat{q},\Big(\Big[\left[ \hat{G}^{(0)},\hat{\rho}%
395: _{1}(t)\right] \Big]  \nonumber \\
396: &&~~~~~~~~~~~~~~~~~~+\mathcal{F}\left( t\right) \Big[\left[ \hat{G}^{(1)},%
397: \hat{N}_{0}(t)\right] \Big]\Big)\Big].  \label{eq:ml1rel}
398: \end{eqnarray}
399: Eq. (\ref{eq:ml0}) is nothing but the rate equation of the qubit for the
400: reservoir in the thermal equilibrium state without the EA effect. The EA
401: effect on the measured qubit is described by Eq. (\ref{eq:ml1rel}). In Eqs. (%
402: \ref{eq:ml0},\ref{eq:ml1rel}), $\hat{L}_{D}$ is the Liouvillian operator for
403: the qubit, $\lambda =\pi g_{L}g_{R}/\hbar $, the double-bracket commutator $%
404: \left[ \left[ \hat{A},\hat{B}\right] \right] \equiv \hat{A}\hat{B}-\left(
405: \hat{A}\hat{B}\right) ^{\dagger }$, and $\mathcal{F}(t)$ represents the
406: contribution of the bias effect to the EA
407: \begin{equation}
408: \mathcal{F}\left( t\right) =\cos \left( \omega _{d}t\right) \frac{e^{-\frac{%
409: \Gamma _{R}t}{\hbar }}+re^{-\frac{\Gamma _{L}t}{\hbar }}}{1+r},r=\frac{%
410: g_{R}\left( 1+e^{-\beta \mu _{L}}\right) }{g_{L}\left( 1+e^{-\beta \mu
411: _{R}}\right) }.  \label{eq:replac3}
412: \end{equation}
413: Also, the operator $\hat{q}$ and $\hat{G}^{(k)}$ for $k=0,1$ in Eqs. (\ref
414: {eq:ml0},\ref{eq:ml1rel}) are defined as
415: \begin{eqnarray}
416: \hat{q} &=&\hat{P}_{0}-\hat{P}_{1}-\hat{P}_{2}, \nonumber  \\
417: \hat{G}^{(k)}
418: &=&\sum_{i=0,1,2}(G_{+,i}^{(k)}+G_{-,i}^{(k)})\hat{P}_{i},
419: \label{eq:qg}
420: \end{eqnarray}
421: and the operators $\hat{P}_{0,1,2}$ \cite{stace} are given by
422: \begin{eqnarray}
423: &&\hat{P}_{0}=\left( \Omega -\frac{\delta \Omega }{2}\right) 1+\frac{\delta
424: \Omega \cos \theta }{2}\left( \left| g\right\rangle \left\langle g\right|
425: -\left| e\right\rangle \left\langle e\right| \right) ,  \nonumber \\
426: &&\hat{P}_{1}=\frac{\delta \Omega \sin \theta }{2}\left| e\right\rangle
427: \left\langle g\right| ,\;\hat{P}_{2}=\frac{\delta \Omega \sin \theta }{2}%
428: \left| g\right\rangle \left\langle e\right| ,  \label{eq:p012}
429: \end{eqnarray}
430: where $\theta =\cos ^{-1}\left[ \left( E_{R}-E_{L}\right) /\gamma
431: \right] $, $E_{R,L}$ denote the energy for a single electron state
432: in right and left dot, and $\gamma =\sqrt{4\Omega _{0}^{2}+\left(
433: E_{R}-E_{L}\right) ^{2}}$ is the energy difference of the two
434: eigenstates of the charge qubit. The coupled dots structure is
435: shown by $\theta $. The operators $\hat{P}_{1,2}$ are responsible
436: for the inelastic excitation and relaxation of the electron in the
437: dots coupling with the QPC. The coefficients $G_{\pm ,i}^{\left(
438: 0\right) }$ and $G_{\pm ,i}^{\left( 1\right) }$ in Eq.
439: (\ref{eq:qg})\ are the reservoir spectrum functions without and
440: with the EA
441: \begin{eqnarray}
442: &&G_{+,0}^{(0)}=\tilde{g}^{(0)}\left( eV_{d}\right) ,\;G_{+,0}^{(1)}=-\bar{%
443: \mu}\tilde{g}^{(1)}\left( eV_{d}\right) ,  \nonumber \\
444: &&G_{+,1}^{(0)}=-\tilde{g}^{(0)}\left( eV_{d}-\gamma \right)
445: ,\;G_{+,1}^{(1)}=\bar{\mu}\tilde{g}^{(1)}\left( eV_{d}-\gamma \right) ,\;
446: \nonumber \\
447: &&G_{+,2}^{(0)}=-\tilde{g}^{(0)}\left( eV_{d}+\gamma \right) ,G_{+,2}^{(1)}=%
448: \bar{\mu}\tilde{g}^{(1)}\left( eV_{d}+\gamma \right) ,  \label{eq:bg1} \\
449: &&G_{-,i}^{(0)}=\left. G_{+,i}^{(0)}\right| _{V_{d}\rightarrow
450: -V_{d}},\;G_{-,i}^{(1)}=-\left. G_{+,i}^{(1)}\right| _{V_{d}\rightarrow
451: -V_{d}},  \label{eq:bgm}
452: \end{eqnarray}
453: and
454: \begin{equation}
455: \tilde{g}^{(0)}(x)=\frac{x}{1-e^{-\beta x}},\;\tilde{g}^{(1)}(x)=\frac{%
456: \partial }{\partial \beta }\left( \frac{\beta }{1-e^{-\beta x}}\right) ,
457: \label{eq:g01}
458: \end{equation}
459: where $V_{d}$ is the bias voltage and $e$ the electron charge. For
460: convenience, $e=1$ is used hereafter. Also, $\hat{N}_{0}(t)$ in Eq. (\ref
461: {eq:ml1rel}) is the number density operator of the tunneling electrons for
462: the reservoir in the thermal equilibrium state \cite{gurvitz}.
463: 
464: Meanwhile, the qubit information can be read out through the QPC\ current $%
465: \hat{I}=\frac{d\hat{N}}{dt}$, where $\hat{N}={\sum_{n=0}^{\infty }}n\mathrm{%
466: Tr}_{D^{(n)}}[\hat{\rho}_{tot}(t)]-{\sum_{m=0}^{\infty }}m\mathrm{Tr}_{\bar{D%
467: }^{(m)}}[\hat{\rho}_{tot}(t)].$ An extra minus sign in the second term of $%
468: \hat{N}$ is added because electrons in the backward tunneling processes
469: tunnel backwardly from drain to source. In our perturbation scheme, $\hat{I}=%
470: \frac{d}{dt}\hat{N}_{0}+\xi \frac{d}{dt}\hat{N}_{1}+\mathcal{O(}\xi
471: ^{2})+\cdots $, and $\hat{N}_{1}(t)$ is the number density operator of the
472: tunneling electrons for the reservoir fluctuated by the EA. The QPC current
473: is governed by
474: 
475: \begin{eqnarray}
476: \xi ^{0} &:&\frac{d}{dt}\hat{N}_{0}(t)=-i\hat{L}_{D}\hat{N}_{0}(t)-\lambda %
477: \Big\{\left[ \hat{q},\left[ \left[ \hat{G}^{(0)},\hat{N}_{0}(t)\right]
478: \right] \right]   \nonumber \\
479: &&~~~~~~~~~~~~~~~~~~~~~~~-\left( \hat{\overline{G}}^{(0)}\hat{\rho}_{0}(t)%
480: \hat{q}+H.c.\right) \Big\},  \label{eq:mc0} \\
481: \xi ^{1} &:&\frac{d}{dt}\hat{N}_{1}(t)=-i\hat{L}_{D}\hat{N}_{1}(t)-\lambda %
482: \Big\{\Big[\hat{q},\Big(\left[ \left[ \hat{G}^{(0)},\hat{N}_{1}(t)\right]
483: \right]   \nonumber \\
484: &&~~~~~~~~~~~~~~~~~~~~~+\mathcal{F}\left( t\right) \left[ \left[ \hat{G}%
485: ^{(1)},\hat{W}_{0}(t)\right] \right] \Big)\Big]  \nonumber \\
486: &&-\Big( \Big( \hat{\overline{G}}^{(0)}\hat{\rho}_{1}(t)+\mathcal{F}(t)\hat{%
487: \overline{G}}^{(1)}\hat{N}_{0}(t)\Big) \hat{q}+H.c.\Big) \Big\},
488: \label{eq:mc1rel}
489: \end{eqnarray}
490: where $\hat{\overline{G}}^{(k)}=\sum_{i=0,1,2}(G_{+,i}^{(k)}-G_{-,i}^{(k)})%
491: \hat{P}_{i}$ for $k=0,1$. The operator $\hat{W}_{0}(t)$ in Eq. (\ref
492: {eq:mc1rel}) corresponding to the zero order perturbation of noise spectrum
493: \cite{stace,li} is determined by the following master equation
494: \begin{eqnarray}
495: &&\frac{d}{dt}\hat{W}_{0}(t)=-i\hat{L}_{D}\hat{W}_{0}(t)-\lambda \Big\{%
496: \left[ \hat{q},\left[ \left[ \hat{G}^{(0)},\hat{W}_{0}(t)\right] \right]
497: \right]   \nonumber \\
498: &&~~~~~-\left( \left( \hat{G}^{(0)}\hat{\rho}_{0}(t)+2\hat{\overline{G}}%
499: ^{(0)}\hat{N}_{0}(t)\right) \hat{q}+H.c.\right) \Big\}.  \label{eq:wl0}
500: \end{eqnarray}
501: In the derivation of Eqs. (\ref{eq:ml1rel},\ref{eq:mc1rel}), the ER decay
502: rates $\Gamma _{R,L}$ have been assumed to be independent of the tunneling
503: electron number $n$ for simplification. The detailed derivation of the above
504: master equations can be found in the Appendix.
505: 
506: As we can see, the EA effect is described by Eqs. (\ref{eq:ml1rel},\ref
507: {eq:mc1rel}). Eqs. (\ref{eq:ml0},\ref{eq:mc0}) are simply the rate equations
508: for the thermal equilibrium state without the EA effect.  If the bias
509: current flows from the source to the drain such that electrons accumulate in
510: the drain, the EA induces a negative effective bias voltage $-\delta V_{EA}$%
511: . Then, an additional current $I_{1}=\frac{d}{dt}$Tr$\left[ \xi \hat{N}%
512: _{1}\left( t\right) \right] $ is induced from the drain to the source. If
513: the bias current flows from the drain to the source such that electrons
514: accumulate in the source, a positive effective bias voltage $\delta V_{EA}$
515: is induced and an additional current flows from the source to the drain.
516: Eq.~(\ref{eq:mc1rel}) shows a modification of the first order perturbation
517: to the transport current in the QPC. Eqs. (\ref{eq:ml0}-\ref{eq:wl0}) are
518: the main result of the theory, which will be used to explore the quantum
519: decoherence of the measured qubit in the rest of the paper.
520: 
521: \section{DEPHASING AND RELAXATION IN THERMAL EQUILIBRIUM LIMIT}
522: 
523: As we have discussed in the previous section, the EA dynamics can be
524: considered as a perturbative modification of the thermal equilibrium
525: dynamics for the reduced system. Without considering the EA effect, the
526: electronic reservoir is treated as the thermal equilibrium state, in which
527: the ER process is assumed to respond fast and effectively. The measured
528: qubit in this case can be studied by the zero-order perturbation master
529: equation Eq. (\ref{eq:ml0}). When the EA effect also becomes significant,
530: the qubit undergoes a local non-equilibrium process initially, and then
531: approaches to a stable state asymptotically due to the bias effect. In this
532: section we will study the quantum decoherence of the qubit in the thermal
533: equilibrium limit. The EA effect on the qubit decoherence will be explored
534: in the next section.
535: 
536: In order to show the intrinsic measurement effect on the decoherence of the
537: qubit, we concentrate first on the qubit with symmetric CQD, namely, both
538: dots have equal energy levels $E_{R}=E_{L}$ and the qubit state is
539: characterized by $\theta =\pi /2$. The qubit with asymmetric CQD will be
540: studied later.
541: 
542: There have been several efforts contributed on this issues using the secular
543: approximation \cite{li} or discussed in the zero temperature limit \cite
544: {stace}. Here, an analytic discussion is presented. To discuss the dephasing
545: and the relaxation of the qubit in the measurement processes, a set of
546: matrix elements for the reduced density operator $\hat{\rho}_{0}$ are
547: introduced : $\rho _{0,r}=\rho _{0,ee}-\rho _{0,gg}$, $\rho _{0,d}=\rho
548: _{0,eg}+\rho _{0,ge}$ and $i\rho _{0,p}=\rho _{0,eg}-\rho _{0,ge}$, where $%
549: \rho _{0,ij}=\left\langle i\left| \hat{\rho}_{0}\right| j\right\rangle $,
550: and $\left| i,j=g\right\rangle $ ($\left| i,j=e\right\rangle $) are the
551: ground (excited) state of the qubit. Eq. (\ref{eq:ml0}) can then be
552: expressed as
553: \begin{eqnarray}
554: &&\dot{\rho}_{0,r}\left( t\right) =-\Gamma _{0,r}\left( \rho _{0,r}\left(
555: t\right) -\rho _{0,r}\left( \infty \right) \right) ,  \nonumber \\
556: &&\Gamma _{0,r}=\eta _{d}G_{+,a}^{\left( 0\right) },\;\rho _{0,r}\left(
557: \infty \right) =\frac{G_{+,b}^{\left( 0\right) }}{G_{+,a}^{\left( 0\right) }}%
558: ,  \label{eq:l0relax}
559: \end{eqnarray}
560: and
561: \begin{equation}
562: \dot{\rho}_{0,d}\left( t\right) =\frac{\gamma }{\hbar }\rho _{0,p}\left(
563: t\right) ,\;\dot{\rho}_{0,p}\left( t\right) =-\frac{\gamma }{\hbar }\rho
564: _{0,d}\left( t\right) -\Gamma _{0,r}\rho _{0,p}\left( t\right) ,
565: \label{eq:l0deph}
566: \end{equation}
567: where $\eta _{d}=\pi g_{L}g_{R}\left( \delta \Omega \right) ^{2}/\hbar $,
568: and
569: \begin{eqnarray}
570: &&G_{+,a}^{(k)}=-\sum_{i=1,2}\left( \frac{G_{+,i}^{\left( k\right)
571: }+G_{-,i}^{\left( k\right) }}{2}\right) ,  \nonumber \\
572: &&G_{+,b}^{\left( k\right) }=\sum_{i=1,2}\left( -1\right) ^{i}\left( \frac{%
573: G_{+,i}^{\left( k\right) }+G_{-,i}^{\left( k\right) }}{2}\right) ,
574: \label{eq:locoeff}
575: \end{eqnarray}
576: with $k=0,1$.
577: 
578: According to Eq. (\ref{eq:l0relax}), the relaxation rate of the qubit is $%
579: \Gamma _{0,r}$. Applying the previous definitions in Eqs.
580: (\ref{eq:bg1}-\ref{eq:g01}), it leads to
581: \begin{equation}
582: \Gamma _{0,r}=\frac{\eta _{d}\left( V_{d}\sinh \beta V_{d}-\gamma \sinh
583: \beta \gamma \right) }{\cosh \beta V_{d}-\cosh \beta \gamma },\;\rho
584: _{0,r}\left( \infty \right) =\frac{-\gamma \eta _{d}}{\Gamma _{0,r}}.
585: \label{eq:monot}
586: \end{equation}
587: A plot of $\Gamma _{0,r}$ and $\rho _{0,r}\left( \infty \right) $ is shown
588: in Fig.~\ref{fig2}. The solid curves denote the symmetric CQD. The
589: relaxation rate $\Gamma _{0,r}$ is a positive-monotonic-increase function of
590: the temperature and bias voltage. Increasing temperature and bias voltage
591: will enhance qubit relaxation, as we expected. The asymptotic matrix element
592: $\rho _{0,r}\left( \infty \right) $ denotes an asymptotic stable state of
593: the qubit. It is a negative increase-monotonic function of temperature and
594: bias voltage with values in the range $-1\sim 0$. Because the external bias
595: can be regarded as a macroscopic reservoir (an effective heat bath) with
596: respect to the QPC, it is expected that there exists an analogy between
597: biased QPC and the heat bath \cite{mozyrsky}. A limited case is $\rho
598: _{0,r}\left( \infty \right) \rightarrow 0$ as $V_{d}>>\gamma $. The qubit
599: tends to stay in a completely random mixed state in the high bias limit.
600: This property is similar to the thermal randomness caused by the heat bath.
601: According to $\rho _{0,r}\left( \infty \right) $ in Eq. (\ref{eq:monot}),
602: the ground state occupation of the qubit shows a similar dependence on the
603: temperature effect and bias effect. Furthermore, it can be checked that $%
604: \Gamma _{0,r}\rightarrow \eta _{d}\gamma $ in the zero bias and zero
605: temperature limit. The relaxation due to QPC can not be stopped by turning
606: off bias. The QPC will exhaust energy and information of the qubit, as
607: pointed out first in \cite{mozyrsky}.
608: 
609: \begin{figure}[tbph]
610: \includegraphics*[angle=90,scale=.35]{fig2}
611: \caption{(color online). (a) The decay rate $\Gamma _{0,r}^{\theta }$ of the
612: measured qubit in the equilibrium state. The solid curves are for symmetric
613: CQD $\left( \gamma =2\Omega _{0}\right) $ and the dash curves for asymmetric
614: CQD $\left( \gamma =10\Omega _{0}\right) $. (b) The asymptotic reduced
615: density matrix $\rho _{0,r}^{\theta }\left( \infty \right) $ of the measured
616: qubit. Two branches corresponds to the symmetric CQD (solid curves) and the
617: asymmetric CQD with $\gamma =10\Omega _{0}$(the dash curves), respectively.}
618: \label{fig2}
619: \end{figure}
620: 
621: The qubit dephasing is governed by the coupled differential equations (\ref
622: {eq:l0deph}), the solutions of which are a linear combination of two decay
623: modes
624: \begin{equation}
625: c_{1}e^{-\Gamma _{0,p}^{+}t}+c_{2}e^{-\Gamma _{0,p}^{-}t}
626: \label{eq:dephas_mode}
627: \end{equation}
628: with dephasing rates $\Gamma _{0,p}^{\pm }=\frac{1}{2}\Gamma _{0,r}\pm \frac{%
629: 1}{2}\Gamma _{0,r}$ $\times \sqrt{1-\left( 2\gamma /\hbar \Gamma
630: _{0,r}\right) ^{2}}$, respectively. The mean dephasing rate is $\Gamma
631: _{0,r}/2$. The coefficients $c_{1,2}$ in Eq. (\ref{eq:dephas_mode}) depend
632: on the initial condition of the qubit. We then have the result
633: \begin{equation}
634: \rho _{0,p}\left( \infty \right) \rightarrow 0,\quad \rho _{0,d}\left(
635: \infty \right) \rightarrow 0.  \label{asy_pd}
636: \end{equation}
637: The qubit completely dephasing asymptotically. In addition, two decay modes
638: show different decay behavior in the regime $\hbar \Gamma _{0,r}/2>\gamma $.
639: A plot of $\Gamma _{0,p}^{\pm }$ versus $V_{d}$ is shown in Fig.~\ref{fig3}.
640: The curves with non-smooth turning point denote that $\Gamma _{0,p}^{\pm }$
641: has a transition from a degenerate regime $\hbar \Gamma _{0,r}/2<\gamma $,
642: in which both modes have the same dephasing rate $\Gamma _{0,r}/2$, to a
643: non-degenerate regime along $V_{d}$ axis. In the degenerate regime, only the
644: real part of $\Gamma _{0,p}^{\pm }$ is plotted. In Fig. \ref{fig3}a, the
645: fast decay mode, $\Gamma _{0,p}^{+}$, shows that the dephasing rate
646: increases as increasing temperature and bias voltage. In the high
647: temperature and high bias voltage limit, it approaches to the linear
648: dependence as the usual expectation. However, for the slow decay mode, $%
649: \Gamma _{0,p}^{-}$, Fig. \ref{fig3}b shows an upside-down dependence on the
650: temperature and bias voltage. At a given bias voltage, the dephasing rate of
651: the slow decay mode decreases with temperature increasing. It also shows
652: that the dephasing rate decreases with the bias voltage increasing at a
653: given temperature. The minimum dephasing rate appears in the large bias
654: voltage and large temperature limit. This phenomenon is very different from
655: the behavior of the fast decay mode that we have known. Even for the case of
656: high density of state (i.e. with small value of $\gamma /\hbar \eta _{d}$),
657: the phenomenon is still preserved. However, it should be noted that for a
658: quantum information process, quantum computations must deal with all
659: possible initial conditions. Therefore, the decoherence of the qubit is
660: characterized by the larger dephasing rate $\Gamma _{0,p}^{+}$ rather than
661: the slow one, $\Gamma _{0,p}^{-}$, unless an ancillary operation on the
662: qubit can be added to limit the initial conditions to a sub-domain belonging
663: to the slow decay mode.
664: \begin{figure}[tbph]
665: \includegraphics*[angle=90,scale=.35]{fig3}
666: \caption{(color online). (a) The dephasing rate $\Gamma _{0,p}^{+}$ of the
667: fast mode of the measured qubit in the equilibrium state. (b) The dephasing
668: rate $\Gamma _{0,p}^{-}$ of the slow mode for the measured qubit. The
669: figures are plotted with the fixed parameter $\gamma /\hbar \eta _{d}=\Omega
670: _{0}$.}
671: \label{fig3}
672: \end{figure}
673: 
674: Next we shall study the decoherence of the qubit with asymmetric CQD. In
675: this case, Eq. (\ref{eq:ml0}) can be expressed as
676: \begin{eqnarray}
677: &&\dot{\rho}_{0,d}(t)=\frac{\gamma }{\hbar }\rho _{0,p}(t)-\eta _{d}\cos
678: \theta \Big[\cos \theta G_{+}^{(0)}\rho _{0,d}(t)  \nonumber \\
679: &&~~~~~~~~~~~~~~~~~~+\sin \theta \Big(G_{+,b}^{(0)}-G_{+,a}^{(0)}\rho
680: _{0,r}(t)\Big) \Big],  \label{eq:asymast1} \\
681: &&\dot{\rho}_{0,r}(t)=\eta _{d}\sin \theta \Big[\cos \theta G_{+}^{(0)}\rho
682: _{0,d}(t)  \nonumber \\
683: &&~~~~~~~~~~~~~~~~~~+\sin \theta \Big(G_{+,b}^{(0)}-G_{+,a}^{(0)}\rho
684: _{0,r}(t)\Big) \Big],  \label{eq:asymast2} \\
685: &&\dot{\rho}_{0,p}(t)=\frac{-\gamma }{\hbar }\rho _{0,d}(t)-\eta _{d}\Big[%
686: G_{+}^{(0)}\cos \theta ^{2}  \nonumber \\
687: &&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~+G_{+,a}^{(0)}\sin \theta ^{2}\Big]\rho
688: _{0,p}(t),  \label{eq:asymast3}
689: \end{eqnarray}
690: where $G_{+}^{\left( 0\right) }=G_{+,0}^{\left( 0\right) }+G_{-,0}^{\left(
691: 0\right) }=V_{d}\coth \left( \beta V_{d}/2\right) $. This is a set of
692: coupled differential equations. There is no analytic form to define the
693: decoherence rate like the case in the symmetric CQD. However, in the large
694: bias limit $\gamma \ll V_{d}$ we can find from the first two master
695: equations the relation $\dot{\rho}_{0,d}\left( t\right) /\dot{\rho}%
696: _{0,r}\left( t\right) \simeq -\cot \theta $. Then the relaxation rate can be
697: obtained
698: \begin{equation}
699: \Gamma _{0,r}^{\theta }=\eta _{d}\left( \cos \theta ^{2}G_{+}^{\left(
700: 0\right) }+\sin \theta ^{2}G_{+,a}^{\left( 0\right) }\right) .
701: \label{eq:relaxasy}
702: \end{equation}
703: It can been shown that three decay rates from master equations (\ref
704: {eq:asymast1}-\ref{eq:asymast3}) correspond to the roots of the following
705: equation
706: \begin{equation}
707: 0=y^{3}-2\Gamma _{0,r}^{\theta }y^{2}+\left( \frac{\gamma ^{2}+\left( \hbar
708: \Gamma _{0,r}^{\theta }\right) ^{2}}{\hbar ^{2}}\right) y-\eta
709: _{d}G_{+,a}^{\left( 0\right) }\left( \frac{\gamma \sin \theta }{\hbar }%
710: \right) ^{2}.  \label{eq:decayroot}
711: \end{equation}
712: The mean value of these roots is $2\Gamma _{0,r}^{\theta }/3$. $\Gamma
713: _{0,r}^{\theta }$ is also plotted in Fig.~\ref{fig2}a. All curves show a
714: positive-monotonic-increase with the temperature and bias voltage. This bias
715: and temperature dependence of the mean decay rate approaches to be linear in
716: the high bias regime. In the low temperature limit, it becomes linear for
717: almost the whole range of the bias voltage. In addition, it can be found in
718: Fig.~\ref{fig2}a that the asymmetric CQD has a longer decay time than the
719: symmetric CQD. It indicates that the influence of the device structure on
720: qubit decoherence can also become significant in the low temperature and low
721: bias regime.
722: 
723: The asymptotic behavior of Eqs. (\ref{eq:asymast1}-\ref{eq:asymast3}) can be
724: solved easily
725: \begin{eqnarray}
726: && \rho _{0,d}^{\theta }\left( \infty \right) =\rho _{0,p}^{\theta }\left(
727: \infty \right) \rightarrow 0,  \label{eq:asymdp} \\
728: && \rho _{0,r}^{\theta }\left( \infty \right) \rightarrow \frac{-\gamma
729: \left( \cosh \beta V_{d}-\cosh \beta \gamma \right) }{V_{d}\sinh \beta
730: V_{d}-\gamma \sinh \beta \gamma }.  \label{eq:asymr}
731: \end{eqnarray}
732: $\rho _{0,r}^{\theta }\left( \infty \right) $ implicitly depends on the
733: device structure through $\gamma $. The plot of $\rho _{0,r}^{\theta }\left(
734: \infty \right) $ in Fig.~\ref{fig2}b shows that there are two branches
735: corresponding to the symmetric (solid curves) and the asymmetric (dash
736: curves) CQD, respectively. At a given bias voltage, the symmetric CQD is
737: asymptotically forced into a more random mixed state in comparison with the
738: asymmetric CQD. This is because the device structure effectively modifies
739: the bias-dependence of the qubit coherence.
740: 
741: \section{EA EFFECT ON QUBIT}
742: 
743: Now, we turn to discuss the properties of the EA and the decoherence of the
744: charge qubit due to the EA effect.
745: 
746: \subsection{Properties of the EA}
747: 
748: The EA can be characterized by the EA induced current $I_{1}\left( t\right) =%
749: \frac{d}{dt}$Tr$\left[ \xi \hat{N}_{1}\left( t\right) \right] $. $%
750: I_{1}\left( t\right) $ is plotted in Fig.~\ref{fig4}, which is calculated
751: according to the resulted master equations. The result can be studied in
752: detail by the analysis of Eq. (\ref{eq:mc1rel}). In the right-hand side of
753: Eq. (\ref{eq:mc1rel}), because the traces of $-i\hat{L}_{D}\hat{N}_{1}\left(
754: t\right) $ and commutators vanish, the terms $\lambda \mathcal{F}\left(
755: t\right) \left( \hat{\overline{G}}^{(1)}\hat{N}_{0}\left( t\right) \hat{q}%
756: +H.c.\right) \equiv \hat{K}_{1}$ and $\lambda \left( \hat{\overline{G}}^{(0)}%
757: \hat{\rho}_{1}(t)\hat{q}+H.c.\right) \equiv \hat{K}_{0}$ completely
758: determine the EA current. Also, the factor $\mathcal{F}\left( t\right) $ of $%
759: \hat{K}_{1}$ is exponential decay in time, and $\hat{\rho}_{1}\left(
760: t_{0}\right) \ll \hat{N}_{0}\left( t_{0}\right) $. Thus, $\hat{K}_{1}$ is
761: initially dominant, and then $\hat{K}_{0}$ becomes significant as $\hat{K}%
762: _{1}$ exponentially decays to zero. The linear increase of the EA current in
763: the beginning is governed mainly by $\hat{K}_{1}$, see Fig.~{\ref{fig4}. }
764: 
765: \begin{figure}[tbph]
766: \includegraphics*[angle=90,scale=.6]{fig4}
767: \caption{The EA current $\emph{I}_{1}\left( t\right) $ in the QPC. The qubit
768: is initially in the $\left| L\right\rangle $ state without energy-level
769: offset. The following measurement parameters are used to plot figures : $%
770: \Omega =\Omega _{0}$, $\delta \Omega =0.1\Omega _{0}$, $\beta =1/\Omega _{0}$
771: , $\mu _{L}=100\Omega _{0}$, $(V_{d}=10\Omega _{0})$ and $g_{L,R}=2/\Omega
772: _{0}$. The initial input number of electrons in the source is set to be $100$%
773: . There is no electron accumulation at initial time. Two cases with
774: different ER decay rate are plotted.}
775: \label{fig4}
776: \end{figure}
777: 
778: Analytically, $\hat{N}_{0}\left( t\right) $ can be solved in the asymptotic
779: limit
780: \begin{eqnarray}
781: &&n_{0,gg}\left( t\right) =c_{g0}+c_{g1}t,\;n_{0,ee}\left( t\right)
782: =c_{e0}+c_{e1}t,  \nonumber \\
783: &&c_{e1}=c_{0}\left( G_{+,1}^{\left( 0\right) }+G_{-,1}^{\left( 0\right)
784: }\right) ,\;c_{g1}=c_{0}\left( G_{+,2}^{\left( 0\right) }+G_{-,2}^{\left(
785: 0\right) }\right) ,  \nonumber \\
786: &&  \label{eq:n1slope}
787: \end{eqnarray}
788: where $c_{g0}$ and $c_{e0}$ are simply constants, $c_{0}=-\lambda \Omega
789: ^{2}V_{d}/G_{+,a}^{\left( 0\right) }$, the higher order contributions of $%
790: \delta \Omega $ to the coefficient $c_{0}$ have also been ignored. For the
791: weak interaction coupling, it can be numerically checked that $%
792: n_{0,ge}\left( t\right) \ll n_{0,gg(ee)}\left( t\right) $ in the asymptotic
793: limit. We then obtain
794: \begin{equation}
795: \text{Tr}\left[ \hat{K}_{1}\right] =-\mathcal{F}\left( t\right) \left(
796: \left( 2\lambda \Omega ^{2}\right) ^{2}V_{d}t+\text{constant}\right) .
797: \label{eq:k1asympt}
798: \end{equation}
799: In the beginning (the $\hat{K}_{1}$ dominant regime), the ER decay is
800: adiabatic, namely, $\mathcal{F}\left( t\right) \approx $ constant ($\equiv
801: \mathcal{F}^{\prime }$). It leads to
802: \begin{equation}
803: I_{1}\left( t\right) \approx -\mathcal{F}^{\prime }\xi \left( 2\lambda
804: \Omega ^{2}\right) ^{2}V_{d}t+\text{constant}.  \label{eq:ea1vnon}
805: \end{equation}
806: This shows why the EA current is linear increasing in time, as one can see
807: in Fig.~\ref{fig4}.
808: 
809: In the crossover regime in Fig.~\ref{fig4}, the EA is slowed down by the
810: bias. The maximum amount of accumulated electrons occurs at the valley of
811: the $I_{1}$-curves, in which the maximum reverse EA current is induced.
812: Obviously, the smaller the ER decay rate $\Gamma _{R,L}$, the stronger the
813: EA as shown in Fig.~\ref{fig4}.
814: 
815: Due to the bias, electrons accumulated in the QPC are completely exhausted
816: through the ER process. Tr$\left[ \hat{K}_{1}\right] $ of Eq. (\ref
817: {eq:mc1rel}) vanishes in this stage, and Tr$\left[ \hat{K}_{0}\right] $
818: becomes dominant. The asymptotic current can then be solved analytically
819: \begin{eqnarray}
820: I_{1}(\infty ) &=&\frac{d}{dt}\text{Tr}\left[ \xi \hat{N}_{1}(t)\right] =\xi
821: \text{Tr}[\hat{K}_{0}(\infty )]  \nonumber \\
822: &=&2\lambda \xi V_{d}\left( \Omega ^{2}-\Omega \delta \Omega \left( 1+\rho
823: _{1,r}\left( \infty \right) \cos \theta \right) \right) ,  \nonumber \\
824: &&  \label{eq:cur1asy}
825: \end{eqnarray}
826: where the higher order contribution from $\delta \Omega ^{2}$ has been
827: ignored, and the first-order relaxation shift $\rho _{1,r}\left( \infty
828: \right) $ is given in Eq. (\ref{eq:l1rasympt}). It is worth noting that the
829: qubit density $\rho _{1,r}\left( t\right) $ records the history of the EA
830: effect. Even the accumulated electrons has been exhausted in this stage, $%
831: \rho _{1,r}\left( t\right) $ in $\hat{K}_{0}(t)$ performs as an EA
832: background field and affects the electron tunneling in the QPC. Also, the
833: time independent relaxation shift $\rho _{1,r}\left( \infty \right) $ leads
834: to a constant transmission probability for the first-order electron
835: tunneling. Finally, the stable behavior of the EA current is reached, see
836: Fig.~\ref{fig4}. As a result, with the EA effect, the asymptotic current up
837: to the first order perturbation is
838: \begin{eqnarray}
839: &&I_{0}(\infty )+I_{1}(\infty )=2\lambda \left( \Omega -\delta \Omega
840: \right) ^{2}V_{d}  \nonumber \\
841: &&~~~~~~~~~~~~~~~~~~~~+\frac{\lambda \delta \Omega ^{2}}{2}\frac{\left(
842: V_{d}^{2}-\gamma ^{2}\right) \sinh \beta V_{d}}{V_{d}\sinh \beta
843: V_{d}-\gamma \sinh \beta \gamma }  \nonumber \\
844: &&~~~~~~~~~ +2\lambda \xi V_{d}\left( \Omega ^{2}-\Omega \delta \Omega
845: \left( 1+\rho _{0,r}^{\theta }\left( \infty \right) \cos \theta \right)
846: \right).  \nonumber \\
847: &&
848: \end{eqnarray}
849: 
850: \subsection{Decoherence induced by EA effect}
851: 
852: The qubit relaxation and dephasing due to the EA effect are studied in this
853: subsection. Eq.~(\ref{eq:ml1rel}) describes this decoherence process. The
854: terms $\lambda \mathcal{F}\left( t\right) \left[ \left[ \left[ \hat{G}%
855: ^{\left( 1\right) },\hat{N}_{0}\left( t\right) \right] \right] ,\hat{q}%
856: \right] \equiv \hat{K}_{1}^{\prime }$ in Eq. (\ref{eq:ml1rel}) is related to
857: $\hat{K}_{1}$ of Eq. (\ref{eq:mc1rel}) through the relation $\hat{N}_{k}={%
858: \sum_{n=0}^{\infty }}n\hat{\rho}_{k,f}^{(n)}(t)-{\sum_{m=0}^{\infty }}m\hat{%
859: \rho}_{k,b}^{(m)}(t)$. This term arises from the EA, and the induced
860: relaxation is suppressed by the bias. To understand this property, let us
861: look at the asymptotic matrix elements $\left\langle i\left| \hat{K}%
862: _{1}^{\prime }(\infty )\right| i\right\rangle $ first. Applying the previous
863: result Eq. (\ref{eq:n1slope}) and the asymptotic analysis, we obtain, up to
864: the order $\delta \Omega $,
865: \begin{eqnarray}
866: &&\left\langle g\left| \hat{K}_{1}^{\prime }(\infty )\right| g\right\rangle
867: =K_{g0},  \nonumber \\
868: &&\left\langle e\left| \hat{K}_{1}^{\prime }(\infty )\right| e\right\rangle
869: =-\left\langle g\left| \hat{K}_{1}^{\prime }(\infty )\right| g\right\rangle ,
870: \label{eq:kg0}
871: \end{eqnarray}
872: where
873: \[
874: K_{g0}\approx \lambda \Omega \delta \Omega \mathcal{F}\left( t\right) V_{d}%
875: \frac{(G_{+,b}^{\left( 0\right) }-G_{+,a}^{\left( 0\right)
876: })(G_{+,a}^{\left( 1\right) }+G_{+,b}^{\left( 1\right) })}{(G_{+,a}^{\left(
877: 0\right) })^{2}},
878: \]
879: and $G_{+,(a,b)}^{\left( 0,1\right) }$ is defined in Eq.~(\ref{eq:locoeff}).
880: It can be checked that $K_{g0}$ is positive because of $(G_{+,b}^{\left(
881: 0\right) }-G_{+,a}^{\left( 0\right) })\leq 0$ and $-1<(G_{+,a}^{\left(
882: 1\right) }+G_{+,b}^{\left( 1\right) })\leq 0$. Since $\hat{\rho}_{1}\left(
883: t_{0}\right) \ll \hat{N}_{0}\left( t_{0}\right) $, $\frac{d}{dt}\left\langle
884: i\left| \hat{\rho}_{1}(t)\right| i\right\rangle \approx \left\langle i\left|
885: \hat{K}_{1}^{\prime }\right| i\right\rangle $ before the bias effect becomes
886: active (i.e. in the EA effect dominant regime). We then have
887: \begin{eqnarray}
888: \left\langle g\left| \hat{\rho}_{1}\left( t\right) \right| g\right\rangle
889: &\approx &\left. K_{g0}\right| _{\mathcal{F}=1}t\text{ + constant},
890: \nonumber \\
891: \left\langle e\left| \hat{\rho}_{1}\left( t\right) \right| e\right\rangle
892: &\approx &-\left. K_{g0}\right| _{\mathcal{F}=1}t\text{ + constant}.
893: \label{eq:relax}
894: \end{eqnarray}
895: This linear time-dependence behavior indeed indicates an extra \textit{qubit
896: relaxation} due to the EA effect. The numerical results in Fig. ~\ref{fig5}a
897: (plotted by green and black curves) roughly show this linear time-dependence
898: of the relaxation. In Fig. \ref{fig5}a, the solid (dash) lines correspond to
899: the ground (excited) state, respectively. Instructively, a numerical result
900: for $\Gamma _{L,R}=0$ is plotted in Fig.~\ref{fig5}b. A perfect linear
901: character is obtained. The induced relaxation from the \textit{pure} EA
902: effect obeys the linear time dependence, as shown by Eq. (\ref{eq:relax}).
903: 
904: \begin{figure}[tbph]
905: \includegraphics*[angle=90,scale=.55]{fig5}
906: \caption{(color online). The time evolution of the reduced density matrix of
907: the first order perturbation. The symmetric CQD is simulated. The qubit is
908: initially in $\left| L\right\rangle $ state with $\xi \mathtt{Tr}[\hat{\rho}%
909: _{1}\left( t=0\right) ]=0.01 $. The input initial electron number is $100$.
910: The following measurement parameters are used : $\Omega =\Omega _{0}$, $%
911: \delta \Omega =0.1\Omega _{0}$, $\beta =1/\Omega _{0}$ , $\mu _{L}=100\Omega
912: _{0}$, $g_{L,R}=2/\Omega _{0}$, $\omega _{d}=0$ and $\mu _{R}=\mu _{L}-V_{d}$
913: with $V_{d}=0$ and $\Gamma _{L,R}=0.15\Omega _{0}$ for red curve, $%
914: V_{d}=10\Omega _{0}$ and $\Gamma _{L,R}=0.15\Omega _{0}$ black curve, $%
915: V_{d}=10\Omega _{0}$ and $\Gamma _{L,R}=0.3\Omega _{0} $ green curve, and $%
916: V_{d}=10\Omega _{0}$ and $\Gamma _{L,R}=0$ cyan curve, respectively. (a) The
917: qubit relaxation. The time evolution of the matrix elements $\rho _{1g}$ $%
918: \left( \rho _{1e}\right) $ for the qubit in the ground (excited) state are
919: plotted by solid (dash) curves, respectively. (b) The qubit relaxation under
920: the pure EA effect ($\Gamma _{L,R}=0$). (c) and (d) The qubit dephasing. The
921: time evolution of the matrix element $\rho _{1d}$ and $\rho _{1p}$ are
922: plotted, respectively.}
923: \label{fig5}
924: \end{figure}
925: 
926: According to Eq.~(\ref{eq:kg0}), it is obvious that this relaxation process
927: is slowed down by the bias because $K_{g0}$ reduces to zero by the decay
928: factor $\mathcal{F}(t)$. This slow-down effect causes the qubit decay to the
929: lowest energy state (in the first order perturbation) in the relaxation
930: process, see Fig.~\ref{fig5}a. Then, the term $\lambda \left[ \left[ \left[
931: \hat{G}^{(0)},\hat{\rho}_{1}(t)\right] \right] ,\hat{q}\right] \equiv \hat{K}%
932: _{0}^{\prime }$ of Eq. (\ref{eq:ml1rel}) becomes dominant. The qubit state
933: undergoes a transition from the relaxation to excitation, see Fig.~\ref{fig5}%
934: a. After the accumulated electrons exhausted, this excitation is completely
935: governed by $\hat{K}_{0}^{\prime }$. In addition, the black and green curves
936: in Fig.~\ref{fig5}a show that a stronger EA effect (i.e. with smaller ER
937: decay rate) forces the qubit to experience a larger relaxation shift (i.e. a
938: lower energy state). The asymptotic relaxation shift $\rho _{1,r}(\infty) $
939: can be easily solved
940: \begin{eqnarray}
941: \rho _{1,r}\left( \infty \right) &=&\left\langle e\left| \hat{\rho}%
942: _{1}\left( \infty \right) \right| e\right\rangle -\left\langle g\left| \hat{%
943: \rho}_{1}\left( \infty \right) \right| g\right\rangle  \nonumber \\
944: &\rightarrow &\frac{-\gamma \left( \cosh \beta V_{d}-\cosh \beta \gamma
945: \right) }{V_{d}\sinh \beta V_{d}-\gamma \sinh \beta \gamma },
946: \label{eq:l1rasympt}
947: \end{eqnarray}
948: which is independent of the ER decay rate. As a result, the total qubit
949: relaxation up to the first order perturbation is
950: \begin{eqnarray}
951: &&\rho _{tot,r}\left( \infty \right) =\frac{-\gamma \left( \cosh \beta
952: V_{d}-\cosh \beta \gamma \right) }{V_{d}\sinh \beta V_{d}-\gamma \sinh \beta
953: \gamma }  \nonumber \\
954: &&\times \Big[1+\frac{2}{\left( \mu _{L}+\mu _{R}\right) }\left( \frac{%
955: 1+e^{-\beta \mu _{L}}}{Ag_{L}}+\frac{1+e^{-\beta \mu _{R}}}{Ag_{R}}\right) %
956: \Big].  \nonumber \\
957: &&
958: \end{eqnarray}
959: The perturbed term comes from the EA effect.
960: 
961: We may also discuss the low bias voltage limit, $V_{d}\ll \gamma $. It can
962: be checked that $\hat{G}^{\left( 1\right) }\rightarrow 0$ as $%
963: V_{d}\rightarrow 0$. Thus, $\hat{K}_{1}^{\prime }$ in Eq. (\ref{eq:ml1rel})
964: vanishes, and the EA makes almost no effect on the qubit dynamics. The
965: result is shown by the red curve in Fig.~\ref{fig5}a. The qubit only
966: experiences a relaxation process without the excitation one.
967: 
968: Next, we shall calculate the decoherence rate of the qubit. For a low bias
969: voltage, because of $\hat{G}^{\left( 1\right) }\rightarrow 0$ as $%
970: V_{d}\rightarrow 0$, the dynamic structure of Eq. (\ref{eq:ml0}) and (\ref
971: {eq:ml1rel}) are almost the same. The decay modes of the qubit contributed
972: by the zero order and the first order perturbation are closer. The EA effect
973: can only enhance the qubit relaxation. Dynamically, the decoherence rate of
974: the qubit does not be speeded up by the EA effect. According to Eq. (\ref
975: {eq:monot}), we obtain the total relaxation time $T_{tot,r}$ and the
976: dephasing time $T_{tot,p}$ of the qubit with the EA effect
977: \begin{equation}
978: T_{tot,r}=\frac{1}{\left. \Gamma _{0,r}\right| _{V_{d}\rightarrow 0}}=\frac{%
979: \cosh \beta \gamma -1}{\eta _{d}\gamma \sinh \beta \gamma }%
980: ,\;T_{tot,p}=2T_{tot,r}  \label{eq:dtsym}
981: \end{equation}
982: for the symmetric CQD in the low bias limit.
983: 
984: With bias voltage increasing, the first-order qubit dephasing rate can not
985: be solved exactly. We study the qubit dephasing according to Eq. (\ref
986: {eq:ml1rel}), or in terms of the set of coupled differential equations for
987: the symmetric CQD
988: \begin{eqnarray}
989: \dot{\rho}_{1,d}\left( t\right) &=&\frac{\gamma }{\hbar }\rho _{1,p}\left(
990: t\right) ,  \nonumber \\
991: \dot{\rho}_{1,p}\left( t\right) &=&-\frac{\gamma }{\hbar }\rho _{1,d}\left(
992: t\right) -\Gamma _{0,r}\rho _{1,p}\left( t\right)  \nonumber \\
993: &&~~~~~~~-\Gamma _{1,r}\mathcal{F}\left( t\right) n_{0,p}\left( t\right) ,
994: \label{eq:1dephas}
995: \end{eqnarray}
996: where
997: \begin{equation}
998: \Gamma _{1,r}=-\eta _{d}G_{+,a}^{\left( 1\right) },  \label{eq:iearr}
999: \end{equation}
1000: $\rho _{1,d}=\rho _{1,eg}+\rho _{1,ge}$, $i\rho _{1,p}=\rho _{1,eg}-\rho
1001: _{1,ge}$ and $in_{0,p}=n_{0,ge}-n_{0,eg}$. The term $\Gamma _{0,r}\rho
1002: _{1,p}\left( t\right) $ in Eq. (\ref{eq:1dephas}) denotes the qubit
1003: dephasing arisen from the QPC-dots interaction associated with the bias,
1004: while the term $\Gamma _{1,r}\mathcal{F}\left( t\right) n_{0,p}\left(
1005: t\right) $ is resulted from the EA. Because $\Gamma _{1,r}$ is proportional
1006: to the mean chemical potential $\bar{\mu}$ $(=\frac{\mu _{L}+\mu _{R}}{2})$%
1007: ,\ $\Gamma _{1,r}$ can be larger than $\Gamma _{0,r}$ for $\bar{\mu}%
1008: >V_{d}\gg \gamma $. The qubit dephasing is mainly determined by the ER decay
1009: rate and the mean chemical potential, it does not so sensitively
1010: depend on the bias voltage. However, in the low bias limit,
1011: especially for $V_{d}\ll \gamma $, $G_{+,a}^{\left( 1\right)
1012: }\rightarrow 0$ (i.e. $\Gamma _{1,r}\rightarrow 0$). The qubit
1013: dephasing rate becomes much smaller. The numerical plot of the
1014: time evolution of the qubit off-diagonal matrix element shows the
1015: coincident result, as one can see in Fig.~\ref{fig5}c and d.
1016: 
1017: \subsection{Decoherenceless EA effect in the low temperature regime}
1018: 
1019: We have pointed out that the qubit undergoes an extra relaxation under the
1020: EA effect. However, this EA induced relaxation can be suppressed in the low
1021: temperature regime. The symmetric CQD is used to check this result. The
1022: qubit relaxation rate in the first order perturbation can be deduced from
1023: Eq. (\ref{eq:ml1rel})
1024: \begin{equation}
1025: \Gamma _{1,r}(t)\equiv \frac{-\dot{\rho}_{1,r}\left( t\right) }{(\rho
1026: _{1,r}\left( t\right) -\rho _{1,r}\left( \infty \right) )}=\Gamma _{0,r}+%
1027: \mathcal{K}\left( t\right) \Gamma _{1,r}
1028: \end{equation}
1029: where $\mathcal{K}\left( t\right) =\mathcal{F}\left( t\right) n_{0,r}\left(
1030: t\right) /(\rho _{1,r}\left( \infty \right) -\rho _{1,r}\left( t\right) )$,
1031: and $n_{0,r}\left( \infty \right) =n_{0}G_{+,b}^{\left( 1\right)
1032: }/G_{+,a}^{\left( 1\right) }=0$ has been used ($n_{0}\equiv $Tr$\left[ \hat{N%
1033: }_{0}(t_{0})\right] $), and $\Gamma _{1,r}$ is defined in Eq. (\ref{eq:iearr}%
1034: ). The relaxation rate $\Gamma _{1,r}(t)$ is time dependent. Due to the EA
1035: effect, initially, the qubit severely relaxes with the relaxation rate $%
1036: \Gamma _{1,r}(t)\approx \Gamma _{1,r}\mathcal{K}\left( t\right) $ for $\bar{%
1037: \mu}\gg V_{d}$. After the EA effect suppressed by the bias, the relaxation
1038: rate asymptotically reduces to the zero-order $\Gamma _{0,r}$. Here, $%
1039: \mathcal{K}\left( t\right) $ is a fluctuant factor. The EA induced
1040: relaxation rate $\Gamma _{1,r}$ is plotted in Fig.~\ref{fig6}. It shows that
1041: $\Gamma _{1,r}$ has a very different behavior from $\Gamma _{0,r}$. For a
1042: large mean chemical potential ($\bar{\mu}\gg V_{d}$) in which $\bar{\mu}$ is
1043: almost bias voltage independent, the electron source of the EA is excited
1044: from the energy levels below the Fermi surface. The variation of the
1045: chemical potential $\delta V_{EA}$ induced by the EA is limited by the
1046: chemical potential. The $\delta V_{EA}$-dependence of the qubit relaxation
1047: rate leads to a bounded phenomenon, as shown in Fig.~\ref{fig6}.
1048: \begin{figure}[tbph]
1049: \includegraphics*[angle=90,scale=.6]{fig6}
1050: \caption{(color online). The EA induced relaxation rate. The curves are
1051: plotted with different temperatures.}
1052: \label{fig6}
1053: \end{figure}
1054: 
1055: The relaxation rate $\Gamma _{1,r}$ also shows an anti-symmetric bias
1056: dependence (see Fig.~\ref{fig6}). This feature can be understood as follows.
1057: There are two kind of electron-tunneling correlations, $G_{+,i}^{(k)}$ and $%
1058: G_{-,i}^{(k)}$, are involved. $G_{+,i}^{(k)}$ ($G_{-,i}^{(k)}$) corresponds
1059: to the electron tunneling that one electron is from the source to the drain
1060: at $\tau $ $(t)$ and the other is from the drain to the source at $t$ $(\tau
1061: )$. Explicitly, $G_{+(-),i}^{(k)}$ is associated with $\mathrm{Tr}_{D^{n}}[%
1062: \hat{f}_{t}^{+}\hat{f}_{\tau }\hat{\rho}_{I}(t)]$ $(\mathrm{Tr}_{D^{n}}[\hat{%
1063: f_{t}}\hat{f}_{\tau }^{+}\hat{\rho}_{I}(t)])$ for forward tunneling
1064: processes and $\mathrm{Tr}_{\bar{D}^{m}}[\hat{f_{t}}^{+}\hat{f}_{\tau }\hat{%
1065: \rho}_{I}(t)]$ $(\mathrm{Tr}_{\bar{D}^{m}}[\hat{f_{t}}\hat{f}_{\tau }^{+}%
1066: \hat{\rho}_{I}(t)])$ for backward tunneling processes. The relaxation rate
1067: is related to $G_{+,i}^{(k)}$ and $G_{-,i}^{(k)}$ through the relation $%
1068: \Gamma _{k,r}=\eta _{d}\sum_{i=1,2}\left( G_{+,i}^{\left( k\right)
1069: }+G_{-,i}^{\left( k\right) }\right) /2$ with $k=0,1$. For $k=0$, the
1070: electron-tunneling correlation is simply governed by the QPC-dots
1071: interaction and the bias, no EA involved. It can be checked that a relation
1072: of the bias symmetry holds
1073: \begin{equation}
1074: \frac{G_{+,i}^{\left( 0\right) }(V_{d})+G_{-,i}^{\left( 0\right) }(V_{d})}{2}%
1075: =\frac{G_{+,i}^{\left( 0\right) }(-V_{d})+G_{-,i}^{\left( 0\right) }(-V_{d})%
1076: }{2}.
1077: \end{equation}
1078: The $\Gamma _{0,r}$ does not be changed by reversing the bias ($%
1079: V_{d}\rightarrow -V_{d}$), due to the fact that the qubit relaxation caused
1080: by the electron tunneling in the QPC only depends on the \textit{amplitude}
1081: of the external bias voltage. However, the electron-tunneling correlation
1082: under the EA effect $(k=1)$ obeys the relation of the bias anti-symmetry
1083: \begin{equation}
1084: \frac{G_{+,i}^{\left( 1\right) }(V_{d})+G_{-,i}^{\left( 1\right) }(V_{d})}{2}%
1085: =-\frac{G_{+,i}^{\left( 1\right) }(-V_{d})+G_{-,i}^{\left( 1\right) }(-V_{d})%
1086: }{2}.  \label{antisym}
1087: \end{equation}
1088: To understand this relation, let us look at the forward tunneling processes,
1089: in which an extra negative bias voltage is induced by the EA, $%
1090: V_{d}\rightarrow V_{d}-\delta V_{EA}$. According to the result in Sec. IV B,
1091: an effective \textit{relaxation} of the qubit is induced by the EA effect.
1092: On the other hand, for a negative bias voltage, an extra negative bias
1093: voltage induced by the EA leads to $-V_{d}\rightarrow -V_{d}-\delta V_{EA}$.
1094: The amplitude of bias voltage is increased. An effective \textit{excitation}
1095: of the qubit is induced by the EA effect. Accordingly, Eq. (\ref{antisym})
1096: indicates that the EA induced relaxation rate of the qubit under the
1097: positive bias voltage ($V_{d}$) is equal to the EA induced excitation rate
1098: of the qubit under the negative bias voltage ($-V_{d}$).
1099: 
1100: In addition, the relaxation process can be separated into two effective
1101: modes
1102: \begin{equation}
1103: \Gamma _{1,r}=\eta _{d}\bar{\mu}(\frac{R(z_{+})+R(z_{-})}{2}),\quad z_{\pm }=%
1104: \frac{\beta (V_{d}\pm \gamma )}{2},
1105: \end{equation}
1106: where the function $R(z)=\coth (z)-z\csc $h$(z)^{2}$ is a step-like function
1107: with bounded values $\pm \frac{1}{2}$. The function preserves the bias
1108: anti-symmetry. As discussed in Sec. III, the relaxation depends on the
1109: structure of the CQD. It leads to two effective modes corresponding to $%
1110: R(z_{+})$ and $R(z_{-})$. In the low temperature regime, these two effective
1111: modes separate. Eq. (\ref{antisym}) shows that the qubit excitation rate in
1112: the $R(z_{-})$ mode cancels the qubit relaxation rate in the $R(z_{+})$ mode
1113: in the regime of $V_{d}$ limited by $\pm \gamma $ . Then a relaxationless EA
1114: effect occurs, as shown in Fig.~\ref{fig6}. However, in the high temperature
1115: regime, the CQD-structure dependence of the relaxation is suppressed, and
1116: two modes $R(z_{\pm })$ become degenerate. No relaxationless EA effect
1117: occurs. As a result, $\Gamma _{1,r}(t)$ reduces to $\Gamma _{0,r}$ (i.e. $%
1118: \Gamma _{1,r}\rightarrow 0$) in low temperature regime. For the qubit
1119: dephasing, it can be checked that the dephasing term $\Gamma _{1,r}\mathcal{F%
1120: }\left( t\right) n_{0,p}\left( t\right) \rightarrow 0$ as $T\rightarrow 0$.
1121: The qubit is also EA-dephasingless in the low temperature regime.
1122: 
1123: \section{SUMMARY}
1124: 
1125: We have developed a perturbative theory to study the EA effect on the
1126: decoherence of the charge qubit, a single electron confined in CQD, measured
1127: by the QPC. The contribution due to the EA is treated perturbatively. A set
1128: of master equations for the reduced density matrix of the qubit has been
1129: obtained in this perturbation scheme. By solving the resulted master
1130: equations, we obtain the decay modes of the dephasing and the relaxation of
1131: the qubit in the thermal equilibrium limit, and study the temperature- and
1132: bias-dependence of the dephasing and the relaxation rate. We find two kinds
1133: of decay modes for the dephasing in the electrical measurement processes:
1134: one decay rate increases as the increase of temperature and bias voltage, as
1135: expected; the other shows an abnormal dependence. The qubit in the later
1136: decay mode preserves the longest decoherence time in the high temperature
1137: and high bias limit, its time scale is much longer than that of the usual
1138: mode. In addition, the EA properties are studied extensively. The EA current
1139: is obtained according to the master equations. We find that the EA current
1140: varies linearly with time under the pure EA effect, and is then slowed down
1141: into an asymptotic stable state by the bias. The qubit decoherence due to
1142: the EA effect is studied based on this analysis. We find that under the EA
1143: effect the qubit relaxation and dephasing rate are much larger than the ones
1144: in the thermal equilibrium limit, unless the bias voltage is turned off.
1145: Also, the master equations show that an extra qubit relaxation is induced by
1146: the EA effect initially, and then be suppressed by the bias. Asymptotically,
1147: the qubit will be forced into the state with a small relaxation shift which
1148: is independent of the ER decay rate. However, in the low bias limit, the
1149: qubit will not be affected by the EA. The qubit decoherence rate will not be
1150: speeded up by the EA effect. Finally, we find a decoherenceless EA effect in
1151: the low temperature regime. A bias anti-symmetry in EA processes suppresses
1152: the qubit decoherence. These decoherent behaviors worth being explored in
1153: experiments.
1154: 
1155: \appendix*
1156: 
1157: \section{DERIVATION OF MASTER EQUATIONS}
1158: 
1159: The derivation of the master equations (\ref{eq:ml0},\ref{eq:ml1rel},\ref
1160: {eq:mc0},\ref{eq:mc1rel}) is presented in this appendix. The total density
1161: operator in the interaction picture of the reservoir satisfies
1162: \begin{equation}
1163: \frac{d}{dt}\hat{\rho}_{I}\left( t\right) =\frac{1}{i\hbar }\left[ \hat{H}%
1164: _{S}+\hat{H}_{I}^{\prime },\hat{\rho}_{I}\left( t\right) \right] ,\;\hat{H}%
1165: _{I}^{\prime }=e^{\frac{i\hat{H}_{B}t}{\hbar }}\hat{H}^{\prime }e^{\frac{%
1166: \hat{H}_{B}t}{i\hbar }}.  \label{eq:A1}
1167: \end{equation}
1168: In the weak coupling regime, the interaction Hamiltonian $\hat{H}^{\prime }$
1169: can be treated by using the second order cummulant expansion technique \cite
1170: {mori,li,goan,stace}. By taking the conditional trace Tr$_{D^{\left(
1171: n\right) }\left( \bar{D}^{\left( m\right) }\right) }$ shown in Sec. II and
1172: the second order cummulant expansion, the master equation becomes
1173: \begin{eqnarray}
1174: &&\frac{d}{dt}\hat{\rho}_{f\left( b\right) }^{\left( n\right) }\left(
1175: t\right) =-i\hat{L}_{D}\hat{\rho}_{f\left( b\right) }^{\left( n\right)
1176: }\left( t\right) -\hat{R}_{f\left( b\right) }\hat{\rho}_{f\left( b\right)
1177: }^{\left( n\right) }\left( t\right) ,  \label{eq:A2} \\
1178: &&\hat{R}_{f\left( b\right) }\hat{\rho}_{f\left( b\right) }^{\left( n\right)
1179: }\left( t\right) =\frac{1}{\hbar ^{2}}\int_{t_{0}}^{t}d\tau \text{Tr}%
1180: _{D^{(n)}(\bar{D}^{(n)})}  \nonumber \\
1181: &&~~~~~~\times \left\{ \Big[\hat{H}_{I}^{\prime }(t),\left[ \hat{G}\left(
1182: t,\tau \right) \hat{H}_{I}^{\prime }(\tau )\hat{G}\left( t,\tau \right) ^{+},%
1183: \hat{\rho}_{I}\left( t\right) \right] \Big]\right\} ,  \nonumber \\
1184: &&  \label{eq:A3}
1185: \end{eqnarray}
1186: where $\hat{\rho}_{f\left( b\right) }^{\left( n\right) }\left( t\right) =%
1187: \mathrm{Tr}_{D^{(n)}(\bar{D}^{(n)})}[\hat{\rho}_{tot}(t)]$ and $\hat{G}%
1188: \left( t,\tau \right) $ the propagator of the CQD. The decoherence
1189: of the measured qubit is governed by the dissipation term
1190: $-\hat{R}_{f\left( b\right) }\hat{\rho}_{f\left( b\right)
1191: }^{\left( n\right) }\left( t\right) $ in Eq. (\ref{eq:A2}).
1192: 
1193: Note that the master equations (\ref{eq:A2}, \ref{eq:A3})
1194: essentially describe a Markovian process. The Markovian
1195: approximation has been introduced in this stage. Therefore, the
1196: time integration in Eq. (\ref{eq:A3}) is replaced by the one be
1197: integrated out along the whole time domain. Carrying out the commuter in Eq. (\ref{eq:A3}), $\hat{R}_{f\left( b\right) }%
1198: \hat{\rho}_{f\left( b\right) }^{\left( n\right) }\left( t\right) $ can be
1199: expressed as
1200: \begin{eqnarray}
1201: &&\frac{1}{2 \hbar ^{2}}\int_{-\infty}^{\infty}d\tau \Big\{\hat{q}(\hat{P}_{0}-e^{%
1202: \frac{i\gamma \left( \tau -t\right) }{\hbar
1203: }}\hat{P}_{1}-e^{\frac{-i\gamma
1204: \left( \tau -t\right) }{\hbar }}\hat{P}_{2})  \nonumber \\
1205: &&~~~~~~~\times (C_{f(b),-}^{n}(t-\tau )+C_{f(b),+}^{n}(t-\tau ))\Big\}+H.c.
1206: \nonumber \\
1207: &&-\frac{1}{2 \hbar ^{2}}\int_{-\infty}^{\infty}d\tau \Big\{(\hat{P}_{0}-e^{\frac{%
1208: i\gamma \left( \tau -t\right) }{\hbar
1209: }}\hat{P}_{1}-e^{\frac{-i\gamma \left(
1210: \tau -t\right) }{\hbar }}\hat{P}_{2})  \nonumber \\
1211: &&~~~~~~~\times (C_{f,+(b,-)}^{n-1}(t-\tau )+C_{f,-(b,+)}^{n+1}(t-\tau ))%
1212: \hat{q}\Big\}+H.c.,  \nonumber \\
1213: &&  \label{eq:A4}
1214: \end{eqnarray}
1215: where the reservoir time correlation functions are defined by
1216: \begin{eqnarray}
1217: C_{f,+}^{n}(t-\tau ) &=&\mathrm{Tr}_{D^{(n)}}[\hat{f}_{t}^{+}\hat{f}_{\tau }%
1218: \hat{\rho}_{I}(t)],  \nonumber \\
1219: C_{f,-}^{n}(t-\tau ) &=&\mathrm{Tr}_{D^{(n)}}[\hat{f_{t}}\hat{f}_{\tau }^{+}%
1220: \hat{\rho}_{I}(t)],  \label{eq:c1}
1221: \end{eqnarray}
1222: for the forward tunneling processes, and
1223: \begin{eqnarray}
1224: C_{b,+}^{m}(t-\tau ) &=&\mathrm{Tr}_{\bar{D}^{(m)}}[\hat{f}_{t}^{+}\hat{f}%
1225: _{\tau }\hat{\rho}_{I}(t)],  \nonumber \\
1226: C_{b,-}^{m}(t-\tau ) &=&\mathrm{Tr}_{\bar{D}^{(m)}}[\hat{f_{t}}\hat{f}_{\tau
1227: }^{+}\hat{\rho}_{I}(t)],  \label{eq:c2}
1228: \end{eqnarray}
1229: for the backward tunneling processes. It can be easily checked from Eqs.~(%
1230: \ref{eq:c1},\ref{eq:c2}) that without the EA effect, the backward tunneling
1231: processes is covered in the forward tunneling processes, and the result is
1232: the same as that in \cite{gurvitz}.
1233: 
1234: Next, the EA effect is taken into account to calculate the
1235: reservoir time correlation functions. It should be noted that the
1236: excitation rate of an electron occupying in lower energy levels
1237: far from the Fermi energy is much smaller than the one near the
1238: Fermi energy. That is, most likely, only the electrons occupying
1239: near the Fermi energy can tunnel through the QPC. If the QPC has a
1240: low transmission, i.e. $n/A,\;m/A\ll \bar{N}$, the states $\left\{
1241: \left| D\left( \bar{L}^{n},R^{n}\right) \right\rangle \right\} $
1242: contributed by all allowed tunnelings in which electrons occupying
1243: near the Fermi energy can tunnel from the source to the drain are
1244: almost the same. Therefore, we have the following approximation
1245: \begin{equation}
1246: \mathrm{Tr}_{D^{(n)}}[\hat{f}_{t}^{+}\hat{f}_{\tau }\hat{\rho}%
1247: _{I}(t)]\approx \hat{\rho}_{f}^{(n)} \sum_{%
1248: \bar{L}^{n}R^{n}} \langle D^{n}|\hat{f}_{t}^{+}\hat{f}%
1249: _{\tau }|D^{n}\rangle ,  \label{eq:assupt1}
1250: \end{equation}
1251: and $C_{f,+}^{n}(t-\tau )$ can be expressed as $\hat{\rho}_{f}^{(n)}(t)\sum_{%
1252: \bar{L}^{n}R^{n}}%
1253: \mathrm{Tr}[\hat{A}_{n}^{+}\hat{f}_{t}^{+}\hat{f}_{\tau }\hat{A}%
1254: _{n}\hat{\rho}_{B}^{(0)}]$, where $\hat{A}_{n}=\hat{a}_{r_{1}}^{+}\cdots
1255: \hat{a}_{r_{n}}^{+}\hat{a}_{\bar{l}_{1}}\cdots \hat{a}_{\bar{l}_{n}}$, and $%
1256: \hat{\rho}_{B}^{(0)}$ is the reservoir vacuum state. Furthermore, we define $%
1257: \hat{\rho}_{B}^{(n)}\equiv \hat{A_{n}}\hat{\rho}_{B}^{(0)}\hat{A_{n}}^{+}$
1258: as the electronic reservoir density operator with $n$ electrons created in
1259: the drain for the forward tunneling processes. Combine these analysis
1260: together, we obtain
1261: \begin{eqnarray}
1262: C_{f,+}^{n}(t-\tau ) &=&\hat{\rho}_{f}^{(n)}(t)\sum_{l^{\prime }r^{\prime
1263: }}\sum_{lr}{e^{\frac{it}{\hbar }\left( \varepsilon _{l^{\prime
1264: }}-\varepsilon _{r^{\prime }}\right) }e^{-\frac{i\tau }{\hbar }\left(
1265: \varepsilon _{l}-\varepsilon _{r}\right) }}  \nonumber \\
1266: &&~~~~~~~\times \sum_{\bar{L}^{n}R^{n}} \mathrm{Tr}\left[ \hat{a}_{r}^{+}\hat{a}_{l}%
1267: \hat{\rho}_{B}^{(n)}\hat{a}_{l^{\prime }}^{+}\hat{a}_{r^{\prime
1268: }}\right]
1269: \nonumber \\
1270: &=&\hat{\rho}_{f}^{(n)}(t)\sum_{lr}e^{\frac{i(t-\tau )}{\hbar }(\varepsilon
1271: _{l}-\varepsilon _{r})}F_{l}^{-}(n,\beta ,t-\tau )  \nonumber \\
1272: &&~~~~~~~~~~~~\times [1-F_{r}^{+}(n,\beta ,t-\tau )],  \label{eq:cfp}
1273: \end{eqnarray}
1274: where the EA-fluctuated Fermi-Dirac functions $F_{l,r}^{\pm }$
1275: have been treated perturbatively and shown
1276: in Eq. (\ref{eq:pf}). Note that Eq.~(\ref{eq:pf}) is valid when $n/A\ll \bar{%
1277: N}$ and $m/A\ll \bar{N}$, which coincides with the approximation used in
1278: Eq.~(\ref{eq:assupt1}). Other reservoir time correlation functions can be
1279: similarly reduced to the following forms
1280: \begin{eqnarray}
1281: &&C_{f,-}^{n}(t-\tau )=\hat{\rho}_{f}^{(n)}(t)\sum_{lr}e^{-\frac{i(t-\tau )}{%
1282: \hbar }(\varepsilon _{l}-\varepsilon _{r})}F_{r}^{+}(n,\beta ,t-\tau )
1283: \nonumber \\
1284: &&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\times [1-F_{l}^{-}(n,\beta ,t-\tau )],
1285: \nonumber \\
1286: &&C_{b,+}^{m}(t-\tau )=\hat{\rho}_{b}^{(m)}(t)\sum_{lr}e^{\frac{i(t-\tau )}{%
1287: \hbar }(\varepsilon _{l}-\varepsilon _{r})}F_{l}^{+}(n,\beta ,t-\tau )
1288: \nonumber \\
1289: &&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\times [1-F_{r}^{-}(n,\beta ,t-\tau )],
1290: \nonumber \\
1291: &&C_{b,-}^{m}(t-\tau )=\hat{\rho}_{b}^{(m)}(t)\sum_{lr}e^{-\frac{i(t-\tau )}{%
1292: \hbar }(\varepsilon _{l}-\varepsilon _{r})}F_{r}^{-}(n,\beta ,t-\tau )
1293: \nonumber \\
1294: &&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\times [1-F_{l}^{+}(n,\beta ,t-\tau )].
1295: \label{eq:cbm}
1296: \end{eqnarray}
1297: Obviously, it shows that, in the above time correlation functions, the
1298: perturbed terms proportional to $\delta \mu _{R,L}(n,\beta ,t)$ come from
1299: the EA effect. Using the perturbation scheme in Sec. II, we then have $\hat{%
1300: \rho}_{f}^{(n)}(t)=\hat{\rho}_{0,f}^{(n)}(t)+\xi \hat{\rho}%
1301: _{1,f}^{(n)}(t)+\xi ^{2}\hat{\rho}_{2,f}^{(n)}(t)+\cdots $ for the forward
1302: tunneling processes, and $\hat{\rho}_{b}^{(m)}(t)=\hat{\rho}%
1303: _{0,b}^{(m)}(t)+\xi \hat{\rho}_{1,b}^{(m)}(t)+\xi ^{2}\hat{\rho}%
1304: _{2,b}^{(m)}(t)+\cdots $ for the backward tunneling processes.
1305: 
1306: In addition, the time integration in Eq. (\ref{eq:A4}) can be
1307: calculated by using the results
1308: \begin{eqnarray}
1309: &&\frac{1}{2\hbar }{\int_{-\infty }^{\infty }d\tau e^{\frac{\gamma (t-\tau )%
1310: }{i\hbar }}C_{f(b),+}^{n}(t-\tau )}=\pi g_{L}g_{R}\hat{\rho}_{f(b)}^{(n)}(t)
1311: \nonumber \\
1312: &&~~~~~~~~~~~\times \left( \tilde{g}^{(0)}(V_{d}-\gamma )\mp n\mathcal{U}%
1313: \tilde{g}^{(1)}(V_{d}-\gamma )\right) ,  \nonumber \\
1314: &&\frac{1}{2\hbar }{\int_{-\infty }^{\infty }d\tau e^{\frac{i\gamma (t-\tau )%
1315: }{\hbar }}C_{f(b),-}^{n}(t-\tau )}=\pi g_{L}g_{R}\hat{\rho}_{f(b)}^{(n)}(t)
1316: \nonumber \\
1317: &&~~~~~~\times \left( \tilde{g}^{(0)}(-V_{d}+\gamma )\pm n\mathcal{U}\tilde{g%
1318: }^{(1)}(-V_{d}+\gamma )\right) ,  \label{eq:ftm}
1319: \end{eqnarray}
1320: and the relation $\delta \mu _{L}\left( n,\beta ,t\right) +\delta \mu
1321: _{R}\left( n,\beta ,t\right) =n\bar{\mu}\xi \mathcal{F}(t)$, the master
1322: equations become
1323: \begin{widetext}
1324: \begin{eqnarray}
1325: &&\frac{d}{dt}\hat{\rho}_{0,f}^{(n)}(t)+\xi \frac{d}{dt}\hat{\rho}%
1326: _{1,f}^{(n)}(t)=-i\hat{L}_{D}\big(\hat{\rho}_{0,f}^{(n)}(t)+\xi \hat{\rho}%
1327: _{1,f}^{(n)}(t)\big) \,\,\,\,  \nonumber \\
1328: &&~~~~~~-\lambda \Big\{\hat{q}\big(\hat{A}_{+}^{(0)}+\hat{A}_{-}^{(0)}\big)%
1329: \hat{\rho}_{0,f}^{(n)}(t)-\big(\hat{A}_{+}^{(0)}\hat{\rho}_{0,f}^{(n-1)}(t)
1330: +\hat{A}_{-}^{(0)}\hat{\rho}_{0,f}^{(n+1)}(t)\big) \hat{q}  \nonumber \\
1331: &&~~~~~~~~~~~~~~~+\xi \Big[\hat{q}\big(\hat{A}_{+}^{(0)}+\hat{A}_{-}^{(0)}\big)\hat{\rho}%
1332: _{1,f}^{(n)}(t)  -\big(\hat{A}_{+}^{(0)}\hat{\rho}_{1,f}^{(n-1)}(t)+\hat{A}%
1333: _{-}^{(0)}\hat{\rho}_{1,f}^{(n+1)}(t)\big)\hat{q}  \nonumber \\
1334: &&~~~~~~~~~~~~~~~~~~~~~~~~~ +\mathcal{F}(t) n\hat{q}\big(\hat{A}%
1335: _{+}^{(1)}+\hat{A}_{-}^{(1)}\big)\hat{\rho}_{0,f}^{(n)}(t)-\mathcal{F}(t)
1336: \big((n-1) \hat{A}_{+}^{(1)}\hat{\rho}_{0,f}^{(n-1)}(t)
1337: +(n+1)\hat{A}_{-}^{(1)}\hat{\rho}_{0,f}^{(n+1)}(t)\big) \hat{q}
1338:  \Big]\Big\} +H.c. \nonumber \\
1339: &&~~~~~~+\mathcal{O}\left( \xi ^{2}\right) +\cdots \label{eq:m_nf}
1340: \end{eqnarray}
1341: for the forward tunneling processes, and
1342: 
1343: \begin{eqnarray}
1344: &&\frac{d}{dt}\hat{\rho}_{0,b}^{(n)}(t)+\xi \frac{d}{dt}\hat{\rho}%
1345: _{1,b}^{(n)}(t)=-i\hat{L}_{D}\big(\hat{\rho}_{0,b}^{(n)}(t)+\xi \hat{\rho}%
1346: _{1,b}^{(n)}(t)\big) \,\,\,\,  \nonumber \\
1347: &&~~~~~~-\lambda \Big\{\hat{q}\big(\hat{A}_{+}^{(0)}+\hat{A}_{-}^{(0)}\big)%
1348: \hat{\rho}_{0,b}^{(n)}(t)-\big(\hat{A}_{-}^{(0)}\hat{\rho}_{0,b}^{(n-1)}(t)
1349: +\hat{A}_{+}^{(0)}\hat{\rho}_{0,b}^{(n+1)}(t)\big) \hat{q}  \nonumber \\
1350: &&~~~~~~~~~~~~~~~+\xi %
1351: \Big[\hat{q}\big(\hat{A}_{+}^{(0)}+\hat{A}_{-}^{(0)}\big)\hat{\rho}%
1352: _{1,b}^{(n)}(t)-\big(\hat{A}_{-}^{(0)}\hat{\rho}_{1,b}^{(n-1)}(t)+\hat{A}%
1353: _{+}^{(0)}\hat{\rho}_{1,b}^{\left( n+1\right) }(t)\big)\hat{q}  \nonumber \\
1354: &&~~~~~~~~~~~~~~~~~~~~~~~~~ -\mathcal{F}(t) n\hat{q}\big(%
1355: \hat{A}_{+}^{(1)}+\hat{A}_{-}^{(1)}\big)\hat{\rho}_{0,b}^{(n)}(t)+\mathcal{F}(t) \big((n-1)%
1356: \hat{A}_{-}^{(1)}\hat{\rho}_{0,b}^{(n-1)}(t)+(n+1)\hat{A}_{+}^{(1)}
1357: \hat{\rho}_{0,b}^{(n+1)}(t)\big) \hat{q}\Big]\Big\}+H.c. \nonumber \\
1358: &&~~~~~~+\mathcal{O}\left( \xi ^{2}\right) +\cdots \label{eq:m_nb}
1359: \end{eqnarray}
1360: \end{widetext}
1361: for the backward tunneling processes, where $\hat{A}_{\pm }^{\left(
1362: 0,1\right) }=\sum_{i=0,1,2}G_{\pm ,i}^{\left( 0,1\right) }\hat{P}_{i}$. All
1363: the definitions of the elements of Eqs.~(\ref{eq:m_nf},\ref{eq:m_nb}) can be
1364: found in Sec. II.
1365: 
1366: Eqs.~(\ref{eq:m_nf},\ref{eq:m_nb}) describes the transport properties of
1367: electrons in the system we concerned, for example, the transport current and
1368: noise spectrum \cite{gurvitz,korotkov,goan,mozyrsky,stace,li}. According to
1369: Eq.~(\ref{eq:tstat}) and recalling that $\hat{\rho}_{f(b)}^{(n)}(t)$
1370: describes the quantum oscillation of the electron in the CQD with the
1371: conditions of $n$ electrons accumulating in the drain (source), the $k$-th
1372: order perturbation of the qubit oscillation is completely described by $\hat{%
1373: \rho}_{k}(t)={\sum_{n=0}^{\infty }}\hat{\rho}_{k,f}^{(n)}(t)+{%
1374: \sum_{m=0}^{\infty }}\hat{\rho}_{k,b}^{(m)}(t).$ Accordingly, the $k$-th
1375: order perturbation of the QPC current operator is given by $\frac{d\hat{N}%
1376: _{k}}{dt},$ where $\hat{N}_{k}={\sum_{n=0}^{\infty }}n\hat{\rho}%
1377: _{k,f}^{(n)}(t)-{\sum_{m=0}^{\infty }}m\hat{\rho}_{k,b}^{(m)}(t),$ and the $k
1378: $-th order perturbation of noise spectrum is defined by $\hat{W}_{k}(t)={%
1379: \sum_{n=0}^{\infty }}n^{2}\left( \hat{\rho}_{k,f}^{(n)}(t)+\hat{\rho}%
1380: _{k,b}^{(k)}(t)\right) $. Combine with Eqs. (\ref{eq:m_nf},\ref{eq:m_nb}),
1381: one can obtain the master equations shown in Sec. II straightforwardly.
1382: 
1383: \acknowledgements
1384: 
1385: One of the authors (M.T.L) would like to thank Profs. B. L. Hu and C. E. Lee
1386: for useful discussions. The work is supported by the National Science
1387: Council of Republic of China under Contract Nos. NSC-93-2119-M-006-002 and
1388: NSC-93-2120-M-006-005, and National Center for Theoretical Science, Republic
1389: of China.
1390: 
1391: \begin{thebibliography}{99}
1392: \bibitem{decoh1}  D. Giulini, E. Joos, C. Kiefer, J. Kupsch, I.-O.
1393: Stamatescu, and H.-D. Zeh, Decoherence and the Appearance of a Classical
1394: World in Quantum Theory (Springer, New York, 1996); W.H. Zurek, Phys. Rev.
1395: D. \textbf{24}, 1516 (1981); E. Joos and H.D. Zeh, Z. Phys. B \textbf{59},
1396: 223 (1985).
1397: 
1398: \bibitem{decoh2}  A.O. Caldeira and A.J. Leggett, Physica \textbf{121A}, 587
1399: (1983); A.O. Caldeira and A.J. Leggett, Ann. Phys. \textbf{149}, 374 (1983).
1400: 
1401: \bibitem{decoh3}  M. Namiki, S. Pascazio, and H. Nakazato, Decoherence and
1402: Quantum Measurements (World Scientific, Sigapore, 1997); C. Anastopoulos and
1403: B.L. Hu, Phys. Rev. A \textbf{62}, 033821 (2000).
1404: 
1405: \bibitem{qis}  M.A. Nielsen and I.L. Chuang, Quantum Computation and Quantum
1406: Information (Cambridge University Press, Cambridge, U.K., 2000); W.-M.
1407: Zhang, Y.-Z. Wu, and C. Soo, quant-ph/0502002; Y.-Z. Wu and W.-M. Zhang,
1408: Europhys. Lett. \textbf{71}, 524 (2005) .
1409: 
1410: \bibitem{gurvitz}  S.A. Gurvitz, Phys. Rev. B \textbf{56}, 15215 (1997); S.
1411: A. Gurvitz, quant-ph/9808058.
1412: 
1413: \bibitem{korotkov}  A.N. Korotkov, Phys. Rev. B \textbf{60}, 5737 (1999);
1414: A.N. Korotkov and D.V. Averin, Phys. Rev. B \textbf{64}, 165310 (2001).
1415: 
1416: \bibitem{goan}  H.-S. Goan, G. J. Milburm, H. M. Wiseman, and H. B. Sun,
1417: Phys. Rev. B \textbf{63}, 125326 (2001); H.-S. Goan and G. J. Milburm, Phys.
1418: Rev. B \textbf{64}, 235307 (2001).
1419: 
1420: \bibitem{mozyrsky}  D. Mozyrsky and I. Martin, Phys. Rev. Lett. \textbf{89},
1421: 018301 (2002).
1422: 
1423: \bibitem{clerk}  A.A. Clerk, S.M. Girvin, and A.D. Stone, Phys. Rev. B 67,
1424: 165324 (2003); S. Pilgram and M. Buttiker, Phys. Rev. Lett. 89, 200401
1425: (2002).
1426: 
1427: \bibitem{stace}  T.M. Stace and S.D. Barrett, Phys. Rev. Lett. \textbf{92},
1428: 136802 (2004).
1429: 
1430: \bibitem{li}  X.Q. Li, W.-K. Zhang, P. Cui, J. Shao, Z. Ma and Y. J. Yan,
1431: Phys. Rev. B \textbf{69}, 085315 (2004); X.Q. Li, P. Cui, and Y. J. Yan,
1432: Phys. Rev. Lett. \textbf{94}, 066803 (2005).
1433: 
1434: \bibitem{free decoh}  G.M. Palma, K.A. Suominen, and A.K. Ekert, Proc. R.
1435: Soc. London, Ser. A \textbf{452}, 567 (1996); P. Zanardi and M. Rasetti,
1436: Phys. Rev. Lett. \textbf{79}, 3306 (1997); D.A. Lidar, I.L. Chuang, and K.B.
1437: Whaley, Phys. Rev. Lett. \textbf{81}, 2594 (1998).
1438: 
1439: \bibitem{dy decoup}  L. Viola and Knill, and S. Lioyd, Phys. Rev. Lett.
1440: \textbf{8}2, 2417 (1999); \textbf{8}5, 3520 (2000); P. Facchi, D.A. Lidar,
1441: and S. Pascazio, Phys. Rev. A \textbf{69}, 032314 (2004).
1442: 
1443: \bibitem{qpc}  M. Field, C. G. Smith, M. Pepper, D. A. Ritchie, J. E. F.
1444: Frost, G. A. C. Jones, and D. G. Hasko, Phys. Rev. Lett. \textbf{70}, 1311
1445: (1993); E. Buks, R. Schuster, M. Heiblum, D. Mahalu, and V. Umansky, Nature
1446: \textbf{391}, 871 (1998); T. Hayashi, T. Fujisawa, H. D. Cheong, Y. H.
1447: Jeong, and Y. Hirayama, Phys. Rev. Lett. \textbf{91}, 226804 (2003); J.M.
1448: Elzerman, R. Hanson, L.H. Willems van Beveren, B. Witkamp, L.M.K.
1449: Vandersypen, and L.P. Kouwenhoven, Nature \textbf{430}, 431 (2004).
1450: 
1451: \bibitem{set}  R.J. Schoelkopf, P. Wahlgren, A.A. Kozhevnikov, P. Delsing,
1452: D.E. Prober, Science, \textbf{280}, 1238 (1998); T.M. Buehler, D.J. Reilly,
1453: R.P. Starrett, A.D. Greentree, A.R. Hamilton, A.S. Dzurak, and R.G. Clark,
1454: App. Phys. Lett \textbf{86}, 143117 (2005).
1455: 
1456: \bibitem{cooper}  Y. Nakamura, Y.A. Pashkin, and J.S. Tsai, Nature \textbf{%
1457: 398}, 786 (1999).
1458: 
1459: \bibitem{mori}  H. Mori, Prog. Theor. Phys. \textbf{33}, 423 (1965); Y.J.
1460: Yan, Phys. Rev. A \textbf{58}, 2721 (1998).
1461: 
1462: \bibitem{elzerman}  J.M. Elzerman, R. Hanson, J. S. Greidanus, L.H. Willems
1463: van Beveren, S. De Franceschi, L.M.K. Vandersypen, S. Tarucha, and L.P.
1464: Kouwenhoven, Phys. Rev. B \textbf{67}, 161308 (2003).
1465: 
1466: \bibitem{averin}  D.V. Averin and E.V. Sukhorukov, Phys. Rev. Lett. \textbf{%
1467: 95}, 126803 (2005).
1468: 
1469: \bibitem{gardiner}  C.W. Gardiner and P. Zoller, Quantum Noise (Springer,
1470: New York, 2000).
1471: \end{thebibliography}
1472: 
1473: \end{document}
1474: