cond-mat0501242/GME.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %                      GME
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: 
5: \chapter{Generalized Master Equation}\label{sec:GME}
6: 
7: The state of a physical system is determined by the measurement of
8: a certain number of observables. Repeated measurements of a given
9: observable always return the same expectation value when the
10: system is in an eigenstate for that particular observable. The
11: uncertainty principle ensures us that, for quantum systems, there
12: are incompatible observables that can not be measured at the same
13: time with indefinite precision.
14: 
15: Given a generic quantum system $\mathcal{S}$ and a complete set of
16: compatible observables $\mathcal{A}_i$ \cite{destri}, an
17: eigenstate of the system for all observables is defined by the set
18: of the corresponding expectation values, i.e. the quantum numbers
19: $a_i$. Each of the possible sets of expectation values is
20: associated with an eigenvector in the Hilbert space of the system.
21: More precisely the Hilbert space of the system is spanned by the
22: eigenvectors of a complete set of compatible observables.
23: 
24: A pure state of the system is represented by a radius (class of
25: equivalence of normalized
26: vectors with arbitrary phase) of this
27: Hilbert space. We call $|\psi\rangle$
28: a representative vector
29: of a radius. Observables are associated to Hermitian operators
30: on the Hilbert space of the system. The dynamics of the quantum
31: system is governed by the Hamiltonian operator, i.e. the operator
32: associated to the observable energy. Given an initial vector, the
33: Schr\"{o}dinger equation prescribes the evolution of this vector at
34: all times:
35: 
36: \be
37:  i \hbar \frac{d}{dt} |\psi(t)\rangle = \mathcal{H} |\psi(t)\rangle
38:  \label{eq:Schroedinger}
39: \end{equation}
40: %
41: with the initial condition $|\psi(0)\rangle = |\psi_0\rangle$. An
42: equivalent formulation of the dynamics can be given in terms of
43: projector operators $|\psi \rangle \! \langle \psi|$. A projector
44: is independent from the arbitrary phase of the vector $|\psi
45: \rangle$, it is then equivalent to a radius of the Hilbert space
46: and represents a \emph{pure state} of the system. Using the
47: Leibnitz theorem for derivatives and the Schr\"odinger equation we
48: derive the equation of Liouville-von Neumann:
49: 
50: 
51: \be
52:  \frac{d \rho}{dt} = -\frac{i}{\hbar}[\mathcal{H},\rho]
53:  \label{eq:von-Neumann}
54: \end{equation}
55: %
56: where $\rho \equiv |\psi \rangle \! \langle \psi|$, $[A,B] \equiv
57: AB - BA$ is the commutator of the operators $A$ and $B$. The
58: operator $\rho$ is usually called \emph{density operator}. For
59: each basis of the Hilbert space all the operators have a matrix
60: representation. The matrix that corresponds to the density
61: operator is called \emph{density matrix}. Each vector of the basis
62: of the Hilbert space corresponds to a particular eigenstate of the
63: system defined by a set of quantum numbers. The diagonal elements
64: of the density matrix are called \emph{populations}. Each
65: population represents the probability that the system in the pure
66: state $|\psi\rangle\!\langle \psi| $ is in the eigenstate defined
67: by the corresponding set of quantum numbers. The trace of the
68: density matrix is one and supports this probabilistic
69: interpretation. The off-diagonal terms of the density matrix are
70: the \emph{coherencies} of the system.  They reflect the linear
71: structure of the Hilbert space. A linear combination of
72: eigenvectors gives rise to a pure state with non-zero coherencies.
73: 
74: Not all density operators correspond to pure states.
75: A convex linear combination of pure states $|\psi_n \rangle
76: \! \langle \psi_n |, n=1,...,N$ is called \emph{statistical
77: mixture}:
78: %
79: \be
80:  \rho = \sum_{n=1}^{N} P_n |\psi_n \rangle \! \langle \psi_n |
81: \end{equation}
82: %
83: where $ P_n \in [0,1),\sum_n P_n = 1$. This is an incoherent
84: superposition of pure states. Also statistical mixtures obey the
85: Liuoville-von Neumann equation of motion (\ref{eq:von-Neumann}).
86: 
87: The \emph{master equation} is an equation of motion for the
88: populations. It is a coarse grained\footnote{In the sense that it
89: describes the effective dynamics on a time scale long compared to
90: the typical times of the fastest processes in the physical
91: system.} equation that neglects coherencies. It was derived the
92: first time by Pauli under the assumption that coherencies have
93: random phases in time due to fast molecular dynamics. It reads:
94: 
95: \be
96:  \frac{dP_n(t)}{dt} = \sum_m[\Gamma_{nm}P_m(t)-\Gamma_{mn}P_n(t)]
97: \end{equation}
98: %
99: where $P_n$ is the population\footnote{Since a density matrix
100: without coherencies is a statistical mixture of eigenstates we
101: have adopted the notation $P_n \equiv \rho_{nn}$} of the
102: eigenstate $n$ and $\Gamma_{nm}$ is the rate of probability flow
103: from eigenstate $n$ to $m$ \cite{huang}.
104: 
105: 
106: \section{Coherent dynamics of small open systems}
107: %
108: The master equation is usually derived for models in which
109: a``small'' system with few degrees of freedom is in interaction
110: with a ``large'' bath with effectively an infinite number of
111: degrees of freedom. The Liouville von-Neumann equation of motion
112: for the total density matrix is very complicated to solve and
113: actually contains too much information since it also takes into
114: account coherencies of the bath. It is useful to average it over
115: bath variables and obtain an equation of motion for the density
116: matrix of the system (the \emph{reduced density matrix}). With no
117: further simplification this equation is called {\bf Generalized
118: Master Equation} (GME) since it involves not only the populations
119: but also the coherencies of the small subsystem. The derivation of
120: the GME from the equation of Liouville-von Neumann is far from
121: trivial and also non-universal: it involves a series of
122: approximations justified by the physical properties of the model
123: at hand. Despite the apparent similarities, the two equations are
124: deeply different: the equation of Liouville-von Neumann describes
125: the reversible dynamics of a closed system; the GME, instead,
126: describes the irreversible dynamics of an open system that
127: continuously exchange energy with the bath\footnote{How can
128: irreversibility be derived from reversibility? The solution of
129: this dilemma lies in time scales: system$+$bath recurrence time is
130: ``infinite'' on the time scale of the system. The GME holds on the
131: time scales of the system.}.
132: 
133: Shuttle devices are small systems coupled to different baths
134: (leads, thermal bath) but they maintain a high degree of
135: correlation between electrical and mechanical degrees of
136: freedom captured by the coherencies of the reduced density matrix.
137: The GME seems to be a good candidate for the description of
138: their dynamics.
139: 
140: In the next two sections we will derive two GMEs using two
141: different approaches. They are both necessary for the description
142: of the shuttling devices since they correspond to the different
143: coupling of the system to the mechanical and electrical baths.
144: 
145: \section{Quantum optical derivation}\label{sec:QOptical}
146: 
147: The harmonic oscillator weakly coupled to a bosonic bath  is a
148: typical problem analyzed in quantum optics. This model well
149: describes in shuttling devices the interaction of the mechanical
150: degree of freedom of the NEMS with its environment. Following
151: section $5.1$ of the book ``Quantum Noise'' by C. W. Gardiner and
152: P. Zoller \cite{gardiner} we start considering a small system S
153: coupled to a large bath B described by the generic Hamiltonian:
154: 
155: \be
156:  H = H_{\rm S} + H_{\rm B} + H_{\rm I}
157: \end{equation}
158: %
159: where $H_{\rm S}$ and $H_{\rm B}$ respectively describe the
160: dynamics of the decoupled system and bath  and $H_{\rm I}$
161: represents the interaction between the two that we assume weak.
162: The density operator $\rho$ (the state of the system$+$bath)
163: satisfies, in the Schr\"odinger picture, the equation of
164: Liouville-von Neumann:
165: 
166: \be
167:  \dot{\rho} = -\frac{i}{\hbar}[H_{\rm S} + H_{\rm B} + H_{\rm I},\rho]
168:  \label{eq:Liouvillespecial}
169: \end{equation}
170: %
171: The state of the system is described by the reduced density matrix $\s$:
172: 
173: \be
174:  \s = {\rm Tr}_{\rm B}\{\rho\}
175:  \label{eq:defreduced}
176: \end{equation}
177: %
178: where ${\rm Tr}_{\rm B}$ indicates the partial trace over the bath
179: degrees of freedom. Our task is to derive from
180: (\ref{eq:Liouvillespecial}) an equation of motion for the reduced
181: density matrix $\s$.
182: 
183: \subsection{Interaction picture}
184: %
185: We start by going to the interaction picture and we use as
186: non-interacting Hamiltonian $H_{\rm S}+H_{\rm B}$. We indicate all
187: the operators in the interaction picture with a tilde. The total
188: density operator in the interaction picture reads:
189: 
190: \be
191: \tilde{\rho}(t) = \exp\left[\frac{i}{\hbar}(H_{\rm S}+H_{\rm B})t\right] \rho(t)
192:                   \exp\left[-\frac{i}{\hbar}(H_{\rm S}+H_{\rm B})t\right]
193: \label{eq:definteraction}
194: \end{equation}
195: %
196: and obeys the equation of motion:
197: 
198: \be
199:  \dot{\tilde{\rho}}(t) = -\frac{i}{\hbar}[\tilde{H}_{\rm I}(t),\tilde{\rho}(t)]
200:  \label{eq:EOMinteraction}
201: \end{equation}
202: %
203: where
204: 
205: \be
206:  \tilde{H}_{\rm I}(t) = \exp\left[\frac{i}{\hbar}(H_{\rm S}+H_{\rm B})t\right] H_{\rm I}
207:                         \exp\left[-\frac{i}{\hbar}(H_{\rm S}+H_{\rm B})t\right]
208: \end{equation}
209: %
210: From (\ref{eq:defreduced}) and (\ref{eq:definteraction}) it follows that
211: 
212: \be
213:  \s(t) = {\rm Tr}_{\rm B}\left\{\exp\left[-\frac{i}{\hbar}(H_{\rm S}+H_{\rm B})t\right] \tilde{\rho}(t)
214:                           \exp\left[\frac{i}{\hbar}(H_{\rm S}+H_{\rm B})t\right] \right\}
215: \end{equation}
216: %
217: The exponentials of $H_{\rm B}$ can be cancelled using the cyclic
218: property of the trace since $H_{\rm B}$ depends only on bath
219: variables. We get
220: 
221: \be
222:  \s(t) = \exp\left[-\frac{i}{\hbar}H_{\rm S}t\right] \tilde{\s}(t)
223:          \exp\left[\frac{i}{\hbar}H_{\rm S}t\right]
224: \end{equation}
225: %
226: where
227: 
228: \be
229:  \tilde{\s}(t) \equiv {\rm Tr}_B\{\tilde{\rho}(t)\}
230: \end{equation}
231: %
232: In other terms the interaction picture for the reduced density
233: matrix is effectively obtained only from the non-interacting
234: Hamiltonian for the system $H_{\rm S}$.
235: 
236: 
237: \subsection{Initial conditions}
238: %
239: We assume that the system and the bath are initially independent,
240: the initial total density operator $\rho$ is then factorized into
241: the tensor product:
242: 
243: \be
244:  \rho(0) = \s(0) \otimes \rho_{\rm B}
245: \end{equation}
246: %
247: For definiteness we assume the bath to be in thermal equilibrium:
248: \be
249:  \rho_{\rm B} = \frac{e^{-\beta H_{\rm B}}}{{\rm Tr_B}e^{-\b H_{\rm B}}}
250: \end{equation}
251: %
252: where $\beta = 1/k_B T$ is the inverse temperature.
253: 
254: \subsection{Reformulation of the equation of motion}
255: %
256: It is most important for the derivation of the GME to recast the
257: original equation of motion in the interaction picture
258: (\ref{eq:EOMinteraction}) into an integro-differential form. The
259: integral from $0$ to $t$ of (\ref{eq:EOMinteraction}) reads
260: 
261: \be
262:  \tilde{\rho}(t) = \tilde{\rho}(0) - \frac{i}{\hbar} \int_0^t dt'
263:                   [\tilde{H}_{\rm I}(t'),\tilde{\rho}(t')]
264: \end{equation}
265: %
266: and inserted back into (\ref{eq:EOMinteraction}) itself gives
267: %
268: \be
269:  \dot{\tilde{\rho}}(t) =
270:   -\frac{i}{\hbar}[\tilde{H}_{\rm I}(t),\tilde{\rho}(0)]
271:   -\frac{1}{\hbar^2}\int_0^t dt'
272:   \tilde{H}_{\rm I}(t),[\tilde{H}_{\rm I}(t'),\tilde{\rho}(t')]]
273:  \label{eq:integrodiff}
274: \end{equation}
275: 
276: \subsection{Average over the bath variables}
277: %
278: We take the partial trace over the bath variables on both sides of
279: (\ref{eq:integrodiff}). From the definition of the reduced density
280: operator (\ref{eq:defreduced}) we obtain, for the LHS,
281: $\dot{\tilde{\s}}$. We assume that the first term in the RHS
282: vanishes, namely
283: 
284: \be
285:  {\rm Tr_B}[\tilde{H}_{\rm I},\tilde{\rho}(0)] = 0
286: \end{equation}
287: %
288: where $\tilde{\rho}(0) = \rho(0) = \s(0) \otimes \rho_{\rm B}$.
289: This means that the interaction Hamiltonian has a bath component
290: with zero average. It is not difficult to fulfill this condition
291: in general by a redefinition of the system and interaction
292: Hamiltonian that subtracts the average of the bath component from
293: the latter.
294: 
295: \subsection{Weak coupling}
296: %
297: We assume that $H_{\rm I}$ is only a small perturbation of $H_{\rm
298: S}$ and $H_{\rm B}$. This condition allows a factorization at all
299: times of the total density operator into its system and bath
300: components. The density operator of the bath is also taken as
301: constant in time:
302: 
303: \be
304:  \tilde{\rho}(t) \approx \tilde{\s}(t) \otimes \rho_{\rm B}
305: \end{equation}
306: 
307: The factorization assumption can be weakened. We introduce for
308: this purpose the notion of correlation function. Given a physical
309: system in the state described by the stationary density operator
310: $\rho$ and two operators $\tilde{O_{\rm 1}}$ and $\tilde{O_{\rm
311: 2}}$ the correlation function between $\tilde{O_{\rm 1}}$ and
312: $\tilde{O_{\rm 2}}$ in this order and at times $t$ and $t'$ is:
313: 
314: \be
315:  C_{O_1O_2}(t,t') \equiv {\rm Tr}\{\rho
316:  \tilde{O_{\rm 1}}(t)
317:  \tilde{O_{\rm 2}}(t')\}
318: \end{equation}
319: %
320: where the trace is taken over the the Hilbert space of the system
321: and the operators are in Heisenberg picture. Returning to our
322: system-bath model, we assume that the interaction Hamiltonian is a
323: sum of operators in the form $F_{\rm S}A_{\rm B}$. The minimal
324: requirement for the weak coupling approximation is that the
325: correlation functions of the bath are not influenced by the state
326: of the system. In formulas:
327: 
328: \be
329:  {\rm Tr_B}\{[\tilde{A_{\rm B}}(t),
330:  [\tilde{A_{\rm B}}(t'),\tilde{\rho}(t')]]\}
331:  \approx  \tilde{\s}(t') \otimes {\rm Tr_B}\{[\tilde{A_{\rm B}}(t),
332:  [\tilde{A_{\rm B}}(t'),\rho_{\rm B}]]\}
333:  \label{eq:weakcouplingapprox}
334: \end{equation}
335: 
336: \subsection{Markov approximation}
337: %
338: The integro-differential equation for $\tilde{\sigma}$ obtained in
339: the weak coupling approximation is non-local in time. The state of the
340: system at time $t$ depends on the history of the model starting
341: from the initial time $t=0$. This is the meaning of the integral
342: on the RHS of the equation\footnote{Note that causality is
343: preserved since the state of the system at times $t'> t$ does
344: \emph{not} enter the integral.}
345: %
346: \be
347:  \dot{\tilde{\s}} = -\frac{1}{\hbar^2}\int_0^t dt'
348:  {\rm Tr_B}\{[\tilde{H_{\rm I}}(t),
349:  [\tilde{H_{\rm I}}(t'),\tilde{\s}(t')\otimes \rho_{\rm B}]] \}.
350: \end{equation}
351: %
352: obtained from (\ref{eq:weakcouplingapprox}) by tracing over bath
353: variables. Due to the different sizes and the weak coupling the
354: effects on the bath of the interaction with the system are
355: negligible. The bath is a classical macroscopic object in thermal
356: equilibrium. Its stationary state is a thermal state: an
357: incoherent statistical mixture of energy  eigenstates. The
358: coherencies in the bath introduced by the interaction with the
359: system decay on a time scale called the correlation time. This is
360: precisely the decaying time of the correlation functions of the
361: bath. If the correlation time of the bath is much shorter than the
362: typical time scale for the system dynamics\footnote{By system
363: dynamics we mean in this case the time evolution of the reduced
364: density operator in the interaction picture $\tilde{\s}(t)$. In
365: this sense it is the weak coupling to keep the system dynamics
366: slow.}, then we can make in (\ref{eq:weakcouplingapprox}) the
367: replacement: \be \tilde{\sigma}(t') \to \tilde{\sigma}(t)
368: \end{equation}
369: %
370: and obtain in this way a differential equation for $\tilde{\s}$.
371: Finally, if we are interested in the dynamics of the system for
372: times much longer than the bath correlation time, the lower
373: integration limit in (\ref{eq:weakcouplingapprox}) can be moved to
374: $-\infty$ since the initial bath correlation are irrelevant. With
375: this set of approximations the knowledge of the state of the
376: system at some time $t_0$ is enough to determine the state at all
377: times $t>t_0$. This property is called Markov property. In the weak
378: coupling limit and assuming short correlation times in the heat
379: bath we have derived the following GME for the reduced density
380: operator in the interaction picture:
381: 
382: \be
383:  \dot{\tilde{\s}} = -\frac{1}{\hbar^2} \int_{0}^{\infty} d\t
384:  {\rm Tr_B}\{[\tilde{H}_{\rm I}(t),[\tilde{H}_{\rm I}(t-\t),
385:  \tilde{\s}(t)\otimes \rho_{\rm B}]]\}
386:  \label{eq:QOpGME}
387: \end{equation}
388: 
389: To proceed further the precise knowledge of the model Hamiltonian
390: is required. In section \ref{subsec:driving_and_damping} we will
391: specialize this derivation of the GME to the description of the
392: dissipative environment of the shuttling devices.
393: 
394: 
395: \section{Derivation ``\`{a} la Gurvitz''}\label{sec:Gurvitz}
396: %
397: The tunneling coupling of the shuttling devices to their
398: electrical baths (the leads) is \emph{not} weak. It sets, on the
399: contrary, the time scale of the electrical dynamics that in the
400: shuttling regime is comparable with the period of the mechanical
401: oscillations in the system.
402: 
403: In the SDQS the tunneling amplitudes depend exponentially on the
404: displacement of the central dot from the equilibrium position. The
405: oscillations of the QD modify correspondingly the tunneling rates.
406: In the shuttling regime the following non-adiabatic condition is
407: fulfilled:
408: %
409: \be
410:  \bar{\Gamma}_{\rm L,R} = \frac{\w}{2 \pi}
411: \end{equation}
412: %
413: where the average can be interpreted as a classical average over
414: the stable limit cycle trajectory or quantum mechanically as an
415: expectation value in the stationary state. In both cases Coulomb
416: blockade must be taken into account for a correct evaluation of
417: the average rate\footnote{We will discuss the details in section
418: \ref{sec:the_three_regimes}.}.
419: 
420: In the TDQS the coupling to the leads is constant and represents a
421: tunable parameter of the model. Also for this device the cleanest
422: shuttling regime is achieved for a rather high coupling (and
423: associated tunneling rates comparable with the mechanical
424: frequency).
425: 
426: In 1996 S. Gurvitz and Ya. S. Prager proposed a microscopic
427: derivation of the GME\footnote{In the article they use the
428: expression ``rate equations''. Nevertheless the equations derived
429: fully involve coherencies.} for quantum transport with an
430: \emph{arbitrary} coupling to the leads \cite{gur-prb-96}.
431: Following their article we give now in detail the derivation of
432: the rate equations for transport of non-interacting spin-less
433: particles through a static single dot connected to
434: leads\footnote{It must be noticed that the problem of calculation
435: of current through arrays of static quantum dots has been analyzed
436: in more or less equivalent approach also by other authors. We cite
437: as example for similarity of results  Wegewijs and Nazarov
438: \cite{weg-prb-99}.}. Even if in this oversimplified case the
439: result is intuitive and could be guessed just using common sense,
440: the generic features of the derivation will appear. We extend the
441: result to particles with spin in strong Coulomb blockade and
442: eventually we conclude the section showing that also coherent
443: transport can be treated in this formalism and we derive the GME
444: for a static double dot device.
445: 
446: Let us consider a quantum dot connected to two electronic leads.
447: We neglect (at the moment) the spin degree of freedom and Coulomb
448: interaction between the electrons. The energy levels in the
449: macroscopic leads are very dense while they are discrete in the
450: microscopic QD. We assume that only one level in the dot
451: participates in the dynamics of the model\footnote{While the non--
452: interacting approximation for the leads can be understood in the
453: frame of Landau theory of quasi-particles, no interaction combined
454: with single level approximation for the QD is an
455: oversimplification with no physical explanation that \emph{must}
456: (and will) be relaxed.}.
457: 
458: \subsection{Many-body basis expansion}
459: 
460: The Hamiltonian for the model reads
461: 
462: \be
463: \bs
464:  H =&\sum_l \eps_l c_l^{\dagger}c_l^{\phd} + \eps_1 c_1^{\dagger}c_1^{\phd}
465:    + \sum_r \eps_r c_r^{\dagger}c_r^{\phd} \\
466:    &+\sum_l \W_l (c_l^{\dagger}c_1^{\phd} + c_1^{\dagger}c_l^{\phd})
467:    + \sum_r \W_r (c_r^{\dagger}c_1^{\phd} + c_1^{\dagger}c_r^{\phd})
468: \end{split}
469: \label{eq:HGurvitz1}
470: \end{equation}
471: %
472: where $c_l^{\dagger}, c_1^{\dagger}, c_r^{\dagger}$ create  a
473: particle in the left lead (energy level $\eps_l$), in the dot and
474: in the right lead (energy level $\eps_r$) respectively. $\eps_1$
475: is the energy level of the dot and $\W_{l\,(r)}$ the tunneling
476: amplitudes between the dot and the left (right) lead. For
477: temperatures much smaller than the Fermi energy of the leads we
478: can approximate their Fermi distributions by step functions. The
479: chemical potential of the left (right) lead is assumed much higher
480: (lower) than the dot energy level.
481: 
482: We identify the empty state $|0\rangle$ for the model  with the
483: condition of empty dot and leads filled up to their Fermi
484: energies. Then we gradually move electrons from the emitter to the
485: dot and finally to the collector and associate a new vector to
486: each state of this ``decaying chain''. We construct in this way an
487: infinite many-body basis that defines the Hilbert space for the
488: model\footnote{Some of the possible states of the device are
489: excluded from this Hilbert space: states with electrons excited
490: above the Fermi level of the left lead and/or holes below the
491: Fermi energy of the right lead. They are neglected because they
492: would anyway be hardly populated due to the fast relaxation of the
493: leads to their thermal states and the very low probability of
494: electron tunneling for energies so far from the resonant level of
495: the dot.}. The first few elements of the basis read:
496: 
497: \be
498:  |0\rangle \qquad
499:  c_1^{\dagger}   c_{l_1}^{\phd}|0\rangle \qquad
500:  c_{r_1}^{\dagger}c_{l_1}^{\phd}|0\rangle \qquad
501:  c_1^{\dagger}c_{r_1}^{\dagger}c_{l_1}^{\phd}c_{l_2}^{\phd}|0\rangle \qquad
502:  c_{r_2}^{\dagger}c_{r_1}^{\dagger}c_{l_1}^{\phd}c_{l_2}^{\phd}|0\rangle \, \ldots
503: \label{eq:basis}
504: \end{equation}
505: %
506: where we choose by convention to move all the creation  operators
507: to the left and the annihilation operators to the right. We also
508: assume with the two groups that $\eps_{l_i} < \eps_{l_{i+1}}$ and
509: $\eps_{r_i} < \eps_{r_{i+1}}$ to avoid double counting.
510: %
511: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
512: % Figure
513: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
514: \begin{figure}
515:  \begin{center}
516:   \includegraphics[angle=0,width=\textwidth]{Figures/basisbis.eps}
517:   \label{fig:basis}
518:   \caption{\small \textit{Schematic representation of the first
519:   elements of the many-body basis for the single dot model. Electrons
520:   are progressively taken from the emitter and moved to the QD and finally
521:   to the collector. The vectors represented are in the order
522: ${\rm a)} = |0\rangle$,
523: ${\rm b)} = c_1^{\dagger}c_{l_1}|0\rangle$,
524: ${\rm c)} = c_{r_1}^{\dagger}c_{l_1}|0\rangle$,
525: ${\rm d)} = c_1^{\dagger}c_{r_1}^{\dagger}c_{l_1}c_{l_2}|0\rangle$,
526: ${\rm e)} = c_{r_2}^{\dagger}c_{r_1}^{\dagger}c_{l_1}c_{l_2}|0\rangle$.}}
527:  \end{center}
528: \end{figure}
529: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
530: The state of the model is described by the many-particle vector
531: $|\Psi(t)\rangle$ which we can expand over the basis:
532: 
533: \be
534: \bs
535:  |\Psi(t)\rangle =&\Big[ b_0(t) +
536:                    \sum_{l_1} b_{1l_1}(t) c_1^{\dagger}c_{l_1}^{\phd}
537: %
538:                    +\sum_{l_1,r_1}b_{l_1r_1}(t) c_{r_1}^{\dagger}c_{l_1}^{\phd} \\
539: %
540:                   &+ \!\sum_{l_1<l_2,r_1}b_{1l_1l_2r_1}(t)
541:                      c_1^{\dagger}c_{r_1}^{\dagger}c_{l_1}^{\phd}c_{l_2}^{\phd}
542: %
543:                    + \! \sum_{l_1<l_2,r_1<r_2}b_{l_1l_2r_1r_2}(t)
544:                      c_{r_2}^{\dagger}c_{r_1}^{\dagger}c_{l_1}^{\phd}c_{l_2}^{\phd}
545: %
546:                    +\ldots \Big]|0\rangle
547: \end{split}
548: \label{eq:expansion}
549: \end{equation}
550: 
551: \subsection{Recursive equation of motion for the coefficients}
552: %
553: The vector $|\Psi(t)\rangle$ obeys the Schr\"odinger equation
554: $i\hbar |\dot{\Psi}(t) \rangle = H |\Psi(t)\rangle$ and we impose
555: the initial condition $|\Psi(0)\rangle = |0\rangle$. In terms of
556: the coefficients of the expansion (\ref{eq:expansion}) we obtain
557: an infinite set of differential equations with the initial
558: condition $b_0(0)= 1$ and all the other coefficients equal to $0$
559: at time $t=0$.
560: 
561: Due to the quadratic form of the Hamiltonian, the infinite set of
562: differential equations for the coefficients $b$'s presents a
563: recursive structure: each coefficient is linked in its equation of
564: motion only to the previous and the next in the ``decaying
565: chain''(\ref{eq:expansion}). Since we are interested in keeping track
566: of the state of the dot we condense the full system of equations
567: into two equations for the generic coefficients
568: $b_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(t)$ and
569: $b_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(t)$:
570: %
571: \be
572: \bs
573:  i\dot{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n} =&
574:         \left[\sum_{k=1}^n(\eps_{r_k}-\eps_{l_k})\right]
575:         b_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}\\
576:         &\!\!\!\!+\sum_{l_{n+1}}\W_{l_{n+1}}
577:         b_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}
578:         +\sum_{k=1}^n(-1)^{k-n}\W_{r_k}
579:         b_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n \backslash \{r_k\}}\\
580: %
581:  i\dot{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n} =&
582:       \left[\eps_1+\sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
583:        -\eps_{l_{n+1}}\right]
584:        b_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}\\
585:       &\!\!\!\!+\sum_{r_{n+1}}\W_{r_{n+1}}
586:        b_{\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^{n+1}}
587:        +\sum_{k=1}^{n+1}(-1)^{k-n-1}\W_{l_k}
588:        b_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}\\
589: \end{split}
590: \label{eq:diffsystem}
591: \end{equation}
592: %
593: where $\hbar=1$ and the sums over labels (e.g. $\sum_{l_{n+1}}$)
594: are continuous sums over all the possible energy levels of the
595: leads. We also used the shortened notations:
596: 
597: \be
598: \bs
599:  b_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n} &\equiv
600:  b_{l_1l_2l_3\ldots l_nr_1r_2r_3\ldots r_n}(t)\\
601:  b_{1\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n} &\equiv
602:  b_{1l_1l_2l_3\ldots l_{k-1}l_{k+1}\ldots l_{n+1}r_1r_2r_3\ldots r_n}(t)
603:  \end{split}
604: \end{equation}
605: 
606: In order to  proceed in the derivation of the rate equations it is
607: most convenient to make the Laplace transform of the system of
608: differential equations (\ref{eq:diffsystem}). We obtain the system
609: of algebraic  equations:
610: 
611: \be
612: \bs
613:    &\left[E + \sum_{k=1}^n (\eps_{l_k}-\eps_{r_k})\right]
614:     \tilde{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)\\
615:    &-\sum_{l_{n+1}}\W_{l_{n+1}}
616:     \tilde{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)
617:     -\sum_{k=1}^n(-1)^{k-n}\W_{r_k}
618:     \tilde{b}_{1\{l_j\}_{j=1}^{n}\{r_j\}_{j=1}^n\backslash \{r_k\}}(E)= i\d_{n0}\\
619: %
620:    &\left[E-\eps_1+\sum_{k=1}^n(\eps_{l_k}-\eps_{r_k})
621:     +\eps_{l_{n+1}}\right]
622:     \tilde{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)\\
623:    &-\sum_{r_{n+1}}\W_{r_{n+1}}
624:     \tilde{b}_{\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^{n+1}}(E)
625:     -\sum_{k=1}^{n+1}(-1)^{k-n-1}\W_{l_k}
626:     \tilde{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)=0\\
627: \end{split}
628: \label{eq:algebsystem}
629: \end{equation}
630: %
631: where the Laplace transformed coefficients are indicated with a
632: tilde and are functions of the variable $E$ (that we assume to
633: have an imaginary part to ensure convergence of the Laplace
634: integral). At this level the left-right asymmetry reveals itself
635: in the number of ``decay channels''. Each state of the chain
636: (\ref{eq:basis}) is coupled to an infinite number of right states
637: and only a finite number of left states. Since the couplings are
638: equivalent this results in a statistically definite direction of
639: motion for the electrons.
640: 
641: 
642: 
643: \subsection{Injection and ejection rates}
644: %
645: The continuous sums in (\ref{eq:algebsystem}) can  be simplified
646: using the recursive structure of the equation of motion. We
647: isolate in the second equation of (\ref{eq:algebsystem}) the coefficient
648: $\tilde{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)$ and insert
649: the result into the first equation of (\ref{eq:algebsystem}). The
650: continuous sum results into two terms:
651: 
652: \be
653:  \sum_{l_{n+1}}\sum_{r_{n+1}} \frac{\W_{l_{n+1}}\W_{r_{n+1}}}
654:  {E -\eps_1+\sum_{k=1}^n (\eps_{l_k}-\eps_{r_k}) +\eps_{l_{n+1}}}
655:  \tilde{b}_{\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^{n+1}}
656:  \label{eq:vanish}
657: \end{equation}
658: %
659: and
660: %
661: \be
662:  \sum_{l_{n+1}}\sum_{k=1}^{n+1} (-1)^{k-n-1} \frac{\W_{l_{n+1}}\W_{l_k}}
663:  {E -\eps_1+\sum_{i=1}^n (\eps_{l_i}-\eps_{r_i}) +\eps_{l_{n+1}}}
664:  \tilde{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}
665:  \label{eq:rate1}
666: \end{equation}
667: 
668: Since the energy levels in the leads are dense we can substitute
669: 
670: \be
671:  \sum_{l_{n+1}} \to
672:  \int_{-\infty}^{+\infty}d\eps_{l_{n+1}}D_L(\eps_{l_{n+1}})
673: \end{equation}
674: %
675: where $D(\eps_{l_{n+1}})$ is the density of states in the left
676: lead calculated at energy $\eps_{l_{n+1}}$ and we have extended
677: the integration limits to infinity in the wide band and high bias
678: approximation. We can evaluate now the sum over $l_{n+1}$ in
679: (\ref{eq:vanish}) and (\ref{eq:rate1}) using residues method.
680: Since all the poles are in the same half plane we can neglect all
681: terms which are asymptotically $o(|\eps_{l_{n+1}}|^{-1})$ for
682: $|\eps_{l_{n+1}}| \to \infty $. It is clear from the algebraic
683: system (\ref{eq:algebsystem}) that a coefficient $\tilde{b}$ that
684: has among its indices $\eps_{l_{n+1}}$ behaves asymptotically at
685: least like $|\eps_{l_{n+1}}|^{-1}$. For this reason
686: (\ref{eq:vanish}) vanishes and only one term is left from
687: (\ref{eq:rate1}):
688: 
689: 
690: \be
691:  \tilde{b}_{\{l_j\}_{j=1}^{n}\{r_j\}_{j=1}^n}
692:  \int_{-\infty}^{+\infty}d\eps_{l+1} D_L(\eps_{l_{n+1}}) \frac{\W_{l_{n+1}}^2}
693:  {E -\eps_1+\sum_{i=1}^n (\eps_{l_i}-\eps_{r_i}) +\eps_{l_{n+1}}}
694:   \label{eq:rate2}
695: \end{equation}
696: %
697: For the evaluation of the integral is enough that the tunneling
698: amplitude $\W_{l_{n+1}}$ and the density of states
699: $D_L(\eps_{l_{n+1}})$ are analytical and non-zero where  the
700: denominator vanishes. We assume that they are constant to avoid
701: $n$-dependence in the tunneling rates and perform  the integral in
702: (\ref{eq:rate2}). The second equation (\ref{eq:algebsystem}) can
703: be treated analogously and the system reads:
704: 
705: \be
706: \bs
707: &\left[E + \sum_{k=1}^n (\eps_{l_k}-\eps_{r_k})
708:      +i\frac{\G_L}{2}\right]
709:      \tilde{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)\\
710:     &\phantom{aaaaa}-\sum_{k=1}^n(-1)^{k-n}\W_{r_k}
711:      \tilde{b}_{1\{l_j\}_{j=1}^{n}\{r_j\}_{j=1}^n\backslash \{r_k\}}(E)= i\d_{n0}\\
712: %
713: &\left[E -\eps_1+\sum_{k=1}^n (\eps_{l_k}-\eps_{r_k}) +\eps_{l_{n+1}}
714:      +i\frac{\G_R}{2}\right]
715:      \tilde{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)\\
716:     &\phantom{aaaaa}-\sum_{k=1}^{n+1}(-1)^{k-n-1}\W_{l_k}
717:      \tilde{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)=0\\
718: \end{split}
719: \label{eq:ratesystem}
720: \end{equation}
721: %
722: where we have introduced the \emph{injection} and \emph{ejection
723: rates} $\G_L$ and $\G_R$
724: 
725: \be
726: \bs
727:  \G_L &\equiv 2\pi D_L\W_L^2\\
728:  \G_R &\equiv 2\pi D_R\W_R^2
729: \end{split}
730: \end{equation}
731: %
732: whith the energy independent tunneling amplitudes ($\W_L$ and
733: $\W_R$) and density of states ($D_L$ and $D_R$)\footnote{We assume
734: real tunneling amplitudes  as it is also implied by the form of
735: the Hamiltonian (\ref{eq:HGurvitz1}). The most general case of
736: complex amplitude would result anyway in the tunneling rates
737: $\G_{L,R} \equiv 2\pi D_{L,R}|\W_{L,R}|^2$.}.
738: 
739: \subsection{The reduced density matrix}\label{sec:ReducedDensityMatrix}
740: %
741: The reduced density operator is defined as the trace over the bath
742: variables of the total density operator:
743: 
744: \be
745:  \s(t) = {\rm Tr}_{B}\{ |\Psi(t) \rangle \! \langle \Psi(t) | \}
746: \end{equation}
747: 
748: The matrix elements of the reduced density operators are explicitly
749: 
750: \be
751:  \s_{ij}(t) = \sum_{\{B\}} \langle i_S,B|\Psi(t)\rangle\!\langle\Psi(t)|B,j_S\rangle
752: \end{equation}
753: %
754: where $|i_S,B \rangle, i = 0,1$ is the vector that corresponds to
755: the empty or charged dot (the system) and a particular
756: configuration $B$ of the leads (the baths). We assume that the
757: bath state $B$ does not contain coherent superpositions of states
758: with different number of particles. This implies the vanishing of
759: coherencies in the reduced density matrix. It is useful to
760: organize the sum over the bath configurations according to the
761: number of extra electrons (holes) collected into the right (left)
762: lead.
763: 
764: \be
765:  \s_{ii}(t) = \sum_{n=0}^{\infty}
766:  \sum_{\{B_n\}}\langle i_S,B_n|\Psi(t)\rangle\!\langle\Psi(t)|B_n,i_S\rangle
767:  = \sum_{n=0}^{\infty}\s_{ii}^{(n)}(t)
768:  \label{eq:genRDM}
769: \end{equation}
770: %
771: where $B_n$ is a configuration of the baths with $n$ extra
772: electrons in the collector and we have introduced the
773: \emph{n-resolved} density matrix $\s^{(n)}$. Using the expansion
774: of the vector $|\Psi(t)\rangle$ in the many-body basis
775: (\ref{eq:basis}) we can express the $n$-resolved density matrix in
776: terms of the coefficients $b$. For the two non-vanishing elements:
777: 
778: \be
779:  \bs
780:   \s_{00}^{(n)} &= \sum_{\{B_n\}}|\langle 0_S,B_n|\Psi(t)\rangle|^2
781:   = \sum_{\{l_k\}\{r_k\}}|b_{\{l_k\}_{k=1}^n\{r_k\}_{k=1}^n}(t)|^2\\
782:   \s_{11}^{(n)} &= \sum_{\{B_n\}}|\langle 1_S,B_n|\Psi(t)\rangle|^2
783:   = \sum_{\{l_k\}\{r_k\}}|b_{1\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}(t)|^2
784:  \end{split}
785:  \label{eq:nRDM}
786: \end{equation}
787: %
788: where the sums are calculated over all the possible configurations
789: of indistinguishable particles (e.g. $\sum_{\{l_k\}} \equiv
790: \sum_{l_1<l_2<l_3<\ldots<l_n}$). The time-dependent matrix
791: elements of the reduced density matrix are connected to the
792: coefficients $\tilde{b}(E)$ by the inverse of the Laplace
793: transform:
794: 
795: \be
796: \bs
797: \s^{(n)}_{00}(t) &=
798:  \sum_{\{l_k\}\{r_k\}}\int\frac{dE dE'}{4 \pi^2}
799:  \tilde{b}_{\{l_k\}_{k=1}^n\{r_k\}_{k=1}^n}(E)
800:  \tilde{b}^*_{\{l_k\}_{k=1}^n\{r_k\}_{k=1}^n}(E')
801:  e^{i(E-E')t}\\
802: %
803: \s^{(n)}_{11}(t) &=
804:   \sum_{\{l_k\}\{r_k\}}\int\frac{dE dE'}{4 \pi^2}
805:   \tilde{b}_{1\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}(E)
806:   \tilde{b}^*_{1\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}(E')
807:   e^{i(E-E')t}\\
808: \end{split}
809: \label{eq:invLaplace}
810: \end{equation}
811: %
812: Apart from being a natural step in the derivation of the GME in
813: the Gurvitz approach, the $n$-resolved density matrix contains the
814: additional information on the number of electrons collected in the
815: resevoir at time $t$. This information is very useful to the
816: calculation of the current noise in the SDQS where the quantum
817: regression theorem can not be applied due to the form of the
818: current operator that involves both system and bath operators.
819: 
820: 
821: \subsection{Generalized Master Equation}\label{subsec:GME}
822: %
823: The equation of motion for the reduced density matrix is obtained
824: combining (\ref{eq:invLaplace}) and (\ref{eq:ratesystem}). First
825: we derive an equation of motion for the $n$-resolved reduced density
826: matrix $\s^{(n)}$. The case of the empty dot population with $n=0$
827: is special due to the particular choice of the initial condition
828: and we treat it separately. The starting point is the first of the
829: equations (\ref{eq:ratesystem}) specialized for $n=0$, namely:
830: 
831: \be
832:  \left( E + i \frac{\G_L}{2}\right)\tilde{b}_0(E) = i
833: \end{equation}
834: %
835: We taking the inverse Laplace transform and obtain
836: 
837: \be
838:  \dot{b}_0(t) = -\frac{\G_L}{2}b_0(t)
839: \end{equation}
840: 
841: The definition of $\s_{00}^{(0)}$ and the Leibnitz theorem for
842: derivatives lead to the conclusion:
843: 
844: \be
845:  \dot{\s}_{00}^{(0)} = \dot{b}_0b_0^* + b_0\dot{b}_0^*
846:                      = -\G_Lb_0b_0^* = -\G_L\s_{00}^{(0)}
847:  \label{eq:nGME0}
848: \end{equation}
849: 
850: This argument is applied also in the case  with $n \ne 0$ but the
851: structure of the equation is more complex and in general a final
852: continuous sum must be evaluated. We take the first equation in
853: (\ref{eq:ratesystem}) and multiply it by
854: $\tilde{b}^*_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E')$$e^{-i(E-E')t}$.
855: Then we subtract side by side the complex conjugate of the first
856: equation of (\ref{eq:ratesystem}) evaluated in $E'$ and multiplied
857: by $\tilde{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)$$e^{-i(E-E')t}$.
858: Finally we integrate in $dE$ and $dE'$ and sum over the bath
859: configurations with $n$ electrons in the collector. We repeat the
860: procedure also for the second equation in (\ref{eq:ratesystem})
861: and obtain:
862: 
863: \be
864: \bs
865:  &\sum_{\{l_j\}\{r_j\}}\!\int\frac{dE dE'}{4 \pi^2}\Big[(E-E'+i\G_L)
866:     \tilde{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)
867:   \tilde{b}^*_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}\!(E')e^{-i(E-E')t}\\
868:  &-2 \Im \Big(
869:  \sum_{k=1}^n(-1)^{k-n}\W_{r_k}
870:   \tilde{b}_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}(E)
871:  \tilde{b}^*_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E')e^{-i(E-E')t}
872:  \Big) \Big]=0
873: \end{split}
874: \label{eq:nGME11}
875: \end{equation}
876: %
877: for the first equation and similarly
878: %
879: \be
880: \bs
881:  &\sum_{\{l_j\}\{r_j\}}\!\int\frac{dE dE'}{4 \pi^2}\Big[(E-E'+i\G_R)
882:     \tilde{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)
883:   \tilde{b}^*_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}\!(E')e^{-i(E-E')t}\\
884:  &-2  \Im \Big(
885:  \sum_{k=1}^{n+1}(-1)^{k-n-1}\W_{l_k}
886:   \tilde{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)
887:  \tilde{b}^*_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E')e^{-i(E-E')t}
888:  \Big) \Big]=0\\
889: \end{split}
890: \label{eq:nGME12}
891: \end{equation}
892: %
893: for the second. $\Im$ indicates the imaginary part. In the
894: definition of the $n$-resolved reduced density matrix the two
895: coefficients $b$ correspond to the same bath configuration. The
896: finite sums in equations (\ref{eq:nGME11}) and (\ref{eq:nGME12})
897: still have coefficients with different bath configuration. Using
898: properties of the Laplace transform, the definition of $\s^{(n)}$
899: and the relations
900: 
901: \be
902: \bs
903:  \tilde{b}^*_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E') &=
904:  \frac{\sum_{k=1}^n (-1)^{k-n}\W_{r_k}
905:        \tilde{b}^*_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}(E')}
906:        {E'+\sum_{k=1}^n(\eps_{l_k} -\eps_{r_k}) -i\frac{\G_L}{2}}\\
907:  %
908:  %
909:  \tilde{b}^*_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E') &=
910:  \frac{\sum_{k=1}^{n+1} (-1)^{k-n-1}\W_{l_k}
911:        \tilde{b}^*_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E')}
912:        {E'-\eps_1+\sum_{k=1}^n(\eps_{l_k} -\eps_{r_k}) + \eps_{l_{n+1}} -i\frac{\G_R}{2}}
913: \end{split}
914: \end{equation}
915: %
916: obtained from (\ref{eq:ratesystem}), we transform
917: (\ref{eq:nGME11}) and (\ref{eq:nGME12}) into:
918: 
919: \be
920: \bs
921:  i(\dot{\s}_{00}^{(n)} + \G_L \s_{00}^{(n)}) =
922:  %%%%%%%%%
923:  \phantom{\Big[ 2 \Im\Big( \sum_{k,k'=1}^n
924:  \frac{(-1)^{k+k'}\W_{r_k}\W_{r_{k'}}}
925:  {E' + \sum_{k=1}^n(\eps_{l_k}-\eps_{r_k}) -i\frac{\G_L}{2}}}&\\
926:  %%%%%%%%%
927:  \sum_{\{l_j\}\{r_j\}} \int\frac{dE dE'}{4 \pi^2}
928:   \Big[ 2 \Im\Big( \sum_{k,k'=1}^n
929:  \frac{(-1)^{k+k'} \W_{r_k}\W_{r_{k'}}}
930:  {E' + \sum_{k=1}^n(\eps_{l_k}-\eps_{r_k}) -i\frac{\G_L}{2}}&\\
931:  \tilde{b}_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}(E)
932:  \tilde{b}^*_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_{k'}\}}(E')
933:          e^{-i(E-E')t}\Big)\Big]&
934: \end{split}
935: \end{equation}
936: 
937: and
938: 
939: \be
940: \bs
941:  i(\dot{\s}_{11}^{(n)} + \G_R \s_{11}^{(n)}) =
942:  %%%%%%%
943:  \phantom{\Big[ 2 \Im\Big( \sum_{k,k'=1}^{n+1}
944:  \frac{(-1)^{k+k'}\W_{l_k}\W_{l_{k'}}}
945:  {E' -\eps_1 + \sum_{k=1}^n(\eps_{l_k}-\eps_{r_k} + \eps_{l_{n+1}})
946:   -i\frac{\G_L}{2}}}&\\
947:  %%%%%%%
948:  \sum_{\{l_j\}\{r_j\}} \int\frac{dE dE'}{4 \pi^2}
949:   \Big[ 2 \Im\Big( \sum_{k,k'=1}^{n+1}
950:  \frac{(-1)^{k+k'}\W_{l_k}\W_{l_{k'}}}
951:  {E' -\eps_1 + \sum_{k=1}^n(\eps_{l_k}-\eps_{r_k} + \eps_{l_{n+1}})
952:   -i\frac{\G_L}{2}}&\\
953:  \tilde{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)
954:  \tilde{b}^*_{\{l_j\}_{j=1}^{n+1}\backslash\{l_{k'}\}\{r_j\}_{j=1}^n}(E')
955:          e^{-i(E-E')t}\Big)\Big]&
956: \end{split}
957: \end{equation}
958: 
959: 
960: It is crucial at this point that $k = k'$. If $k \ne k'$ we can
961: eliminate the variables $r_{k}$ ($l_{k}$) from $\tilde{b}$ and
962: $r_{k'}$ ($l_{k'}$) from $\tilde{b}^*$, perform the integral over
963: one of the now common ``missing'' variables\footnote{Missing in
964: the sense that they have been eliminated from the coefficients
965: subcripts.} and obtain zero. We are left with the case $k = k'$.
966: We transform the sum over the ``missing'' variable $r_k$ ($l_k$)
967: into an integral in the corresponding energy. The discrete sum in
968: the index $k$ takes care of the integration limits and sets them
969: to infinity. The integral can be performed using residues methods
970: to get:
971: 
972: \be
973: \bs
974:  i[\dot{\s}_{00}^{(n)}(t) + \G_L \s_{00}^{(n)}(t)] &
975:  = i\G_R \sum_{\{l_j\}\{r_j\}}
976:  |b_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^{n-1}}(t)|^2\\
977:  i[\dot{\s}_{11}^{(n)}(t) + \G_R \s_{11}^{(n)}(t)] &
978:  = i\G_L \sum_{\{l_j\}\{r_j\}}
979:  |b_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(t)|^2\\
980: \end{split}
981: \end{equation}
982: 
983: Finally we use the representation of the $n$-resolved reduced
984: density matrix (\ref{eq:nRDM}) and obtain the master equation:
985: 
986: \be
987: \bs
988:  \dot{\s}_{00}^{(n)} &= -\G_L \s_{00}^{(n)} + \G_R \s_{11}^{(n-1)}\\
989:  \dot{\s}_{11}^{(n)} &= -\G_R \s_{11}^{(n)} + \G_L \s_{00}^{(n)}
990: \end{split}
991: \label{eq:nGMEsimple}
992: \end{equation}
993: %
994: where we assume that $\s_{11}^{(-1)}\equiv 0$ to include into the
995: same compact form also the equation (\ref{eq:nGME0}) for $n=0$.
996: From this set of equations it is possible to determine the current
997: in the left and right leads. The current in the right lead is the
998: time derivative of the total number of electrons collected in the
999: right lead at time $t$:
1000: %
1001: \be
1002:  I_R(t) = \dot{N}_R(t) = \sum_{n=0}^{\infty}
1003:  n [\dot{\s}_{00}^{(n)}(t) + \dot{\s}_{11}^{(n)}(t)]
1004: \end{equation}
1005: %
1006: Inserting (\ref{eq:nGMEsimple}) we obtain the intuitive result:
1007: 
1008: \be
1009:  I_R(t) = \sum_{n=0}^{\infty}
1010:  n \G_R [\s_{11}^{(n-1)}(t) - \s_{11}^{(n)}(t)] =
1011:  \G_R \sum_{n=0}^{\infty} \s_{11}^{(n)}(t) = \G_R \s_{11}(t)
1012:  \label{eq:e-current}
1013: \end{equation}
1014: 
1015: For the calculation of the left lead current we have to start with
1016: the analog of (\ref{eq:nGMEsimple}) but this time resolved for the
1017: number of holes $h$ accumulated in the emitter:
1018: 
1019: \be
1020: \bs
1021:  \dot{\s}_{00}^{(h)} &= -\G_L \s_{00}^{(h)} + \G_R \s_{11}^{(h)}\\
1022:  \dot{\s}_{11}^{(h)} &= -\G_R \s_{11}^{(h)} + \G_L \s_{00}^{(h-1)}
1023: \end{split}
1024: \label{eq:hGMEsimple}
1025: \end{equation}
1026: %
1027: The left lead current reads:
1028: 
1029: \be
1030:  I_L(t) = \dot{N}_L(t) = \G_L \s_{00}(t).
1031:  \label{eq:h-current}
1032: \end{equation}
1033: 
1034: 
1035: The average over the bath degrees of freedom is completed by
1036: summing (\ref{eq:nGMEsimple}) or (\ref{eq:hGMEsimple}) over all
1037: the possible number of electrons (holes) collected in the right
1038: (left) lead.
1039: 
1040: \be
1041: \bs
1042:  \dot{\s}_{00}&=-\G_L \s_{00} + \G_R \s_{11}\\
1043:  \dot{\s}_{11}&=-\G_R \s_{11} + \G_L \s_{00}
1044: \end{split}
1045: \label{eq:GMEsimple}
1046: \end{equation}
1047: 
1048: \vspace{1cm}
1049: 
1050: The system (\ref{eq:GMEsimple}) is a set of rate equations for a
1051: two-state model. The empty and charged states are connected by
1052: charging and discharging rates ($\G_L$ and $\G_R$ respectively)
1053: and the variations in the populations of the two states is given
1054: by a balance of incoming and outgoing currents. The stationary
1055: solution of (\ref{eq:GMEsimple}) is achieved for $I_L =
1056: I_R$\footnote{It is easy to verify that the condition of balanced
1057: current is equivalent to the stationary condition $\dot{\s}_{00} =
1058: \dot{\s}_{11} = 0$ }. This condition and the general sum rule
1059: $\s_{00} + \s_{11} = 1$ give the stationary populations:
1060: %
1061: \be
1062: \bs
1063:  \s_{00}^{st} &= \frac{\G_R}{\G_L + \G_R}\\
1064:  \s_{11}^{st} &= \frac{\G_L}{\G_L + \G_R}
1065: \end{split}
1066: \end{equation}
1067: %
1068: and the stationary current:
1069: %
1070: \be
1071:  I^{st} = \frac{\G_L \G_R}{\G_L + \G_R}.
1072: \end{equation}
1073: 
1074: \subsection{Spin and strong Coulomb blockade}
1075: %
1076: The rate equations (\ref{eq:GMEsimple}) are an intuitive result
1077: that can be written simply using common sense. Nevertheless the
1078: effort spent for their microscopic derivation is justified by the
1079: possible generalizations that will lead us to the GME for shuttle
1080: devices. First we want to relax the assumption of spin-less
1081: non-interacting particles. The spin of the electrons can be very
1082: easily taken into account if we assume strong Coulomb repulsion in
1083: the dot. Due to a charging energy much larger than any other
1084: energy in the model we assume that only one electron at a time can
1085: occupy the dot. The Hamiltonian reads:
1086: 
1087: \be
1088: \bs
1089:  H =&\sum_{l,s} \eps_l c_{ls}^{\dagger}c_{ls}^{\phd} + \eps_1 c_{1s}^{\dagger}c_{1s}^{\phd}
1090:    + \sum_{r,s}\eps_r c_{rs}^{\dagger}c_{rs}^{\phd} \\
1091:    &+\sum_{l,s} \W_l (c_{ls}^{\dagger}c_{1s}^{\phd} + c_{1s}^{\dagger}c_{ls}^{\phd})
1092:    + \sum_{r,s} \W_r (c_{rs}^{\dagger}c_{1s}^{\phd} + c_{1s}^{\dagger}c_{rs}^{\phd})\\
1093:    &+ U c_{1s}^{\dagger}c_{1s}^{\phd}c_{1-s}^{\dagger}c_{1-s}^{\phd}
1094: \end{split}
1095: \label{eq:HGurvitz2}
1096: \end{equation}
1097: %
1098: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1099: % Figure
1100: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1101: \begin{figure}
1102:  \begin{center}
1103:   \includegraphics[angle=-90,width=.7\textwidth]{Figures/spindyn2.eps}
1104:   \caption{\small  \textit{Schematic representation of the dynamics of the
1105:   single dot device with spin. The two spin species can tunnel
1106:   to and from the two degenerate spin levels of the quantum dot with rates
1107:   $\G_{L\uar},\G_{L\dar}$ and $\G_{R\uar},\G_{R\dar}$, respectively and without
1108:   influencing each other. Due to Coulomb
1109:   blockade only one electron at a time can occupy the dot.}}
1110:  \end{center}
1111: \end{figure}
1112: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1114: % Figure
1115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1116: \begin{figure}
1117:  \begin{center}
1118:   \includegraphics[angle=-90,width=.7\textwidth]{Figures/spindyn3.eps}
1119:   \caption{\small  \textit{If we neglect the information of the spin specie the
1120:   system dynamics can be reduced to the one of an effective spinless system with asymmetric
1121:   injection and ejection rates.}}
1122:  \end{center}
1123: \end{figure}
1124: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1125: which is an extension of the Hamiltonian (\ref{eq:HGurvitz1})
1126: where $s=\pm 1/2$ is the spin degree of freedom and $U$ the
1127: charging energy of the double occupied dot. We take into account
1128: the interaction in the definition of the Hilbert space by
1129: discarding all the many-body states with double occupied dot. The
1130: effective Hamiltonian that we consider is then quadratic, and the
1131: Schr\"odinger equation projected onto the many-body basis gives rise
1132: to a recursive set of equations similar to (\ref{eq:diffsystem}).
1133: We have in this case three general equations corresponding to the
1134: three different states of the quantum dot:
1135: 
1136: \be
1137: \bs
1138:  i\dot{b}_{\{\uar l_j\}_{j=1}^{n_{\uar}}
1139:            \{\dar l_j\}_{j=1}^{n_{\dar}}}&_{
1140:            \{\uar r_j\}_{j=1}^{n_{\uar}}
1141:            \{\dar r_j\}_{j=1}^{n_{\dar}}} =\\
1142:         &\left[\sum_{k=1}^{n_{\uar}+n_{\dar}}(\eps_{r_k}-\eps_{l_k})\right]
1143:         b_{\{\uar l_j\}_{j=1}^{n_{\uar}}
1144:            \{\dar l_j\}_{j=1}^{n_{\dar}}
1145:            \{\uar r_j\}_{j=1}^{n_{\uar}}
1146:            \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1147: %
1148:         &+\sum_{l_{n_{\uar}+1}}\W_{l_{n_{\uar}+1}}
1149:         b_{\uar\{\uar l_j\}_{j=1}^{n_{\uar}+1}
1150:                \{\dar l_j\}_{j=1}^{n_{\dar}}
1151:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1152:                \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1153:         &+\sum_{l_{n_{\dar}+1}}\W_{l_{n_{\dar}+1}}
1154:         b_{\dar\{\uar l_j\}_{j=1}^{n_{\uar}}
1155:                \{\dar l_j\}_{j=1}^{n_{\dar}+1}
1156:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1157:                \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1158: %
1159:         &+\sum_{k=1}^{n_{\uar}}(-1)^{k-n_{\uar}}\W_{r_k}
1160:         b_{\uar\{\uar l_j\}_{j=1}^{n_{\uar}}
1161:                \{\dar l_j\}_{j=1}^{n_{\dar}}
1162:                \{\uar r_j\}_{j=1}^{n_{\uar}}\backslash \{\uar r_k\}
1163:                \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1164: %
1165:         &+\sum_{k=1}^{n_{\dar}}(-1)^{k-n_{\dar}}\W_{r_k}
1166:         b_{\dar\{\uar l_j\}_{j=1}^{n_{\uar}}
1167:                \{\dar l_j\}_{j=1}^{n_{\dar}}
1168:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1169:                \{\dar r_j\}_{j=1}^{n_{\dar}}\backslash \{\dar r_k\}}\\
1170: \end{split}
1171: \label{eq:diffsystem21}
1172: \end{equation}
1173: %
1174: for the empty dot coefficient,
1175: %
1176: 
1177: \be
1178: \bs
1179:  i\dot{b}_{\uar\{\uar l_j\}_{j=1}^{n_{\uar}+1}
1180:                \{\dar l_j\}_{j=1}^{n_{\dar}}}&_{
1181:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1182:                \{\dar r_j\}_{j=1}^{n_{\dar}}} =\\
1183:         &\left[\eps_1 + \sum_{k=1}^{n_{\uar}+n_{\dar}}
1184:         (\eps_{r_k}-\eps_{l_k})-\eps_{l_{n_{\uar}+1}}\right]
1185:             b_{\{\uar l_j\}_{j=1}^{n_{\uar}+1}
1186:                \{\dar l_j\}_{j=1}^{n_{\dar}}
1187:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1188:                \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1189: %%
1190:         &+\sum_{r_{n_{\uar}+1}}\W_{r_{n_{\uar}+1}}
1191:             b_{\{\uar l_j\}_{j=1}^{n_{\uar}+1}
1192:                \{\dar l_j\}_{j=1}^{n_{\dar}}
1193:                \{\uar r_j\}_{j=1}^{n_{\uar}+1}
1194:                \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1195: %%
1196:         &+\sum_{k=1}^{n_{\uar}+1}(-1)^{k-n_{\uar}-1}\W_{l_k}
1197:             b_{\{\uar l_j\}_{j=1}^{n_{\uar}+1}\backslash \{\uar l_k\}
1198:                \{\dar l_j\}_{j=1}^{n_{\dar}}
1199:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1200:                \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1201: %%
1202: \end{split}
1203: \label{eq:diffsystem22}
1204: \end{equation}
1205: %%
1206: for the spin-up and finally
1207: 
1208: \be
1209: \bs
1210: %%
1211:  i\dot{b}_{\dar\{\uar l_j\}_{j=1}^{n_{\uar}}
1212:                \{\dar l_j\}_{j=1}^{n_{\dar}+1}}&_{
1213:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1214:                \{\dar r_j\}_{j=1}^{n_{\dar}}} =\\
1215:        &\left[\eps_1 + \sum_{k=1}^{n_{\uar}+n_{\dar}}
1216:        (\eps_{r_k}-\eps_{l_k})-\eps_{l_{n_{\dar}+1}}\right]
1217:             b_{\{\uar l_j\}_{j=1}^{n_{\uar}}
1218:                \{\dar l_j\}_{j=1}^{n_{\dar}+1}
1219:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1220:                \{\dar r_j\}_{j=1}^{n_{\dar}}}\\
1221: %%
1222:         &+\sum_{r_{n_{\uar}+1}}\W_{r_{n_{\dar}+1}}
1223:             b_{\{\uar l_j\}_{j=1}^{n_{\uar}+1}
1224:                \{\dar l_j\}_{j=1}^{n_{\dar}}
1225:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1226:                \{\dar r_j\}_{j=1}^{n_{\dar}+1}}\\
1227: %%
1228:         &+\sum_{k=1}^{n_{\dar}+1}(-1)^{k-n_{\dar}-1}\W_{l_k}
1229:             b_{\{\uar l_j\}_{j=1}^{n_{\uar}}
1230:                \{\dar l_j\}_{j=1}^{n_{\dar}+1}\backslash \{\dar l_k\}
1231:                \{\uar r_j\}_{j=1}^{n_{\uar}}
1232:                \{\dar r_j\}_{j=1}^{n_{\dar}}}
1233: \end{split}
1234: \label{eq:diffsystem23}
1235: \end{equation}
1236: %
1237: for the spin-down coefficient. In the last three differential
1238: equations (\ref{eq:diffsystem21}), (\ref{eq:diffsystem22}) and
1239: (\ref{eq:diffsystem23}) we have extended the notation used in
1240: equation (\ref{eq:diffsystem}) to take into account the spin
1241: degree of freedom. Despite the heavy but complete notation that
1242: keeps track of the four baths (two leads with two spin species per
1243: lead) and the state of the dot, the same kind of arguments that we
1244: used for the spin-less case bring us to the set of rate equations:
1245: %
1246: \be
1247: \bs
1248: \dot{\s}_{00}^{(n)} &= -(\G_{L\uar}+\G_{L\dar})\s_{00}^{(n)} +
1249:                          \G_{R\uar}\s_{\uar\uar}^{(n-1)} +
1250:                          \G_{R\dar}\s_{\dar\dar}^{(n-1)}\\
1251: \dot{\s}_{\uar\uar}^{(n)} &= -\G_{R\uar}\s_{\uar\uar}^{(n)} +
1252:                          \G_{L\uar}\s_{00}^{(n)}\\
1253: \dot{\s}_{\dar\dar}^{(n)} &= -\G_{R\dar}\s_{\dar\dar}^{(n)} +
1254:                          \G_{L\dar}\s_{00}^{(n)}
1255: \end{split}
1256: \label{eq:nGMEspin}
1257: \end{equation}
1258: %
1259: where $\s_{\uar\uar}^{(n)}$ ($\s_{\dar\dar}^{(n)}$) is the
1260: population of spin up (down) in the dot with $n = n_{\uar} +
1261: n_{\dar}$ electrons in the collector and we have introduced the
1262: spin-dependent injection and ejection rates:
1263: 
1264: \be
1265: \bs
1266:  \G_{L,R\uar} &= 2 \pi D_{L,R\uar}\W_{L,R}^2\\
1267:  \G_{L,R\dar} &= 2 \pi D_{L,R\dar}\W_{L,R}^2
1268: \end{split}
1269: \end{equation}
1270: %
1271: The sum over the number of electrons in the collector gives:
1272: 
1273: \be
1274: \bs
1275: \dot{\s}_{00} &= -(\G_{L\uar}+\G_{L\dar})\s_{00} +
1276:                          \G_{R\uar}\s_{\uar\uar} +
1277:                          \G_{R\dar}\s_{\dar\dar}\\
1278: %
1279: \dot{\s}_{\uar\uar} &= -\G_{R\uar}\s_{\uar\uar}+
1280:                         \G_{L\uar}\s_{00}\\
1281: %
1282: \dot{\s}_{\dar\dar} &= -\G_{R\dar}\s_{\dar\dar} +
1283:                          \G_{L\dar}\s_{00}\\
1284: \end{split}
1285: \label{eq:GMEspin}
1286: \end{equation}
1287: 
1288: The coherencies between different spin species in the QD (e.g.
1289: $\s_{\uar\dar}$) vanish in this model because different spin
1290: states of the dot correspond to different bath states and the only
1291: way to have coherent superpositions in the dot would be to
1292: maintain the same in the leads which instead are assumed (as
1293: macroscopic objects) incoherent\footnote{Non-trivial spin
1294: coherencies can be achieved for example by introducing a spin
1295: dynamics in the dot. In that case different spin states on the dot
1296: could correspond to the same bath state.}. If we are not
1297: interested in the spin information on the dot we can introduce the
1298: population for the charged dot $\s_{11}^{(n)} \equiv
1299: \s_{\uar\uar}^{(n)} + \s_{\dar\dar}^{(n)}$. Assuming also
1300: non-polarized leads (i.e. $D_{L,R\uar}=D_{L,R\dar}$) and,
1301: consequently, tunneling rates independent of different spin
1302: species the system of rate equations (\ref{eq:GMEspin}) becomes:
1303: %
1304: \be
1305: \bs
1306:  \dot{\s}_{00} &= -2\G_L\s_{00} + \G_R\s_{11}\\
1307:  \dot{\s}_{11} &=  -\G_R\s_{11} +2\G_L\s_{00}
1308: \end{split}
1309: \label{eq:GMEspin2}
1310: \end{equation}
1311: %
1312: where $\G_L = \G_{L\uar} = \G_{L\dar}$ and $\G_R = \G_{R\uar} =
1313: \G_{R\dar}$. Comparing these rate equations with the ones
1314: derived for the spin-less non-interacting model
1315: (\ref{eq:GMEsimple}) we note that the only remaining signature of
1316: the spin degree of freedom is in the injection rate. In the case
1317: of identical leads the injection rate doubles the ejection rate.
1318: This behaviour can be interpreted in terms of tunneling channels:
1319: both spin species can tunnel in when the dot is empty, but once
1320: the dot is charged with an electron of specific spin only that
1321: species can tunnel out. At this level the spin degree of freedom
1322: is just renormalizing the injection rate. Since this argument can
1323: be repeated for any model in strong Coulomb blockade we will
1324: restrict the derivation of the GME for shuttling devices to
1325: spin-less non-interacting particles.
1326: 
1327: \subsection{Coherencies and double-dot model}
1328: %
1329: A simple example of a device that exhibits coherent transport is
1330: represented by an array of two quantum dots located between a
1331: source and a drain lead. We assume that the device is working in
1332: strong Coulomb blockade (i.e. only one electron at a time can occupy
1333: the device, either in the left or in the right dot). Electrons can
1334: tunnel in the device from the emitter only to the left dot while
1335: tunneling off is allowed only from the right dot. This condition
1336: can be achieved due to the fact that the tunneling coupling to the
1337: leads decreases exponentially with the distance and can be
1338: neglected for the far lead. Also the two dots are in tunneling
1339: contact. Since the transport must happen via tunneling between the
1340: discrete levels of the dots we expect coherencies to play a role.
1341: The Hamiltonian for the model reads:
1342: 
1343: \be
1344: \bs
1345:  H =& \eps_L |L\rangle\!\langle L|
1346:      + \W_0( |L\rangle\!\langle R| + |R\rangle\!\langle L| )
1347:      + \eps_R |R\rangle\!\langle R|\\
1348:     &+ \sum_l \eps_l c_l^{\dagger}c_l^{\phd}
1349:      + \sum_r \eps_r c_r^{\dagger}c_r^{\phd} \\
1350:     &+ \sum_l \W_l
1351:     (|L\rangle\!\langle 0|c_l^{\phd} + |0\rangle\!\langle L|c_l^{\dag})
1352:      + \sum_r \W_r
1353:     (|R\rangle\!\langle 0|c_r^{\phd} + |0\rangle\!\langle R|c_r^{\dag})
1354: \end{split}
1355: \label{eq:HGurvitzdouble}
1356: \end{equation}
1357: %
1358: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1359: % Figure
1360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1361: \begin{figure}
1362:  \begin{center}
1363:   \includegraphics[angle=0,width=.8\textwidth]{Figures/basis2.eps}
1364:   \caption{\small  \textit{Schematic representation of the first many-body basis elements
1365:   for the double dot system.}}
1366:  \end{center}
1367: \end{figure}
1368: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1369: We identify in the first line the system Hamiltonian with the
1370: single energy levels of the left and right dot ($\eps_L$ and
1371: $\eps_R$) and the tunneling amplitude $\W_0$. The second and third
1372: lines describe respectively the electronic baths (the leads) and
1373: their coupling to the device. The system can be found, as in the
1374: spin model, in three different states: empty or occupied with an
1375: electron either in the left or right dot. We associate to each of
1376: these states the vectors $|0\rangle$, $|L\rangle$ and $|R\rangle$.
1377: The Schr\"odinger equation projected onto the many-body basis for
1378: the system+baths Hilbert space can be then represented by the
1379: following three recursive differential equations for the expansion
1380: coefficients:
1381: 
1382: \be
1383: \bs
1384:  i\dot{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n} =&
1385:         \left[\sum_{k=1}^n(\eps_{r_k}-\eps_{l_k})\right]
1386:         b_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}\\
1387:         &\!\!\!+\sum_{l_{n+1}}\W_{l_{n+1}}
1388:         b_{L\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}
1389:         \!\!+\sum_{k=1}^n(-1)^{k-n}\W_{r_k}
1390:         b_{R\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}\\
1391: %
1392:  i\dot{b}_{L\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n} =&
1393:       \left[\eps_L+\sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1394:        -\eps_{l_{n+1}}\right]
1395:        b_{L\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}\\
1396:       &\!\!\!+\W_0
1397:        b_{R\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}
1398:        +\sum_{k=1}^{n+1}(-1)^{k-n-1}\W_{l_k}
1399:        b_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}\\
1400: %
1401: i\dot{b}_{R\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n} =&
1402:       \left[\eps_R+\sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1403:        -\eps_{l_{n+1}}\right]
1404:        b_{R\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}\\
1405:       &\!\!\!+\W_0
1406:        b_{L\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}
1407:        +\sum_{r_{n+1}}\W_{r_{n+1}}
1408:        b_{\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^{n+1}}
1409: \end{split}
1410: \label{eq:diffsystemdouble}
1411: \end{equation}
1412: 
1413: %
1414: \noindent The Laplace  transform can be taken and the continuous
1415: sums in the first and third equation be performed to give the set
1416: of algebraic equations:
1417: %
1418: \be
1419: \bs
1420: &\left[E +\sum_{k=1}^n(\eps_{r_k}-\eps_{l_k}) + i\frac{\G_L}{2}\right]
1421:         \tilde{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)\\
1422:         &\phantom{aaaaa}-\sum_{k=1}^n(-1)^{k-n}\W_{r_k}
1423:         \tilde{b}_{R\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}(E)= i\d_{n0}\\
1424: %
1425: &\left[E + \eps_L+\sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1426:        -\eps_{l_{n+1}}\right]
1427:        \tilde{b}_{L\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)\\
1428:        &\phantom{aaaaa}-\W_0
1429:        \tilde{b}_{R\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)
1430:        -\sum_{k=1}^{n+1}(-1)^{k-n-1}\W_{l_k}
1431:        \tilde{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)=0\\
1432: %
1433: &\left[E + \eps_R+\sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1434:        -\eps_{l_{n+1}} + i\frac{\G_R}{2}\right]
1435:        \tilde{b}_{R\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)\\
1436:        &\phantom{aaaaa}-\W_0
1437:        \tilde{b}_{L\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)=0
1438: \end{split}
1439: \label{eq:ratesystemdouble}
1440: \end{equation}
1441: 
1442: In the double dot model some coherencies of the reduced density
1443: matrix do not vanish since they correspond to different
1444: ``internal'' states of the dot and can be combined with the same
1445: state of the baths. For example:
1446: %
1447: \be
1448: \bs
1449:  \s_{LR}^{(n)}(t) &=
1450:  \sum_{\{B_n\}} \langle L, B_n | \Psi(t) \rangle \!
1451:  \langle \Psi(t) | R, B_n \rangle\\
1452:   &= \sum_{\{l_k\}\{r_k\}}
1453:  b_{L\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}(t)
1454:  b_{R\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}^*(t)
1455: \end{split}
1456: \end{equation}
1457: 
1458: The next step is an $n$-resolved GME for the reduced density
1459: matrix $\s$. The equations for the populations are derived
1460: following the procedure we explained for the single dot model:
1461: %
1462: \be
1463: \bs
1464:  \dot{\s}_{00}^{(n)} &= -\G_L \s_{00}^{(n)} + \G_R \s_{RR}^{(n-1)}\\
1465:  \dot{\s}_{LL}^{(n)} &= -i\W_0\left(\s_{RL}^{(n)} -\s_{LR}^{(n)}\right)
1466:                + \G_L \s_{00}^{n}\\
1467:  \dot{\s}_{RR}^{(n)} &= -\G_R \s_{RR}^{(n)}
1468:                         -i\W_0\left(\s_{LR}^{(n)} -\s_{RL}^{(n)}\right)
1469: \end{split}
1470: \label{eq:populationdouble}
1471: \end{equation}
1472: %
1473: 
1474: %
1475: The coherencies  play an active role in the transport through the
1476: quantum dots: the second and third equations in
1477: (\ref{eq:populationdouble}) show that the left (right) dot can be
1478: discharged (charged) only via  coherent transport. We concentrate
1479: now on the equation for the coherence $\s_{LR}^{(n)}$. We take the
1480: second equation in (\ref{eq:ratesystemdouble}) and multiply it by
1481: $\tilde{b}_{R\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}^*(E')
1482: e^{-i(E-E')t}$, then we subtract side by side the complex
1483: conjugate of the third equation in (\ref{eq:ratesystemdouble})
1484: evaluated in $E'$ and multiplied by
1485: $\tilde{b}_{L\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)
1486: e^{-i(E-E')t}$. Finally we integrate over $dE$ and $dE'$ and sum
1487: over the baths configurations with $n$ electrons in the collector.
1488: Using the properties of the Laplace transform and the
1489: representation of the reduced density matrix in terms of the
1490: coefficients $b$ of the many-body expansion we obtain:
1491: %
1492: \be
1493: \bs
1494:  &i\dot{\s}_{LR}^{(n)}(t)
1495:   + \Big(\eps_L -\eps_R +i\frac{\G_R}{2}\Big)\s_{LR}^{(n)}(t)
1496:   -\W_0\Big[\s_{RR}^{(n)}(t) -\s_{LL}^{(n)}(t)\Big]\\
1497:   &-\sum_{\{l_j\}\{r_j\}}\int\frac{dE dE'}{4 \pi^2}\Big[\sum_{k=1}^{n+1}
1498:   (-1)^{k-n-1}\W_{l_k}
1499:   \tilde{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)\\
1500:   &\tilde{b}_{R\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^{n}}^*(E')
1501:   e^{-i(E-E')t}\Big] = 0
1502: \end{split}
1503: \label{eq:coherence}
1504: \end{equation}
1505: %
1506: The last term in the LHS of equation (\ref{eq:coherence}) vanishes
1507: since the integrand behaves asymptotically as
1508: $o(|\eps_{l_j,r_j}|^{-1})$ in the limit $|\eps_{l_j,r_j}| \to
1509: \infty$  for all the integrated variables $\eps_{l_j,r_j}$. The
1510: average over the bath degrees of freedom is completed by the sum
1511: over the number of electrons in the collector $n$ that leads to
1512: the GME:
1513: %
1514: \be
1515: \bs
1516:  \dot{\s}_{00} &= -\G_L \s_{00}+\G_R \s_{RR}\\
1517:  \dot{\s}_{LL} &= -i\W_0(\s_{RL}-\s_{LR}) + \G_L \s_{00}\\
1518:  \dot{\s}_{RR} &= -\G_R \s_{RR}-i\W_0(\s_{LR}-\s_{RL})\\
1519:  \dot{\s}_{LR} &= -i(\eps_L -\eps_R)\s_{LR}
1520:                   -\textstyle{\frac{1}{2}}\G_R\s_{LR}
1521:                   -i\W_0(\s_{RR}-\s_{LL})
1522: \end{split}
1523: \label{eq:GMEdouble}
1524: \end{equation}
1525: %
1526: where we have omitted the equation for ${\s}_{RL}$ since $\s$ is a
1527: hermitian operator on the Hilbert space of the system and $\s_{RL}
1528: = \s_{LR}^*$. We can better visualize the coherent and incoherent
1529: contribution to the GME using a matrix representation:
1530: 
1531: \be
1532:  \dot{\s} = -i[H_{sys},\s] + \Xi[\s]
1533: \end{equation}
1534: %
1535: where $\s$ is the density matrix
1536: 
1537: \be
1538: \s =\left(
1539: \begin{array}{ccc}
1540: \s_{00} &\s_{0L} &\s_{0R} \\
1541: \s_{L0} &\s_{LL} &\s_{LR} \\
1542: \s_{R0} &\s_{RL} &\s_{RR} \\
1543: \end{array}
1544: \right),
1545: \end{equation}
1546: %
1547: $H_{sys}$ is the Hamiltonian for the system extended to the empty
1548: state for the system
1549: 
1550: \be
1551: H_{sys} = \left(
1552: \begin{array}{ccc}
1553: 0    & 0     & 0 \\
1554: 0    &\eps_L &\W_0 \\
1555: 0    &\W_0   &\eps_R \\
1556: \end{array}
1557: \right)
1558: \end{equation}
1559: %
1560: and $\Xi[\bullet]$ is a linear super-operator that transform
1561: operators on the Hilbert space of the system into operators on the
1562: same space and acts on the density matrix:
1563: 
1564: \be
1565: \Xi[\s] =\left(
1566: \begin{array}{ccc}
1567: -\G_L \s_{00} +\G_R \s_{RR}   & 0  & 0 \\
1568: 0  & \G_L \s_{00}  & -\frac{1}{2}\G_R\s_{LR} \\
1569: 0    &-\frac{1}{2}\G_R\s_{RL}   & -\G_R\s_{RR} \\
1570: \end{array}
1571: \right)
1572: \end{equation}
1573: 
1574: 
1575: \section{GME for shuttle devices}
1576: \label{subsec:driving_and_damping}
1577: 
1578: 
1579: The shuttle devices are in contact with two different kinds of
1580: bath: two electrical baths (the leads) and a mechanical bath. We
1581: assume that the electrical and mechanical baths act independently
1582: on the device. This assumption splits the GME into two additive
1583: components, one for each kind of bath. Due to the different
1584: coupling strengths we derive the electrical component extending
1585: the method proposed by Gurvitz and extensively presented in
1586: Section \ref{sec:Gurvitz} while for the mechanical component we
1587: adopt the weak coupling quantum optical derivation presented in
1588: Section \ref{sec:QOptical}.
1589: 
1590: \subsection{Single Dot Quantum Shuttle}\label{sec:GMESDQS}
1591: 
1592: We start recalling the Hamiltonian for the SDQS:
1593: 
1594: \be
1595:     H =H_{\rm sys}+H_{\rm leads}+H_{\rm bath}
1596:       +H_{\rm tun}+H_{\rm int}
1597: \end{equation}
1598: where
1599: \be
1600:  \bs
1601:   &H_{\rm sys} =\frac{\hat{p}^2}{2 m}
1602:                +\frac{1}{2}m \w^2 \hat{x}^2
1603:                +(\eps_1- e\mathcal{E} \hat{x})c_1^{\dag}c_1^{\phd}\\
1604:  &H_{\rm leads} = \sum_{k}(\eps_{l_k}
1605:                c^{\dagger}_{l_k}c^{\phd}_{l_k}
1606:                +\eps_{r_k}
1607:                c^{\dagger}_{r_k}c^{\phd}_{r_k})\\
1608:  &H_{\rm tun} = \sum_{k}[T_{l}(\hat{x}) c^{\dagger}_{l_k}c_1^{\phd} +
1609:                          T_{r}(\hat{x}) c^{\dagger}_{r_k}c_1^{\phd}] + h.c.\\
1610:  &H_{\rm bath} + H_{\rm int }= {\rm generic \; heat \; bath}
1611:  \end{split}
1612: \end{equation}
1613: 
1614: 
1615: %
1616: Also the mechanical degree of freedom is treated quantum
1617: mechanically. For example the position operator can be expressed
1618: in the form:
1619: %
1620: \be
1621: \hat{x} = \sqrt{\frac{\hbar}{2 m \w}}(d^{\dag}+d)
1622: \end{equation}
1623: %
1624: where $d^{\dag}$ and $d$
1625: are respectively the creation and annihilation operators for
1626: the harmonic oscillator.
1627: We neglect for a moment the mechanical bath and its coupling to
1628: the system and start the Gurvitz analysis of the model dynamics.
1629: The many-body basis introduced in (\ref{eq:basis}) must be
1630: extended to take into account also the phononic excitations of the
1631: system. For definiteness we choose the eigenvectors of the
1632: oscillator Hamiltonian as a basis for the mechanical part. We
1633: display the basis elements in the following table:
1634: 
1635: 
1636: \begin{equation}
1637: \begin{tabular}{|l|l|l|l}
1638: \hline
1639:    \multicolumn{1}{|c}{r=0}
1640:  & \multicolumn{1}{|c}{r=1}
1641:  & \multicolumn{1}{|c}{r=2}
1642:  & \multicolumn{1}{|c}{$\ldots$}\\
1643:  \hline
1644:    $|0\rangle$
1645:  &
1646: $d^{\dag}|0\rangle$
1647:  & ${d^{\dag}}^2|0\rangle$
1648:  & $\ldots$\\
1649:    $c_1^{\dagger}   c_{l_1}^{\phd}|0\rangle$
1650:  & $c_1^{\dagger}c_{l_1}^{\phd} d^{\dag}|0\rangle$
1651:  & $c_1^{\dagger}c_{l_1}^{\phd}{d^{\dag}}^2|0\rangle$
1652:  & $\ldots$\\
1653:    $c_{r_1}^{\dagger}c_{l_1}^{\phd}|0\rangle$
1654:  & $c_{r_1}^{\dagger}c_{l_1}^{\phd} d^{\dag}|0\rangle$
1655:  & $c_{r_1}^{\dagger}c_{l_1}^{\phd}{d^{\dag}}^2|0\rangle$
1656:  & $\ldots$\\
1657:    \multicolumn{1}{|c}{$\vdots$} &
1658:    \multicolumn{1}{|c}{$\vdots$} &
1659:    \multicolumn{1}{|c|}{$\vdots$} &
1660: \end{tabular}
1661: \label{eq:SDQSbasis}
1662: \end{equation}
1663: 
1664: %
1665: \noindent where $r$ is the number of excitations  of the
1666: oscillator and the empty state $|0\rangle$ represents the empty
1667: dot in its \emph{mechanical ground state} with leads filled up to
1668: their Fermi energies. It is convenient to organize the
1669: coefficients of the expansion of the state vector $|\Psi\rangle$
1670: in the basis (\ref{eq:SDQSbasis}) in vectors: one for each
1671: electronic configuration. The different elements of the vectors
1672: refer to the different excited states for the oscillator. The
1673: Schr\"odinger equation for the state vector $|\Psi\rangle$ is
1674: represented in the basis (\ref{eq:SDQSbasis}) by two recursive
1675: differential equations for the vector coefficients $\vet{b}$'s:
1676: 
1677: \be
1678: \bs
1679:  i\dot{\vet{b}}_{\{l_j\}_{j=1}^n}&_{\{r_j\}_{j=1}^n} =
1680:          \left[\hat{H}_{\rm osc}
1681:          +\sum_{k=1}^n(\eps_{r_k}-\eps_{l_k})\right]
1682:          \vet{b}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}\\
1683:         &+\sum_{l_{n+1}}T_{l}(\hat{x})
1684:          \vet{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}
1685:          +\sum_{k=1}^n(-1)^{k-n}T_{r}(\hat{x})
1686:          \vet{b}_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}\\
1687: %
1688:  i\dot{\vet{b}}_{1\{l_j\}_{j=1}^{n+1}}&_{\{r_j\}_{j=1}^n} =
1689:          \left[\hat{H}_{\rm osc} + \eps_1- e\mathcal{E}\hat{x}
1690:          +\sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1691:          -\eps_{l_{n+1}}\right]
1692:          \vet{b}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}\\
1693:         &+\sum_{r_{n+1}}T_{r}(\hat{x})
1694:          \vet{b}_{\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^{n+1}}
1695:          +\sum_{k=1}^{n+1}(-1)^{k-n-1}T_{l}(\hat{x})
1696:          \vet{b}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}\\
1697: \end{split}
1698: \label{eq:SDQSdiffsystem}
1699: \end{equation}
1700: %
1701: where $\hat{x}$ is given in its  matrix representation in terms of
1702: the occupation number basis ($\hbar = 1$):
1703: 
1704: \be
1705: \hat{x}_{rs} = \sqrt{\frac{r}{2 m \w}}(\d_{r,s+1}+\d_{r,s-1})
1706: \end{equation}
1707: %
1708: and the Hamiltonian for the harmonic oscillator in the same basis
1709: reads\footnote{The representation given in equation
1710: (\ref{eq:SDQSdiffsystem}) is actually independent of the basis
1711: for the oscillator Hilbert space. The $\vet{b}$ vectors are
1712: projections of the state vector $|\Psi\rangle$ on the particular
1713: subspace given by the electronic configuration specified by the
1714: subscript.}:
1715: 
1716: \be
1717:  \hat{H}_{\rm osc} = \w\left(\frac{1}{2} + r\d_{rs}\right)
1718: \end{equation}
1719: 
1720: All the constants in equation (\ref{eq:SDQSdiffsystem}) are
1721: identity operators in the mechanical Hilbert space.
1722: 
1723: One of the key assumptions in the derivation of the GME in the
1724: Gurvitz approach is the position of the energy levels of the
1725: system: they must lie well inside the transport window open
1726: between the chemical potentials of the leads. Since the oscillator
1727: spectrum is \emph{not} bounded from above we assume that only a
1728: finite number of mechanical excitations are involved in the
1729: dynamics of the system. We will see that, at least in the presence
1730: of a mechanical bath, this assumption is numerically fulfilled. In
1731: any case a violation of this condition in the final result would
1732: be unacceptable since it would violate the validity condition of
1733: the GME. From the point of view of the experimental realization of
1734: the device this limit is imposed at least by the leads that set an
1735: upper bound to the amplitude of the dot oscillations.
1736: 
1737: The Laplace transform of (\ref{eq:SDQSdiffsystem}) with the
1738: initial condition $|\Psi(t=0)\rangle = |0\rangle$ reads:
1739: 
1740: \be \bs &\left[E + \hat{H}_{\rm osc} + \sum_{k=1}^n
1741: (\eps_{r_k}-\eps_{l_k})\right]
1742:          \tilde{\vet{b}}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)
1743:          -\sum_{l_{n+1}}T_{l}(\hat{x})
1744:          \tilde{\vet{b}}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)\\
1745:          &\phantom{a}-\sum_{k=1}^n(-1)^{k-n}T_{r}(\hat{x})
1746:          \tilde{\vet{b}}_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}(E) =
1747:          i\d_{n0}\vet{v}_0\\
1748: %
1749: &\left[E + \hat{H}_{\rm osc} +\eps_1- e\mathcal{E}\hat{x}
1750:          + \sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1751:          -\eps_{l_{n+1}}\right]
1752:          \tilde{\vet{b}}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)\\
1753:          &\phantom{a}-\!\sum_{r_{n+1}}T_{r}(\hat{x})
1754:          \tilde{\vet{b}}_{\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^{n+1}}(E)
1755:          -\!\sum_{k=1}^{n+1}(-1)^{k-n-1}T_{l}(\hat{x})
1756:          \tilde{\vet{b}}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)=0
1757: \end{split}
1758: \label{eq:SDQSalgebsystem}
1759: \end{equation}
1760: %
1761: where $\vet{v}_0$ is the initial condition of the oscillator.
1762: 
1763: The continuous sums in the  system (\ref{eq:SDQSalgebsystem}) can
1764: be performed with an argument similar to the one used for the
1765: static QD. We have just to be careful with the matrix notation and
1766: change the basis to diagonalize the matrix:
1767: 
1768: \be
1769:  M = \left[E + \hat{H}_{\rm osc} +\eps_1- e\mathcal{E}\hat{x}
1770:          + \sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1771:          -\eps_{l_{n+1}}\right]
1772: \end{equation}
1773: %
1774: before taking the integral. The sum $\sum_{r_{n+1}}$ in the second
1775: Eq.\ of (\ref{eq:SDQSalgebsystem}) can be treated analogously. As
1776: in the static case the continuous sums are condensed into
1777: ``rates''\footnote{These ``rates'' are position dependent and then
1778: in our quantum treatment they are operators. Actual rates can be
1779: recovered by averaging these operators on a given the quantum
1780: state.}:
1781: 
1782: \be
1783: \bs
1784: &\left[E + \hat{H}_{\rm osc} + \sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1785:          -i\frac{\G_L(\hat{x})}{2} \right]
1786:          \tilde{\vet{b}}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)\\
1787:          &\phantom{aaaaaa}-\sum_{k=1}^n(-1)^{k-n}T_{r}(\hat{x})
1788:          \tilde{\vet{b}}_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}(E) =
1789:          i\d_{n0}\vet{v}_0\\
1790: %
1791: &\left[E + \hat{H}_{\rm osc} +\eps_1- e\mathcal{E}\hat{x}
1792:          + \sum_{k=1}^n (\eps_{r_k}-\eps_{l_k})
1793:          -\eps_{l_{n+1}} -i\frac{\G_R(\hat{x})}{2}\right]
1794:          \tilde{\vet{b}}_{1\{l_j\}_{j=1}^{n+1}\{r_j\}_{j=1}^n}(E)\\
1795:          &\phantom{aaaaaa}-\sum_{k=1}^{n+1}(-1)^{k-n-1}T_{l}(\hat{x})
1796:          \tilde{\vet{b}}_{\{l_j\}_{j=1}^{n+1}\backslash \{l_k\}\{r_j\}_{j=1}^n}(E)=0
1797: \end{split}
1798: \label{eq:SDQSrates}
1799: \end{equation}
1800: %
1801: where
1802: 
1803: \be
1804:  \G_{L,R}(\hat{x}) = 2\pi D_{L,R}T^2_{l,r}(\hat{x})
1805:                    = \G_{L,R}\, e^{\mp 2\hat{x}/\l}
1806: \end{equation}
1807: %
1808: and we have introduced  the tunneling length $\lambda$ and the
1809: bare injection (ejection) rate $\G_L$ ($\G_R$).
1810: 
1811: The reduced density matrix for the system contains information
1812: about the electrical occupation of the QD \emph{and} its
1813: mechanical state. Coherencies between occupied and empty state
1814: vanish because they imply coherencies between states with
1815: different particle number in the baths. The equation
1816: (\ref{eq:genRDM}) for the non-vanishing elements written in the
1817: static QD case still holds:
1818: 
1819: \be
1820:  \s_{ii}(t) = \sum_{n=0}^{\infty}
1821:  \sum_{\{B_n\}}\langle i_S,B_n|\Psi(t)\rangle\!\langle\Psi(t)|B_n,i_S\rangle
1822:  = \sum_{n=0}^{\infty}\s_{ii}^{(n)}(t)
1823: \end{equation}
1824: %
1825: with the difference that the  ``elements'' $\s_{ii}$ are now full
1826: operators in the mechanical Hilbert space. It is useful to express
1827: them in terms of the vectors $\vet{b}$:
1828: 
1829: \be
1830:  \bs
1831:   \s_{00}^{(n)}(t) &=\sum_{\{l_k\}\{r_k\}}
1832:    \vet{b}_{\{l_k\}_{k=1}^n\{r_k\}_{k=1}^n}(t)
1833:    \vet{b}_{\{l_k\}_{k=1}^n\{r_k\}_{k=1}^n}^{\dag}(t)\\
1834:   \s_{11}^{(n)}(t) &=\sum_{\{l_k\}\{r_k\}}
1835:    \vet{b}_{1\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}(t)
1836:    \vet{b}_{1\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}^{\dag}(t)
1837:  \end{split}
1838:  \label{eq:SDQSnRDM}
1839: \end{equation}
1840: %
1841: The notation of  equation (\ref{eq:SDQSnRDM}) can be understood in
1842: terms of the Dirac notation: $\vet{b}^{\dag}$ is the \emph{bra} of
1843: the corresponding vector $\vet{b}$ (the \emph{ket})\footnote{In
1844: other terms the linear operator $\s_{ii}$ is the tensor product of
1845: the vector $\vet{b}$ and the linear form $\vet{b}^{\dag}$.}. The
1846: inverse Laplace transform brings us back to the vectors
1847: $\tilde{\vet{b}}$:
1848: 
1849: \be
1850:  \bs
1851:   \s_{00}^{(n)}(t) &=\int \frac{dE dE'}{4 \pi^2}\sum_{\{l_k\}\{r_k\}}
1852:    \tilde{\vet{b}}_{\{l_k\}_{k=1}^n\{r_k\}_{k=1}^n}(E)
1853:    \tilde{\vet{b}}_{\{l_k\}_{k=1}^n\{r_k\}_{k=1}^n}^{\dag}(E')
1854:    e^{-i(E-E')t}\\
1855:   \s_{11}^{(n)}(t) &=\int \frac{dE dE'}{4 \pi^2}\sum_{\{l_k\}\{r_k\}}
1856:    \tilde{\vet{b}}_{1\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}(E)
1857:    \tilde{\vet{b}}_{1\{l_k\}_{k=1}^{n+1}\{r_k\}_{k=1}^n}^{\dag}(E')
1858:    e^{-i(E-E')t}
1859:  \end{split}
1860: \end{equation}
1861: 
1862: The case with $n=0$ must be treated separately. The inverse
1863: Laplace transform of the first equation in the system
1864: (\ref{eq:SDQSrates}) specialized for $n=0$ reads:
1865: 
1866: \be
1867:  i\dot{\vet{b}}_0 = \hat{H}_{\rm osc}\vet{b}_0
1868:                     -i\frac{\G_L(\hat{x})}{2}\vet{b}_0
1869: \label{eq:dotb0}
1870: \end{equation}
1871: %
1872: and its Hermitian conjugate:
1873: %
1874: \be
1875:  -i\dot{\vet{b}}_0^{\dag} = \vet{b}_0^{\dag}\hat{H}_{\rm osc}
1876:                     +i\vet{b}_0^{\dag}\frac{\G_L(\hat{x})}{2}
1877: \label{eq:dotb0dag}
1878: \end{equation}
1879: %
1880: where we have used the property of adjoint of vectors and
1881: operators $(AB)^{\dag} = B^{\dag}A^{\dag}$ and the fact that the
1882: oscillator Hamiltonian and position operator are Hermitian on
1883: the mechanical Hilbert space. The combination of equations
1884: (\ref{eq:SDQSnRDM}), (\ref{eq:dotb0}), (\ref{eq:dotb0dag}) and the
1885: Leibnitz rule for derivatives extended to the tensor product
1886: between vectors and linear forms lead to the first component of
1887: the GME for the SDQS:
1888: 
1889: \be
1890:  \dot{\s}_{00}^{(0)}(t) = -i\Big[\hat{H}_{\rm osc},\s_{00}^{(0)}\Big]
1891:                           -\frac{\G_L}{2}
1892:                           \Big\{e^{-2\hat{x}/\lambda},\s_{00}^{(0)}\Big\}
1893: \label{eq:SDQSGME0}
1894: \end{equation}
1895: %
1896: where $[A,B]\equiv AB - BA$ is the commutator and $\{A,B\} \equiv
1897: AB + BA$ the anticommutator between the operators $A$ and $B$.
1898: Equation (\ref{eq:SDQSGME0}) already contains the essence of the
1899: driving part of the GME: a coherent evolution represented by the
1900: commutator with the oscillator Hamiltonian and a non-coherent term
1901: due to the interaction with the bath. In this second contribution
1902: the quantum features are given by the particular ordering of the
1903: operators.
1904: 
1905: For the general case with $n \ne 0$ the procedure is to take the
1906: first equation of (\ref{eq:SDQSrates}) evaluated in $E$ and make
1907: the tensor product with
1908: $\tilde{\vet{b}}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}^{\dag}(E')$
1909: $e^{-i(E - E')t}$, subtract side by side the adjoint of the first
1910: equation in (\ref{eq:SDQSrates}) evaluated in $E'$ multiplied
1911: (from the right) with the vector
1912: $\tilde{\vet{b}}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)$
1913: $e^{-i(E-E')t}$, integrate in $dE$ and $dE'$ and sum over the
1914: possible bath configuration with $n$ extra electrons in the right
1915: lead. Using the properties of the Laplace transform and the
1916: representation of the reduced density matrix in terms of the
1917: vectors $\vet{b}$ (\ref{eq:SDQSnRDM}) we get:
1918: 
1919: \be
1920: \bs
1921:  i\dot{\s}&_{00}^{(n)}- \Big[\hat{H}_{\rm osc},\s_{00}^{(n)}\Big]
1922:                         -i\frac{\G_L}{2} \Big\{e^{-2\hat{x}/\l},\s_{00}^{(n)}\Big\}\\
1923:   &-\!\!\sum_{\{l_j\}\{r_j\}}\!\!\int\frac{dE dE'}{4 \pi^2}\!\sum_{k=1}^{n}
1924:   (-1)^{k-n}\Big[
1925:   T_r(\hat{x})
1926:   \tilde{\vet{b}}_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}(E)
1927:   \tilde{\vet{b}}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}^{\dag}(E')\\
1928:   &-\tilde{\vet{b}}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}(E)
1929:   \tilde{\vet{b}}_{1\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n\backslash \{r_k\}}^{\dag}(E')
1930:   T_r(\hat{x})
1931:   \Big]e^{-i(E-E')t} = 0
1932:   \end{split}
1933: \label{eq:SDQSrate2}
1934: \end{equation}
1935: 
1936: We solve the first equation in (\ref{eq:SDQSrates}) with respect
1937: to $\tilde{\vet{b}}_{\{l_j\}_{j=1}^n\{r_j\}_{j=1}^n}$. Then we
1938: insert the result and its adjoint in (\ref{eq:SDQSrate2}) and, as
1939: in the static QD, we are left with the only non-vanishing
1940: continuous sum in the ``missing'' variable $r_n$. The result is
1941: the matrix equation:
1942: 
1943: \be
1944:  \dot{\s}_{00}^{(n)}= -i\Big[\hat{H}_{\rm osc},\s_{00}^{(n)}\Big]
1945:  -\frac{\G_L}{2}\Big\{e^{-2\hat{x}/\l},\s_{00}^{(n)}\Big\}
1946:  +\G_R e^{\hat{x}/\l}\s_{11}^{(n-1)}e^{\hat{x}/\l}
1947: \end{equation}
1948: 
1949: The treatment of the second equation in (\ref{eq:SDQSrates}) is
1950: totally analogous and brings us to  an equation of motion for the
1951: charged component of the density matrix ($\s_{11}^{(n)}$).
1952: Collecting all the results we can write the $n$-resolved GME:
1953: 
1954: \be
1955: \bs
1956:  \dot{\s}_{00}^{(n)}&= -i\Big[\hat{H}_{\rm osc},\s_{00}^{(n)}\Big]
1957:         -\frac{\G_L}{2}\Big\{e^{-2\hat{x}/\l},\s_{00}^{(n)}\Big\}
1958:         +\G_R e^{+\hat{x}/\l}\s_{11}^{(n-1)}e^{+\hat{x}/\l}\\
1959:  \dot{\s}_{11}^{(n)}&= -i\Big[\hat{H}_{\rm osc} -e\mathcal{E}\hat{x},\s_{11}^{(n)}\Big]
1960:         -\frac{\G_R}{2}\Big\{e^{+2\hat{x}/\l},\s_{11}^{(n)}\Big\}
1961:         +\G_L e^{-\hat{x}/\l}\s_{00}^{(n)}e^{-\hat{x}/\l}
1962: \end{split}
1963: \label{eq:nGMESDQS}
1964: \end{equation}
1965: 
1966: In order to complete the description of the dynamics  of the SDQS
1967: we have to take into account the mechanical bath and its
1968: interaction with the system. We derive the mechanical component of
1969: the GME starting from the general formulation for the equation of
1970: motion of the reduced density matrix (\ref{eq:QOpGME}). We
1971: consider the problem described by the Hamiltonian (at the moment
1972: independent of the electronic dynamics):
1973: 
1974: \be
1975:  H = H_{\rm sys} + H_{\rm bath} + H_{\rm int}
1976: \end{equation}
1977: %
1978: where
1979: %
1980: \be
1981: \bs
1982:  H_{\rm sys} &= \hbar \w \left(\frac{1}{2} + d^{\dag}d\right)\\
1983:  H_{\rm bath} &= \sum_{\vet{q}} \hbar \w_{\vet{q}}
1984:  {d_{\vet{q}}}^{\dagger} d_{\vet{q}}\\
1985: \end{split}
1986: \end{equation}
1987: %
1988: and the generic interaction contribution:
1989: 
1990: \be
1991:  H_{\rm int} = \hbar\sum_a G_a F_a
1992:  \label{eq:GenInt}
1993: \end{equation}
1994: %
1995: $G_a$ being a system operator and $F_a$ a bath  operator and $a$ a
1996: generic quantum number. We will specialize later the interaction in the
1997: form we have introduced in the chapter dedicated to the model:
1998: 
1999: \be
2000:  H_{\rm int} = \hbar \sum_{\vet{q}}
2001:               g(d_{\vet{q}} + d^{\dagger}_{\vet{q}})
2002:               (d + d^{\dagger})
2003: \label{eq:nRWAint}
2004: \end{equation}
2005: %
2006: and with the rotating wave approximation
2007: 
2008: \be
2009:  H_{\rm int}^{\rm (RWA)} = \hbar \sum_{\vet{q}}
2010:               g(d^{\dagger}_{\vet{q}}d + d^{\dagger}d_{\vet{q}})
2011: \label{eq:RWAint}
2012: \end{equation}
2013: 
2014: 
2015: We start recalling the GME \eqref{eq:QOpGME} derived  in the weak
2016: coupling using the quantum optical formalism:
2017: 
2018: \be
2019:  \dot{\tilde{\s}} = -\frac{1}{\hbar^2} \int_{0}^{\infty} d\t
2020:  {\rm Tr_B}\{[\tilde{H}_{\rm int}(t),[\tilde{H}_{\rm int}(t-\t),
2021:  \tilde{\s}(t)\otimes \rho_{\rm B}]]\}
2022: \end{equation}
2023: %
2024: With the generic form for the interaction Hamiltonian (\ref{eq:GenInt}) we get:
2025: 
2026: \be
2027: \bs
2028:  \dot{\tilde{\s}}(t) = - \int_0^{\infty}\!\!d\t \sum_{ab}
2029:  \{
2030:   &[\tilde{G}_a(t)\tilde{G}_b(t-\t) \tilde{\s}(t)
2031:  -\tilde{G}_b(t-\t) \tilde{\s}(t) \tilde{G}_a(t)]
2032:   \langle \tilde{F}_a(\t)\tilde{F}_b(0) \rangle\\
2033:  +&[\tilde{\s}(t)\tilde{G}_a(t-\t)  \tilde{G}_b(t)
2034:   -\tilde{G}_b(t) \tilde{\s}(t) \tilde{G}_a(t-\t)]
2035:   \langle \tilde{F}_a(0)\tilde{F}_b(\t) \rangle
2036:  \}
2037: \end{split}
2038: \end{equation}
2039: %
2040: where the tilde indicates the  interaction picture and $\langle
2041: \bullet \rangle \equiv {\rm Tr_B}\{\rho_B \bullet\}$. We can
2042: easily go to the Schr\"odinger picture:
2043: 
2044: \be
2045: \bs
2046:  \dot{\s} = -\frac{i}{\hbar}[H_{\rm sys},\s] +
2047:   \sum_{ab}\int_0^{\infty} d \t
2048:   \{
2049:   &[\tilde{G}_b(-\t)\langle \tilde{F}_a(\t)\tilde{F}_b(0)\rangle\s,G_a]\\
2050:    +
2051:   &[G_b,\s\tilde{G}_a(-\t)\langle \tilde{F}_a(\t)\tilde{F}_b(0)\rangle]
2052:   \}
2053: \end{split}
2054: \end{equation}
2055: %
2056: Following \cite{koh-jcp-97} we introduce the compact notation:
2057: 
2058: \be
2059:  \dot{\s} = -\frac{i}{\hbar}[H_{\rm sys},\s] +
2060:   \sum_{a}\{[G_a^+\s,G_a] + [G_a,\s G_a^-]\}
2061: \label{eq:KohGME}
2062: \end{equation}
2063: %
2064: where
2065: 
2066: \be \bs G_a^+ &= \sum_b\int_0^{\infty} d\tau
2067:          \tilde{G}_b(-\t)\langle \tilde{F}_a(\t)\tilde{F}_b(0)\rangle\\
2068: G_a^- &= \sum_b\int_0^{\infty} d\tau
2069:          \tilde{G}_b(-\t)\langle \tilde{F}_b(0)\tilde{F}_a(\t)\rangle\\
2070: \end{split}
2071: \end{equation}
2072: %
2073: This formalism is very efficient since,  for a given interaction,
2074: it requires for the GME simply the calculation of the two
2075: operators $G^+$ and $G^-$. For the interaction Hamiltonian
2076: (\ref{eq:nRWAint}) we identify
2077: 
2078: \be
2079: \bs
2080:  G_{\vet{q}} &= d + d^{\dag}\\
2081:  F_{\vet{q}} &=g(d_{\vet{q}}^{\phd} + d_{\vet{q}}^{\dag})
2082: \end{split}
2083: \end{equation}
2084: %
2085: and calculate
2086: 
2087: \be
2088: \bs
2089:  G^+_{\vet{q}} =& \sum_{\vet{q}'}\int_0^{\infty} d\t
2090:  \tilde{G}_{\vet{q}'}(-\t)
2091:  \langle \tilde{F}_{\vet{q}}(\t)\tilde{F}_{\vet{q}'}(0)\rangle\\
2092:  =& d^{\phd}
2093:  g^2 \int_0^{\infty} d\t e^{i\w\t}
2094:  \{e^{ i\w_{\vet{q}}\t }n_B(\w_{\vet{q}}) +
2095:    e^{-i\w_{\vet{q}}\t }[1+ n_B(\w_{\vet{q}})] \}\\
2096:  &+ d^{\dag}
2097:  g^2  \int_0^{\infty} d\t e^{-i\w\t}
2098:  \{e^{ i\w_{\vet{q}}\t }n_B(\w_{\vet{q}}) +
2099:    e^{-i\w_{\vet{q}}\t }[1+ n_B(\w_{\vet{q}})] \}\\
2100:    =& \pi g^2 \{d [1+n_B(\w_{\vet{q}})] + d^{\dag}n_B(\w_{\vet{q}}) \}
2101:    \d(\w - \w_{\vet{q}})
2102: \end{split}
2103: \end{equation}
2104: %
2105: where we assumed the bath in thermal equilibrium and calculated the average:
2106: 
2107: \be
2108:  \langle d_{\vet{q}}^{\dag}d_{\vet{q}}^{\phd}\rangle
2109:  = n_B(\w_{\vet{q}}) \equiv \frac{1}{e^{\b\hbar\w_{\vet{q}}}-1}
2110: \end{equation}
2111: %
2112: and we integrated the exponentials
2113: %
2114: \be
2115:  \int_0^{\infty} d\t e^{i(\w \pm \w_{\vet{q}})\t}
2116:  = \pi\d(\w \pm \w_{\vet{q}})
2117:  + \mathcal{P} \left(\frac{i}{\w \pm \w_{\vet{q}}}\right)
2118:  \approx \pi\d(\w \pm \w_{\vet{q}})
2119: \end{equation}
2120: %
2121: where $\mathcal{P}$ denotes the principal  value. We neglected the
2122: contribution due the principal value that only slightly shifts the
2123: oscillator frequency $\w$. Terms proportional to $\d(\w +
2124: \w_{\vet{q}})$ vanish since both $\w$ and $\w_{\vet{q}}$ are
2125: positive. The operator $G^-$ can be calculated in a similar way
2126: and reads:
2127: 
2128: \be
2129:  G^-_{\vet{q}}= \pi g^2 \{ d n_B(\w_{\vet{q}}) +
2130:    d^{\dag} [1+n_B(\w_{\vet{q}})] \}
2131:    \d(\w - \w_{\vet{q}}) = (G^+_{\vet{q}})^{\dag}
2132: \end{equation}
2133: %
2134: Substituting all the $G$ operators, the GME (\ref{eq:KohGME}) takes the form:
2135: 
2136: \be
2137:  \dot{\s} = -\frac{i}{\hbar}[H_{\rm sys},\s] +
2138:  \frac{\g}{2}[1 + n_B(\w)] [d + d^{\dag}, \s d^{\dag} -d \s ]
2139:  +
2140:  \frac{\g}{2}n_B(\w) [d + d^{\dag}, \s d -d^{\dag} \s ]
2141: \label{eq:GMEnRWA}
2142: \end{equation}
2143: %
2144: where $\gamma = 2 \pi D(\w) g^2$ is the damping  rate and $D(\w)$
2145: is the density of states of the phonon bath at the frequency of
2146: the oscillator. We rewrite the previous equation in the form:
2147: 
2148: \be
2149:  \dot{\s} = \mathcal{L}[\s] = \mathcal{L}_{\rm coh}[\s]
2150:  + \mathcal{L}_{\rm damp} [\s]
2151: \end{equation}
2152: %
2153: The linear (super) operator $\mathcal{L}$ is  also known as
2154: the Liouvillean and is a linear operator defined on the space of
2155: linear operators on the Hilbert space of the system. The first
2156: term $\mathcal{L}_{\rm coh} = -\frac{i}{\hbar}[H_{\rm sys},\s]$
2157: describes the coherent dynamics of the isolated system. The terms
2158: proportional to $\gamma$ and grouped in $\mathcal{L}_{\rm damp}
2159: \s$ represent the interaction with the bath which is damping the
2160: oscillator. This interaction introduces decoherence in the system
2161: in the sense that no matter how we prepare the initial quantum
2162: state of the oscillator ($\s(t=0)$), in absence of other driving
2163: forces, the stationary state reached at long times ($\s(t =
2164: \infty)$) is a thermal distribution that corresponds to a diagonal
2165: density matrix with no coherencies left. The thermal stationary
2166: solution is a typical property required from a generalized master
2167: equation. The GME (\ref{eq:GMEnRWA}) is also translationally
2168: invariant as can be more directly checked introducing the position
2169: and momentum operators for the oscillator $x,p$\footnote{From now
2170: on we drop for simplicity the hat for the operators. It will
2171: be clear from the context if we are dealing with operators or
2172: with classical variables.}:
2173: 
2174: \be
2175:  \dot{\s} = -\frac{i}{\hbar}[H_{\rm sys},\s]
2176:             -\frac{i\g}{2\hbar} [x,\{p,\s\}]
2177:             -\frac{\g m \w}{\hbar}\left[n_B(\w)+\frac{1}{2}\right]
2178:              [x,[x,\s]]
2179: \label{eq:TrasInvLiou}
2180: \end{equation}
2181: 
2182: 
2183: Unfortunately a translationally invariant  GME with a thermal
2184: stationary solution is generating density matrices which are not
2185: \emph{a priori} always positive definite. This is a general
2186: problem of the GME \cite{koh-jcp-97}. In our specific case though
2187: we checked numerically that in the relevant cases the positivity
2188: was not broken within numerical errors.
2189: 
2190: The interaction Hamiltonian in the rotating wave approximation
2191: (\ref{eq:RWAint}) can be treated in a similar way. One has to
2192: extend the space of quantum numbers and define:
2193: 
2194: \be
2195: \begin{array}{ll}
2196:  G_{\vet{q}_1} = d & G_{\vet{q}_2} = d^{\dag}\\
2197:  F_{\vet{q}_1} = g d_{\vet{q}}^{\dag} & F_{\vet{q}_2} = g d_{\vet{q}}
2198: \end{array}
2199: \end{equation}
2200: %
2201: The corresponding $G^+$ and $G^-$ operators read:
2202: 
2203: \be
2204: \bs
2205: G_{\vet{q}_1}^+ &= d^{\dag}
2206: \pi g^2 n_B(\w_{\vet{q}}) \d(\w - \w_{\vet{q}})\\
2207: G_{\vet{q}_2}^+ &= d
2208: \pi g^2 [1 + n_B(\w_{\vet{q}})] \d(\w - \w_{\vet{q}})\\
2209: G_{\vet{q}_1}^- &=(G_{\vet{q}_2}^+)^{\dag}\\
2210: G_{\vet{q}_2}^- &=(G_{\vet{q}_1}^+)^{\dag}
2211: \end{split}
2212: \end{equation}
2213: %
2214: We insert the operators in the general GME (\ref{eq:KohGME}) and
2215: obtain:
2216: 
2217: \be
2218: \bs
2219:  \dot{\s} = -\frac{i}{\hbar}[H_{\rm sys},\s]
2220:  &+ \frac{\g}{2} n_B(\w)([d^{\dag}\s,d] + [d^{\dag},\s d])\\
2221:  &+ \frac{\g}{2} [1+ n_B(\w)]([d,\s d^{\dag}] + [d\s, d^{\dag}])
2222: \end{split}
2223: \label{eq:RWAdampkernel}
2224: \end{equation}
2225: %
2226: where we have defined as usual the damping rate $\g \equiv 2 \pi D g^2$.
2227: Collecting all the terms we can write the GME for the SDQS in the form:
2228: 
2229: \be
2230: \begin{tabular}{|rl|}
2231: \hline
2232:  \phantom{a} & \phantom{a}\\
2233:  $\dot{\s}_{00}=$& $\!\!\!-i\Big[H_{\rm osc},\s_{00}\Big]
2234:         -\frac{\G_L}{2}\Big\{e^{-\frac{2x}{\l}},\s_{00}\Big\}
2235:         +\G_R e^{\frac{x}{\l}}\s_{11}e^{\frac{x}{\l}}
2236:         + \mathcal{L}_{\rm damp}[\s_{00}]$\\
2237:  \phantom{a} & \phantom{a}\\
2238:  $\dot{\s}_{11}=$& $\!\!\!-i\Big[H_{\rm osc} -e\mathcal{E}x,\s_{11}\Big]
2239:         -\frac{\G_R}{2}\Big\{e^{\frac{2x}{\l}},\s_{11}\Big\}
2240:         +\G_L e^{-\frac{x}{\l}}\s_{00}e^{-\frac{x}{\l}}
2241:         + \mathcal{L}_{\rm damp}[\s_{11}]$\\
2242:  \phantom{a} & \phantom{a}\\
2243: \hline
2244: \end{tabular}
2245: \label{eq:SDQSGME}
2246: \end{equation}
2247: %
2248: where we have taken the sum over the number of electrons  collected
2249: in the resevoir and we have introduced the generic damping
2250: Liouvillean $\mathcal{L}_{\rm damp}$. One can use for example one
2251: of the two damping  Liouvillean we have just derived. In the rest
2252: of the thesis we will always adopt for the SDQS the
2253: translationally invariant damping Liouvillean (\ref{eq:GMEnRWA}).
2254: We will also refer to the electronic part of the equation
2255: (\ref{eq:SDQSGME}) as to the driving term. In a compact form:
2256: 
2257: \be
2258:  \dot{\s} = \mathcal{L}_{\rm coh}[\s] + \mathcal{L}_{\rm driv}[\s] +
2259:  \mathcal{L}_{\rm damp}[\s]
2260: \end{equation}
2261: %
2262: where we have introduced a block matrix structure to  take care of
2263: the mechanical and electrical degrees of freedom simultaneously.
2264: 
2265: \subsection{Triple Dot Quantum Shuttle}
2266: %
2267: The driving term of the Liouvillean operator for the Triple  Dot
2268: Quantum Shuttle can be derived in strict analogy with the fixed double
2269: dot. The major simplification with respect to the SDQS is
2270: the drop of the position dependence in the coupling to the leads
2271: as one can see from the Hamiltonian for the model:
2272: 
2273: \be
2274:     H =H_{\rm sys}+ H_{\rm leads} + H_{\rm bath}
2275:       +H_{\rm tun}+ H_{\rm int}
2276: \end{equation}
2277: %
2278: where
2279: %
2280: \be
2281:  \bs
2282:   H_{\rm sys} =&
2283:    \e_0 |0\rangle\!\langle 0|
2284:    + \frac{\D V}{2} |L\rangle \! \langle L|
2285:    - \frac{\D V}{2 x_0}x  |C\rangle\!\langle C|
2286:    - \frac{\D V}{2} |R\rangle \! \langle R|
2287:    + \hbar \w \left(d^{\dag}d + \frac{1}{2}\right)\\
2288:    &+t_R(x)(|C \rangle\!\langle R| + |R \rangle\!\langle C|)
2289:     +t_L(x)(|C \rangle\!\langle L| + |L \rangle\!\langle C|)\\
2290: %
2291:   H_{\rm leads} =&
2292:   \sum_{k}(\eps_{l_k}
2293: c^{\dagger}_{l_k}c^{\phd}_{l_k}
2294:           +\eps_{r_k}
2295: c^{\dagger}_{r_k}c^{\phd}_{r_k})\\
2296: %
2297: H_{\rm bath} =&
2298:  \sum_{\vet{q}} \hbar \w_{\vet{q}}{d_{\vet{q}}}^{\dagger} d_{\vet{q}}\\
2299: %
2300:   H_{\rm tun} =&
2301:   \sum_{k}T_{l}( c^{\dagger}_{l_k}|0\rangle\!\langle L| +
2302:                  c^{\phd}_{l_k}   |L\rangle\!\langle 0|)
2303:          +T_{r}( c^{\dagger}_{r_k}|0\rangle\!\langle R| +
2304:                  c^{\phd}_{r_k}   |R\rangle\!\langle 0|) \\
2305: %
2306:  H_{\rm int } =&
2307:  \hbar \sum_{\vet{q}} g(d^{\dagger}_{\vet{q}}d + d^{\dagger}d_{\vet{q}})
2308:  \end{split}
2309: \end{equation}
2310: %
2311: The reduced density matrix for the  triple dot system takes into
2312: account mechanical and electrical degrees of freedom. As in the
2313: case of the fixed double dot we can organize the density matrix
2314: in a block representation:
2315: 
2316: \be
2317: \s =
2318: \left[
2319: \begin{array}{cccc}
2320: \s_{00} & \s_{0L} & \s_{0C} & \s_{0R}\\
2321: \s_{L0} & \s_{LL} & \s_{LC} & \s_{LR}\\
2322: \s_{C0} & \s_{CL} & \s_{CC} & \s_{CR}\\
2323: \s_{R0} & \s_{RL} & \s_{RC} & \s_{RR}
2324: \end{array}
2325: \right]
2326: \end{equation}
2327: %
2328: Due to the incoherent leads the  elements $\s_{0\a}$ and
2329: $\s_{\a0}$ with $\a= L,C,R$ identically vanish\footnote{The
2330: incoherent leads do not have coherent superposition of states with
2331: different particle number. Only this kind of ``forbidden'' bath
2332: states would allow coherent superposition of states with 0 and 1
2333: particle in the device and thus non-vanishing coherencies
2334: $\s_{\a0}$ or $\s_{0\a}$.}. In the same matrix representation we
2335: write\footnote{This equation was first derived in the Gurvitz
2336: scheme by Armour and MacKinnon in \cite{arm-prb-02}.} the driving
2337: Liouvillean:
2338: 
2339: \be
2340:  \mathcal{L}_{\rm driv}[\s] =
2341:  \left[
2342:  \begin{array}{cccc}
2343:   \G_R \s_{RR}-\G_L \s_{00} & 0            & 0 & 0 \\
2344:   0 & \G_L \s_{00} & 0 & -\textstyle{\frac{1}{2}} \G_R \s_{LR} \\
2345:   0 & 0 & 0 & -\textstyle{\frac{1}{2}}\G_R \s_{CR} \\
2346:   0 & -\textstyle{\frac{1}{2}}\G_R \s_{RL}
2347:     & -\textstyle{\frac{1}{2}}\G_R \s_{RC}
2348:     & -\G_R \s_{RR}
2349:  \end{array}
2350:  \right]
2351: \end{equation}
2352: %
2353: where the injection and ejection rates have the form:
2354: 
2355: \be
2356:  \G_{L,R} = 2 \pi D_{L,R} T_{l,r}^2
2357: \end{equation}
2358: 
2359: The overall structure of the driving  component of the Liouvillean
2360: can be understood in terms of ``decaying channels'', since the
2361: interaction with the continuum of states in the leads gives rise
2362: to incoherent tunneling processes. This concept is underlying the
2363: following formulation of the incoherent dynamics
2364: \cite{gur-prb-96}:
2365: 
2366: \be
2367: \bs
2368:  \Big(\mathcal{L}_{\rm driv}[\s]\Big)_{\a\b} =&
2369:   -\frac{1}{2}\s_{\a\b}\left(
2370:   \sum_{\d \ne \a} \G_{\a \to \d}+
2371:   \sum_{\d \ne \b} \G_{\b \to \d}
2372:   \right)\\
2373:   &+\frac{1}{2}\sum_{\a'\b' \ne \a\b} \s_{\a'\b'}
2374:   (\G_{\a' \to \a} + \G_{\b' \to \b})
2375:  \end{split}
2376:  \label{eq:genrates}
2377: \end{equation}
2378: %
2379: where $\a,\a',\b,\b',\d = 0,L,C,R$ and $\G_{\a \to \b}$  is the
2380: probability per unit time for the system to make a transition from
2381: state $|\a\rangle$ to state $|\b\rangle$. The generic state $|\a
2382: \rangle \! \langle \b|$ is emptied (first line in equation
2383: \ref{eq:genrates}) and pumped (second line) with different $\G$
2384: rates. The central dot does not contribute to this incoherent
2385: dynamics since it is coupled only to the left and right dot
2386: \emph{discrete} states. Due to the left-right asymmetry only the
2387: rates $\G_{0 \to L} \equiv \G_L$ and $\G_{R \to 0} \equiv \G_R$
2388: are non-zero. The mechanical state of the system is not involved
2389: in the equation (\ref{eq:genrates}) but plays an active role in
2390: the coherent dynamics. In block matrix notation the system
2391: Hamiltonian takes the form:
2392: %
2393: \be
2394:  H_{\rm sys} = \left[
2395:  \begin{array}{cccc}
2396:  \eps_0 + H_{\rm osc} & 0 & 0 & 0\\
2397:  0 & \frac{\D V}{2}+ H_{\rm osc} & t_L(x) & 0\\
2398:  0 & t_L(x) & -\frac{\D V}{2 x_0}x + H_{\rm osc}& t_R(x)\\
2399:  0 & 0 & t_R(x) & -\frac{\D V}{2}+ H_{\rm osc}
2400:  \end{array}
2401:  \right]
2402: \end{equation}
2403: %
2404: The corresponding coherent Liouvillean reads:
2405: %
2406: \be
2407:  \Big(\mathcal{L}_{\rm coh}[\s]\Big)_{\a\b} =
2408:  -i\sum_{\d}\left[
2409:  (H_{\rm sys})_{\a\d} \s_{\d\b} -  \s_{\a\d}(H_{\rm sys})_{\d\b}
2410:  \right]
2411: \end{equation}
2412: %
2413: We assume for the damping term the same used by Armour and
2414: MacKinnon.  It is the standard quantum optical damping in RWA that
2415: we derived in the previous section (see eq.
2416: \eqref{eq:RWAdampkernel}):
2417: %
2418: \be
2419: \bs
2420:  \mathcal{L}_{\rm damp}[\s] =&
2421:   -\frac{\g}{2}n_B(\w)(dd^{\dag}\s - 2d^{\dag}\s d + \s dd^{\dag})\\
2422:  &-\frac{\g}{2}[n_B(\w)+ 1](d^{\dag}d\s - 2d\s d^{\dag} + \s d^{\dag}d)
2423: \end{split}
2424: \end{equation}
2425: %
2426: We write for completeness the system of equations for the 10
2427: non-vanishing sub-matrices of the reduced density matrix:
2428: %
2429: \be
2430: \begin{tabular}{|rl|}
2431: \hline
2432: \phantom{.} & \\
2433:  $\dot{\s}_{00} =$ & $-i[H_{\rm osc},\s_{00}]
2434:                       +\G_R \s_{RR} - \G_L \s_{00}
2435:                       +\mathcal{L}_{\rm damp}[\s_{00}]$\\
2436: \phantom{.} & \\
2437: %%
2438:  $\dot{\s}_{LL} =$ & $-i[H_{\rm osc},\s_{LL}]
2439:                       -it_L(x)\s_{CL} + i\s_{LC}t_L(x)
2440:                       +\G_L \s_{00}
2441:                       +\mathcal{L}_{\rm damp}[\s_{LL}]$\\
2442: \phantom{.} & \\
2443: %%
2444:  $\dot{\s}_{CC} =$ & $-i\left[H_{\rm osc}
2445:                       - \frac{\D V}{2 x_0}x,\s_{CC}\right]
2446:                       -it_L(x)\s_{LC} + i\s_{CL}t_L(x)$\\
2447: %\phantom{.} &  \\
2448:                 &$-it_R(x)\s_{RC} + i\s_{CR}t_R(x)
2449:                  +\mathcal{L}_{\rm damp}[\s_{CC}]$\\
2450: \phantom{.} & \\
2451: %%
2452:  $\dot{\s}_{RR} =$ & $-i[H_{\rm osc},\s_{RR}]
2453:                       -it_R(x)\s_{CR} + i\s_{RC}t_R(x)
2454:                       -\G_R \s_{RR}
2455:                       +\mathcal{L}_{\rm damp}[\s_{RR}]$\\
2456: \phantom{.} & \\
2457: %%
2458:  $\dot{\s}_{LC} =$ & $-i[H_{\rm osc},\s_{LC}]
2459:                       -i\frac{\D V}{2}\s_{LC}
2460:                       \left(1 + \frac{x}{x_0}\right)
2461:                       -it_L(x)\s_{CC} + i\s_{LL}t_L(x)$\\
2462: %\phantom{.} & \\
2463:                 &$+i\s_{LR}t_R(x) +\mathcal{L}_{\rm damp}[\s_{LC}]$\\
2464: \phantom{.} & \\
2465: %%
2466:  $\dot{\s}_{CL} =$ & $-i[H_{\rm osc},\s_{CL}]
2467:                       +i\frac{\D V}{2}\left(1 + \frac{x}{x_0}\right)\s_{CL}
2468:                       +i\s_{CC}t_L(x) - it_L(x)\s_{LL}$\\
2469: %\phantom{.} & \\
2470:                 &    $-it_R(x)\s_{RL} +\mathcal{L}_{\rm damp}[\s_{CL}]$\\
2471: \phantom{.} & \\
2472: %%
2473:  $\dot{\s}_{LR} =$ & $-i[H_{\rm osc},\s_{LR}]
2474:                       -i\D V\s_{LR} -i t_L(x)\s_{CR}
2475:                       +i\s_{LC}t_R(x)$\\
2476: %\phantom{.} & \\
2477:                 &$-\frac{\G_R}{2}\s_LR
2478:                  +\mathcal{L}_{\rm damp}[\s_{LR}]$\\
2479: \phantom{.} & \\
2480: %%
2481:  $\dot{\s}_{RL} =$ & $-i[H_{\rm osc},\s_{RL}]
2482:                       +i\D V\s_{RL} +i\s_{RC} t_L(x)
2483:                       -it_R(x)\s_{CL}$\\
2484: %\phantom{.} & \\
2485:                 & $-\frac{\G_R}{2}\s_RL
2486:                    +\mathcal{L}_{\rm damp}[\s_{RL}]$\\
2487: \phantom{.} & \\
2488: %%
2489:  $\dot{\s}_{CR} =$ & $-i[H_{\rm osc},\s_{CR}]
2490:                       -it_L(x)\s_{LR}
2491:                      -i\frac{\D V}{2}\left(1-\frac{x}{x_0}\right)\s_{CR}$\\
2492: %\phantom{.} & \\
2493:                 &$-i t_R(x)\s_{RR} + i \s_{CC}t_R(x)
2494:                  +\mathcal{L}_{\rm damp}[\s_{CR}]$\\
2495: \phantom{.} & \\
2496: %%
2497:  $\dot{\s}_{RC} =$& $-i[H_{\rm osc},\s_{RC}]
2498:                      +i\s_{RL}t_L(x)
2499:                      +i\frac{\D V}{2}\s_{RC}\left(1-\frac{x}{x_0}\right)$\\
2500: %\phantom{.} & \\
2501:                 &$+i \s_{RR}t_R(x) - i t_R(x)\s_{CC}
2502:                  +\mathcal{L}_{\rm damp}[\s_{RC}]$\\
2503: \phantom{.} & \\
2504: \hline
2505: \end{tabular}
2506: \label{eq:TDQSGME}
2507: \end{equation}
2508: 
2509: \section{The stationary solution: a numerical challenge}\label{sec:numericalGME}
2510: 
2511: The master equation generally describes the irreversible dynamics
2512: due to the coupling between the system and the infinite number of
2513: degrees of freedom of the environment. It is reasonable to require
2514: that in absence of a driving mechanism the system tends
2515: asymptotically to thermalize with the bath. This property is
2516: reflected in the evolution of the reduced density matrix that in
2517: this case has a thermal stationary solution. In the case of
2518: shuttle devices the oscillator is driven by the electrical
2519: dynamics: every time an electron jumps onto the moving island it
2520: feels an electrostatic force that excites the oscillator. For this
2521: reason, at least for small enough damping we do \emph{not} expect
2522: the oscillator to relax to the stationary thermal distribution.
2523: Nevertheless since both the electronic and the mechanical degrees
2524: of freedom of the system are coupled to baths we \emph{do} expect
2525: a stationary solution for the GMEs (\ref{eq:SDQSGME}) and
2526: (\ref{eq:TDQSGME}), i.e. a matrix $\s^{stat}$ that fulfills the
2527: condition:
2528: 
2529: \be
2530:  \mathcal{L} \s^{stat} = 0
2531: \label{eq:stationary}
2532: \end{equation}
2533: 
2534: \subsection{A matter of matrix sizes}
2535: We calculate the stationary matrix numerically: we have to find
2536: the null vector of the matrix representation for the Liouvillean
2537: super-operator $\mathcal{L}$. The challenge arises from the matrix
2538: sizes. In principle the reduced density matrix has infinite size
2539: due to the mechanical degree of freedom. In order to treat the
2540: problem numerically we truncate the corresponding Hilbert space
2541: retaining only the first $N$ states of the harmonic oscillator
2542: basis\footnote{The case of a different mechanical potential is not
2543: more difficult in principle, once we know the eigenvectors and
2544: eigenvalues of the corresponding one-dimensional Schr\"odinger
2545: equation.}. The size of the reduced density matrix $\s$ is in the
2546: SDQS and TDQS respectively $2N \times 2N$ and $4N \times 4N$:
2547: prefactors 2 and 4 come from the size of the electrical Hilbert
2548: space which is spanned by the vectors $|0\rangle, |1\rangle$ for the
2549: single-dot device and $|0\rangle, |L\rangle,|C\rangle,|R\rangle$
2550: for the triple dot.
2551: 
2552: The Liouvillean is a linear operator on the space of
2553: Hilbert-Schmidt operators\footnote{Given some Hilbert space
2554: $\mathcal{H}$ an operator  $A$ is of Hilbert-Schmidt if it is
2555: linear and  ${\rm Tr}\{A^{\dag}A\}$ is finite.} on the Hilbert
2556: space for the system (Liouville space). Equipped with the scalar
2557: product:
2558: 
2559: \be
2560:  (A,B) \equiv {\rm Tr}\{A^{\dag}B\}
2561: \end{equation}
2562: %
2563: the Liouville space  becomes a Hilbert space. One can therefore
2564: introduce an orthonormal basis and represent linear operators as
2565: matrices. The truncated Hilbert space for the system gives rise to
2566: finite size Liouvillean matrices: for the SDQS we reach the size
2567: of $4N^2 \times 4N^2$ while the richer electronic structure of
2568: the TDQS is reflected in a $16N^2 \times 16N^2$ elements
2569: Liouvillean. Even with the condition of incoherent baths that
2570: prevents coherencies within states with different electron number
2571: in the system and therefore sets to zero all sub-matrices in the
2572: form $\s_{01}$ or $\s_{10}$ in the SDQS and $\s_{0\a}$ or $\s_{\a
2573: 0}$ with $\a = L,C,R$ in the TDQS we can not reduce the size of
2574: the Liouvillean matrix to more than $2N^2 \times 2N^2$ and $10N^2
2575: \times 10N^2$ respectively.
2576: 
2577: The description of the shuttle device dynamics requires
2578: (especially in the shuttling regime) amplitude oscillations of the
2579: vibrating dot between 5 and 10 times larger than the zero point
2580: fluctuations. For this reason, in both devices, we are left to
2581: study the null space of matrices of typical size of $2 \cdot 10^4 \times
2582: 2\cdot 10^4$. To indicate that we are treating the truncated matrix
2583: representation of the original stationary state problem we change
2584: slightly notation and formulate the numerical problem:
2585: 
2586: \be
2587:   \vet{L}\vet{p}^{stat} = 0
2588: \end{equation}
2589: %
2590: with ${\rm Tr}\{\vet{p}^{stat}\} = 1$. The solution to this
2591: numerical problem came from prof.~Timo Eirola, Helsinki University
2592: of Technology, in the form of a suggestion and implementation for
2593: the SDQS of the iterative Arnoldi scheme. We successfully extended
2594: the method to the TDQS.
2595: 
2596: The Arnoldi scheme applied to the calculation of the null space
2597: has advantages with respect to the singular value decomposition
2598: both in terms of computational speed and memory
2599: consumption\footnote{For the theory and implementation of the
2600: Arnoldi scheme we refer to the lecture notes ``Numerical Linear
2601: Algebra; Iterative Methods" by Eirola and Nevanlinna
2602: \cite{eirola}.}. First we do not need to store the Liouvillean
2603: matrix and we can always work with operators on the system Hilbert
2604: space only; second we iteratively look for the best approximation
2605: to the null vector $\vet{p}^{stat}$ in spaces which are typically
2606: much smaller that the Liouville space. Good introductions to the
2607: Arnoldi scheme can be found in different places in the literature
2608: (for example \cite{golub} or \cite{eirola}). We dedicate the next
2609: section to a detailed analysis of the Arnoldi iteration scheme
2610: referring for examples to the SDQS Liouvillean. Some of the
2611: material can be found also in the appendix of
2612: \cite{fli-preprint-04}.
2613: 
2614: \subsection{The Arnoldi scheme}
2615: 
2616: The central r\^ole in the Arnoldi scheme is played by Krylov spaces.
2617: For a given Liouvillean $\vet{L}$ and a vector $\vet{x}_0$ of the
2618: Liouville space we define the Krylov space as:
2619: %
2620: \be
2621:  \mathcal{K}_j(\vet{L},\vet{x}_0)
2622:  \equiv {\rm span}(\vet{x}_0,\vet{Lx}_0,\ldots,\vet{L}^{j-1}\vet{x}_0)
2623: \end{equation}
2624: %
2625: where $j$ is a small\footnote{We mean small compared to the
2626: dimension of the Liouville space. We used  $j = 20$ for a
2627: Liouville space dimension of roughly $2\cdot 10^4$.} natural
2628: number. It is important to note that for the construction of the
2629: Krylov space all what we need are the vectors $\vet{x}_0$,
2630: $\vet{Lx}_0$, $\vet{L}^2\vet{x}_0$, $\ldots$ and not explicitly
2631: the matrix $\vet{L}$. In the SDQS device we would take an
2632: arbitrary state represented by the two matrices $\s_{00}$ and
2633: $\s_{11}$ and choose a reshaping procedure to map them into a
2634: single vector $\vet{x}_0 \in \mathbb{C}^{2N^2 \times 1}$. The
2635: vector $\vet{Lx}_0$ is obtained applying the operator defined in
2636: equation (\ref{eq:SDQSGME}) to the density matrices $\s_{00}$ and
2637: $\s_{11}$ and then reshaping with the same procedure used for
2638: $\vet{x}_0$.
2639: 
2640: The Arnoldi iteration starts with the construction of an
2641: orthonormal basis $\{\vet{q}_k\}_{k=1}^j$ for the Krylov space
2642: using standard Gram-Schmidt orthonormalization:
2643: 
2644: \be \bs
2645:  \vet{q}_1 &= \frac{\vet{x}_0}{\|\vet{x}_0\|}\\
2646:  \vet{q}_{k+1} &= \frac
2647:    {\vet{Lq}_k-\sum_{i=1}^k[\vet{q}_i^{\dag}\cdot \vet{Lq}_k]\vet{q}_i}
2648:  {\|\vet{Lq}_k-\sum_{i=1}^k[\vet{q}_i^{\dag}\cdot \vet{Lq}_k]\vet{q}_i\|}
2649:  ,k=1,\ldots,j-1
2650: \end{split}
2651: \end{equation}
2652: %
2653: where the norm in the Liouville space is defined from the
2654: canonical Hermitian product  $\|\vet{a}\|_2 \equiv
2655: \sqrt{\vet{a}^{\dag}\vet{a}}$ and has nothing to do with the
2656: scalar product we introduced to demonstrate that the Liouville
2657: space is a Hilbert space. Each new tentative basis vector
2658: $\vet{Lq}_k$ is first orthogonalized (by subtracting all
2659: components in the preceding vectors of the basis) and then
2660: normalized. This requires the calculation of the quantities:
2661: 
2662: \be
2663:  h_{i,k} = \vet{q}_i^{\dag}\cdot \vet{Lq}_k,
2664:  i=1,\ldots,k;k=1,\ldots,j
2665: \end{equation}
2666: %
2667: and
2668: %
2669: \be
2670:  h_{k+1,k} = \|\vet{Lq}_k-\sum_{i=1}^k[\vet{q}_i^{\dag}\cdot
2671:  \vet{Lq}_k]\vet{q}_i\|, k=1,\ldots,j
2672: \end{equation}
2673: %
2674: that are stored into the upper Hessenberg matrix
2675: 
2676: \be
2677:  \vet{H}_j =
2678:  \left[
2679:  \begin{array}{ccccc}
2680:  h_{1,1} & h_{1,2} & h_{1,3} & \cdots & h_{1,j}\\
2681:  h_{2,1} & h_{2,2} & h_{2,3} & \cdots & h_{2,j}\\
2682:  0 & h_{3,2} & h_{3,3} & \cdots & h_{3,j}\\
2683:  0 & 0 & h_{4,3} & \cdots & h_{4,j}\\
2684: \vdots & \vdots & \vdots & \ddots & \vdots\\
2685: 0 & 0 & 0 & \cdots & h_{j+1,j}
2686:  \end{array}
2687:  \right] \in \mathbb{C}^{j+1 \times j}
2688: \end{equation}
2689: %
2690: while the basis vectors are stored as columns in the matrix
2691: 
2692: \be
2693:  \vet{Q}_j = [\vet{q}_1,\vet{q}_2, \ldots, \vet{q}_j] \in
2694:  \mathbb{C}^{2N^2 \times j}
2695: \end{equation}
2696: %
2697: which fulfills $\vet{Q}_j^{\dag}\vet{Q}_j = \vet{I}$, $\vet{I} \in
2698: \mathbb{C}^{j \times j}$ being the identity matrix, since the
2699: basis is orthonormal. The method proceeds by looking for the best
2700: approximation of the null vector for the Liouvillian within the
2701: Krylov space $\mathcal{K}_j(\vet{L},\vet{x}_0)$. We call this
2702: vector $\vet{x}_j$. In terms of its $j$ components in the
2703: orthonormal basis $\vet{x}_j = \vet{Q}_j\vet{v}_j$. The $j$
2704: coordinates in the Krylov space $\vet{v}_j$ solve the minimum
2705: problem:
2706: 
2707: \be
2708:  \min_{\scriptsize \begin{array}{c}
2709:    \vet{x}\in \mathcal{K}_j \\
2710:    \|x\|_2 = 1 \\
2711:  \end{array} }  \!\!\! \|\vet{Lx}\|_2 = \|\vet{Lx}_j\|_2 =
2712:  \|\vet{LQ}_j\vet{v}_j\|_2
2713: \end{equation}
2714: %
2715: In the process of finding these coordinates a key r\^ole is played
2716: by the Hessenberg matrix since the following relation holds:
2717: 
2718: \be
2719:  \vet{LQ}_j = \vet{Q}_{j+1}\vet{H}_j
2720: \label{eq:Arnoldikernel}
2721: \end{equation}
2722: %
2723: We refer to the notes by Eirola \cite{eirola} for a rigorous
2724: mathematical proof of the relation (\ref{eq:Arnoldikernel}) and we
2725: only limit ourselves to exploit here some of its consequences:
2726: 
2727: 
2728: \be \bs
2729:   \min_{\scriptsize \begin{array}{c}
2730:    \vet{x}\in \mathcal{K}_j \\
2731:    \|\vet{x}\|_2 = 1 \\
2732:  \end{array} }  \!\!\!  \|\vet{Lx}\|_2
2733:  &= \|\vet{LQ}_j\vet{v}_j\|_2
2734:   = \|\vet{Q}_{j+1}\vet{H}_j\vet{v}_j\|_2\\
2735:  &=\sqrt{(\vet{Q}_{j+1}\vet{H}_j\vet{v}_j)^{\dag}\cdot
2736:          (\vet{Q}_{j+1}\vet{H}_j\vet{v}_j)}\\
2737:  &=\sqrt{(\vet{H}_j\vet{v}_j)^{\dag}
2738:          (\vet{Q}_{j+1}^{\dag}\vet{Q}_{j+1})
2739:          (\vet{H}_j\vet{v}_j)}\\
2740:  &=\sqrt{(\vet{H}_j\vet{v}_j)^{\dag}\vet{I}
2741:          (\vet{H}_j\vet{v}_j)} = \|\vet{H}_j\vet{v}_j\|_2
2742:          =  \min_{\scriptsize \begin{array}{c}
2743:    \vet{v}\in \mathbb{C}^j \\
2744:    \|\vet{v}\|_2 = 1 \\
2745:  \end{array} }  \!\!\! \|\vet{H}_j\vet{v}\|_2
2746: \end{split}
2747: \end{equation}
2748: 
2749: We started with a minimum problem involving the vector $\vet{x}$
2750: of length $2N^2$ and a matrix $\vet{L}$ of size $2N^2 \times 2N^2$
2751: and we have reduced it to the minimum problem in the last line
2752: that only involves a vector $\vet{v}$ of length $j$ and a matrix
2753: $\vet{H}_j$ of size $(j+1) \times j$. Since $j=20$ the latter is
2754: absolutely not demanding neither from a computational or a memory
2755: point of view on a modern computer. We proceed to the minimization
2756: using singular value decomposition (SVD) \cite{eirola}. Given a
2757: complex matrix $\vet{A} \in \mathbb{C}^{m \times n}$ with $m \geq
2758: n$, there exist two unitary matrices $\vet{U} \in \mathbb{C}^{m
2759: \times m}$ and $\vet{V}^{n \times n}$ such that $\vet{A} =
2760: \vet{U\Sigma V}^{\dag}$, and $\vet{\Sigma} \in \mathbb{R}^{m
2761: \times n}$ is diagonal with non-negative diagonal elements
2762: (conventionally in decreasing order starting with the highest
2763: singular value in the upper corner \cite{golub}). Thus for the
2764: Hessenberg matrix $\vet{H}_j$
2765: %
2766: \be
2767:  \vet{H}_j = \vet{U\Sigma V^{\dagger}}
2768: \end{equation}
2769: %
2770: where $\vet{U} \in \mathbb{C}^{(j+1) \times (j+1)}$, $\vet{V} \in
2771: \mathbb{C}^{j \times j}$ and
2772: 
2773: \be
2774:  \vet{\Sigma} =
2775: \left[
2776: \begin{array}{ccc}
2777: \l_1 & \cdots & 0 \\
2778: \vdots & \ddots & \vdots \\
2779: 0 & \cdots & \l_j \\
2780: 0 & \cdots & 0
2781: \end{array}
2782: \right] \in \mathbb{R}^{(j+1) \times j}
2783: \end{equation}
2784: %
2785: contains the singular values $\l_j \geq 0$. The minimum problem is
2786: solved as follows:
2787: 
2788: \be
2789:   \min_{\scriptsize \begin{array}{c}
2790:    \vet{v}\in \mathbb{C}^j \\
2791:    \|\vet{v}\|_2 = 1 \\
2792:  \end{array} }  \!\!\!  \|\vet{H}_j\vet{v}\|_2 =
2793:   \min_{\scriptsize \begin{array}{c}
2794:    \vet{v}\in \mathbb{C}^j \\
2795:    \|\vet{v}\|_2 = 1 \\
2796:  \end{array} }  \!\!\!  \|\vet{U \Sigma V}^{\dag}\vet{v}\|_2 =
2797:   \min_{\scriptsize \begin{array}{c}
2798:    \vet{v}\in \mathbb{C}^j \\
2799:    \|\vet{v}\|_2 = 1 \\
2800:  \end{array} }  \!\!\!  \|\vet{\Sigma V}^{\dag}
2801:  \vet{v}\|_2= \l_j
2802: \end{equation}
2803: %
2804: where we have used the unitarity  of $\vet{U}$ and $\vet{V}$ and
2805: we have chosen the minimizing vector $\vet{v}_j$ to be the column
2806: of $\vet{V}$ corresponding to the smallest singular value. Having
2807: found $\vet{v}_j$ the best approximation of the null vector within
2808: the Krylov space can be calculated $\vet{x}_j = \vet{Q}_j
2809: \vet{v}_j$. Finally we test the result and compare
2810: $\|\vet{L}\vet{x}_j\|_2$ with a given tolerance. If the test is
2811: positive we accept the result and reshape the vector $\vet{x}_j$
2812: into the matrix form as an approximation of the stationary
2813: solution of the GME. Otherwise we use $\vet{x}_j$ as the starting
2814: vector for a new iteration of the Arnoldi scheme. We have chosen
2815: as tolerance parameter $10\eps\|\vet{L}\|_2$ where $\eps$ is the
2816: machine precision and the norm\footnote{We used for the
2817: Liouvillean the norm $\|\vet{L}\|_2 \equiv \ \max_{\|\vet{x}\|_2 =
2818: 1}\frac{\|\vet{Lx}\|_2}{\|\vet{x}\|_2}$.} of the Liouvillean was
2819: estimated by T.~Eirola as $\|\vet{L}\|_2 \approx \exp(N/\log N)$.
2820: 
2821: \subsection{Preconditioning}
2822: 
2823: The Arnoldi scheme is iterative and can suffer from convergence
2824: problems. It is not a priori clear how many iterations one needs
2825: to converge and fulfill the convergence criterion. A possible
2826: answer to a non-convergent code is to reformulate the problem into
2827: an equivalent and (hopefully) convergent form. This was exactly
2828: the situation we had to face with the Arnoldi scheme applied to
2829: the problem of finding the stationary reduced density matrix for
2830: shuttle devices. The solution to this problem came again from
2831: prof. T.~Eirola in the form of a preconditioner. The basic idea is
2832: to find a regular operator\footnote{To be precise the
2833: preconditioner should be regular only on the image of the
2834: Liouvillian.} $\mathcal{M}$ on the Liouville space, invertible,
2835: easy to implement, such that the original problem
2836: $\mathcal{L}[\s^{stat}] = 0$ can be recast into the form:
2837: %
2838: \be
2839:  \mathcal{M}[\mathcal{L}[\s^{stat}]] = 0
2840: \end{equation}
2841: %
2842: and that the finite version of the operator
2843: $\mathcal{M}\mathcal{L}$ gives rise to a (fast) convergent
2844: iteration scheme. The operator $\mathcal{M}$ is also known as the
2845: \emph{preconditioner}.
2846: 
2847: The Arnoldi scheme is particularly efficient in finding the best
2848: approximation of the eigenvalues and corresponding eigenvectors
2849: for those eigenvalues that are separated from the rest of the
2850: spectrum. Since we want to calculate the null vector it is
2851: important that the preconditioner  moves the non-vanishing part of
2852: the spectrum far from the origin\footnote{In this sense the
2853: philosophy is: ``Invert as much as you can!''}.
2854: 
2855: In order to understand  the preconditioner that we used for the
2856: shuttling problem, we introduce the generic Sylvester operator
2857: $\phi : \mathbb{C}^{n \times m} \to \mathbb{C}^{n \times m}$:
2858: 
2859: \be
2860:  \phi(\vet{X})= \vet{AX - XB}
2861: \end{equation}
2862: %
2863: where $\vet{A} \in \mathbb{C}^{n \times n}$, $\vet{B} \in
2864: \mathbb{C}^{m \times m}$, $\vet{X} \in \mathbb{C}^{n \times m}$.
2865: This operator is invertible if and only if the spectrua of
2866: $\vet{A}$ and $\vet{B}$ have empty intersection and the inversion
2867: routine is computationally light. A part of the Liouvillean
2868: operator is of Sylvester form. In the SDQS for example:
2869: 
2870: \be
2871:  \mathcal{L}[\s] = \mathcal{L}_{\rm Sylv}[\s] +
2872:  \mathcal{L}_{\rm rest}[\s]
2873: \end{equation}
2874: %
2875: where
2876: %
2877: \be
2878:  \mathcal{L}_{\rm Sylv} = \vet{A}\s + \s\vet{A}^{\dag} =
2879:  \left[
2880:  \begin{array}{cc}
2881: A_{00} \s_{00} + \s_{00}A_{00}^{\dag} & 0\\
2882: 0 & A_{11} \s_{11} + \s_{11}A_{11}^{\dag}
2883:  \end{array}
2884:  \right]
2885: \end{equation}
2886: %
2887: and
2888: %
2889: \be
2890:  \bs
2891:   A_{00} &= -\frac{i}{\hbar}H_{\rm osc} -
2892:   \frac{\G_L}{2}e^{-\frac{2x}{\l}}-  \frac{i \g}{2 \hbar} xp -
2893:   \frac{\g m \w}{\hbar}\left(n_B+ \frac{1}{2}\right)x^2\\
2894:   A_{11} &= -\frac{i}{\hbar}(H_{\rm osc} - e \mathcal{E}x)
2895:   -\frac{\G_R}{2}e^{\frac{2x}{\l}} -  \frac{i \g}{2 \hbar} xp -
2896:   \frac{\g m \w}{\hbar}\left(n_B+ \frac{1}{2}\right)x^2\\
2897:  \end{split}
2898: \end{equation}
2899: %
2900: where $n_B$ is the average occupation number of the energy states
2901: of the harmonic oscillator in equilibrium with the bath. The rest
2902: of the Liouvillean in the non-Sylvester form reads:
2903: 
2904: \be
2905:  \mathcal{L}_{\rm rest}
2906:  \left[\begin{array}{c}
2907:  \s_{00}\\
2908:  \s_{11}
2909:  \end{array} \right]
2910:  =
2911:  \left[ \begin{array}{c}
2912:  \G_R e^{\frac{x}{\l}}\s_{11}e^{\frac{x}{\l}}
2913:  -\frac{i\g}{2\hbar}(x\s_{00}p - p\s_{00}x)
2914:  +\frac{\g m \w}{\hbar}(2n_B + 1) x\s_{00}x\\
2915:  \G_L e^{-\frac{x}{\l}}\s_{00}e^{-\frac{x}{\l}}
2916:  -\frac{i\g}{2\hbar}(x\s_{11}p - p\s_{11}x)
2917:  +\frac{\g m \w}{\hbar}(2n_B + 1) x\s_{11}x\\
2918:  \end{array}\right]
2919: \end{equation}
2920: 
2921: We use as the preconditioner for the Arnoldi iteration scheme
2922: $\mathcal{M} = \mathcal{L}_{\rm Sylv}^{-1}$. The check of empty
2923: intersection between the spectrum of $\vet{A}$ and
2924: $-\vet{A}^{\dag}$ must be done numerically due to the presence of
2925: the exponentials of position operators.
2926: 
2927: The stationary solution of the GME represents the starting point
2928: for the analysis of the properties of shuttle devices that we will
2929: present in the following chapters. The Arnoldi scheme allows the
2930: calculation of the stationary solution $\s^{stat}$ of the GME for
2931: a given set of parameters and a fixed dimension $N$ of the Hilbert
2932: space of the oscillator. The number $N$ has been chosen from
2933: considerations on the phase space distribution for the stationary
2934: solution\footnote{See next chapter for details on the Wigner
2935: distribution and its meaning.}.
2936: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2937: % Figure
2938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2939: \begin{figure}
2940:  \begin{center}
2941:   \includegraphics[angle=0,width=\textwidth]{Figures/Lmatrices.eps}
2942:   \caption{\small  \textit{Example of matrix representation of the
2943:   Liouvillean for the TDQS. The Hilbert space of the harmonic
2944:   oscillator has dimension respectively $N = 2,5,10$ ($nz$ is the number of no-zero elements
2945:   of the mtrix).
2946:    The non-zero elements are represented by a blue
2947:    dot. The exponentials of position operator present
2948:    in the tunneling amplitudes make the Liouvillean
2949:    non-sparse (large squares progressively emerging
2950:    in the figures from left to right). For large $N$ the SVD is not
2951:    viable and we use the Arnoldi iteration scheme to find the null
2952:     space of the Liouvillean.}}
2953:  \end{center}
2954: \end{figure}
2955: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2956: 
2957: 
2958: 
2959: \clearpage{\pagestyle{empty}\cleardoublepage}
2960: