1: \chapter{Linear open quantum systems}
2:
3: \index{Open quantum system}
4:
5: A relativistic particle propagating in a background can be studied from the perspective of the theory of open quantum systems \cite{Davies,BreuerPetruccione,GardinerZoller}. The mode corresponding to the momentum of the relativistic particle constitutes the open quantum system, whereas the environment is formed by the rest of the closed quantum system, namely, the other modes of the particle field and any other fields in interaction. The dynamics of the relativistic mode can be fully characterized by the reduced density matrix.
6:
7: In the next chapter we will pursue this point of view by showing that the dynamics of a given particle mode can, under certain assumptions, be treated as if it were an open quantum system interacting linearly with some environment. It is therefore of interest to begin this thesis by studying the theory of linear open quantum systems.
8:
9: \index{Quantum Brownian motion}
10: \index{QBM models}
11: \index{Caldeira-Leggett model|see{QBM models}}
12:
13: The paradigm of linear open quantum system is the quantum Brownian motion (QBM) model,
14: whose system of interest is a non-relativistic massive particle interacting linearly with an infinite bath of
15: harmonic oscillators. This model has had many applications in different contexts, among which one may mention the quantum to classical transition \cite{UnruhZurek89,RomeroPaz97}, the escape from a potential well \cite{CaldeiraLeggett81,CaldeiraLeggett83a,ArteagaEtAl03,CalzettaVerdaguer06}, the Unruh effect \cite{MassarParentaniBrout93} or quantum optics \cite{WallsMilburn,GardinerZoller}. In an influential paper Caldeira and Leggett
16: \cite{CaldeiraLeggett83b} applied the influence functional model
17: of Feynman and Vernon \cite{FeynmanVernon63,FeynmanQMPI} to the QBM model and
18: computed in closed analytic form the propagator for the reduced density
19: matrix.
20:
21: Aside from the main motivation for studying linear open quantum systems expressed above, there are additional reasons that make worth studying a QBM model. First, many of the concepts and techniques that apply to a mode by mode description of field theory can already be discussed within a simpler quantum mechanical system, and the field theory language can be compared to that of open quantum systems. Second, the system considered is linear and can be solved in a closed analytic form, so that, on the one hand, exact and perturbative solutions can be shown to match, and, on the other hand, properties related to the linear character of the system can be discussed. Third, the QBM model can be generalized to encompass the general class of linear open quantum systems. Finally, the QBM system clearly illustrates the need for a closed time path approach.
22:
23:
24:
25: % The QBM we will consider is
26: % that of the quantum oscillator linearly coupled to a
27: % one-dimensional massless scalar field, which represents an
28: % infinite set of harmonic oscillators. This model was used by Zurek
29: % and Unruh \cite{UnruhZurek89} to study the interface between the
30: % quantum and the classical limit an loss of quantum coherence. The
31: % same model was used by Massar \emph{et al.}\
32: % \cite{MassarParentaniBrout93} to represent a point-like oscillator
33: % linearly coupled to a radiation field to study the Unruh effect,
34: % \ie, the quantum state for a uniformly accelerated observer.
35: % This model has been also extended to
36: % semiclassical gravity
37: % \cite{CalzettaHu94,HuMatacz95,HuSinha95,CamposVerdaguer96,CalzettaEtAl97,MartinVerdaguer99a,MartinVerdaguer99c,MartinVerdaguer00,HuVerdaguer03,BarrabesFrolovParentani00}.
38:
39:
40: It should be mentioned that we will not provide a full account of the theory of linear open quantum systems. Although the basic formalism will be presented in order to make the presentation as self-contained as possible, only those aspects relevant for the rest of the thesis will be discussed in detail. The reader interested in a more detailed presentation can check refs.~\cite{Davies,BreuerPetruccione,GardinerZoller,CaldeiraLeggett83a,CalzettaRouraVerdaguer03,RouraThesis,Weiss}. In particular, we shall not discuss the important aspects of decoherence and time evolution of the reduced density matrix.
41:
42:
43:
44:
45:
46: \section{A quantum Brownian motion model}
47:
48: We shall consider an open quantum system composed of a harmonic
49: oscillator $q(t)$, which will be the subsystem under study,
50: linearly coupled to a free massless field $\varphi(t,x)$, which will
51: act as environment or reservoir. The action for the full system
52: can be decomposed as
53: \begin{subequations}
54: \begin{equation}
55: S[q,\varphi] = S_\mathrm{sys}[q] + S_\mathrm{env}[\varphi] + S_{\text{int}}[q,\varphi], \label{S}
56: \end{equation}
57: where the terms on the right-hand side, which correspond
58: respectively to the action of the harmonic oscillator, the action
59: of the scalar field and the interaction term are given by
60: \begin{align}
61: S_\mathrm{sys}[q] & = \int \ud{t} \left[ \fud \dot q^2 - \fud \omega_0^2 q^2 \right], \\
62: S_\mathrm{env}[\varphi] & = \int \ud{t} \ud{x} \left[ \fud(\partial_t
63: \varphi)^2-\fud(\partial_x \varphi)^2 \right], \\
64: S_{\text{int}}[q,\varphi] & = g \int \ud{t} \ud{x} \delta(x) \dot q
65: \varphi, \label{SInt}
66: \end{align}
67: \end{subequations}
68: with $\omega_0$ being the bare frequency of the
69: harmonic oscillator and $g$ being the coupling
70: constant. The oscillator is taken to have unit mass.
71: % The derivative coupling between the oscillator and the
72: % field ensures a friction term proportional to the velocity, as it
73: % will be shown later on.
74:
75: We use a one-dimensional free field as the environment, following the treatment of ref.~\cite{UnruhZurek89}. This is equivalent to the alternative representation
76: \cite{CaldeiraLeggett83b} in which the environment is modelled by a
77: large ensamble of harmonic oscillators. This
78: equivalence can be seen performing a mode decomposition in the
79: interaction term \eqref{SInt},
80: \begin{equation}
81: S_{\text{int}}[q,\varphi] = \sqrt{L} \int \ud t \frac{\vd p}{2\pi} g \dot q
82: \varphi_p,
83: \end{equation}
84: where $\varphi_p(t)$ is proportional to the spatial Fourier transform of the
85: scalar field,
86: \begin{equation*}
87: \varphi_p(t) = \frac{1}{\sqrt{L}} \int \ud x \expp{- i p x} \varphi(t,x),
88: \end{equation*}
89: where $L$ is the length of the real axis (formally infinite).
90:
91: \index{Distribution of frequencies}
92: \index{Distribution of frequencies!ohmic}
93: The model can be generalized by replacing the delta interaction of
94: equation \eqref{SInt} by a function $f(x)$. In this case the interaction term is,
95: \begin{subequations}
96: \begin{equation}
97: S_\text{int}[q,\varphi] = \int \ud t \ud x f(x)g \dot q(t) \varphi(t,x).
98: \end{equation}
99: or equivalently in the Fourier space,
100: \begin{equation}
101: S_\text{int}[q,\varphi] = \sqrt{L} \int \ud t \frac{\ud p}{2\pi} \tilde f(-p) g \dot{q}(t)
102: \varphi_p(t).
103: \end{equation}
104: \end{subequations}
105: We shall see that, when working in Fourier space, one can go from the standard to the generalized model by making the replacement
106: \[
107: \frac{\ud p}{2\pi} \ \to \ \frac{\mathcal I(p) \ud p }{2\pi},
108: \]
109: where $\mathcal I(p) := \tilde f(p) \tilde f(-p)$, whenever there is a momentum
110: integration over the environment field. The real even function $\mathcal I(p)$ will be called the \emph{distribution of frequencies}.\footnote{In the literature the distribution of frequencies is frequently defined as $\omega\mathcal I(p)$.} The product $g^2 \mathcal I(p)$ characterizes the properties of the coupling with the environment. The QBM model this way generalized encompasses the entire class of linearly coupled environments. Following the literature, the standard distribution of frequencies $\mathcal I(p)=1$ will be hereafter referred as the ``ohmic'' environment model.
111:
112: %\section{Open quantum system description} \label{sect:QBMdescr}
113:
114: \section{The influence functional}
115: \index{Influence functional}
116: \index{Influence action}
117: \index{Density matrix!reduced}
118:
119: The \emph{reduced density matrix} for an open quantum system is defined
120: from the density matrix $\hat\rho$ of the whole system by tracing out
121: the environment degrees of freedom:
122: \begin{equation}
123: \hat\rho_\text{s}(t) = \Tr_\text{env} \hat\rho(t),
124: \end{equation}
125: or equivalently, in a coordinate representation,
126: \begin{equation}
127: \rho _\text{s}(q_\mathrm f,q_\mathrm f^{\prime },t_\mathrm f)=\int \widetilde{\mathrm d}
128: \varphi \, \rho (q_\mathrm f,[\varphi],q_\mathrm f^{\prime
129: },[\varphi],t_\mathrm f)\text{.}
130: \end{equation}
131: The functional measure $\widetilde{\mathrm d} \varphi$ goes over all field configurations at a
132: given time $t_\mathrm f$. When the system and the environment are
133: initially uncorrelated, {\em i.e.}, when the initial density
134: matrix
135: factorizes ---$\hat{\rho}(t_\mathrm i)=\hat{\rho}_\mathrm{s}(t_\mathrm i)\otimes \hat{\rho}%
136: _\mathrm{e}(t_\mathrm i)$, where $\hat{\rho}_\text{s}(t_\mathrm i)$
137: and $\hat{\rho}_\text{e}(t_\mathrm i)$ mean, respectively, the density
138: matrix operators of the system and the environment at the initial
139: time---, the evolution for the reduced density matrix, which is in
140: general nonunitary and even non-Markovian, can be written as
141: \begin{equation}
142: \rho _\text{s}(q_\mathrm f,q_\mathrm f^{\prime },t_\mathrm f)=\int \ud{q_\mathrm i}\ud{q_\mathrm i^{\prime
143: }}J(q_\mathrm f,q_\mathrm f^{\prime },t_\mathrm f;q_\mathrm i,q_\mathrm i^{\prime },t_\mathrm i)\rho
144: _{r}(q_\mathrm i,q_\mathrm i^{\prime },t_\mathrm i)\text{,}
145: \end{equation}
146: where the propagator $J$ is defined in a path integral
147: representation by
148: \begin{equation}
149: J(q_\mathrm f,q_\mathrm f^{\prime },t_\mathrm f;q_\mathrm i,q_\mathrm i^{\prime
150: },t_\mathrm i)=\int\limits_{q(t_\mathrm i)=q_\mathrm i}^{q(t_\mathrm f)=q_\mathrm f}{\cal D}%
151: q\int\limits_{q^{\prime }(t_\mathrm i)=q_\mathrm i^{\prime }}^{q^{\prime
152: }(t_\mathrm f)=q_\mathrm f^{\prime }}{\cal D}q^{\prime } \expp{i(S[q]-S[q^{\prime
153: }]+S_{\mathrm{IF}}[q,q^{\prime }]) }\text{,}
154: \end{equation}
155: with $S_{\mathrm{IF}}[q,q^{\prime }]$ being the \emph{influence action},
156: which is related to the the influence functional introduced by Feynman and
157: Vernon \cite{FeynmanVernon63,FeynmanQMPI} through $F[q,q^{\prime
158: }]=\expp {iS_{\mathrm{IF}}[q,q^{\prime }]}$. In turn, the
159: influence functional can be expressed in the following way:
160: \begin{equation} \label{FInfl}
161: \begin{split}
162: F[q,q^{\prime }] = \iint & \D \varphi \, \D \varphi'
163: \rho_\mathrm{e}([\varphi_\mathrm i],[\varphi'_\mathrm i],t_\mathrm i) \\
164: & \times \exp{\left[ i \left(
165: S[\varphi] - S[\varphi'] + S_{\text{int}}[q,\varphi] -
166: S_{\text{int}}[q',\varphi'] \right) \right]}.
167: \end{split}
168: \end{equation}
169: The path integral has the boundary conditions
170: $\varphi(x,t_\mathrm i)=\varphi_\mathrm i(x)$, $\varphi'(x,t_\mathrm
171: i)=\varphi'_\mathrm i(x)$, $\varphi(x,t_\mathrm f)=\varphi'(x,t_\mathrm
172: f)=\varphi_\mathrm f(x)$; there is also an implicit sum over initial
173: and final states, $\varphi_\mathrm i(x)$, $\varphi_\mathrm i'(x)$ and
174: $\varphi_\mathrm f(x)$.
175: The influence functional can also be expressed in operator language \cite{RouraThesis}:
176: \begin{equation}
177: F[q,q'] = \Tr_\text{env} \left( \hat \rho\, U^\dag(\tf,\ti;[ q]) U(\tf,\ti;[q'])\right),
178: \end{equation}
179: where $U(\tf,\ti,[q])$ is the time evolution operator for the environment with $q(t)$ regarded as an external classical source.
180:
181: \index{Preparation functions} \label{par:InitialConditions}
182: Considering a factorized initial state is a rather unphysical hypothesis that leads to suprising results in many circumstances (see for instance sect.~\ref{sect:QBMdynamics} and ref. \cite{HuPazZhang92}). The methods presented in this chapter can be generalized to more natural initial density matrices by the use of the so-called preparation functions \cite{Weiss,GrabertEtAl88}. However the preparation function method does not completely solve all the problems because it is based in a sudden change of the density matrix. A more physical approach involves a continous preparation of the system \cite{AnglinPazZurek96}. In any case, these techniques are increasingly more involved, and we shall be mostly interested in studying the situation in which initial conditions are set in the remote past. In this case the system and environment have had enough time to interact and become entangled, and the precise form of the initial state becomes unimporant.
183:
184: \index{Gaussian state}
185: Besides demanding that the total state is a factorized product state, we shall also require stationary and isotropic Gaussian states for the environment. Gaussian states are analyzed in appendix \ref{app:Gaussian}. Since the environment field is one-dimensional, the requeriment of isotropy simply translates in the equivalence of the positive and negative field modes. The general class of linear systems can be characterized by isotropic environment states; there is no loss of generality in this hypothesis.
186:
187: When the initial density matrix of the environment $\rho
188: _\mathrm{env}([\varphi_\mathrm i],[\varphi_\mathrm i'],t_\mathrm i)$ is
189: Gaussian, the path integrals can be exactly performed and one
190: obtains \cite{FeynmanVernon63,CaldeiraLeggett83a,RouraThesis}:
191: \begin{equation}
192: \begin{split}
193: S_{\mathrm{IF}}[q,q^{\prime}]
194: =&- 2\int_{t_\mathrm i}^{t_\mathrm f} \ud{t} \int_{t_\mathrm i}^{t} \ud{t^{\prime
195: }}\Dot \Delta (t) \mathcal D(t,t^{\prime })\dot Q(t^{\prime }) \\ & + \frac{i}{2}\int_{t_\mathrm i}^{t_\mathrm f}\ud{t}%
196: \int_{t_\mathrm i}^{t_\mathrm f} \ud{t^{\prime }}\Dot \Delta (t) \mathcal N(t,t^{\prime })\Dot \Delta
197: (t^{\prime })\text{,}
198: \end{split}
199: \end{equation}
200: where $\Delta(t) := q(t)- q'(t)$ and $Q(t) := [ q(t)+
201: q'(t)]/2$.
202: \index{Dissipation kernel}
203: \index{Noise kernel}
204: In the ohmic case the kernels $\mathcal D(t,t')$ and $\mathcal N(t,t')$, which are related to the
205: dissipation and noise kernels (defined below), are given by
206: \cite{CalzettaRouraVerdaguer03}:
207: \begin{subequations} \
208: \begin{align}
209: \mathcal D(t,t')&=\frac{ig^2}{2} \av{ [\hat \varphi_\mathrm I (t,0), \hat
210: \varphi_\mathrm I (t',0)] } = \frac{ig^2}{2} \int \frac{\vd p}{2\pi} \av{ [\hat \varphi_{\mathrm I(-p)} (t), \hat
211: \varphi_{\mathrm Ip} (t')] }
212: ,\\
213: \mathcal N(t,t')&=\frac{g^2}{2} \av { \{\hat \varphi_\mathrm I (t,0), \hat \varphi_\mathrm I (t',0)
214: \}} = \frac{g^2}{2} \int \frac{\vd p}{2\pi} \av{ \{\hat \varphi_{\mathrm I(-p)} (t), \hat
215: \varphi_{\mathrm Ip} (t')\} } ,
216: \end{align}
217: \end{subequations}
218: where $\hat \varphi_\mathrm I (x,t)$ is the field operator in the
219: interaction picture, $\hat \varphi_{\mathrm Ip} (t)$ is the $p$-mode of the same field operator, and we recall that $\av{\cdot} = \Tr( \hat \rho \ \cdot )$
220: means quantum mechanical average. In the generic case the kernels can be expressed as
221: \begin{subequations} \label{mathcals}
222: \begin{align}
223: \mathcal D(t,t')&= \frac{ig^2}{2} \int \frac{\vd p}{2\pi} \mathcal I(p) \av{ [\hat \varphi_{\mathrm I(-p)} (t), \hat
224: \varphi_{\mathrm Ip} (t')] }
225: , \label{mathcalD} \\
226: \mathcal N(t,t')&= \frac{g^2}{2} \int \frac{\vd p}{2\pi} \mathcal I(p) \av{ \{\hat \varphi_{\mathrm I(-p)} (t), \hat
227: \varphi_{\mathrm Ip} (t')\} } . \label{mathcalN}
228: \end{align}
229: \end{subequations}
230: It is convenient to interpret
231: these kernels as bidistributions or generalized
232: functions of two variables.
233:
234:
235:
236: \index{Dissipation kernel}
237: \index{Noise kernel}
238: By integration by parts, the influence action can also be
239: expressed as
240: \begin{equation} \label{S_IF2}
241: \begin{split}
242: S_{\mathrm{IF}}[q,q^{\prime}]=
243: & \int_{t_\mathrm i}^{t_\mathrm f} \ud{t} \int_{t_\mathrm i}^{t_\mathrm f} \ud{t'} \Delta(t) H(t,t') Q(t') \\
244: & + \frac{i}{2} \int_{t_\mathrm i}^{t_\mathrm f} \ud{t} \int_{t_\mathrm i}^{t_\mathrm f} \ud{t'} \Delta(t) N(t,t')
245: \Delta(t'),
246: \end{split}
247: \end{equation}
248: or as
249: \begin{equation} \label{S_IF3}
250: \begin{split}
251: S_{\mathrm{IF}}[q,q^{\prime}]
252: =&- 2\int_{t_\mathrm i}^{t_\mathrm f} \ud{t} \int_{t_\mathrm i}^{t} \ud{t^{\prime
253: }}\Delta (t) D(t,t^{\prime }) Q(t^{\prime }) + \int_{t_i}^{t_f} \ud{t} \delta\omega^2_0 \Delta(t)Q(t) \\ & + \frac{i}{2}\int_{t_\mathrm i}^{t_\mathrm f}\ud{t}%
254: \int_{t_\mathrm i}^{t_\mathrm f} \ud{t^{\prime }} \Delta (t) N(t,t^{\prime }) \Delta
255: (t^{\prime })\text{,}
256: \end{split}
257: \end{equation}
258: where the different kernels are defined as
259: \begin{subequations} \label{HDN}
260: \begin{align}
261: H(t,t') &:=-2 \derp{}{t} \derp{}{t'} [\theta(t-t') \mathcal D(t,t')] \label{kernelH1} \\ &= -2\theta(t-t') D(t-t') + \delta \omega_0^2 \delta(t-t') \label{kernelH},
262: \\
263: D(t,t') &:= \derp{}{t} \derp{}{t'} \mathcal D(t,t') \label{kernelD},\\
264: N(t,t') &:= \derp{}{t} \derp{}{t'} \mathcal N(t,t') \label{kernelN}.
265: \end{align}
266: The kernels $D(t,t')$ and $N(t,t')$ are called respectively
267: \emph{dissipation} and \emph{noise kernels}. The frequency shift $\delta\omega^2_0$ is a formally divergent quantity given by
268: \begin{equation}\label{FreqShift}
269: \delta \omega_0^2 := 2 \lim_{t\to t'} \derp{\mathcal D(t,t')}{t} .
270: \end{equation}
271: \end{subequations}
272:
273: There is a subtle technical point involving the frequency shift and the dissipation kernel. The kernel $H(t,t')$ involves the product of two distributions [see eqs.~\eqref{kernelH1} and \eqref{kernelH}] and such a product is strictly speaking not well-defined. In practice this means that one has to be very careful when considering contributions coming from the coincidence limit of the dissipation kernel. As we have seen, depending on the way calculations are done the divergent frequency shift $\delta\omega_0^2$ may show up or not. We shall see that, at least with the ohmic environment, the final results are well-defined and
274: ultraviolet-finite.\footnote{\emph{Note added in the second printing}. We have recently become aware that QBM models are affected by logarithmic ultraviolet divergences, other than the associated to the initial state preparation \cite{FlemingEtAl07}. In any case, the results presented here are most likely not affected.}
275: This is in contrast with the original presentation of the QBM model in terms of a collection of harmonic oscillators \cite{CaldeiraLeggett83b}, in which the frequency shift $\delta\omega^2_0$ has to be introduced in the original action in order to get finite results. (Except for this point, both treatments of the QBM model are completely equivalent.) In any case it is worth recalling that, as a matter of principle, the physical frequency of the interacting system needs not coincide with $\omega_0$.%, and that a frequency shift can always be introduced.%Therefore it will also prove useful to us to consider that the frequency shift $\delta\omega_0$
276:
277:
278: % this translates into the fact that two formally equivalent expressions for the kernel may differ in a constant contribution related to the coincidence limit of the kernel. This constant contribution turns out to be divergent, and coinciding precisely with the value of $\delta\omega_0$. All the expressions written above are finite, at least once a proper ultraviolet regularization is considered; however, the constant divergent term will reappear when going to the Fourier space. Anyway, this constant term can always be reguralized by introducing an ultraviolet cuttoff, and renormalized away by considering a suitable counterterm to the frequency of the oscillator in the original action. This amounts to considering $\delta\omega_0$ an additional free parameter of the theory. Let us also comment that, while in our presentation of the QBM model the constant divergent can be avoided depending in the way we do the calculations, it always appears in the original representation in terms of an infinite collection of harmonic oscillators \cite{CaldeiraLeggett83b},
279:
280:
281: Let us end this section by presenting yet another alternative equivalent expression for the influence action:
282: \begin{equation}
283: \begin{split} \label{S_IF-Alt}
284: S_{\mathrm{IF}}[q,q']= \frac{ig^2}{2} \int_\ti^\tf \ud{t} \int_\ti^\tf \ud{t'}\Big[ & \dot q(t) \Delta_\mathrm F(t,t') \dot q(t')
285: + \dot q'(t) \Delta_\mathrm D(t,t') \dot q'(t') \\
286: &- \dot q(t) \Delta_{-}(t,t') \dot q'(t')
287: - \dot q'(t) \Delta_{+}(t,t') \dot q(t') \Big].
288: \end{split}
289: \end{equation}
290: In the ohmic case the different kernels are given by
291: \begin{subequations}
292: \begin{gather}
293: \Delta_\mathrm F(t,t') = \av{ T \hat \varphi_\mathrm I(t,0) \hat \varphi_\mathrm I(t',0)}, \\
294: \Delta_\mathrm D(t,t') = \av{ \widetilde T \hat \varphi_\mathrm I(t,0) \hat \varphi_\mathrm I(t',0)}, \\
295: \Delta_{-}(t,t') = \Delta_{+}(t',t) = \av{\hat \varphi_\mathrm I(t',0) \hat
296: \varphi_\mathrm I(t,0)},
297: \end{gather}
298: \end{subequations}
299: and correspond to the Feynman propagator, the Dyson propagator and
300: the Whightman functions of the environment, respectively. In the generic case the above expressions should be generalized following the lines of eqs.~\eqref{mathcals}.
301:
302:
303: \index{Dissipation kernel}
304: \index{Noise kernel}
305: \section{Dissipation and noise kernels}\label{sect:DisNoise}
306:
307: Let us explicitly compute the dissipation and noise kernels following eqs.~\eqref{mathcals} and \eqref{HDN}. Although the results are standard, at least in the ohmic case, the details of the calculation method will be useful for us afterwards. We need the expression of the one-dimensional field
308: operator in the interaction picture in terms of anihilation and creation operators,\footnote{Recall that, for an interacting theory,
309: operators in the interaction picture follow the same relations as
310: the operators of the free theory in the Heisenberg picture.}
311: \begin{equation}
312: \hat \varphi_\mathrm I(t,x) = \sqrt{L} \int \frac{\ud{p}}{2\pi} \frac
313: {1}{\sqrt{2|p|}}
314: ( \hat a_p
315: \expp{-i |p| t + i p x} + \hat a_p^\dag \expp {i |p|t-ipx}
316: ),
317: \end{equation}
318: where $\hat a_p$ and $\hat a^\dag_p$ satisfy
319: $[\hat a_p, \hat a^\dag_q] = \delta_{pq} = (2\pi/L) \delta(p-q)$, and the expression for the mode-decomposed field operator in analogous terms,
320: \begin{equation} \label{ModeDecomp}
321: \hat \varphi_{\mathrm Ip}(t) = \frac{1}{\sqrt{L}} \int \ud x \varphi_\mathrm I(t,x) \expp{-ipx} = \frac{1}{\sqrt{2|p|}} (\hat a^\dag_p + \hat a^\dag_ {-p} ).
322: \end{equation}
323:
324: Since the system is linear, the dissipation kernel \eqref{kernelD}, which corresponds to the field anticonmutator, is state independent. Let us check it explicitly. According to eqs.~\eqref{mathcalD} and \eqref{kernelD} the dissipation kernel can be computed as
325: \begin{equation}
326: D(t,t')= \frac{ig^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \mathcal I(p) \Tr_\text{env}{\Big(\hat\rho_\text{e} \big[\hat \varphi_{\text Ip}(t), \hat \varphi_{\text I(-p)}(t')\big]\Big)}.
327: \end{equation}
328: Introducing the explicit evolution operators one gets
329: \begin{equation*}
330: \begin{split}
331: D(t,t')&= \frac{ig^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \mathcal I(p) \\
332: &\quad\times\Tr_\text{env} {\Big(\hat\rho_\text{e} \expp{-i\hat H_\text{env} (t_\text i-t)} \hat\varphi_{p} \expp{-i\hat H_\text{env}(t-t')} \hat \varphi_{-p}\expp{-i\hat H_\text{env}(t'-t_\text i)}\Big)}\\
333: &- \text{($t \leftrightarrow t'$)}.
334: \end{split}
335: \end{equation*}
336: where $\hat H_\text{env}$ is the Hamiltonian operator for the environment field and the initial time $t_\text i$ is taken in the remote past (the result will be independent of the initial time). Since it is sufficient to trace over the two-mode reduced state, we introduce two resolutions of the identity in the subspace of these two modes, $1_{\pm p} = \sum_{n,n'} |n_pn'_{-p}\rangle \langle n_pn'_{-p} |$, and get
337: \begin{equation*}
338: \begin{split}
339: D(t,t')&= \frac{ig^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \mathcal I(p) \sum_{n,n',m,m'} \expp{ -i (n+n'-m-m')|p|(t-t')}\\&\qquad\times \rho_{m,m'} |\langle m_p m'_{-p}| \hat \varphi_{p} | n_p n'_{-p}\rangle |^2- \text{($t \leftrightarrow t'$)},
340: \end{split}
341: \end{equation*}
342: where $\rho_{m,m'}= \Tr_\text{other modes}\langle m_pm'_{-p} |\hat\rho_e|m_pm'_{-p}\rangle$. Since the environment is stationary the density matrix operator conmutes with the Hamiltonian and is diagonal in the basis of eigenstates of the environment Hamiltonian
343: %\footnote{One could be worried by the following subtle point. Since the environment is stationary, in principle the density matrix should conmute with the full Hamiltonian and be diagonal in the basis of eigenstates of the full Hamiltonian, and not necessarily in the basis of eigenstates of the environment Hamiltonian; both statements are not equivalent in presence of interaction. One should understand that the density matrix elements that we are quoting refer to the density matrix at remote times, when interaction was still switched off and the density matrix was simultaneously diagonal in the basis of eigenstates of the system and environment Hamiltonians. We assume that later on interaction was slowly switched in a way that all observables remained time translation-independent. Notice that an adiabatic switching of the interaction is needed in order to reconcile the assumption of factorized initial conditions with the assumption of stationarity of the environment.}
344: $|m_pm'_{p'}\rangle$. Expressing the field operator in terms of the creation and anihilation operators [see \Eqref{ModeDecomp}] the above equation can be reexpressed as
345: \begin{equation*}
346: \begin{split}
347: D(t,t')&= \frac{ig^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \frac{\mathcal I(p)}{2|p|} \\ &\qquad \times \sum_{m,m'} \left[ \rho_{m,m'}(m+1) \expp{ -i |p|(t-t')}+ \rho_{m,m'} m' \expp{ i |p|(t-t')}\right]
348: \\ &\quad - \text{($t \leftrightarrow t'$)}.
349: \end{split}
350: \end{equation*}
351: Considering the 1-mode reduced density matrix $\rho_n = \Tr_\text{other modes}\langle n_p |\hat \rho_\text{env} | n_p\rangle$, and taking into account the property of isotropy of the environment, the above equation can be simplified to
352: \begin{equation}
353: \begin{split}
354: D(t,t')&= \frac{ig^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \frac{\mathcal I(p)}{2|p|} \sum_n \rho_n \left[ (n+1) \expp{- i |p|(t-t')}+ n \expp{ i |p|(t-t')}\right]
355: \\ &\quad - \text{($t \leftrightarrow t'$)}.
356: \end{split}
357: \end{equation}
358: or, substracting the time-reversed part,
359: \begin{equation*}
360: D(t,t') = \frac{ig^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \frac{\mathcal I(p)}{2|p|} \left[ \expp{ -i |p|(t-t')}- \expp{ i |p|(t-t')}\right] \sum_n \rho_n.
361: \end{equation*}
362: %\shorfage[4]
363: Since the density matrix is normalized, $\sum_n \rho_n=1$, it can be explicitly seen that the dissipation kernel is indeed state-independent:
364: \begin{equation*}
365: D(t,t')= \frac{ig^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \frac{\mathcal I(p)}{2p} (-2i)\sin {[p(t-t')]}.
366: \end{equation*}
367: Introducing the Fourier transform we get the value of the dissipation kernel in the frequency space.
368: \begin{equation} \label{DisQBM}
369: D(\omega) = \frac{i \omega g^2}{2} \mathcal I(\omega).
370: \end{equation}
371: \pagebreak
372: The dissipation kernel is closely related to the kernel $H(\omega)$ [see eqs.~\eqref{kernelH1} and \eqref{kernelH}], which is also state-independent and given by:
373: \begin{equation}
374: H(\omega) = g^2 \int \frac{\vd p}{2\pi} \frac{\omega \mathcal I(\omega)}{\omega-\omega'+i\epsilon} + \delta \omega^2_0.
375: \end{equation}
376:
377: \index{Occupation number}
378: In contrast, the noise kernel \eqref{kernelN} is state dependent. Repeating an analogous calculation, we get the following expression for the noise kernel:
379: \begin{equation}
380: \begin{split}
381: N(t,t')&= \frac{g^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \frac{\mathcal I(p)}{2|p|} \sum_n \rho_n \left[ (n+1) \expp{ -i |p|(t-t')}+ n \expp{ i |p|(t-t')}\right]
382: \\ &\quad + \text{($t \leftrightarrow t'$)}.
383: \end{split}
384: \end{equation}
385: Adding the time-reversed part yields:
386: \begin{equation}
387: \begin{split}
388: N(t,t')&= \frac{g^2}{2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \frac{\mathcal I(p)}{2|p|} \left[ \expp{ i |p|(t-t')}+ \expp{ -i |p|(t-t')}\right] \sum_n \rho_n (2n+1).
389: \end{split}
390: \end{equation}
391: Defining the environment \emph{occupation number} as
392: \begin{equation}
393: n(|p|) := \Tr{ \left( \hat\rho_\text e \hat a^\dag_p \hat a_p \right)}=\Tr{ \left( \hat\rho_\text e \hat a^\dag_{-p} \hat a_{-p} \right)} = \sum_n \rho_n n ,
394: \end{equation}
395: the above equation can be reexpressed as
396: \begin{equation}
397: \begin{split}
398: N(t,t')&= {g^2} \partial_t \partial_{t'} \int \frac{\vd p}{2\pi} \frac{\mathcal I(p)}{2|p|} \left[ \expp{ i p(t-t')}+ \expp{ -i p(t-t')}\right] \left[ n(|p|) + \frac{1}{2} \right].
399: \end{split}
400: \end{equation}
401: For a general Gaussian stationary environments, characterized by the occupation numbers $n(|\omega|)$, the noise kernel in Fourier space is given by
402: \begin{equation}
403: N(\omega)= {g^2|\omega|}\mathcal I(\omega)\left[\frac12 + n(|\omega|)\right].
404: \end{equation}
405:
406:
407:
408: %\shortpage[4]
409: \index{Fluctuation-dissipation theorem}
410: For the particular case of an environment in thermal equilibrium at a temperature $T$ the occupation numbers are given by $n(|\omega|) = 1/(\expp{|\omega|/T}-1)$ and the noise kernel is given by
411: \begin{equation} \label{NoiseQBM}
412: N(\omega) = \frac{g^2|\omega| \mathcal I(\omega)}{2} \coth \left(\frac{|\omega|}{2 T}\right).
413: \end{equation}
414: Therefore in a thermal state the noise and dissipation kernels are related through the
415: \emph{fluctuation-dissipation theorem}, (see also
416: appendix \ref{app:GenRel}):
417: \begin{equation}\label{FluctDisTh}
418: N(\omega) = - i \sign(\omega) \coth \left(\frac{|\omega|}{2
419: T}\right)
420: D(\omega).
421: \end{equation}
422:
423: We shall next quote explicit time expressions in the case of ohmic distribution of frequencies. The dissipation kernel is given by:
424: \begin{equation}
425: D(t,t') = \frac{g^2}{4} \delta' (t-t'), \qquad H(t,t') = - \frac{g^2}{2} \delta' (t-t').
426: \end{equation}
427: The noise kernel in the vacuum is
428: \begin{subequations}
429: \begin{equation}
430: N(t,t') = \frac{g^2}{2\pi} \Pf \frac{-1}{(t-t')^2}.
431: \end{equation}
432: For thermal states, a closed analytic expression for the noise kernel in time space can only be given in the high temperature limit $T \gg \omega$:
433: \begin{equation}
434: N(t,t') \approx g^2 T \delta(t-t').
435: \end{equation}
436: \end{subequations}
437: %Again, details are given in appendix \ref{app:DissipationNoise}.
438:
439:
440: By considering an arbitrary distribution of frequencies $\mathcal I(\omega)$ and an arbitrary Gaussian state for the environment $\hat\rho_\text{e}$ the dissipation and noise kernels may adopt almost any value. In the rest of the chapter we shall try to express all results in terms of the dissipation and noise kernels. Only two restrictions will be imposed: (1) we will consider environments with an infinite continous number of degrees of freedom (\ie, avoiding singular delta-like distributions of frequencies), and (2) we will only consider stationary Gaussian states for the environment (see appendix~\ref{app:Gaussian} for a description of the Gaussian states). Under these assumptions, generic exact results can be obtained by working in the Fourier space. To this end, it will prove useful to reexpress \Eqref{kernelH} in Fourier space:
441: \begin{subequations}
442: \begin{equation}
443: H(\omega) =
444: - 2 \int \frac{\ud {\omega'}}{2\pi}
445: \frac{ i D(\omega')}
446: {\omega-\omega'+i\epsilon} + \delta\omega^2_0.
447: \end{equation}
448: The kernel $H(\omega)$ can be decomposed in its real and imaginary parts as:
449: \begin{align}
450: H_\mathrm R(\omega) &:= \Re H(\omega)= -2\PV \int \frac{\ud {\omega'}}{2\pi} \frac{
451: i D(\omega')}{\omega-\omega'}+ \delta\omega^2_0, \label{RealImOm} \\
452: H_\mathrm I(\omega) &:= \Im H(\omega) = i D(\omega).
453: \end{align}
454: \end{subequations}
455: We have used the property $1/(x+i\epsilon) = \PV(1/x) - i \pi
456: \delta(x)$. Notice that $H(-\omega)=H^*(\omega)=
457: H_\mathrm R(\omega)-iH_\mathrm I(\omega)$. The real and imaginary parts of the kernel in frequency space correspond, respectively, to the even and odd parts in time space. \index{Kramers-Kronig relation} Notice also the Kramers-Kronig relation betewen the real and imaginary parts of the kernel $H(\omega)$:
458: \begin{equation}
459: H_\mathrm R(\omega) = -2\PV \int \frac{\ud {\omega'}}{2\pi} \frac{
460: H_\mathrm I(\omega')}{\omega-\omega'}+ \delta\omega^2_0.
461: \end{equation}
462:
463:
464: We end up signaling that the noise kernel in the frequency space is positive-defined,
465: \begin{subequations}
466: \begin{equation}
467: N(\omega)\geq 0,
468: \end{equation}
469: and that the dissipation kernel is imaginary negative for positive frequencies and imaginary positive for negative frequencies,
470: \begin{equation}
471: \Im H(\omega)= iD(\omega) \leq 0 \quad \text{if } \omega > 0, \qquad \Im H (\omega)=iD(\omega) \geq 0, \quad \text{if }\omega<0.
472: \end{equation}
473: \end{subequations}
474: %These properties follow from the spectral representation of the propagators that shall be analyzed in the next chapter.
475:
476:
477:
478: \section{The closed time path generating functional}
479:
480: \index{Generating functional!in the QBM models}
481:
482: As explained in appendix~\ref{app:CTP} [see \Eqref{ZCTPOper}], the CTP generating functional can be expressed as
483: \begin{equation}
484: Z[j_1,j_2] = \Tr \left[ \hat \rho\, \widetilde T \expp {- i \int \ud{t} j_2(t)
485: \hat q(t) }T \expp {i \int \ud{t} j_1(t)
486: \hat q(t)} \right],
487: \end{equation}
488: where $\hat \rho$ is the initial density matrix for
489: the whole system and the trace is also taken over the whole
490: system. In terms of path integrals the generating functional is
491: written, after integrating out the environment,
492: \begin{equation}
493: \begin{split} \label{ZCTP2}
494: Z[j_1,j_2] = & \int \ud{q_\mathrm f} \int \ud{q_\mathrm i} \ud{q'_\mathrm i}
495: \int \limits_{q(t_\mathrm i)=q_\mathrm i}^{q(t_\mathrm f)=q_\mathrm f} \uD{q}
496: \int \limits_{q'(t_\mathrm i)=q'_\mathrm i}^{q'(t_\mathrm f)=q_\mathrm f} \uD{q'}
497: \rho_\mathrm r(q_\mathrm i,q'_\mathrm i,t_\mathrm i) \\
498: & \times \expp{ i(S[q] - S[q'] + S_{\mathrm{IF}}[q,q'])} \expp{ i\left( \int \ud{t} j_1(t) q(t) - \int \ud{t}
499: j_2(t) q'(t) \right)}.
500: \end{split}
501: \end{equation}
502: At least for the moment, we will be interested in the case in which initial
503: conditions are set in the remote past, so that we shall take the limit
504: $t_\mathrm i \to -\infty$.
505:
506: %We will assume an influence action of the form
507: %\begin{equation}
508: %\begin{split}
509: % S_{\mathrm{IF}}[q,q^{\prime}]=
510: % \int \ud{t} \ud{t'} \Delta(t) H(t,t') Q(t') + \frac{i}{2} \int \ud{t} \ud{t'} \Delta(t) N(t,t')
511: % \Delta(t').
512: %\end{split}
513: %\end{equation}
514: \index{Propagator!retarded}
515: \index{Propagator!in the QBM models}
516: Once the path integrals of
517: \Eqref{ZCTP2} are performed one can show that, for a Gaussian environment and asymptotic initial boundary
518: conditions the generating functional can be expressed as\footnote{In the general case there would be a prefactor taking into account the initial conditions for the system. However, since the system has a dissipative dynamics and the initial conditions are given in the remote past, initial conditions for the system turn out to be completely irrelevant.} \cite{CalzettaRouraVerdaguer03}
519: \begin{equation}
520: \begin{split} \label{ZCTP}
521: Z[j_1,j_2] = & \
522: \expp{ \fud \int \ud{t_1} \ud{t_2} \ud{t_3}
523: \ud{t_4}
524: j_\Delta(t_1)
525: \Gret(t_1,t_2)N(t_2,t_3)j_\Delta(t_4)\Gret(t_4,t_3) } \\
526: & \times
527: \expp{ -\int \ud{t_1} \ud{t_2} j_\Delta(t_1) \Gret(t_1,t_2)
528: j_\Sigma(t_2)},
529: \end{split}
530: \end{equation}
531: where $j_\Sigma(t) := [j_1(t)+j_2(t)]/2$, $j_\Delta(t) :=
532: j_1(t) - j_2(t)$ and $\Gret(t,t')$ is the retarded propagator of the
533: kernel
534: \begin{equation} \label{KernelL}
535: L(t,t') = \left( \dert[2]{}{t} + \omega_0^2 \right)
536: \delta(t-t') + H(t,t'),
537: \end{equation}
538: \ie, is the kernel which verifies
539: \begin{equation} \label{RetEqDif}
540: \int \ud s \Gret(t,s) L(s,t') = - i \delta(t-t'),
541: \qquad \Gret(t,t') = 0 \quad \text{if } t<t'.
542: \end{equation}
543: Thus in order to have an explicit expression for $Z_{\mathrm{CTP}}$ we only need to know the dissipation kernel (from which the retarded propagator of the kernel $L(t,t')$ can be obtained) and the noise kernel. With expressions in next section is easy to check that the retarded propagator of the kernel $L(t,t')$ also happens to be the retarded propagator of the system:
544: \begin{equation}
545: \GR(t,t') = \av{ [\hat q(t), \hat q(t')]} \theta(t-t').
546: \end{equation}
547:
548:
549: \index{Propagator!Feynman}
550: \index{Propagator!Dyson}
551: \index{Propagator!Whightman}
552: \index{Propagator!in the QBM models}
553:
554: Differentiating the CTP generating functional we obtain the
555: different correlation functions. In particular, the
556: Feynman and Dyson propagators, and the Whightman functions are given by:
557: \begin{subequations}\label{PropDivers}
558: \begin{gather}
559: G_\mathrm F(t,t') = \av{T \hat q(t) \hat q(t') } =- \derff{Z[j_1,j_2]}{j_1(t)}
560: {j_1(t')} \bigg|_{j=j_2=0}, \label{GF} \\
561: G_\mathrm D(t,t') = \av{\widetilde T \hat q(t) \hat q(t') } = - \derff{Z[j_1,j_2]}{j_2(t)}
562: {j_2(t')} \bigg|_{j=j_2=0}, \\
563: G_-(t,t') = G_+(t',t) := \av{ \hat q(t') \hat q(t) } = \derff{Z[j_1,j_2]}{j_1(t)}
564: {j_2(t')} \bigg|_{j=j_2=0}.
565: \end{gather}
566: \end{subequations}
567: The CTP generating functional can also be
568: expressed as a function of these correlators:
569: \begin{equation} \label{ZQBMDirectBasis}
570: \begin{split}
571: Z[j_1,j_2] = \exp\bigg[&- \frac{1}{2} \int \ud t \ud {t'}
572: \big[ j_1(t) G_\mathrm F(t,t') j_1(t')
573: - j_1(t) G_-(t,t') j_2(t') \\
574: & - j_2(t) G_+(t,t') j_1(t)
575: + j_2(t) G_\mathrm D(t,t') j_2(t') \big]\bigg].
576: \end{split}
577: \end{equation}
578:
579:
580:
581:
582: \section{Correlation functions}
583: \index{Propagator!in the QBM models}
584:
585: Let us start by considering the retarded propagator. According to \Eqref{RetEqDif} the computation of the retarded propagator is trivial in Fourier space:
586: \begin{equation}\label{GretGen}
587: \GR(\omega) = \frac{-i}{L(\omega)} = \frac{-i}{-\omega^2 + \omega_0^2 + H(\omega)}.
588: \end{equation}
589: The fact that the imaginary part of the kernel $H(\omega)$ is negative (positive) for positive (negative) frequencies, as stated before, implies that the above propagator is indeed retarded. For the ohmic distribution of frequencies one gets
590: \begin{subequations}
591: \begin{equation} \label{GretFourier}
592: \Gret(t,t')=-i\int \frac{ \ud{\omega} }{2\pi} \frac{1}{L(\omega)} = \int \frac{ \ud{\omega} }{2\pi}
593: \frac{i
594: \expp{-i\omega(t-t')}}{\omega^2 +i \omega g^2 /2 -
595: \omega_0^2},
596: \end{equation}
597: or, equivalently, after performing the integral,
598: \begin{equation}
599: \Gret(t,t')= \frac{-i}{\omega_1}\expp{-\gamma(t-t')/2} \sin{
600: [\omega_1(t-t')]}
601: \: \theta(t-t'), \label{Gret}
602: \end{equation}
603: \end{subequations}
604: where the coefficients $\gamma$ and $\omega_1$ are defined through
605: \begin{equation} \label{gammaomega}
606: \gamma := \frac{g^2}{2}, \qquad \omega_1 := \sqrt{ \omega_0^2 -
607: (\gamma/2)^2}.
608: \end{equation}
609:
610:
611:
612: Introducing eq.~\eqref{GretGen} into \Eqref{ZCTP}
613: the CTP generating functional of the
614: system is obtained as a function of the noise and dissipation kernels. From the generating functional the full set
615: of propagators of the system can be obtained.
616:
617: \index{Propagator!Feynman}
618: \index{Propagator!in the QBM models}
619:
620: As an example, we proceed now to the computation
621: of the Feynman propagator. Applying \Eqref{GF},
622: \begin{equation}
623: \begin{split}
624: G_\mathrm F&(t,t') = \frac{1}{2} \Gret(t,t') + \frac{1}{2}\Gret(t',t)
625: - \int \ud{s} \ud{s'}
626: \Gret(t,s) N(s,s') \Gret(t',s').
627: \end{split}
628: \end{equation}
629: This last expression is more easily computed in
630: Fourier space,
631: \begin{equation*} \label{MagicGF}
632: \begin{split}
633: G_\mathrm F(\omega) = & \frac{1}{2} \Gret(\omega)
634: +\frac{1}{2} \Gret(-\omega) - \fud \Gret(\omega) {N}(\omega)
635: \Gret(-\omega) \\
636: = & \frac{1}{2} \Gret(\omega)
637: -\frac{1}{2} \Gret^*(\omega) +\fud \Gret(\omega) {N}(\omega)
638: \Gret^*(\omega) \\
639: = & \ i \Im \Gret(\omega) + N(\omega)
640: \lvert \Gret(\omega) \rvert^2.
641: \end{split}
642: \end{equation*}
643: With the aid of \Eqref{GretGen}, the Feynman propagator reads
644: \Eqref{GretGen}:
645: \begin{equation} \label{FeynmanGenT}
646: \begin{split}
647: G_\mathrm F(\omega)
648: &= \frac{- i \left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega) \right]
649: + N(\omega)}{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2 +
650: \left[H_\mathrm I(\omega)\right]^2}.
651: \end{split}
652: \end{equation}
653: This last expression is completely general and valid for any state.
654: If we now introduce the fluctuation-dissipation at zero
655: temperature [see \Eqref{FluctDisTh}], $[ H_\mathrm I(\omega)]^2=
656: [N(\omega)]^2$, we may write
657: \begin{equation*}
658: G_\mathrm F(\omega) = i\frac{ \omega^2 - \omega_0^2 - H_\mathrm R(\omega)
659: -i N(\omega)}{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2 +
660: \left[N(\omega)\right]^2}.
661: \end{equation*}
662: Factorizing this last equation we find the final expression for
663: the Feynman propagator of the QBM system at zero temperature:
664: \begin{subequations}
665: \begin{equation} \label{FeynmanGen0}
666: G_\mathrm F(\omega) = \frac{i }{\omega^2-\omega_0^2 -H_\mathrm R(\omega)
667: +i N(\omega)}.
668: \end{equation}
669: Note that this expression is valid for any distribution of
670: frequencies in the environment. For the ohmic environment
671: one get:
672: \begin{equation}
673: G_\mathrm F(\omega) =
674: \frac{i }
675: {\omega^2 + i(g^2/2) |\omega| -\omega_0^2}.
676: \end{equation}
677: \end{subequations}
678:
679: The other correlators may similarly be computed. Either
680: we may functionally differentiate the CTP generating functional,
681: see \Eqref{PropDivers}, or we may apply the relations in appendix \ref{app:GenRel}. In this latter case it is useful to take the retarded propagator \eqref{GretGen} as one of the basic building blocks of the propagators, the other being the Hadamard function,
682: \begin{subequations} \label{HadamardQBM}
683: \begin{align}
684: G^{(1)} (t,t') &= 2\int \ud{s} \ud{s'}
685: \Gret(t,s) N(s,s') \Gret(t',s') \\
686: G^{(1)} (\omega ) &=2 |\Gret(\omega)|^2 N(\omega).
687: \end{align}
688: \end{subequations}
689: General expressions
690: valid for any state are:
691: \begin{subequations} \label{GOtherFourier}
692: \begin{align}
693: G_\mathrm D(\omega) &= \frac{ i \left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)
694: \right]
695: + N(\omega)}{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2 +
696: \left[H_\mathrm I(\omega)\right]^2}, \\
697: G_-(\omega) &= \frac{ N(\omega) + H_\mathrm I(\omega)}{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2
698: +
699: \left[H_\mathrm I(\omega)\right]^2}, \\
700: G_+(\omega) &= \frac{ N(\omega) - H_\mathrm I(\omega)}{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2
701: +
702: \left[H_\mathrm I(\omega)\right]^2}.
703: \end{align}
704: \end{subequations}
705: The Hadamard functions and Pauli-Jordan propagators are given by:
706: \index{Propagator!Hadamard}
707: \index{Propagator!Pauli-Jordan}
708: \index{Propagator!in the QBM models}
709: \begin{subequations} \label{GOtherFourier2}
710: \begin{align}
711: G^{(1)}(\omega) &= \frac{ 2 N(\omega)}{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2 +
712: \left[H_\mathrm I(\omega)\right]^2}, \\
713: G(\omega) &= \frac{ -2H_\text I(\omega)}{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2 +
714: \left[H_\mathrm I(\omega)\right]^2}.
715: \end{align}
716: \end{subequations}
717: % In the zero temperature case we may further simplify these
718: % expressions to:
719: % \begin{subequations}
720: % \begin{align}
721: % G_-(\omega) &= \frac{ 2N(\omega) \theta(-\omega) }{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2 +
722: % \left[N(\omega)\right]^2}, \\
723: % G_+(\omega) &= \frac{ 2 N(\omega) \theta(\omega) }{\left[ - \omega^2 + \omega_0^2 + H_\mathrm R(\omega)\right]^2 +
724: % \left[N(\omega)\right]^2}.
725: % \end{align}
726: % \end{subequations}
727:
728: Notice that in general only the retarded (and advanced) propagators follow the typical structure $G(\omega)=-i/[-\omega^2+\omega_0^2 + \Sigma(\omega)]$:
729: \begin{equation} \label{QBMSigma}
730: \GR(\omega) = \frac{-i}{-\omega^2+\omega_0^2+\SigmaR(\omega)}, \quad \SigmaR(\omega) = H(\omega).
731: \end{equation}
732: The Feynman propagator can only be put in this form in a vacuum state.
733:
734:
735:
736: \section{The stochastic approach} \label{sect:stoch}
737:
738: It is a well known fact that the evolution of the state of a closed quantum system composed entirely by free or linearly coupled harmonic oscillators is completely determined once the solution of the corresponding classical problem is known. This does not mean however that the behavior of the system is classical: it still exhibits genuine quantum properties such as the level discretization or the zero point energy. However, the time-evolution of the quantum system can be expressed in terms of the corresponding classical trajectories. Obviously, these classical trajectories do not have a physical realization but must be regarded as computational tools.
739:
740: Similarly, it can be shown that for linearly coupled open quantum systems the evolution of the reduced density matrix can be fully determined by the solution of a corresponding classical stochastic equation. In fact, the correspondence between quantum and stochastic dynamics for linear systems is also well known \cite{GardinerZoller} and was already noticed by Feynman long ago \cite{FeynmanVernon63,FeynmanQMPI}. It was further investigated in refs.~\cite{CalzettaRouraVerdaguer03,RouraThesis}. Again, the fact that the evolution of the system can be determined by stochastic methods does not mean that the system does not exhibit generic quantum properties. The stochastic trajectories do not have a direct physical meaning and should also be regarded as computational tools.
741:
742: Here we will not give a complete account of the stochastic description for open quantum systems, but instead we shall introduce the stochastic method via a heuristic trick, and show its usefulness with a particular example. The stochastic approach will be also used later in section \ref{sect:QBMdynamics}.
743:
744:
745: \index{Effective action}
746: There is a direct relation between the influence functional and
747: the CTP effective action for an open quantum system:
748: \cite{CamposVerdaguer96,MartinVerdaguer99b}
749: \begin{equation} \label{ActionCTPTree}
750: \Gamma^{(0)}[q,q']=S[q]-S[q']+S_{\mathrm{IF}}[q,q'],
751: \end{equation}
752: where $\Gamma^{(0)}[q,q']$ is the CTP tree-level
753: effective action, which for a linear open quantum system coincides
754: with the complete CTP effective action (see appendix \ref{app:CTP}). Thus, it is possible to
755: obtain equations of motion for the mean value of the Heisenberg
756: operator $\hat q$, which we will denote simply as $q$, by
757: demanding
758: \begin{equation}
759: \frac{\delta}{\delta q(t)} \Gamma^{(0)}_{\mathrm{CTP}}[q,q']
760: \Big|_{q'=q} = 0.
761: \end{equation}
762:
763: \index{Langevin equation}
764: \index{Effective action!stochastic}
765: Now here comes the trick
766: \cite{CamposVerdaguer96,MartinVerdaguer99b}: We define a
767: \emph{stochastic effective action} as
768: \begin{equation} \label{StochAction}
769: S_{\text{eff}}[q,q';\xi] = S[q] - S[q'] + \Re S_{\mathrm{IF}}[q,q']+\int \ud t \xi(t)\left[q'(t)- q(t)\right].
770: \end{equation}
771: In this expression $\xi(t)$ must be interpreted as a stochastic Gaussian process defined
772: by the correlators
773: \begin{subequations} \label{StochasticCorr}
774: \begin{align}
775: \av{ \xi(t) }_\xi &=0, \\
776: \av{ \xi(t)\xi(t') }_\xi &= N(t,t'),
777: \end{align}
778: \end{subequations}
779: where $\av{\cdot}_\xi$ means stochastic average, or equivalently
780: by the probability density functional
781: \begin{equation}
782: P[\xi] = \exp{\left( - \frac{1}{2} \int \ud t \ud {t'}
783: \xi(t) N^{-1}(t,t') \xi(t') \right)}.
784: \end{equation}
785: The stochastic average of a functional of $\xi$ is defined as \[\av{A([\xi],t)}_\xi = \int \tilde{\mathrm d} \xi\, P[\xi] A([\xi],t).\]
786: The effective action may be written as a statistical average,
787: \begin{equation}
788: \av{ \expp{ {i} S_{\text{eff}}[q,q';\xi]}}_\xi =
789: \expp{ i \Gamma^{(0)}_{\mathrm{CTP}}[q,q']}.
790: \end{equation}
791: Then we may obtain a stochastic equation of motion for the system,
792: the Langevin equation:
793: \begin{equation}
794: \frac{\delta}{\delta q(t)} S_{\text{eff}}[q,q';\xi]
795: \Big|_{q'=q} = 0.
796: \end{equation}
797: The introduction of the stochastic approach with this heuristic trick, although useful and appealing, can be misleading, since one could be tempted to think that the validity of the method is restricted to a classical or semiclassical regime, whereas in fact the stochastic method is valid in the fully quantum regime.
798:
799: According to \Eqref{ZCTP} the corresponding Langevin equation is
800: \begin{subequations}
801: \begin{equation} \label{LangevinMPB}
802: \ddot q(t) + \int \ud{t'} H(t,t') q(t') +\omega_0^2 q(t) = {\xi(t)},
803: \end{equation}
804: or, particularizing for an ohmic frequency distribution,
805: \begin{equation} \label{LangevinMPBExpl}
806: \ddot q(t) + \frac{g^2}{2} \dot q(t) +\omega_0^2 q(t) = \xi(t),
807: \end{equation}
808: \end{subequations}
809: where $\xi$ is a stochastic process of zero mean and correlation
810: function
811: \begin{equation}
812: \av{ \xi(t) \xi(t')}_\xi = N(t,t').
813: \end{equation}
814:
815: The Langevin equation, \Eqref{LangevinMPB}, can be formally
816: solved:
817: \begin{subequations}
818: \begin{equation} \label{LangevinSolution}
819: q(t) = q_\mathrm h(t;q_0,v_0) - i \int_{-\infty}^{\infty} \ud {s} \Gret(t,s)
820: \xi(s),
821: \end{equation}
822: where $q_\mathrm h(t;q_0,v_0)$ is a homogeneous solution of the differential equation \eqref{LangevinMPB} subject to the initial conditions $q_0=q(t_0;q_0,v_0)$ and $v_0=\dot q(t_0;q_0,v_0)$. Particularizing for the ohmic frequency distribution one gets
823: \begin{equation} \label{SolutionLangevin}
824: \begin{split}
825: q(t)= &\ q_0 \expp{-\gamma (t-t_0)/2} \cos \omega_1 (t-t_0) + \frac{(v_0+ \gamma q_0)}{\omega_1}
826: \expp{-\gamma (t-t_0)/2} \sin \omega_1 (t-t_0) \\
827: & - i \int_{-\infty}^{\infty} \ud {s} \Gret(t,s)
828: \xi(s),
829: \end{split}
830: \raisetag{1.2\baselineskip}
831: \end{equation}
832: \end{subequations}
833: where $\gamma$ and $\omega$ are defined in \eqref{gammaomega}, and $\Gret(t,t')$ is given by \Eqref{Gret}.
834: In the above equation we have assumed that we are in the weak dissipation case,
835: $\omega_0 > \gamma$.
836:
837: Generically speaking, the system has a dissipative dynamics, so that it will decay
838: to its equilibrium state
839: $q=\dot q=0$ if the stochastic source is turned off. For the ohmic environment the decay time is of order $\gamma^{-1}$. Consequently,
840: initial conditions will turn out to be unimportant if they are set in
841: the remote past, $t_0 \to -\infty$. Then we may forget about the
842: homogeneous part of the solution and concentrate in the last term
843: of \Eqref{SolutionLangevin}. In this situation we may compute the stochastic correlation
844: functions for the position $q$
845: \begin{equation*}
846: \begin{split}
847: \av{ q(t) q(t') }_\xi & = - \left\langle
848: \int \ud {s} \Gret(t,s) \xi(s)
849: \int \ud {s'} \Gret(t',s')
850: \xi(s')\right\rangle_\xi \\
851: & = -\int \ud{s} \ud{s'} \Gret(t,s) \av{\xi(s) \xi(s')}_\xi
852: \Gret(t',s').
853: \end{split}
854: \end{equation*}
855: Taking into account that $\av{\xi(s) \xi(s')}_\xi = N(s,s')$ [see \Eqref{StochasticCorr}], we can also write
856: \begin{equation}
857: \av{ q(t) q(t') }_\xi = - \int \ud{s} \ud{s'} \Gret(t,s) N(s,s')
858: \Gret(t',s'),
859: \end{equation}
860: and comparing with \Eqref{HadamardQBM} we conclude that
861: \begin{equation}
862: \av{ q(t) q(t') }_\xi = \fud G^{(1)}(t,t') = \fud \av{ \{ \hat q(t), \hat q(t')\} }.
863: \end{equation}
864: Thus, the stochastic two-point correlation corresponds to the symmetrized two-point correlation function (the Hadamard function).
865:
866: \index{Stochastic gravity}
867: The equivalence between stochastic correlation functions and some
868: quantum correlation functions, which we have presented through a
869: particular example, is indeed completely general. It can be shown on general grounds that stochastic correlation functions correspond to a subset of quantum correlation functions \cite{CalzettaRouraVerdaguer03,RouraThesis}.
870: The correspondence between quantum and stochastic correlation functions is at the heart of the theory of stochastic gravity \cite{CalzettaHu94,HuMatacz95,HuSinha95,CamposVerdaguer96,
871: CalzettaEtAl97,MartinVerdaguer99a,MartinVerdaguer99c,MartinVerdaguer00,HuVerdaguer03,HuVerdaguer04}.
872:
873: If initial conditions are set at some finite time\footnote{Recall however that factorized initial conditions are assumed, and that this assumption sometimes leads to unphysical results for finite initial times. See the comments in page \pageref{par:InitialConditions}.} $t_0$, then additionally to the average over the Gaussian stochastic process, an average over the initial conditions must be implemented. Explicitly, in this case the equivalence between quantum and stochastic correlation functions is given by \cite{RouraThesis,CalzettaRouraVerdaguer03}
874: \begin{equation}
875: \fud \av{ \{ \hat q(t), \hat q(t')\} } = \Av{ \av{ q(t) q(t') }_\xi }_{q_0,v_0},
876: \end{equation}
877: where the average over the initial conditions is defined as
878: \begin{equation}
879: \av{ q(t) q(t') }_{q_0,v_0} := \int \ud {q_0} \ud {v_0} q(t;q_0,v_0)q(t';q_0,v_0) W_\text s(q_0,v_0;t_0),
880: \end{equation}
881: where $ W_\text s(q_0,v_0;\ti)$ is the \emph{reduced Wigner function}, which is an alternative representation of the reduced density matrix $\hat\rho_\text{s}$.
882:
883: \index{Wigner function}
884: In general, given a density matrix $\hat \rho$, the \emph{Wigner function} is defined as \cite{HilleryEtAl84,Wigner32}
885: \begin{equation} \label{Wigner}
886: W(q,p;t):= \frac{1}{2\pi} \int \ud\Delta \expp{ip\Delta} \rho(q-\Delta/2,q+\Delta/2;t).
887: \end{equation}
888: The Wigner function has some similarities with a classical distribution function, although it cannot be interpreted as a probability density since the uncertainity principle prevents a simultaneous measure of the position and the momentum. In fact, the Wigner function can adopt negative values. However, the partial distributions $\int \ud q W(q,p,t)$ and $\int \ud p W(p,q,t)$ are true probability densities. Notice the similarity between the definition of the Wigner function and the definition of the inhomogeneous Fourier transformed propagators, \Eqref{MidTime}.
889:
890:
891: \section{The perturbative approach}\label{sect:QBMPert}
892: \begin{fmffile}{mpb}
893:
894: Let us try to recover the results of the previous sections by
895: using perturbation theory. Since the system
896: is linear, we expect the perturbative calculation to reproduce the exact results found in the previous sections. For the remaining of this section we restrict
897: ourselves to the ohmic distribution of frequencies with $\mathcal I (\omega)=1$. That the results can be easily generalized to an arbitrary distribution of frequencies by including an $\mathcal I(p)$ whenever there is a momentum integration.
898:
899: The reader must be warned that in this section we shall make use of methods and techniques typical of field theory, which will be described in the next chapter.
900:
901:
902: \subsection{Environment in a vacuum state}
903:
904: \index{Self-energy!in the vacuum}
905:
906: \index{Schwinger-Dyson equation}
907: The perturbative expansion for the dressed propagator of the
908: system propagator (indicated by a double line) is given by
909: \begin{equation*}
910: \parbox{12mm}{
911: \begin{fmfgraph}(12,5)
912: \fmfleft{i1}
913: \fmfright{o1}
914: \fmf{double}{i1,o1}
915: \end{fmfgraph}}
916: \ = \ %
917: \parbox{12mm}{
918: \begin{fmfgraph}(12,5)
919: \fmfleft{i1}
920: \fmfright{o1}
921: \fmf{plain}{i1,o1}
922: \end{fmfgraph}}\ + \ %
923: \parbox{28mm}{
924: \begin{fmfgraph}(28,5)
925: \fmfleft{i1}
926: \fmfright{o1}
927: \fmf{plain}{i1,v1}
928: \fmf{dashes}{v1,v2}
929: \fmf{plain}{v2,o1}
930: \fmfdot{v1,v2}
931: \end{fmfgraph}}
932: \ + \ %
933: \parbox{46mm}{
934: \begin{fmfgraph}(46,5)
935: \fmfleft{i1}
936: \fmfright{o1}
937: \fmf{plain}{i1,v1}
938: \fmf{dashes}{v1,v2}
939: \fmf{plain}{v2,v3}
940: \fmf{dashes}{v3,v4}
941: \fmf{plain}{v4,o1}
942: \fmfdot{v1,v2,v3,v4}
943: \end{fmfgraph}}
944: \ + \ \cdots,
945: \end{equation*}
946: where the single line indicates the free system propagator,
947: \begin{equation}
948: G^{(0)}_\mathrm F(\omega)= \frac{i}{m(\omega^2-\omega_0^2)+i\epsilon},
949: \end{equation}
950: and the dashed line indicates the free environment propagator,
951: \begin{equation}
952: \Delta_\mathrm F(\omega) = \int \udpi{p}
953: \frac{i}{\omega^2-p^2+i\epsilon} = \frac{1}{2|\omega|}.
954: \end{equation}
955: The Feynman rules are completed by adding a $i e (\pm i \omega)$
956: in each vertex linking a system (environment) with an environment
957: (system) propagator. The above series can be resumed with the aid
958: of the \emph{self-energy} $\Sigma(\omega)$, which is defined through the
959: \emph{Schwinger-Dyson equation}
960: \begin{equation} \label{SD}
961: G_\mathrm F(\omega) = G_\mathrm F^{(0)}(\omega) + G_\mathrm F^{(0)}(\omega)
962: [-i\Sigma(\omega)] G_\mathrm F(\omega),
963: \end{equation}
964: or alternatively,
965: \begin{equation}
966: G_\mathrm F(\omega)= \frac{i}{m(\omega^2-\omega_0^2)-\Sigma(\omega)}.
967: \end{equation}
968: Diagrammatically the self-energy is the sum of all one particle
969: irreducible diagrams with external legs amputated. In our
970: case there is just one of these diagrams,
971: \begin{equation*}
972: -i \Sigma(\omega) = \ \parbox{20mm}{\begin{fmfgraph}(20,5)
973: \fmfleft{i1}
974: \fmfright{o1}
975: \fmf{plain}{i1,v1}
976: \fmf{dashes}{v1,v2}
977: \fmf{dashes}{v2,v3}
978: \fmf{dashes}{v3,v4}
979: \fmf{plain}{v4,o1}
980: \fmfdot{v1,v4}
981: \end{fmfgraph}}
982: \end{equation*}
983: (it must be understood amputated) which can be straightforward
984: computed to give
985: \begin{equation}
986: -i\Sigma{(\omega
987: )}= (ie)(i\omega) \Delta_\text{F}(\omega) (ie)
988: (-i\omega) = - \frac{g^2 |\omega|}{2},
989: \end{equation}
990: and hence the self-energy coincides with that computed in the
991: previous sections, $\Sigma(\omega) = -{i g^2 |\omega|}/{2}$.
992:
993: \subsection{Environment in an arbitrary state}
994:
995: As explained in appendix \ref{app:CTP}, in an arbitrary state one has to consider the CTP
996: doubling of the degrees of freedom. There are two kind of
997: vertices, 1 and 2 vertices, and four kinds of propagators, 11
998: (Feynman), 22 (Dyson), 12 (negative Whightman) and 21 (positive
999: Whightman). When doing perturbation theory all
1000: Feynman diagrams with internal 1 and 2 vertices must be summed, taking into
1001: account that type 2 vertices carry an additional minus sign with
1002: respect to type 1 vertices. Graphically, the perturbative expansion of one of the propagators (\eg\ the Feynman propagator) can be represented as:
1003: \begin{equation*}
1004: \begin{split}
1005: \parbox{12mm}{
1006: \begin{fmfgraph*}(12,5)
1007: \fmfleft{i1}
1008: \fmfright{o1}
1009: \fmf{double}{i1,o1}
1010: \fmfv{label=\ssmall 1,label.angle=-90}{i1,o1}
1011: \end{fmfgraph*}}
1012: \ &= \ %
1013: \parbox{12mm}{
1014: \begin{fmfgraph*}(12,5)
1015: \fmfleft{i1}
1016: \fmfright{o1}
1017: \fmf{plain}{i1,o1}
1018: \fmfv{label=\ssmall 1,label.angle=-90}{i1,o1}
1019: \end{fmfgraph*}}\ + \ %
1020: \parbox{28mm}{
1021: \begin{fmfgraph*}(28,5)
1022: \fmfleft{i1}
1023: \fmfright{o1}
1024: \fmf{plain}{i1,v1}
1025: \fmf{dashes}{v1,v2}
1026: \fmf{plain}{v2,o1}
1027: \fmfv{label=\ssmall 1,label.angle=-90}{i1,o1}
1028: \fmfv{label=\ssmall 1,label.angle=-90}{v1}
1029: \fmfv{label=\ssmall 1,label.angle=-90}{v2}
1030: \fmfdot{v1,v2}
1031: \end{fmfgraph*}}
1032: \ + \ \parbox{28mm}{
1033: \begin{fmfgraph*}(28,5)
1034: \fmfleft{i1}
1035: \fmfright{o1}
1036: \fmf{plain}{i1,v1}
1037: \fmf{dashes}{v1,v2}
1038: \fmf{plain}{v2,o1}
1039: \fmfv{label=\ssmall 1,label.angle=-90}{i1,o1}
1040: \fmfv{label=\ssmall 1,label.angle=-90}{v1}
1041: \fmfv{label=\ssmall 2,label.angle=-90}{v2}
1042: \fmfdot{v1,v2}
1043: \end{fmfgraph*}}%
1044: \\[.3cm]
1045: &\quad + \ \parbox{28mm}{
1046: \begin{fmfgraph*}(28,5)
1047: \fmfleft{i1}
1048: \fmfright{o1}
1049: \fmf{plain}{i1,v1}
1050: \fmf{dashes}{v1,v2}
1051: \fmf{plain}{v2,o1}
1052: \fmfv{label=\ssmall 1,label.angle=-90}{i1,o1}
1053: \fmfv{label=\ssmall 2,label.angle=-90}{v1}
1054: \fmfv{label=\ssmall 1,label.angle=-90}{v2}
1055: \fmfdot{v1,v2}
1056: \end{fmfgraph*}}
1057: \ + \ \parbox{28mm}{
1058: \begin{fmfgraph*}(28,5)
1059: \fmfleft{i1}
1060: \fmfright{o1}
1061: \fmf{plain}{i1,v1}
1062: \fmf{dashes}{v1,v2}
1063: \fmf{plain}{v2,o1}
1064: \fmfv{label=\ssmall 1,label.angle=-90}{i1,o1}
1065: \fmfv{label=\ssmall 2,label.angle=-90}{v1}
1066: \fmfv{label=\ssmall 2,label.angle=-90}{v2}
1067: \fmfdot{v1,v2}
1068: \end{fmfgraph*}} \ + \ \cdots
1069: \end{split}.
1070: \end{equation*}
1071:
1072: \vspace{3mm}
1073:
1074: \index{Self-energy!in the QBM models}
1075: \index{Propagator!in the QBM models}
1076: The self-energy is a matrix-valued quantity related with the propagator through equation \eqref{SelfEnergyGeneralApp}. It can be computed as the following amputated diagram,
1077: \begin{equation*}
1078: -i \Sigma^{ab}(\omega) = \ \parbox{20mm}{\begin{fmfgraph*}(20,5)
1079: \fmfleft{i1}
1080: \fmfright{o1}
1081: \fmf{plain}{i1,v1}
1082: \fmf{dashes}{v1,v2}
1083: \fmf{dashes}{v2,v3}
1084: \fmf{dashes}{v3,v4}
1085: \fmf{plain}{v4,o1}
1086: \fmfdot{v1,v4}
1087: \fmfv{label=\ssmall $a$,label.angle=-90}{v1}
1088: \fmfv{label=\ssmall $b$,label.angle=-90}{v4}
1089: \end{fmfgraph*}},
1090: \end{equation*}
1091: or, symbollically, as
1092: \begin{equation}
1093: -i\Sigma^{ab}(\omega) = c^{ac} c^{bd} (ie) (i\omega) \Delta_{cd} (\omega) (-i\omega) (ie)
1094: \end{equation}
1095: where $\Delta_{ab}(\omega)$ is the field propagator with the momentum integrated out.
1096:
1097: \index{Self-energy!retarded}
1098: \index{Self-energy!Hadamard}
1099: In order to get explicit results, let us particularize to the case of a thermal state at temperature $\beta^{-1}$:
1100: \begin{equation}
1101: \begin{split}
1102: \Delta_{ab}(\omega) &= \int \udpi{p} \left[
1103: \begin{pmatrix}
1104: \dfrac{i}{\omega^2-p^2+i\epsilon} & 2\pi \delta(\omega^2-p^2) \theta(-\omega) \\
1105: 2\pi \delta(\omega^2-p^2) \theta(\omega) & \dfrac{-i}{\omega^2-p^2-i\epsilon}
1106: \end{pmatrix} \right. \\[2mm]
1107: &\qquad\qquad\left. + 2\pi \begin{pmatrix}
1108: 1 & 1\\
1109: 1 & 1
1110: \end{pmatrix} n(|\omega|) \delta(p^2-\omega^2) \right]\\
1111: &=
1112: \frac{1}{|\omega|}\begin{pmatrix}
1113: \fud + n(|\omega|) & - \theta(-\omega) - n(|\omega|)\\
1114: - \theta(\omega)-n(|\omega|) & \fud+n(|\omega|)
1115: \end{pmatrix},
1116: \end{split}
1117: \end{equation}
1118: where
1119: \begin{equation*}
1120: n(E) = \frac{1}{\expp{\beta E} - 1}.
1121: \end{equation*}
1122: Therefore the $ab$ component of the self-energy is given by
1123: \begin{equation}
1124: \begin{split}
1125: \Sigma^{ab}{(\omega)} &= - i {g^2 |\omega|} \begin{pmatrix}
1126: \fud + n(|\omega|) & - \theta(-\omega) - n(|\omega|)\\
1127: - \theta(\omega)-n(|\omega|) & -\fud+n(|\omega|)
1128: \end{pmatrix}.
1129: \end{split}
1130: \end{equation}
1131: Notice that the different self-energy components verify the relations found in appendix \ref{app:CTP}. We can extract the retarded self-energy.
1132: \begin{equation}
1133: \SigmaR(\omega) = \Sigma^{11}( \omega) + \Sigma^{12}(\omega) = - i {g^2 |\omega|}\left[\fud - \theta(-\omega)\right] = - i \frac{g^2 \omega}{2} = H(\omega),
1134: \end{equation}
1135: in agreement with \Eqref{QBMSigma}. Let us also mention that the Hadamard self-energy $\SigmaN(\omega) = \Sigma^{11}( \omega) + \Sigma^{22}(\omega)$ is proportional to the noise kernel:
1136: \begin{equation}
1137: \begin{split}
1138: \SigmaN(\omega) = - i g^2 |\omega| 2 \left[\fud +n(|\omega|)\right] = -i g^2 |\omega| \coth{\left(\frac{\beta|\omega|}{2}\right)} =-2i N(\omega).
1139: \end{split}
1140: \end{equation}
1141: The usefulness of the Hadamard self-energy and the relations between the self-energy and the propagator will be explored in more depth in the next chapter.
1142:
1143: \end{fmffile}
1144:
1145:
1146:
1147: % \section{Non-linear interaction case}
1148: %
1149: % %\subsection{General comments}
1150: %
1151: % \subsection{Non-linear in the environment}
1152: %
1153: % Let us consider the same MPB model but replacing the linear
1154: % interaction of \Eqref{SInt} by
1155: % \begin{equation} \label{IntInt}
1156: % S_{\text{int}}[q,\varphi] = \int \ud{t} \ud{x} \delta(x) e q
1157: % \varphi^2.
1158: % \end{equation}
1159: % In this case the general formalism is still valid, but the path
1160: % integrals of \Eqref{FInfl} must be evaluated perturbatively. Up to
1161: % quadratic order in $e$ the influence functional is given by
1162: % \begin{equation} \label{S_IF2}
1163: % \begin{split}
1164: % S_{\mathrm{IF}}[q,q^{\prime}]
1165: % =&\ \int_{t_i}^{t_f} \ud t Z_2(t) \Delta(t) -2\int_{t_{i}}^{t_{f}} \ud{t} \int_{t_{i}}^{t} \ud{t^{\prime
1166: % }} \Delta (t)D_2(t,t^{\prime }) Q(t^{\prime }) \\ & + \frac{i}{2}\int_{t_{i}}^{t_{f}}\ud{t}%
1167: % \int_{t_{i}}^{t_{f}} \ud{t^{\prime }} \Delta (t)N_2(t,t^{\prime }) \Delta
1168: % (t^{\prime }) + O(e^3)\text{,}
1169: % \end{split}
1170: % \end{equation}
1171: % where
1172: % \begin{align}
1173: % Z_2(t)&= - e \av{ \hat \varphi^2_I(t,0) }, \label{Z2}\\
1174: % D_2(t,t')&=\frac{ie^2}{2} \av{ [\hat \varphi^2_I (t,0), \hat
1175: % \varphi^2_I (t',0)] \label{D2}
1176: % },\\
1177: % N_2(t,t')&=\frac{e^2}{2} \av { \{\hat \varphi^2_I (t,0), \hat \varphi^2_I (t',0) \}}- Z_2(t)Z_2(t')
1178: % .\label{N2}
1179: % \end{align}
1180: % In the previous case the term linear in $e$ did not appear since
1181: % $\langle \hat \varphi_I(t,0) \rangle=0$. We shall use the shorthand
1182: % notation $\hat \varphi_I(t) \equiv \hat \varphi_I(t,0)$. %The alternative expression of
1183: % %\Eqref{S_IF-Alt} is still valid perturbatively making the
1184: % %corresponding changes and introducing a term linear in x
1185: %
1186: % $Z_2$ is the product at the same space-time point of two
1187: % operator-valued distributions, $\hat \varphi_I (t) \hat \varphi_I (t)$,
1188: % product which is not well defined in general \cite{Schwartz}. At a
1189: % practical level this means that $Z_2$ will be divergent, and a
1190: % regularization-renormalization process will have to be introduced,
1191: % adding the suitable counterterms to the action. On the other hand,
1192: % expressions \eqref{D2} and \eqref{N2} involve 4 point free
1193: % correlation functions, which can be explicitly evaluated with the
1194: % aid of Wick's theorem. One could think that that coefficients
1195: % $N_2$ and $D_2$ will be also divergent. However, the ultraviolet
1196: % singular behavior of $\langle \hat \varphi^2_I(t) \hat\varphi^2_I(t')
1197: % \rangle $ is the same as that of $\langle \hat \varphi^2_I(t) \rangle
1198: % \langle \hat\varphi^2_I(t') \rangle$, so that, in equations
1199: % \eqref{D2} and \eqref{N2}, all divergences exactly cancel. Of
1200: % course one regularization procedure is still needed to give the
1201: % bracketed expressions sense, but once these regularization has
1202: % been performed, the result does not depend on the regulator
1203: % chosen, and one is not obliged to introduce additional
1204: % counterterms in the action\footnote{Although the dissipation
1205: % kernel $D_2$ is free of divergences, the kernel $ H_2(t,t') \equiv
1206: % -2 D_2(t,t') \theta(t-t')$, which implicitly appears in
1207: % \eqref{S_IF2}, is again a product of two distributions, not well
1208: % defined. New divergences may arise from this term. However, note
1209: % that both divergences arising from $Z_2$ and $H_2$ enter into the
1210: % semiclassical equation of motion and not in the stochastic term.}.
1211: %
1212: %
1213: % Results of section 1.2 will continue to be correct in a
1214: % perturbative sense up to order $e^2$. The definition of the CTP
1215: % generating functional of \Eqref{ZCTP2} will be of course
1216: % applicable to this case. Since the influence action is known only
1217: % up to order $e^2$, all the results of section 1.3 will be still
1218: % valid perturbatively up to order $e^2$, replacing $\mathcal
1219: % N(t,t')$ by $N_2(t,t')$ and $\mathcal H(t,t')$ by $H_2(t,t')
1220: % \equiv -2 D_2(t,t') \theta(t-t')$, and dealing with divergences if
1221: % needed.
1222: %
1223: % We now proceed to evaluate explicitly the influence functional. We
1224: % start by evaluating the coefficient $Z_2$. As we do not want to
1225: % deal with infrared divergences, we will suppose that our scalar
1226: % field has a small mass $m$, much lower than the typical energies
1227: % in which we are interested, as we did in section 2. From
1228: % \Eqref{Prop} we get
1229: % \[
1230: % \langle 0 | \hat\varphi^2_I(t)|0 \rangle =
1231: % \int \udpi{p} \frac{1}{2\sqrt{p^2+m^2}}
1232: % \]
1233: % This integral is ultraviolet-divergent. We regularize it by
1234: % introducing a cutoff $\Lambda$ in the momentum. Thus
1235: % \begin{equation}
1236: % Z^{\Lambda}_2(t) = -e \langle 0 | \hat\varphi^2_I(t)|0 \rangle_\Lambda = - e \int_0^\Lambda \frac{\ud{p}}{2\pi} \frac{1}{\sqrt{p^2+m^2}}
1237: % = -\frac{e}{2\pi} \ln \frac{\Lambda}{2m} + O(m/\Lambda).
1238: % \end{equation}
1239: %
1240: % Next we compute the kernels $D_2(t,t')$ and $N_2(t,t')$. We need
1241: % the expression for the 4-point correlation function $\langle 0 |
1242: % \hat\varphi_I(t) \hat\varphi_I(t) \hat\varphi_I(t') \hat\varphi_I(t') | 0
1243: % \rangle$. Decomposing the fields in terms of the creation and
1244: % anihilation operators, or, alternatively, appealing to the Wick's
1245: % theorem, we can express this correlation function as
1246: % \begin{equation}
1247: % \begin{split}
1248: % \langle 0 |\hat\varphi^2_I(t) \hat\varphi^2_I(t')| 0 \rangle &=
1249: % \langle 0 |\hat\varphi^2_I(t) | 0 \rangle \langle 0 |\hat\varphi^2_I(t') | 0 \rangle + 2\langle 0 |\hat\varphi_I(t) \hat\varphi_I(t') | 0
1250: % \rangle^2 \\
1251: % &= \frac{1}{e^2} Z_2(t) Z_2(t') + \frac{2}{e^4}\left[ - i D(t,t') +
1252: % N(t,t') \right]^2 ,
1253: % \end{split}
1254: % \end{equation}
1255: % where $D(t,t')$ and $N(t,t')$ are the dissipation and noise
1256: % kernels of the linear case, with the addition of a small mass. We
1257: % have already computed the values of those kernels, see
1258: % Eqs.~\eqref{D_NR}. Decomposing this last expression in its
1259: % symmetric and antisymmetric parts we get the values of $N_2(t,t')$
1260: % and $D_2(t,t')$:
1261: % \begin{align}
1262: % N_2(t,t') &= \frac{e^2}{2} \langle 0 |\{ \hat\varphi_I^2(t) , \hat\varphi_I^2(t') \} | 0
1263: % \rangle - Z_2^2(t)
1264: % = \frac{2}{e^2} \left[ N^2(t,t') - D^2(t,t') \right], \\
1265: % D_2(t,t') &= \frac{ie^2}{2} \langle 0 |[ \hat\varphi_I^2(t) , \hat\varphi_I^2(t') ] | 0
1266: % \rangle = \frac{4}{e^2} D(t,t') N(t,t').
1267: % \end{align}
1268: % Introducing the explicit expressions for $D(t,t')$ and $N(t,t')$,
1269: % which are given by \eqref{D_NR}, we get
1270: % \begin{align}
1271: % N_2(t,t') &= \frac{e^2}{2\pi^2} \ln^2 \mu
1272: % |t-t'| - \frac{e^2}{8} , \\
1273: % D_2(t,t') &= \frac{e^2}{2\pi} \ln \mu |t-t'| \sign (t-t').
1274: % \end{align}
1275: % Recall that $\mu$ is a small mass scale of order $m$. Both kernels
1276: % $N_2(t,t')$ and $N_2(t,t')$ are well defined as bidistributions,
1277: % in the sense that they give a finite result when integrated with
1278: % test functions\footnote{In this particular case the kernel
1279: % $H(t,t') = e^2 (2\pi)^{-1} \ln \mu|t-t'| \theta(t-t')$ is also
1280: % well defined (see previous footnote).}.
1281: %
1282: % In order to renormalize the system we need to introduce the
1283: % following counterterm to the action,
1284: % \begin{equation}
1285: % S^\Lambda_\text{count}[q] = \int \ud t \left[ Z^\Lambda_2(t) - Z_0 \right]
1286: % q(t),
1287: % \end{equation}
1288: % where $Z_0$ is an arbitrary finite renormalization constant. The
1289: % bare action of the whole quantum system will read
1290: % $S_\text{bare}^\Lambda [q,\varphi] = S[q] + S[\varphi] +
1291: % S_\text{int}[q,\varphi] + S_\text{count}^\Lambda[q] $.
1292: %
1293: % Instead of introducing a hard cut-off in the momentum integrals,
1294: % it is also possible to make use of an alternative regulator,
1295: % smearing the delta function interaction through a function $f(x)$,
1296: % \ie, replacing the interaction term of \Eqref{IntInt} by
1297: % \begin{equation}
1298: % \begin{split}
1299: % S_\text{int}[q,\varphi] &= \int \ud t \ud x \ud{x'} f(x) f(x') e
1300: % q(t) \varphi(t,x) \varphi(t,x') \\
1301: % &= \int \ud t \int \frac{\ud p}{2\pi} \frac{\ud{p'}}{2\pi} \tilde f(-p) \tilde f(-p') e
1302: % q(t) \tilde \varphi(t,p) \tilde \varphi(t,p')
1303: % \end{split}
1304: % \end{equation}
1305: % Tildes indicate spatial Fourier transforms. With this regulator we
1306: % can readapt all above expressions replacing $\varphi(0,t)$ by $\int
1307: % \ud p \tilde f(p) \tilde \varphi(t,p) /(2\pi)$. We will require the
1308: % function $\tilde f(p)$ to be even in $p$ and rapidly decaying when
1309: % $p \to \infty$. The regulated $Z_2$ function will be given by
1310: % \begin{equation}
1311: % \begin{split}
1312: % Z^{f}_2(t) &= \av{ \int \frac{\ud p}{2\pi}
1313: % \frac{\ud{p'}}{2\pi} \tilde f(p) \tilde f(p') \tilde{\hat \varphi}_I(t,p) \tilde{\hat
1314: % \varphi}_I(t,p')}
1315: % = - e \int_0^\infty \frac{\ud{p}}{2\pi} \frac{\tilde f^2(p)}{\sqrt{p^2+m^2}}
1316: % \end{split}
1317: % \end{equation}
1318: % Choosing, for instance, a regulator $\tilde f^2(p) =
1319: % {\Lambda^2}/(p^2+\Lambda^2)$, which verifies $\tilde f(p=0)=1$ and
1320: % $\tilde f(p \gg \Lambda) \approx 0$, we get
1321: % \begin{equation}
1322: % Z^{f}_2(t) = - \frac{e^2}{2\pi} \ln \left( \frac{\Lambda}{2m} \right) +
1323: % O(m/\Lambda).
1324: % \end{equation}
1325: % In this case the result of employing this soft regulator coincides
1326: % with the use of a hard cut-off. As we have seen, the terms
1327: % $D_2(t,t')$ and $N_2(t,t')$ are independent of the regularization
1328: % chosen, so that the above calculations are still aplicable. If we
1329: % would like our results to be independent of the specific form of
1330: % the regulator $f(x)$ chosen, we should still perform a
1331: % renormalization procedure, which is exactly the same as the one we
1332: % have already sketched.
1333: %
1334: % %For the moment we do not want to consider it,
1335: % %so that we shall postpone the explicit evaluation of this case.
1336: %
1337: % \subsection{Non-linear in the system}
1338: %
1339: % Now we are going to consider the same model but with an
1340: % interaction term which is given by
1341: % \begin{equation}
1342: % S_{\text{int}}[q,\varphi] = \int \ud{t} \ud{x} \delta(x) e q^2
1343: % \varphi.
1344: % \end{equation}
1345: % In this case the general formalism is also correct and environment
1346: % path integrals of \Eqref{FInfl} can be performed exactly, so that
1347: % we may write
1348: % \begin{equation} \label{S_IF3}
1349: % \begin{split}
1350: % S_{\mathrm{IF}}[q,q^{\prime}]
1351: % =&-2\int_{t_{i}}^{t_{f}} \ud{t} \int_{t_{i}}^{t} \ud{t^{\prime
1352: % }}\Delta_2(t)D(t,t^{\prime })Q_2(t^{\prime }) \\ & + \frac{i}{2}\int_{t_{i}}^{t_{f}}\ud{t}%
1353: % \int_{t_{i}}^{t_{f}} \ud{t^{\prime }}\Delta_2 (t)N(t,t^{\prime })
1354: % \Delta_2
1355: % (t^{\prime })\text{,}
1356: % \end{split}
1357: % \end{equation}
1358: % with $Q_2(t) \equiv [q'^2(t)+q^2(t)]/2$ and $\Delta_2(t) \equiv
1359: % q'^2(t) - q^2(t)$. The values of $D$ and $N$ are given by
1360: % eqs.~\eqref{D} and \eqref{N}.
1361: %
1362: % Equation \eqref{ActionCTPTree} is valid and exact in this case,
1363: % but now we must take into account that, since there are quartic
1364: % terms in the action, the CTP effective action will have radiative
1365: % corrections. Therefore results of section 1.2 should be
1366: % interpreted as tree-level results.
1367: %
1368: % Since the influence action is quartic, one cannot evaluate the
1369: % path integrals of \Eqref{ZCTP2} exactly, and a perturbative
1370: % diagrammatic expansion is needed. It is possible to evaluate the
1371: % tree-level generating functional,
1372: % \begin{equation}
1373: % Z_{\mathrm{CTP}}^{(0)}[j_\Sigma,j_\Delta] = \expp{ - i \int \ud{t_1}
1374: % \ud{t_2}j_\Delta(t_1) \Gret^{(0)}(t_1,t_2)j_\Sigma(t_2)},
1375: % \end{equation}
1376: % being $\Gret^{(0)}(t,t')$ the retarded propagator of the kernel
1377: % \[
1378: % L^{(0)}(t,t') = m\left( \dert[2]{}{t} + \omega_0^2 \right)
1379: % \delta(t-t'),
1380: % \]
1381: % which is given by
1382: % \begin{equation}
1383: % \begin{split}
1384: % \Gret^{(0)}(t,t') &= \int \frac{\ud{\omega}}{2\pi}
1385: % \frac{\expp{-i \omega (t-t')}}
1386: % {m(\omega+\omega_0 + i \epsilon)(\omega-\omega_0+i\epsilon)} \\
1387: % &= \frac{1}{m\omega_0} \sin [\omega_0(t-t')] \theta(t-t').
1388: % \end{split}
1389: % \end{equation}
1390: % Then one can evaluate perturbatively the CTP generating functional
1391: % with the aid of
1392: % \begin{equation}
1393: % Z_{\mathrm{CTP}}[j, j'] = \exp \left( i S_{\mathrm{IF}} \left[i
1394: % \derf{}{j},i\derf{}{j'}\right] \right) Z_{\mathrm{CTP}}^{(0)}[j,j'].
1395: % \end{equation}
1396: % This formula is the usual starting point for the perturbative
1397: % evaluation of the generating functional in QFT \cite{Ramond}. If
1398: % we want to develop explicitly this last expression, we will find
1399: % divergences, so that again we shall need to renormalize
1400: % quantities.
1401: %
1402: %
1403: % \subsection{Non linear both in the system and in the
1404: % environment}
1405: %
1406: % Let us consider now the following interaction term:
1407: % \begin{equation}
1408: % S_{\text{int}}[q,\varphi] = \int \ud{t} \ud{x} \delta(x) e q^2
1409: % \varphi^2.
1410: % \end{equation}
1411: % This case is a combination of the two previous ones. Environment
1412: % path integrals must be performed perturbatively:
1413: % \begin{equation} \label{S_IF4}
1414: % \begin{split}
1415: % S_{\mathrm{IF}}[q,q^{\prime}]
1416: % =&\int_{t_i}^{t_f} \ud t Z_2(t) \Delta_2(t) -2\int_{t_{i}}^{t_{f}} \ud{t} \int_{t_{i}}^{t} \ud{t^{\prime
1417: % }}\Delta_2(t)D_2(t,t^{\prime })Q_2(t^{\prime }) \\ & + \frac{i}{2}\int_{t_{i}}^{t_{f}}\ud{t}%
1418: % \int_{t_{i}}^{t_{f}} \ud{t^{\prime }}\Delta_2 (t)N_2(t,t^{\prime })
1419: % \Delta_2
1420: % (t^{\prime }) + O(e^3)\text{,}
1421: % \end{split}
1422: % \end{equation}
1423: % with $Q_2(t) \equiv [q'^2(t)+q^2(t)]/2$ and $\Delta_2(t) \equiv
1424: % q'^2(t) - q^2(t)$. The values of $Z_2$, $D_2$ and $N_2$ are given
1425: % by eqs.~\eqref{Z2}--\eqref{N2}.
1426: %
1427: % Now the CTP effective action will have two kind of corrections:
1428: % first, we only know the tree-level CTP effective action
1429: % perturbatively up to order $e^2$; second, since there are quartic
1430: % terms in the action, this tree-level CTP effective action does not
1431: % coincide with the full CTP effective action due to the radiative
1432: % corrections.
1433: %
1434: % For the generating functional we also need a perturbative
1435: % treatement in this case:
1436: % \begin{equation}
1437: % Z_{\mathrm{CTP}}[j, j'] = \exp \left( i S_{\mathrm{IF}} \left[i
1438: % \derf{}{j},i\derf{}{j'}\right] \right) Z_{\mathrm{CTP}}^{(0)}[j,j'] + O(e^2).
1439: % \end{equation}
1440: % The difference with the previous case is that now we only know the
1441: % influence action up to order $e^2$, so that to be consistent we
1442: % should develop the exponential to linear order in
1443: % $S_{\mathrm{IF}}$.
1444: %
1445: % In this case and the previous ones, if we would like to obtain
1446: % higher precision results, we could do a systematic diagrammatic
1447: % expansion of all quantities.
1448:
1449: \section{Dynamics of perturbations}\label{sect:QBMdynamics}
1450:
1451: So far in this chapter we have only considered equilibrium properties, this is to say, we have considered systems having time-independent density matrices. In this section we would like to go a step further, and study the behavior of the reduced system when it is taken out of equilibrium. In fact, as it is shown in refs.~\cite{RouraThesis,CalzettaRouraVerdaguer03}, the methods and techniques presented in this chapter are already suited to treat this more general generic case, although our presentation was restricted to the equilibrium situation. For reasons that will become clear in the next chapter, we are interested in studying the behavior of the energy of systems exhibiting some particular properties, and in this situation it proves more useful for us to make a separate analysis, which will be based on the stochastic methods presented in sect.~\ref{sect:stoch}.
1452:
1453: At some time $t_0$ the system is perturbed by displacing the reduced system from its equilibrium situation. We shall assume that the system is such that the perturbation is long-lived, \ie, the characteristic lifetime of the perturbation is much longer than the characteristic oscillation time. Aside from this assumption, the frequency distribution and the state for the environment will be arbitrary. We shall try to compute, first, the energy of the perturbation and, second, its time-evolution. We are interested in studying properties for long times as compared to the characteristic oscillation time.
1454:
1455: Regarding the first point, a couple of comments are in order. First, notice that the energy of the perturbation is not defined unambiguously since the perturbation eventually decays, or, in other words, it does not correspond to the eigenstate of any Hamiltonian. However, if the perturbation is sufficiently long-lived we expect the energy to be still a physically meaningful quantity, defined up to some uncertainty given by the decay rate. A sensible measure for the energy would be the expectation value of some Hamiltonian. However, the naive guess that the energy is given by the expectation value of
1456: \begin{equation*}
1457: \hat H_\text{naive}(t) = \fud \dot{\hat q}^2(t) + \fud \omega_0^2 \hat q^2(t)
1458: \end{equation*}
1459: is obviously not necessarily correct, since the frequency $\omega_0$ need not coincide with the the physical frequency of the interacting oscillators, as there might be (finite or infinite) renormalization effects. For this reason, we shall instead compute the expectation value of the operator
1460: \begin{equation}
1461: \hat H_\text{sys}(t) = \fud \dot{\hat q}^2(t) + \fud \Omega_1^2 \hat q^2(t)
1462: \end{equation}
1463: with $\Omega_1$ being yet undetermined. We expect to find a physically reasonable criterion for choosing the value of $\Omega_1$.
1464:
1465: Since the Hamiltonian operator is a quadratic operator, and since it can be seen as the coincidence limit of a symmetrized correlation function, its quantum expectation value coincides with the stochastic expectation value:
1466: \begin{equation}
1467: E(t):=\av{\hat H_\text{sys}(t)} = \Av{\av{H_\text{sys}(t)}_\xi}_{q_0,v_0},
1468: \end{equation}
1469: where we recall that the stochastic average goes over the different realizations of the stochastic field and also over the initial Wigner function of the perturbation. The corresponding Langevin equation \eqref{LangevinMPB} is solved by \Eqref{LangevinSolution}.
1470:
1471: If the system is in equilibrium with the environment
1472: the expectation value of the reduced subsystem is given by the inhomogeneous part of the solution of the Langevin equation: \cite{GardinerZoller}
1473: \begin{equation}
1474: \begin{split}
1475: E_0 &= \av{\hat H_\text{sys}(t)}_0= \av{H_\text{sys}}_\xi = \fud \int \frac{\vd\omega}{2\pi} (\omega^2 + \Omega_1^2) |\Gret(\omega)|^2 N(\omega) \\ &= \fud \int \frac{\vd\omega}{2\pi} \frac{(\omega^2 + \Omega_1^2)N(\omega)}{[\omega^2 - \omega_0^2 - H_\text{R}(\omega)]^2 + H^2_\text{I}(\omega)}
1476: \end{split}
1477: \end{equation}
1478: See ref.~\cite{GardinerZoller} for a further developement of the above equation and its connection with statistical mechanics. As a side note, let us mention that when the system is at equilibrium the density matrix for the reduced subsystem does not correspond to the eigenstate of a Hamiltonian, so that actually there are fluctuations on the energy of the equilibrium system \cite{NagaevButtiker02}.
1479:
1480: When perturbations are included, the energy of the system is given by
1481: \begin{equation}
1482: E(t)=E_0 + \frac{1}{2} \av{\dot q_\text h^2(t;q_0,v_0)}_{q_0,v_0} + \fud \Omega_1^2 \av{ q_\text h^2(t;q_0,v_0)}_{q_0,v_0},
1483: \end{equation}
1484: where we recall that $q_\mathrm h(t;q_0,v_0)$ is a homogeneous solution of the Langevin equation \eqref{LangevinMPB} with initial conditions $q_0$ and $v_0$.
1485:
1486: Let us find explicitly the homogeneous solution of the Langevin equation. In order to deal with the non-local term in the Langevin equation, some additional information is needed on the form of the solution. Since we are considering slowly decaying perturbations, we assume that the solution is of the form
1487: \begin{equation} \label{HomAnsatz}
1488: q_\text h(t;v_0,q_0) = A(t;v_0,q_0) \expp{i\Omega t} +A^*(t;q_0,v_0) \expp{-i\Omega t} ,
1489: \end{equation}
1490: where the phase $\Omega t$ changes much more rapidly than the modulus $A(t;v_0,q_0)$. Under these assumptions the non-local term can be expressed as
1491: \begin{equation*}
1492: \int \ud s H(t,s) q_\mathrm h(s) = \int \ud s \frac{\vd \omega}{2\pi} \expp{-i\omega t}
1493: \left[ A(s) \expp{i(\Omega-\omega) s} + A^*(s) \expp{-i(\Omega +\omega)s}\right] H(\omega),
1494: \end{equation*}
1495: where, for simplicity, we omitted the dependence on the initial conditions. Let us concentrate on the first term on the right hand side of this equation. Taking into account that $A(s)$ is a slow function of $s$, the time integral only contributes significantly when $\omega \approx \Omega$; otherwise the phase oscillates too rapidly. Therefore $H(\omega)$ can be approximated by $H(\Omega)$ in the first term. Similarly, in the second term $H(\omega)$ can be approximated by $H(-\Omega)$. Hence,
1496: \begin{equation*}
1497: \begin{split}
1498: \int \ud s H(t,s) q_\mathrm h(s) &= H(\Omega) \int \ud s \frac{\vd \omega}{2\pi}A(s) \expp{i(\Omega-\omega) s} \\
1499: &\quad + H(-\Omega) \int \ud s \frac{\vd \omega}{2\pi}A^*(s) \expp{-i(\Omega+\omega) s}\\
1500: &= H(\Omega) A(t) \expp{i\Omega t} + H(-\Omega) A^*(t) \expp{-i\Omega t}.
1501: \end{split}
1502: \end{equation*}
1503: According to \Eqref{RealImOm}, the real part of the kernel $H(\omega)$ is even while the imaginary part is odd. Therefore the above equation can be reexpressed as
1504: \begin{equation*}
1505: \begin{split}
1506: \int \ud s H(t,s) q_\mathrm h(s) &= \Re{H(\Omega)} \left[ A(t) \expp{i\Omega t} + A^*(t) \expp{-i\Omega t} \right] \\ &\quad+ i \Im{H(\Omega)} \left[ A(t) \expp{i\Omega t}- A^*(t) \expp{-i\Omega t} \right],
1507: \end{split}
1508: \end{equation*}
1509: or alternatively, neglecting the time-derivatives of $A(t)$ in front of $\Omega A(t)$,
1510: \begin{equation}
1511: \begin{split}
1512: \int \ud s H(t,s) q_\mathrm h(s) &= \Re{H(\Omega)} q_\mathrm h(t) - \frac{1}{\Omega} \Im{H(\Omega)} \dot q_\mathrm h(t).
1513: \end{split}
1514: \end{equation}
1515: Notice that this amounts to a local expansion of the non-local term:
1516: \begin{subequations}
1517: \begin{equation}
1518: \int \ud s H(t,s) q_\text h(s) = \delta \Omega^2 q_\text h(t) + \Gamma \dot q_\text h(t)
1519: \end{equation}
1520: with
1521: \begin{equation}
1522: \delta \Omega^2 := \Re{H(\Omega)}, \qquad \Gamma := - \frac{1}{\Omega} \Im{H(\Omega)}.
1523: \end{equation}
1524: \end{subequations}
1525:
1526:
1527: Within this approximation the equation of motion is given by:
1528: \begin{equation}
1529: \ddot q (t) + \Gamma \dot q(t) + (\omega_0^2 + \delta\Omega^2) q(t) = 0
1530: \end{equation}
1531: and the corresponding solution is [see also \Eqref{SolutionLangevin}]
1532: \begin{equation}
1533: q_\mathrm h(t;q_0,v_0)= q_0 \expp{-\Gamma (t-t_0)/2} \cos{ \Omega(t-t_0)} + \frac{v_0}{\Omega} \expp{-\Gamma(t-t_0)/2} \sin{\Omega(t-t_0)},
1534: \end{equation}
1535: where we have identified $\Omega$ through the following self-consistent equation
1536: \begin{equation}
1537: \Omega^2 = \omega_0^2 + \delta\Omega^2 = \omega_0^2 + \Re H(\Omega),
1538: \end{equation}
1539: and where have made the approximation that $\Omega \gg \Gamma$. Notice that the solution is in accordance with the ansatz \eqref{HomAnsatz}, and notice also that the above equation applies to the ohmic case if $\gamma$ is small in comparison with $\omega_0$. In this case $\delta\Omega=0$ and $\Gamma=\gamma$.
1540:
1541: With this solution, the expectation value of the energy is given by
1542: \begin{equation}\label{EnergyPertLarge}
1543: \begin{split}
1544: E(t) = E_0 &+ \fud \expp{-\Gamma(t-t_0)} \bigg[ \av{q_0^2}_{q_0,v_0} \left[ \Omega^2 \sin^2{\Omega(t-t_0)} + \Omega_1^2 \cos^2{\Omega(t-t_0)} \right] \\
1545: & \quad+\Av{ \frac{v_0^2}{\Omega^2}}_{q_0,v_0}\left[ \Omega^2 \cos^2{\Omega(t-t_0)} + \Omega_1^2 \sin^2{\Omega(t-t_0)} \right]\\
1546: &\quad + \Av{ \frac{v_0 q_0}{\Omega}}_{q_0,v_0} (\Omega_1^2 -\Omega^2) \cos{\Omega(t-t_0)} \sin{\Omega(t-t_0)} \bigg].
1547: \end{split}
1548: \end{equation}
1549: In general the above function is rapidly oscillating with a frequency given by $\Omega$. This is not what is physically expected: the energy should be a smooth function, slowly decaying at long times but approximately constant at short timescales. However, recall that the parameter $\Omega_1$ is still undetermined. If we choose $\Omega_1=\Omega$ the time-evolution of the energy is greatly simplified, and we get the expected behavior:\footnote{There is a subtle point concerning \Eqref{QBMEt} and the factorized initial conditions. This equation implies that any perturbation increases the energy of the system. This is true even if we choose as the reduced state for the system one that it is apparently identical to the equilibrium state. The reason for this behavior is the assumption of uncorrelated initial states. In a short time scale the system will become correlated with the environment, thereby inducing a rapid change in the state of the system, after which it will undergo the slow decay described by \eqref{QBMEt}. The fast decoherence process is not described by \eqref{QBMEt} because of our initial adiabatic-like assumption.}
1550: \begin{equation}\label{QBMEt}
1551: E(t)=E_0 + \Av{ \frac{\Omega^2 q_0^2+ v^2_0}{2\Omega^2}}_{q_0,v_0} \expp{-\Gamma(t-t_0)}.
1552: \end{equation}
1553: Notice that this is similar to what would happen if we tried to compute the energy of an isolated harmonic oscillator using a Hamiltonian with an arbitrary frequency. When the Hamiltonian frequency differs from the physical frequency the expectation value of the fictitious Hamiltonian is not a constant but oscillates with the physical frequency of the oscillator. Only when the frequency coincides with the physical frequency the energy is constant. Similarly, we shall adopt the criterion that the physical frequency of the oscillator is that which makes the energy constant over short timescales.
1554:
1555: Let us end up by summarizing the main results of this section. We have seen that for those systems exhibiting long-lived excitations the dynamics of the perturbations is governed by the kernel $H(t,t')$, which is related to the dissipation kernel and which corresponds to the retarded self-energy. At first approximation, the excitations decay exponentially with a decay rate given by the imaginary part of $H(\Omega)$ (which coincides with the value of the Fourier-transformed dissipation kernel). Moreover, the effective frequency of the perturbation is determined by the real part of $H(\Omega)$.
1556:
1557: In the next chapter the Brownian particle will play the role of the quasiparticle mode, $\Omega$ will correspond to the quasiparticle energy, and $\Gamma$ will correspond to the quasiparticle decay rate.
1558:
1559:
1560: \index{Green function|see{propagator}}
1561: \index{Hadamard function|see{propagator}}
1562: \index{Whightman function|see{propagator}}
1563: