0803.2808/TBH.tex
1: %\documentclass[prl,twocolumn,aps,superscriptaddress,showpacs]{revtex4}
2: \documentclass[pre,twocolumn,aps,showpacs]{revtex4}
3: %\documentclass[prl,preprint,aps,showpacs]{revtex4}
4: %\documentclass[pre,preprint,showpacs]{revtex4}
5: \usepackage{graphicx}
6: \usepackage{amsmath}
7: \usepackage{amsfonts}
8: %\usepackage[left]{lineno}
9: %\linenumbers
10: %\usepackage{showlabels}
11: \bibliographystyle{apsrev}
12: \def\btt#1{{\tt$\backslash$#1}}
13: \renewcommand{\textfraction}{0}
14: \def\be{\begin{equation}}
15: \def\ba{\begin{eqnarray}}
16: \def\a{\alpha}
17: \def\b{\beta}
18: \def\g{\gamma}
19: \def\G{\Gamma}
20: \def\d{\delta}
21: \def\D{\Delta}
22: \def\e{\epsilon}
23: \def\z{\zeta}
24: \def\h{\eta}
25: \def\th{\theta}
26: \def\Th{\Theta}
27: \def\k{\kappa}
28: \def\l{\lambda}
29: \def\La{\Lambda}
30: \def\m{\mu}
31: \def\n{\nu}
32: \def\x{\xi}
33: \def\p{\pi}
34: \def\r{\rho}
35: \def\s{\sigma}
36: \def\t{\tau}
37: \def\f{\varphi}
38: \def\c{\chi}
39: \def\ps{\psi}
40: \def\o{\omega}
41: \def\vo{\varpi}
42: \def\O{\Omega}
43: \def\i{\int}
44: \def\lg{\langle \g \rangle}
45: \def\bA{{\mathbf A}}
46: \def\bB{{\mathbf B}}
47: \def\bx{{\mathbf x}}
48: \def\bv{{\mathbf v}}
49: \def\bD{{\mathbf D}}
50: \def\bM{{\mathbf M}}
51: \def\bb{{\mathbf b}}
52: \def\bff{{\mathbf f}}
53: \def\bG{{\mathbf G}}
54: \def\bV{{\mathbf V}}
55: \def\bq{{\mathbf q}}
56: \def\bp{{\mathbf p}}
57: \def\bz{{\mathbf z}}
58: \def\bZ{{\mathbf Z}}
59: \def\bp{{\mathbf p}}
60: \def\bq{{\mathbf q}}
61: \def\cP{{\mathcal P}}
62: \def\cG{{\mathcal G}}
63: \def\cT{{\mathcal T_>}}
64: \def\cH{{\mathcal H}}
65: \def\br{\mbox{\boldmath{$\r$}}}
66: \def\tr{\text{tr}}
67: %\def\bfi{\mbox{\boldmath{$\f$}}}
68: %\def\bg{\mbox{\boldmath{$\g$}}}
69: %\def\bps{\mbox{\boldmath{$\ps$}}}
70: %\def\bz{\mbox{\boldmath{$\z$}}}
71: \def\Tr{\mbox{Tr}}
72: \def\tc{\mbox{TC}}
73: %\def\bLa{\mbox{\boldmath{$\La$}}}
74: \def\ee#1{\label{#1}\end{equation}}
75: \def\ea#1{\label{#1}\end{eqnarray}}
76: %\usepackage[left]{lineno}
77: %\linenumbers
78: \begin{document}
79: %\draft
80: %\preprint{HEP/123-qed}
81: \title{Statistics of work performed on a forced quantum  
82: oscillator}
83: \author{Peter Talkner, P. Sekhar Burada, and Peter H\"anggi}
84: \affiliation{Institute of Physics, University of Augsburg,
85: D-86135 Augsburg, Germany}
86: \date{\today}
87: \begin{abstract}
88: Various aspects of the statistics of work performed by an external
89: classical force on a quantum
90: mechanical system are elucidated for a driven harmonic
91: oscillator. In this special case two parameters are introduced that 
92: are sufficient to completely characterize the force protocol. 
93: Explicit results for the
94: characteristic function of work and the respective probability
95: distribution are provided and discussed for three different types of
96: initial states of the oscillator: microcanonical, canonical and coherent
97: states.  Depending on the choice of the initial state the probability
98: distributions of the performed work may grossly differ. This result in
99: particular holds also true for
100: identical force protocols. General fluctuation and work theorems holding for
101: microcanonical and canonical initial states are confirmed.
102: \end{abstract}
103: \pacs{05.30.-d,05.70.Ln,05.40.-a}
104: \maketitle
105: \section{Introduction}
106: During the last decade various fluctuation and work theorems
107: \cite{fwt,ECM} have been formulated and discussed. They inter
108: alia characterize
109: the full nonlinear response of a system under the action of a time
110: dependent force \cite{J,J07}.
111: These theorems have been derived and experimentally
112: confirmed primarily for classical systems \cite{ex,BLR,BHSB}. 
113: Quantum mechanical generalizations
114: were proposed recently \cite{qm,M,RM,EM,TLH,TH,TMH,DL}. 
115: 
116: Conceptual problems though arise 
117: in the context of quantum mechanics 
118: if one tries to generalize those classical
119: relations that require for example the specification of a system's trajectory
120: extending over  
121: some interval of time, or the simultaneous measurement of
122: noncommuting observables.
123: For example, the measurement of work performed by an external force on
124: an otherwise isolated system may be accomplished in the framework of
125: classical physics in principle in two different ways. The first method
126: is based on two measurements of the energy, one at the beginning and
127: the second at the end of the considered process. This method becomes
128: unreliable in practice if the system is large and the work performed on the
129: system is negligibly small compared to the total energy of the system.
130: Such a situation typically arises if the system of interest, on which the
131: force exclusively acts, interacts with its environment. In order to
132: retain an isolated system, the large system made of the open system and
133: its environment must be considered. Again, the work performed on the
134: system results as the difference of the energies of the total system,
135: which both may be very large. 
136: 
137: For classical systems, this
138: unfortunate situation can be circumvented by a second method,  
139: by monitoring the state of
140: the relevant small system during the time when the force is acting. 
141: Having this information at hand one can  
142: determine the work by 
143: integrating the power supplied to the system at each instant of time. 
144: The respective power can be inferred from the registered state of
145: the system and the known force protocol. In a quantum
146: system a continuous measurement of even a single observable would strongly
147: influence and possibly manifestly distort the system's dynamics.
148: Apparently, only the first method of two energy measurements is
149: feasible, at least in principle, in the quantum context.  
150: 
151: An alternative method based on a continuous monitoring 
152: has recently
153: been suggested by Esposito and Mukamel \cite{EM} for open quantum
154: systems described
155: by Markovian quantum master equations. 
156: There the dynamics of the density matrix is 
157: mapped onto a classical rate process for which known fluctuation
158: theorems can be applied \cite{Seif}. This provides an interesting
159: formal approach but its physical meaning has remained unclear
160: \cite{EM}. 
161: Moreover,
162: this approach is restricted to open systems that only weakly
163: interact with their respective environments. 
164: 
165: In the present paper the distribution of work is discussed for the
166: exactly solvable system of a driven harmonic oscillator \cite{DL,H}. In
167: this case, 
168: the distribution of work is discrete. We provide formal expressions
169: for this distribution and its corresponding characteristic function
170: which are valid for all initial states of the system as well as for
171: all possible kinds of force protocols. In particular, we determine the
172: characteristic functions and distributions of the work for
173: microcanonical, canonical and coherent initial states which lead to
174: qualitatively different work distributions.
175: 
176: The paper is organized as follows. In Sect.~\ref{II} we review the general form
177: of the characteristic function of work performed on a system in terms
178: of a correlation function of the exponentiated Hamiltonians at the
179: initial and final time of the force protocol. We prove that this
180: particular expression indeed always represents a characteristic
181: function, i.e. the Fourier transform of a probability density.
182: Sect.~\ref{III} presents various fluctuation and work theorems for
183: canonical and microcanonical initial states. In Sect.~\ref{IV} general
184: expressions for the characteristic function and the corresponding
185: probability distribution of work are derived for a driven harmonic
186: oscillator. Moreover, the expressions for the first four cumulants are
187: derived. The dependence of the work distribution on the force protocol 
188: for microcanonical,
189: canonical and coherent initial states as well as its dependence on the
190: specific parameters of these initial states are investigated.
191:  
192:  
193: %{\it main aspects of the paper, short description of content }
194:            
195: 
196: \section{Characteristic function of work} 
197: \label{II}
198: The response of a quantum system on a perturbation by a classical,
199: external force can be characterized by the change of energies
200: contained in the total system. The energy as an observable coincides
201: with its Hamiltonian $H(t)$  of
202: the total system. It includes
203: the external force and therefore depends on time. We will consider the
204: dynamics of the system only within a finite window of time $[t_0,t_f]$
205: during which the force is acting in a prescribed way, resulting in a
206: protocol of Hamiltonians which is denoted by $\{H(t)\}_{t_f,t_0}$.
207: Apart from the action of the external force the system is assumed to
208: be closed. Its dynamics is consequently governed
209: by a unitary time evolution $U_{t,t_0}$,
210: which is the solution of the Sch\"odinger equation
211: \be
212: \begin{split}
213: i \hbar \partial U_{t,t_0} / \partial t &= H(t) U_{t,t_0}, \\
214: U(t,t_0)&=1. 
215: \end{split}
216: \ee{SE}
217: As explained in the introduction, the work $w$ is measured as the
218: difference of the energies of the system at the final and initial times
219: $t_{f}$ and $t_{0}$. In a single measurement  the work is given by
220: the difference of two eigenvalues $e_{n}(t_{f})$ and $e_{m}(t_{0})$ of the
221: Hamiltonians $H(t)$ at the respective times $t_{f}$ and $t_{0}$,
222: i.e. by $w=e_{n}(t_{f})-e_{m}(t_{0})$. The inherent randomness of the
223: outcome of a quantum measurement in general leads to a  measured work
224: that is random.   
225: A complete description of the statistical properties of the work
226: performed on the system is provided by the characteristic function
227: $G_{t_{0},t_{f}}(u)$ 
228: which presents the Fourier transform of the probability density of the
229: work $p_{t_{f},t_{0}}(w)$, i.e.
230: \be
231: G_{t_{0},t_{f}}(u) = \int dw\: e^{iuw} p_{t_{f},t_{0}}(w).
232: \ee{Gp}
233: It can be expressed as a quantum correlation function of the
234: exponentiated Hamiltonian at the initial and the final time
235: \cite{TMH}, i.e.  
236: \be
237: \begin{split}
238: G_{t_{0},t_{f}}(u) &= \langle e^{iu H(t_{f})} e^{-iu H(t_{0})} \rangle
239: \\ 
240: & \equiv \Tr e^{i u H_{H}(t_{f})} e^{-iu H(t_{0})} \bar{\r}(t_{0}),
241: \end{split}
242: \ee{GC}
243: where 
244: \be
245: H_{H}(t_{f}) = U^{+}_{t_{f},t_{0}} H(t_{f}) U_{t_{f},t_{0}}
246: \ee{HH}
247: denotes the Hamiltonian in the Heisenberg picture. The density
248: matrix $\bar{\r}(t_{0})$ from the initial density matrix $\r(t_{0})$
249: as a result of the
250: measurement of the Hamiltonian $H(t_{0})$. It is given by
251: \be
252: \bar{\r}(t_{0}) = \sum_{n} P_{n}(t_{0}) \r(t_{0}) P_{n}(t_{0}),    
253: \ee{br}
254: where the operators $P_{n}(t_{0})$ denote the eigenprojection operators of the
255: Hamiltonian at time $t_{0}$, which present a partition of the unity
256: \be
257: \sum_{k} P_{n}(t_{0}) = 1.
258: \ee{P1}  
259: Before we apply the general expression (\ref{GC}) 
260: to a particular system and investigate its dependence on the
261: initial state $\r(t_{0})$, we discuss three general properties of the
262: correlation expression (\ref{GC}) which guarantee that the resulting
263: function $G_{t_{f},t_{0}}(u)$ indeed always presents a proper characteristic
264: function of a classical random variable $w$. This is the consequence
265: of the three following properties: \\ 
266: (i) $G_{t_f,t_{0}}(u)$ is a continuous function of $u$.\\
267: (ii)  $G_{t_f,t_0}(u)$ is a positive definite function of $u$, i.e. 
268: for all integer numbers $n$, 
269: all real sequences $u_1,u_2,\ldots u_n$, and all complex numbers
270: $z_{i}$, $i=1,2 \ldots n$ 
271: \be
272: \sum_{i,i'}^n  G_{t_f,t_0}(u_i-u_{i'})
273: z^*_{i} z_{i'} \geq 0 
274: \ee{pd}
275: holds. Here, the asterisk $z_{i}^*$ denotes the complex conjugate of
276: $z_{i}$.\\
277: (iii) $G_{t_f,t_0}(0) = 1$\\
278: According to a theorem by Bochner \cite{Boch} the properties (i-iii)
279: are necessary and sufficient conditions in order that the function
280: $G_{t_f,t_0}(u)$ 
281: is the Fourier transform of the probability measure of a random
282: variable. In short, the first condition insures that, strictly speaking,
283: the function $G_{t_{f},t_{0}}(u)$ is the Fourier transform of a
284: measure, the second condition assures that this measure is positive
285: and the third condition that it is normalized.
286: Hence the correlation expression eq. (\ref{GC}) always
287: defines a proper characteristic function. For a
288: proof of the properties (i-iii) we refer the reader to the appendix \ref{AA}.
289: 
290: \section{Canonical and microcanonical initial states}
291: \label{III}
292: In  experiments an external force is often applied on a system that
293: initially is found in a 
294: thermodynamic equilibrium state. Depending on whether the system was
295: in weak contact with a heat bath or was totally isolated from its
296: environment, the 
297: initial state of the system is described either by a canonical or a 
298: microcanonical density matrix. For both situations fluctuation and
299: work theorems are known. We will shortly review these relations.
300: \subsection{Work and fluctuation theorems for canonical initial
301:   states} 
302: 
303: If the initial density matrix is canonical, i.e. 
304: if 
305: \be
306: \r(t_0) = Z^{-1}(t_0) \exp \{ -\b H(t_0) \},
307: \ee{cs} 
308: where 
309: \be
310: Z(t_0) = \Tr \exp \{ - \b H(t_0) \} = e^{-\b F(t_{0})}
311: \ee{Z} 
312: denotes the partition function and $F(t_{0})$ the free energy, 
313: then $[H(t_0),\r(t_0)]=0$
314: and the first measurement of
315: the energy leaves the density matrix unchanged, such that
316: $\bar{\r}(t_0) = \r(t_0)$. With eq. (\ref{GC}) this leads to the
317: characteristic function of work for a canonical initial state which was
318: derived in Ref.~\cite{TLH}.
319: In this case, $G_{t_f,t_0}(u)$ can be
320: continued to an analytic function of $u$ for all $0\leq \Im u \leq
321: \b$ \cite{TH}. For the particular value $u=i\b$ the characteristic function
322: yields the mean value of the exponentiated work, $\langle \exp\{-\b w\}
323: \rangle$ and the 
324: correlation function expression (\ref{GC}) simplifies to the ratio of
325: the partition functions at the times $t_f$ and $t_0$, resulting in the
326: Jarzynski work theorem
327: \be
328: \langle e^{-\b w} \rangle    = Z(t_f)/Z(t_0) = \exp\left \{-\b
329:   (F(t_{f})-F(t_{0})) \right \},
330: \ee{J}
331: where $Z(t_f) = \tr \exp \{-\b H(t_f) \} = \exp \{- \b F(t_{f}) \}$.
332: 
333: Within the domain of analyticity $\mathcal{S} = \{u | 0 \leq \Im u
334: \leq \b\}$ the characteristic functions for the original and the time reversed
335: protocol are related to each other by the following formula, cf. \cite{TH} 
336: \be
337: G_{t_f,t_0}(u) = \frac{Z(t_f)}{Z(t_0)} G_{t_0,t_f}(-u+i\b),
338: \ee{GTC} 
339: where $G_{t_0,t_f}(u)$ refers to processes under the time reversed
340: protocol $\{H(t)\}_{t_o,t_f}$ starting from the canonical
341: state $Z^{-1}(t_f) \exp \{ -\b H(t_f)\}$. An inverse Fourier
342: transform leads to the Tasaki-Crooks fluctuation theorem, which
343: relates the probability
344: densities  of work $p_{t_f,t_0}(w)$ for a given protocol to
345: the density of
346: the work $p_{t_0,t_f}(w)$ for the time reversed protocol. This theorem
347: explicitly reads \cite{TH}
348: \be
349: \frac{p_{t_f,t_0}(w)}{p_{t_0,t_f}(-w)} = \frac{Z(t_f)}{Z(t_0)} e^{\b
350:   w} = e^{-\b(F(t_{f})-F(t_{0})-w)}.
351: \ee{TC}  
352: \subsection{Fluctuation theorems for microcanonical initial states}
353: A system in a microcanonical state is described by the density matrix
354: \be
355: \r(t_{0}) = \vo_{E}^{-1}(t_{0}) \d (H(t_{0})-E),
356: \ee{rmc}
357: where 
358: \be
359: \vo_{E}(t_{0}) = \Tr\: \d (H(t_{0})-E) = \exp\left \{S(E,t_{0})/k_{B}
360: \right \}
361: \ee{oE}
362: denotes the density of states as a function of the energy $E$ of the
363: system. The density of states can be expressed in terms of the
364: entropy of the system $S_{E}(t_{0})$ provided the spectrum of the
365: system Hamiltonian is sufficiently dense such that the density of
366: states becomes a smooth function on a coarsened energy scale. 
367: The microcanonical density
368: matrix commutes with the Hamiltonian $H(t_{0})$. Consequently,
369: $\bar{\r}(t_{0})$ and $\r(t_{0})$ coincide. 
370: 
371: The microcanonical quantum Crooks theorem
372: assumes the form \cite{TMH}
373: \be
374: \begin{split}
375: \frac{p_{t_{f},t_{0}}(E,w)}{p_{t_{0},t_{f}}(E+w,-w)} &=
376: \frac{\vo_{E+w}(t_{f})}{\vo_{E}(t_{0})}\\
377: & =
378: \exp \left \{\frac{S(E+w,t_{f})-S(E,t_{0})}{k_{B}} \right \}.
379: \end{split}
380: \ee{mcC}
381: Analogous to the canonical case it relates the probability density
382: $p_{t_{0},t_{0}}(E,w)$ of work $w$,
383: for a system starting in a microcanonical state with energy $E$, to the
384: respective quantity for the time reversed process starting
385: at energy $E+w$. This quantum theorem is formally identical to the
386: respective classical theorem \cite{CBK}.     
387: 
388: From the microcanonical Crooks theorem the probability density
389: relating to the time reversed process can be eliminated to yield the
390: so-called entropy-from-work theorem \cite{TMH}, reading:
391: \be
392: \vo_{E_{f}}(t_{f}) = \i dw\: \vo_{E_f-w}(t_{0})
393: p_{t_{f},t_{0}}(E_{f}-w,w) \;.
394: \ee{efw}  
395: This theorem allows one to determine the unknown density of states of a system
396: with Hamiltonian $H(t_{f})$ from the known density of states of a reference
397: system $H(t_{0})$ by means of the statistics of the work that is
398: performed on the system in a process that leads from the reference
399: system to the final system with unknown density of states. In the case
400: of systems with a sufficiently smooth density of states
401: the respective entropy can be determined. For further details
402: see in Ref.~\cite{TMH}.   
403: \section{Driven harmonic oscillator}
404: \label{IV}
405: To illustrate these concepts we consider 
406: an example which allows the analytical construction of the
407: probability of work. Specifically we consider a harmonic oscillator 
408: on which a time
409: dependent force acts during a finite interval of time.  
410: Its time evolution is governed 
411: by the Hamiltonian
412: \be
413: H(t) = \hbar \o a^+ a + f^*(t) a + f(t) a^+,
414: \ee{hoH}
415: where $\o$ denotes the angular frequency, and $a^+$ and $a$ creation and
416: annihilation operators, respectively, which obey the usual commutation
417: relation, i.e. $[a,a^+]=1$. The complex driving force $f(t)$ allows
418: for a coupling
419: to position and/or momentum of the oscillator.  
420: We assume that $f(t)$ vanishes for times $t\leq t_0=0$. 
421: It is our aim to study the influence of the initial state
422: $\r(t_{0})$ on the statistics of work performed on the oscillator. 
423: The measurement of
424: $H(t_{0}) = \hbar \o a^{+} a$
425: at time $t_{0}=0$ then yields the result $\hbar \o n$ with probability 
426: \be
427: p_{n} = \langle n|\r(t_{0})|n\rangle.
428: \ee{pn}
429: Accordingly, the oscillator is found in the state
430: \be
431: \bar{\r}(t_{0}) = \sum_{n} p_{n} |n\rangle \langle n|
432: \ee{bro}
433: immediately after this measurement. Putting this density matrix in the
434: general expression for the characteristic function, eq.~(\ref{GC}) one
435: obtains
436: \be
437: G_{t_f,t_{0}}(u) = \sum_n p_n e^{-iu\hbar \o
438:   n} \langle n |
439: e^{iu H_H(t_f)} |n \rangle\;.
440: \ee{G0} 
441: For the driven harmonic oscillator the diagonal matrix element of the
442: exponentiated Hamiltonian $H_{H}(t_{f})$ can be determined \cite{H}. For
443: details see the Appendix~\ref{B}. With the expression (\ref{nxy}) for
444: the matrix element $\langle n| \exp \left \{ iH_{H}(t_{f})\right \} |n
445: \rangle$ we
446: find
447: \begin{widetext}
448: \be
449: \begin{split}
450: G_{t_f,t_{0}}(u) =& \:  e^{iu |f(t_f)|^2/(\hbar \o)} 
451: \exp \left \{ \left ( e^{iu\hbar \o}-1
452:   \right )|z|^2 \right \}   \sum_{n=0}^\infty\sum_{k=0}^n p_n
453: \binom{n}{k} \frac{|z|^{2(n-k)}}{(n-k)!} e^{-iu\hbar \o (n-k)} \left
454:     (e^{iu\hbar \o}-1 \right )^{2(n-k)} \\
455: &=\:  e^{iu |f(t_f)|^2/(\hbar \o)} 
456: \exp \left \{ \left ( e^{iu\hbar \o}-1
457:   \right )|z|^2 \right \} \sum_{n=0}^{\infty}p_{n}L_{n}\!\left (4 |z|^{2}
458: \sin^{2}\frac{\hbar \o u}{2}\right )\;,
459: \end{split}
460: \ee{Gho} 
461: \end{widetext}
462: where $|f(t_f)|^2/(\hbar \o)$ denotes a uniform shift of the 
463: spectrum of the harmonic
464: oscillator due to the presence of the external force, cf.
465: eq. (\ref{L}), and 
466: \be
467: z = \frac{1}{\hbar \o} \i_0^{t_f} ds \dot{f}(s) \exp \{i \o s\} 
468: \ee{zf}
469: is a dimensionless functional of the driving force $f(t)$, cf. eq. (\ref{z}). 
470: This dimensionless quantity vanishes in particular for all-quasi
471: static forcings,
472: i.e. if the force changes only very slowly with 
473: $f(t)= g(t/t_{f})$ for $t_{f} \to \infty$,  
474: where $g(\t)$ is a continuously differentiable
475: function for $\t \in [0,1]$. We hence call $z(t)$ the {\it rapidity
476: parameter} of the force protocol.  
477: Finally, $L_{n}(x) =\sum_{k=0}^{n} \binom{n}{k} (-x)^{k}/k!$ denotes
478: the Laguerre polynomial of order $n$ \cite{RG}.
479: 
480: 
481: Introducing the cumulant generating function $K(u)= \ln G(u)$ one
482: obtains 
483: the cumulants of work $k_{n}$  
484: as the $n$th derivatives of $K(u)$  with respect to $u$ taken at $u=0$
485: \cite{vK}, i.e. $k_{n} = (-i)^{n} d^{n }K(0)/d u^{n}$. The 
486: first four cumulants become:
487: \begin{align}
488: k_{1} &= \langle w \rangle \nonumber\\ &= \frac{|f(t_f)|^2}{\hbar \o} +\hbar \o
489: |z|^2, \label{aw}\\
490: k_{2} &=\langle w^{2} \rangle -\langle w \rangle^{2}\nonumber\\
491:  &= 2 (\hbar \omega)^2 |z|^2
492: \left ( \langle a^+ a \rangle_0 +\frac{1}{2} \right ),\label{ww}\\
493: k_{3} &=\langle w^{3} \rangle- 3 \langle w^{2} \rangle \langle w
494: \rangle +2 \langle w
495: \rangle^{3}\nonumber\\ 
496: &= \left (\hbar \o \right )^{3} |z|^{2},\\
497: k_{4}&=  \langle w^{4} \rangle -4 \langle w^{3} \rangle\langle w
498: \rangle -3 \langle w^{2} \rangle^{2} \nonumber \\
499: & \quad +12 \langle w^{2} \rangle \langle
500: w \rangle^{2} - 6 \langle w \rangle^{4}\nonumber\\ 
501: &= \left (\hbar \o \right )^{4}
502: |z|^{2} \left \{ 1+4\langle a^{+}a \rangle_{0} + 6 \left [\langle
503:   a^{+}a(a^{+}a-1)\rangle_{0} \right . \right .\nonumber \\
504: &\left . \left .\quad - 2\langle a^{+}a \rangle_{0} \right ]|z|^{2}\right \}.   
505: \label{cu}
506: \end{align}
507: The odd cumulants of the work are independent of the
508: initial preparation. The even cumulants depend on the
509: factorial moments of the number operator $a^{+}a$ with respect to
510: the initial state $\bar{\r}(t_{0})$ such as $\langle
511: a^{+}a \rangle_{0} = \sum_{n} n p_{n}$ and $\langle
512: a^{+}a(a^{+}a -1) \rangle_{0} = \sum_{n} n(n-1) p_{n}$, where $p_{n}$
513: is defined in eq.~(\ref{pn}).  Moreover, all
514: cumulants apart from the first one 
515: vanish for forcings with $z=0$. This holds true in particular for all quasi-static
516: force characteristics. The underlying work probability density then
517: shrinks to a delta function at $w=|f(t_{f})|^{2}/(\hbar \o)$. 
518: 
519: In general, the work probability density follows from the characteristic
520: function by means of an inverse Fourier transformation. Rather than
521: the characteristic function itself we first consider the function
522: $\mathcal{G}(u)\equiv \exp \left \{ -iu|f(t_{f})|^{2}/(\hbar \o)\right\}
523: G_{t_{f},t_{0}}(u)$. Upon expanding $\exp\left \{|z|^{2} \exp\left
524:   \{iu\hbar \o\right \} \right \}$
525: into a series of powers of $|z|^{2}$ we obtain for $\mathcal{G}(u)$ a
526: Laurent series in the variable $\exp\left\{iu\hbar \o\right\}$. The
527: inverse Fourier transformation is given by a series of delta
528: functions $\d(w-\hbar \o r)$, with $r \in \mathbb{Z}$, with
529: weights
530: \be
531: \begin{split}
532: q_r =& e^{-|z|^2 } \sum_{m,n=0}^\infty \sum_{k=0}^n
533: \sum_{l=0}^{2k}(-1)^{2k-l} p_n  \\
534: &\quad\times \frac{|z|^{2(k+m)}}{m!\: k!}
535: \binom{n}{k} \binom{2k}{l} \d_{l+m,k+r}\\
536: =& e^{-|z|^2}\sum_{n=0}^\infty \sum_{k=0}^n \sum_{l=0}^{\text{min}
537:   (k+r,2k)}
538: (-1)^{2k-l}  p_n  \\
539: &\quad \times \frac{|z|^{2(2k+r-l)}}{(k+r-l)!\:k!} \binom{n}{k}
540: \binom{2k}{l}\;. 
541: \end{split}
542: \ee{q}    
543: The factor  $\exp \left \{ -iu|f(t_{f})|^{2}/(\hbar \o)\right\}$, by
544: which $\mathcal{G}(u)$ has to be multiplied to yield
545: $G_{t_{f},t_{0}}(u)$, gives
546: rise to a constant shift such that the probability density of work
547: performed on a harmonic oscillator assumes the result 
548: \be
549: p_{t_f,0}(w) = \sum_r q_{r} \:\d\!\left (w-(\hbar \o r
550: +\frac{|f(t_{f})|^{2}}{\hbar \o})\right ) \;.
551: \ee{pt0}
552: In the next Section we will investigate the influence of the initial
553: state on the statistics of the work.
554: \subsection{Distributions of work for different initial states}
555: As particular examples of initial states we will discuss
556: microcanonical, canonical and coherent states.\\
557: \subsubsection{Microcanonical initial state}
558: For a microcanonical initial state with energy $\hbar \o n_{0}$ the
559: density matrix becomes
560: \be
561: \r(t_{0}) = \bar{\r}(t_{0}) = |n_{0}\rangle \langle n_{0} |.
562: \ee{mco}
563: The characteristic function then reads
564: \be
565: \begin{split}
566: G^{\text{mc}}_{t_{f},t_{0}}(n_{0},u) &= e^{iu|f(t_{f})|^{2}/(\hbar
567:   \o)} \exp\left 
568:   \{ \left( e^{iu\hbar \o} -1\right)|z|^{2}\right \} \\
569: &\quad \times L_{n_{0}}\!\left ( 4|z|^{2} \sin^{2} \frac{\hbar \o u}{2}\right )
570: \end{split}
571: \ee{mccf0}   
572: and, accordingly, the probability $q^{\text{mc}}_{r}(n_{0})$ to find a change of
573: energy by $w=\hbar \o r +|f(t_{f})|^{2}/(\hbar \o)$ emerges as 
574: \be
575: \begin{split}
576: q^{\text{mc}}_{r}(n_{0})& = e^{-|z|^{2}} \sum_{k=0}^{n_{0}} \sum_{l=0}^{\min
577:   (k+r,2k)}\frac{(-1)^{2k-l}}{(k+r-l)!\: k!}\\
578: &\quad \times \binom{n}{k} \binom{2k}{l}
579: |z|^{2(2k+r-l)}.
580: \end{split}
581: \ee{mcq}
582: As expected from the behavior of the moments, all probabilities
583: $q^{\text{mc}}_{r}(n_{0})$ with $r\neq 0$ vanish for quasi-static forcing, i.e. if
584: $z \to 0$. The dependence of $q^{\text{mc}}_{r}(n_{0})$ for $n_{0}=0$ and $3$ 
585: as well as for the eight lowest values of
586: $r$ on the parameter $z$ is displayed in Fig.~\ref{fz}. With
587: increasing values of the rapidity parameter $z$ the distribution is
588: becoming broader.
589: \begin{figure}
590: \includegraphics[width=8cm]{Fig1-a.eps}\hfill 
591: \includegraphics[width=8cm]{Fig1-b.eps} 
592: \caption{The probabilities $q^{\text{mc}}_{r}(n_{0})$ for two
593: microcanonical initial states with $n_{0}=0$ (a) and $n_{0}=3$ (b) are 
594: depicted for $r=-3
595: \ldots 4$, as functions of the rapidity parameter $z$ in
596: eq.~(\ref{z}).  In both cases the
597: distribution collapses at $r=0$ for adiabatic forcing corresponding to
598: $|z|=0$ and broadens with increasing $|z|$. Obviously, when starting in
599: the ground state the oscillator cannot deliver work whence the
600: probability for negative $r$ strictly vanishes. ``Stimulated
601: emission'' becomes possible from an excited state at finite driving
602: rapidity $z$ leading to nonzero probabilities
603: $q^{\text{mc}}_{r}(n_{0})$ at negative values of $r$ in panel (b).}      
604: \label{fz}
605: \end{figure}
606: For the fixed value of $|z|=2$ the distribution
607: $q^{\text{mc}}_{r}(n_{0})$ is compared for the three initial states  with
608: $n_{0}=0$ , $10$ and $30$ in
609: Fig.~\ref{fn}. 
610: \begin{figure}
611: \includegraphics[width=8cm]{Fig2.eps}
612: \caption{(Color online) The probabilities $q^{\text{mc}}_{r}(n_{0})$ for a
613: microcanonical initial state with $n_{0}=0$ (circles) and $n_{0}=10$
614: (diamonds) and $n_{0}=30$ (crosses) are
615: compared for a fixed rapidity parameter $|z|=2$ and $r=-22 \ldots
616: 30$. The lines serve as a guide for the eye.}
617: \label{fn}
618: \end{figure}
619: With increasing value of $n_{0}$ the distributions
620: become broader. They develop a slightly asymmetric shape with 
621: higher peaks at negative values of $r$ compared to those at positive
622: $r$ values. Between these dominant peaks the probability still
623: displays pronounced variations.\\
624: For a harmonic oscillator, the microcanonical Crooks theorem reduces
625: to the relation $q^{\text{mc}}_{r}(n) = q^{\text{mc}}_{-r}(n+r)$. One can show that
626: this symmetry is fulfilled by the probabilities $q^{\text{mc}}_{r}(n)$
627: given by eq.~(\ref{mcq}).
628: \subsubsection{Canonical initial state}
629: For a canonical density matrix 
630: \be
631: \r(t_{0}) = (1-e^{-\b \hbar \o }) e^{-\b \hbar \o a^{+} a}
632: \ee{cdm}
633:  the initial states are
634: distributed according to $p_{n} = e^{-\b
635:   \hbar \o n}/(1- e^{-\b \hbar \o})$. This allows one to perform the
636: sum over $n$ in the expression for the characteristic function
637: (\ref{Gho}) in
638: terms of the generating function of the Laguerre polynomials, cf. \cite{RG}
639: yielding the expression
640: \begin{widetext}
641: \be
642: G^{\text{c}}_{t_{f},t_{0}}(\b,u) = \exp \left \{ \frac{iu
643:     |f(t_{f})|^{2}}{\hbar \o} +\left(e^{iu\hbar
644:       \o}-1 \right )|z|^{2} - 4|z|^{2} \frac{\sin^{2}(\hbar \o u/2)}{e^{\b \hbar
645:       \o}-1}   \right \}. 
646: \ee{Gca}
647: \end{widetext}
648: Putting $u=i\b$ one finds that the two terms in the
649: exponent which are proportional to $|z|^{2}$ cancel each other, such
650: that one obtains 
651: \be
652: \langle e^{\b w} \rangle = G^{\text{c}}_{t_{f},t_{0}}(\b,i\b) =
653: \exp\left \{-\b
654:   |f(t_{f}|^{2}/(\hbar \o)\right \}. 
655: \ee{Jho}
656:  The free energy difference of two oscillators with Hamiltonians
657:  $H(t_{0}) =\hbar \o a^{+}a$ and  $H(t_{f}) =\hbar \o a^{+}a
658:  +f^{*}(t_{f}) a +f(t) a^{+})$ each one staying in a canonical state at the
659:   temperature $\b$ is given by $\D F = F(t_{f}) - F(t_{0}) =
660:  |f(t_{f})|^{2}/(\hbar \o)$ in accordance with Jarzynski's work theorem.
661: 
662: The probability $q^{\text{c}}_{r}(\tilde{\b})$ to find the work $w= \hbar \o r
663: +|f(t_{f})|^{2}/(\hbar \o)$ if the system starts in a canonical state becomes
664: \be
665: \begin{split}
666: q^{\text{c}}_{r}(\tilde{\b})& = e^{-|z|^{2}} \left (1 - e^{-\tilde{\b}
667:   } \right )\sum_{n=0}^{\infty} \sum_{k=0}^{n} \sum_{l=0}^{\min
668:   (k+r,2k)} 
669:   (-1)^{l} e^{-\tilde{\b}  n} \\
670: &\quad \times \frac{|z|^{2(2k+r-l)}}{(k+r-l)!
671:     k!} \binom{n}{k}
672: \binom{2k}{l}
673: \end{split}
674: \ee{qc}
675: where $\tilde{\b} = \b \hbar \o$ denotes the inverse dimensionless
676: temperature of the initial state.
677: The expression for $q^{c}_{r}(\tilde{\b})$ can be further simplified to read
678: \be
679: q^{\text{c}}_{r}(\tilde{\b})= e^{-|z|^{2} \coth
680:   (\tilde{\b}/2)} e^{\tilde{\b}r/2} I_{r}\left (\frac{|z|^{2}}{\sinh
681:     \tilde{\b}/2} \right )
682: %\sum_{m=0}^{\infty}  
683: %    \frac{|z|^{2m}}{m!}  I_{|m-r|}\left ( \frac{2|z|^{2}}{e^{\tilde{\b}
684: %       }-1}\right )
685: \ee{qc2}
686: where $I_{\n}(x)$ denotes the modified Bessel function of first kind of
687: order $\n$ \cite{RG}. For details of the derivation see Appendix~\ref{C}.
688: 
689: Note that the following detailed balance like symmetry relation exists,
690: \be
691: q^{\text{c}}_{-r}(\tilde{\b}) = e^{-\tilde{\b}r} q^{\text{c}}_{r}(\tilde{\b}),
692: \ee{qrmr}
693: relating the occurence of 
694: positive and negative work.   
695: In Fig.~\ref{fcz} the $z$ dependence of $q^{c}_{r}(\tilde{\b})$ for
696: $\tilde{\b}=\ln(4/3)$ is compared for a few small values of $r$. One finds
697: that due to the average over the canonical initial distribution 
698: the multipeaked structure of the microcanonical distribution as a
699: function of the rapidity parameter $|z|$ disappears and only
700: a single peak remains for each value of $r$.
701: The temperature dependence of the work distribution is illustrated
702: in Fig.~\ref{fcb}.
703: \begin{figure}
704: \includegraphics[width=8cm]{Fig3.eps}
705: \caption{The probabilities $q^{\text{c}}_{r}(\tilde{\b}=\ln(4/3))$ for a
706: canonical initial state are displayed for $r=-3
707: \ldots 4$ as a function of the parameter $z$. For the sake of
708: comparability the dimensionless
709: inverse temperature is chosen such that the average energy in the
710: initial state coincides with the energy $3 \hbar \omega$ 
711: of the microcanonical state in Fig.~\ref{fz} (b).   }
712: \label{fcz}
713: \end{figure}
714: \begin{figure}
715: \includegraphics[width=8cm]{Fig4.eps}
716: \caption{(Color online) The probabilities $q^{\text{c}}_{r}(\tilde{\b})$ for a
717: canonical initial state are displayed as functions of  $r$ 
718: for  $|z|=2$ and different values of the dimensionless inverse
719: temperature $\tilde{\b} = 0.5$ (boxes), $1$ (circles) and $5$
720: (crosses). The lines serve as a guide for the eye.}  
721: \label{fcb}
722: \end{figure}
723: Finally, we verify the validity of the Tasaki-Crooks theorem (\ref{TC})
724: for a driven oscillator. For this purpose we
725: consider the probability density $p_{t_{0},t_{f}}(-w)$ for the time
726: reversed protocol. Since the absolute values of the rapidity parameters
727: coincide for the original and the time reversed protocol the
728: probability density of work for the reversed protocols becomes
729: \be
730: p_{t_{0},t_{f}}(-w) = \sum_{r=-\infty}^{\infty} q_{r} \d\left (-w-(\hbar \o r
731: -\frac{|f(t_{f})|^{2}}{\hbar \o}) \right)\; ,
732: \ee{p0f}
733: where we took into account the overall shift of the spectrum by the
734: reversed protocol as
735: $-|f(t_{f})|/(\hbar \o)$.
736: Multiplying both sides of eq.~(\ref{p0f}) with $\exp \left \{ -\b( \D
737:   F -w) \right \}=\exp \left \{ -\b( |f(t_{f})|^{2}/(\hbar \o) -w)
738: \right \} $  one obtains 
739: \be
740: \begin{split}
741: e^{  -\b( \D
742:   F -w) } p_{t_{0},t_{f}}(-w) &= \sum_{r} e^{ -\b
743:   \left (
744:   |f(t_{f})|^{2}/(\hbar \o)  -w \right )}
745: q^{\text{c}}_{r}(\tilde{\b}) \\
746: &\quad \times \d\left (-w-(\hbar \o r 
747: -\frac{|f(t_{f})|^{2}}{\hbar \o}) \right)\\
748: &=\sum_{r} e^{\tilde{\b}r} q^{c}_{r}(\b)\\
749: &\quad \times \d \left (w +(\hbar
750:   \o r -\frac{|f(t_{f})|^{2}}{\hbar \o}) \right )\\
751: &= p_{t_{f},t_{0}}(w),
752: \end{split}
753: \ee{TCho}
754: in accordance with the Tasaki-Crooks theorem (\ref{TC}).     
755:  
756: 
757: 
758: \subsubsection{Coherent initial state}
759: An oscillator prepared in a coherent state $|\a \rangle$ is described
760: by the density matrix
761: \be
762: \r(t_{0}) = |\alpha \rangle \langle \alpha |
763: \ee{rcs}
764: where 
765: \be
766: |\a \rangle = e^{\a a^{+} + \a a} |0 \rangle
767: \ee{acs}
768: and $|0 \rangle $ is the normalized ground state of the oscillator
769: satisfying $a |0 \rangle =0$.
770: Note that the coherent state density matrix does {\it not} commute with the
771: Hamiltonian $H(t_{0})$. The measurement of $H(t_{0})$ modifies the
772: coherent state (\ref{rcs})  by
773: projecting it onto the eigenstates $|n \rangle = (a^{+})^{n}
774: /\sqrt{n!} |0 \rangle$ of this Hamiltonian leading to 
775: \be
776: \bar{\r}(t_{0}) = e^{-|\a|^{2}} \sum_{n} \frac{|\a|^{2n}}{n!} |n
777: \rangle \langle n |
778: \ee{brcs}
779: This implies a Poissonian distribution of the respective  
780: energy eigenvalues $\hbar \o n$ 
781: \be
782: p^{\text{cs}}_{n} = \frac{|\a|^{2n}}{n!} e^{-|\a|^{2}}, 
783: \ee{pcs}
784: which yields for the characteristic function of work (\ref{Gho}) a
785: closed expression of the form
786: \be
787: \begin{split}
788: G^{\text{cs}}_{t_{f},t_{0}}(\a,u) &= \exp\left
789:   \{\frac{iu|f(t_{f})|^{2}}{\hbar \o} +|z|^{2} \left(e^{i\hbar \o u}-1
790:   \right ) \right \}\\ 
791: & \quad \times J_{0}\left (4\left |\a z \sin 
792:   \frac{\hbar \o u}{2} \right | \right ) 
793: \end{split}
794: \ee{Gcs}  
795: where $J_{0}(x)$ is the Bessel function of order zero, cf. Ref.~\cite{RG}.
796: For the probability $q^{\text{cs}}_{r}(\a)$ of work one obtains with
797: eq.~(\ref{q})
798: \begin{widetext}
799: \be
800: q^{\text{cs}}_{r}(\a) = \;e^{-|z|^{2}} \sum_{m=0}^{\infty} 
801:   \frac{|z|^{2m} \:|\a z|^{2|m-r|}}{m! \left( |m-r|!\right)^{2}}
802:   {}_{1}F_{2}
803:   \left ( |m-r|+\frac{1}{2};|m-r|+1,2|m-r|+1;-4|\a z|^{2}\right )
804: \ee{qcs}
805: \end{widetext}
806: where $_{1}F_{2}(a;b,x;x)$ denotes a generalized hypergeometric
807: function \cite{RG}. For details of the derivation see the Appendix~\ref{D}. 
808:  
809: The dependence of  the probabilities $q^{\text{cs}}_{r}(\a)$ on the
810: rapidity parameter $|z|$ is illustrated in Fig.~\ref{fcsz} for $r$
811: values ranging from $-10$ to $20$. 
812: \begin{figure}
813: \includegraphics[width=8cm]{Fig5.eps}
814: \caption{The probabilities $q^{\text{cs}}_{r}(\a)$ for a
815: coherent state with parameter $|\a|=3$ are displayed 
816: for $r = -10 \ldots 20$ as functions of  $z$.}  
817: \label{fcsz}
818: \end{figure}
819: \begin{figure}
820: \includegraphics[width=8cm]{Fig7.eps}
821: \hfill
822: \includegraphics[width=8cm]{Fig6.eps}
823: \caption{(Color online) The distribution of work performed on an
824:   oscillator which 
825:   initially is prepared in a coherent state $|\a \rangle$ for
826:   different values of $\a$ in panel (a) and of the rapidity parameter
827:   $z$ in panel (b).
828: In panel (a) the rapidity parameter has the value $|z|=2$. In panel
829: (b) the coherent state parameter has the value  
830:  $|\a|^{2} =1$. }
831: \label{f2}
832: \end{figure}
833: Increasing values of $z$ lead to a broadening of the distribution
834: and also to a shift towards larger values of $r$, see also panel (b)
835: of Fig.~\ref{f2}.
836: This is in accordance with eq. (\ref{aw}) and (\ref{ww})
837: for the first two cumulants of the work, which both increase with
838: $|z|^{2}$.
839: %\begin{figure}
840: %\includegraphics[width=8cm]{Fig7.eps}
841: %\caption{The distribution of work performed on a harmonic oscillator
842: %  which is
843: %  initially prepared in a coherent state $|\a\rangle$, is displayed for
844: %  different
845: %values of the mean excitation $|\a|^2$. Different phases of $\a$ do
846: %not matter. The force protocols of the three cases are characterized
847: %by the parameter $|z| = 1$, see eqs. (\ref{FG}) and (\ref{z}). Details
848: %of the protocol which leave this parameter unchanged have no influence
849: %on the weight factors $q_n$.
850: %In accordance with eq. (\ref{ww}) the distribution becomes broader for
851: %larger values of $\langle a^{+}a \rangle_{0} =|\a|^2$.}
852: %\label{f1}
853: %\end{figure}
854: Panel (a) of Fig.~\ref{f2} shows the dependence of the probabilities
855: $q^{\text{cs}}_{r}(\a)$ on the parameters $\a$. Increasing
856: $\a$ also leads to a broadening of the work
857: distribution without influencing its mean value, cf. also eq. (\ref{aw}). 
858: 
859: %In Fig.~\ref{f3} the work distributions are compared for a canonical
860: %and a coherent initial state both having the same first two
861: %cumulants. Yet the two distributions are quite different: The work
862: %distribution for the canonical initial state is unimodal 
863: %resembling a Gaussian distribution whereas the coherent state leads to
864: %a bimodal slightly asymmetric distribution. 
865:  
866:   
867: \begin{figure}
868: \includegraphics[width=8cm]{Fig8.eps}
869: \caption{(color online) The distribution of work is compared for a
870:   canonical and a  
871:   coherent initial state subject to the same force protocol with
872:   rapidity parameter $z=2$. With $\b \hbar \o  =0.1$ and  $|\a|^2\approx
873:   9.51$ the expectation values of the energies agree in the two initial
874:   states such that according to eqs.~(\ref{aw}) and (\ref{ww}) the
875:   first and second moments of the work also coincide. Still the
876:   distributions of work grossly differ from each other. }
877: \label{f3}
878: \end{figure}  
879: In Fig.~\ref{f3} the probabilities $q_r$ are depicted for different
880: initial states. In one case the oscillator is initially prepared in a
881: canonical state at inverse dimensionless temperature $\tilde{\b}=\b
882: \hbar \o =0.1$. 
883: In the other case, the oscillator stays in a coherent state
884: $|\a\rangle$, where the absolute value of $|\a|$ is chosen such that
885: the mean excitation number is the same for both states, i.e. $|\a|^2 =
886: \exp\{ -\b \hbar \o\} / (1- \exp\{ -\b \hbar \o\} )$. For $\b \hbar \o
887: =0.1$ one finds $|\a|^2 \approx 9.51$. The two
888: oscillators then are subjected to protocols with the same
889: rapidity parameter 
890: $|z|=2$. According to eqs. (\ref{aw}) and (\ref{ww}) the first two
891: moments of the work performed on the oscillator coincide. Yet the
892: distribution of weight factors $q^{\text{c}}_r(\tilde{\b})$ and
893: $q^{\text{cs}}_r(\a)$ distinctly differ. Whereas the
894: distribution is pronouncedly bimodal in case of the coherent state, it
895: is unimodal for the canonical state. The weight factors
896: $q^{\text{c}}_r(\tilde{\b})$ almost 
897: perfectly fall onto the Gaussian probability density which has the
898: same first two
899: moments as the discrete distribution given by $q_r$.   
900: 
901: 
902: 
903:  
904: 
905: \section{Conclusions}
906: In this work we studied the statistics of work performed on an
907: externally driven quantum mechanical oscillator by means of a
908: correlation function expression of the work. We demonstrated that this
909: particular expression indeed always represents a proper characteristic
910: function of a random variable, which is the performed work in the present
911: context. The proof given here is based on Bochner's theorem. It holds
912: for general 
913: quantum mechanical systems, not only for harmonic oscillators.
914: 
915: The considered force linearly couples to the position and momentum of
916: the oscillator. It may describe the influence of an electric
917: field on charged particles in a parabolic trap or the external forcing
918: of a single electromagnetic cavity mode. For this
919: type of additive forcing, the frequency of the oscillator remains unchanged and
920: therefore the level spacing of the eigenvalues of the Hamiltonian is
921: not influenced by the force. The spectrum is only shifted as a whole.
922: As a consequence the work performed on the oscillator is, as a positive
923: or negative integer of the level spacing, 
924: a discrete random variable. We determined the first few cumulents of
925: the work for arbitrary force protocols and initial states. A
926: complementary study for a parametrically forced oscillator was
927: recently 
928: performed by Deffner and Lutz \cite{DL}.     
929: 
930: It turns out that for the harmonic oscillator 
931: the statistics of work depends on the force
932: protocol $\{f(t)\}_{t_{f},t_{0}}$ only through two real parameters, 
933: which are (i) the shift of the spectrum, given by $L(t_{f})=
934: |f(t_{f})|^{2}/(\hbar \o)$, 
935: and (ii) the absolute value of the dimensionless quantity $z =
936: \i_{t_{0}}^{t_{f}}\dot{f}(s) \exp \{i \o s \}$. 
937: This parameter vanishes for all quasi-static processes
938: and therefore presents a measure of the rapidity of the force
939: protocol.
940: While the presence of $L(t_{f})$ only causes an overall shift of the
941: possible values of the work, the rapidity parameter $|z|$ also
942: influences its distribution. Typically, the distributions move towards 
943: larger values of work $w$ and become broader with increasing rapidity $|z|$.
944: 
945: We also demonstrated that different initial states of the system such
946: as microcanonical, canonical or coherent states, have a large influence on
947: the work statistics. We further note that two different initial
948: density matrices with the 
949: same diagonal elements with respect to the energy eigenbasis of the
950: Hamiltonian $H(t_{0})$ lead to identical work distributions
951: even though the two density matrices may be very different in other
952: respects. For example, the coherent pure state $|\a\rangle \langle \a|$
953: and the mixed state $\exp\{-|\a|^{2 n}\}\sum_{n} |\a|^{2}/n! |n\rangle
954: \langle n| $ cannot be distinguished by means of their respective work
955: statistics. This statistics is also insensitive to the phase of a
956: coherent state.  
957: 
958: \acknowledgments
959: This work has been supported by the Deutsche Forschungsgemeinschaft
960: via the Collaborative Research Centre SFB-486, project A10. Financial
961: support of the German Excellence Initiative via the {\it Nanosystems
962:   Initiative Munich} (NIM) is gratefulle acknowledged as well. 
963: 
964: 
965: 
966: \appendix
967: \section{Proof of the properties of $G_{t_f,t_0}(u,v)$}
968: \label{AA}
969: We prove that the conditions of Bochner's theorem are fulfilled, and
970: consequently $G_{t_{f},t_{0}}(u)$ is a proper characteristic function.
971:   
972: Proof of property (i): $G_{t_{f},t_{0}}(u)$ is a continuous function
973: of $u$. The Hamiltonian operators at the two times of measurement
974: $t_{0}$ and $t_{f}$ are selfadjoint operators. According to the
975: theorem of Stone \cite{Yosida}, each of the exponential operators $\exp \left
976:   \{ -iu H(t_{0})\right \}$ and  $\exp \left
977:   \{ iu H_{H}(t_{f})\right \}$ forms a strongly continuous one-parameter 
978: group of unitary operators with parameter $u$. As the trace of a
979: product of two strongly continuous operator valued functions of $u$
980: with the density operator $\bar{\r}(t_{0})$, which is a trace class
981: operator and independent of $u$, the
982: characteristic function (\ref{GC}) is a continuous function of $u$.
983: 
984: Proof of property (ii): $G_{t_{f},t_{0}}(u)$ is a positive definite
985: function of $u$.
986: Using the cyclic invariance of the trace and the  fact that $H(t_{0})$
987: and $\bar{\r}(t_{0})$ commute with each other, we can rewrite the left
988: hand side of the inequality (\ref{pd}) as
989: \begin{widetext}
990: \be
991: %\begin{split}
992: \sum_{i,j}^{n} G_{t_{f},t_{0}}(u_{i}-u_{j}) z^{*}_{i} z_{j} =
993: \sum_{i,j}^{n} \Tr \:e^{i(u_{i}-u_{j})H_{H}(t_{f})} \:
994: e^{-i(u_{i}-u_{j})H(t_{0})} \bar{\r}(t_{0}) z_{i}^{*} z_{j}
995: = \Tr A^{+} A \:\bar{\r}(t_{0}) \geq 0\;,  
996: %\end{split}
997: \ee{pos}   
998: \end{widetext}
999: where 
1000: \be
1001: A= \sum_{i}z_{i}^{n} e^{-iu_{i}H_{H}(t_{f})} e^{iu_{i} H(t_{0})}
1002: \ee{A}
1003: is a bounded operator and $A^{+}$ its adjoint. The last inequality in 
1004: (\ref{pos}) immediately 
1005: follows with the positivity of $A^{+}A$ and of the density matrix
1006: $\bar{\r}(t_{0})$.   
1007: 
1008: Proof of property (iii): $G_{t_{f},t_{0}}(0) =1$. 
1009: For $u=0$ the exponential operators $\exp \left
1010:   \{ -iu H(t_{0})\right \}$ and  $\exp \left
1011:   \{ iu H_{H}(t_{f})\right \}$ become unity. The trace over the
1012: density matrix $\bar{\r}(t_{0})$ reduces by means of
1013: eqs. (\ref{br}), (\ref{P1}) to the trace of the initial
1014: density matrix $\r(t_{0})$, which is one.
1015: \section{The matrix element $\langle n| \exp \{iu H_H(t_f)\} |n\rangle
1016:   $}
1017: \label{B} 
1018: The total time rate of change of the Hamiltonian $H_H(t)$ coincides
1019: with its partial derivative with respect to the time which for the
1020: driven oscillator becomes, cf. eq. (\ref{hoH}),
1021: \be
1022: \frac{d H_H(t)}{dt} = \dot{f}^* (t) a_H(t) - \dot{f}(t) a^+_H(t),
1023: \ee{dH}
1024: where $a_H(t)$ and $a^+_H(t)$ denote annihilation and creation
1025: operators, respectively, in the Heisenberg picture, which are given by
1026: \be
1027: \begin{split}
1028: a_H(t) &=   e^{-i \o t} a -\frac{i}{\hbar} \i_0^t ds e^{-i\o(t-s)} f(s)\\
1029: a^+_H(t) &=  e^{i \o t} a^+ +\frac{i}{\hbar} \i_0^t ds e^{i\o(t-s)} f^*(s)  
1030: \end{split}
1031: \ee{aH}
1032: This yields for $H_H(t_f)$
1033: \be
1034: H_H(t_f) = \hbar \o a^+ a + B^*(t_{f})a + B(t_{f}) a^+ + C(t),
1035: \ee{HtH}
1036: where
1037: \be
1038: \begin{split}
1039: B(t_{f}) &= \i_0^{t_{f}} ds \dot{f}(s) e^{i \o s}\\
1040: C(t_{f})& = \frac{i}{\hbar} \i_0^{t_{f}} ds \i_0^s ds' \left [ \dot{f}(s)
1041:   f^*(s') e^{i\o(s-s')} \right .\\
1042: &\quad \left . - \dot{f}^*(s) f(s') e^{-i\o (s-s')} \right ].
1043: \end{split}
1044: \ee{FG}
1045: The unitary operator
1046: \be
1047: V= e^{z a^+ -z^* a}
1048: \ee{V}
1049: with 
1050: \be
1051: z = \frac{B(t_{f})}{\hbar \o}
1052: \ee{z}
1053: transforms $H_H(t)$ into
1054: \be
1055: V H_H(t_{f}) V^+ = \hbar \o a^+ a + L(t_{f}),
1056: \ee{VHV}
1057: where 
1058: \be
1059: L(t_{f}) = C(t_{f})-\frac{|B(t_{f})|^2}{\hbar \o}
1060: =\frac{|f(t_{f})|^2}{\hbar \o} .
1061: \ee{L}
1062: Note that $V$ induces a shift of the creation and annihilation
1063: operators
1064: \be
1065: \begin{split}
1066: V a V^+ = a - z, \quad
1067: V a^+ V^+ = a^+ - z^*
1068: \end{split}
1069: \ee{Va}
1070: and further note that, when acting on the groundstate $|0\rangle$ with
1071: $a|0\rangle =0$, the operator $V$ yields the coherent state $|z\rangle$, i.e.
1072: \be
1073: V|0\rangle = | z \rangle.
1074: \ee{V0}
1075: One finds with these properties
1076: %\begin{widetext}
1077: %\be
1078: %\begin{split}
1079: \begin{align}{\allowdisplaybreaks}
1080: \langle n | e^{iu H_H(t_{f})}|n\rangle &= \frac{1}{n!} \langle
1081: z|(a-z)^n \nonumber \allowdisplaybreaks\\
1082: &\quad\times  e^{iu\hbar \o a^+a +i u L(t_{f})} (a^+-z^*)^n |
1083: z \rangle \nonumber \allowdisplaybreaks \\
1084: &=\frac{1}{n!} e^{iu L(t_{f})} \frac{\partial^{2n}}{\partial x^n \partial
1085:   y^n} \nonumber \allowdisplaybreaks \\
1086: &\quad \langle z| e^{x(a-z)} e^{iu \hbar \o a^+a} e^{y(a^+-z^*)} | z
1087: \rangle|_{x=y=0}
1088: \label{nz}
1089: \end{align}
1090: %\end{split}
1091: %\ee{nz}
1092: %\end{widetext}
1093: Here we have introduced the auxiliary variables $x$ and $y$ which allow
1094: to represent the $n$th powers of shifted creation and annihilation
1095: operators by derivatives of respective order. The scalar
1096: function $e^{-i(xz + y z^*)}$ can be taken out of the scalar product
1097: and the remaining operator can be brought into normal order. It then
1098: becomes \cite{Wi}
1099: %\begin{widetext}
1100: \be
1101: \begin{split}
1102: e^{xa} e^{iu \hbar \o a^+ a} e^{ya^+} &= \mathcal{N}\left \{
1103:   \exp\left [(e^{iu\hbar \o} -1) a^+a \right . \right .\\
1104: &\quad \left . \left . + e^{iu\hbar\o} \left (x a + y
1105:       a^+ + x y \right ) \right ] \right \}\; ,  
1106: \end{split}
1107: \ee{N} 
1108: %\end{widetext}
1109: where  under the normal ordering operator $\mathcal{N}$ all
1110: creation operators stand left of the annihilation operators. The
1111: matrix element with respect to the coherent state $|z\rangle$ can be
1112: read off, yielding,
1113: \begin{widetext}
1114: \be
1115: \begin{split}
1116: \langle n | e^{iu H_H(t_{f})} |n\rangle =&\: \frac{1}{n!}  e^{iu L(t_{f})} \exp
1117: \left \{ \left ( e^{iu \hbar \o}-1 \right ) |z|^2 \right \}
1118: \frac{\partial^{2n}}{\partial x^n \partial y^n} \exp \left \{\left 
1119:     (e^{iu \hbar \o} -1 \right )(xz +y z^*) +e^{iu\hbar \o} xy \right
1120: \}|_{x=y=0}\\
1121: =&\:\frac{1}{n!}  e^{iu L(t_{f})} \exp
1122: \left \{ \left ( e^{iu \hbar \o}-1 \right ) |z|^2 \right \} 
1123: \frac{\partial^n}{\partial y^n} \left [ \left (e^{iu\hbar \o} -1
1124:   \right )z + e^{iu \h \o} y \right ]^n \exp \left \{ \left ( e^{iu\hbar
1125:       \o}-1 \right ) \left ( |z|^2 + y z^* \right ) \right \}|_{y=0}\\
1126: =& \: e^{iu |f(t_{f})|^{2}/(\hbar \o)} \exp
1127: \left \{ \left ( e^{iu \hbar \o} -1 \right ) |z|^2 \right \} 
1128: \sum_{k=0}^n \binom{n}{k} \frac{|z|^{2(n-k)}}{(n-k)!}
1129: e^{iu\hbar \o k} \left ( e^{iu\hbar \o} -1 \right )^{2(n-k)} \; .
1130: \end{split}
1131: \ee{nxy}
1132: \end{widetext}
1133: \section{Work distribution for a canonical initial state}
1134: \label{C}
1135: To determine the expression (\ref{qc2}) for the work
1136: distribution $q^{\text{c}}(\tilde{\b})$ we start from 
1137: the general expression given in the first line of 
1138: eq. (\ref{q}). Interchanging 
1139: the summation over the indices $n$ and $k$ we obtain
1140: %\begin{widetext}
1141: %\be
1142: %\begin{split}
1143: \begin{align}
1144: q^{c}_{r}(\tilde{\b})& = e^{-|z|^{2}}\sum_{m,k=0}^{\infty} \sum_{l=0}^{2k}
1145: (-1)^{l}\frac{|z|^{2(k+m)}}{m!\:k!} \binom{2k}{l}\allowdisplaybreaks
1146: \nonumber \\
1147: &\quad \times \d_{l+m,k+r}
1148: \sum_{n=k}^{\infty} \frac{e^{-\tilde{\b} n}}{1-e^{-\tilde{\b} }}
1149: \binom{n}{k} \allowdisplaybreaks\nonumber \\
1150: &\stackrel{(1)}{=}e^{-|z|^{2}} \sum_{m,k=0}^{\infty} \sum_{l=0}^{2k}
1151: (-1)^{l}\frac{|z|^{2(k+m)}}{m!\:k!} \binom{2k}{l}\nonumber \allowdisplaybreaks\\
1152: &\quad \times  \left
1153:   ( \frac{1}{e^{\tilde{\b} }-1}\right )^{k}
1154: \d_{l+m,k+r}\allowdisplaybreaks \nonumber \\
1155: &\stackrel{(2)}{=}(-1)^{r}e^{-|z|^{2}}
1156: \sum_{m=0}^{\infty}\frac{(-|z|^{2})^{m}}{m!}\nonumber \allowdisplaybreaks\\
1157: &\quad \times \sum_{k=|m-r|}^{\infty}  
1158: \frac{\left (-|z|^{2}/(e^{\tilde{\b}}-1) \right )^{k}}{k!}
1159: \binom{2k}{k+r-m}\allowdisplaybreaks\nonumber \\
1160: &\stackrel{(3)}{=}(-1)^{r}e^{-|z|^{2}}
1161: \sum_{m=0}^{\infty}\frac{(-|z|^{2})^{m}}{m!}
1162: e^{-2|z|^{2}/(e^{\tilde{\b}}-1)} \nonumber \allowdisplaybreaks\\
1163: &\quad \times I_{|m-r|}\left (- \frac{2
1164:     |z|^{2}}{e^{\tilde{\b}}-1}\right )\nonumber \allowdisplaybreaks\\ 
1165: & \stackrel{(4)}{=}e^{-|z|^{2} \coth (\tilde{\b}/2)}
1166: \sum_{m=0}^{\infty} \frac{|z|^{2m}}{m!} I_{|m-r|} \left (\frac{2
1167:   |z|^{2}}{e^{\tilde{\b}}-1} \right )\nonumber \allowdisplaybreaks\\
1168: & \stackrel{(5)}{=} e^{-|z|^{2} \coth (\tilde{\b}/2)} \left \{ 
1169:  \sum_{m=0}^{\infty} \frac{|z|^{2}}{m!} I_{r-m} \left (\frac{2
1170:   |z|^{2}}{e^{\tilde{\b}}-1} \right ) \right .\nonumber\allowdisplaybreaks\\
1171: &\quad \left . +\sum_{m=r+1}^{\infty} 
1172: \frac{|z|^{2}}{m!}\left [I_{m-r} \left (\frac{2
1173:   |z|^{2}}{e^{\tilde{\b}}-1} \right ) \right . \right
1174: .\nonumber\allowdisplaybreaks\\ 
1175: &\left . \left .  \quad -I_{r-m} \left (\frac{2
1176:   |z|^{2}}{e^{\tilde{\b}}-1} \right ) \right] \right \}
1177: \allowdisplaybreaks\nonumber\\ 
1178: & %\stackrel{(6)}{=}  
1179: = e^{-|z|^{2} \coth (\tilde{\b}/2)}
1180: e^{\tilde{\b}r/2} I_{r}\left(\frac{|z|^{2}}{\sinh (\tilde{\b}/2)} \right).
1181: \label{qcp}
1182: \end{align}
1183: %\end{split}
1184: %\ee{qcp}   
1185: %\be
1186: %q \stackrel{(5)}{=} e^{-|z|^{2} \coth (\tilde{\b}/2)} \left \{ 
1187: %\sum_{m=0}^{\infty} \frac{|z|^{2}}{m!} I_{r-m} \left (\frac{2
1188: %  |z|^{2}}{e^{\tilde{\b}}-1} \right ) + %e^{-| q| () /2} 
1189: %\sum_{m=r+1}^{\infty} 
1190: %\frac{|z|^{2}}{m!}\left [I_{m-r} \left (\frac{2
1191: %  |z|^{2}}{e^{\tilde{\b}}-1} \right )-I_{r-m} \left (\frac{2
1192: %  |z|^{2}}{e^{\tilde{\b}}-1} \right ) \right] \right \} 
1193: %\ee{test}
1194: %\end{widetext} 
1195: In the first step ($\stackrel{(1)}{=}$) we  performed the sum on $n$
1196: according to 
1197: \be
1198: \sum_{n=k}^{\infty} \frac{x^{k}}{1-x} \binom{n}{k} = \left (
1199:   \frac{x}{1-x}\right)^{k},
1200: \ee{s1} 
1201: cf. Ref.~\cite{PBM}, 5.2.11.3. In the second step $\stackrel{(2)}{=}$
1202: the Kronecker delta is used 
1203: to perform the sum over k. The third step $\stackrel{(3)}{=}$ is based
1204: on the relation 
1205: \be
1206: \sum_{k=|l|}^{\infty} \frac{x^{k}}{k!} \binom{2k}{k+l} = e^{2x}
1207: I_{|l|}(2x)
1208: \ee{s2}
1209: valid for integer $l$. Here $I_{\n}(x)$ denotes the modified Bessel
1210: function of the first kind of order $\n$. 
1211: With $I_{\n}(-x) =(-1)^{\n}I_{\n}(x)$ where
1212: $\n$ is an integer, we come to the right hand side of the equality 
1213: $\stackrel{(4)}{=}$. In the next step the sum on $m$ is rewritten. The
1214: term in the square brackets vanishes because $I_{\n}(x)$ is an even
1215: function of the order $\n$. The remaining sum can be performed by
1216: means of the identity
1217: \be
1218: \sum_{k=0}^{\infty} \frac{t^{k}}{k!} I_{\n -k} (x) = \left ( \frac{2
1219:     t}{x} +1\right )^{\n/2} I_{\n} \left( \sqrt{x^{2}+2 t x}\right )\; ,
1220: \ee{sB}
1221: cf. \cite{PBM} 5.8.3.1.
1222: This leads to the final 
1223: result given in eq. (\ref{qc2}).
1224: \section{Work distribution for a coherent initial state eq. (\ref{qcs})}
1225: \label{D}
1226: Starting from eq.~(\ref{q}) we may proceed in an analogous way as 
1227: in the case of a canonical
1228: initial state, cf. the Appendix~\ref{C}. 
1229: According to eq. (\ref{pcs}) 
1230: a Poissonian average over the binomial
1231: $\binom{n}{k}$ has to be performed
1232: instead of the geometric average in the first step of eq. (\ref{qcp}).
1233: This yields
1234: \be
1235: \sum_{n=k}^{\infty} \frac{|\a|^{2}}{k!} e^{-|\a|^{2}} \binom{n}{k}
1236: = \frac{|\a|^{2k}}{k!} \; .
1237: \ee{qcs1}
1238: Next the Kronnecker delta is used to perform the sum over $l$ leaving
1239: one with two  sums of which the inner one over k can be expressed in
1240: terms of a generalized hypergeometric function, \cite{RG}, to become
1241: \begin{widetext}
1242: \be
1243: \sum_{k=|m-r|}^{\infty} \frac{(-|\a z|^{2})^{k}}{(k!)^{2}}
1244: \binom{2k}{k+r-m} = \frac{(-|\a z|^{2})^{|m-r|}}{(|m-r|!)^{2}}\\
1245: {}_1F_{2} \left( |m-r|+\frac{1}{2};|m-r|+1,2|m-r|+1;-4|\a
1246:   z|^{2}\right) \; .
1247: \ee{qcs2}
1248: \end{widetext}
1249: This immediately leads to the expression in eq. (\ref{qcs}).
1250: \begin{thebibliography}{99}
1251: \bibitem{fwt} G.N. Bochkov, Yu.E. Kuzovlev, Sov. Phys. JETP {\bf 45},
1252:   125 (1977).
1253: \bibitem{ECM}
1254: D.J. Evans, E.G.D. Cohen, G.P. Morriss,
1255:   Phys. Rev. Lett. {\bf 71}, 2401 (1993).
1256: \bibitem{J} C. Jarzynski, Phys. Rev. Lett. {\bf 78}, 2690 (1997).
1257: \bibitem{J07} C. Jarzynsky, C. R. Physique {\bf 8}, 495 (2007).
1258: \bibitem{ex} F. Douarche, S. Ciliberto, A. Petrosyan, I. Rabbiosi,
1259:   Europhys. Lett. {\bf 70}, 593 (2005).
1260: \bibitem{BLR}
1261:  C. Bustamante, J. Liphardt,
1262:   F. Ritort, Physics Today {\bf 58} (7), 43 (2005).
1263: \bibitem{BHSB}
1264: V. Blickle,
1265:   T. Speck, L. Helden, U. Seifert, C. Bechinger, Phys. Rev. Lett. {\bf
1266:   96}, 070603 (2006).
1267: \bibitem{qm} H. Tasaki, cond-mat/0009244.
1268: \bibitem{M}
1269: S. Mukamel,
1270:   Phys. Rev. Lett. {\bf 90}, 170604 (2003). 
1271: \bibitem{RM}
1272: W. De Roeck, C. Maes,
1273:   Phys. Rev. E {\bf 69}, 026115 (2004).
1274: \bibitem{EM} M. Esposito, S. Mukamel, Phys. Rev. E {\bf 73}, 046129 (2006).
1275: \bibitem{TLH} P. Talkner, E. Lutz, P. H\"anggi, Phys. Rev. E {\bf 75},
1276:   050102(R) (2007).
1277: \bibitem{TH} P. Talkner, P. H\"anggi, J. Phys. A {\bf 40}, F569 (2007).
1278: \bibitem{TMH} P. Talkner, M. Morillo, P. H\"anggi, arXiv:0707.2307v2.
1279: \bibitem{DL} S. Deffner, E. Lutz, Phys. Rev. E {\bf 77}, 021128 (2008).
1280: \bibitem{Seif} U. Seifert, J. Phys. A: Math. Gen. {\bf 37}, L517 (2004).
1281: \bibitem{H} K. Husimi, Prog. Theor. Phys. {\bf 9}, 381 (1953).
1282: \bibitem{Boch} E. Lukacs, {\it Characteristic Functions}, Griffin,
1283:   London, 1970.
1284: \bibitem{CBK} B. Cleuren, C. Van den Broeck, R. Kawai,
1285:   Phys. Rev. Lett. {\bf 96}, 050601 (2006).
1286: \bibitem{Yosida} K. Yosida, Functional Analysis, Springer Verlag, Berlin, 1971.
1287: \bibitem{RG} L.S. Gradshteyn, I.M. Ryzhik, {\it Table of Integrals,
1288:     Series and Products}, Academic Press, San Diego (2000).
1289: \bibitem{vK} N.G. van Kampen, {\it Stochastic Processes in Physics and
1290:   Chemistry}, North Holland, Amsterdam, 1992.
1291: \bibitem{Wi} R.M. Wilcox, J. Math. Phys. {\bf 8}, 962 (1967).
1292: \bibitem{PBM} A.P. Prudnikov, Yu.A. Brychkov, O.I. Marichev, {\it
1293:     Integrals and Series}, Vol. 1, Gordon and Breach, New York, 1986.
1294: \end{thebibliography} 
1295: 
1296: \end{document}