61
Annals of Physics 281, 547607 (2000) The Theory of a General Quantum System Interacting with a Linear Dissipative System R. P. Feynman California Institute of Technology, Pasadena, California and F. L. Vernon, Jr. 1 Aerospace Corporation, El Segundo, California Received April 5, 1963 A formalism has been developed, using Feynman's space-time formulation of nonrelativistic quantum mechanics whereby the behavior of a system of interest, which is coupled to other external quantum systems, may be calculated in terms of its own variables only. It is shown that the effect of the external systems in such a formalism can always be included in a general class of functionals (influence functionals) of the coordinates of the system only. The proper- ties of influence functionals for general systems are examined. Then, specific forms of influence functionals representing the effect of definite and random classical forces, linear dissipative systems at finite temperatures, and combinations of these are analyzed in detail. The linear system analysis is first done for perfectly linear systems composed of combinations of harmonic oscillators, loss being introduced by continuous distributions of oscillators. Then approximately linear systems and restrictions necessary for the linear behavior are considered. Influence func- tionals for all linear systems are shown to have the same form in terms of their classical response functions. In addition, a fluctuation-dissipation theorem is derived relating temperature and dissipation of the linear system to a fluctuating classical potential acting on the system of interest which reduces to the NyquistJohnson relation for noise in the case of electric circuits. Sample calculations of transition probabilities for the spontaneous emission of an atom in free space and in a cavity are made. Finally, a theorem is proved showing that within the require- ments of linearity all sources of noise or quantum fluctuation introduced by maser-type amplification devices are accounted for by a classical calculation of the characteristics of the maser. 1963 Academic Press I. INTRODUCTION Many situations occur in quantum mechanics in which several systems are coupled together but one or more of them are not of primary interest. Problems in the theory of measurement and in statistical mechanics present good examples of doi:10.1006aphy.2000.6017, available online at http:www.idealibrary.com on 547 0003-491663 35.00 Copyright 1963 by Academic Press All rights of reproduction in any form reserved. Reprinted from Volume 24, pages 118173. 1 This report is based on a portion of a thesis submitted by F. L. Vernon, Jr. in partial fulfilment of the requirements for the degree of Doctor of Philosophy at the California Institute of Technology, 1959.

Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

  • Upload
    tjny699

  • View
    60

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

Annals of Physics 281, 547�607 (2000)

The Theory of a General Quantum System Interacting witha Linear Dissipative System

R. P. Feynman

California Institute of Technology, Pasadena, California

and

F. L. Vernon, Jr.1

Aerospace Corporation, El Segundo, California

Received April 5, 1963

A formalism has been developed, using Feynman's space-time formulation of nonrelativisticquantum mechanics whereby the behavior of a system of interest, which is coupled to otherexternal quantum systems, may be calculated in terms of its own variables only. It is shownthat the effect of the external systems in such a formalism can always be included in a generalclass of functionals (influence functionals) of the coordinates of the system only. The proper-ties of influence functionals for general systems are examined. Then, specific forms of influencefunctionals representing the effect of definite and random classical forces, linear dissipativesystems at finite temperatures, and combinations of these are analyzed in detail. The linearsystem analysis is first done for perfectly linear systems composed of combinations of harmonicoscillators, loss being introduced by continuous distributions of oscillators. Then approximatelylinear systems and restrictions necessary for the linear behavior are considered. Influence func-tionals for all linear systems are shown to have the same form in terms of their classical responsefunctions. In addition, a fluctuation-dissipation theorem is derived relating temperature anddissipation of the linear system to a fluctuating classical potential acting on the system ofinterest which reduces to the Nyquist�Johnson relation for noise in the case of electric circuits.Sample calculations of transition probabilities for the spontaneous emission of an atom in freespace and in a cavity are made. Finally, a theorem is proved showing that within the require-ments of linearity all sources of noise or quantum fluctuation introduced by maser-typeamplification devices are accounted for by a classical calculation of the characteristics of themaser. � 1963 Academic Press

I. INTRODUCTION

Many situations occur in quantum mechanics in which several systems arecoupled together but one or more of them are not of primary interest. Problems inthe theory of measurement and in statistical mechanics present good examples of

doi:10.1006�aphy.2000.6017, available online at http:��www.idealibrary.com on

5470003-4916�63 �35.00

Copyright � 1963 by Academic PressAll rights of reproduction in any form reserved.

Reprinted from Volume 24, pages 118�173.1 This report is based on a portion of a thesis submitted by F. L. Vernon, Jr. in partial fulfilment of

the requirements for the degree of Doctor of Philosophy at the California Institute of Technology, 1959.

Page 2: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601702 . By:SD . Date:31:03:00 . Time:14:44 LOP8M. V8.B. Page 01:01Codes: 3585 Signs: 2932 . Length: 46 pic 0 pts, 194 mm

such situations. Suppose, for instance, that the quantum behavior of a system is tobe investigated when it is coupled to one or more measuring instruments. Theinstruments in themselves are not of primary interest. However, their effects arethose of perturbing the characteristics of the system being observed. A moreconcrete example is the case of an atom in an excited state which interacts with theelectromagnetic field in a lossy cavity resonator. Because of the coupling there willbe energy exchange between the field and the atom until equilibrium is reached. If,however, the atom were not coupled to any external disturbances, it would simplyremain unperturbed in its original excited state. The cavity field, although not ofcentral interest to us, influences the behavior of the atom.

To make the discussion more definite, let us suppose there are two nonrelativisticquantum systems whose coordinates are represented in a general way by Q and X,as in Fig. 1, coupled together through some interaction potential which is a func-tion of the parameters of the two systems. It is desired to compute the expectationvalue of an observable which is a function of the Q variables only. As is wellknown, the complete problem can be analyzed by taking the Hamiltonian of thecomplete system, forming the wave equation as follows,

[H(Q)+H(X)+V(Q, X)] �(Q, X)=&\�

i+��t

�(Q, X),

and then finding its solution. In general, this is an extremely difficult problem. Inaddition, when this approach is used, it is not easy to see how to eliminate thecoordinates of X and include its effect in an equivalent way when making computa-tions on Q. A satisfactory method of formulating such problems as this in a generalway was made available by the introduction of the Lagrangian formulation of quantummechanics by Feynman. He applied the techniques afforded by this method extensivelyto studies in quantum electrodynamics. Thus, in a problem where several chargesparticles interact through the electromagnetic field, he found that it was possible toeliminate the coordinates of the field and recast the problem in terms of the coor-dinates of the particles alone. The effect of the field was included as a delayedinteraction between the particles [1, 2].

The central problem of this study is to develop a general formalism for findingall of the quantum effects of an environmental system (the interaction system) upona system of interest (the test system), to investigate the properties of this formalismand to draw conclusions about the quantum effects of specific interaction systemson the test system. Cases where the interaction system is composed of variouscombinations of linear systems and classical forces will be considered in detail. Forthe case in which the interaction system is linear, it will be found that parameters such

FIG. 1. General quantum systems Q and X coupled by a potential V(Q, X, t).

548 FEYNMAN AND VERNON

Page 3: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

as impedance, which characterize its classical behavior, are also important in deter-mining its quantum effect on the observed system. Since this linear system may includedissipation, the results have application in a study of irreversible statistical mechanics.

In Section II, after a brief discussion of the Lagrangian formulation of quantummechanics, a general formulation of the problem is made and certain functionals,called influence functionals, will be defined, which contain the effect of the interactionsystem (such as system X in Fig. 1) on the test system in terms of the coordinatesof the test system only. General properties of these functionals will be derived andtheir relationship to statistical mechanics will be discussed. To obtain more specificinformation about the properties of the formalism, we then specialize the discussionto cases where well-defined systems are involved. In Section III, the special cases areconsidered in which the interaction system is a definite classical force and a randomclassical force. In Section IV, the influence functionals for exactly linear systems atzero temperatures are derived and then extended to the case that the linear systemsare driven by classical forces. In addition, the effect of finite temperatures of linearsystems is considered. Then, in Section V, the unobserved systems are againassumed to be general but weakly coupled to the observed system. Within theapproximation of weak coupling these general systems also behave as if they werelinear. Then finally in Section VI, the results of the analysis are used to prove ageneral result concerning maser noise.

It is to be emphasized that although we shall talk of general test and interactionsystems, the Lagrangian formulation is restricted to cases involving momentum orcoordinate operators. Therefore, strictly speaking, systems in which the spin is ofimportance are not covered by this analysis. However, this has no bearing on theresults since their nature is such that their extension to the case where spins areimportant can be inferred.

An equivalent approach can be made to the problem using the Hamiltonianformulation of quantum mechanics by making use of the ordered operator calculusdeveloped by Feynman [3]. This approach has been used to some extent by Fano [4]and has been developed further by Hellwarth [5].2 Some advantages of this method arethat many results may be obtained more simply than by the Lagrangian method andnonclassical concepts such as spin enter the formalism naturally. However, the physicalsignificance of the functions being dealt with are often clearer in the Lagrangian method.

II. GENERAL FORMULATION�INFLUENCE FUNCTIONAL

A. Lagrangian Formulation of Quantum Mechanics

We shall begin the discussion with a brief introduction to the Lagrangian orspace-time approach to quantum mechanics and the formal way in which one may

549INTERACTION OF SYSTEMS

2 Many of the results obtained in this work have also been btained by him using ordered operatortechniques.

Page 4: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601704 . By:SD . Date:31:03:00 . Time:14:45 LOP8M. V8.B. Page 01:01Codes: 2857 Signs: 1866 . Length: 46 pic 0 pts, 194 mm

FIG. 2. Space-time diagram showing possible paths for particle to proceed from Q{ to QT .

set up problems of many variables.3 Let us suppose that we are considering a singlesystem which has coordinates that are denoted by Q and that for the time being itis not acted on by any other quantum system. It can be acted on by outside forces,however. The system may be very complicated, in which case Q represents all thecoordinates in a very general way. If at a time t the variable Q is denoted by Qt ,then the amplitude for the system to go from position Q{ at t={ to QT at t=T isgiven by

K(QT , T; Q{ , {)=| exp[(i��) S(Q)] DQ(t) (2.1)

in integral which represents the sum over all possible paths Q(t) in coordinate spacefrom Q{ to QT of the functional exp[(i��) S(Q)].4 S(Q)=�T

{ L(Q4 , Q, t) dt is theaction calculated classically from the Lagrangian for the trajectory Q(t). For the casethat Q is a single linear coordinate of position, this is represented in the diagramin Fig. 2. The magnitude of the amplitude for all paths is equal but the phase foreach path is given by the classical action along that path in units of �. Thus,amplitudes for neighboring paths which have large phases tend to cancel. The pathswhich contribute the greatest amount are those whose amplitudes have stationaryphases for small deviations around a certain path. This is the path for which theclassical action is at an extremum and is, therefore, the classical path. Remarkablyenough, for free particles and harmonic oscillators, the result of the path integrationis

K(QT , Q{)=(Smooth Function) exp[(i��) Scl],

where Scl is the action evaluated along the classical path between the two endpoints Q{ , QT . However, for more complicated systems this simple relation does

550 FEYNMAN AND VERNON

3 For a more complete treatment, see ref. [1].4 In subsequent equations K(QT , T; Qt , t) will be written K(QT , Qt).

Page 5: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

not hold. A discussion of the methods of doing integrals of this type is not includedhere since methods appropriate for the purposes here are already contained in theliterature [1, 2].

Since K(QT , Q{) is the amplitude to go from coordinate Q{ to QT , it follows thatat t=T the amplitude that the system is in a state designated by ,m(QT) wheninitially in a state ,n(Q{) is given by

Amn =| ,m*(QT) K(QT , Q{) ,n(Q{) dQT dQ{

=| ,m*(QT) exp[(i��) S(Q)] ,n(Q{) DQ(t) dQT dQ{ . (2.2)

The probability of the transition from n � m is given by |Amn |2 and from Eq. (2.2)this can be written in the form of multiple integrals as follows:

Pmn =| ,m*(QT) ,m(Q$T) exp[(i��)[S(Q)&S(Q$)]]

_,n(Q{) ,n*(Q${) DQ(t) DQ$(t) dQ{ dQ${ dQT dQ$T . (2.3)

As an example of a more complicated case let us consider two systems whosecoordinates are Q and X.5 The systems are coupled by a potential which can bedesignated as V(Q, X) and incorporated in the total Lagrangian. We assume thatwhen V=0 the states of Q and X can be described by sets of wave functions ,k(Q)and /p(X) respectively. If, initially, Q is in a state ,n(Q{) and X is in a state /i (X{),then the amplitude that Q goes from state n to m while X goes from state i to f canbe formed in a similar way to that of Eq. (2.2),

Amf, ni =| ,m*(QT) /f*(XT) exp[(i��) S(Q, X)] ,n(Q{) / i (X{)

_DQ(t) DX(t) dQ{ dX{ dQT dXT , (2.4)

where S(Q, X) represents the classical action of the entire system including both Qand X. The important property of separability afforded by writing the amplitude inthis way is now apparent.6 For instance, if one wishes to know the effect that X hason Q when X undergoes a transition from state i to f, then all of the integrals onthe X variables may be done first. What is left is an expression for Amn for Q andin terms of Q variables only but with the effect of X included. The extension of

551INTERACTION OF SYSTEMS

5 Each system will be denoted by the coordinates that characterize it. Where Q or X means specificallya coordinate, it will be so designated by a statement if it is not obvious.

6 If system Q represents a harmonic oscillator and the interaction of Q with X was linear and of theform &#(t, X) Q(t), then that part of Eq. (2.4) which involves the Q variables corresponds to the func-tion Gmn defined and used by Feynman to eliminate the electromagnetic field oscillators. See ref. [2].

Page 6: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

writing transition amplitudes for large numbers of systems is obvious. In principlethe order in which the variables are eliminated is always arbitrary.

B. Definition of Influence Functional

A functional can now be defined which can be used to describe mathematicallythe effect of external quantum systems upon the behavior of a quantum system ofinterest.7

The fundamental theorem for this work may be stated as follows: For anysystem, Q, acted on by external classical forces and quantum mechanical systems asdiscussed above, the probability that it makes a transition from state �n(Q{) at t={to �m(QT) at t=T can be written

Pmn =| �m*(QT) �m(Q$T) exp[(i��)[S0(Q)&S0(Q$)]] F(Q, Q$)

_�n*(Q${) �n(Q{) DQ(t) DQ$(t) dQ{ dQ${ dQT dQ$T , (2.5)

where F(Q, Q$) contains all the effects of the external influences on Q, and S0(Q)=�T

{ L(Q4 , Q, t) dt, the action of Q without external disturbance. The proof of thisis straightforward. Let us examine two coupled systems characterized by coor-dinates Q and X as represented diagrammatically in Fig. 1. Q will represent the testsystem and X the quite general interaction system (excepting only the effects ofspin) coupled by a general potential V(Q, X, t) to Q. Assume Q to be initially(t={) in state �n(Q{) and X to be in state /i (X{), a product state. The probabilitythat Q is found in state �m(QT) while X is in state /f (XT) at t=T can be writtenin the manner discussed above and is

Pmf, ni =|Amf, ni |2

=| �m*(QT) �m(Q$T) /f*(XT) /f (X$T)

_exp[(i��)[S0(Q)&S0(Q$)+S(X)&S(X$)+SI (Q, X)&SI (Q$, X$)]]

_�n *(Q${) �n(Q{) /i*(X${) /i (X{) dx{ } } } dQ$T DX$(t) } } } DQ(t).

(2.6)

The primed variables were introduced when the integrals for each Amf, ni

were combined. Now if all of Eq. (2.6) which involves coordinates other than Q or

552 FEYNMAN AND VERNON

7 Hereafter, the system of interest will be referred to as the test system. Conversely, the system not ofprimary interest will be called the interaction or environmental system.

Page 7: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

Q$ is separated out and designated as F(Q, Q$), then the following expression isobtained

F(Q, Q$)=| /f*(XT) /f (X$T)

_exp[(i��)[S(X)&S(X$)+SI (Q, X)&SI (Q$, X$)]]

_/i *(X${) /i (X{) dX{ dX${ DX(t) DX$(t). (2.7)

Incorporation of this expression into Eq. (2.6) yields the desired form of Eq. (2.5).If the path integrals are written in terms of kernels, Eq. (2.7) becomes

F(Q, Q$)=| /f*(XT) /f (X$T) KQ(XT , X{) K*Q$(X$T , X${)

_/i *(X${) /i (X{) dX{ } } } dX$T , (2.8)

where the subscript Q means that the kernel includes the effect of a potentialV(Q, X) acting on X during the interval T>t>{. As can be seen, F is a functionalwhose form depends upon the physical system X, the initial and final states of X,and the coupling between Q and X.

It is to be emphasized that the formulation of F is such that it includes all theeffects of the interaction system in influencing the behavior of the test system. Thus,if there are two systems A and B which can act on Q, and if

FA on Q=FB on Q ,

then the effects of A on Q are the same as those of B on Q. It follows that if sim-plifying assumptions are necessary in finding FA on Q and FB on Q (due to the com-plicated nature of A and B) and if the resulting functions are equal, then within theapproximations the effects of A and B on the test system are the same. In the situa-tion where the interaction system is composed of a linear system or combinationsof linear systems we shall see that the same form of F is always appropriate. Toadapt this general form of F to a particular linear system it is only necessary toknow such quantities as impedance and temperature which determine its classicalbehavior. In still other situations, very weak coupling between systems is involved.The approximate F which can be used in this case to represent the effect of theinteraction systems has a form which is independent of the nature of the interactionsystem. This form is the same as for linear systems. These cases will be consideredin more detail in later sections.

C. General Properties of Influence Functionals

There are several general properties of influence functionals which are of interestand which will be useful in subsequent arguments. The first three of these (1, 2, 3)

553INTERACTION OF SYSTEMS

Page 8: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

follow directly from the definition of F(Q, Q$). The last two (4, 5) will require morediscussion.

1. If the physical situation is unsure (as for instance if the type of interactionsystem X, or the initial or final states are not known precisely) but if the probabilityof the p th situation is wp and the corresponding influence functional is Fp , then theeffective F is given by

Feff=:p

wpFp #(F). (2.9)

Thus, in Eq. (2.6) if the initial state of X was not certain but the probability of eachinitial state was wi , then Pmn for system Q would be given by �i wiPmf, ni . Since thesummation involves only the part of Eq. (2.6) involving the X variables, it is a sumover the influence functions for each possible initial state and results in an averageinfluence functional of the type given above.

2. If a number of statistically and dynamically independent partial systemsact on Q at the same time and if F(k) is the influence of the k th system alone, thetotal influence of all is given by the product of the individual F(k):

F= `N

k=1

F(k). (2.10)

Again referring to Eq. (2.6), if there were N subsystems interacting with Q, then theprobability that Q makes a transition from state n to m while each of the sub-systems makes a transition from its initial to its final state is given by an expressionof the same form as Eq. (2.6). The difference in this case being that the term involv-ing the X variables would be replaced by a product of N similar terms��one foreach subsystem. Thus, when the term involving all the X (k) variables is separatedout the complete influence functional is recognized as a product of the functionalsF(k)(Q, Q$) for each subsystem.

Definition. In many cases it will be convenient to write F in the formexp[i8(Q, Q$)]. 8 is then called the influence phase. For independent disturbancesas considered in 2, the influence phases add. In the event that i8(Q, Q$) is a realnumber we will continue to use the notation 8; the phase simply becomesimaginary. It will frequently be more convenient to work with 8 rather than F.

3. The influence functional has the property that

F*(Q, Q$)=F(Q$, Q). (2.11)

Referring to Eq. (2.7), the definition of the influence functional, this fact followsimmediately upon interchanging Q and Q$.

554 FEYNMAN AND VERNON

Page 9: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

4. In the class of problems in which the final state of the interaction systemis arbitrary, which means the final states are to be summed over, then F(Q, Q$) isindependent of Q(t) if Q(t)=Q$(t) for all t. All of the problems we will be concernedwith here are of this type.

The validity of this statement can be ascertained by observing Eq. (2.7), thegeneral definition of the influence functional. In particular, for the case where theinitial and final states of the interaction system X are i and f respectively, as inEq. (2.7), we denote the influence functional by f Fi (Q, Q$). Let us assume we haveno interest in the final state of X which means that f Fi (Q, Q$) must be summedover all such states. The initial state i can be quite general. Thus, the influence func-tional for the case of an arbitrary final state is

Fi (Q, Q$)=:f

f Fi (Q, Q$).

For clarity in finding the result of letting Q(t)=Q$(t) for all t in Fi (Q, Q$), we willwrite out the expression explicitly from Eq. (2.7). It is

Fi (Q, Q)=| :f

/f*(XT) /f (X$T)_exp[(i��)[S(X)&S(X$)+SI (Q, X)&SI (Q, X$)]]

_/i *(X${) /i (X{) dX{ } } } DX$(t).

Since Q appears in the interaction potentials acting on the X and X$ variablesrespectively, it loses its identity as the coordinate of a quantum system and becomesjust a number (which may be, of course, a function of time). Thus SI (Q, X) maybe interpreted as the action of an external potential which drives the X system. Theabove expression then represents the probability that X, which is in state i initially,is finally in any one of its possible states after being acted on by an external poten-tial (as, for instance, in Eq. (2.3) summed over the final states, m). This result isunity. We have then that F(Q, Q)=1 and is independent of Q(t).

5. A more restrictive statement of the property in the above paragraph (4)can be made. In this same class of problems in which the final states are summedover, if Q(t)=Q$(t) for all t>r then F(Q, Q$) is independent of Q(t) for t>r. Tosee this we write down the influence functional from Eq. (2.8) breaking up the timeinterval into two parts, before and after r. Setting Q=Q$ for t>r and utilizing theclosure relation for the sum over final states we have

Fr(Q, Q$)=| $(XT&X$T) KQ(XT , Xr) KQ*(X$T , X$r)

_KQ(Xr , X{) K*Q$(X$r , X${) / i*(X${) /i (X{) dX{ } } } dX$T .

555INTERACTION OF SYSTEMS

Page 10: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

Examining the parts of the above integral which contain the effects of t>r:

| $(XT&X$T) KQ(XT , Xr) KQ *(X$T , X$r) dXT dX$T

=| KQ(X$r , XT) KQ(XT , Xr) dXT

=KQ(X$r , Xr)=$(Xr&X$r).

The expression for Fr(Q, Q$) becomes then

Fr(Q, Q$)=| $(Xr&X$r) KQ(Xr , X{) K*Q$(X$r , X${)

_/i *(X${) /i (X{) dX{ } } } dX$r ,

which is independent of Q(t) for t>r. As will be seen later in the specific case oflinear systems, this leads to a statement of causality.

D. Statistical Mechanics

Finally it is appropriate to point out explicitly the significance of the influencefunctional in a study of quantum statistical mechanics. In the class of problemsconsidered here we are only interested in making measurements on the test systemand not on the interacting system. Thus, when the expectation value of an operatorwhich acts only on the test system variables is taken, the final states of the inter-action system must be summed over. It is equivalent to taking the expectation valueof the desired operator in the test system and simultaneously the unit operator inthe interaction system. Therefore, only the influence functional where the final statesof the interaction system are summed over will be of interest to us.

Starting with the coordinate representation of the density matrix [6] for the testand interaction systems, \(Q, X; Q$, X$), we will show the part played by theinfluence functional in obtaining an expression for \(QT , Q$T), that is, with the Xcoordinates eliminated, in terms of its value at an earlier time {, \(Q{ , Q${). First,we recall that the definition of \ is as follows,

\(Q, X; Q$, X$)=(�(Q, X) �*(Q$, X$)) av , (2.12)

where �(Q, X) represents the wave function for one of the systems in an ensembleof systems each representing one of the possible states of the Q, X system [7]. Theaverage, represented by ( ) av , is taken over the ensemble. The trace of the densitymatrix is

Tr \(Q, X; Q$, X$)=|| \(Q, X; Q, X) dQ dX (2.13)

556 FEYNMAN AND VERNON

Page 11: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

and the expectation value of an operator A which operates on the Q variables onlyis

(A)=||| \(Q, X; Q$, X) A(Q, Q$) dQ dQ$ dX. (2.14)

In the above

A(Q$, Q)=:i, j

Aij, i*(Q) , j (Q$),

(2.15)Aij =| ,i*(Q) A,j (Q) dQ,

and ,i (Q) is one of a set of complete orthonormal eigenfunctions. From Eq. (2.14)then we see that the formal expression which we wish to derive is

| \(QT , XT ; Q$T , XT) dXT=\(QT , Q$T) (2.16)

in terms of \(Q{ , Q${). From the rules given in Section II.A for propagation of awave function with time we can easily find \T in terms of \{ . Thus

\(QT , XT ; Q$T , X$T)=| exp[(i��)[S0(Q)&S0(Q$)

+S(X)&S(X$)+SI (Q, X)&SI (Q$, X$)]]

_\(Q{ , X{ ; Q${ , X${) DQ(t) } } } dX${ . (2.17)

Now, for simplicity let us assume that initially the two systems are independent sothat

\(Q{ , X{ ; Q${ , X${)=\(Q{ , Q${) \(X{ , X${).

Then eliminating the XT coordinate as indicated in Eq. (2.16) we have

\(QT , Q$T)=| {| $(XT&X$T) exp[(i��)[S(X)&S$X$)

+SI (Q, X)&SI (Q$, X$)]] \(X{ , X${) DX(t) } } } dX${=_exp[(i��)[S0(Q)&S0(Q$)]] \(Q{ , Q${) DQ(t) } } } dQ${ .

557INTERACTION OF SYSTEMS

Page 12: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

The expression inside the braces is identified as F(Q, Q$) for the case in whichthe final state of X is summed over. Therefore, the following result is obtained:

\(QT , Q$T)=| F(Q, Q$) exp[(i��)[S0(Q)&S0(Q$)]] \(Q{ , Q${) DQ(t) } } } dQ${ .

(2.18)

Thus, if the density matrix of the test system Q is represented by \(Q{ , Q${) atsome initial instant {, the density matrix \(QT , Q$T) at some later time T is givenby Eq. (2.18). The entire influence of the interaction system is contained in F(Q, Q$).

E. Use of Influence Functionals

At this point we need to consider how influence functionals can be used in theanalysis of a problem. For clarity the discussion will be specialized to a particularproblem but the principle is valid more generally. Suppose we wish to know theprobability that a test system Q makes a transition from an initial state ,n(Q{)_exp[(&i��) En{] to a final state ,m(QT) exp[(&i��) EmT] when coupled to aninteraction system. The formal expression for this probability is, from Eq. (2.5),

Pnm =| ,m*(QT) ,m(Q$T) exp[(i��)[S0(Q)&S0(Q$)]] F(Q, Q$)

_,n*(Q${) ,n(Q{) dQ{ } } } dQ$(t). (2.19)

This is formally exact but except in special cases it cannot be evaluated exactly.Furthermore, to obtain any specific answers to the problem the characteristics of Qmust be known as well as knowing the influence functional. However, by usingperturbation theory we may find general expressions for transition probabilities toas many orders as desired. For example, if the interaction system is a linear systemat zero temperature, we will find that F0(Q, Q$) is of the form exp[i80(Q, Q$)]. Theperturbation expansion is obtained by writing exp[i80(Q, Q$)] in terms of a powerseries and evaluating the path integral corresponding to each term in the expansion.In many cases the coupling between Q and the interaction system is small enoughthat only a few terms in the expansion are necessary. In Appendix I the basic proce-dure for finding the perturbation expansion is demonstrated by finding the specificexpression up to second order in the potentials involved for transition probabilityof a test system when acted on by a linear interaction system at zero temperature.Calculation of transition probabilities represent only one piece of information thatone might desire to know about a test system. For instance, it is more usuallydesired to find the expectation value of an operator in the test system. To calculatethis one needs to know the density matrix describing the test system when it iscoupled to an interaction system. The exact expression for the required densitymatrix is given in Section II.D. Again in the general case, one runs into the difficultyof making an exact calculation and is forced to make calculations using perturbation

558 FEYNMAN AND VERNON

Page 13: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

theory. The same procedure of expanding the influence functional into a power seriesand performing the required path integrations yields useful perturbation expressions.

III. INFLUENCE FUNCTIONALS FOR CLASSICAL POTENTIALS

In this section we will derive specific forms and properties of influence functionalsfor the effects of classical potentials on the test system. These represent the simplestform of influence functionals and their properties follow directly from the generalproperties obtained in the previous section. These forms will then be extended tothe case where the classical potential represents Brownian noise.

A. Properties of Influence Functionals for Classical Potentials

The first step is to find the influence functional for a definite classical potentialacting on the test system, Q. If the potential energy term in the Lagrangian is of theform V(Q, t), then it can be ascertained readily by referring to the fundamentaldefinition of F(Q, Q$) that

F(Q, Q$)=exp {&(i��) |T

{[V(Q, t)&V(Q$, t)] dt= (3.1)

or equivalently the influence phase is

8(Q, Q$)=&(1��) |T

{[V(Q, t)&V(Q$, t)] dt. (3.2)

The next degree of complication is to have several potentials, �k Vk(Q, t) actingon Q simultaneously. However, since the sum of all these potentials represents anequivalent potential, say V(Q, t)=�k Vk(Q, t), then it is obvious that the totalinfluence function F(Q, Q$) is the product of the individual Fk(Q, Q$). Morespecifically,

F(Q, Q$)=exp {&(i��) |T

{[V(Q, t)&V(Q$, t)] dt=

=exp :k {&(i��) |

T

{[Vk(Q, t)&Vk(Q$, t)] dt=

=`k

Fk(Q, Q$), (3.3)

559INTERACTION OF SYSTEMS

Page 14: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

or

8(Q, Q$)= &(1��) :k|

T

{[Vk(Q, t)&Vk(Q$, t)] dt

=:k

8k(Q, Q$). (3.4)

The same result follows directly from Section II.C.3 which gives the total influencefunctional for several statistically and dynamically independent systems acting on Q.The total influence functional for all the systems (in this case potentials) is the productof the functionals for the individual systems.

Another property of the classical influence functionals is obtained by inspectionof Eq. (3.1). We notice that for any classical F(Q, Q$) if conditions are such thatQ(t)=Q$(t), then F(Q, Q$)=1 and is independent of t for all times that the twovariables are equal. It follows that the influence phase is zero for this condition.

Finally, from Section II.C.1 we find that if the potential is uncertain but theprobability of each Vr(Q, t) is wr then the average functional is by

(F(Q, Q$))=:r

wr exp {&(i��) |T

{[Vr(Q, t)&Vr(Q$, t)] dt=

=:r

wrFr(Q, Q$). (3.5)

In the following sections we will assume a probability distribution wr appropriateto Brownian noise and will be able to derive a specific form for the averageinfluence functional.

B. Specific Functionals for Random Potentials

Let us now suppose that the potential has known form, V(Q), but unknownstrength C(t) as a function of time so that the total potential is V(Q, t)=C(t) V(Q).The average influence functional for two cases involving this type of potential willbe particularly useful in the discussion contained in Sections IV and V. These casesare: (1) when C(t) is characterized by any coupling strength (average magnitude ofC) with a purely Gaussian distribution, and (2) when C(t) is composed of largenumber of very weak potentials (acting on the test system simultaneously) whosedistributions are stationary but not necessarily Gaussian.

1. Gaussian Noise

First, we consider the situation when C(t) is Gaussian noise with a powerspectrum 8(&) and a correlation function R({)=(2�?) ��

0 8(&) cos &{ d& then (F)is given by

560 FEYNMAN AND VERNON

Page 15: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

(F)=�exp {(i��) |T

{C(t)[V(Q)&V(Q$)] dt=�

=exp {&�&2 |T

{|

t

{R(t&s)[V(Qt)&V(Q$t)][V(Qs)&V(Q$s)] ds dt= . (3.6)

Expressed in Fourier transform notation this becomes

(F) =exp {&(?�2)&1 |�

0,(&) |V&(Q)&V&(Q$)|2 d&=, (3.7)

where

V&(Q)&V&(Q$)=|T

{[V(Q)&V(Q$)] e&i&t dt. (3.8)

Expressions of the type given in Eq. (3.6) are common for operations in which it isrequired to find the characteristic function, F(i`)=(ei` f (T )) for f (T ) represented byintegrals of the form f (T )=�b(T )

a(T ) A(T, t) x(t) dt where x(t) is a Gaussian process.The result will not be worked out here as it may be found in standard references[8].8 The equivalent expression for F in terms of frequency components, Eq. (3.7),is obtained from Eq. (3.6) in a direct manner using the definitions for R(t) andEq. (3.8).

2. Brownian Noise

The Gaussian behavior of Brownian noise, characterized by the typical Gaussianprobability distribution, may be the result of the cumulative effects of many smallstatistically independent sources, none of which is truly Gaussian. How that comesabout can be seen as follows. The effect of these small sources on a test system maybe represented by an influence functional of the same form as that of Eq. (3.6)where now C(t)=�N

i=1 Ci (t), N is a very large number, and the Ci (t) are inde-pendent random variables. Application of the central-limit theorem to this situationshows that the probability distribution appropriate to C(t) is asymptotically normalsubject to the following conditions [9]:

(a) The average values,

(Ci (t)) <�

and

+i, 2 #( |Ci (t)&(Ci (t)) |2)<�,

561INTERACTION OF SYSTEMS

8 See, for example, pp. 372�373 where it is shown that the characteristic function F(i!) appropriate tothe integral given above for a Gaussian process x(t) is exp[&1

2 !2 � �b(T )a(T ) } A(T, t) A(T, s) Kx(t, s) ds dt]

where the covariance Kx(t, s)=(xtxs) is the correlation function corresponding to R(t&s) in Eq. (3.6).

Page 16: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

(b) the absolute moments

+i, 2+$ #( |Ci (t)&(Ci (t)) | 2+$)

exist for some $>0, and

(c) making use of the definition

+l= :N

i=1

+i, l ,

then

limN � �

+2+$

(+2)1+$�2 � 0.

The condition of independence on the large number of variables and the finiteaverage values required by (a) and (b) above ensures that no one componentdominates the total distribution. Condition (c) is sufficient to ensure that all higherorder correction terms tending to deviate from a normal distribution vanish in thelimit of large N. It should be recognized that if the Ci possesses finite third moments+i, 3 the correction terms arising from these moments decrease as N&1�2. However,for the cases in which we are interested, the number of the component forces Ci isessentially infinite and higher order terms are negligible.

IV. INFLUENCE FUNCTIONALS FOR LINEAR SYSTEMS

Linear systems are of considerable interest both because of the large number ofsituations in which they are involved and because they are amenable to exactcalculation. In this section the influence functional for arbitrary combinations ofoscillators will be found by direct extension of the analysis of a single oscillator. Alllinear systems which are lossless and those which contain certain kinds of loss canbe represented by distributions of oscillators. Situations in which dissipation arisesfrom sources other than distributions of perfect oscillators will be covered inSection V. The same conclusions apply for all linear systems, however, as will bediscussed subsequently. For clarity, we will restrict our attention initially to linearinteraction systems at zero temperature and not acted on by classical forces. Theeffects of finite temperature and forces can then be included so that their significanceis more apparent.

A. Zero Temperature Linear System

The result to be proven involves the assumption that the interaction system (X)is linearly coupled to the test system (Q). The total Lagrangian for the system is

Ltotal=L0(Q4 , Q, t)+L(X4 , X, t)+LI (Q, X) (4.1)

562 FEYNMAN AND VERNON

Page 17: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

where LI=#QX, and L(X4 , X, t) is the part of the Lagrangian involving the Xsystem above. The situation is the same as that shown in Fig. 1 except the interaction

potential is given by V(Q, X, t)=&#QX and the X system is linear. Our fundamentaltheorem for linear systems is as follows:

The influence phase for the effect of X on Q can be written as follows:

8(QQ$)=1

2?� |�

0 _Q$&(Q&&&Q$&&)(i&Z&)

+Q&&(Q&&Q$&)

(&i&Z&&) & d&. (4.2)

80(Q, Q$) is found by studying the properties of X alone.9 Q& is the Fourier transformof #(t) Q(t) and Z& is a classical impedance function which relates the reaction of Xto an applied force. Z& is found by taking the classical system corresponding to X(that is, whose Lagrangian is L(X4 , X, t)) and finding the response of the coordinateX to a driving force f (t) which is derived from the potential & f (t) X(t) } f (t) isconsidered to be applied at T=0 subject to the initial conditions that X(0)=X4 (0)=0.Z& is defined by the expression

Z&= f&�(i&X&), (4.3)

where

f&=|�

0f (t) e&i&t dt and X&=|

0X(t) e&i&t dt.

In the time domain, Eq. (4.2) can be expressed as

i8(Q, Q$)=&(1�2�) |�

&�|

t

&�#t #s(Qt&Q$t)

_[QsF*(t&s)&Q$sF(t&s)] ds dt, (4.4)

where Im F(t), which we will call B(t) is, for t>0, the classical response of X to aforce f (t)=$(t). Re F(t), which for this zero temperature case we call A0(t), is thecorrelation function for the zero point fluctuation of the variable X, a point discussedat more length below. The relations connecting these quantities are then

F(t)=A0(t)+iB(t)

(4.5a)1i&Z&

=|�

0B(t) e&i&t dt

563INTERACTION OF SYSTEMS

9 More generally, the part of the interaction represented by Q could be represented by function of Qsuch as V(Q). In this case Q in the influence phase would be replaced by V&(Q), the Fourier transformof V(Q(t)).

Page 18: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

and the inverse relations

A0(t)=&2? |

0Im \ 1

i&Z&+ cos &t d&

(4.5b)

B(t)=&2? |

0Im \ 1

i&Z&+ sin &t d&.

A0(t) and B(t) are related as follows:

A0(t)=2? |

0

sB(s)s2&t2 ds. (4.5c)

These relationships may be written in many forms. Two additional forms are

F(t)=2i? |

0 \ 1i&Z&+ cos &t d&

and

|�

&�F( |t| ) e&i&t dt=2i(i&Z&)&1 for &>0

=2i(&i&Z&&)&1 for &<0 (4.5d)

All the poles of 1�i&Z& have positive imaginary parts and this impedance functionhas the additional property that

(1�i&Z&)=(1�&i&Z&&)*.

In the case of finite temperatures, the influence phase can be written in the sameform as Eq. (4.4) except that Re F(t)=A(t), that is, without the subscript 0, and amore general relation exists connecting A(t) and Im(1�i&Z&) (see Section IV.C).

F(Q, Q$) for Single Lossless Harmonic Oscillator

To prove the above theorem, we consider first a test system, Q, which is coupledto a simple harmonic oscillator whose mass is m, characteristic frequency |, anddisplacement coordinate X. The complete Lagrangian for X and Q can be written

Ltotal=L0(Q4 , Q, t)+ 12 mX4 2& 1

2 m|2X2+QX (4.6)

and the total action is written similarly:10

Stotal=S0(Q)+|T

{( 1

2 mX4 2& 12 m|2X2+QX) dt.

564 FEYNMAN AND VERNON

10 The interaction Lagrangian QX could be written more transparently as #QX where Q and X are thecoordinates of the system involved and # is a coupling factor which may or may not be a function oftime. For simplicity in writing the lengthy expressions to follow, # has been incorporated into an effectivecoordinate Q since no loss in generality results.

Page 19: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

If X is assumed to be initially in the ground state (corresponding to zero temperature)then to within a normalizing constant /i (X)=e&m|X2�2�. The final state of X is assumedto be arbitrary which means the final states are to be summed over. Therefore, inEq. (2.7), the definite state /f *(XT) /f (X$T) will be replaced by the sum �n 8n*_(XT) 8n(X$T)=$(XT&X$T). The 8n(X) represent the energy eigenfunctions of theharmonic oscillator. With this information available the influence functional iscompletely defined and from Section II.B can be written

F0(Q, Q$)=| $(XT&X$T) KQ(XT , X{) K*Q$(X$T , X${)

_exp[&(m|�2�)(X{2+X$2

{ )] dX{ } } } dX$T , (4.7a)

where the subscripts Q, Q$ refer to the interaction potentials &QX and &Q$X$acting on the X and X$ systems respectively. The subscript 0 on F0(Q, Q$) indicateszero temperature. For the harmonic oscillator

KQ(XT , X{)=N exp[(i��)[S(X)&SI (Q, X)]classical

=N exp[[i|�2� sin |(T&{)][(XT2+X{

2) cos |(T&{)

&2XTX{+(2XT �|) |T

{Qt sin |(t&{) dt+(2X{�|)

_|T

{Qt sin |(T&t) dt&(2�|2)

_|T

{|

t

{Qt Qs sin |(T&t) sin |(s&{) ds dt, (4.7b)

where N is a normalizing factor depending only on | and the time interval T&{.11

Thus, Eq. (4.7a) represents a Gaussian integral over the four X variables since S isitself quadratic in the X variables. When the integrals are carried out the followingresult is obtained for the influence phase:

i80(Q, Q$)=&(2�m|)&1 |T

{|

t

{(Qt&Q$t)

_(Qse&i|(t&s)&Q$sei|(t&s)) ds dt. (4.8)

565INTERACTION OF SYSTEMS

11 See ref. [2, Section 3].

Page 20: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

Thus, F(t&s) in Eq. (4.4) corresponds in this case to e+i|(t&s)�m| and from thedefinition given above B(t&s)=(1�m|) sin |(t&s).12 Rewriting Eq. (4.8) in trans-form notation we have

80(Q, Q$)=(2?�)&1 |�

0 { Q$&(Q&&&Q$&&)&m[(&&i=)2&|2]

+Q&&(Q&&Q$&)

&m[(&+i=)2&|2]= d&, (4.9)

where

Q&=|�

&�Qte&i&t dt.

The function [&m[(&&i=)2&|2]]&1=(m|)&1 ��0 sin |te&i&t dt corresponds to

1�i&Z& of Eq. (4.2).13

Having obtained the expression for the influence phase we now turn to the classicalproblem of finding the response of X(t) to a driving force, f (t), applied at t=0 withthe initial conditions X(0)=X4 (0)=0. Starting with the Lagrangian on the unper-turbed oscillator from Eq. (4.6), we add to it a potential term & f (t) X(t). Thispotential has the same form as the coupling potential &QX used in the quantumcalculation. However, it is to be emphasized that the response of X to a force hasnothing to do with the system Q outside of the type of coupling involved; therefore,f (t) will symbolize the force in the classical problem. The complete Lagrangian is

L(X4 , X, t)= 12 mX4 2& 1

2 m|2X2+ fX (4.10)

and the equation of motion derived from it is

mX� +m|2X= f. (4.11)

Its solution under the initial conditions states above is

X(t)=(m|)&1 |t

0f (s) sin |(t&s) ds (4.12)

or alternatively, in terms of Fourier transforms, is

X&= f&[&m[(&&i=)2&|2]]&1. (4.13)

566 FEYNMAN AND VERNON

12 The finite time interval indicated by the limits T and { can be interpreted as turning the coupling(between Q and X) on at t={ and off at t=T. However, since the interaction system is to be consideredin most cases as part of the steady-state environment of Q, it is really more meaningful to extend theselimits over an infinite range of time ({ � &�, T � +�). The possibility of allowing X to interact withQ over a finite range of time can be taken care of by giving the coupling factor (already included in thevariable Q) the proper time dependence.

13 = which occurs in i&Z& is a convergence factor which was inserted in taking the Fourier transform(1(t)�m|) sin |t where 1(t) is the unit step function and is kept to show the location of the poles withrespect to the & axis when doing integrations of the type ��

0 B(&) [i&Z&]&1 d&.

Page 21: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601721 . By:SD . Date:31:03:00 . Time:14:45 LOP8M. V8.B. Page 01:01Codes: 2772 Signs: 1718 . Length: 46 pic 0 pts, 194 mm

FIG. 3. Test system Q coupled to a distribution of oscillators.

Therefore, B(t&s) in this case is a Green's function which yields the response ofX(t) to an impulse force f (s)=$(s) and its transform yields 1�i&Z& . Thus a classicalcalculation of the ratio X&�f& under quiescent initial conditions yields the properfunction for 1�i&Z& .

Distribution of Oscillators��Representation of Loss

The results of the preceding section are easily extended to the situation where theinteraction system is a distribution of oscillators. First, we consider the case ofindependent oscillators coupled to the test system. It is assumed that there is adistribution of oscillators such that G(0) d0 is the weight of oscillators whosenatural frequency is in the range between 0 and 0+d0. More specifically, G(0) d0is the product of the number of oscillators and the square of their coupling constantsdivided by the mass in d0. Thus, we have a situation represented by the diagramof Fig. 3. Each oscillator is assumed to be initially in the ground state and finallyin an arbitrary state; the coupling is again assumed to be linear. The total actionis then given by

S[Q, X(0)]=S0(Q)+|T

{|

0G(0)[ 1

2X4 2& 12X202+QX] d0 dt. (4.14)

For the general properties of influence functionals already described we know thatwhen independent disturbances act on Q the influence functional is a product of theones for each individual disturbance. Since F0(Q, Q$)=exp[i80(Q, Q$)] for thecase of a single oscillator, the total influence phase for the distribution is the sumof the individual phases,

80(Q, Q$)=|�

0G(0) 80, 0(Q, Q$) d0. (4.15)

More explicitly,

80 =(2?�)&1 |�

0G(0) d0

_|�

0 { Q$&(Q&&&Q$&&)&[(&&i=)2&02]

+Q$&&(Q&&Q$&)

&[(&+i=)2&02]= d&. (4.16)

567INTERACTION OF SYSTEMS

Page 22: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

For this case then, the form of Eq. (4.2) is obtained if we put

(i&Z&)&1= lim

= � 0 |�

0G(0)[(&&i=)2&02]&1 d0 (4.17)

or14

(Z&)&1=(?�2) G(&)&i& |�

0G(0)(&2&02)&1 d0. (4.18)

Thus the effects of all the oscillators are included in the influence phase through theexpression for Z& , Eq. (4.17). Now, however, because of the continuous distributionof oscillators, Z& has a finite real part. We will now show that this real partrepresents dissipation by arriving at the same impedance function classically.

As before, we take the part of the Lagrangian from Eq. (4.14) having to do withthe oscillators, except that the coupling potential &Q(t) ��

0 G(0) X(0, t) d0 isreplaced by & f (t) ��

0 G(0) X(0, t) d0, a classical potential. X(0, t) is the coor-dinate of the oscillator in the distribution whose frequency is 0 while the totalcoordinate of the complete linear system with which f (t) is interacting is��

0 G(0) X(0, t) d0=X(t). It is the relationship between f (t) and X(t) in which weare interested in this classical case:

L[X4 (0), X(0), t]=|�

0G(0) d0[ 1

2X4 (0)2& 12 02X(0)2]

+ f (t) |�

0G(0) X(0) d0. (4.19)

The equations of motion are the infinite set represented by

X� (0)+02X(0)= f (t). (4.20)

They result from varying L with respect to the independent variables X(0). Forquiescent initial conditions and for f (t) applied at t=0, this solution is expressed

X&(0)� f&=&[(&&i=)2&0]&1=[i&Z&(0)]&1.

The relation of the total coordinate X& to f& is obtained simply,

X&� f& =_|�

0X&(0) G(0) d0& ( f&)&1

=&|�

0G(0)[(&&i=)2&02]&1 d0=(i&Z&)

&1. (4.21)

568 FEYNMAN AND VERNON

14 Eq. (4.18) is obtained from (4.17) using the identity

lim= � 0

[(&&i=)2&02]&1=(&2&02)&1+(i?�20)[$(&&0)&$(&+0)].

Page 23: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

Referring to Eq. (4.17) it is seen that the same expression for Z& is obtained in thequantum and classical cases. In addition, since Z& is now identified with a classicalimpedance, the real part represents resistance while the imaginary part correspondsto reactance. Therefore, at least for the case that loss in represented by distributionsof oscillators, its effect can be included in the influence functional by using theappropriate impedance expression. The spontaneous emission of a particle in freespace represents a good example of such a loss mechanism. A demonstration ofthis point is included in Appendix II where the oscillator distribution is relatedto the probability of spontaneous emission starting from the influence functionalrepresenting the effect of free space.

The relationship

(i&Z&)&1=|�

0B(t) e&i&t dt

has already been established during the course of the derivation of the influencephase for the single oscillator. Now the inverse relation between F(t) and 1�i&Z& canbe written for the zero temperature case. In the time domain the influence phase forthe distribution of oscillators is

i8(Q, Q$)=&(2�)&1 |�

0G(0) 0&1 d0 |

&�|

t

&�(Qt&Q$t)

_(Qse&i0(t&s)&Q$sei0(t&s)) ds dt.

Comparing this with Eq. (4.4) it is evident that

F(t)=|�

0G(0) 0&1ei0t d0.

But from Eq. (4.18),

Im(i&Z&)&1=&?G(&)�2&.

Therefore, it can be immediately written that

F(t)=&(2�?) |�

0Im(i&Z&)&1 ei&t d& (4.22)

as was given in Eq. (4.5b).The results above can now be extended by a simple argument to include all linear

systems composed entirely of distributions of oscillators. To do this it need only beshown that the general system can be reduced to a distribution of oscillators independ-ently coupled to the test system, which was the situation just considered. Explicitly,suppose there exists a test system Q, coupled to an assemblage of oscillators whichare also interconnected with each other. For instance, the situation might be as in

569INTERACTION OF SYSTEMS

Page 24: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601724 . By:SD . Date:31:03:00 . Time:14:45 LOP8M. V8.B. Page 01:01Codes: 4052 Signs: 3070 . Length: 46 pic 0 pts, 194 mm

FIG. 4. Test system coupled to an arbitrary assemblage of oscillators.

Fig. 4, where each of the Xn components of the total interaction system could alsorepresent a system of oscillators. However, it is well known15 that such a linearsystem may be represented by an equivalent set of oscillators (the normal modes ofthe total system) independently coupled to Q.16 Or, stated another way, the classicalrepresentation of the Lagrangian in normal modes finds new linear combinationsof the Xn which makes the total Lagrangian, except for the coupling, a sum ofindividual quadratic forms with no cross terms. But, this same transformationof variables can be made on the expression for F(Q, Q$) (see Eq. (2.7)). The effectof this transformation is to change the DX(t) volume by a numerical factor, sincethe transformation is linear.17 Thus, in effect, we get the sum of independentsystems in the quantum mechanical case also. From this argument it is concludedthat the results above regarding a distribution of independent oscillators coupled to atest system apply to any linear interaction system. Therefore, it has been found that theinfluence functional for all linear systems has exactly the same form exp[i80(Q, Q$)]where 80(Q, Q$) is a quadratic functional of the Q and Q$. 80(Q, Q$) is adapted toa particular linear system only through the classical response of that linear systemto a force. Thus, the procedure for finding the influence functional for a linearsystem has been reduced to a classical problem. The fact that eliminating the coor-dinate of an oscillator always yields an influence functional which is quadratic inthe potential applied to that oscillator is a basic property of linear systems. Forexample, where the coupling Lagrangian is linear between an oscillator of coor-dinate X and another system of coordinate Q, the elimination of the X coordinateyields an influence phase which is quadratic in Q as has already been shown. If Qwere the coordinate of another oscillator coupled to P, then elimination of the Qcoordinates would yield an influence phase quadratic in P, etc. This can be under-stood mathematically by observing that the Lagrangian for all the oscillators withlinear coupling is always quadratic. Doing the path integral to eliminate a coor-dinate is basically a process of completing the square and performing Gaussian

570 FEYNMAN AND VERNON

15 This point is considered more fully in Section IV.B on classical forces.16 The fact that one or more of the Xn might represent continuous distributions of oscillators need not

be bothersome since in principle they represent the behavior of the total system in terms of its infiniteset of normal modes.

17 The only result of such a numerical factor would be to change the normalization of F(Q, Q$).However, we already know that for the case that the final states of the interaction system are summedover F(Q, Q$)=1. Therefore the normalization of F(Q, Q$) is not changed by the transformation andthus is not dependent upon the coordinates chosen to represent the interaction system.

Page 25: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

integrals. This process of completing the square also yields quadratic terms. It istherefore not surprising that the influence phase for any linear system should bealways of the same quadratic form.

It is to be emphasized that the analysis so far presented has been concernedentirely with systems whose complete behavior can be described by combinationsof lossless oscillators at zero temperature. The only example of such a system is theelectromagnetic field in free space. In all other physical situations linear behavior isan approximation to the actual behavior. However, this approximation may bevery good over a wide range of operating conditions. In Section V the problem ofapproximately linear systems will be considered in detail. The results will be foundto be the same as for perfect oscillators to the extent that linear behavior is realized.

Form of Influence Functionals for Linear Systems and Classical Forces as Deducedfrom Properties of Influence Functionals

So far, we have found the influence phase for classical potentials, uncertain classicalpotentials, and linear systems at zero temperature. By studying Eqs. (3.1), (3.6), and(4.4), we see that the general form for the influence functional in which all three ofthese were acting on Q is

F(Q, Q$)=exp {|T

{iC1(t)(Qt&Q$t) dt&|

T

{|

t

{A1(t&s)(Qt&Q$t)

_(Qs&Q$s) ds dt&|T

{|

t

{iB1(t&s)(Qt&Q$t)(Qs+Q$s) ds dt= . (4.23)

The exponent is written solely in terms of Q for simplicity although when thepotentials are not linear in Q (as XV(Q)), the same general form exists, except thatit is written in terms of V(Q). We now observe that there are other possible com-binations of the Q, Q$ variables not represented here such as terms in (Qt+Q$t),(Qt+Q$t)(Qs+Q$s). To see if such terms are possible, let us form a hypotheticalfunctional containing all possible forms up to second order in Q:

F(Q, Q$)=exp {|T

{[iC1(t)(Qt&Q$t)+D1(t)(Qt+Q$t)] dt

&|T

{|

t

{[A1(t&s)(Qt&Q$t)(Qs&Q$s)

+iB1(t&s)(Qt&Q$t)(Qs+Q$s)+iD2(t&s)(Qt+Q$t)(Qs&Q$s)

+D3(t&s)(Qt+Q$t)(Qs+Q$s)] ds dt= . (4.24)

That the coefficients of the Q's inside the double integrals should be functions of(t&s) is evident since the functional should not depend on the absolute time. Wenow will try to eliminate terms in the exponent by using the general properties of

571INTERACTION OF SYSTEMS

Page 26: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

F(Q, Q$) given in Section II. First, we know F(Q, Q$)=F*(Q$, Q). This impliesthat all the functions A1 , B1 , C1 , D1 , D2 , and D3 are real. Next, we know thatF(Q, Q$)=1 if Q$(t)=Q(t). Hence, D1 , D3 are zero. This leaves only one termwhich we did not have before, that of D2(t&s). Now we apply the property of thesefunctionals requiring that if Qt=Q$t for t>t0 then F(Q, Q$) is independent of Qfor t>t0 . This statement is obviously true for the C1(t) and A1(t&s) terms.Consider now the B1(t&s) term. For t>t0 , Qt&Q$t=0 and therefore this term isalso legitimate. As for the D2(t&s) term, let us consider t>t0 , but s<t0 . ThenQs&Q$s {0. Furthermore, Qt+Q$t=2Qt {0. Therefore, D2(t&s) must also bezero. The fact that D2(t&s)=0 is actually a statement of causality, i.e., that theeffect due to an applied force cannot precede the time the force was applied. To seethis, let us change the limits of integration on this term:

|T

{|

t

{D2(t&s)(Qt+Q$t)(Qs&Q$s) ds dt

=|T

{|

T

tD2(s&t)(Qt&Q$t)(Qs+Q$s) ds dt.

Now the integrand is of the same form as that of the B(t&s) term. However, fora fixed t the integration over s is over the range of s>t. This amounts to a sumover the future rather than a sum over past histories of the variable Q.

The conclusion to be drawn is that there are three possible types of terms up tosecond order in Q and Q$ when definite classical forces, indefinite classical forces,and linear systems act on Q. Terms of this type have already been derived duringthe course of our analysis. Therefore, there are no major types of phenomenawhich have not been noticed. In light of the above discussion we would expect theeffects of additional phenomena, if they are described by terms of second order orless in Q and Q$, to be contained in one or more of the three forms of exponentsshown in Eq. (4.23). For instance, in the case of a linear system at finite temperature,it will be found that the effect of temperature is to change the effective value of A1(t&s)in the exponent of Eq. (4.23) from its minimum value which occurs at zero tem-perature. It should be pointed out that although A0(t&s) (see Eq. (4.5b)) occurs ina term which has the form of an uncertain classical potential acting on the testsystem, at zero temperature one must be careful about this interpretation, for theexistence of a random classical potential implies a random fluctuation of thevariables of the interaction system which could induce transitions in the test systemeither upward or downward in energy. However, if the interaction system isalready in its lowest state it can only induce downward transitions in the testsystem so that the term in A0(t&s) by itself is not sufficient. Thus, as has alreadybeen found the exponent of the zero temperature influence functional contains twoterms, one in A0(t&s) and the other in B(t&s), which are related through Eq. (4.5c).Together they give the whole picture, i.e., that there is a zero point, random fluctuationof the variables of the interaction system but that this fluctuation can induce only

572 FEYNMAN AND VERNON

Page 27: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

those transitions in the test system which, through spontaneous emission, give upenergy to the interaction system.

B. Influence Functionals for Driven Linear Systems

It is to be expected that if a classical force is applied to a linear interaction systemwhich in turn is coupled to a test system, the effect of the interaction system is tomodify the character of the force applied to the test system. In this section we willfind the exact form for the influence functional of this effective force. If a linearsystem is coupled to Q(t) through one of its coordinates X(t) and if a classical forceC(t) is coupled to another coordinate Y(t), then 8(Q, Q$) representing the effect ofboth the linear system and the force is

8(Q, Q$)=80(Q, Q$)+(2?�)&1 |�

0 _C&(Q&&&Q$&&)i&z&

+C&&(Q&&Q$&)

&i&z&& & d&, (4.25)

where 80 is the influence phase of the linear system in the absence of a classical forceC(t), and z& is a transfer impedance function which modifies the effect of C& on Q.The impedance z& is found by computing the classical response of the coordinate X tothe force C with all other potentials acting on the linear system (including those dueto coordinates of external systems such as Q) set equal to zero. The result of thecalculation yields i&z&=C& �X& . Alternatively, in the time domain,

8(Q, Q$)=80(Q, Q$)+�&1 |�

&�|

t

&�(Qt&Q$t) b(t&s) Cs ds dt (4.26)

where18

(i&z&)&1=|

0b(t) e&i&t dt.

The theorem can be stated in the form of a diagram as shown in Fig. 5. In thisfigure f (t)=� t

&� F(s) b(t&s) ds. It will be convenient to work in the frequencydomain.

First we recall the influence phase for a classical force acting directly on Q. Fromthis expression we will be able to identify the character of the force acting on Q in

573INTERACTION OF SYSTEMS

18 The notation z& , b(t&s) was chosen to avoid confusion with Z& and B(t&s) which are theimpedance and response function respectively of the linear system as seen by the test system.

Page 28: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601728 . By:SD . Date:31:03:00 . Time:14:45 LOP8M. V8.B. Page 01:01Codes: 2661 Signs: 1417 . Length: 46 pic 0 pts, 194 mm

FIG. 5. Equivalent influences of a linear system and a force, F(t), acting on a test system, Q.

more complicated expressions. If the potential is of the form &C(t) Q(t), we have,from Section III,

8(Q, Q$)=(2?�)&1 |�

0[C&(Q&&&Q$&&)+C&&(Q&&Q$&)] d&. (4.27)

Classical Potential and Q Coupled to the Same Coordinate

Before developing the general situation we consider the simpler situation whereQ is coupled to a linear system through the potential &QX, and a force F(t) isapplied to the same system through the potential &FX. The Lagrangian for thecomplete system is

L(system)=L0(Q4 , Q, t)+(F+Q) X+L(X4 , Y4 , ..., X, Y, t), (4.28)

where X, Y, ... represent all the coordinates of the linear system. If F=0,

8(Q, Q$)#80(Q, Q$)

=(2?�)&1 |�

0 _Q$&(Q&&&Q$&&)(i&Z&)

+Q&&(Q&&Q$&)

(&i&Z&&) & d&. (4.29)

If F{0 it is evident from Eq. (4.28) that the required influence phase can be foundby replacing Q by Q+F and Q$ by Q$+F. Notice that F does not carry the primenotation since it is not a coordinate. If this substitution is made in Eq. (4.29) wehave

8(Q, Q$)=(2?�)&1 |�

0 _(Q$&+F&)(Q&&&Q$&&)i&Z&

+(Q&&+F&&)(Q&&Q$&)

(&i&Z&&) & d&

=80(Q, Q$)+(2?�)&1

_|�

0 _\ F&

i&Z&+ (Q&&&Q$&&)+\ F&&

&i&Z&&+ (Q&&Q$&)& d&. (4.30)

As might be expected the total effect of the linear system and the driving forceconsists of two separate terms, one describing the effect of the linear system aloneand the other describing the effect of the driving force. Comparison of Eqs. (4.26)

574 FEYNMAN AND VERNON

Page 29: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

and (4.30) shows that the effective force applied to Q is, in transform language,F& �i&Z& and, further, shows that F& is modified by 1�i&Z& , the classical impedancefunction of the interaction system. In this special example where F and Q are bothcoupled to the same coordinate, Z& is both the correct impedance to be usedin 80(Q, Q$), i.e., that impedance seen by the test system, and the transferimpedance z& which modifies F& . This is not true generally as we shall see in thenext section. In addition, it is interesting to observe that no unexpected quantumeffects appear because of the addition of a force to the interaction system. The onlyeffect of the interaction system is to modify the characteristics of F(t) in an entirelyclassical way.

Classical Forces Acting through a General Linear System

Having obtained an idea of the type of results to expect in the above simplifiedanalysis we now proceed to the more general case. Let the N coordinates of theinteraction system be represented by Xi , i=1 } } } N. Its coupling to the test system,Q(t) and to the driving force C(t) is given by the potentials &Xn Q and &XkC,respectively. Thus we are assuming, for simplification in writing, that Q is coupledonly to the variable Xn and the force C(t) is applied just to the variable Xk . Againthe interaction system is assumed to be composed entirely of harmonic oscillators.The Lagrangian is

L(system)=L(Q4 , Q, t)+:i, j

[ 12 (TijX4 iX4 j&VijXiXj)]

+XnQ+XkC. (4.31)

It is well known in the theory of linear systems that new coordinates may bedefined by means of a linear transformation of the Xi . These new coordinates willbe chosen as the eigenvectors, Yl , of the interaction system [10]. Thus,

Xi= :n

l=1

a ilY l i=1, 2, ..., n.

Assuming the ail to be properly normalized, the Lagrangian may be rewritten asfollows:

L(Q4 , Q, Y4 i , Yl , t)=L0(Q4 , Q, t)

+:l

[ 12 (Y4 2

l &|2l Y 2

l )+Yl (anlQ+akl C)]. (4.32)

Since these are now independent oscillators coupled to Q the influence phase canbe written down immediately,

575INTERACTION OF SYSTEMS

Page 30: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

8(Q, Q$)=:l

8l (Q, Q$)

=:l

12?� |

0 _Q$&(Q&&&Q$&&) \ a2nl

i&Zl (&)++Q&&(Q&&Q$&) \ a2

nl

&i&Z l (&&)+& d&+:l

12?�

_|�

0 _anlakl C&

i&Zl (&)(Q&&&Q$&&)+

anlaklC&&

&i&Zl (&&)(Q&&Q$&)& d&, (4.33)

where

1i&Zl (&)

=Yl (&)anlQ& }C&=0

#Yl (&)akl C& }Q&=0

(4.34)

calculated classically.This can be written in the form of Eq. (4.25) if we make the correspondence

1i&Z&

=:l

a2nl

i&Zl (&)

and

1i&z&

=:l

anlakl

i&Zl (&). (4.35)

Using Eqs. (4.34) and (4.35) we can show that 1�i&Z& and 1�i&z& are equivalent toXn(&)�Q& and Xn(&)�C& , respectively. Thus

1i&Z&

=:l

a2nl

i&Zl (&)=:

l

a2nl \Yl (&)

anlQ&+ }C&=0

=:l

anlYl (&)Q& }C&=0

=Xn(&)

Q& }C&=0

(4.36)

and

1i&z&

=:l

anlakl

i&Zl (&)=:

l

anlaklYl (&)aklC& }Q&=0

=� l anlY l (&)

C& }Q&=0

=Xn(&)

C& }Q&=0

. (4.37)

Equation (4.36) is a mathematical expression of the argument used earlier to findthe appropriate impedance function to be used in the influence functional for alinear system acting on Q. The additional information obtained here in this regard

576 FEYNMAN AND VERNON

Page 31: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

is that when other forces are present they are to be set equal to zero when this com-putation is made. Equation (4.37) states the new result that the transfer impedancewhich modifies C& in its effect on the test system is to be found by computing theratio of C& to the coordinate Xn(&) to which the test system is coupled. The totalforce acting on the test system when several forces are acting on the interactionsystem is simply the sum of these forces each modified by the appropriate transferimpedance determined in the above described manner.

C. Linear Systems at Finite Temperatures

The forms of influence functionals which are possible for linear systems havebeen established by an argument which utilized the general properties discussed inSection II. Each of these forms has already occurred in the analyses of classicalpotentials, random potentials, and zero temperature linear systems. Therefore, theresults to be expected here are one or more of the forms already obtained.

The discussion is begun again with a single oscillator as our linear system, forsimplicity. From this, the extension to distributions of oscillators is immediate, asit was in Section IV.A. The complete problem is set up in the same way as for zerotemperature except that the initial state of the oscillator is not simply the groundstate or any definite eigenstate. The effect of temperature is to make the initial stateuncertain and it is properly represented by a sum over all states weighted by theBoltzmann factor e&;En where ;=1�kT, T being the temperature in this case. Thefinal state is again arbitrary: therefore, a formal expression for the influence func-tional is

F(Q, Q$)=| $(XT&X$T) KQ(XT , X{) K*Q$(X$T , X${)

_:n

N&1e&;En ,n(X{) ,n *(X${) dX{ } } } dX$T , (4.38)

where N, the normalization constant, is �n e&;En. The ,n represent the energyeigenfunctions of the oscillator unperturbed by external forces. The first problem isto find a closed form for the expression �n ,n(X) ,n(X$) e&;En. This can be done bynoticing that its form is identical with the kernel which takes a wave function fromone time to another if we make the correspondence that ; represents an imaginarytime interval. If the times involved are t2 and t1 , this kernel is

K0(X2 , X1)=:n

,n(X2) ,n *(X1) exp[&(i��) En(t2&t1)]

=exp[(i��) Scl] (4.39)

577INTERACTION OF SYSTEMS

Page 32: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

for the harmonic oscillator, and where the subscript 0 indicates the absence ofexternal forces. For the harmonic oscillator the expression for S is easily obtainedin terms of the initial and final positions X1 and X2 [2]. Thus,

Scl=m|[2 sin |(t2&t1)]&1 [(X12+X2

2) cos |(t2&t1)&2X1X2]. (4.40)

Utilizing Eqs. (4.42) and (4.43) and making the correspondence ;=i(t2&t1)��,X1=X{ , and X2=X{ $, we find that

: ,n(X{) ,n*(X${) exp(&;En)

=exp[&m|[2� sinh(;�|)]&1 [(X{2+X$2

{ ) cosh(;�|)&2X{ X${]]. (4.41)

Using this closed expression for the average initial state of the oscillator and thekernel for the driven harmonic oscillator (Eq. (4.7b)) the influence functional canbe evaluated by evaluation of a series of Gaussian integrations just as in the zerotemperature case. The result expressed in the frequency domain is

i8(Q, Q$)=i80(Q, Q$)&(?�2)&1 |�

0?�[2m|(e;�|&1)]&1 $(&&|) |Q&&Q$& |2 d&

(4.42)

Thus, the influence phase is made up of two terms, the first of which is the effectof the oscillator at zero temperature. The second is recognized as having the sameform which was found for an uncertain classical potential with a Gaussian distribu-tion. Therefore, the effect of finite temperature is to introduce a noisy potentialacting on Q at the frequency of the original oscillator. The power spectrum of thenoise produced by the finite temperature is

,(&)=?�[2m&(e;�&&1)]&1 $(&&|). (4.43)

To indicate more clearly the relationship of ,(&) to the characteristics of the linearsystem it is instructive to extend this expression to the case of a distribution ofoscillators G(0) all at the same finite temperature. The resulting influence phase is

i8(Q, Q$)=i |�

080(Q, Q$) G(0) d0

+(?�2)&1 |�

0?�G(&)[2&(e;�&&1)]&1 |Q&&Q$& |2 d&, (4.44)

where in the distribution m has been set equal to unity. The first term is again theinfluence phase for zero temperature, while the second term again has the form ofa noisy potential whose power spectrum

,(&)=(�?�2) G(&)[&(e;�&&1)]&1.

578 FEYNMAN AND VERNON

Page 33: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

Recalling the analysis of the distribution of oscillators, it is found from Eq. (4.18)that ?G(&)�2=Re(1�Z&). Therefore, the power spectrum can be written

,(&)=� Re(1�Z&)[&(e;�&&1)]&1. (4.45)

In the time domain the influence phase is

i8=i80&(?�2)&1 |�

0d& G(&) �[&(e;�&&1)]

_|�

&�|

t

&�(Qt&Q$t)(Qs&Q$s) cos &(t&s) ds dt. (4.46)

Comparing this with Eq. (3.6) for random classical forces we see that the correlationfunction of the noise due to the finite temperature is

R(t&s)=|�

0�G(&)[&(e;�&&1)]&1 cos &(t&s) d&

=&(2�?) |�

0� Im(i&Z&)&1 (e;�&&1)&1 cos &(t&s) d&. (4.47)

Finally, if we write Eq. (4.46) in terms of F(t&s) as, for instance, in Eq. (4.4), wefind that

F(t)=|�

0[G(&)�&] coth (;�&�2) cos &t d&+i |

0[G(&)�&] sin &t d&.

Thus,

A(t)=&(2�?) |�

0Im(i&Z&)&1 coth (;�&�2) cos &t d&

=&(2�?) |�

0Im(i&Z&)&1 [1+2(e;�&&1)&1] cos &t d& (4.48)

and

B(t)=&(2�?) |�

0Im(i&Z&)&1 sin &t d&

which are the more general counterparts of Eq. (4.5b). Notice, however, that onlythe relation for A(t) changes with temperature.

Thus, if a linear interaction system is initially at a finite temperature, the resultingeffect is the same, as far as the test system is concerned, as if the linear system wereat zero temperature and, in addition, a random classical potential were connectedindependently to the test system. The power spectrum of the random potential is

579INTERACTION OF SYSTEMS

Page 34: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601734 . By:SD . Date:31:03:00 . Time:16:25 LOP8M. V8.B. Page 01:01Codes: 3327 Signs: 2467 . Length: 46 pic 0 pts, 194 mm

FIG. 6. Equivalent influences of a linear system at a finite temperature acting on a test system.

given by Eq. (4.45) and is related both to the temperature and to the dissipativepart of the impedance of the linear system. The theorem is stated in terms of adiagram in Fig. 6 where the power spectrum ,(&) of the random force C& is definedby

,(&)=4? (C&C&&$) $(&+&$). (4.49)

This fluctuation dissipation theorem has a content which is different from those statedby Callen and Welton [11], Kubo [12], and others. It represents still anothergeneralization of the Nyquist theorem which relates noise and resistance in electriccircuits [13]. These previously stated fluctuation dissipation theorems related thefluctuations of some variable in an isolated system, which is initially at thermalequilibrium, to the dissipative part of the impedance of the isolated system. Thiswould be equivalent in our case to relating (Q2) when the test system is at equi-librium (F=0) to the dissipative part of F& �Q& where F is a classical force actingon Q through a linear potential as shown in Fig. 7. However, we have shown thatthe effect of an external quantum system at thermal equilibrium on a test systemcan be separated into two effects, a zero point quantum term, which cannot beclassified as pure noise, and a random potential term. Using this influence func-tional approach we can find (Q2) due not only to the internal fluctuations of Qbut also due to the effect of X.

The fact that all the derivations so far have been exact, which is a consequenceof dealing only with systems made up of distributions of oscillators, brings up twointeresting aspects of the theory. The first one is that Z& does not depend on thetemperature, only on the distribution of oscillators. Yet in any real finite system, itis to be expected that the temperature of a system does affect its impedance. Thesecond aspect is that when a force f (t), of any magnitude whatever, is applied tothe linear system, its temperature does not change although it is obvious that if z&

has a finite real part the linear system must absorb energy. For example, we haveshown that the effect on a test system of a linear system at a finite temperatureacted on by a force f (t) is the same as the effect of a linear system at zero temperature,a force dependent only on the temperature of the linear system, and the force f (t)

FIG. 7. Linear classical force F acting on a test system.

580 FEYNMAN AND VERNON

Page 35: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601735 . By:SD . Date:31:03:00 . Time:14:46 LOP8M. V8.B. Page 01:01Codes: 3106 Signs: 2332 . Length: 46 pic 0 pts, 194 mm

FIG. 8. Equivalent influences of a linear system at a finite temperature and a force acting on Q.

modified by the transfer characteristics of the linear system. Figure 8 shows thesituation where f&=F&�i&z& and C(t) is the random force with a power spectrumgiven in Eq. (4.49). The influence phase acting on Q is

i8=i80+i(2?�)&1 |�

&�( f& �i&z&)(Q&&&Q$&&) d&

&(?�2)&1 |�

0,(&) |Q&&Q$& | 2 d&.

As can be seen, the addition of the driving force f (t) (which is denoted in the aboveexpression by its Fourier transform f&) does not change either the impedancecharacteristics or the temperature of the interaction system (which would be reflectedas changes in 80(Q, Q$), z& , and ,(&)).

The fact that the impedance characteristics of the interaction system are nottemperature dependent is a direct result of linearity. The incremental response of aperfectly linear system due to a driving force is independent of its initial state ofmotion or of that induced by other linearly coupled forces. Therefore, the randommotion of the coordinate of the interaction system implied by a finite temperaturedoes not affect its response to a driving force.

To see the reason that temperature of the linear interaction system has nodependence on the applied force F(t), let us consider a linear system represented bya very large box whose dimensions we will allow to become infinite. In addition, weassume that a test system is located in the box and is receiving portions of a classical(very large amplitude) electromagnetic wave which is being transmitted by an antennaalso within the box. As long as the box is of finite size and has lossless walls, any signaltransmitted by the antenna will find its way either to the test system or back to theantenna. Transmission of the signal will be effected through the modes of the cavitywhich have a discrete frequency distribution. This system exhibits no loss since eachmode represents a lossless harmonic oscillator. For the rest of the discussion it isconvenient to separate the energy transmitted by the antenna into parts whichreach the test system directly or through reflection from the walls. As the dimen-sions of the box are allowed to become infinite (i.e., the distribution of oscillatorsdescribing the electromagnetic behavior of the box becomes continuous), the time

581INTERACTION OF SYSTEMS

Page 36: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601736 . By:SD . Date:31:03:00 . Time:14:46 LOP8M. V8.B. Page 01:01Codes: 3067 Signs: 2279 . Length: 46 pic 0 pts, 194 mm

required for the energy to reach the test system by reflection from the walls alsobecomes infinite. Therefore, since part of the energy is lost from the antenna andtest systems, the box has, in effect, dissipation. However, because of the volume ofthe box is infinite, its temperature is not changed by this lost energy. In otherwords, the box has an infinite specific heat. In addition, the energy transmitted bythe antenna will generally have different characteristics from that of the Gaussiannoise associated with temperature and even if the average background energy con-tent of the box were changed it could not be properly described by a temperatureparameter.

All the results so far suggest that any linear system can be handled by the samerules that have been developed for systems of oscillators. This will be developedfully in the next section. However, we will assume this to be true now and concludethis section by applying the theorem just derived to obtain Nyquist's result fornoise from a resistor. Take as an example an arbitrary circuit as the test systemconnected to a resistor RT (&) at a temperature T as shown in Fig. 9. The resistorcomprises the interaction system. The interaction between the test system and theresistor is characterized by a charge Q(t) flowing through the test system andresistor and a voltage V(t) across the terminals. Let us associate Q(t) with the coor-dinate of the test system and V(t) with the coordinate of R(&). The interaction partof the Lagrangian is symbolically represented by &Q(t) V(t) since the currentvoltage relationship in R(&) is opposite to that of a generator. The quantity i&Z&

appearing in the influence functional is given by

&[Q& �V&]=&(i&R&)&1=i&Z& .

Thus, it follows that

Re(Z&)&1=&2R& .

Then the results of this section tell us that this situation, as shown in Fig. 9, maybe replaced by a resistor at zero temperature (i.e., a resistor with thermal fluctua-tions appropriate to zero temperature but with the same magnitude of resistance ithas at temperature T ) and a random classical voltage whose power spectrum is

,(&)=�&R&(e;�&&1)&1 (4.50)

as shown in Fig. 10.

FIG. 9. Test system acted on by a linear system represented by an electrical resistance R at a finitetemperature.

582 FEYNMAN AND VERNON

Page 37: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601737 . By:SD . Date:31:03:00 . Time:14:46 LOP8M. V8.B. Page 01:01Codes: 2975 Signs: 1958 . Length: 46 pic 0 pts, 194 mm

FIG. 10. Equivalent interaction system of a resistor at finite temperature acting on the test system Q.

The mean value of this voltage is

(v2(t))=(2�?) |�

0,(&) d&

=|�

02�&R&[?(e;�&&1)]&1 d&. (4.51)

for high temperature ;<<1, and we find that over the frequency range where;�&<<1, ,(&)=kTR& . If R& is constant over the frequency range of interest whoselimits are &2 and &1 , then as a result of the noise power in this range

(v2(t)) =4kTR( f2& f1), (4.52)

where f is the circular frequency. This is the famous Nyquist result for noise froma resistor. Notice that the noise voltage generator must be placed in series with theresistor as a consequence of interacting directly with the coordinate of the testsystem, which in this case is Q(t), the charge flowing in the circuit. If the coor-dinates were chosen such that Q(t) were the coordinate of the resistor and V(t) thatof the test system, then the noise generator would become a current source interact-ing with the voltage V(t) so that the situation would be as shown in Fig. 11, where(i2) =(4kT�R)( f2& f1) in the high temperature limit.

It is worth mentioning, but obvious from the derivation, that if there were manysources of dissipation coupled to the test system, each at different temperatures,then there would be a fluctuating potential associated with each source of dissipa-tion with a power spectrum characteristic to the temperature involved. The caseof different temperatures represents a nonequilibrium condition in that the hotresistors are always giving up energy to the cold ones. However, when the resistorsare represented by continuous distributions of oscillators as in the case of an infinitebox (free space), the temperatures do not change because of the infinite specific heatof the ensemble of oscillators.

FIG. 11. Equivalent interaction system of a resistor at finite temperature when the coordinateassociated with the resistor is the charge flowing through it.

583INTERACTION OF SYSTEMS

Page 38: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

V. WEAKLY COUPLED SYSTEMS

We are now faced with the problem of finding influence functionals whose behavioris in some sense linear but whose total behavior is not representable by systems ofperfect oscillators. There are many examples of this. The concepts of resistance,electric and magnetic polarizations, etc. are basic quantities which characterize theclassical electrical behavior of matter. However, for an accurate description of thisbehavior, these quantities can be constant, independent of the applied electro-magnetic field only in the range of approximation that the magnitude of thisapplied field does not become too large. We will consider here the analysis of suchsystems and the approximations which make linear analysis valid. Insofar as linearbehavior is obtained, the results of this section are basically the same as thoseobtained in previous sections with regard to finding influence functionals. However,it is interesting to see the same principles come out of the analysis another way. Inaddition, it will be found that expressions corresponding to Z& which appear in theinfluence functional are actually closed forms which can be used to compute suchquantities as the conductivity from a knowledge of the unperturbed quantumcharacteristics of a system. These expressions have been derived before by severalauthors but it is interesting to find that they also appear in the influence functionalquite naturally. The results will then be applied to the case of a beam of non-inter-acting particles passing near a test system such as a cavity. This analysis naturallylends itself to a discussion of noise in beam-type maser amplifiers. In Appendix III,the results of this and Section IV are used to compute the spontaneous emission ofa particle in a cavity.

A. Interaction Systems with Coupling Potentials of the Form &V(Q) U(P)

The specific result to be shown here can be stated as follows: If a general interac-tion system, P, is coupled to a test system Q so that the interaction potential issmall and of the form &V(Q) U(P), then the effect of the test system is that of a sum ofoscillators whose frequencies correspond to the possible transitions of the interactionsystem. Therefore, to the extent that second order perturbation theory yields sufficientaccuracy, the effect of an interaction system is that of a linear system.

To show this, we shall first assume the interaction system to be in an eigenstate,a(P{) exp[&(i��) Ea {] at the beginning of the interaction and in an arbitrarystate at the end of the interaction consistent with the usual procedure we havefollowed. Also, for convenience in writing, the interaction potential will be assumed&U(P) Q. The influence functional is then

F(Q, Q$)=| $(PT&P$T) exp {(i��) _S(P)&S(P$)

+|T

{(QU&Q$U$) dt&= ,a*(P${) ,a(P{) dP{ } } } DP$(t), (5.1)

584 FEYNMAN AND VERNON

Page 39: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

where in the above we have written U for U(P) and U$ for U(P$). Since themagnitude of the interaction is assumed to be small, the perturbation approach canbe used to good advantage. Thus, expanding the intersection part of the exponentand keeping terms to second order in Q only,

F(Q, Q$)=| $(PT&P$T) exp[(i��)[S(P)&S(P$)]]

_{1+(i��) |T

{(QU&Q$U$) dt+(i��)2 |

T

{|

t

{(Qt Ut&Q$t U$t)

_(QsUs&Q$sU$s) ds dt= ,a*(P${) ,a(P{) dP{ } } } DP$(t). (5.2)

The evaluation of this is done in an entirely straightforward manner and thefollowing is obtained,19

F(Q, Q$)=1+(iUaa��) |T

{(Qt&Q$t) dt&:

b

|Uab��|2

_|T

{|

t

{(Qt&Q$t)[Qs exp[&i|ba(t&s)]

&Q$s exp[i|ba(t&s)]] ds dt, (5.3)

where

Uab=| ,a*(P) U,b(P) dP and |ba=(Eb&Ea)��.

585INTERACTION OF SYSTEMS

19 A typical term in Eq. (5.2) is as follows:

| $(PT&P$T) exp[(i��)[S(P)&S(P$)]] (i��)2 |T

{|

t

{QtUtQs Us ds dt

_,a*(P${) ,a(P{) dP{ } } } DP(t).

Taking the time integrations outside the path integral and replacing the path integrations by propagat-ing kernels, this becomes

(i��)2 |T

{|

t

{QtQs ds dt | $(PT&P$T) K*(P$T , P${) K(PT , Pt)

_UtK(Pt , Ps) Us K(Ps , P{) ,a*(P${) ,a(P) dP{ } } } dP$T .

Remembering that K(PT , Pt)=� ,n(PT) ,n*(Pt) exp[&iEn(T&t)��], this expression becomes, simply,

&:b

|Uab��| 2 |T

{|

t

{QtQs exp[&i|ba(t&s)] ds dt.

Page 40: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

All terms which involve Uaa will be disregarded. This is because Uaa representsthe average value of the operator U(P) in an eigenstate of the interaction systemalone. Even if it is not zero, it will be noted from Eq. (5.2) that terms involving Uaa

can be written

1+(iUaa ��) |T

{(Qt&Q$t) dt+(1�2!)[(iUaa ��) |

T

{(Qt&Q$t) dt]2+ } } }

=exp {(i��) |T

{Uaa(Qt&Q$t) dt= .

When F(Q, Q$) is used to make a calculation on the test system, this term has theeffect of adding a constant potential V(Q(t))=&Uaa Q(t) to the unperturbed testsystem. Disregarding Uaa , the influence functional then becomes

F(Q, Q$)=1+:a , (5.4)

where

:a=\1�+ |

&�|

t

&�#t#s(Qt&Q$t)[QsFa*(t&s)&Q$s Fa(t&s)] ds dt

and

Fa(t&s)=:b

( |Uba | 2��) exp[i|ba(t&s)]

where the limits have been extended and the factors #t#s inserted to allow finitecoupling time if necessary. If the strength of coupling is sufficiently weak thenEq. (5.4) can be rewritten

Frexp(:a). (5.5)

In this form we recognize F(Q, Q$) as that describing the effect of a sum ofharmonic oscillators independently connected to the test system each of whose``weights'' is UbaUab��2. The complete response function for the system of oscillators is

Ba(t&s)=Im Fa(t&s)=:b

(2 |Uba |2��) sin |ba(t&s),

where the subscript a on Ba refers to the initial eigenstate. According to previousdefinition, the mass of each individual oscillator is identified by m=�(2 |Uba |2 |ba)&1

and its characteristic frequency by |ba . Therefore, to the extent that second orderperturbation theory yields satisfactory accuracy, any system may be considered asa sum of harmonic oscillators. This is equivalent in classical mechanics to thetheory of the motion of a particle having small displacements around an equi-librium position. Its motion, to a first approximation, is also that of a harmonic

586 FEYNMAN AND VERNON

Page 41: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

oscillator, if the first effective term in a power series expansion of the potentialaround that equilibrium position is quadratic in the displacement.

In this part of the discussion we should again point out the motivation forwriting the approximate influence function, Eq. (5.4), in terms of the approximateexponential of Eq. (5.5), apart from the obvious advantage of making the formagree with that of exactly linear systems. Frequently, we deal with a test systemwhich is influenced by another system which is actually made up of a large numberof very small systems. Examples of such an interaction system would be a beam ofatoms or the electrons in a metal. Although the expression for the influence func-tional for any one of the subsystems is only good to second order, their individualeffects are so small that this accuracy is very good and Eqs. (5.4) or (5.5) is equallyvalid. However, when the sum of the effects of the subsystems is not small, then thetwo forms above do not describe the situation equally insofar as the compositeeffect of the interaction system is concerned. We know that when these subsystemsare dynamically and statistically independent, the total influence functional is simply aproduct of the individual ones. In such a case the influence functional obtained byusing Eq. (5.5) as follows

Fk(Q, Q$)rexp {i :k

8k(Q, Q$)=yields much greater accuracy than that obtained from Eq. (5.4) where we wouldfind

F(Q, Q$)r1+i :k

8k(Q, Q$)

8k(Q, Q$) being the influence phase for the k th subsystem (see Appendix IV).The expression given by Eq. (5.5) has additional implications which are not

immediately apparent from the analogy with the harmonic oscillator. No assump-tion was made as to whether the initial state was necessarily the lowest of thesystem. Therefore, |ba could be either positive or negative. Suppose for a momentthat the interaction system has only two states (,a(P) and ,b(P)) and that theinitial state, a, is the lower one (|ba>0). It is obvious that the only effect it canhave on the test system Q is to absorb energy from Q. However, if the initial state,a, is the upper state (corresponding to |ba<0) then the interaction system can onlygive up energy to the test system. It can do this in two ways, through spontaneousemission or through coherent emission due to some coherent driving force exertedon it by the test system. So for the case |ba>0 we expect the influence functionalto show that the interaction system has the effect of a cold system characterized bya dissipative impedance (or positive resistance). Conversely, for |ba<0 it is expectedthat F(Q, Q$) will be characterized by a negative resistance and a random potentialdue to the spontaneous emission transitions. This situation is made more obviousif we translate Eq. (5.5) into transform notation. Thus,

587INTERACTION OF SYSTEMS

Page 42: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

8(Q, Q$)=(2?�)&1 |�

0 _Q$&(Q&&&Q$&&)(i&Zba, &)

+Q&&(Q&&Q$&)(&i&Zba, &&) & d&

&(?�2)&1 |�

0|Q&&Q$& |2 ? |Uba |2 $(&+|ba) d&, (5.6)

where

i&Zbc, &=&�(2|ba |Uba |2)&1 [(&&i=)2&|2ba]. (5.7)

First of all we notice that the sign of Zba, & changes with that of |ba and thereforeits dissipative part can be positive or negative as was argued above. Secondly, for|ba<0 there is a random potential acting on Q whose power spectrum is given by,(&)=? |Uba |2 $(&+|ba). Of course when |ba>0, the integral involving this termdisappears indicating the noise potential does not exist and the effect is the sameas that of a harmonic oscillator initially in the ground state with 1�2m| identifiedwith |Uba |2��.

In a real physical situation it is not likely that the interaction system will be ina definite state initially. So to extend the above results to a more general case weassume that the initial state is described as a sum over states weighted by a densitymatrix \({) which is diagonal in the energy eigenstates of the system. For example,if the system is initially in temperature equilibrium, \=e&;H�Tr(e&;H) where His the Hamiltonian operator such that \mn=$mn exp(&;En)��n exp(&;En). Theinfluence functional becomes

F(Q, Q$)=| $(PT&P$T) exp {(i��) _S(P)&S(P$)

+|T

{(UQ&U$Q$) dt&= :

a

\aa,a*(P${) ,a(P{) dP{ } } } DP$(t). (5.8)

Within the limit of small coupling, then we can simply extend the influence phaseof Eq. (5.5) by summing over all initial states weighted by the initial density matrix\aa . If this is done, we obtain the usual form of the influence phase, Eq. (5.4), witha response function B(t&s) given by

B(t&s)= :a, b

(2\aa |Uab | 2��) sin |ba(t&s). (5.9)

Again 8(Q, Q$) is the phase for a sum of oscillators, each of whose weights is\aa |Uab |2��.20 In Section IV it was shown that B(t&s) was the classical response of

588 FEYNMAN AND VERNON

20 It is interesting to notice that the relative populations of any two levels may be described by aneffective temperature Te=1�k;e . For instance if the probabilities of occupation of states a and b are \aa

and \bb respectively, we use the definition \aa �\bb=exp ;e(Eb&Ea). If \aa=0 this is described by settingTe=0+, meaning to approach zero from the positive side. Similarly, if the two states were inverted\bb=0 and Te=0&. This device has been used widely in the description of such situations.

Page 43: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

the linear system to an impulse of force applied to the coordinate U(P). However,the expression above is in a form which is familiar to us only when we think of theinteraction system, P, as consisting of a sum of oscillators. In this connectionB(t&s) is the total classical response of the oscillators describing the system P toan impulse of force applied to the ``coordinate'' U(P).21 To obtain a more directinterpretation of Eq. (5.9) we now calculate the linear classical response B(t&s) ofthe interaction system to an applied impulse of force in terms of its unperturbedquantum characteristics. In so doing we will show that Eq. (5.9) is indeed thisexpression. Therefore, we will again have the result that the influence functionalfor a general, linear interaction system is formed simply from a knowledge of itsclassical characteristics just as in the case of systems of perfect oscillators. By the``classical characteristics'' of the quantum mechanical system we mean the expectedvalue of U(P) as a function of time after a potential & f (t) U(P) is applied startingat t=0. Thus

(U(Pt)) =| �*(Pt) U(Pt) �(Pt) dPt , (5.10)

where �(Pt) represents the state of system P at t. Using the path integral represen-tation for the development of a wave function with time, as outlined in Section II,this can be written

(U(Pt))=| Ut $(Pt&P$t) exp {(i��) _S0(P)&S0(P$)

+|t

0f (s)(Us&U$s) ds&= �*(P$0) �(P0) dP0 } } } DP$(t).

Alternatively, this expression can evidently be written

(U(Pt)) =&i���f

F( ft ; f $t)| f =f $ .

The initial definite state �0 *(P$) �0(P) will be replaced by an average statedescribed by the following density matrix diagonal in the energy representation,� \aa ,a *(P$0) ,a(P0) so that

�*(P$0) �(P0)=:a

\aa,a*(P$0) ,a(P0).

589INTERACTION OF SYSTEMS

21 U(P) may be regarded as a coordinate which is a function of other coordinates P in terms of whichwe choose to describe the interaction system.

Page 44: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

Assuming f (s) to be small in magnitude, Eq. (5.10) can be written to first order

(Ut) =:a

\aa | Ut$(Pt&P$t) exp[(i��)[S0(P)&S0(P$)]]

_{1+(i��) |t

0fsUs ds&(i��) |

t

0fsU$s ds=

_,a*(P$0) ,a(P0) dP0 } } } DP$(t)

=:a

\aa {Uaa+:b

(i��) |Uab | 2 |t

0fs exp[&i|ba(t&s)] ds

&:b

(i��) |Uab |2 |t

0fs exp[i|ba(t&s)] ds= . (5.11)

Again assuming Uaa=0, Eq. (5.11) becomes

(Ut)= :a, b

(2\aa |Uab |2��) |t

0fs sin |ba(t&s) ds. (5.12)

For f (s)=$(s), then, the classical response function is22

B(t)= :a, b

(2\aa |Uab | 2��) sin |ba t (5.13)

which is identical with the response function found in the influence functional.Since we have found expressions identifiable as classical response functions and

impedances, it remains to show that associated with the dissipative part of theimpedance is a noise potential. The impedance is simply obtained from23

(i&Z&)&1=|

0B(t) e&i&t dt

= :a, b

(&2\aa|ba |Uba | 2��)[(&&i=)2&|2ba]&1. (5.14)

To obtain the power spectrum it is only necessary to sum the influence phase ofEq. (5.6) over all initial states weighted by \a . Thus we find

,(&)= :a, b

?\aa |Uab |2 $(&+|ba) (5.15)

590 FEYNMAN AND VERNON

22 It should be noted that implicit in the use of first order perturbation theory to obtain Eq. (5.13) isthe fact that for this relation for the response function to hold as a steady-state description of the linearsystem, the initial distribution must not be significantly disturbed by the application of the driving force.

23 Expressions of this kind have been used by several authors to compute quantities such as theconductivity of materials. See, for instance, ref. [14].

Page 45: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

and we now wish to relate this to the real part of 1�Z& . From Eq. (5.14) it is foundthat

Re(Z&)&1= :

a, b

(?& |Uba |2��)(\bb&\aa) $(&+|ba). (5.16)

Rewriting ,(&)

,(&)= :a, b

? |Uab | 2 (\bb&\aa)[(\bb �\aa)&1]&1 $(&+|ba). (5.17)

If the average initial state of the interaction system is one of temperature equilibrium,then \aa=e&;Ea��n e&;En and \bb�\aa=e;(Ea&Eb)#e;�|ab. Taking advantage of thecharacteristics of $(&+|ba) so that & can replace |ab , from Eqs. (5.16) and (5.17)24

,(&)= :a, b

(?��) |Uba | 2 &(\bb&\aa) $(&+|ba) �[&(e;�&&1)]&1

=� Re(Z&)&1 [&(e;�&&1)]&1. (5.18)

Therefore, again we find that the power spectrum is related to the dissipative partsof the impedance when the initial state is one of temperature equilibrium.25 Thepower spectrum given in the form of Eq. (5.17) again illustrates the origin of thermalnoise and identifies it as being just another aspect of spontaneous emission. Pound hasalso discussed this [15]. The only contribution to the noise power spectrum is throughthe possible downward transitions of each possible state, a, weighted by the statisticalfactor, \aa .

B. Beam of Particles Interacting with a Cavity

In the cases just considered the interaction system provided a steady-stateenvironment for the test system. In contrast, let us now examine the situation wherethe interaction system is made up of a large number of independent particlescoupled to the test system at different times. As an example consider a beam of non-interacting, identical particles which interact weakly with a resonant cavity as mightoccur, for instance, in a gas maser. We assume that the beam is not necessarily intemperature equilibrium but that the initial state of the particles entering the cavity

591INTERACTION OF SYSTEMS

24 Notice that if P were a two-level system initially in the lower state, then \bb=0 and Te=0+. Inthis case ,(&)=0 which agrees with the required result for &>0 (Eq. (5.6) for |ba>0). If initially P werein the upper state then \aa=0 and Te=0& yielding ,(&)=&(��&) Re(1�Z&) agreeing with Eq. (5.6) for|ba<0. This is the power spectrum of the so-called spontaneous emission noise from an inverted two-level system.

25 It may be disturbing that Re(1�Z&) contains singular forms such as $(&+|ba). However, the infinitesums over the distribution of states of which it is a coefficient can be replaced by integrals over densitiesof states in most practical situations and as part of an integrand $(&+|ba) is not unrealistic.

Page 46: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

would be properly represented by a density matrix diagonal in the energy represen-tation. Such a situation would occur if the beam were prepared by passing itthrough a beam separator whose function would be to eliminate certain particlesfrom the beam depending on their energy levels. For the purposes of simplifying theanalysis we assume the molecules to be two-level quantum systems and that beforethey enter the cavity all of them are in the lower state or all in the upper state. Itis easy to extend these results to the case where the beam is mixed with a certainfraction in the upper state and a certain fraction in the lower state initially. Sincethe beam is assumed to be composed of noninteracting particles, we can considerthe total beam as composed of two independent beams appropriate to the twopossible initial states of the constituent particles. The influence phase for the com-plete beam is simply the sum of the influence phases for the two beams. In addition,we assume that the beam is characterized by a spatial density such that the numberof particles passing a given point along the beam in a time dt is N dt and if t0 isthe time a molecule passes a reference point in the cavity, #(t&t0) describes thecoupling between the molecule and cavity. Thus, the beam is a univelocity beam.Again, in a real case where the beam is characterized by a distribution of velocities,the total beam may be split up into many univelocity beams. The total influencephase is simply a sum of those for each component beam.

Let us call the coordinates of the cavity Q, representing the test system and thecoordinates of a particle in the beam P. The interaction between beam and particlesis given by LI (Q, P)=#(t&t0) QP. Under these circumstances the influence func-tional for the effect of the beam on the cavity can be written down immediately:

i8B(Q, Q$)=&N |Pab��|2 |�

&�|

t

&�ds dt _|

&�#(t&t0) #(s&t0) dt0&

_(Qt&Q$t)(Qse&i|ba(t&s)&Q$sei|ba(t&s)). (5.19)

The integral involving the coupling parameters can be used to define a new function 1,

|�

&�#(t&t0) #(s&t0) dt0 =|

&�#(!) #(!&t+s) d!

#1(t&s). (5.20)

Therefore,

i8(Q, Q$)=&N |Pab ��|2 |�

&�|

t

&�1(t&s)

_(Qt&Q$t)(Qse&i|ba (t&s)&Q$sei|ba (t&s)) ds dt. (5.21)

From this we can identify the response function of the beam as

B(t&s)=(2��) N |Pab |2 1(t&s) sin |ba(t&s). (5.22)

592 FEYNMAN AND VERNON

Page 47: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

In transform notation the influence functional for the effect of the beam has thesame form as has been previously derived with

(i&Z&)&1=(i��) N |Pab |2 (1&&|ba&1&+|ba

) (5.23)

where

1&+|ba=|

&�1(r) 1(r) e&i(&+|ba) r dr (5.24)

and with a power spectrum

,(&)= 12N |Pab | 2 [1&+|ba

+1&&&|ba]. (5.25)

Previously, we have found that for a linear system initially in the ground state thepower spectrum is zero over the range of positive & thus indicating a zero noisepotential due to the linear system. However, ,& is not necessarily zero for &>0 inEq. (5.25) in the case that the beam is initially in its lowest state (|ba>0). This isbecause no restrictions were placed on the time variation of #(t&t0). However,in practical situations the coupling between a cavity and a particle in a beampassing through the cavity varies adiabatically so that for all practical purpose ,(&)is really zero for &>0. To point this out more clearly, let us assume that #(t&t0)=1(t&t0) 1(t0+{0&t), that is, the coupling is turned on at t0 and off at t0+{0 , thetime of transit in the cavity being {0 . We find by evaluating Eq. (5.20) that

1(t&s)=(s&t+{0) 1(s&t+{0)

and from this we can find

1&+|ba=|

&�({0&r) 1(r0&r) 1(r) exp[&i(&+|ba) r] dr

=[1&i{0(&+|ba)&exp[&i(&+|ba) {0]](&+|ba)&2.

Therefore, from Eq. (5.25),

,(&)=( 12) N |Pab | 2 {0

2(sin2 %�%2), (5.26)

where

%= 12 (&+|ba) {0 .

To find the effect of suddenly turning the coupling on and off we find the ratio ofthe total noise power to the noise power for &>0. This is given by

��&� ,(&) d&��

0 ,(&) d&>

��&� (sin2 %�%2) d%

��|ba{0

d%�%2 =?|ba{0 .

593INTERACTION OF SYSTEMS

Page 48: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

For ammonia molecules at a temperature of T=290%K, the average velocityvr6_104 cm�sec. For a microwave cavity of 10 cm length {0=2_104 sec. For the3�3 line of ammonia |ba r1.5_1011 rad�sec. For our case then (?|ba{0)r108.From this, we can conclude that even in this unfavorable case of coupling timevariation, the ,(&) is negligible for &>0.

Examination of Re(1�Z&), derived from Eq. (5.23), reveals terms of the same formas those just discussed,

Re(1�Z&)= 12 [(1�Z&)+(1�Z&*)]

=(&N�2�) |Pab |2 [1&+|ba&1&&|ba

&1&&+|ba+1&&&|ba

]. (5.27)

By the same type of argument as above, the terms in 1&&|bacan be neglected when

&>0, |ba<0 and conversely, when &>0, |ba>0 the terms in 1&+|bacan be neglected.

Since this is the case, we can write

,(&)r� Re(1�Z&)[&(e&;e�|ba&1)]&1, (5.28)

where ;e=1�kTe describes the relative initial populations of the two states. There-fore, we conclude that in most practical cases the power spectrum can be writtenin the form of Eq. (5.28). In cases where the transients cannot be neglected for verylow |ba and very short transit times, however, ,(&) is not so simply related toRe(1�Z&) and must be written in the form given by Eq. (5.25).

It is to be noted that ,(&) of Eq. (5.28) is not precisely of the form for theNyquist relation because of the appearance of e;�|ba in the denominator rather thane;�&. This is a consequence of the finite coupling time between each part of the beamand the cavity which results in a nonequilibrium condition. If the coupling timeswere infinite, the expression for Re(1�Z&) would contain forms such as $(&+|ba),a situation discussed earlier, so that the Nyquist form then results. However, whenthe coupling time is long as in masers,

1&+|ba+1&&&|ba

r4?$(&+|ba)

so that the true Nyquist relation may be used with negligible error.

VI. SOURCES OF NOISE IN MASERS

Having developed the theory of linear systems in detail we are now in a positionto discuss the sources of noise in linear quantum-mechanical devices such as maseramplifiers. The subject of maser noise has been explored by many authors [15�19].The details of the treatment given the subject differ, but the principles are essen-tially the same. The amplifiers are considered as operating at signal levels highenough (classical) that a signal entering a maser may be considered as a group ofphotons whose number is large enough that the amplification process increases thesignal in a continuous fashion. The sources of noise were found to be those derivedfrom the thermal noise arising from the sources of dissipation and those derived

594 FEYNMAN AND VERNON

Page 49: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

File: 595J 601749 . By:SD . Date:31:03:00 . Time:14:47 LOP8M. V8.B. Page 01:01Codes: 3679 Signs: 3133 . Length: 46 pic 0 pts, 194 mm

from spontaneous emission from the ``active'' quantum material. They go furtherand define an effective temperature of the active quantum system so that the noiseit produces is related to the negative resistance of the active materials. Our analysisof linear systems has also shown that these same sources of noise exist. However,if the signal level entering the maser is very small such that its strength can becharacterized by a few quanta per second, a serious question arises as to the natureof the signal out of the maser (here assumed to be at a classical level). An addi-tional fluctuation of the output signal or ``quantum'' noise might be expected duesolely to what might be termed a ``shot noise'' effect created by each individualphoton entering the cavity. It has been shown that no such signal exists by otherauthors [20, 21]. We wish to show how the same result follows from the theorydeveloped here in an extremely transparent way. We will show that the only fluctuationsin output signal which are to be expected are those noise sources computed classically.No additional ``shot noise'' does, in fact, appear.

Let us suppose that we have a beam-type maser amplifier in which all participatingsystems used meet the requirement of linearity. There may be one or more beamsinteracting with various electromagnetic resonators which can be coupled togetherin any way desired. The output of the maser is connected to a detector of some sortwhich perhaps consists of a resistor in which the current is to be measured. Tothe input of the maser system we now apply an incoming classical signal of largemagnitude and of frequency | through an attenuation whose value of attenuationmay be varied at will as shown in Fig. 12. In practice such a situation could ariseif the classical wave originates from a distant antenna with a very large magnitudeof output, so large that all quantum effects in the wave are effectively obscured. Thelong distance would then play the role of the attenuator.

Now, if the classical wave was attenuated by a large amount so that only a fewphotons�sec were entering the maser, the only uncertainty in the signal in the out-put of the maser caused by the maser itself arises from those sources of noise whichcan be arrived at by a classical calculation of the characteristics of the maser. Thereis no extra quantum fluctuation introduced by the maser into the output signal dueto the small number of quanta entering the maser.

It is true that the amplitude of the signal output from the maser might itself beso small that it is still on a quantum level. In this case the detector output wouldbe uncertain due to the inherently small magnitude of the signal from the maser.However, if this is the case, we may put as many amplifiers in series as necessaryto bring the output signal back to a classical level. When this is done the signalapplied to the detector consists of the original signal modified by the transfer

FIG. 12. System in which a classical wave is attenuated to a very low level (a few photons�sec), thenamplified by a maser and detected.

595INTERACTION OF SYSTEMS

Page 50: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

characteristics of the maser system (and attenuator) and noise signals which arisefrom all the possible sources computes classically. The proof of this assertion is notdifficult. We divide up the total system into a test system, here the detector, and aninteraction system which consists of the maser, attenuator, and classical signal,C(t). The, to find the effect of the interaction system on the detector we need onlyto look at the influence functional. However, we already know that it can be writtenas follows,

F(Q, Q$)=exp {\ i2?�+ |

0 _Q$&(Q&&&Q$&&)(i&Z&)

+Q&&(Q&&Q$&)

(&i&Z&&) & d&

+\ i2?�+ |

&� \C&

i&Z&+ (Q&&&Q$&&) d&&\ 1?�2+ |

0|Q&&Q$&|

2 :i

,i (&) d&= ,

where , i (&) represents the power spectrum of the noise from the i th source, z& isthe classical transfer characteristic of the complete interaction system, and Z& is theimpedance of the maser system as seen by the detector. All the terms in theinfluence functional are familiar in view of the derivations which have been presentedpreviously. The first term in the exponent of the influence functional is recognized asdescribing a linear system at zero temperature (Section IV), the linear system in thiscase being the maser. This term describes the spontaneous emission of the detectorback into the maser (see Appendix I). Furthermore, it can be deduced that thisspontaneous emission can be thought of as resulting from a noise generator createdby the detector (test system) acting on the interaction system in the usual classicalway, i.e., whose power spectrum was related to the dissipative part of the detectorimpedance and to the temperature by the Nyquist relation

,(&)=�&R&[exp(�&�kT )&1]&1,

where R& is the detector resistance and T its temperature. The second term in theabove composite influence functional is easily interpreted and is simply the effect ofa classical voltage, related to the input voltage by the classical transfer charac-teristic of the maser, acting on the detector (see Section IV.B). The last termrepresents the effect on the detector of random noise voltages associated with thevarious classical noise sources in the maser (Sections IV.C and V). Both positiveand negative resistances are such noise sources. In either case the power spectrumof the noise from a particular resistance is computed from the same relation asgiven above.

If R& is negative the effective temperature of R& will also be negative alwaysgiving a positive power spectrum. Therefore, if we were to compute the current inthe detector due to the interaction system (maser) using the influence functional wewould find components of current due to: (1) the noise voltage generated by thedetector itself, the power spectrum of which is related to the resistance of the detector

596 FEYNMAN AND VERNON

Page 51: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

by the generalized Nyquist relation given above; (2) a classical voltage relatedto the input voltage C(t) by the classical transfer characteristic of the maser;and (3) random noise voltages associated with the various classical noise sources(resistances) in the maser. Therefore, the maser simply acts as a classical amplifierwith sources of noise which can be predicted from considerations of its classicalcharacteristics.

APPENDIX I

In Section II.E the problem of making perturbation calculations using influencefunctionals was outlined. Here we will calculate in detail the probability, to secondorder in the coupling potentials, that a test system which is in a definite state att={ and finds itself in another state at t=T when interacting with a zero temperatureinteraction system. Let us call the initial and final states ,n(Q{) and ,m(QT) respectivelyand, for simplicity, assume that these are eigenstates. The formal expression for thetransition probability is given by Eq. (4.24) and for this case,

F(Q, Q$)=exp {&(2�)&1 |�

&�|

t

&�#t #s(Qt&Q$t)

_[QsF*(t&s)&Q$s F(t&s)] ds dt= . (I.1)

Then to second order

Pnm r| ,m*(QT) ,m(Q$T) exp[(i��)[S0(Q)&S0(Q$)]]

_{1&(2�)&1 |�

&�|

t

&�#t#s(Qt&Q$t)[QsF*(t&s)&Q$sF(t&s)] ds dt=

_,n*(Q${) ,n(Q{) dQ{ } } } DQ$(t). (I.2)

Making use of the fact that

| exp[(i��) S0(Q)] DQ(t)=K(QT , Q{)

=| K(QT , Qt) K(Qt , Q{) dQt (where T<t<{)

597INTERACTION OF SYSTEMS

Page 52: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

and writing (I.2) as a sum of integrals

Pnm r| ,m*(QT) ,m(Q$T) exp[(i��)[S0(Q)&S0(Q$)]]

_,n*(Q${) ,n(Q{) dQ{ } } } DQ$(t)

&(2�)&1 |�

&�|

t

&�ds dt #s#t | ,m*(QT) ,m(Q$T)

_exp[(i��)[S0(Q)&S0(Q$)]] (Qt&Q$t)[QsF*(t&s)&Q$sF(t&s)]

_,n*(Q${) ,n(Q{) dQ{ } } } DQ$(t). (I.3)

Replacing K(QT , Qt) by �k ,k(QT) ,k*(Qt) exp[&iEk(T&t)��] and taking matrixelements we have

Pnm=$nm _1&:k

(2�)&1 |Qnk |2 f (&nk)&+(2�)&1 |Qnm |2 f (&nm), (I.4)

where

f (&nm)=|�

&�|

t

&�#t#s[F*(t&s) ei&nm (t&s)+F(t&s) e&i&nm (t&s)] ds dt. (I.5)

f (&) can be simplified by restating the integral over t and s in terms of frequency.To do this we replace the upper limit t by +� and multiply the integrand by a stepfunction 1(t&s). Then utilizing the convolution theorem in the form

|�

&�|

&�M(t) N(s) R(t&s) ds dt=(2?)&1 | M&&N&R& d&,

where

M(&)=|�

&�M(t) e&i&t dt.

f (&) becomes

f (&)=(2?)&1 |�

&�|#&&| |2 (F&+F&*) d&,

where

F&=|�

&�1(t) F (t) e&i&t dt.

598 FEYNMAN AND VERNON

Page 53: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

From Eqs. (4.5a) and (4.5b) we have that

F(t)=&(2�?) |�

0Im(1�i'Z') ei't d'.

Therefore,

F&+F&*= &(2�?) |�

0d' Im(1�i'Z') |

0dt(ei('&&) t+e&i('&&) t)

= &4 Im(1�i&Z&)=(4�&) Re(1�Z&)

and

f (&)=(2�?) |�

&�|#'&& |2 '&1 Re(1�Z') d'. (I.6)

Thus, to second order

Pnm =$nm {1&:k

(?�)&1 |Qnk |2 |�

0&&1 |#&&&nk

| 2 Re(Z&)&1 d&=+(?�)&1 |Qnm |2 |

0&&1 |#&&&nm

|2 Re(Z&)&1 d&. (I.7)

For the special case that the coupling # is

#=0 for t>T�2 and t<&T�2

#=1 for &T�2<t<T�2

|#&&&nk| 2= } |

T�2

&T�2e&i(&&&nk) t dt }

2

=4(&&&nk)&2 sin2(&&&nk) T�2 � 2?T$(&&&nk) for large T.

Then,

Pnm =$nm {1& :k, &nk>0

2T(�&nk)&1 |Qnk |2 Re(Z&nk)&1=

+2T(�&nm)&1 |Qnm |2 Re(Z&nm)&1 for &nm>0

=0 for &nm<0. (I.8)

Thus, it is seen that the Pnm now is proportional to the matrix element and to thedissipative part of the impedance Z&nm

. Appropriately enough, no transition ispossible to energy states such that &nm<0 since the interaction system, being atzero temperature initially, can give up no energy to the test system.

599INTERACTION OF SYSTEMS

Page 54: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

APPENDIX II

A. Influence Phase for Effect of Free Space on an Atom

As an illustration, the influence phase for the effect of free space on an atom willbe calculated. This problem is more complicated than the idealized systems con-sidered in deriving the formalism since the interaction here is of the form Q } X, Qand X being vectors rather than QX where Q and X are scalars. This difficultycould be overcome by writing the influence phase in tensor notation or by recastingthe problem so that the interaction is of the form QX. The latter will be done toadhere more closely to the point of view of the derivations. Since a linear systemis being dealt with, it is only necessary to determine a suitable coordinate for theatom and find the impedance function Z& for the effect of free space. It is assumedthat the atom is made up of a system of particle of mass mn , charge en , andposition rA+xn where rA is the position of the center of charge of the atom. If thetransverse part of the radiation field in the box is expanded into a series of planewaves each representing independent harmonic oscillations [22], then the non-relativistic Lagrangian for the complete system consisting of the atom and the fieldin the box can be written [2]

L(X4 , X, q* k , qk , t)=LA+c&1 :n

en X4 nAtr(rA+rn)

+ 12 :

k

:4

r=1

[(q* (r)k )&c2k2(q (r)

k )2], (II.1)

where LA is the Lagrangian of the atom unperturbed by outside forces and

Atr(X)=(8?c2)1�2 :k

[e*(q (1)k cos(k } X)+q (3)

k sin(k } X))

+el (q(2)k cos(k } X)+q (4)

k sin(k } X))]. (II.2)

Here e* and el are two mutually orthogonal polarization vectors, each orthogonalto the propagation vector k. Now we assume that the radiation field of the box isconstant over the particle, i.e., that A(rA+rn)rA(rA), the dipole approximation.26

This permits one to replace �n enX4 n by j, the current operator for the atom. Inaddition, even though e* } el=0, which fixes their relative orientations, their absolutedirections in a plane perpendicular to k are still arbitrary. Choosing el so that

el } j=0. (II.3)

600 FEYNMAN AND VERNON

26 This is equivalent to taking A(kA+xn), expanding it in a series of k } xn since this is assumed small,and keeping only those terms which keep the interaction term of the Lagrangian linear. Since the inter-action is of the form enxn } A(kA+xn) for the n th particle, then A can only contain constant terms.

Page 55: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

We assume the box to be very large so that �k � (16?3)&1 � d 3 k, because the modecorresponding to k and &k is the same. Combining (II.1), (II.2), and (II.3) thefollowing total Lagrangian is obtained:

LT =LA+ :r=1, 3

|d 3k16?3 _1

2(q* (r)

k )2&12

k2c2(q (r)k )2&

+(8?)1�2 |d 3k16?3 (j } e*) q (1)

k cos(k } rA)

+(8?)1�2 |d 3k16?3 (j } e*) q (3)

k sin(k } rA). (II.4)

Thus, the number of each of the two sets of oscillators (q (1)k and q (3)

k ) in a volumeof k space d 3k is d 3k�16?3. The coupling strength of the q (1)

k and q (3)k oscillators

with the atom is (8?)1�2 (j } e*) q (1)k cos(k } rA) and (8?)1�2 (j } e*) q (3)

k cos(k } rA) respec-tively. If j is oriented along the %=0 axis in polar coordinate representation,j } e*= j sin %. Then choosing j as the atom coordinate, the impedance for the twooscillators of frequency kc can be found from the rule found in Section IV to be

[i&Z&(k)]&1=(1�j&)[(8?)1�2 sin % cos(k } rA) q (1)k +(8?)1�2 sin % sin(k } rA) q (3)

k ]&

=&8? sin2 % cos2(k } rA)

(&&i=)2&k2c2 &8? sin2 % sin2(k } rA)

(&&i=)2&k2c2

=&8? sin2 %[(&&i=)2&k2c2]&1| = � 0 . (II.5)

The total effect of all the oscillators is

(Z&)&1=(16?3)&1 | Z&1& (k) d 3 k=(16?3)&1 | k2 sin % d% d, dk(Z&1

& )

=&i4&(3?c3)&1 |�

002 d0[(&&i=)2&02]&1| = � 0

=(2&2�3c3)&i(4&�3?c3) |�

002 d0(&2&02)&1, (II.6)

where the substitution 0=kc has been made. Thus, the effect of free space ischaracterized by (i&Z&)&1. The equivalent distribution of oscillators coupled to j& is

G(0)=(402�3?c3). (II.7)

B. Spontaneous Emission Probability of an Atom in Free Space

To compute the transition probability for this atom, we use second order pertur-bation theory developed in Appendix I for a system initially in state ,n(X{) and

601INTERACTION OF SYSTEMS

Page 56: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

finally in state ,m(XT) when acted on by an influence functional for a linear systemat zero temperature. The expression is

Pnm =2T | jnm |2 (�&nm)&1 Re(Z&nm)&1, &nm>0, n{m

=0, &nm<0. (II.8)

From Eq. (II.7) we find

Re(Z&nm)&1=(?�2) G(&nm)=2&2

nm �3c3. (II.9)

Using this in Eq. (II.8)

Pnm =4&nm | jnm |2 T�3�c3

=4&3nme2 |Xnm |2 T�3�c3, (II.10)

where to obtain the last, more familiar form, the substitution jnm=&nmXmn has beenmade. This is the first order spontaneous emission probability for an atom in free space.

Now we can form an expression for the intensity of radiation per unit time. Thepower radiated from the dipole is

�&nmPnm�T=2&2nm e2 |Xnm |2 Re(Z&nm)&1

=4e2 |Xnm | 2 &4nm �3c2, (II.11)

an expression which is almost the same as that for power radiated from a classicaldipole. The expression becomes exactly the same if we apply the correspondenceprinciple by replacing the matrix element of the time average of the coordinate ofthe oscillator by its corresponding classical quantity. Thus, if X is the coordinate ofthe corresponding classical oscillator (X0 is its maximum value) then27

2 |Xnm |2 � (X2)= 12X0

2.

602 FEYNMAN AND VERNON

27 That this correspondence is true can be seen easily as follows. Consider the dipole, a harmonic oscillatoras above, to be in a high quantum state, ,n . Classically the motion of the dipole can be described as X=X0 sin |t. We wish to relate the classical value of (X2) to its matrix element. In a quantum mechanical sense,

(X 2)=| ,n*(X) X2,n(X) dX

=|| ,n *(X) X :k

,k(X) ,k*(X$) X$,n(X$) dX dX$.

Since matrix elements exist, in the case of a harmonic oscillator only for k=n&1, k=n+1, we have

(X 2) =|Xn, n&1| 2+ |Xn, n+1|2.

For very high quantum numbers these two terms become nearly equal since

|Xn, n&1| 2=n��2m|, |Xn, n+1| 2=(n+1) ��2m|.

Thus, as n � � (X 2)=2 |Xn, n&1| 2. But in the classical case, X� 2= 12 X0

2. Therefore, |Xn, n&1|2 O 14X0

2.

Page 57: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

If |Xmn |2 is replaced by X02�4 then Eq. (II.11) becomes the expression for the power

radiated from a classical dipole. Our purpose, however, in doing this example, wasto show for a specific problem that the effect of a distribution of oscillators interact-ing on a system is the same as the effect of loss on the system.28 This has been doneby relating the energy lost from the radiating dipole to the distribution. It is notsurprising that a sea of oscillators should give this effect. If the dimensions of thebox are allowed to be finite then energy emitted from the system under observationis reflected from the walls and eventually finds its way back to be absorbed again.This is equivalent to saying that the number of oscillators comprising the electro-magnetic field in the box is infinite with a finite frequency spacing between themodes. Since the oscillators are independent there is no coupling between them andenergy coupled into one of the oscillators from the test system must eventuallyreturn to it. If the dimensions of the box are allowed to get infinitely large, energyemitted from the test system never gets reflected and thus never returns. In oscillatorlanguage this means that the frequency spacing between oscillators has becomeinfinitesimal, so close that a little of the energy absorbed by each one graduallyleaks into nearby modes and eventually is completely gone.

APPENDIX III: SPONTANEOUS EMISSION OF AN ATOM IN A CAVITY

In this calculation as in the free space calculation the dipole approximation willbe used in computing the spontaneous emission probability. The linear coordinatesinside the cavity will be represented by the vector Q while the time varying coor-dinates of the single cavity mode being considered will be X(t). The Lagrangian ofthe system may be written

LT #L(Q4 n , Qn , t)+c&1 :n

enQ4 n } A(Ql+Qn , t)+Lcavity , (III.1)

where Qp is the atom coordinate, Qn+Qp is the particle coordinate in the atom,and A is the vector potential of the cavity field. The interaction term is the one of

603INTERACTION OF SYSTEMS

28 If the power radiated from the oscillator is related to the classical expression 12I 2R then from

Eq. (II.11) it can be seen that R is proportional to Re[1�Z(&)] which in turn is related to the distribu-tion of oscillators. One might expect Im[1�Z(&)] to be replaced to the reactance seen by an oscillatingdipole, a quantity which is known to be infinite classically. From Eq. (II.7),

Im(1�Z&)=4&�3?c3 |�

0

02(&2&02)&1 d0=4&0�3?c3| 0 � � .

The integral is linearly divergent. This factor is also related to the infinite self energy of a point chargewhich occurs both classically and in quantum electrodynamics. Here this divergence does not bother ussince it never enters into the calculation.

Page 58: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

interest, since from it we find the terms that we wish to solve for classically. Thisterm will be put into more convenient form. Let us write

A(Q, t)=a(Q) X(t), (III.2)

where

| a(Q) } a(Q) d 3Q=4?c2. (III.3)

If A does not vary much over the atom, then A(Qp)rA(Qp+Qn) and the inter-action term is written

c&1 :n

en Q4 n } a(Qp) X(t)=( j�c) |a(Qp)| X(t), (III.4)

where j=(� enQ4 n) } a(Qp)� |a(Qp)|, the component of the atom current in thedirection of the cavity field.

Let us now determine the ratio

j&[|a(Qp)| X&c&1]&1=i&Z& (III.5)

classically. The wave equation appropriate for this calculation (high Q)29 is

{2A&(1�c2) A� +(|�c2Q) A4 &(4?�c) P4 =0. (III.6)

The atom is located at Qp and, since the dipole moment is induced, its direction onthe average is the same as that of the field A in the cavity. We have then

P4 = j$(Q&Qp) a(Qp)�|a(Qp)|. (III.7)

Substituting (III.7) and (III.2) into (III.6) we obtain

[&|2X+(|�Q) X4 &X� ] c&2a(Q)&(4?�c) j$(Q&Qp) a(Qp)�|a(Qp)|=0, (III.8)

where | is the resonant frequency of the cavity. Multiplying by a(Q), integratingover Q, and taking Fourier transforms, (III.8) becomes

(&2&|2+i&|�Q) X&&c&1 |a(Qp)| j&=0. (III.9)

We find that the ratio (III.5)

i&Z&=[&2&|2+(i&|�Q)] c2�|a(Qp)|2.

604 FEYNMAN AND VERNON

29 Q is used here as the dissipation factor of the cavity, |0 L�R, while Q is a vector representating thelinear coordinates inside the cavity.

Page 59: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

The influence phase for this is (although it is unnecessary to write it)

8( j, j $)=1

2?� |�

0 _j $&( j&&& j $&&)(i&Z&)

+j&&( j&& j $&)(&i&Z&&) & d&.

From second order perturbation theory we know

Pnm=2T | jnm | 2 (�nm)&1 Re(Znm)&1 for &nm>0.

Noting that

Re(Z&nm)&1=|&2nm |a(Qp)| 2 Q&1c&2[(&2

nm&|2)2+|2&2nm Q&2]&1

and defining a cavity from factor f 2=V |a(Qp)|2�4?c2

Pnm=(8? | jnm |2 f 2|&nmT )(�VQ)&1 [(&2nm&|2)+|2&2

nmQ&2]. (III.10)

At resonance this expression reduces to

Pnm=8? | jnm |2 f 2QT��V&2nm . (III.11)

The quantity usually computed is the ratio of the transition probability in thecavity at resonance to that in free space. This ratio is

Pnm(cavity)Pnm(free space)

=[8? | jnm |2 Qf 2�&2

nmV�] T[4 | jnm |2 &nm�3�c3] T

=6?c3Qf 2

V&3nm

.

At resonance, the ratio increases with respect to Q as one might expect anddecreases with respect to the cavity volume V and &3

nm . This expression agrees withthe one given by E. M. Purcell [23] although the form factor in his calculation wasleft out. This does not matter since for a particle located near the maximum fieldpoint in a cavity the magnitude of f is on the order of unity.

APPENDIX IV

It is to be demonstrated that

G(X)= limN large

`N

k=1

(1+Xk) � exp _:k

Xk& , (IV.1)

605INTERACTION OF SYSTEMS

Page 60: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

where the Xk are small but not necessarily equal to each other and where the totalsum �k Xk is finite. Rewriting the expression for G(X) we have (where the summa-tions on all indices go from 1 to N)

G(X)=1+:k

Xk+ 12 :

j{k

XjXk+( 16) :

j{k{l

XjXk Xl+ } } }

=1+:k

Xk+ 12 ! :

j, k

Xj Xk(1&$ jk)

+( 13 !) :

j, k, l

XjXkXl(1&$ jk&$kl&$ jl+2$ jk$kl). (IV.2)

As N is allowed to get very large the contribution of the terms involving quantitiessuch as $jk becomes less significant. For instance, in the third term

:j, k

(Xj Xk)(1&$ jk)t(NX)2&(NX)2�N (IV.3)

and for very large N only the leading term in this sum is important. Thus, we havethe result that

G(X) large N r1+:k

Xk+ 12 ! \:

k

Xk +2

+ 13 ! \:

k

Xk+3

rexp _:k

Xk & . (IV.4)

REFERENCES

1. R. P. Feynman, Rev. Mod. Phys. 20 (1948), 367.2. R. P. Feynman, Phys. Rev. 80 (1950), 440.3. R. P. Feynman, Phys. Rev. 84 (1951), 108.4. V. Fano, Rev. Mod. Phys. 29 (1957), 74.5. R. W. Hellwarth (private discussion).6. R. McWeeny, Rev. Mod. Phys. 32 (1960), 335.7. R. C. Tolman, in ``The Principles of Statistical Mechanics,'' Chap. 9, Oxford Univ. Press, New York,

1930.8. D. Middleton, ``An Introduction to Statistical Communication Theory,'' McGraw�Hill, New York,

1960.9. D. Middleton, in ``An Introduction to Statistical Communication Theory,'' pp. 362�364, McGraw�Hill,

New York, 1960.10. H. Goldstein, in ``Classical Mechanics,'' Chap. 10, Addison�Wesley, Reading, MA, 1953.11. H. B. Callen and J. H. Welton, Phys. Rev. 83 (1951), 34.12. R. Kubo, J. Phys. Soc. Japan 12 (1957), 570.13. H. Nyquist, Phys. Rev. 32 (1928), 110.14. H. Nakano, Progr. Theoret. Phys. (Kyoto) 17 (1957), 145.15. R. V. Pound, Ann. Phys. (N.Y.) 1 (1957), 24.16. M. W. Mueller, Phys. Rev. 106 (1957), 8.17. K. Shinoda, T. C. Wang, and C. H. Townes, Phys. Rev. 102 (1956), 1308.

606 FEYNMAN AND VERNON

Page 61: Feynman and Vernon-The Theory of a General Quantum System Interacting with a Linear Dissipative System

18. M. W. P. Stranberg, Phys. Rev. 106 (1957), 617.19. K. Shinoda, H. Takahasi, and C. H. Townes, J. Phys. Soc. Japan 12 (1957), 686.20. W. H. Wells, Ann. Phys. (N.Y.) 12 (1961), 1.21. J. P. Gordon, Proc. IRE 50 (1962), 1898.22. E. Fermi, Rev. Mod. Phys. 4 (1932), 87.23. E. M. Purcell, Phys. Rev. 69 (1946), 681.

607INTERACTION OF SYSTEMS