Você está na página 1de 12

Chapter 2

Non-equilibrium Thermodynamics
2.1 Response, Relaxation and Correlation
Atthebeginningofthe21stcentury,thethermodynamicsofsystemsfarfromequilibriumremains
poorly understood. However, it turns out that many nonequilibrium phenomena can be described
rather well in terms of equilibrium uctuations; this is especially true of systems near equilibrium
[1, 2].
By designating a system as near equilibrium, we mean that the system is perturbed from its
equilibrium state by some time-dependent external force f(t). The external force is deterministic,
not random; typical examples include mechanical forces and forces due to an applied electric or
magnetic eld. This force drives the expectation values of some of the systems observables away
from their equilibrium values. For example, a typical observable A aected by the external force
might be the systems velocity or its magnetic moment. If the response of the observable A to the
external force f(t) satises the linearity property
A(f(t), t) =A(f(t), t) (2.1)
where A=A A
eq
and describes the strength of the force, then we call the time-dependent
behavior of A the linear response of A to the external force f(t). The linearity property Eq.(2.1)
impliesthatthe shape of theresponsecurve Avs. tisindependentofthe valueofinthe caseof
linear response.
After achieving a short-lived nonequilibrium steady state (between t
2
and t
3
in Figure 2.1), the
system is allowed to relax back to equilibrium. This process is also known as regression. Linear
response and regression of a system driven from equilibrium are both described in terms of the
time correlation function of the observable A, and so we turn rst to the denition and properties
of the time correlation function [3, 4].
The time correlation function C
AA
(t,t

) of the observable A is dened by


Tr[A(t)A(t

)
eq
]
C
AA
(t,t

) =A(t)A(t

)= (2.2)
Tr[
eq
]
Here,
eq
denotes the equilibrium density matrix of the system; hence the average denoted by
is the ensemble average. This function describes how the value of A at time t is correlated to its
1
2 Chapter 2. Non-equilibrium Thermodynamics
Figure 2.1: Response, nonequilibrium steady state, and relaxation
value at time t

; it is sometimes referred to as the autocorrelation function of A to distinguish it


from correlation functions between A and other observables.
For a system which is time-translational invariant, we often choose for convenience to set t

= 0
and to drop the subscript on C
AA
, so that the time correlation function becomes simply
C(t) =A(t)A(0) (2.3)
The correlation function may in general take on complex values. This result is in keeping with
ourphenomenologicalunderstandingofquantummechanicsinthefollowingway. Inordertomea-
sure the correlation function of an observable A, the quantity A must be measured twice (rst at
time zero, then again at time t). However, the rst measurement at t = 0 collapses the system
wavepacket, and the state that would have been exhibited by the unperturbed system at time t
becomes irrecoverable.
We now identify some important features and properties of correlation functions.
1. All inner products X|Y satisfy the Schwarz inequality
|X|Y|
2
X
2
Y
2
(2.4)
Thus the correlation function for any relaxation process has the property
C
2
(t) =|A(t)|A(0)|
2
A
2
(t)A
2
(0) A
2
(0)
2
=C
2
(0) (2.5)
The second inequality above arises from the fact that A
2
(t)< A
2
(0) for relaxation processes
when t >0. More concisely, the Schwarz inequality implies that
|C(t)| C(0) (2.6)
2. Correlationfunctionsaretime-invariant,thatis,theirvaluedependsonlyonthetimeinterval
between the two measurements of the observable:
A(t)A(0)=A(tt
0
)A(t
0
)=A(0)A(t) (2.7)
5.72, Spring 2008 J. Cao
3 Chapter 2. Non-equilibrium Thermodynamics
3. Time-invariance imparts the following identities on the time derivative of a time correlation
function:
C

(t) =A

(t)A(0)=A(0)A

(t)=A(t)A

(0) (2.8)
4. IftheequilibriumvalueofAisA
eq
=0,thenthelong-timelimitofthecorrelationfunction
is zero,
limA(t)A(0)=A
eq
A(0)=0 (2.9)
t
5. Forquantum systems, the time-invariance properties imply that C(t) = C

(t). In the
classicallimit,thecorrelationfunctionisalwaysreal-valued,sothisrelationbecomesC(t) =
C(t)andC(t)isthuseven. Thefactthatclassicalcorrelationfunctionsarereal-valuedshould
seem sensible because we can (and do) measure correlation functions every day for classical
systems, for example, when we try to steady a cord dangling from the ceiling. In this case,
wedeterminetheappropriatetimeandplacetoapplyanexternalsteadyingforcebylooking

fortimecorrelationsbetweenthevariousmotionsthecordundergoes. NotethatC(t)isodd

with C(0)=0 for classical time correlation functions.


6. Forergodicsystems,thetimecorrelationfunctioncanbecalculatedasatimeaverageinstead
of an ensemble average:

A(t)A(0)= lim
1
A(t+

)A(

)d (2.10)

0
Since most systems amenable to analysis by the methods of statistical mechanics are inher-
ently ergodic, we are generally free to choose whichever formulation is easier to work with.
The time average is often easier to implement experimentally because it only requires inte-
gration along a trajectory rather than a simultaneous sampling of every state accessible to
the system.
Example: The classical linear harmonic oscillator with mass m and frequency obeys the equation of
motion
x +
2
x= 0
If we provide initial conditions x(0) and x (0) = v(0), then this equation of motion has the closed-form
solution
v(0)
x(t) =x(0)cost+ sint

Taking the inner product of x(t) with the initial value x(0), we nd
v(0)x(0)
x(t)x(0)=x
2
(0)cost+ sint

The second term is zero because x(0)=0, so the time correlation function is just
C(t) =x
2
(0)cost
kT
Finally,invokingtheequipartitionresultx
2
(0)= ,wherek istheBoltzmannconstant,thecorrelation
m
2
function for the classical linear harmonic oscillator is
kT
C(t)= cost
m
2
5.72, Spring 2008 J. Cao
4 Chapter 2. Non-equilibrium Thermodynamics
2.2 Onsager Regression Theory
Atrstglance,therelaxationofmacroscopicnon-equilibriumdisturbancesinasystemmightseem
completelyunrelatedtotheregressionofmicroscopicuctuationsinthecorrespondingequilibrium
system. However, they are intimately related by so-called uctuation-dissipation theorems. The
existenceofthislinkbetweenmicroscopicuctuationsandmacroscopicrelaxationwasconjectured
byLarsOnsagerin1931,sometwentyyearsbeforeitwasnallyproventobetrue;henceitisoften
referred to as the Onsager regression hypothesis.
To formulate the hypothesis, we consider an observable A with A
eq
= 0 that takes on a non-
equilibriumaveragevalueAduetoanappliedexternalforcef whichactsduringthetimeinterval
t0 but becomes identically zero for t >0.
For t0, the ensemble average of A can be expressed as
Ae
(HfA)

[ ]
A= f A(0)A(0) A(0)
2
=fC(0) (2.11)
(HfA)
e
wheretheapproximationbeingmadeistruncationoftheTaylorseriesforeachexponentialtorst
order.
For t > 0, the system evolves according to H instead of H fA, so A is no longer stationary,
but acquires a time-dependence:
(HfA)
A(t)e
[ ]
A= f A(t)A(0) A(0)
2
=fC(t) (2.12)
(HfA)
e
Onsagershypothesisstatesthattherelaxationofthenon-equilibriumvalueofAisrelatedtoits
value at t=0 in the same way that the time correlation function for a spontaneous uctuation is
related to its value at t=0:
A(t) C(t)
= (2.13)
A(0) C(0)
Example: ThetransitionstatetheoryofchemicalkineticscanbeformulatedthroughtheOnsagerrelation
weve just presented. Consider a chemical equilibrium established between two species A and B,
k
f
AB
k
b
with forward rate constant k
f
and backward rate constant k
b
.
1. Equilibrium popuations WecandescribethepopulationdynamicsofAandBdeterministicallyinthe
macroscopic limit through a pair of coupled dierential equations,
{

P
A
= k
f
P
A
+k
b
P
B

P
B
= k
f
P
A
k
b
P
B
The equilibrium state of this system satises the detailed balance condition
k
f
P
A
=k
b
P
B

5.72, Spring 2008 J. Cao


5
x
Chapter 2. Non-equilibrium Thermodynamics
where the angle brackets denote the equilibrium values of the populations. Taking the populations to be
normalized to unity, P
A
+P
B
=1, we can express P
A
in terms of the rate constants:
P
A
k
b
P
A
= =
P
A
+P
B
k
f
+k
b
For notational simplicity, we introduce k=k
f
+k
b
and refer to the equilibrium populations P
A
and P
B

byq
A
andq
B
,respectively. Withthisnewnotation,wecanexpresstheequilibriumpopulationsofAandB
as

k
b
q
A
=
k
k
f
q
B
=
k
If the initial state is all species A, the solution to the coupled dierential equations indicates a decay to
equilibrium with rate constant k, which we can write in terms of P
A
(t) =P
A
(t)q
A
as
P
A
(t) = P
A
(0)e
kt
Settingthisresultasideforamoment,notethatifweconsidertheenergiesofspeciesAandBtobepotential
wells connected along a reaction coordinate x, then we can write down an expression for the uctuation in
occupation number n for each species as a function of x. The barrier between the A and B potential wells
is a maximum at x=x
b
; see Figure 2.2.
Figure2.2: ProjectionofthepotentialenergysurfaceconnectingspeciesAandBalongreactioncoordinate
2. Application of Onsager Regression hypothesis To reect the fact that a particle to the left of the
barrierisspeciesAandaparticletotherightisspeciesB,wewritetheoccupationnumbersintermsofthe
Heaviside step function,
{
n
A
=(x
b
x)
n
B
=(xx
b
)
where n
A
= q
A
and n
B
= q
B
. Applying Onsagers regression hypothesis to this example, we can relate
the dissipation of P
A
to the uctuations in occupation number as follows:
C(t) n
A
(t)n
A
(0) P
A
(t)
kt
= = =e
C(0) n
2
(0) P
A
(0)
A
5.72, Spring 2008 J. Cao
6 Chapter 2. Non-equilibrium Thermodynamics
The second equality arises from our integrated rate equation for the dissipation of P
A
. Also note that
2 2
n
2
A
n
A

2
=q
A
q =q
A
q
A
(1q
B
) =q
A
q
B A
=n
A
Dierentiating the uctuation-dissipation relation above with respect to t and invoking the identity just
shown, we nd
n
A
(t)n
A
(0) n
A
(t) n
A
(0)
ke
kt
= =
n
2
(0) q
A
q
B
A
Recasting this equation in terms of the reaction coordinate x, we arrive at an expression for the time
dependence of the forward rate constant k
f
(t),
(x(t)x
b
)(x
b
x(0))v
k
f
(t) =k
f
e
kt
=
(x
b
x(t))
where v=n
A
(0) is the initial rate of reaction.
3. Expression for the TST rate constant Finally, to determine the transition state theory (TST) rate
constant, we consider our time-dependent expression for k
f
in the short-time limit, since transition states
typically only survive a few molecular vibrations. In this limit,
(x(0
+
)x
b
)(x
b
x(0))v (v)(x
b
x(0))v
lim k
f
(t) = =
t0
+
n
A
n
A

From the kinetic theory of gases, we recognize that

k
B
T
(v)v= = (2m)
1/2
2m
IfwestipulatenowthattheheightofthebarrierisE
b
,somerearrangementoftheprecedingformulasreveals
that

(x
b
x) m
2

E
b
= e
(x
b
x) 2
where is the fundamental frequency of the left potential well. It follows that the TST rate constant takes
on the simple form

E
b
k
TST
= e
2
To conclude our excursion into TST kinetics, note that the ratio
k(t) (x(t)x
b
)(x(0)x
b
)v
=
k
TST
(x(0
+
)x
b
)(x(0)x
b
)v
is always less than or equal to one. This result indicates that the TST ux is partially trapped in the
product well while part of the TST ux recrosses back to the reactant state. This result is in keeping
with our intuition of chemical dynamics in that every macroscopic reaction is, to some degree, a process of
establishing equilibrium rather than a perfect ow from all reactants to all products.
2.3 Linear Response Theory and Causality
The concept of linear response was introduced in section 2.1. Here, we explore further how the
linear response of a system is quantied by considering the important relations regularly invoked
by practitioners of linear response theory.
5.72, Spring 2008 J. Cao
7 Chapter 2. Non-equilibrium Thermodynamics
2.3.1 Response Functions
The motivating idea behind linear response is that the response of a system to an external force
dependsonthestrengthofthatforceatall timesduringwhichtheforceactsonthesystem. Thatis,
theresponseattimetdependsonthehistoryoftheforcesactiononthesystem. Anappropriately
weighted sum of the strength of the external force at each moment during the interaction will
describe the overall response. Mathematically, therefore, we express the response as an integral
over the history of the interaction,

A(t) = K(t,)f()d (2.14)

The kernel K(t,) in this expression, which provides the weight for the strength of the external
force at each time, is called the response function. The response function has two very important
properties:
Timeinvariance: K dependsonlyonthetimeintervalbetween andt,notonthetwotimes
independently.More succinctly,
K(t,) =K(t) (2.15)
Causality: Thesystemcannotresponduntiltheforcehasbeenapplied. Thisplacesanupper
limit of t on the integration over the history of the external force.
With these observations in place, we arrive at the standard formula describing the linear response
of an observable A to an external force f(t),

t
A(t) = K(t)f()d (2.16)

Linear response described by the response function K(t) and linear regression described by
the time correlation function C(t) are directly related to one another. To see the connection,
consider a force f(t) which is constant with strength f for t 0 and is zero for t > 0. We have
established two ways to describe the response of an observable A to this force:
Linear regression: A(t) =fC(t)

0
Linear response: A(t) =

K(t)f()d
Fromthisinformation,weconcludethatthecorrelationfunctionandresponsefunctionarerelated
by
K(t) =C

(t)(t) (2.17)
where (t) is the Heaviside function.
Sometimesthelinearresponsefunctionismoreconvenientlyexpressedinthefrequencydomain,in
which case it is called the frequency-dependent response function. In many physical situations, it
plays the role of a susceptibility to a force and consequently is denoted by (),

it
K(t)dt () = e (2.18)
0
Thisresponsefunctionisoftenpartitionedintorealandimaginaryparts,whichcanalsobethought
of as even and odd parts, respectively,
() =

() +i

() (2.19)
5.72, Spring 2008 J. Cao
8 Chapter 2. Non-equilibrium Thermodynamics
Example: The response function for the classical linear harmonic oscillator can be quickly deduced from
itstimecorrelationfunction. Recallfromtherstexampleinthischapterthatthetimecorrelationfunction
for the classical linear harmonic oscillator is
k
B
T
C(t)= cost
m
2
1
Applying Eq.(2.17), we dierentiate with respect to t and multiply by =
k
B
T
to determine that
1
K(t)= sin (t)(t)
m
This is the response function for the classical linear harmonic oscillator.
2.3.2 Absorption Power Spectra
Thefrequency-dependentresponsefunctionisdirectlyrelatedtotheabsorptionspectrum: infact,
knowledge of () and the time-dependent external force f(t) is sucient to fully describe the
absorption spectrum.
The rate at which work is done on a system by a generalized external force f(t) is f(t)A

(t),
where A is the observable corresponding to the generalized force f. This quantity has units of
power, so we can calculate the total absorption energy by integrating this power over time,

P(t)dt= f(t)A

(t)dt (2.20)
Torecastthisresultintermsof(),werstconsidertheFouriertransformofthetime-dependent
observable A,

i
A

() = e
t
A(t)dt (2.21)
Applying Eq.(2.16), we have

it
A() = e K(t)f()ddt (2.22)

The following rearrangements allow us to express A() entirely in terms of frequency-dependent


functions:

i(t) i
A

() = e e

K(t)f()ddt (2.23)

i i
= e
(t)
K(t)d(t) e

f()dt (2.24)
=()f

() (2.25)
where f

() is the Fourier transform of f(t). Returning to Eq.(2.20) for the absorption energy,

1

P(t)dt= f(t)A

(t)dt= f

()A

()d (2.26)
2
Rearranging the expression once again and evaluating the integral over time yields

P(t)dt=
1
(i)f

()A

()d (2.27)
2

1
= (i)()|f

()|
2
d (2.28)
2

1
=

()|f

()|
2
d (2.29)
2
5.72, Spring 2008 J. Cao

9 Chapter 2. Non-equilibrium Thermodynamics


Hence the absorption power spectrum P() has the form
P() =

()|f

()|
2
(2.30)
Example: For a monochromatic force F(t) =F cos
0
t, the Fourier transform of F(t) is given by

F() =F [(
0
) +(+
0
)]
HenceourexpressionfortheabsorptionpowerspectrumEq.(2.30)tellsusthattheabsorptionrate(i.e. the
time average of P(t)) for such a system is

()|F |
2
lim P(t)dt=
2
0
Furthermore, using Eq.(2.17) and the fact that

() is odd, we nd that

()= sintK(t)dt
0

( )
= sint C

(t) dt
0

= C(t)costdt
0

i
=

e
t
C(t)dt
2

= C()
2
Thissimplerelationshipillustratesthecloseconnectionbetweenfrequency-dependentresponsefunctions(or
susceptibilities) and time correlation functions.
2.3.3 Causality and the Kramers-Kronig Relations
Our nal consideration in this chapter is the relationship between the real and imaginary parts of
the frequency-dependent response function as dened in Eq.(2.19). The equations relating these
twofunctionsareknownastheKramers-Kronigrelations. Intheirmostgeneralform,theygovern
theresponsefunctionasafunctionofthecomplex frequencyz=+i,thoughundermostphysical
circumstances of interest they can be expressed in terms of real-valued frequencies alone.
The relations arise from the causality requirement, which we originally expressed by requiring

K(t) = 0,t < 0. It turns out that this requirement, along with the assumption that K(t)dt
0
converges,impliesthattheresponsefunction(z)isanalyticontheupperhalfofthecomplexplane.
Consider integrating the function
(z)
z
0
around the contour in the complex plane shown in Figure 2.3. The contour has been explicitly
constructed to avoid the pole at z=
0
, so Cauchys integral theorem implies that

(z)
dz=0 (2.31)
z
0
5.72, Spring 2008 J. Cao
10 Chapter 2. Non-equilibrium Thermodynamics
Figure2.3: ContourinthecomplexfrequencyplaneusedtomotivatetheKramers-Kronigrelations
We can also, however, integrate piecewise over each part of the contour; some manipulation with
the residue theorem is required, but the nal result is

(z) (z)
dz=P dzi(
0
) (2.32)
z
0
z
0
where P denotes the Cauchy principal value of the integral. Setting the two results above equal
and solving for the response function,

1 (z)
() = P dz (2.33)
i z

ThedecompositionofEq.(2.33)intorealandimaginarypartsyieldstheKramers-Kronigrelations,

(z)

() =
1
P dz (2.34)
z

(z)

() = P dz (2.35)
z

Aspromised,thispairofequationsprovidesaconciserelationshipbetweentherealandimaginary
parts of any response function ().
5.72, Spring 2008 J. Cao
References
[1] N. Kubo, R.; Saito and N. Hashitsume. Statistical Physics II: Nonequilibrium Statistical Me-
chanics. Springer, 2003.
[2] Robert Zwanzig. Nonequilibrium Statistical Mechanics. New York: Oxford University Press,
2001.
[3] David Chandler.Introduction to Modern Statistical Mechanics. New York: Oxford University
Press, 1987.
[4] L.E. Reichl. A Modern Course in Statistical Physics. New York: Wiley-Interscience, 1998.
11
MIT OpenCourseWare
http://ocw.mit.edu
5.72 Statistical Mechanics
Spring 2012
For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

Você também pode gostar