<<

Chapter 2

Outline of density theory

2.1 The concept of a density matrix

Aphysical exp eriment on a usually consists in observing in a macroscopic

1

ensemble a quantity that dep ends on the particle's degrees of freedom (like the

measure of a sp ectrum, that derives for example from the absorption of

due to vibrational transitions). When the interactions of each molecule with any

2

other physical system are comparatively small , it can b e describ ed by a state vector

j (t)i whose time evolution follows the time{dep endent Schrodinger equation

i

_

^

j (t)i = iH j (t)i; (2.1)

i i

^

where H is the system Hamiltonian, dep ending only on the degrees of freedom of

the molecule itself.

Even if negligible in rst approximation, the interactions with the environment

usually bring the particles to a distribution of states where many of them b ehave

in a similar way b ecause of the p erio dic or uniform prop erties of the reservoir (that

can just consist of the other of the gas). For numerical purp oses, this can

1

There are interesting exceptions to this, the \single molecule sp ectroscopies" [44,45].

2

Of course with the exception of the electromagnetic or any other eld probing the system!

10 Outline of density matrix theory

b e describ ed as a discrete distribution. The interaction b eing small compared to the

internal dynamics of the system, the macroscopic e ect will just be the incoherent

sum of the computed for every single group of similar particles times its

3

weight :

1

X

^ ^

< A>(t)=  h (t)jAj (t)i (2.2)

i i i

i=0

If we skip the time dep endence, we can think of (2.2) as a thermal distribution, in

this case the  would b e the Boltzmann factors:

i

E

i

k T

b

e

 := (2.3)

i

Q

where E is the energy of state i, and k and T are the Boltzmann constant and the

i b

temp erature of the reservoir, resp ectively . Q is the partition function de ned as

1

X

E

i

k T

b

Q := e :

i=0

With this formalism, the weights  must b e left time indep endent b ecause we do not

i

have an equation for the distribution but only a formula to derive the exp ectation

values (2.2).

What we need is a mechanical equivalent of the distribu-

tion or micro- of classical statistical . Because of the

sup erp osition principle of , a simple sum generates interferences,

that should not exist b etween the elements of a statistical ensemble (using again the

time indep endent formalism for states in equilibrium):

1 1

X X



^ ^ ^

Aj i: a h j j i = a j i ) < A> = h jAj i = a

i i j stat i i stat stat

j

i=0 i;j =0

To derive the quantum mechanical equivalent of a Liouville distribution, we may

^ ^

intro duce an identity op erator I on the left of A in (2.2)

1 1 1

X X X

^ ^ ^

 h j ih jAj i j ih jAj i =  h j < A> =

i i j j i j j i i i

i;j =0 j =0 i=0

3

The numb er of states is set to in nity, for a nite numb er of states N , it is enough to set the

co ecients with index larger than N to zero.

2.1 The concept of a density matrix 11

and rearrange the order of the terms

1 1

X X



^ ^

< A>= h jA  j ih j j i :

j i i i j

j=0 i=0

^

Accordingly, exp ectation values of quantum mechanical op erators A for ensembles

can be derived from the op eration

 

^ ^

hAi =tr A^ ; (2.4)

where the equilibrium statistical op erator ^ is de ned as

1

X

^ =  j ih j : (2.5)

i i i

i=0

It is p ossible to generalize the de nition (2.5) by intro ducing a and time de-

4

p endent co ecients :

1

X

^(t):=  (t)j ih j : (2.6)

ij i j

i;j =0

Tohave the prop erties equivalent to a micro canonical ensemble, the density op erator

5

must b e Hermitean and p ositive semide nite b ecause when brought back to diago-

nal form the co ecients  (t)=   (t) have the meaning of probabilities, so they

ij ij i

must be real and p ositive. Once the density op erator is generalized, the problem

arises whether (2.6) can always b e rewritten as (2.5); b eing Hermitean, it can always

b e diagonalized; but if there are states with the same p opulation, the transformation

is not unique anymore. Any orthogonal set of state vectors spanning the subspace

of the degenerate eigenvectors can b e used in (2.5); but all of them will give rise to

the same exp ectation values as can b e veri ed by rewriting the equations for ^ and

using di erent sets for the functions b elonging to the degenerate eigenvalues. So the

orthogonalization pro cedure may b e problematic, but the resulting density op erator

will have anyway the correct physical prop erties.

If the coupling with the bath is negligible or its co ordinates are included in (2.6),

the Schrodinger dynamics can b e mapp ed directly onto the density matrix space to

4

The basis set b eing complete a time dep endence on the functions is not necessary.

5

Positive semide nite op erators have eigenvalues real p ositiveornull.

12 Outline of density matrix theory

get the nondissipative Liouville{von Neumann equation [37, 46]

h i

_

^

^(t)= L^(t) L ^(t)i H; ^(t) ; (2.7)

H

where L is the Hamiltonian Liouvillian sup erop erator. The dynamics are unitary

H

and the eigenvalues consequently do not change in time, keeping the distribution

constant. The equations for isolated systems follow always the Hamiltonian dy-

namics, b eing by de nition indep endent of external forces. Eqn.(2.7) is completely

equivalent to a set of time dep endentSchrodinger equations plus the averaging (2.2).

The practical solution of (2.7), however, requires the propagation of matrices rather

than state vectors. This disadvantage of density matrices is outweighed only in rare

cases by the advantage that b oth the propagation of mixed and pure initial states

can be done with the same e ort.

When the strength of the interactions with the external mo des cannot be ne-

glected (like in condensed phases) each molecule will b ehave according to the slightly

di erent environment it is in, thus mo difying the distribution (2.6). The dynamics

is even more p erturb ed when we lo ok at the prop erties of small parts of molecules

like the chromophores of a macromolecule or the absorption prop erties of the ac-

6

tive groups of chlorophyll or rho dopsin, b oth emb edded in large biological systems.

Here with the time evolution app ears the phenomenon of IVR (Intramolecular Vi-

brational Redistribution). This could be mo deled using a distribution of molecular

Hamiltonians (source of the inhomogeneous broadening in sp ectroscopy) and/or

uctuating forces (homogeneous broadening). All these phenomena can again be

more elegantly, and often more eciently, be handled in the framework of density

matrix theory. Here, the bath can b e left out of the dynamics, but its e ects on the

system can b e repro duced directly in the equations of of the system via the

(non-Markovian) dissipative Liouville-von Neumann equation [37]

Z

h i

t

_

^

^(t)=L^(t)= i H ; ^(t) + d L (t;  )f^( )g: (2.8)

s D

0

The relevant observables are then directly computed with ^(t) and relation (2.4).

^

In (2.8), the Hamiltonian is indicated with H to stress that it refers only to the

s

co ordinates of the smaller system. One usually talks ab out the \reduced dynamics"

6

Molecules involved in the photosynthesis and in the visual pro cess resp ectively.

2.2 Equations of motion 13

^

of the \reduced density op erator". H is an e ective Hamiltonian b ecause even if

s

it dep ends only on the system mo des, it includes the static (averaged) distortion of

the system dynamics that is due to the environmental degrees of freedom, like the

additional electrostatic eld of a crystal cavity.

The term

Z

t

d L (t;  )f^( )g

D

0

enters in (2.8) as a dissipative corrective to the Hamiltonian evolution to describ e the

dynamical coupling of the system to an unobserved environment and dep ends on the

uctuations of the bath variables. It accounts for energy and phase relaxation,i.e.,

the mo di cation of distribution (2.5), or equivalently, of the eigenvalues of (2.6).

The derivation of the dissipative term may lead to a non-Markovian evolution (as in

Eqn.(2.8)), then the time evolution of the density op erator dep ends up on its past,

the system develops a \memory".

Under the Markov approximation (i.e., when memory e ects are neglected), the

equations of motion turn into [37, 46, 4]:

h i

_

^

^(t)=L^(t)=i H ; ^(t) +L (^(t)) : (2.9)

s D

Here, L (^(t)) is a function of ^ at time t only and is usually linear. The expression

D

\dissipative Liouville-von Neumann equation" normally refers to this equation and

not to (2.8).

The prop er choice of this dissipative part is still a matter of dispute, so we

dedicate to this issue a separate section.

2.2 Equations of motion

The Liouville{von Neumann equation with dissipative terms do es not have a unique

7

form , and even its general prop erties are not yet clearly understo o d. Already in

7

At least there are the non{Markovian (2.8) and the Markovian (2.9) alternatives.

14 Outline of density matrix theory

the derivation of the equations of motion, two main approaches are p ossible: One

is to start from the complete system and to eliminate through pro jection op erator

techniques [47] the environmental degrees of freedom, leaving them only as op era-

tors of the Hilb ert space describing the \essential" mo des. The other one consists in

taking a phenomenological viewp oint, that is devising prop er dissipation op erators

for the Liouville{von Neumann equation, mainly using empirical knowledge on the

system. The two techniques should be equivalent in principle, given that the un-

derlying physics is the same, but the e ects of the environment are often not only

8 9

theoretically complex to disentangle , but also p o orly known . Moreover, coarse

approximations have to b e made on the system and bath mo dels just to be able to

derive the equations of motion, thus making the two approaches very di erent. Both

solutions have their advantages and shortcomings, and b oth rely on the separability

of the system from the bath. Unfortunately, this assumption contradicts the quan-

tum mechanical character of the dynamics b ecause, when systems get in contact

with one another, they are entangled and \cannot be divided" (as tell the famous

Einstein{Po dolski{Rosen correlations). Indeed, it was recently shown by Lindblad

[48] that a reduced dynamics cannot have all the mathematical prop erties wewould

10

exp ect for its physical interpretability .

Nevertheless, the complexity of many reactions makes the use of a reduced dy-

namics interesting, if not unavoidable, b ecause the approximations that have to be

intro duced to describ e the complete dynamics of the system plus reservoir would b e

to o p o or. So, we follow the idea of eliminating the bath degrees of freedom from

the dynamics, while keeping only the co ordinates necessary to compute the phys-

ical observables relevant to the phenomenon of interest. First, the environmental

degrees of freedom are pro jected out of the dynamics, then an inverse mapping is

constructed to the complete space in order to determine the equations of motion (in

simple words: \What would do the bath if it would have b een propagated together

8

For instance, where the asymptotic evolution should lead, and which are the mathematical

restrictions that the dynamics should satisfy.

9

The evaluation of p otential energy surfaces for the strongly coupled co ordinates is already a

very complicated numerical task.

10

This dep ends on the basic algebraic structure of quantum mechanics, the so called von Neu-

mann (noncommutative with involution) algebras, see in [48] and references for a more detailed discussion of this issue.

2.2 Equations of motion 15

with the system"). This is the source of the non-Markovian terms b ecause the

present \answer of the environment" dep ends on how the smaller system b ehaved in

the past. Therefore, the elimination of the bath from the dynamics creates a link to

the past, a memory, that otherwise wouldn't b e there, b ecause quantum mechanics

11

is Markovian . This is why in order to have a reduced dynamics lo cal in time, an

additional approximation has to be made, namely the already mentioned Markov

approximation. It is valid when the correlation time of the reservoir is much smaller

than the characteristic time scale of the system (that is the bath needs much less time

to go back to equilibrium than the system to change signi cantly); in other cases,

e.g., when the bath correlation times are long (slowly moving solvent molecules, low

frequency phonons), it breaks down. This assumption dep ends also on the weakness

of the coupling, b eing exact in the zero coupling limit. An example of the failure

of the Markov approximation is when the system dynamics is very fast b ecause

ultrashort lasers are used in the system Hamiltonian [49].

Many of the mo dels based on pro jection op erators have to make the assumption

that the bath and the system are initially uncorrelated; this may turn out to be

a ma jor source of errors in addition to the neglected non-Markovian character of

the reduced dynamics. The so called build up of memory e ects or \system slips"

[50, 51], are thus generated. In this case the reduced dynamics resembles the true

dynamics (for the solvable cases) only after an initial stage, and if the reduced

op erator phases are changed (slipp ed) in the derivation of the equations. Physically

this corresp onds to set a correction for the initial correlations in the global system.

An extensive study aimed at nding a consistent solution for these problems is b eing

done in the group of Tannor [52].

The size and complexityofmany condensed phase problems makes probably the

use of a Markovian dynamics the of the passable approximations, esp ecially

in the case of the gas surface problems we are interested in. The Red eld theory

[53, 54] and the mathematical, phenomenological approach of Lindblad and Gorini{

11

This happ ens also in the case of Time Dep endent Self Consistent Field equations [20], for cases

where some of the coupled equations are analytically solvable (like harmonic oscillators sub ject to

a time dep endent force) and can b e included in the other equations as non-Markovian nonlinear

terms. In [20], a nice example is rep orted.

16 Outline of density matrix theory

Kossakowski{Sudarshan [55, 56, 46] are the most prominent Markovian approaches

to op en quantum systems [57].

The former b elongs to the class of metho ds that combine pro jection op erator

techniques [47] with p erturbation theory to treat the system bath coupling micro-

scopically. Red eld theory can be shown to converge to the exact dynamics in the

12

weak coupling limit [58]. It is also known to generate negative eigenvalues (proba-

bilities) in case of violation of the p erturbativehyp othesis or for a time scale shorter

than the bath correlation time [59]. This is the \coarse graining" assumption,i.e.,

the dynamics is meaningful only if averaged over timesteps larger than the bath

correlation time.

The Red eld equations of motion, derived for the matrix elements of ^(t)inthe

basis of the eigenfunctions of the e ective Hamiltonian, are

X

_ (t)=i!  (t)+ R  (t) (2.10)

ij ij ij ij;k l kl

kl

where ! = E E is the transition frequency between states i and j , and R is

ij i j ij;k l

the Red eld tensor, a matrix with the square of the of the density matrix.

The Lindblad approach b elongs to the branch of the completely p ositive dynami-

13

cal semigroups . Here, the so called Lindblad op erators are used in the equations of

motion to treat the relaxation in the framework of the system only. One asks which

are the p ossible forms for L in (2.9) that give to the dynamics of the physical sys-

D

tem the correct mathematical and physical prop erties; it is, consequently, a semiphe-

nomenological mathematical approach. Information from microscopic knowledge on

the system bath coupling can be included also in this mo del [43, 60]. This second

approach allows for the probabilistic interpretation of the diagonal elements of the

density matrix at any instantof time, while the Red eld theory do es not.

One of the requirements made in the Lindblad approach is the restrictive com-

plete p ositivity [46, 55]; this can intensify the coarse character of the dynamics as

12

Through the use of numerical simulations and for an harmonic oscillator coupled with a bath

of harmonic oscillators, that is the only system for which a direct solution is p ossible.

13

\Semigroups", b ecause the inverse of the dynamical map do es not exist [46], re ecting an

irreversible evolution.

2.2 Equations of motion 17

illustrated in numerous works [61, 62, 63, 64], b ecause of the weakness of the re-

duced dynamics approach in quantum systems [48]. Nevertheless, this approach is

general and allows to intro duce the known prop erties of the system under study

phenomenologically, even when a prop er microscopical mo del approach is imprac-

tical. The mathematical form of the Lindblad generator was derived by Gorini,

Kossakowski, and Sudarshan [56] for nite dimensional Hilb ert spaces and by Lind-

blad [55] for in nite Hilb ert spaces:

 

h i

X

1

y y

^ ^ ^ ^

L ^ = C ^C : (2.11) C C ; ^

D i i

i i

2 +

i=0

^

The C are the Lindblad op erators. They b elong to the algebra of the b ounded linear

i

P

y

^ ^

op erators acting on Hilb ert space of the system with the condition that C C

i

i

i

must be b ounded [55].

There are other approaches derived from the pro jection op erator technique, the

Caldeira{Leggett mo del [65, 66] b eing one example of these. As far as shortcomings

are concerned it shares the nonp ositive character with the Red eld theory. As a

matter of fact, the original derivation of the Caldeira{Leggett mo del is made in the

Feynmann-Vernon formalism from which one gets a path integral formulation for the

system density matrix dynamics with so called in uence functionals that represent

the e ect of the environment [67]. This last formalism is often used to mo del small

systems in condensed phases, but b ecause of its complexity it cannot be directly

applied to large systems.

An interesting discussion on Markovian dynamics is given in the series of pap ers

of Tannor and coworkers [63, 64].

To deal with stronger couplings between system and bath, time dep endent op-

erator techniques can be used. Here the p erturbation approach is not needed; they

generate a Markovian, but nonlinear dynamics. These metho ds are more accurate,

but computationally intense and were therefore used only for small and simple sys-

tems [68]. Another way to deal with strongly interacting baths is to use higher

order p erturbation theory (the Red eld theory b eing only second order in the bath

coupling); master equations of this kind are derived and discussed in [69].

18 Outline of density matrix theory

We have used Lindblad forms for almost all our dissipation mo dels, b ecause of

the complexityofthe system studied (with electron hole pairs, phonons and multi-

ple electronic surfaces) that renders almost imp ossible to derive sensible dissipation

op erators microscopically. Moreover, we often need a strict p ositive density ma-

trix, b ecause we are interested in small probabilities, and the only alternative to

a Lindblad form would be to abandon the Markovian or the linear approaches in

the hop e that the less approximate dynamics would correct the arising of negative

probabilities [52]. This approach is presently technically imp ossible for our large

systems. Moreover our dissipation is often connected with the motion of electrons,

whose sp eed compared to the nuclei renders the Markov approach very reasonable.