You are on page 1of 9

PHYSICAL REVIEW B 90, 125140 (2014)

Derivation of nonlinear single-particle equations via many-body Lindblad superoperators:


A density-matrix approach

Roberto Rosati,1 Rita Claudia Iotti,1 Fabrizio Dolcini,1,2 and Fausto Rossi1,*
1
Department of Applied Science and Technology, Politecnico di Torino, C.so Duca degli Abruzzi 24, 10129 Torino, Italy
2
CNR-SPIN, Monte S. Angelo-via Cinthia, I-80126 Napoli, Italy
(Received 8 August 2014; published 23 September 2014)

A recently proposed Markov approach provides Lindblad-type scattering superoperators, which ensure the
physical (positive-definite) character of the many-body density matrix. We apply the mean-field approximation
to such a many-body equation, in the presence of one- and two-body scattering mechanisms, and we derive
a closed equation of motion for the electronic single-particle density matrix, which turns out to be nonlinear
as well as non-Lindblad. We prove that, in spite of its nonlinear and non-Lindblad structure, the mean-field
approximation does preserve the positive-definite character of the single-particle density matrix, an essential
prerequisite of any reliable kinetic treatment of semiconductor quantum devices. This result is in striking contrast
with conventional (non-Lindblad) Markov approaches, where the single-particle mean-field equations can lead
to positivity violations and thus to unphysical results. Furthermore, the proposed single-particle formulation is
extended to the case of quantum systems with spatial open boundaries, providing a formal derivation of a recently
proposed density-matrix treatment based on a Lindblad-like system-reservoir scattering superoperator.

DOI: 10.1103/PhysRevB.90.125140 PACS number(s): 72.10.−d, 73.63.−b, 85.35.−p

I. INTRODUCTION may lead to unphysical results. As originally pointed out by


Spohn and coworkers [24], the positive-definite character of
The microscopic derivation of suitable scattering super-
the density-matrix operator may be violated. In particular, in
operators is one of the most challenging problems in quan-
Ref. [24] the author pointed out that the choice of the adiabatic
tum physics. For purely atomic and/or photonic systems,
decoupling strategy is definitely not unique. Only the case
dissipation and decoherence phenomena may successfully be
discussed by Davies [3] of a “small” subsystem interacting
described via adiabatic-decoupling procedures [1,2] in terms
with a thermal environment could be shown to preserve
of extremely simplified models based on phenomenological
positivity. However, such a result was restricted to finite-
parameters; within such effective treatments, the main goal is
dimensional subsystems (i.e., N -level atoms), and to the
to identify a suitable form of the Liouville superoperator, able
particular projection scheme of the partial trace. Thus, it
to preserve the positive-definite character of the corresponding
cannot be straightforwardly extended to the study of solid-state
density-matrix operator [3]. This is usually accomplished
systems.
by identifying proper Lindblad superoperators [4] expressed
To overcome this serious limitation, an alternative and more
in terms of a few crucial system-environment coupling
general Markov procedure has recently been proposed [25],
parameters. In contrast, in solid-state materials and devices
offering the following advantages: (i) in the discrete-spectrum
the complex many-electron quantum evolution results in a
case it coincides with the Davies model mentioned above,
nontrivial interplay between coherent dynamics and energy-
(ii) in the semiclassical limit it reduces to the well-known
dissipation and decoherence processes [5–10], which has to be
Fermi’s golden rule, and (iii) it holds also in the case
treated via fully microscopic approaches.
of a continuous-spectrum. Most importantly, this approach
Based on the pioneering works by Van Hove [11], Kohn
describes a genuine Lindblad evolution, thereby ensuring the
and Luttinger [12], and Zwanzig [13], several adiabatic or
positivity of the many-electron density matrix and providing
Markov approximation schemes have been developed, which
a reliable and robust treatment of energy-dissipation and
can be grouped in two main categories: approaches based on
decoherence processes in semiconductor quantum devices.
semiclassical (i.e., diagonal) scattering superoperators, also re-
Once the evolution for the many-electron density matrix
ferred to as Pauli master equations [14–16], and fully quantum-
is thus solved, its positive-character—-ensured by Lindblad
mechanical (i.e., nondiagonal) dissipation models [17–21].
evolution—directly transfers via a trace-type projection onto
These approaches have been widely applied to quantum-
the single-particle density matrix, which is needed to obtain
transport and coherent-optics phenomena in semiconductor
various physical observables such as the carrier density, the
materials and devices.
average kinetic energy, and the charge current. However,
When the system-environment coupling becomes strong
despite the conceptual importance of such an alternative
and/or the excitation timescale is extremely short, Marko-
Markov approach, its practical implementation is limited by
vian approaches are known to be unreliable, and memory
the fact that the many-body evolution is in general not exactly
effects have to be taken into account via quantum-kinetic
solvable, so that the single-particle density matrix cannot
approaches [22,23]. However, even in regimes where the
be extracted either. Indeed, so far, all relevant applications
Markov limit is applicable, conventional Markov approaches
of such Markov treatment to semiconductor nanosystems
are limited to the low-density limit [26–29], where one can
explicitly show that the scattering-induced time evolution
*
fausto.rossi@polito.it; staff.polito.it/fausto.rossi is Lindblad-type also for the single-particle density matrix.

1098-0121/2014/90(12)/125140(9) 125140-1 ©2014 American Physical Society


ROSATI, IOTTI, DOLCINI, AND ROSSI PHYSICAL REVIEW B 90, 125140 (2014)

At high carrier concentrations the problem is by far more operator ρ̂ can be written in operatorial form as
complicated. A direct solution of the many-electron problem 
d ρ̂  1  s s†
is in general too demanding, and approximations have to be  = (â ρ̂ b̂ − âs† b̂s ρ̂) + H.c., (2)
introduced. The most straightforward way to obtain a closed dt scat 2 s
equation for the single-particle density matrix is to apply the Ĥs
mean-field scheme to the many-electron equation. In doing where âs = 
,
so, however, the Lindblad structure of the original equation is  +∞
1 Ĥ◦ t  Ĥ◦ t 
lost. The crucial question arises whether the positive-character b̂ = s
e− i Ĥs e i dt  , (3)
of the single-particle equation is violated by the mean-field  −∞

approximation. and H.c. denotes the Hermitian conjugate. The scattering


Primary goal of this paper is to address this problem. By superoperator in Eq. (2) is definitely non-Lindblad, and
applying the conventional mean-field approximation to the therefore does not necessarily preserve the positive-definite
many-electron dynamics obtained via the alternative Markov character of the global density matrix ρ̂ [31].
limit recalled above, we derive a closed equation of motion for In contrast, the alternative Markov procedure proposed
the electronic single-particle density matrix, in the presence in Ref. [25] is based on a time symmetrization between
of one- as well as two-body scattering mechanisms. While microscopic and macroscopic scales and enables one to
in the low-density limit the Lindblad form is preserved, at express the scattering contribution in terms of the following
finite or high carrier concentrations the equation turns out Lindblad superoperator:

to be non-Lindblad and highly nonlinear. Nevertheless, we d ρ̂   1 s† s

= Â ρ̂ Â − {Â Â ,ρ̂} ,
s s†
(4)
are able to prove that the mean-field approximation does dt scat s
2
preserve the positive-definite character of the single-particle
density matrix, an essential prerequisite of any reliable kinetic where
treatment of semiconductor quantum devices. Finally, the   14  +∞
2 2 Ĥ◦ t  Ĥ◦ t  t  2
proposed single-particle formulation is extended to the case of  = lim
s
e− i Ĥs e i e−(  ) dt  , (5)
a quantum system with spatial open boundaries; this provides →0 π 6 −∞
a formal derivation of a recently proposed density-matrix with the energy  playing the role of the level broadening
treatment based on a Lindblad-like system-reservoir scattering corresponding to a finite collision duration and/or to a finite
superoperator [26]. single-particle lifetime [10].
The paper is organized as follows: In Sec. II, after briefly Both the non-Lindblad superoperator in Eq. (2) and the
recalling the main ingredients of the alternative many-body Lindblad one in Eq. (4) may suitably be expressed in terms
Markov approach, we derive a nonlinear equation for the of generalized scattering rates P s . More specifically, denoting
electronic single-particle density matrix in the presence of by {|i} and {i } the eigenstates and the energy levels of the
one- as well as two-body scattering mechanisms. Section III noninteracting Hamiltonian Ĥ◦ , one obtains
presents a detailed investigation, where we show that the pro- 
dρ i1 i2  1  s 
posed single-particle equation preserves the positive-definite
 = Pi1 i2 ,i  i  ρ i1 i2 − Pis∗
 
i ,i i  ρ i  i2 + H.c.,
character of the single-particle density matrix. In Sec. IV our dt scat2 1 2 1 1 1 2 2
s,i1 i2
single-particle treatment is extended to the case of a quantum (6)
system with spatial open boundaries. Finally, in Sec. V we
summarize and draw our conclusions. where for the non-Lindblad model in Eq. (2),
Pis1 i2 ,i  i  = asi1 i  bs∗
i2 i  , (7)
1 2 1 2

whereas for the Lindblad model in Eq. (4) one has


II. FROM A MANY-BODY DESCRIPTION TO A
SINGLE-PARTICLE PICTURE Pis1 i2 ,i  i  = Asi1 i  As∗
i2 i  . (8)
1 2 1 2

Within the spirit of the usual perturbation theory, the global While for both models their diagonal (i.e., semiclassical)
semiconductor Hamiltonian (electrons plus various crystal elements (i1 i1 = i2 i2 ) coincide with the standard Fermi’s-
excitations, e.g., phonons, plasmons, etc.) may be written as golden-rule prescription,

 Pii,i
s
i = |i|Ĥs |i  |2 δ(i − i  ), (9)

Ĥ = Ĥ◦ + Ĥs , (1)
s the Lindblad form in Eq. (8) exhibits a more symmetric
structure, a clear fingerprint of the time symmetrization
previously mentioned.
where the first term Ĥ◦ is the unperturbed contribution that can The study of electro-optical processes in semiconductors
be treated exactly, and the second term describes a number mainly relies on physical quantities that depend on the
of perturbations Ĥs , corresponding to various interaction electronic-subsystem coordinates only. It is thus customary
mechanisms (e.g., carrier-phonon, carrier-carrier, etc.), which to introduce a many-electron density-matrix operator,
are typically treated within some approximation scheme [30]. .
ρ̂ c = tr{ρ̂}p , (10)
In the conventional approaches to the Markov limit, the
second-order (or scattering) contribution to the time evolution where the nonrelevant phononic (p) degrees of freedom have
of the global (e.g., carriers plus phonons) density-matrix been traced out of the global density-matrix operator ρ̂ [32]. By

125140-2
DERIVATION OF NONLINEAR SINGLE-PARTICLE . . . PHYSICAL REVIEW B 90, 125140 (2014)

.
denoting with ρ̂ ◦p the equilibrium density-matrix operator of s = q,± corresponds to the emission (+) or absorption (−) of
the phononic subsystem, and by assuming a state factorization a phonon with wavevector q.
of the form By inserting the carrier-phonon Lindblad operator Eq. (17)
into Eq. (16) and by employing the fermionic anticommutation
ρ̂ = ρ̂ c ⊗ ρ̂ ◦p , (11)
relations, it is easy to show (see Appendix A for details) that the
it is possible to show [25] that the reduced dynamics dictated contribution to the system dynamics due to the generic carrier-
by the Lindblad global evolution in Eq. (4) is still of Lindblad phonon interaction mechanism s involves average values of
type: four fermionic operators of the form
  

d ρ̂ c  1 s† s

hα3 α4 ,α3 α4 = tr ĉα† 3 ĉα4 ĉα ĉα ρ̂ c c . (18)
= Âs
ρ̂ Âs†
− Â Â ,ρ̂ . (12)
dt 
3 4
c c c
scat s
2 c c c
For the carrier-carrier (cc) interaction, the Lindblad opera-
Here the explicit form of the reduced or electronic operators tor has the general two-body form
Âsc can be derived starting from the global Lindblad operators 1  cc †
Âs in Eq. (5) [33]. Âsc = Aαα,α α ĉα† ĉα ĉα ĉα , (19)
2 
Within the above description, although a statistical average  αα,α α
over the phononic degrees of freedom has been performed, the
which describes the transition of the electronic pair from the
electronic subsystem is still treated via a many-body picture.
initial (two-body) state α  α  to the final state αα.
Nevertheless, in the investigation of semiconductor-based
As shown in Appendix A, by inserting Eq. (19) in Eq. (16),
quantum materials and devices, many of the physical quantities
the contribution to the system dynamics due to carrier-carrier
of interest are described via single-particle electronic operators
interaction (s = cc) involves average values of eight fermionic
of the form
 operators of the form
Ĝc = Gα1 α2 ĉα† 1 ĉα2 , (13) † †

α1 α2
kα5 α6 α7 α8 ,α5 α6 α7 α8 = tr ĉα† 5 ĉα† 6 ĉα7 ĉα8 ĉα ĉα ĉα ĉα ρ̂ c c . (20)
5 6 7 8

ĉα†
where and ĉα denote the usual creation and destruction op- As anticipated, the crucial step in order to get a closed
erators over the electronic single-particle states |α. Recalling equation of motion for the single-particle density ma-
that, for any electronic operator one has Gc  = tr{ρ̂ Ĝc } = trix consists of performing the well-known mean-field (or
tr{ρ̂ c Ĝc }c , the average value of the single-particle operator in correlation-expansion) approximation [34–36]; as discussed
Eq. (13) can be written as in Appendix A, employing this approximation scheme and
 omitting renormalization terms [35], for both carrier-phonon
Gc  = ρα1 α2 Gα2 α1 , (14) and carrier-carrier scattering, the resulting single-particle
α1 α2 equation is given by

where dρα1 α2  1   

 = δα1 α − ρα1 α Pαs  α2 ,α α ρα1 α2
ρα1 α2 = tr ĉα† 2 ĉα1 ρ̂ c c (15) dt scat 2   
α α1 α2
1 2

 
is the single-particle density matrix. − δα α1 − ρα α1 Pαs∗ α ,α1 α ρα2 α2 + H.c., (21)
1 2
For the study of the time evolution of single-particle
quantities, such as total carrier density, mean kinetic energy, with generalized carrier-phonon scattering rates
charge current, and so on, it is then crucial to derive a closed s=cp cp cp∗
equation of motion for the above single-particle density matrix. Pα1 α2 ,α α = Aα1 α Aα2 α , (22)
1 2 1 2

Combining its definition in Eq. (15) with the many-electron and generalized carrier-carrier scattering rates
Lindblad dynamics in Eq. (12), and employing the cyclic   
property of the trace, one obtains Pαs=cc   = 2
1 α2 ,α α
δα2 α1 − ρα2 α1
 1 2

dρα1 α2  1  s† †
α 1 α 2 ,α 1 α 2
 = tr Âc ,ĉα2 ĉα1 Âsc ρ̂ c c + H.c. (16)
dt 2
scat α1 α 1 ,α  α  Aα2 α 2 ,α  α  ρα 1 α 2 ,
× Acc cc∗   (23)
s 1 1 2 2

In order to get a closed equation of motion for the single- where


particle density matrix, it is now crucial to specify the form 1  cc 
of our many-electron Lindblad operators Âsc , which, in turn, Acc
αα,α  α  = Aαα,α α − Acc
αα,α  α  − Aαα,α  α  + Aαα,α  α 
cc cc
4
depends on the specific interaction mechanism considered.
(24)
For the case of a generic carrier-phonon (cp) interaction
mechanism, the corresponding (one-body) Lindblad operator denote the totally antisymmetric parts of the two-body coeffi-
is always of the form cients in Eq. (19).
 cp It is worth stressing that, differently from the generalized
Âsc = Aαα ĉα† ĉα . (17) carrier-phonon rates in Eq. (22), the generalized carrier-carrier
αα 
rates in Eq. (23) are themselves a function of the single-particle
Equation (17) describes the phonon-induced carrier transition density matrix; this is a clear fingerprint of the two-body nature
from the initial state α  to the final state α. In this case the label of the carrier-carrier interaction (see below).

125140-3
ROSATI, IOTTI, DOLCINI, AND ROSSI PHYSICAL REVIEW B 90, 125140 (2014)

The single-particle scattering superoperator in Eq. (21) where Î denotes the identity operator of the one-electron
is the result of positive-like (in-scattering) and negative-like Hilbert space. Importantly, due to the quantum-mechanical
(out-scattering) contributions, which are nonlinear functions of Pauli factors (Î − ρ̂), the above scattering superoperator
the single-particle density matrix. Indeed, in the semiclassical Eq. (31) is nonlinear in ρ̂ and non-Lindblad [37]. Only in
limit [10], the low-density limit, i.e., Î − ρ̂ → Î, the nonlinear equation
ρα1 α2 = fα1 δα1 α2 , (25) in Eq. (31) reduces to the Lindblad superoperator

the density-matrix Eq. (21) assumes the expected nonlinear d ρ̂  1
=  ρ̂ † − {† Â,ρ̂}, (32)
Boltzmann-type form dt scat 2

dfα  
= (1 − fα )Pαα
s
 fα  − (1 − fα  )Pα  α fα , (26)
s and the positive-definite character of ρ̂ is thereby ensured.
dt 
scat α At finite or high densities, no straightforward conclusion can
be drawn about the positive-definite character of the generic
with semiclassical carrier-phonon scattering rates time-dependent solution ρ̂(t).
 cp 2
P  = P   = A  
s=cp s=cp
αα αα,α α αα (27) Nevertheless, we show now that the proposed nonlinear
single-particle equation in Eq. (21) does preserve the positive-
and semiclassical carrier-carrier scattering rates definite character of ρ̂. In order to prove that, let us describe
  2
P s=cc

αα = P s=cc  = 2
αα,α α (1 − fα )Acc    fα . αα,α α (28) the single-particle density matrix ρα1 α2 via the corresponding
αα  operator ρ̂ in Eq. (29); at any time t it is possible to define
its instantaneous (i.e., time-dependent) eigenvalues λ and
The above semiclassical limit clearly shows that the nonlin-
eigenvectors |λ according to
earity factors (δα1 α2 − ρα1 α2 ) in Eq. (21) as well as in Eq. (23)
can be regarded as the quantum-mechanical generalization ρ̂|λ = λ |λ, (33)
of the Pauli factors (1 − fα ) of the conventional Boltzmann
theory (see also Sec. III). which implies that
A closer inspection of Eqs. (21) and (23)—together with λ = λ|ρ̂|λ. (34)
their semiclassical counterparts in Eqs. (26) and (28)—
confirms the two-body nature of the carrier-carrier interaction. The eigenvalues λ in Eq. (33) of a single-particle density
Indeed, differently from the carrier-phonon scattering, in this matrix ρ̂ describing a physical state are necessarily positive-
case the density-matrix equation describes the time evolution definite and not greater than one (Pauli exclusion principle). In
of a so-called “main carrier” α interacting with a so-called order to preserve such positive-definite nature, it is imperative
“partner carrier” α. that the scattering-induced time evolution maintains the values
of the eigenvalues within the physical interval [0, 1]; this
can be verified by studying the time derivative of the generic
III. POSITIVITY ANALYSIS OF THE PROPOSED
eigenvalue in Eq. (34), namely:
SINGLE-PARTICLE EQUATION
dλ dλ| d ρ̂ d|λ
The primary goal of this section is to face the most = ρ̂|λ + λ| |λ + λ|ρ̂ . (35)
important issue related to the proposed kinetic treatment: the dt dt dt dt
positivity analysis of the nonlinear density-matrix equation In view of the completeness of the basis set {|λ}, the time
in Eq. (21). Indeed, if the single-particle density matrix derivative in Eq. (35) can also be written as
describes a physical state, its eigenvalues are necessarily dλ  dλ| d ρ̂
positive-definite and not greater than one (Pauli exclusion = |λ λ |ρ̂|λ + λ| |λ
principle); in order to preserve such physical nature, it is dt λ
dt dt
imperative that the scattering-induced time evolution preserves  d|λ
the values of the density-matrix eigenvalues within the interval + λ|ρ̂|λ λ | . (36)
λ
dt
[0, 1].
To this aim, let us start considering the case of carrier- Recalling that
phonon interaction previously discussed, whose nonlinear
equation in Eq. (21) [equipped with the generalized rates in λ|ρ̂|λ  = λ δλλ , (37)
Eq. (22)] may also be easily rewritten in a more compact way the result in Eq. (36) reduces to
via the one-electron operators
 dλ dλ| d ρ̂ d|λ
ρ̂ = |α1 ρα1 α2 α2 | (29) = λ |λ + λ| |λ + λ λ| . (38)
dt dt dt dt
α1 α2
Taking into account that
and
 dλ| d|λ dλ|λ
 = |α1 Acp
α1 α2 α2 |, (30) |λ + λ| = = 0, (39)
α1 α2 dt dt dt
as the first and third term in Eq. (38) cancel out exactly, and one

d ρ̂  1 finally concludes that
= [(Î − ρ̂) ρ̂ † − † (Î − ρ̂) ρ̂] + H.c.,
dt scat 2 dλ d ρ̂ dρλλ
(31) = λ| |λ = . (40)
dt dt dt

125140-4
DERIVATION OF NONLINEAR SINGLE-PARTICLE . . . PHYSICAL REVIEW B 90, 125140 (2014)

This shows that the time variation of the eigenvalues λ (a) (b)
coincides with the time variation of the diagonal elements ρλλ 1.4 1.4

of the operator ρ̂ within the instantaneous eigenbasis {|λ}. 1.2


1.2
In order to evaluate the above time derivative, the crucial
step is to analyze the explicit form of the proposed single- 1.0 1.0
particle scattering superoperator written in the density-matrix
eigenbasis of Eq. (33). Taking into account that the generic 0.8 0.8

density-matrix Eq. (21) is basis-independent, by replacing the


0.6 0.6
original single-particle basis {|α} with the density-matrix
eigenbasis {|λ} and making use of Eq. (37), its diagonal 0.4 0.4
elements turn out to be
dρλλ  0.2 0.2
= (1 − λ )Pλλ
s
 λ − (1 − λ )Pλ λ λ ,
s
(41)
dt λ
0.0 0.0

where -0.2 -0.2

 = Pλλ,λ λ
s s -0.4 -0.4
Pλλ (42)
ω0 t ω0 t
are positive-definite quantities given by the diagonal elements
of the generalized scattering rates [see Eqs. (22) and (23)] FIG. 1. (Color online) Density-matrix eigenvalues as a function
written in our instantaneous density-matrix eigenbasis. By of time for a subset of 25 randomly generated evolutions (see Ap-
inserting this last result into Eq. (40), one finally gets pendix B) corresponding to a simple two-level system in the presence
of carrier-phonon interaction. Comparison between the proposed
dλ 
= (1 − λ )Pλλ single-particle model in Eq. (31) (panel a) and the conventional model
 λ − (1 − λ )Pλ λ λ . (43)
s s
dt λ
in Eq. (44) (panel b) (see text).

This last result is highly nontrivial: it states that, in spite of the reduces to the following non-Lindblad form:
partially coherent nature of the carrier dynamics in Eq. (21), 
d ρ̂  1
the time evolution of the eigenvalues λ is governed by a = (â ρ̂ b̂† − â † b̂ ρ̂) + H.c. (45)
nonlinear Boltzmann-type equation, formally identical to the dt scat 2
semiclassical result in Eq. (26). In analogy to Eq. (7), the generalized scattering rates (within
We are now in the position to state that the physical interval the eigenbasis {|λ}) corresponding to the above non-Lindblad
[0, 1] is the only possible variation range of our eigenvalues superoperator are always of the form
λ . To this end, one can show that, when the latter approach
the extremal values, 0 or 1, their time derivatives do not allow Pλs1 λ2 ,λ λ = aλ1 λ bλ∗2 λ , (46)
1 2 1 2
them to exit the interval. Indeed, a closer inspection of the implying that in this case their diagonal elements,
Boltzmann-like equation in Eq. (43) shows that:

 = Pλλ,λ λ = aλλ bλλ ,
s s
(i) if one of the eigenvalues λ is equal to zero, the corre- Pλλ (47)
sponding time derivative in Eq. (43) is always nonnegative;
(ii) if one of the eigenvalues λ is equal to one, its time are not necessarily positive-definite. This is the reason why,
derivative in Eq. (43) is always nonpositive. starting from a non-Lindblad many-body scattering model,
This leads us to the important conclusion that, for both the system dynamics may exit the physical eigenvalue region,
carrier-phonon and carrier-carrier scattering, the proposed giving rise to positivity violations also in the low-density
nonlinear single-particle Eq. (21) preserves the positive- limit [25,31].
definite character of the single-particle density matrix. To emphasize this point, in Fig. 1 we report the time evolu-
We finally stress that the above positivity analysis is tion of the density-matrix eigenvalues for a subset of simulated
based on the fact that the scattering rates in Eq. (42) are experiments (see Appendix B) in the simple case of a two-level
positive-definite quantities. This property, which applies to the system. As one can see, while for the proposed nonlinear
proposed single-particle equation [obtained starting from the equation in Eq. (31) all the eigenvalue trajectories fall within
Lindblad-type scattering superoperator in Eq. (4)], is generally the physical interval [0, 1] [see Fig. 1(a)], for the nonlinear
not fulfilled by conventional Markov models. In particular, equation in Eq. (44) a significant number of simulated
for the case of carrier-phonon scattering, starting from the eigenvalue trajectories exit the physical interval [Fig. 1(b)].
non-Lindblad scattering superoperator in Eq. (2) and applying
again the mean-field approximation, it is possible to derive a IV. GENERALIZATION TO QUANTUM SYSTEMS
nonlinear single-particle equation of the form WITH SPATIAL OPEN BOUNDARIES

d ρ̂  1 In what follows we extend the proposed single-particle
= [(Î − ρ̂)â ρ̂ b̂† − â † (Î − ρ̂)b̂ ρ̂] + H.c. (44)
dt scat 2 treatment to quantum systems with spatially open boundaries,
namely to the case of a quantum device electrically connected
This nonlinear equation is not intrinsically positive-definite, to one or more external carrier reservoirs. To this end, in
as confirmed by the fact that in the low-density limit the latter analogy to the system factorization between electronic and

125140-5
ROSATI, IOTTI, DOLCINI, AND ROSSI PHYSICAL REVIEW B 90, 125140 (2014)

phononic degrees of freedom in Eq. (11), we shall describe the with generalized scattering rates
global carrier system (device plus reservoirs) as the product
of a device density-matrix operator times a quasiequilibrium Pαs 1 α2 ,β1 β2 = Adr dr∗
α1 β1 Aα2 β2 . (54)
density-matrix operator corresponding to one or more carrier In the semiclassical limit [see Eq. (25)], the above device-
reservoirs: reservoir scattering superoperator reduces to the relaxation-
ρ̂ c = ρ̂ d ⊗ ρ̂ ◦r . (48) time model

dfα    
The quantum-mechanical coupling between device and
 =− s
Pαβ fα − fβ◦ , (55)
external reservoirs (s ≡ dr) may conveniently be described dt scat β
via the following interaction Hamiltonian:
 with device-reservoir scattering rates
∗ †
Ĥs = (γαβ ĉα† ξ̂β + γαβ ξ̂β ĉα ), (49)  2
αβ
s
Pαβ = Pαα,ββ
s
= Adr 
αβ . (56)

where ĉα† (ĉα ) are now creation (destruction) operators acting It is worth stressing that the above semiclassical equation,
† usually referred to as the injection-loss model, has been widely
on the device single-particle states α, while ξ̂β (ξ̂β ) denote
creation (destruction) operators acting on the reservoir single- employed in the semiclassical modeling of optoelectronic
particle states β. Here, the first contribution describes carrier semiconductor devices [38].
injection (β → α) via the destruction of a carrier in state β In order to gain more insight on the structure of the density-
and the creation of a carrier in state α, while the second matrix Eq. (53), the latter may conveniently be rewritten
one describes carrier loss (α → β) via the inverse process. in a compact operatorial form; more specifically, recalling
Moreover, the physical properties of the device-reservoir the definition of the single-particle density-matrix operator
interaction Hamiltonian in Eq. (49) are dictated by the explicit in Eq. (29) and introducing the device-reservoir coupling
form of the coupling matrix elements γαβ ; the latter, in general, operators
are given by a properly weighted spatial overlap between 
Âβ = |αAdr
αβ β| (57)
device and reservoir single-particle wavefunctions.
α
Following the general prescription in Eq. (5), the Lindblad
operator corresponding to the device-reservoir interaction as well as the reservoir density-matrix operator
Hamiltonian Eq. (49) depends on the carrier coordinates only, 
and is always of the form ρ̂ ◦ = |βfβ◦ β|, (58)
  β
† dr∗ †
Âsc = αβ ĉα ξ̂β + Aαβ ξ̂β ĉα .
Adr (50)
one gets
αβ

d ρ̂   1

◦ † †
The evaluation of the single-particle dynamics induced = Â ρ̂ Â − { Â Â ,ρ̂} . (59)
by the above device-reservoir coupling may be performed dt scat β
β β
2 β β
following the very same steps of the corresponding carrier-
phonon and carrier-carrier treatments previously considered Equation (59) should be compared to the Lindblad super-
and described in Appendix A. In particular, it is easy to operator in Eq. (32) describing energy exchange with the
realize that the single-particle contribution in Eq. (16) due phononic excitations. On the one hand, the device-reservoir
to device-reservoir coupling involves average values of two superoperator Eq. (59) is inhomogeneous, due to the presence
device plus two reservoir creation and destruction operators. of the density-matrix operator ρ̂ ◦ of the external reservoirs.
More specifically, in view of the device-reservoir factorization This implies that the trace of the device density matrix ρ̂ is not
in Eq. (48) as well as of the typical quasiequilibrium nature of conserved, as expected in a system that can exchange particles
the reservoirs, one obtains with the reservoirs. On the other hand, the coupling term in

Eq. (59) is linear in ρ̂ and has a Lindblad-like form, which
tr ĉα† 2 ĉα1 ξ̂β2 ξ̂β1 ρ̂ c c = ρα1 α2 ρβ◦1 β2 , (51) ensures the positive-definite character of ρ̂.
where ρα1 α2 is the single-particle density matrix of the device The analysis presented so far can be regarded as a formal
and derivation of the Lindblad-like device-reservoir scattering su-
peroperator recently proposed in Ref. [26], where the reservoir
ρβ◦1 β2 = fβ◦1 δβ1 β2 (52) states are plane waves (|β = |k) and the device single-
is the (diagonal) single-particle density matrix of the particle states are the scattering states of the confinement
quasiequilibrium carrier reservoirs. potential profile (|α = |αk ).
Employing the device-reservoir factorization result in
Eq. (51), a straightforward calculation shows that the con- V. SUMMARY AND CONCLUSIONS
tribution to the system evolution due to the device-reservoir
Exploiting a recent reformulation of the Markov limit,
coupling Hamiltonian Eq. (49) is
  which enables one to provide genuine Lindblad-type scattering
dρα1 α2  1 ◦
 superoperators for the many-body density matrix, we have
= Pα1 α2 ,ββ fβ −
s
Pα1 α ,ββ ρα α2
s
dt scat 2 β α
applied the mean-field approximation to the many-electron
dynamics, and we have derived a closed equation of motion
+ H.c., (53) for the electronic single-particle density matrix, in the presence

125140-6
DERIVATION OF NONLINEAR SINGLE-PARTICLE . . . PHYSICAL REVIEW B 90, 125140 (2014)

of carrier-phonon as well as carrier-carrier scattering mecha- The remaining step in order to get a closed equation of
nisms. While in the low-density limit the equation exhibits motion for the single-particle density matrix consists of per-
a Lindblad form—like for the many-body density matrix—at forming the well-known mean-field (or correlation-expansion)
finite carrier concentrations the resulting time evolution for approximation [34–36]; in this particular case this allows one
the single-particle density matrix turns out to be nonlinear and to express the two-body correlation function Eq. (18)—given
non-Lindblad. by the average value of four fermionic operators—in terms of
We have proven [see Eq. (43)] that, despite the lack products of two single-particle density-matrix elements; more
of a Lindblad form, the mean-field approximation does specifically, omitting renormalization terms [35], one gets
preserve the positive-definite character of the single-particle  
hα3 α4 ,α3 α4 = δα4 α3 − ρα4 α3 ρα4 α3 . (A3)
density matrix, an essential prerequisite of any reliable and
robust kinetic treatment of semiconductor quantum devices. Employing this approximation scheme, the carrier-phonon
This result is in striking contrast with the case of mean- scattering contribution in Eq. (A2) reduces to the single-
field approximation applied to conventional (non-Lindblad) particle density-matrix Eq. (21) equipped with the generalized
Markov approaches [see Eq. (2)], where the corresponding carrier-phonon rates in Eq. (22).
single-particle equations may lead to positivity violations and Let us finally come to the case of carrier-carrier interaction.
thus to unphysical results. In view of the usual anticommutation properties, the original
The proposed single-particle formulation has then been carrier-carrier Lindblad operator in Eq. (19) can also be written
extended to the case of quantum systems with spatial open in terms of the fully antisymmetric coefficients in Eq. (24) as
boundaries; such microscopic treatment can be regarded as
1  cc †
a formal derivation of a recently proposed density-matrix Âsc = Aαα,α α ĉα† ĉα ĉα ĉα . (A4)
treatment [26] based on a Lindblad-like system-reservoir 2   αα,α α
scattering superoperator.
By inserting this alternative form of the Lindblad operator
into Eq. (16) and employing once again the usual fermionic
ACKNOWLEDGMENTS anticommutation relations, it is easy to show that
We are extremely grateful to David Taj for stimulating and s† †  cc∗ †
Âc ,ĉα2 ĉα1 = Aα α2 ,αα ĉα† ĉα ĉα1 ĉα
fruitful discussions. We gratefully acknowledge funding by
ααα 
the Graphene@PoliTo laboratory of the Politecnico di Torino,  †

operating within the European FET-ICT Graphene Flagship − Acc∗
α  α  ,α1 α ĉα2 ĉα ĉα  ĉα  ,
project (www.graphene-flagship.eu). Computational resources αα  α 
were provided by HPC@PoliTo, a project of Academic (A5)
Computing of the Politecnico di Torino (www.hpc.polito.it).
F.D. also acknowledges financial support from FIRB 2012 and therefore

Project HybridNanoDev (Grant No.RBFR1236VV). dρα1 α2 
dt scat

APPENDIX A: DERIVATION OF THE NONLINEAR
1 
SINGLE-PARTICLE SCATTERING SUPEROPERATOR =⎝ Acc∗
α  α2 ,αα Aα  α  α  α  kααα1 α  ,α1 α2 α4 α3
cc    
4 1 2 3 4
In this Appendix we recall the main steps involved in ααα  ,α1 α2 α3 α4
the derivation of the nonlinear single-particle equation in ⎞
1 
   ⎠
Eq. (21). To this aim, let us start by considering the case − Acc∗
α  α  ,α1 α Aα  α  α  α  kα2 αα α  ,α1 α2 α4 α3
cc 
of carrier-phonon interaction. By inserting into Eq. (16) the 4 1 2 3 4
αα  α  ,α1 α2 α3 α4
carrier-phonon Lindblad operator in Eq. (17) and employing
the usual fermionic anticommutation relations, it is easy to + H.c., (A6)
show that where k is the four-particle correlation function introduced in
s† †   cp∗ †  Eq. (20).
Aα2 α ĉα ĉα1 − Aα α1 ĉα† 2 ĉα , (A1)
cp∗
Âc ,ĉα2 ĉα1 = Similar to the case of carrier-phonon coupling, in order to
α
get a closed equation of motion for the single-particle density
and therefore matrix, one is forced to adopt the mean-field approximation
⎛ scheme previously introduced. In this case, the latter amounts

dρα1 α2  1  cp∗ cp to writing the average values of eight fermionic operators
 =⎝ A  A   hα α1 ,α1 α2
dt scat 2    α2 α α1 α2 in Eq. (20) as products of four single-particle density-matrix
α α1 α2
elements. More specifically, neglecting again renormalization

contributions, one gets
1  cp∗ cp   
− A  A   hα α ,α α ⎠ + H.c., kα5 α6 α7 α8 ,α5 α6 α7 α8 = δα8 α5 − ρα8 α5 δα7 α6 − ρα7 α6
2    α α1 α1 α2 2 1 2
α α1 α2  
× ρα7 α6 ρα8 α5 − ρα8 α6 ρα7 α5
(A2)   
− δα7 α5 − ρα7 α5 δα8 α6 − ρα8 α6
where h is the two-particle correlation function introduced in  
Eq. (18). × ρα7 α6 ρα8 α5 − ρα8 α6 ρα7 α5 . (A7)

125140-7
ROSATI, IOTTI, DOLCINI, AND ROSSI PHYSICAL REVIEW B 90, 125140 (2014)

Inserting the above mean-field factorization into Eq. (A6), generalized carrier-phonon scattering rates of the form
after a straightforward calculation one gets again the nonlinear
Pα1 α2 ,α α = Aα1 α1 Bα∗2 α .
s=cp
(B3)
single-particle Eq. (21), equipped with the generalized carrier- 1 2 2

carrier rates in Eq. (23). Our simulated experiments are based on a numerical solu-
tion of the density-matrix Eq. (B1); more specifically, the free
single-particle rotation in Eq. (B2) has been treated exactly,
APPENDIX B: SIMULATION STRATEGY
while the scattering-induced dynamics has been evaluated via
Aim of this Appendix is to describe the simulated experi- a standard time-step integration scheme [10].
ments reported in Fig. 1. We have performed a random generation of several possible
We have considered a two-level system in the presence time evolutions, via a random selection of the initial single-
of carrier-phonon interaction, whose single-particle density particle density matrix as well as of the single-particle energy
matrix will be governed by a nonlinear kinetic equation of the levels in Eq. (B2) and of the carrier-phonon operators in
form Eq. (B3). More specifically, we start each simulated experi-
  ment from a randomly selected single-particle density matrix
dρα1 α2 dρα1 α2  dρα1 α2  ρα1 α2 (t = 0) = Pα1 δα1 α2 , where Pα are randomly generated
= + . (B1)
dt dt Ĥ◦ dt scat (with uniform distribution) between 0 and 1. Similarly, the
single-particle eigenfrequencies ωα in Eq. (B2) are randomly
Here the first term, generated (with uniform distribution) between 0 and ω0 (ω0
 denoting the typical single-particle energy scale in units of
dρα1 α2    ), while in Eq. (B3), Aα1 α2 and Bα1 α2 are random complex
 = −i ωα1 − ωα2 ρα1 α2 , (B2)
dt Ĥ◦ numbers, whose modulus and phase are uniformly generated

between 0 and ω0 and 0 and 2π , respectively.
describes the single-particle rotation induced by the noninter- In particular, while for conventional Markov models the
acting Hamiltonian Ĥ◦ in Eq. (1) (ωα denoting single-particle two carrier-phonon operators A and B are generated inde-
energy levels), while the explicit form of the last term is given pendently [see Eqs. (7) and (46)], in the proposed scattering
by the nonlinear scattering superoperator in Eq. (21) with superoperator one has Aα1 α2 = Bα1 α2 , as shown in Eq. (22).

[1] M. Scully and S. Zubairy, Quantum Optics (Cambridge Univer- [17] M. Lindberg and S. W. Koch, Phys. Rev. B 38, 3342 (1988).
sity Press, Cambridge, 1997). [18] T. Kuhn and F. Rossi, Phys. Rev. Lett. 69, 977 (1992).
[2] H. Breuer and F. Petruccione, The Theory of Open Quantum [19] U. Hohenester and W. Pötz, Phys. Rev. B 56, 13177 (1997).
Systems (Oxford University Press, Oxford, 2007). [20] H. B. Sun and G. J. Milburn, Phys. Rev. B 59, 10748 (1999).
[3] E. Davies, Quantum Theory of Open Systems (Academic Press, [21] C. Flindt, T. Novotný, and A.-P. Jauho, Phys. Rev. B 70, 205334
New York, 1976). (2004).
[4] G. Lindblad, Commun. Math. Phys. 48, 119 (1976). [22] D. B. Tran Thoai and H. Haug, Phys. Rev. B 47, 3574 (1993).
[5] M. Bonitz, Quantum Kinetic Theory (Teubner-Texte zur Physik, [23] J. Schilp, T. Kuhn, and G. Mahler, Phys. Rev. B 50, 5435
Teubner, 1998). (1994).
[6] H. Haug and S. Koch, Quantum Theory of the Optical and [24] H. Spohn, Rev. Mod. Phys. 52, 569 (1980).
Electronic Properties of Semiconductors (World Scientific, [25] D. Taj, R. C. Iotti, and F. Rossi, Eur. Phys. J. B 72, 305 (2009).
Singapore, 2004). [26] F. Dolcini, R. C. Iotti, and F. Rossi, Phys. Rev. B 88, 115421
[7] S. Datta, Quantum Transport: Atom to Transistor (Cambridge (2013).
University Press, Cambridge, 2005). [27] R. Rosati and F. Rossi, Appl. Phys. Lett. 103, 113105
[8] H. Haug and A. Jauho, Quantum Kinetics in Transport and (2013).
Optics of Semiconductors (Springer, Berlin, 2007). [28] R. Rosati and F. Rossi, Europhys. Lett. 105, 17010 (2014).
[9] C. Jacoboni, Theory of Electron Transport in Semiconductors: [29] R. Rosati and F. Rossi, Phys. Rev. B 89, 205415 (2014).
A Pathway from Elementary Physics to Nonequilibrium Green [30] In what follows we shall limit ourselves to the case of time-
Functions (Springer, Berlin, 2010). independent interaction Hamiltonians; the generalization to
[10] F. Rossi, Theory of Semiconductor Quantum Devices: Micro- time-dependent coupling mechanisms is straightforward.
scopic Modeling and Simulation Strategies (Springer, Berlin, [31] R. C. Iotti, E. Ciancio, and F. Rossi, Phys. Rev. B 72, 125347
2011). (2005).
[11] L. Van Hove, Rev. Mod. Phys. 29, 200 (1957). [32] We stress that the proposed treatment of carrier-phonon interac-
[12] W. Kohn and J. M. Luttinger, Phys. Rev. 108, 590 (1957). tion applies to other bosonic degrees of freedom as well (e.g.,
[13] R. Zwanzig, Phys. Rev. 124, 983 (1961). photons, plasmons, etc.).
[14] M. V. Fischetti, Phys. Rev. B 59, 4901 (1999). [33] In spite of their very same formal structure, Eqs. (4) and
[15] R. Gebauer and R. Car, Phys. Rev. Lett. 93, 160404 (2004). (12) describe the system dynamics at different levels; this
[16] I. Knezevic, Phys. Rev. B 77, 125301 (2008). is confirmed by the fact that, while the global operators in

125140-8
DERIVATION OF NONLINEAR SINGLE-PARTICLE . . . PHYSICAL REVIEW B 90, 125140 (2014)

Eq. (5) are always Hermitian, the electronic operators Âsc are [36] V. M. Axt and T. Kuhn, Rep. Prog. Phys. 67, 433 (2004).
usually non-Hermitian, a clear fingerprint of dissipation-versus- [37] Indeed, at any time t, Eq. (31) can always be locally linearized
decoherence processes induced by the phononic subsystem on [10], treating the two Pauli factors (Î − ρ̂) as input parameters;
the carrier one. however, the resulting (time-dependent) linear superoperator is
[34] V. M. Axt and S. Mukamel, Rev. Mod. Phys. 70, 145 (1998). definitely non-Lindblad.
[35] F. Rossi and T. Kuhn, Rev. Mod. Phys. 74, 895 (2002). [38] R. C. Iotti and F. Rossi, Rep. Prog. Phys. 68, 2533 (2005).

125140-9

You might also like