You are on page 1of 12

This article was downloaded by: [York University Libraries]

On: 13 August 2014, At: 13:19


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Journal of Modern Optics


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/tmop20

Comparison and unification of non-Hermitian and


Lindblad approaches with applications to open
quantum optical systems
a ab
Konstantin G. Zloshchastiev & Alessandro Sergi
a
School of Chemistry and Physics, University of KwaZulu-Natal, Pietermaritzburg, South
Africa.
b
National Institute for Theoretical Physics (NITheP), KwaZulu-Natal Node,
Pietermaritzburg, South Africa.
Published online: 20 Jun 2014.

Click for updates

To cite this article: Konstantin G. Zloshchastiev & Alessandro Sergi (2014) Comparison and unification of non-Hermitian and
Lindblad approaches with applications to open quantum optical systems, Journal of Modern Optics, 61:16, 1298-1308, DOI:
10.1080/09500340.2014.930528

To link to this article: http://dx.doi.org/10.1080/09500340.2014.930528

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Journal of Modern Optics, 2014
Vol. 61, No. 16, 1298–1308, http://dx.doi.org/10.1080/09500340.2014.930528

Comparison and unification of non-Hermitian and Lindblad approaches with applications to


open quantum optical systems
Konstantin G. Zloshchastieva∗ and Alessandro Sergia,b
a School of Chemistry and Physics, University of KwaZulu-Natal, Pietermaritzburg, South Africa; b National Institute for Theoretical
Physics (NITheP), KwaZulu-Natal Node, Pietermaritzburg, South Africa
(Received 2 April 2014; accepted 27 May 2014)

We compare two approaches to open quantum systems, namely, the non-Hermitian dynamics and the Lindblad master
equation. In order to deal with more general dissipative phenomena, we propose the unified master equation that combines
the characteristics of both of these approaches. This allows us to assess the differences between them as well as to clarify
Downloaded by [York University Libraries] at 13:19 13 August 2014

which observed features come from the Lindblad or the non-Hermitian part, when it comes to experiment. Using a generic
two-mode single-atom laser system as a practical example, we analytically solve the dynamics of the normalized density
matrix operator. We study the two-level model in a number of cases (depending on parameters and types of dynamics),
compute different observables and study their physical properties. It turns out that one can not only able to describe the
different types of damping in dissipative quantum optical systems but also mimic the undamped anharmonic oscillatory
phenomena which happen in quantum systems with more than two levels (while staying within the framework of the
analytically simple two-mode approximation).
Keywords: quantum optics; open quantum systems; non-Hermitian dynamics; laser theory; two-mode approximation;
master equation; density operator

1. Introduction In the present work, we compare the evolution arising


There are various theoretical methods that can be used in from the Lindblad approach to the one from the non-
order to study open quantum systems. One can start by con- Hermitian approach. Moreover, we propose a “hybrid” for-
sidering the totality of the degrees of freedom and malism that combines features of both of them, and, as
afterwards focus, either analytically [1] or numerically [2], such, it is expected to have a wider range of applicability.
on a subset of relevant coordinates. However, the approach In order to illustrate the theory, we consider quantum two-
that is currently more popular, immediately integrates over level system (TLS) and study its time evolution while mim-
the degrees of freedom of the environment and, for the icking the coupling to a dissipative environment through
Markovian systems, establishes the Lindblad master equa- non-Hermitian dynamics, Lindblad master equation or their
tion [3,4]. Originally introduced in spin physics and quan- combination. The dynamics of the density matrix is solved
tum optics [5,6], master equations have been applied to analytically in a number of relevant cases, the behavior of
many dissipative quantum phenomena [7–9]. primary observables (averages) is studied accordingly.
There is yet another approach to open quantum systems This paper is structured as follows. In Section 2 we give a
that is based on a non-Hermitian extension of quantum me- brief outline of the Lindblad and non-Hermitian approaches.
chanics (NHQM) [10–22]. In this approach the Hamiltonian In Section 3 we combine them into a unified approach,
is assumed to acquire an anti-Hermitian part which can be named “hybrid” throughout the paper, and adopt a two-
associated with dissipative effects. Both the Lindblad and level system as a practical example. In Sections 4–6 we
non-Hermitian approaches are based on certain simplify- consider different limits of the hybrid dynamics of the two-
ing assumptions and have their own range of applicability. level model, derive their analytical solutions and calculate
Moreover, one can find different physical motivations and the relevant observable properties. Some facts about two-
theoretical advantages in their respective use: the Lindblad level systems in quantum optics and relevant notations are
master equation is linear and permits the simple calculation reminded in the Appendix 1. The definitions of the Fourier
of averages whereas non-Hermitian dynamics possesses a transforms used in the paper are given in the Appendix 2.
generalized canonical structure that leads to a well-defined Conclusions are drawn in Section 7.
classical limit [21].

∗ Corresponding author. Emails: k.g.zloschastiev@gmail.com, zloschastievk@ukzn.ac.za

© 2014 Taylor & Francis


Journal of Modern Optics 1299

2. Open quantum system dynamics atoms interacting with the electromagnetic field in the pres-
In this section we give a brief description of two popu- ence of a thermal reservoir of radiation modes [7–9].
lar density-operator-based approaches which are used in a Using the notation of Appendix 1, we can write the master
theory of open quantum systems. In both approaches, one equation in the interaction picture as
distinguishes the coordinates of a subsystem from those d i  
of the environment so that the total Hamiltonian can be ρ̂(I ) (t) = σ̂+ + σ̂− , ρ̂(I ) (t) + D̂(ρ(I ) (t)), (5)
dt 2
written as the sum HT = HS + HB of the Hamiltoni- where the dissipator is given by
ans of the relevant subsystem, HS , and of the environment
1
(or bath), HB . Accordingly, the total Hilbert space becomes D̂(ρ) = γ0 (N + 1) σ̂− ρ̂ σ̂+ − {σ̂+ σ̂− , ρ̂}
the product of two composing Hilbert spaces, 2
1
HT = HS ⊗ HB , (1) + γ0 N σ̂+ ρ̂ σ̂− − {σ̂− σ̂+ , ρ̂} , (6)
2
where HS and HB are the Hilbert spaces of the relevant where γ0 is the spontaneous emission rate and N = N (ω0 )
subsystem and bath, respectively. The density operator of denotes the Planck distribution at the transition frequency
the relevant subsystem is obtained by tracing out the degrees [9].
of freedom of the environment from the total density matrix:
Downloaded by [York University Libraries] at 13:19 13 August 2014

ρ̂S ≡ ρ̂ = tr B ρ̂T , (2)


2.2. Non-Hermitian approach
where tr B denotes the partial trace over the degrees of free- Non-Hermitian dynamics has found various applications in
dom of the environment B. Despite the common basis, the the study of open quantum system [10–20]. Recently, it has
Lindblad and non-Hermitian approaches describe different been shown [19] that this approach is capable of describing
effects of the environment onto the subsystem and produce the evolution of pure states into mixed ones – since the
different properties of the latter. purity is not necessarily conserved if the dimension of the
corresponding Hilbert space is larger than two. In turn,
such a feature can be used in the quantitative modeling of
2.1. Lindblad master equation
the observer-related phenomena in quantum mechanics –
This approach, while maintaining a Hamiltonian contribu- such as the problem of measurement and phenomenon of
tion to the dynamics of the subsystem, modifies the evolu- decoherence. In the simplest formulation of this approach, it
tion equation of the subsystem density operator by adding is assumed that the dissipative effects of the environment are
dissipative terms (2). Upon using the Markov approxima- somehow encoded in anti-Hermitian terms of the subsys-
tion and some auxiliary simplifications, one can show that tem Hamiltonian which appear after averaging (“integrating
the Lindblad master equation takes the form: out”) the degrees of freedom of environment. Hence, in
d i   this approach one deals exclusively with the degrees of
ρ̂(t) = − Ĥ+ , ρ̂(t) + D̂(ρ(t), Ak ), (3)
dt  freedom of the subsystem, which is in turn described by
† a non-Hermitian Hamiltonian.
where Ĥ+ = Ĥ+ is a Hamiltonian that commutes with that
The non-Hermitian Hamiltonian operator can be always
of the subsystem ĤS . The dissipator D̂(ρ, Ak ) in partitioned into Hermitian and anti-Hermitian parts
Equation (3) is a linear operator in ρ̂(t) and quadratic in
the Lindblad operators Âk , k = 1, . . . , N 2 − 1 and N = Ĥ = Ĥ+ + Ĥ− , (7)
dim(H S ). The dissipator must be traceless for the trace †
where we denoted Ĥ± = ± Ĥ±= 12 ( Ĥ ± Ĥ † ). For further
of the density operator to be conserved during evolution.
study, it is convenient to introduce also the self-adjoint
The Lindblad operators Ak describe directly the dissipative
operator ˆ ≡ i Ĥ− which will be referred as the decay rate
effects of the environment. The most general quantum dy-
operator throughout the paper. Starting from the Schrödinger
namical semigroup form of the dissipator [9] can be written
equation, it is easy to show that the evolution equation for
as
the density operator acquires an anti-commutator term:
 
i   i
N2 −1
† 1 † 1 † d
D̂(ρ, Ak ) = γk Âk ρ̂ Âk − Âk Âk ρ̂ − ρ̂ Âk Âk , ρ̂(t) = − Ĥ+ , ρ̂(t) − Ĥ− , ρ̂(t) . (8)
2 2 dt  
k=1
(4) Equation (8) can also be written in matrix form [23], directly
where γ ’s are non-negative quantities which can be implementing, within a quantum framework, the original
expressed in terms of certain correlation functions of the geometric ideas of Grmela [24] about dissipation.
environment and play the role of relaxation rates for differ- While equations of the form (8) find some applications
ent decay modes of the open subsystem. [14,18], they also possess certain features (which also man-
Simple examples of the application of the Lindblad mas- ifest themselves when working with the state vectors) that
ter equation are found when studying models of two-level narrow their applicability range. For instance, the trace of
1300 K.G. Zloshchastiev and A. Sergi

the density operator determined by such equations is not 3. “Hybrid” formalism


preserved in general: In this section, we unify the Lindblad master equation with
d  2   the non-Hermitian equation for the density matrix. Such a
tr ρ̂(t) = tr ρ̂(t) Ĥ− . (9) hybrid equation is postulated replacing the usual
dt i
Hamiltonian contribution to the evolution of the non-
This renders the usual probabilistic interpretation of quan- normalized density matrix of the quantum subsystem in
tum mechanics more problematic to achieve. Another issue the Lindblad master equation with a more general non-
arises if one studies the invariance of the evolution equation Hermitian one, taken from the NH Equation (8). In such
under the Hamiltonian “gauge” shift a way one obtains the following equation

Ĥ → Ĥ + c0 Iˆ, d i   i
(10) ρ̂(t) = − Ĥ+ , ρ̂(t) − Ĥ− , ρ̂(t) + D̂(ρ(t), Ak ),
dt  
(14)
where c0 is the c-number and Iˆ is an identity operator. As
where
it happens in conventional quantum mechanics, one would
Ĥ = Ĥ+ + Ĥ− = Ĥ+ − i . ˆ (15)
like that such a shift should affect neither the observable
averages nor the evolution equation. However, according Upon substituting the normalized density operator (11) into
Downloaded by [York University Libraries] at 13:19 13 August 2014

to (8), this invariance gets broken if c0 has an imaginary Equation (14), one obtains a non-linear evolution equation,
part.
d  i   i
In view of these circumstances, in our previous work [19], ρ̂ (t) = − Ĥ+ , ρ̂  (t) − Ĥ− , ρ̂  (t)
we proposed to consider the normalized density operator, dt  
2i  
 + D̂(ρ  (t), Ak ) + tr ρ̂  (t) Ĥ− ρ̂  (t). (16)
ρ̂  = ρ̂/tr ρ̂ , (11) 
It is shown below that this non-linearity makes the models,
as a primary physical object of theory. Following this idea, based on the hybrid equations, substantially more interest-
the quantum average of an observable Ô = Ô(0) is defined ing than those obtained from the Lindblad or non-Hermitian
in terms of the normalized density operator in Equation (11) equations alone.
as In what follows, we study a two-level optical quantum
     system which is both an instructive example and a fruitful
Oobs ≡ tr ρ̂  Ô(0) = tr ρ̂ Ô(0) /tr ρ̂(t) . (12)
physical application. Using the notation and the definition of
the system given in Appendix 1, we assume that the evolu-
This idea was also adopted in [20] where the evolution
tion is governed by Equations (14), (15) and (6). The model
equation, which can be derived for the normalized density
we study is defined by the following Hermitian Hamiltonian
operator in our approach, was favored over the equations for
the non-normalized operator and state vectors which were Ĥ+ = Ĥ0 + ĤL , (17)
used previously (cf. [18], for instance). It turns out that the
normalized density operator approach automatically solves where
the above-mentioned issues of norm conservation and gage 1
Ĥ0 = ω0 σ̂3 , (18)
invariance: using the evolution equation which follows from 2
1  
(8) and (11), ĤL =  e−iω0 t σ̂+ + eiω0 t σ̂− . (19)
2
d  i   i
ρ̂ (t) = − Ĥ+ , ρ̂  (t) − Ĥ− , ρ̂  (t) The unperturbed Hamiltonian Ĥ0 can represent the two
dt  
2i   energy levels of a free dipole. In such a case, the perturba-
+ tr ρ̂  (t) Ĥ− ρ̂  (t), (13) tion ĤL would describe the interaction between the dipole

and a single-mode electromagnetic wave. More details and
one can easily check that the normalization property tr(ρ̂  ) = corresponding notations are provided in the Appendix 1.
1 is conserved and that the “gauge” invariance under the The anti-Hermitian Hamiltonian, which must be added
transformation (10) is achieved for arbitrary c0 . to Ĥ+ to give the total Hamiltonian Ĥ of the model, is
To conclude, the main advantage of the normalized
Ĥ− = Ĥ + Ĥ D + Ĥ00 , (20)
density-operator approach in NHQM is that it handles in
a unified way not only the pure states but also the mixed where
ones. Moreover, the emerging non-linearity of the evolution 1
Equation (13) provides yet another example of a profound Ĥ = i σ̂3 , (21)
2
interplay between the physics of open quantum systems 1  
and non-linear quantum mechanics: the environment is able Ĥ D = − iα e−iω0 t σ̂+ + eiω0 t σ̂− , (22)
2
to induce effective non-linearities in quantum evolution 1
equations [25–40]. Ĥ00 = − iT Iˆ, (23)
2
Journal of Modern Optics 1301

where , α and T are real-valued free parameters. The a single matrix


 equation for the four  unknown functions,
Hamiltonian Ĥ ∝ σ̂+ σ̂− is motivated by the physics of σμ (t)(I ) ≡ σ (t)(I ) , tr(ρ̂(I ) (t)) (μ = 1, . . . , 4):
photodetection and continuous measurements in presence d
of radiation modes, cf. Section 6.3.1 of [9]. The term Ĥ D is σμ (t)(I ) = Mσμ (t)(I ) , (31)
dt
the anti-Hermitian counterpart of ĤL ; therefore, it is sup-
where
posed to describe the dissipative processes accompanying ⎛ ⎞
the dipole interaction of the atom and electromagnetic field. − 12 γ − T 0 0 −α
⎜ 0 − 12 γ − T  0 ⎟
The term Ĥ00 is a “gauge” term – as mentioned in the M=⎜

⎟,

Section 2.2, it does not affect observable values (as defined 0 − −γ − T  − γ0
by (11) and (12)); however, it can be used for simplifying −α 0  −T
⎛ ⎞
or regularizing intermediate expressions. σ̂1
In terms of the above, the evolution equation of the model ⎜ σ̂2 ⎟
σ̂μ = ⎜

⎟. (32)
in the interaction picture reads σ̂3 ⎠
d i   α  Iˆ
ρ̂(I ) = σ̂+ + σ̂− , ρ̂(I ) − σ̂+ + σ̂− , ρ̂(I )
dt 2  2  Equation (31) is in a matrix form that is useful in order
1 to search for solutions. Nevertheless, it is instructive to
Downloaded by [York University Libraries] at 13:19 13 August 2014

+ γ0 (N + 1) σ̂− ρ̂(I ) σ̂+ − {σ̂+ σ̂− , ρ̂(I ) }


2 re-write it in terms of normalized averages. Using Equa-
 
1 tions (11) and (25), one can write the normalized density
+ γ0 N σ̂+ ρ̂(I ) σ̂− − {σ̂− σ̂+ , ρ̂(I ) }
2 operator in a decomposed form
   
1 ˆ  1 ˆ 
3
+ σ̂3 , ρ̂(I ) − T ρ̂(I ) . (24)  
2 ρ̂(I ) (t) = I+ σ̂i σi (t)(I ) = I + σ̂3 σ3 (t)(I )
2 2
It is convenient to search for solutions of this equation in i=1
the form + σ̂+ σ− (t)(I ) + σ̂− σ+ (t)(I ) , (33)
 
1 ˆ 3 where
ρ̂(I ) (t) = I tr(ρ̂(I ) (t)) + σ̂i σi (t)(I )
2
i=1
σi (t)(I ) = tr(σ̂i ρ̂(I

) (t)) = σi (t)(I ) /tr(ρ̂(t)),(34)
1 ˆ 
(i = 1, . . . , 3) are the observable average values in the
= I tr(ρ̂(I ) (t)) + σ̂3 σ3 (t)(I )
2 interaction picture which satisfy the equation
+ σ̂+ σ− (t)(I ) + σ̂− σ+ (t)(I ) , (25) d 
σ (t)(I ) = Geff (t)σ  (t)(I ) + b, (35)
where dt
σi (t)(I ) = tr(σ̂i ρ̂(I ) (t)), (26) where

with i = 1, . . . , 3, are auxiliary average values that are Geff (t) = G + F(I ) (t) + T Iˆ
⎛ ⎞
regarded as unknown functions of time, together with F(I ) (t) − 12 γ 0 0
tr(ρ̂(I ) (t)) = tr(ρ̂(t)). The equations for the average values =⎝ 0 F(I ) (t) − 12 γ  ⎠,
easily follow from Equation (24) 0 − F(I ) (t) − γ
d (36)
σ (t)(I ) = Gσ (t)(I ) + b tr(ρ̂(I ) (t)), (27)
dt and F(I ) (t) ≡ ασ1 (t)(I ) − σ3 (t)(I ) . While this form
d of evolution equation is somewhat unsuitable for searching
tr(ρ̂(I ) (t)) = −ασ1 (t)(I ) + σ3 (t)(I )
dt for analytical solutions (due to its non-linearity with respect
− T tr(ρ̂(I ) (t)), (28) to unknown functions σi (t)(I ) ), it allows us to demon-
strate a feature mentioned in Section 2.2: the contribution
where we have introduced the matrix
⎛ 1 ⎞ from the “gauge” term Ĥ00 disappears when one deals with
−2γ − T 0 0 observable values.
G=⎝ 0 − 12 γ − T  ⎠, (29)
0 − −γ − T
4. General solution
and three-dimensional vector
⎛ ⎞ In this section, we search for solutions of Equation (31)
−α in the zero-temperature limit. This is equivalent to setting
b=⎝ 0 ⎠. (30) γ = γ0 . Imposing the gage condition T = 0 and rescaling
 − γ0 the time variable τ = t, we reduce (31) to the form:
We have also adopted the vector notation σ = (σ1 , σ2 , σ3 ). d  σμ (τ )(I ) ,
σμ (τ )(I ) = M (37)
It is very convenient to combine Equations (27) and (28) into dτ
1302 K.G. Zloshchastiev and A. Sergi

(μ = 1, . . . , 4) where
⎛ ⎞ 0.8
−2γ̃0 0 0 −α̃
⎜ 0 −2 γ̃ 1 0 ⎟
 =⎜
M
0 ⎟,
⎝ 0 ˜
−1 −4γ̃0  − 4γ̃0 ⎠ 0.6

−α̃ 0 ˜ 0
⎛ ⎞
σ̂ (τ ) 0.4
 1 ⎟
  ⎜ ⎜ σ̂2 (τ ) ⎟
σ̂μ (τ ) = ⎜ ⎟, (38)
⎝ σ̂3 (τ ) ⎠ 0.2

trρ(τ )
 we have introduced the symbols
In the definition of M, 0
5 10 15 20
γ̃0 = γ0 /4, α̃ = α/  and ˜ = / . As in the previous
sections, one should keep in mind that the non-normalized Figure 1. The population of the upper level pe as a function of
values σμ  are auxiliary quantities that are used for com- time τ = t for the parameters α̃ = ˜ = 0 and γ̃0 = 1 (solid
puting the observables, Equations (33) and (34). curve), γ̃0 = 1/4 (dashed curve), γ̃0 = 1/8 (dash-dotted curve),
and γ̃0 = 1/40 (dotted curve). (The color version of this figure is
Downloaded by [York University Libraries] at 13:19 13 August 2014

The solution of Equation (37) can be formally written in


included in the online version of the journal.)
a matrix exponential form:

σμ (τ )(I ) = eMτ σμ (0)(I ) , (39)
where 0.2

σμ (0)(I ) = σμ (0) = σμ (0)(I ) = σμ (0) (40)


are initial values. We used the property ρ̂(0) = ρ̂  (0), 0

which is valid both in the Schrödinger and in the interaction


picture. Naturally, one also finds that
0.2
σ4 (0)(I ) = σ4 (0) = σ4 (0)(I ) σ4 (0) = tr ρ̂(0) = 1,
(41)
for all physical situations. 0.4

Furthermore, if the matrix M  is diagonalizable, the gen-


5 10 15 20
eral solution (39) can be written in the more convenient
form 
Figure 2. The imaginary part of the coherence, Im σ+ obs , as
σμ (τ )(I ) = Sσμ (0)(I ) , (42) a function of time τ = t for the parameters α̃ = ˜ = 0 and
⎛ M τ ⎞ γ̃0 = 1 (solid curve), γ̃0 = 1/4 (dashed curve), γ̃0 = 1/8 (dash-
e (1) 0 0 0
⎜ 0 M(2) τ ⎟ −1 dotted curve), and γ̃0 = 1/40 (dotted curve). (The color version
e 0 0
S = P⎜ ⎝ 0
⎟P ,

of this figure is included in the online version of the journal.)
0 e M(3) τ 0
0 0 0 e M(4) τ
(43)
 and P is the matrix whose the Lindblad term (i.e. α =  = 0) and the case when only
where M(μ) are eigenvalues of M,
 The four-by-four matrix the anti-Hermitian term (i.e. γ0 = 0) is present. This will
columns are the eigenvectors of M.
 allow us to obtain a clearer understanding of the capabilities
M, with its four eigenvalues, arises from the hybrid master
of the hybrid formalism.
equation. Such an equation is defined in terms of both the
dissipator and the total Hamiltonian Ĥ+ + Ĥ− (which has
just two eigenvalues for the model under study). Hence, 5.1. Lindblad-driven dissipation
the four eigenvalues of M carry more physical information  has the following four
When α =  = 0, the matrix M
about the studied model system than the information carried eigenvalues:
by the eigenvalues of the total Hamiltonian alone.
(L) (L)
M1 = −2γ̃0 , M2 = −3γ̃0 − iκ,
(L) (L)
M3 = −3γ̃0 + iκ, M4 = 0, (44)
5. Limit cases 
In this section, we consider two special (limit) cases of the where the quantity κ = 1 − γ̃02 can be imaginary or real-
general solution found in Section 4. In particular, we treat valued. Correspondingly, the solution from Section 4 can
the case when dissipative effects are modeled either only by be written as
Journal of Modern Optics 1303

σμ (τ )(I ) = S(L) σμ (0)(I ) ,


⎛ −2γ̃ τ ⎞
e 0 0 0 0
⎜ 0 
⎜ f 1 (τ )e−3γ̃0 τ sin (κτ ) −3γ̃0 τ
κ e 4γ̃0
f 3 (τ )e−3γ̃0 τ − 1 ⎟ ⎟
⎜ 8γ̃02 +1 ⎟
S(L) = ⎜ ⎟ , (45)
⎜ 0 − sin κ(κτ ) e−3γ̃0 τ f −1 (τ )e−3γ̃0 τ
8γ̃02
f −ν 2 (τ )e −3γ̃ τ −1 ⎟
⎝ 8γ̃02 +1
0

0 0 0 1

where f k (τ ) ≡ cos (κτ ) + k γ̃κ0 sin (κτ ) and ν 2 = 1 + The large-times asymptotic (steady-state) values of the
1/(2γ̃02 ). From the last row of this matrix it follows that spin averages can be found by taking an appropriate limit
σ4 (τ )(I ) = tr(ρ̂(τ )) = 1. Therefore, the physical (nor- in Equation (45). One obtains
⎛ ⎞
malized) averages coincide with the auxiliary ones: 0
4 γ̃
σi (t) = σi (t), i = 1 . . . 3. It is easy to check that this σ  (+∞)(I ) = σ (+∞)(I ) = − 2
0 ⎝ 1 ⎠.
solution coincides with a textbook example – see, for in- 8γ̃0 + 1 2γ̃
0
Downloaded by [York University Libraries] at 13:19 13 August 2014

stance, Section 3.4.5.1 of [9]. The profiles of most important (46)


observables are given in the Figures 1 and 2. These values are indeed stationary points of the system.
Needless to say, they coincide with the textbook values, cf.
1 Section 3.4.5.1 of [9].

0.8
5.2. Anti-Hermitian-driven dissipation
When γ̃0 = 0 then the matrix M  has the following four
0.6
eigenvalues:
 
(A) (A) (A) (A)
M1 , M2 , M3 , M4 = (λ1 , λ2 , λ3 , λ4 )
0.4   
R+ − R1 R+ − R1
= − , ,
2 2
0.2   
R+ + R1 R+ + R1
− , , (47)
0 2 2
5 10 15 20
 
where we have denoted R1 = R+ 2 +4α̃ 2 = R 2 +4α̃ 2 

˜2
Figure 3. The population of the upper level pe as a function of
time τ = t for the parameters γ̃0 = ˜ = 0 and α̃ = 4 (solid and R± = α̃ ± (˜ − 1). One can see that R1  R±
2 2

curve), α̃ = 1 (dashed curve), α̃ = 1/2 (dash-dotted curve), so that the first two eigenvalues (47) are always imaginary-
and α̃ = 1/20 (dotted curve). (The color version of this figure
is included in the online version of the journal.)
1

0.4
0.8

0.2
0.6

0
0.4

0.2

0.2

0.4

0
5 10 15 20 5 10 15 20


Figure 4. The imaginary part of the coherence, Im σ+ obs , as Figure 5. The population of the upper level pe as a function of
a function of time τ = t for the parameters γ̃0 = ˜ = 0 and time τ = t for the parameters α̃ = γ̃0 = 0 and ˜ = 2 (solid
α̃ = 4 (solid curve), α̃ = 1 (dashed curve), α̃ = 1/2 (dash-dotted curve), ˜ = 1 (dashed curve), ˜ = 0.9 (dash-dotted curve),
curve), and α̃ = 1/20 (dotted curve). (The color version of this and ˜ = 1/2 (dotted curve). (The color version of this figure
figure is included in the online version of the journal.) is included in the online version of the journal.)
1304 K.G. Zloshchastiev and A. Sergi

valued. Correspondingly, the solution from Section 4 can σ  (+∞)(I )


⎛ ⎞
be written as 1 α̃ K α̃ 2 λ−1
4 
˜ −α̃ 2 ˜ α̃ 2 λ̄4
1 ⎜ 22 + ⎟
= ⎝ λ2 λ4 ˜ −α̃λ2 λ̄2 α̃λ2 λ̄2 λ4 α̃ ˜ ⎠σ (0),
T̃(A)
−α̃ ˜ −α̃λ2 λ̄2 λ4 2 α̃ K − α̃λ4 ˜
2 1

σμ (τ )(I ) = S(A) σμ (0)(I ) ,


⎛ ˜ ˜ 1,−1 (τ ) ⎞
2 C K − ,K + (τ ) α̃ S −1 (τ ) α̃ C α̃Sλ̄2 ,−λ̄4 (τ )
1
−λ−1
2 ,λ4
⎜ λ λ ˜
 ⎟
1 ⎜ ˜ 1,−1 (τ ) ⎟
α̃ S−λ4 ,λ2 (τ ) Cλ4 λ̄4 ,−λ2 λ̄2 (τ ) −λ2 λ4 Sλ̄4 ,−λ̄2 (τ ) −C
2 4
S(A) = ⎜ ⎟, (48)
R1 ⎝ ˜ 1,−1 (τ )
α̃ C λ2 λ4 Sλ̄4 ,−λ̄2 (τ ) 1
C (τ ) ˜
S −λ2 ,λ4 (τ ) ⎠
2 K + ,K −
− 2α̃ Sλ2 K − ,λ4 K + (τ )
1 ˜
C1,−1 (τ ) ˜
S−λ2 ,λ4 (τ ) C−λ2 λ̄2 ,λ4 λ̄4 (τ )

where we have denoted


where we have denoted T̃(A) = − 12 λ4 K + σ1 (0) − α̃ ˜
Ck1 ,k2 (τ ) ≡ k1 cosh (λ2 τ ) + k2 cosh (λ4 τ ) ˜ 3 (0) + α̃λ4 λ̄4 . Further simplifying this
σ2 (0) + α̃λ4 σ
Downloaded by [York University Libraries] at 13:19 13 August 2014

= k1 cos (|λ2 |τ ) + k2 cosh (|λ4 |τ ), expression, we eventually obtain


Sk1 ,k2 (τ ) ≡ k1 sinh (λ2 τ ) + k2 sinh (λ4 τ ) ⎛ ⎞
−α̃ λ̄4
1 ⎝ ˜ ⎠ ,
= ik1 sin (|λ2 |τ ) + k2 sinh (|λ4 |τ ), σ  (+∞)(I ) = (50)
 λ4 λ̄4
λ4 ˜
and K ± = R1 ± R− = R− ˜ 2 ± R− and λ̄k =
2 + 4α̃ 2 

λk +1/λk . As before, the physical values are the normalized which means that the asymptotic (steady-state) averages do
ones: not depend on the initial values, as in the Lindblad case. One

σ  (τ )(I ) = S(A) σμ (0)(I ) ,


⎛ 1 ⎞
˜ ˜ 1,−1 (τ )
2 C K − ,K + (τ ) α̃ S −λ−1
2 ,λ4
−1 (τ ) α̃ C α̃Sλ̄2 ,−λ̄4 (τ )
1 ⎜ ⎟
S(A) = ⎝ λ2 λα̃4 ˜ S−λ4 ,λ2 (τ ) Cλ λ̄ ,−λ λ̄ (τ ) −λ2 λ4 Sλ̄ ,−λ̄ (τ ) ˜ 1,−1 (τ ) ⎠ ,
−C (49)
T(A) 4 4 2 2 4 2
˜ 1,−1 (τ )
α̃ C λ2 λ4 Sλ̄4 ,−λ̄2 (τ ) ˜ −λ2 ,λ4 (τ )
2 C K + ,K − (τ ) S
1

where T(A) is an internal product of the fourth row of S(A)


(omitting the overall factor) and the four vectors of initial can see that the important difference from the Lindblad case
values (40): is that σx (I ) does not vanish at large times.
1 Another distinctive feature of the anti-Hermitian-driven
T(A) = − Sλ2 K − ,λ4 K + (τ )σ1 (0) + C ˜ 1,−1 (τ )σ2 (0)
2α̃ dynamics can be found if one computes the Fourier trans-
+ S˜ −λ2 ,λ4 (τ )σ3 (0) + C−λ λ̄ ,λ λ̄ (τ ). form of the observables, such as the population of an upper
2 2 4 4
level pe . The informative part of the Fourier transform is,
Depending on whether the eigenvalue λ4 vanishes or not, according to (B3),
one can consider the following two cases. !∞
1
[ pe (ω̃)]reg ∝ √ ( pe (τ ) − pe (∞)) ei ω̃τ dτ, (51)

5.2.1. Exponential damping 0
This case takes place when the parameters of the model where ω̃ = ω/ , and its functional dependence can be
are such that λ4 = 0, or, alternatively, sign (R+ ) + derived with the use of the Equations (49), (A2) and (A14).
1 + (2α̃/R+ )2 = 0. Then the system exhibits the expo- Since the Fourier transforms are essentially complex-valued,
nential damping which is qualitatively, but not necessarily one should consider separately moduli and phases.
quantitatively, similar to the Lindblad-driven dynamics. It turns out that the absolute value of [ pe (ω̃)]reg behaves
Some generic profiles of most important observables are in a qualitative similar way to its Lindblad-driven counter-
given in Figures 3 and 4 (all curves), as well as in Figures part, under the same initial conditions. The differences arise
5 and 6 (solid and dashed curves only). One can notice that when one consider the phase of [ pe (ω̃)]reg . While in the
Figures 3 and 4 qualitatively resemble the Lindblad ones. Lindblad-driven case, it is a smooth function for positive ω̃
The large-times asymptotic values of the spin averages which is bound between −π/2 and π/2 with the asymptotic
can be found by taking an appropriate limit in (49). At first value π/2 (see Figure 7), in the anti-Hermitian-driven case
one obtains it is exactly opposite. From the Figure 8 one can see that
Journal of Modern Optics 1305

the phase of [ pe (ω̃)]reg varies between π/2 and −π/2 with !To
1
the asymptotic value −π/2. [ pe (n)]reg ∝ pe (τ )e2πi(n/To )τ dτ, (53)
To
0
5.2.2. Anharmonic oscillations
In this case, the model parameters are such that λ4 = 0, where To = 2π/ 1 − ˜ 2 , and its specific functional form
which is equivalent to the following two conditions: α̃ = 0 can be derived with the use of the Equations (52), (A2) and
and ˜ 2 < 1. The solution exhibits purely oscillatory (A14). The profile of the computed modulus of [ pe (n)]reg
behavior: is shown in the Figure 9.

⎛ ⎞
ω2 σ1 (0)
1 ⎝
σ  (τ )(I ) =  ˜ − ˜ 2 σ2 (0) − (˜ − σ2 (0)) cos (ω τ ) + ω σ3 (0) sin (ω τ ) ⎠ , (52)
T(A)
ω2 σ3 (0) cos (ω τ ) + (˜ − σ2 (0)) sin (ω τ )

where the oscillation frequency ω = 1 − ˜ 2 , and T(A)  = To summarize, we have shown that the anti-Hermitian
1 − σ˜ 2 (0) − (˜ ˜ − σ2 (0)) cos (ω τ ) + ω σ3 (0) two-level models of this type can actually mimic the prop-
sin(ω τ ). erties of quantum systems with more than two levels. In
Downloaded by [York University Libraries] at 13:19 13 August 2014

Some profiles of most important observables are given in this picture, the parameter ˜ turns out to be a qualitative
Figures 5 and 6 (dotted and dash-dotted curves only). They measure of the number of the additional (effective) levels.
exhibit interesting asymmetric oscillatory patterns which For example, Figure 9 shows that the TLS with ˜ = 1/2
do not appear in the Lindblad case. Such patterns indicate can be used to mimic the four-level or five-level system (if
a few important things. For instance, they show that the one neglects the “transitions” with the wavenumbers larger
anti-Hermitian terms in the Hamiltonian can induce not than four). The further decreasing of ˜ reduces the number
only the standard decay effects (such as the asymptotic of additional wave frequencies.
damping at large times) but also more sophisticated effects.
Indeed, in this case the oscillations are not damped, the 6. Approximate solution
role of the anti-Hermitian parameter ˜ is that it introduces
the asymmetry between the pumping and discharging of the The general solution derived in Section 4 becomes extremely
two-level system. In terms of frequency, it means that the bulky when expressed in terms of radicals. Luckily, in some
pumping frequency is larger than the discharge one. It is physical cases, one could use certain approximations which
similar to what happens in higher-than-two-level systems: drastically simplify final formulae. Indeed, in quantum-
first a system is pumped into the highest excited state, then optical two-level systems, the Rabi frequency usually takes
it spontaneously cascades down to its ground state, passing large values, up to the megahertz scale, whereas the dis-
the intermediate levels on its way. This is particularly clear sipative effects are small. It is thus natural to make the
to see when one takes a look at the Fourier transform of the assumption
population of the upper level pe . The informative part of γ̃0  1, α̃  1, ˜  1, (54)
the Fourier transform is, according to (B4),
1.5

0.4
1.0

0.2 0.5

0.0
0

0.5

0.2
1.0

0.4 1.5
0 1 2 3 4 5
5 10 15 20

 Figure 7. The phase of the Fourier transform pe (ω) versus the


Figure 6. The imaginary part of the coherence, Im σ+ obs , as frequency ω̃ = ω/ , for the parameters α̃ = ˜ = 0 and γ̃0 = 1/2
a function of time τ = t for the parameters α̃ = γ̃0 = 0 and (solid curve), γ̃0 = 1/4 (dashed curve), γ̃0 = 1/8 (dash-dotted
˜ = 2 (solid curve), ˜ = 1 (dashed curve), ˜ = 0.9 (dash-dotted curve), and γ̃0 = 1/100 (dotted curve). Two horizontal thin dotted
curve), and ˜ = 1/2 (dotted curve). (The color version of this lines mark the values ±π/2. (The color version of this figure is
figure is included in the online version of the journal.) included in the online version of the journal.)
1306 K.G. Zloshchastiev and A. Sergi

which corresponds to the strong driving limit, using the 3

textbook terminology [9]. However, this condition is not


2
enough since the perturbation theory has two different sec-
tors – Lindblad-dominated (when the approximate series 1
solution must converge to the solution from Section 5.1
when taking the limit α =  = 0) and anti-Hermitian- 0

dominated (when the series solution must converge to the


solution from Section 5.2 in the limit γ = 0). In the former 1

case, which is of main interest here, one should supplement


2
(54) with the assumption
ᾱ ≡ α/γ0  1, ¯ ≡ /γ0  1, (55) 3
0 1 2 3 4
then expand the exact solution (42) and the related observ-
ables in series with respect to these five small parameters, Figure 8. The phase of the Fourier transform pe (ω) (51) versus
and keep the leading-order terms. By doing so, we obtain the frequency ω̃ = ω/ , for the parameters γ̃0 = ˜ = 0 and
 has the following four eigenvalues, in the
that the matrix M α̃ = 2 (solid curve), α̃ = 1 (dashed curve), α̃ = 1/2 (dash-dotted
leading-order approximation: curve), and α̃ = 1/5 (dotted curve). Two horizontal thin dotted
Downloaded by [York University Libraries] at 13:19 13 August 2014

  lines mark the values ±π/2. (The color version of this figure is
(L D) (L D) (L D) (L D)
M1 , M2 , M3 , M4 included in the online version of the journal.)
≈ (−2γ̃0 − 2γ̃0 χ2 , −3γ̃0 − iχ1 , −3γ̃0 + iχ1 , 2γ̃0 χ2 ),
(56)

where we denoted χ1 = 1 + (2¯ − 1/2)γ̃02 and χ2 = ᾱ 2 /4, The large-times asymptotic values of the spin averages can
both being positive-definite values. Correspondingly, the be found by taking an appropriate limit in (59). Hence, we
solution is given by obtain in the leading-order approximation

σμ (τ )(I ) = S(L D) σμ (0)(I ) , (57)


(0) (α) ()
S(L D) ≈ S(L D) + S(L D) + S(L D) ,
⎛ −2γ̃0 (1+χ2 )τ ⎞
e 0 0 0
(0) ⎜ 0 h + (τ )e−3γ̃0 τ sin (χ1 τ )e−3γ̃0 τ 4γ̃0 h 2 (τ ) ⎟
S(L D) =⎜⎝ −3 γ̃ τ −3γ̃ τ −3γ̃ τ
⎟,

0 − sin (χ1 τ )e 0 h − (τ )e 0 −4γ̃0 sin (χ1 τ )e 0

0 0 0 2γ̃
e 0 2χ τ
⎛ ⎞
0 0 0 h 1 (τ )
⎜ −4γ̃0 h 1 (τ ) 0 0 0 ⎟
S(α) = ᾱ ⎜
⎝ −4γ̃ 2 h 2 (τ ) 0 0
⎟,
(L D)
0 0 ⎠
h 1 (τ ) 0 0 0
⎛ ⎞
0 −ᾱh 1 (τ ) α̃h 2 (τ ) 0
() ⎜ ᾱh 1 (τ ) 0 0 −h 2 (τ ) ⎟
S(L D) = ˜ ⎜
⎝ α̃h 2 (τ )
⎟, (58)
0 0 sin (χ1 τ )e−3γ̃0 τ ⎠
0 h 2 (τ ) sin (χ1 τ )e−3γ̃0 τ 0

where h 1 (τ ) = 12 e−2γ̃0 (1+χ2 )τ − e2γ̃0 χ2 τ , h 2 (τ ) = cos
⎛ ⎞
(χ1 τ )e−3γ̃0 τ − e2γ̃0 χ2 τ , and h ± (τ ) = cos (χ1 τ ) ± γ̃0 sin − 12 ᾱ − 2α̃ ˜
(χ1 τ ). As in Section 5.2, the physical values are the nor- ⎜ ⎟
σ  (+∞)(I ) = ⎝ ˜ − 4γ̃0 ⎠ . (60)
malized ones:
2γ̃0 (˜ − 4γ̃0 )
1
σ  (τ )(I ) = S σμ (0)(I ) , (59)
T(L D) (L D)
where S(L D) is the matrix S(L D) without the bottom row,
7. Conclusion
T(L D) is a product of the bottom row of S(L D) and the four
vectors of initial values (40): In this paper, we have compared two approaches to describ-
ing the effects of a general dissipative environment. Namely,
˜ 2 (τ )σ2 (0)
T(L D) = ᾱh 1 (τ )σ1 (0) + h we considered both the approaches based on the Lindblad
+ ˜ sin (χ1 τ )e 0 τ σ3 (0) + e2γ̃0 χ2 τ .
−3γ̃
master equation and the formalism based on introducing
Journal of Modern Optics 1307

Funding
0.4
This work was supported by the National Research Foundation of
South Africa.

0.3
References
[1] Attal, S.; Joye, A. In Open Quantum Systems I: The
0.2 Hamiltonian Approach; Pillet, C.A., Ed.; Springer: Berlin,
2006.
[2] Sergi, A.; Sinayskiy, I.; Petruccione, F. Phys. Rev. A 2009,
0.1
80, 012108.
[3] Gorini, V.; Kossakowski, A.; Sudarshan, E.C.G. J. Math.
0.0
Phys. 1976, 17, 821–825.
2 4 6 8 10 [4] Lindblad, G. Commun. Math. Phys. 1976, 48, 119–130.
[5] Bausch, R. Z. Phys. 1966, 193, 246–265.
[6] Haake, F. Springer Tracts Mod. Phys. 1973, 66, 98–168.
Figure 9. The modulus of the Fourier transform pe (n) (53) versus
[7] Carmichael, H.J. An Open Systems Approach to Quantum
the wavenumber n, for the parameters α̃ = γ̃0 = 0 and ˜ = Optics; Lecture Notes in Physics, Springer-Verlag, Berlin,
0.999 (dots), ˜ = 0.9 (squares), and ˜ = 1/2 (diamonds). (The 1993.
Downloaded by [York University Libraries] at 13:19 13 August 2014

color version of this figure is included in the online version of the [8] Gardiner, C.W.; Zoller, P. Quantum Noise; Springer-Verlag,
journal.) Berlin, 2000.
[9] Breuer, H.-P.; Petruccione, F. The Theory of Open Quantum
Systems; Oxford University Press, Oxford, 2002.
anti-Hermitian terms into the Hamiltonian. In Section 3, [10] Feshbach, H. Ann. Phys 1958, 5, 357–390; Feshbach, H.
we have proposed a “hybrid” formalism that unifies the 1962, 19, 287–313.
Lindblad and non-Hermitian approaches. This allowed us [11] Wong, J. J. Math. Phys. 1967, 8, 2039–2042.
[12] Faisal, F.H.M.; Moloney, J.V. J. Phys. B At. Mol. Opt. Phys.
not only to reveal the distinctive features of the approaches 1981, 14, 3603–3620.
but also to expand the range of dissipative phenomena that [13] Dattoli, G.; Torre, A.; Mignani, R. Phys. Rev. A 1990, 42,
can be accounted for. 1467–1475.
Using a two-level single-atom model as a practical appli- [14] Hegerfeldt, G.C. Phys. Rev. A 1993, 47, 449.
cation, we have obtained solutions of the hybrid equation [15] Baskoutas, S.; Jannussis, A.; Mignani, R.; Papatheou, V. J.
Phys. A Math. Gen. 1993, 26, L819–L824; Angelopoulou,
for the normalized density matrix operator. In Sections 4–6, P.; Jannussis, A.; Mignani, R.; Papatheou, V. Int. J. Mod.
we have also considered special (limit) cases and physically Phys. B 1995, 9, 2083–2104.
admissible approximations. Using the analytical solutions [16] Rotter, I. J. Phys. A 2009, 42, 153001. arXiv:0711.2926.
of all these cases, we have calculated those properties of [17] Geyer, H.B.; Scholtz, F.G.; Zloshchastiev, K.G. In Proceed-
the model that can be compared with experiments in order ings of 12th International Conference on Mathematical
Methods in Electromagnetic Theory, 250–252, 2008;
to assess whether a specific feature is either non-Hermitian Odessa.
driven or Lindblad driven. [18] Graefe, E.-M.; Schubert, R. Phys. Rev. A 2011, 83, 060101;
Remarkably, we have found that the anti-Hermitian terms J. Phys. A 2012, 45, 244033.
in the Hamiltonian can describe not only the mere dissipa- [19] Sergi, A.; Zloshchastiev, K.G. Int. J. Mod. Phys. B 2013, 27,
tive damping but also undamped anharmonic oscillatory 1350163. [arXiv:1207.4877].
[20] Brody, D.C.; Graefe, E.-M. Phys. Rev. Lett. 2012, 109,
phenomena. Such results are reported in detail in Section 230405. [arXiv:1208.5297].
5.2 where we also showed that this kind of anharmonicity [21] Graefe, E.-M.; Höning, M.; Korsch, H.J. J. Phys. A 2010,
can be used to mimic the cascaded quantum systems with 43, 075306.
more than two levels. In future, it would be interesting to [22] Thilagam, A. J. Chem. Phys. 2011, 136, 065104.
apply the hybrid formalism to those multi-level lasers or [23] Sergi, A. Comm. Theor. Phys. 2011, 56, 96–98.
[24] Grmela, M. Phys. Lett. A 1984, 102, 355–358.
spin systems that can be modeled in the leading order of [25] Gisin, N. J. Phys. A 1981, 14, 2259–2267.
approximation by means of only two states. In particular, [26] Gisin, N. Physica A 1982, 111, 364–370.
we have in mind those systems where one of the two energy [27] Gisin, N. J. Math. Phys. 1983, 24, 1779–1782.
levels is actually a band or a bundle of a few closely situated [28] Korsch, H.J.; Steffen, H. J. Phys. A 1987, 20, 3787–3803.
levels. [29] Kostin, M.D. J. Chem. Phys. 1972, 57, 3589–3591.
[30] Kostin, M.D. J. Stat. Phys. 1975, 12, 145–151.
[31] Bialynicki-Birula, I.; Mycielski, J. Ann. Phys. 1976, 100,
62–93.
Acknowledgements [32] Yasue, K. Ann. Phys. 1978, 114, 479–496.
This article is based on the talks given at the conferences “12th [33] Lemos, N.A. Phys. Lett. A 1980, 78, 239–241.
International Workshop on Pseudo-Hermitian Hamiltonians in [34] Brasher, J.D. Int. J. Theor. Phys. 1991, 30, 979–984.
Quantum Physics” (02–06 July, 2013, Koç University, Istanbul, [35] Schuch, D. Phys. Rev. A 1997, 55, 935–940.
Turkey) and “Quantum Information Processing, Communication [36] Davidson, M.P. Nuov. Cim. B 2001, 116, 1291–1294.
and Control 2” (25–29 November, 2013, KwaZulu-Natal, South [37] Lopez, J.L. Phys. Rev. E. 2004, 69, 026110.
Africa). [38] Zloshchastiev, K.G. Grav. Cosmol. 2010, 16, 288–297.
1308 K.G. Zloshchastiev and A. Sergi

[39] Avdeenkov, A.V.; Zloshchastiev, K.G. J. Phys. B At. Mol. picture to the interaction one. One starts with the unitary transfor-
Opt. Phys. 2011, 44, 195303. [arXiv:1108.0847]. mation of the density operator
[40] Zloshchastiev, K.G. Eur. Phys. J. B 2012, 85, 273.
[arXiv:1204.4652]. ρ̂(I ) = ei Ĥ0 t/ ρ̂ e−i Ĥ0 t/ , (A8)
where Ĥ0 is chosen as in (A6). This implies the transition formulae
Appendix 1. Two-level systems in quantum optics tr ρ̂(I ) = tr ρ̂, (A9)
The two-mode open quantum system is a basic yet very instructive tr(ρ̂(I ) Ô (I ) ) = tr(ρ̂ Ô), (A10)
example of an open quantum (sub)system. In quantum optics, its
most obvious manifestation is the two-level atom interacting with Ô (I ) = ei Ĥ0 t/ Ô e−i Ĥ0 t/ , (A11)
the external electromagnetic field (such as the laser field) and dis- where Ô refers to an observable’s operator, the label (I ) indicates
sipative environment (heat bath, noise, etc). However, two-level the interaction picture with respect to Ĥ0 and absence of the label
models can also serve as a decent first-order approximation for denotes the Schrödinger picture presentation. Using the expres-
those physical phenomena whose dynamics is effectively confined sions above and Pauli matrices’ properties, we can write down
to a two-dimensional subspace, one example to be the systems for the following transformation chart between the Schrödinger and
which one can neglect the influence of excited levels above the first interaction pictures to be used in the evolution equations of the
excited one. Here, we outline the basic notions used in a theory of type (3) or (8):
two-level quantum optical systems.  
For a general two-level quantum system, the Hilbert space has ρ̂  → ρ̂(I ) , Ĥ0 , ρ̂  → 0, (A12)
Downloaded by [York University Libraries] at 13:19 13 August 2014

the dimensionality two, and it is spanned by just two states, a


1  1
ground state |g and an excited state |e. An arbitrary quantum ĤL  →  σ̂+ + σ̂− = σ̂1 , (A13)
state of such system can be written in the basis of the Pauli and unit ⎛ 2 2 ⎞
matrices which form a complete set. In quantum optics, one is often cos (ω0 t) sin (ω0 t) 0
interested in such averaged values as the population difference σ  → ⎝ − sin (ω0 t) cos (ω0 t) 0 ⎠ σ , (A14)
  ρ − ρ22 0 0 1
σ3 obs ≡ tr ρ̂ σ̂3 /tr ρ̂ = 11 , (A1)
ρ11 + ρ22 σ̂±  → e±iω0 t σ̂± , (A15)
the population of the excited-state (upper) level where by σ and σ (I ) we denote a set of the three Pauli operators
 in the Schrödinger and interaction picture, respectively.
ρ11 1
pe = = 1 + σ3 obs , (A2)
ρ11 + ρ22 2 Appendix 2. Fourier transform in open quantum
the population of the ground-state level systems
1  Let us consider the following setup: some observable, F(t), evolves
pg = 1 − pe = 1 − σ3 obs , (A3) according to quantum evolution equations. Suppose that in the
2
absence of background effects its value is trivial: F(t < 0) =
and the coherence f c = const. Then at a certain moment of time, say t = 0, one
  “switches on” the background effects, such that the total function
σ+ obs ≡ tr ρ̂ σ̂+ /tr ρ̂ , (A4)
becomes the following:
where ρi j are the i jth components of the density matrix. One
F(t) = f c θ(−t) + f (t)θ(t), (B1)
can check that during the evolution, the spin averages obey the
following identity where θ is the Heaviside step function and f (t) = F(t  0).
For practical purposes, we will be interested in the following two
σ1 2obs + σ2 2obs + σ3 2obs = 1 − 4 det(ρ̂/trρ̂)  1, (A5) scenarios:
which means that for pure states the averages lie on the Bloch
sphere σ1 2obs + σ2 2obs + σ3 2obs = 1. (a) function f (t) tends to a constant value f ∞ at t → +∞.
The unperturbed Hamiltonian of a quantum-optical two-level In this case, the Fourier transform of the global function
system is usually a linear combination of the operators |gg| and can be written as
|ee|. Up to an additive constant it equals to 1
F(ω) = √ ( f c + f ∞ ) δ(ω) + [F(ω)]reg , (B2)
1 2π
Ĥ0 = ω0 σ̂3 , (A6) where [F(ω)]reg is the regular part of the Fourier transform:
2
where ω0 is the transition frequency. If the system is put into !∞
1
contact with a monochromatic electromagnetic wave of frequency [F(ω)]reg = √ ( f (t) − f ∞ ) eiωt dt, (B3)
ω0 then in the leading order, we can restrict ourselves with the 2π
0
dipole interaction. In the rotation-wave approximation (RWA),
the corresponding Hamiltonian can be reduced to the form which is going to be the most informative for our purposes.
  (b) function f (t) oscillates with a period T .
1
ĤL =  e−iω0 t σ̂+ + eiω0 t σ̂− , (A7) In this case, the regular part of the Fourier transform of the
2 global function can be computed as
where  is the Rabi frequency which measures the strength of the !T
1
interaction of the system’s dipolemoment with the electromag- [F(n)]reg = f (t)e2πi(n/T )t dt, (B4)
netic field, and σ̂± = 12 σ̂1 ± i σ̂2 . T
For the purposes of simplifying the evolution equations, it is 0
often very convenient to perform a transition from the Schrödinger where n is an integer positive number.

You might also like