SlideShare a Scribd company logo
1 of 4
Download to read offline
Modeling decoherence in a driven two-level system
using random matrix theory
Jin Wang
Physics, Department of Natural Sciences, The University of Michigan, Dearborn, Michigan 48128, USA
(jinwang@umd.umich.edu)
Received August 9, 2011; revised September 16, 2011; accepted October 19, 2011;
posted October 20, 2011 (Doc. ID 152624); published December 9, 2011
Random matrix theory is used to model a two-level quantum system driven by a laser and coupled to a reservoir
with N degrees of freedom in both Markovian and non-Markovian regimes. Decoherence is naturally included
in this model. The effect of reservoir dimension and coupling strength between the system and reservoir is
explored. © 2011 Optical Society of America
OCIS codes: 270.2500, 000.5490.
1. INTRODUCTION
The modeling of quantum systems as they interact with the
reservoir is important to the fields of quantum information
and computing. One important phenomenon to model is deco-
herence [1,2]. Decoherence occurs when a quantum system’s
information is dissipated into its reservoir. To describe deco-
herence, the master equation model of quantum systems uses
the Lindblad superoperator [3–6] in order to phenomenologi-
cally add the decay parameters. The reduced density matrix ρs
of the master equation is solved by converting it into a set of
optical Bloch equations.
The random matrix theory (RMT) described in this article
was originally used to successfully model the complex dy-
namics of spectral fluctuations of quantum systems [7]. The
theory was then adapted for use in a number of areas, from
chemical physics to quantum systems [8–11]. Recently the
ability of the RMT combined with the Schrödinger wave func-
tion to solve a quantum system weakly coupled to its reservoir
(Markovian) has been demonstrated [10]. Two new features
that are investigated in this paper are varying the coupling
strength between the system and the environment from weak
(Markovian) to strong (non-Markovian), and study the effect
of varying the reservoir dimension N on the decoherence rate.
This article is organized as follows. In Section 2, a model of
a two-level quantum system coupled to a random reservoir
with dimension size N is solved using the Schrödinger wave
function approach. Decoherence is obtained using a partial
trace over the reservoir. In Section 3, the reservoir dimension
size N and coupling strength Esr are varied to show their ef-
fects on the magnitude and rise time of the steady-state coher-
ence. The effect of reservoir dimension N and the number of
simulations M averaged together is discussed. A discussion
of how non-Markovian behavior can be modeled using the
master equation [4] is given in the concluding section.
2. THE MODEL
The RMT can simulate quantum systems with multiple energy
levels interacting with a random reservoir. In this article, an
example of a two-level system driven by a coherent source
will be investigated. The system consists of a ground state
jgi and excited state jei coupled to a monochromatic laser
field and a random reservoir with finite dimension N set by
cavity parameters.
The full Hamiltonian is given by
H ¼ Hs ⊗ Ir þ Is ⊗ Hr þ HI; ð1Þ
where Hs corresponds to the atom-laser coupling and is
defined as
Hs ¼ EΩðjeihgj þ jgihejÞ: ð2Þ
EΩ represents the Rabi frequency that characterizes the
coupling between the atomic dipole and the laser field. The
terms Ir and Is represent the unit operators in the Hilbert
space of the reservoir and the system, respectively. The reser-
voir Hamiltonian is given by
Hr ¼ Er × R: ð3Þ
The term HI is the Hamiltonian describing the system–
reservoir interaction, which is given by
HI ¼ EsrðHe ⊗ jeihgj þ Hg ⊗ jgihejÞ: ð4Þ
The coefficients Er and Esr in Eqs. (3) and (4) represent
the strength of the reservoir and interaction Hamiltonians,
respectively.
The elements of the matrix R are chosen such that the
matrix is a Gaussian unitary ensemble (GUE) with associated
invariance properties [10]. The GUE was chosen over the
Gaussian orthogonal ensemble due to the reservoir having
complex states as will be defined later. Practically, the real
and imaginary parts of each element in the Hermitian matrix
R is chosen using complex Gaussian random numbers with
variance 0.083 and mean of zero. This variance corresponds
to the variance of a uniform distribution on the interval of −0:5
to 0.5. The Hamiltonian R is determined at the beginning and
held fixed over the time evolution of each single trajectory.
The matrix He is constructed using Gaussian random numbers
Jin Wang Vol. 29, No. 1 / January 2012 / J. Opt. Soc. Am. B 75
0740-3224/12/010075-04$15.00/0 © 2012 Optical Society of America
of variance 0.083 and mean of zero where each element has a
real and imaginary part. The matrix Hg is created by taking the
Hermitian conjugate of He, i.e., Hg ¼ H†
e. This ensures that the
interaction Hamiltonian HI is a Hermitian matrix. The idea
behind the form of interaction Hamiltonian is to set up differ-
ent correlations between the system and the reservoir depend-
ing on whether the system is transitioning from the excited to
ground state and visa versa. The square matrices Hr, He, and
Hg have dimension N. These transition operators are defined
as the lowering jeihgj and raising jgihej operators, respectively,
in Eq. (4).
The initial combined state of the system and reservoir jψ0i
is a product of the pure states of the two-level system jψs0i and
the random reservoir states jψr0i:
jψ0i ¼ jψs0i ⊗ jψr0i: ð5Þ
By keeping track of the reservoir state the system evolution is
dependent on its history, thus allowing non-Markovian sys-
tems to be simulated. The initial random state of the reservoir
jψr0i is created using Gaussian complex random numbers
with variance 0.083 and mean zero. The initial state of the
system jψs0i is the excited state jei, but it can take on any
superposition of ground and excited states.
The wave function of the combined system and reservoir
state jψti is found using unitary time evolution, and a partial
trace to remove the reservoir states yielding the decoherence
of the two-level system [12].
The standard master equation of a two-level atom driven by
a laser interacting with a reservoir is given by the following:
_ρ ¼ −i½H; ρŠ þ γD½jeiðhgjŠ: ð6Þ
Here, H is the atom–laser interaction Hamiltonian and the γ is
the decay rate. The decoherence due to the interaction with
the reservoir is given by the Lindblad form [3]. The Lindblad
superoperator D for arbitrary operators A and B is defined
as D½AŠB ≡ ABA† − 1
2 fA†A; Bg.
3. DECOHERENCE IN BOTH MARKOVIAN
AND NON-MARKOVIAN REGIONS
The amplitude of the Rabi oscillation is set to be EΩ ¼ 1 in this
article. In Fig. 1 the coupling strength is Esr ¼ 0:15, which re-
presents weak coupling between the system and the reservoir.
The coupling is classified as weak or strong with respect to
the strength of the driving field EΩ. This corresponds to a
Markovian process. The reservoir dimension is N ¼ 12. The
decoherence analysis here is based on the ensemble average
evolution of 30 individual simulations of the system coupled
with the reservoir. By averaging over 30 time evolutions the
standard deviation has 5% error relative to the steady-state
value. In each of the 30 simulations the Hamiltonian and
the reservoir state are generated as random matrices and
fixed for the simulation time. The decoherence is measured
using the excited-state population decay rate, the system en-
tropy S, and the ratio between the system’s large and small
eigenvalues.
One advantage of the RMT can be seen in the plot of the
system excited-state population ρee in the bottom of Fig. 1.
Specifically, the decoherence of the system occurs naturally
without the need for adding decoherence phenomenologically
as is the case for the master equation approach. Assuming the
damping of the system modeled using RMT is of the form
e−3γt=4
[13], the decoherence rate γ is found to be 0.06160
through curve fitting. Using this decoherence rate in the mas-
ter equation yields a close match to the random matrix ap-
proach. The difference is that the random matrix approach
contains small random fluctuations due to the randomness
in the reservoir Hamiltonian. The χ2
measure of the curve
fit between the solutions is 0.00051.
Decoherence can also be measured using the system entro-
py. The entropy of the reduced density matrix ρs of the system
is found with S ≡ −tr½ρsðlog2ðρsÞÞŠ. The entropy curve in the
bottom of Fig. 1 starts at 0 and rises to a steady-state value.
This indicates that the system starts from an initially pure
state with zero entropy, and then becomes a more mixed state
with a steady-state entropy of around 0.9 during a time scale of
a few γ−1.
In addition, decoherence can be measured inversely using
the ratio between the large and small eigenvalues of the re-
duced density of the system, as shown in the top half of Fig. 1.
Since the system has only two eigenvalues, the minimum
value with this measure is 1 and corresponds to a completely
mixed state. The maximum value of ∞ corresponds to a com-
pletely pure state. The time evolution of the ratio between the
system large and small eigenvalues yields the Schmidt path,
which gives an alternative way to measure the decoherence
of the system state over time.
One parameter that controls the decoherence rate of the
quantum system is the reservoir dimension N. In the case
Fig. 1. (Color online) Plot of the excited-state population versus time
from the random matrix model (the oscillatory line with randomness)
is on the bottom plot of the figure. The dots that overlap the oscillatory
line represent the excited-state population versus time as found using
the master equation approach. The plot of the system entropy is
shown on the bottom plot as a line starting at zero, and increasing
to a steady-state value of 0.9. The top plot shows the ratio between
the system large and small eigenvalues versus time in units of 1=γ
starting at ∞ and decaying to 0.7. The parameters for all the plots
are EΩ ¼ 1, Er ¼ 1, Esr ¼ 0:15, N ¼ 12. All of the plot lines are an
average of the results of 30 simulations.
76 J. Opt. Soc. Am. B / Vol. 29, No. 1 / January 2012 Jin Wang
of the master equation this corresponds to the number of har-
monic oscillators, in the case of the random matrix model this
corresponds to the number of elements in the reservoir model.
The steady-state entropy of a system coupled to a reservoir of
dimension N from 2 to 50 is shown in Fig. 2. The system will
decohere faster and reach a more mixed steady-state as the
dimension N increases. Also, the time required to reach the
steady state decreases as the reservoir dimension N increases,
as indicated in Fig. 3.
The decoherence behavior of the system also depends on
the system-reservoir coupling strength Esr. Up to this point
weak coupling (Esr=EΩ ¼ 0:15) has been assumed. This
causes the system to evolve under Markovian dynamics,
and is well modeled by the master equation. Now the coupling
strength is increased in order to move the system dynamics
into the non-Markovian region. As shown in Fig. 4, as the
coupling strength Esr is increased, the system exhibits nonex-
ponential decay behavior, which is an indication of non-
Markovian dynamics [14–17] . In the extreme case when
Esr=EΩ ¼ 50, the initial excited system state is frozen. This ap-
plies to any other initial system states, and is an example of
the quantum zenolike effect [18–20]. The reason the method
can model non-Markovian systems comes from the ability of
the reservoir to store information that evolves over time while
dynamically interacting with the quantum system.
Finally, the number of simulated time evolutions averaged
together in the above figures is chosen to be 30. By averaging
over 30 time evolutions, as shown in Fig. 5, the standard de-
viation of the steady-state value is around 0.022. This trans-
lates to a 5% error relative to the steady-state value of 0.5.
Additionally, the standard deviation decreases with increasing
reservoir dimension N, as seen in Fig. 6, thus allowing fewer
simulations to get the same percent error. From the results of
Figs. 2 and 6, an increasing reservoir dimension N causes both
an increase in decoherence and a decrease in the standard
deviation.
Fig. 2. (Color online) Plot of steady-state system entropy versus
reservoir dimension size N. The parameters are EΩ ¼ 1, Er ¼ 1,
Esr ¼ 0:15.
Fig. 3. (Color online) Plot of the time in units of 1=γ needed to reach
90% of the steady-state value of the entropy versus dimension N of the
reservoir. The parameters used in the plot are EΩ ¼ 1, Er ¼ 1, and
Esr ¼ 0:15.
Fig. 4. (Color online) Plot of the excited-state population versus time
in units of 1=γ for different coupling strengths. From bottom to top,
the coupling strengths are Esr ¼ 0:15, 10, 30, 50.
Fig. 5. (Color online) Plot of the system excited-state population
variance after reaching the steady state versus the number of simula-
tions for a reservoir dimension of 12. The parameters used in the plot
are Ωs ¼ 1, Er ¼ 1, Esr ¼ 0:15, N ¼ 12.
Fig. 6. (Color online) Plot of the system steady-state population
variance after averaging 10 runs versus the reservoir dimension N.
The mean steady-state value is around 0.5. The parameters used in
the plot are Ωs ¼ 1, Er ¼ 1, Esr ¼ 0:15.
Jin Wang Vol. 29, No. 1 / January 2012 / J. Opt. Soc. Am. B 77
4. DISCUSSION AND SUMMARY
Up to this point this article has compared the RMT approach
with the standard master equation in the Markovian regime.
The RMT was able to model non-Markovian processes. A few
other methods used to simulate non-Markovian include varia-
tions of unravelling of the stochastic Schrödinger equation.
Some of these methods include using doubled or trippled Hil-
bert space methods [21,22], a non-Markovian quantum state
distribution method [14,23], and a non-Markovian quantum
jump method [24,25]. Each of these methods adds the memory
effect of the non-Markovian process in different ways. How-
ever, all of these methods based on stochastic simulations
make various approximations that may reduce the ability to
use them in general quantum system cases. Another advan-
tage of RMT is that it solves the Schrödinger wave equation
without any approximations besides making the reservoir of
finite size.
One other method used by Budini involves using random
Lindblad operators to model non-Markovian systems [4]. In
the paper, the complex reservoir is split into a number of
subreservoirs each with a random decay rate in order to
observe non-Markovian effects. The ability to simulate the
non-Markovian effects comes about in this model because
the system evolution depends on the history of which sub-
reservoirs are visited on a particular simulation. From the
perspective of describing non-Markovian behavior, both ap-
proaches are similar; however, the random matrix model is
easier to implement and naturally includes decoherence.
In summary, this article has used RMT to model a quantum
system interacting with a reservoir in both the Markovian and
non-Markovian regimes. The results from RMT and the master
equation agreeing in the Markovian region. The wave function
random matrix theory exhibits non-Markovian dynamics
when the atomic system is strongly coupled to the reservoir.
The decoherence increases with reservoir dimension and
decreases with coupling strength. The standard deviation of
the population decreases as the reservoir dimension N and
number of simulations increase.
ACKNOWLEDGMENTS
The author acknowledges valuable discussions with J. Eberly,
H. Batelaan, and A. Marino.
REFERENCES
1. W. H. Zurek, “Decoherence and the transition from quantum to
classical,” Phys. Today 44, 36–44 (1991).
2. C. Kiefer, D. J. W. Giulini, J. Kupsch, I.-O. Stamatescu, E. Joos,
and H. D. Zeh, Decoherence and the Appearance of a Classical
World in Quantum Theory (Springer, 2010).
3. G. Lindblad, “On the generators of quantum dynamical semi-
groups,” Commun. Math. Phys. 48, 119–130 (1976).
4. A. A. Budini, “Random Lindblad equations from complex envir-
onments,” Phys. Rev. E 72, 056106 (2005).
5. X. L. Huang, H. Y. Sun, and X. X. Yi, “Non-markovian quantum
jump with generalized Lindblad master equation,” Phys. Rev. E
78, 041107 (2008).
6. M. Genkin, E. Waltersson, and E. Lindroth, “Estimation of the
spatial decoherence time in circular quantum dots,” Phys.
Rev. B 79, 245310 (2009).
7. T. A. Brody, J. Flores, J. B. French, P. A. Mello, A. Pandey, and
S. S. M. Wong, “Random-matrix physics: spectrum and strength
fluctuations,” Rev. Mod. Phys. 53, 385–479 (1981).
8. J. Gong and P. Brumer, “Coherent control of quantum chaotic
diffusion,” Phys. Rev. Lett. 86, 1741–1744 (2001).
9. F. Franchini and V. E. Kravtsov, “Horizon in random matrix the-
ory, the hawking radiation, and flow of cold atoms,” Phys. Rev.
Lett. 103, 166401 (2009).
10. C. Pineda, T. Gorin, and T. H. Seligman, “Decoherence of
two-qubit systems: a random matrix description,” New J. Phys.
9, 1–35 (2007).
11. M. Sadgrove, S. Wimberger, S. Parkins, and R. Leonhardt,
“Scaling law and stability for a noisy quantum system,” Phys.
Rev. E 78, 025206 (2008).
12. H. J. Carmichael, Statistical Methods in Quantum Optics
(Springer, 1996).
13. G. J. Milburn and D. F. Walls, Quantum Optics (Springer, 1994).
14. W. T. Strunz, L. Diósi, and N. Gisin, “Open system dynamics
with non-Markovian quantum trajectories,” Phys. Rev. Lett.
82, 1801–1805 (1999).
15. J. Piilo, S. Maniscalco, K. Härkönen, and K.-A. Suominen, “Non-
Markovian quantum jumps,” Phys. Rev. Lett. 100, 180402 (2008).
16. R. Vasile, S. Olivares, M. A. Paris, and S. Maniscalco,
“Continuous-variable quantum key distribution in non-
Markovian channels,” Phys. Rev. A 83, 042321 (2011).
17. K.-L. Liu and H.-S. Goan, “Non-Markovian entanglement
dynamics of quantum continuous variable systems in thermal
environments,” Phys. Rev. A 76, 022312 (2007).
18. B. Misra and E. C. G. Sudarshan, “The zeno’s paradox in quan-
tum theory,” J. Math. Phys. 18, 756–763 (1977).
19. F. M. Spedalieri, H. Lee, M. Florescu, K. T. Kapale, U. Yurtsever,
and J. P. Dowling, “Exploiting the quantum zeno effect to beat
photon loss in linear optical quantum information processors,”
Opt. Commun. 254, 374–379 (2005).
20. S. Mancini and R. Bonifacio, “Quantum zeno-like effect due to
competing decoherence mechanisms,” Phys. Rev. A 64, 042111
(2001).
21. H.-P. Breuer, B. Kappler, and F. Petruccione, “Stochastic wave-
function method for non-Markovian quantum master equations,”
Phys. Rev. A 59, 1633–1643 (1999).
22. H.-P. Breuer, “Genuine quantum trajectories for non-Markovian
processes,” Phys. Rev. A 70, 012106 (2004).
23. J. T. Stockburger and H. Grabert, “Exact c-number representa-
tion of non-Markovian quantum dissipation,” Phys. Rev. Lett. 88,
170407 (2002).
24. J. Piilo, S. Maniscalco, K. Härkönen, and K.-A. Suominen, “Non-
Markovian quantum jumps,” Phys. Rev. Lett. 100, 180402 (2008).
25. J. Piilo, K. Härkönen, S. Maniscalco, and K.-A. Suominen, “Open
system dynamics with non-Markovian quantum jumps,” Phys.
Rev. A 79, 062112 (2009).
78 J. Opt. Soc. Am. B / Vol. 29, No. 1 / January 2012 Jin Wang

More Related Content

What's hot

Tsunami Wave Simulation Models Based on Hexagonal Cellular Automata
Tsunami Wave Simulation Models Based on Hexagonal Cellular AutomataTsunami Wave Simulation Models Based on Hexagonal Cellular Automata
Tsunami Wave Simulation Models Based on Hexagonal Cellular AutomataDr.E.Syed Mohamed
 
Articulo no polares
Articulo no polaresArticulo no polares
Articulo no polareslau808
 
General Concepts in QSAR for Using the QSAR Application Toolbox Part 2
General Concepts in QSAR for Using the QSAR Application Toolbox Part 2General Concepts in QSAR for Using the QSAR Application Toolbox Part 2
General Concepts in QSAR for Using the QSAR Application Toolbox Part 2International QSAR Foundation
 
Modification factor
Modification factorModification factor
Modification factorDhruv Shah
 
Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...
Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...
Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...IJERA Editor
 
General Concepts in QSAR for Using the QSAR Application Toolbox Part 1
General Concepts in QSAR for Using the QSAR Application Toolbox Part 1General Concepts in QSAR for Using the QSAR Application Toolbox Part 1
General Concepts in QSAR for Using the QSAR Application Toolbox Part 1International QSAR Foundation
 
47rrkmtheory11 190428142321
47rrkmtheory11 19042814232147rrkmtheory11 190428142321
47rrkmtheory11 190428142321NehaDhansekar
 

What's hot (17)

Tsunami Wave Simulation Models Based on Hexagonal Cellular Automata
Tsunami Wave Simulation Models Based on Hexagonal Cellular AutomataTsunami Wave Simulation Models Based on Hexagonal Cellular Automata
Tsunami Wave Simulation Models Based on Hexagonal Cellular Automata
 
Articulo no polares
Articulo no polaresArticulo no polares
Articulo no polares
 
General Concepts in QSAR for Using the QSAR Application Toolbox Part 2
General Concepts in QSAR for Using the QSAR Application Toolbox Part 2General Concepts in QSAR for Using the QSAR Application Toolbox Part 2
General Concepts in QSAR for Using the QSAR Application Toolbox Part 2
 
Linear free energy relationships
Linear free energy relationshipsLinear free energy relationships
Linear free energy relationships
 
Modification factor
Modification factorModification factor
Modification factor
 
Metadynamics
MetadynamicsMetadynamics
Metadynamics
 
Lecture 12
Lecture 12Lecture 12
Lecture 12
 
Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...
Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...
Effects oftheExchange Interaction Parameter onthePhase Diagrams fortheTransve...
 
Lecture 6
Lecture 6Lecture 6
Lecture 6
 
Energy minimization
Energy minimizationEnergy minimization
Energy minimization
 
SpecMech
SpecMechSpecMech
SpecMech
 
On the dynamics of distillation processes
On the dynamics of distillation processesOn the dynamics of distillation processes
On the dynamics of distillation processes
 
Hinshel wood theory
Hinshel wood   theoryHinshel wood   theory
Hinshel wood theory
 
MOLECULAR MODELLING
MOLECULAR MODELLINGMOLECULAR MODELLING
MOLECULAR MODELLING
 
General Concepts in QSAR for Using the QSAR Application Toolbox Part 1
General Concepts in QSAR for Using the QSAR Application Toolbox Part 1General Concepts in QSAR for Using the QSAR Application Toolbox Part 1
General Concepts in QSAR for Using the QSAR Application Toolbox Part 1
 
47rrkmtheory11 190428142321
47rrkmtheory11 19042814232147rrkmtheory11 190428142321
47rrkmtheory11 190428142321
 
Research Project
Research ProjectResearch Project
Research Project
 

Viewers also liked

Viewers also liked (14)

Jin_single_photon(4)
Jin_single_photon(4)Jin_single_photon(4)
Jin_single_photon(4)
 
Resume
ResumeResume
Resume
 
Question 1
Question 1Question 1
Question 1
 
ATTS_Presentation_Call Tracking
ATTS_Presentation_Call TrackingATTS_Presentation_Call Tracking
ATTS_Presentation_Call Tracking
 
Question 2
Question 2Question 2
Question 2
 
nanno_actuator
nanno_actuatornanno_actuator
nanno_actuator
 
KECK STUDY
KECK STUDYKECK STUDY
KECK STUDY
 
Maria pina.tarea5.genomahumano
Maria pina.tarea5.genomahumanoMaria pina.tarea5.genomahumano
Maria pina.tarea5.genomahumano
 
Technology-Trends_2016
Technology-Trends_2016Technology-Trends_2016
Technology-Trends_2016
 
Evolucion del-periodismo-digital
Evolucion del-periodismo-digitalEvolucion del-periodismo-digital
Evolucion del-periodismo-digital
 
Reuse Permit Program June 2015
Reuse Permit Program June 2015Reuse Permit Program June 2015
Reuse Permit Program June 2015
 
interaction_free_measurement
interaction_free_measurementinteraction_free_measurement
interaction_free_measurement
 
Historia del periodismo digital
Historia del periodismo digitalHistoria del periodismo digital
Historia del periodismo digital
 
10 stills mnek at night
10 stills mnek at night10 stills mnek at night
10 stills mnek at night
 

Similar to josab-29-1-75

CHE 611 Presentation
CHE 611 PresentationCHE 611 Presentation
CHE 611 PresentationDhruv Jain
 
Computational methodologies
Computational methodologiesComputational methodologies
Computational methodologiesMattSmith321834
 
International Journal of Engineering Research and Development
International Journal of Engineering Research and DevelopmentInternational Journal of Engineering Research and Development
International Journal of Engineering Research and DevelopmentIJERD Editor
 
Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...
Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...
Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...ijctcm
 
Deal_Final_Paper_PY525f15
Deal_Final_Paper_PY525f15Deal_Final_Paper_PY525f15
Deal_Final_Paper_PY525f15Jacob Deal
 
Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...
Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...
Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...ijcsa
 
Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...
Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...
Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...IJECEIAES
 
Hopf-Bifurcation Ina Two Dimensional Nonlinear Differential Equation
Hopf-Bifurcation Ina Two Dimensional Nonlinear Differential  EquationHopf-Bifurcation Ina Two Dimensional Nonlinear Differential  Equation
Hopf-Bifurcation Ina Two Dimensional Nonlinear Differential EquationIJMER
 
Power System State Estimation - A Review
Power System State Estimation - A ReviewPower System State Estimation - A Review
Power System State Estimation - A ReviewIDES Editor
 
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...Josemar Pereira da Silva
 
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...Josemar Pereira da Silva
 
Quark Model Three Body Calculations for the Hypertriton Bound State
Quark Model Three Body Calculations for the Hypertriton Bound StateQuark Model Three Body Calculations for the Hypertriton Bound State
Quark Model Three Body Calculations for the Hypertriton Bound StateIOSR Journals
 
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...ieijjournal
 
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...ieijjournal1
 
The phase plane of moving discrete breathers
The phase plane of moving discrete breathersThe phase plane of moving discrete breathers
The phase plane of moving discrete breathersPaul Houle
 
Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...
Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...
Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...ijctcm
 
Cone crusher model identification using
Cone crusher model identification usingCone crusher model identification using
Cone crusher model identification usingijctcm
 

Similar to josab-29-1-75 (20)

CHE 611 Presentation
CHE 611 PresentationCHE 611 Presentation
CHE 611 Presentation
 
Computational methodologies
Computational methodologiesComputational methodologies
Computational methodologies
 
International Journal of Engineering Research and Development
International Journal of Engineering Research and DevelopmentInternational Journal of Engineering Research and Development
International Journal of Engineering Research and Development
 
Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...
Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...
Hybrid Chaos Synchronization of Hyperchaotic Newton-Leipnik Systems by Slidin...
 
Deal_Final_Paper_PY525f15
Deal_Final_Paper_PY525f15Deal_Final_Paper_PY525f15
Deal_Final_Paper_PY525f15
 
Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...
Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...
Sliding Mode Controller Design for Hybrid Synchronization of Hyperchaotic Che...
 
Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...
Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...
Modified Projective Synchronization of Chaotic Systems with Noise Disturbance,...
 
47032417.pdf
47032417.pdf47032417.pdf
47032417.pdf
 
Hopf-Bifurcation Ina Two Dimensional Nonlinear Differential Equation
Hopf-Bifurcation Ina Two Dimensional Nonlinear Differential  EquationHopf-Bifurcation Ina Two Dimensional Nonlinear Differential  Equation
Hopf-Bifurcation Ina Two Dimensional Nonlinear Differential Equation
 
Power System State Estimation - A Review
Power System State Estimation - A ReviewPower System State Estimation - A Review
Power System State Estimation - A Review
 
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
 
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
1 s2.0-s037838121100207 x-main.correlation of thermodynamic modeling and mole...
 
Quark Model Three Body Calculations for the Hypertriton Bound State
Quark Model Three Body Calculations for the Hypertriton Bound StateQuark Model Three Body Calculations for the Hypertriton Bound State
Quark Model Three Body Calculations for the Hypertriton Bound State
 
Three level atom OTE systems
Three level atom OTE systemsThree level atom OTE systems
Three level atom OTE systems
 
40220140501006
4022014050100640220140501006
40220140501006
 
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
 
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
THE LEFT AND RIGHT BLOCK POLE PLACEMENT COMPARISON STUDY: APPLICATION TO FLIG...
 
The phase plane of moving discrete breathers
The phase plane of moving discrete breathersThe phase plane of moving discrete breathers
The phase plane of moving discrete breathers
 
Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...
Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...
Cone Crusher Model Identification Using Block-Oriented Systems with Orthonorm...
 
Cone crusher model identification using
Cone crusher model identification usingCone crusher model identification using
Cone crusher model identification using
 

josab-29-1-75

  • 1. Modeling decoherence in a driven two-level system using random matrix theory Jin Wang Physics, Department of Natural Sciences, The University of Michigan, Dearborn, Michigan 48128, USA (jinwang@umd.umich.edu) Received August 9, 2011; revised September 16, 2011; accepted October 19, 2011; posted October 20, 2011 (Doc. ID 152624); published December 9, 2011 Random matrix theory is used to model a two-level quantum system driven by a laser and coupled to a reservoir with N degrees of freedom in both Markovian and non-Markovian regimes. Decoherence is naturally included in this model. The effect of reservoir dimension and coupling strength between the system and reservoir is explored. © 2011 Optical Society of America OCIS codes: 270.2500, 000.5490. 1. INTRODUCTION The modeling of quantum systems as they interact with the reservoir is important to the fields of quantum information and computing. One important phenomenon to model is deco- herence [1,2]. Decoherence occurs when a quantum system’s information is dissipated into its reservoir. To describe deco- herence, the master equation model of quantum systems uses the Lindblad superoperator [3–6] in order to phenomenologi- cally add the decay parameters. The reduced density matrix ρs of the master equation is solved by converting it into a set of optical Bloch equations. The random matrix theory (RMT) described in this article was originally used to successfully model the complex dy- namics of spectral fluctuations of quantum systems [7]. The theory was then adapted for use in a number of areas, from chemical physics to quantum systems [8–11]. Recently the ability of the RMT combined with the Schrödinger wave func- tion to solve a quantum system weakly coupled to its reservoir (Markovian) has been demonstrated [10]. Two new features that are investigated in this paper are varying the coupling strength between the system and the environment from weak (Markovian) to strong (non-Markovian), and study the effect of varying the reservoir dimension N on the decoherence rate. This article is organized as follows. In Section 2, a model of a two-level quantum system coupled to a random reservoir with dimension size N is solved using the Schrödinger wave function approach. Decoherence is obtained using a partial trace over the reservoir. In Section 3, the reservoir dimension size N and coupling strength Esr are varied to show their ef- fects on the magnitude and rise time of the steady-state coher- ence. The effect of reservoir dimension N and the number of simulations M averaged together is discussed. A discussion of how non-Markovian behavior can be modeled using the master equation [4] is given in the concluding section. 2. THE MODEL The RMT can simulate quantum systems with multiple energy levels interacting with a random reservoir. In this article, an example of a two-level system driven by a coherent source will be investigated. The system consists of a ground state jgi and excited state jei coupled to a monochromatic laser field and a random reservoir with finite dimension N set by cavity parameters. The full Hamiltonian is given by H ¼ Hs ⊗ Ir þ Is ⊗ Hr þ HI; ð1Þ where Hs corresponds to the atom-laser coupling and is defined as Hs ¼ EΩðjeihgj þ jgihejÞ: ð2Þ EΩ represents the Rabi frequency that characterizes the coupling between the atomic dipole and the laser field. The terms Ir and Is represent the unit operators in the Hilbert space of the reservoir and the system, respectively. The reser- voir Hamiltonian is given by Hr ¼ Er × R: ð3Þ The term HI is the Hamiltonian describing the system– reservoir interaction, which is given by HI ¼ EsrðHe ⊗ jeihgj þ Hg ⊗ jgihejÞ: ð4Þ The coefficients Er and Esr in Eqs. (3) and (4) represent the strength of the reservoir and interaction Hamiltonians, respectively. The elements of the matrix R are chosen such that the matrix is a Gaussian unitary ensemble (GUE) with associated invariance properties [10]. The GUE was chosen over the Gaussian orthogonal ensemble due to the reservoir having complex states as will be defined later. Practically, the real and imaginary parts of each element in the Hermitian matrix R is chosen using complex Gaussian random numbers with variance 0.083 and mean of zero. This variance corresponds to the variance of a uniform distribution on the interval of −0:5 to 0.5. The Hamiltonian R is determined at the beginning and held fixed over the time evolution of each single trajectory. The matrix He is constructed using Gaussian random numbers Jin Wang Vol. 29, No. 1 / January 2012 / J. Opt. Soc. Am. B 75 0740-3224/12/010075-04$15.00/0 © 2012 Optical Society of America
  • 2. of variance 0.083 and mean of zero where each element has a real and imaginary part. The matrix Hg is created by taking the Hermitian conjugate of He, i.e., Hg ¼ H† e. This ensures that the interaction Hamiltonian HI is a Hermitian matrix. The idea behind the form of interaction Hamiltonian is to set up differ- ent correlations between the system and the reservoir depend- ing on whether the system is transitioning from the excited to ground state and visa versa. The square matrices Hr, He, and Hg have dimension N. These transition operators are defined as the lowering jeihgj and raising jgihej operators, respectively, in Eq. (4). The initial combined state of the system and reservoir jψ0i is a product of the pure states of the two-level system jψs0i and the random reservoir states jψr0i: jψ0i ¼ jψs0i ⊗ jψr0i: ð5Þ By keeping track of the reservoir state the system evolution is dependent on its history, thus allowing non-Markovian sys- tems to be simulated. The initial random state of the reservoir jψr0i is created using Gaussian complex random numbers with variance 0.083 and mean zero. The initial state of the system jψs0i is the excited state jei, but it can take on any superposition of ground and excited states. The wave function of the combined system and reservoir state jψti is found using unitary time evolution, and a partial trace to remove the reservoir states yielding the decoherence of the two-level system [12]. The standard master equation of a two-level atom driven by a laser interacting with a reservoir is given by the following: _ρ ¼ −i½H; ρŠ þ γD½jeiðhgjŠ: ð6Þ Here, H is the atom–laser interaction Hamiltonian and the γ is the decay rate. The decoherence due to the interaction with the reservoir is given by the Lindblad form [3]. The Lindblad superoperator D for arbitrary operators A and B is defined as D½AŠB ≡ ABA† − 1 2 fA†A; Bg. 3. DECOHERENCE IN BOTH MARKOVIAN AND NON-MARKOVIAN REGIONS The amplitude of the Rabi oscillation is set to be EΩ ¼ 1 in this article. In Fig. 1 the coupling strength is Esr ¼ 0:15, which re- presents weak coupling between the system and the reservoir. The coupling is classified as weak or strong with respect to the strength of the driving field EΩ. This corresponds to a Markovian process. The reservoir dimension is N ¼ 12. The decoherence analysis here is based on the ensemble average evolution of 30 individual simulations of the system coupled with the reservoir. By averaging over 30 time evolutions the standard deviation has 5% error relative to the steady-state value. In each of the 30 simulations the Hamiltonian and the reservoir state are generated as random matrices and fixed for the simulation time. The decoherence is measured using the excited-state population decay rate, the system en- tropy S, and the ratio between the system’s large and small eigenvalues. One advantage of the RMT can be seen in the plot of the system excited-state population ρee in the bottom of Fig. 1. Specifically, the decoherence of the system occurs naturally without the need for adding decoherence phenomenologically as is the case for the master equation approach. Assuming the damping of the system modeled using RMT is of the form e−3γt=4 [13], the decoherence rate γ is found to be 0.06160 through curve fitting. Using this decoherence rate in the mas- ter equation yields a close match to the random matrix ap- proach. The difference is that the random matrix approach contains small random fluctuations due to the randomness in the reservoir Hamiltonian. The χ2 measure of the curve fit between the solutions is 0.00051. Decoherence can also be measured using the system entro- py. The entropy of the reduced density matrix ρs of the system is found with S ≡ −tr½ρsðlog2ðρsÞÞŠ. The entropy curve in the bottom of Fig. 1 starts at 0 and rises to a steady-state value. This indicates that the system starts from an initially pure state with zero entropy, and then becomes a more mixed state with a steady-state entropy of around 0.9 during a time scale of a few γ−1. In addition, decoherence can be measured inversely using the ratio between the large and small eigenvalues of the re- duced density of the system, as shown in the top half of Fig. 1. Since the system has only two eigenvalues, the minimum value with this measure is 1 and corresponds to a completely mixed state. The maximum value of ∞ corresponds to a com- pletely pure state. The time evolution of the ratio between the system large and small eigenvalues yields the Schmidt path, which gives an alternative way to measure the decoherence of the system state over time. One parameter that controls the decoherence rate of the quantum system is the reservoir dimension N. In the case Fig. 1. (Color online) Plot of the excited-state population versus time from the random matrix model (the oscillatory line with randomness) is on the bottom plot of the figure. The dots that overlap the oscillatory line represent the excited-state population versus time as found using the master equation approach. The plot of the system entropy is shown on the bottom plot as a line starting at zero, and increasing to a steady-state value of 0.9. The top plot shows the ratio between the system large and small eigenvalues versus time in units of 1=γ starting at ∞ and decaying to 0.7. The parameters for all the plots are EΩ ¼ 1, Er ¼ 1, Esr ¼ 0:15, N ¼ 12. All of the plot lines are an average of the results of 30 simulations. 76 J. Opt. Soc. Am. B / Vol. 29, No. 1 / January 2012 Jin Wang
  • 3. of the master equation this corresponds to the number of har- monic oscillators, in the case of the random matrix model this corresponds to the number of elements in the reservoir model. The steady-state entropy of a system coupled to a reservoir of dimension N from 2 to 50 is shown in Fig. 2. The system will decohere faster and reach a more mixed steady-state as the dimension N increases. Also, the time required to reach the steady state decreases as the reservoir dimension N increases, as indicated in Fig. 3. The decoherence behavior of the system also depends on the system-reservoir coupling strength Esr. Up to this point weak coupling (Esr=EΩ ¼ 0:15) has been assumed. This causes the system to evolve under Markovian dynamics, and is well modeled by the master equation. Now the coupling strength is increased in order to move the system dynamics into the non-Markovian region. As shown in Fig. 4, as the coupling strength Esr is increased, the system exhibits nonex- ponential decay behavior, which is an indication of non- Markovian dynamics [14–17] . In the extreme case when Esr=EΩ ¼ 50, the initial excited system state is frozen. This ap- plies to any other initial system states, and is an example of the quantum zenolike effect [18–20]. The reason the method can model non-Markovian systems comes from the ability of the reservoir to store information that evolves over time while dynamically interacting with the quantum system. Finally, the number of simulated time evolutions averaged together in the above figures is chosen to be 30. By averaging over 30 time evolutions, as shown in Fig. 5, the standard de- viation of the steady-state value is around 0.022. This trans- lates to a 5% error relative to the steady-state value of 0.5. Additionally, the standard deviation decreases with increasing reservoir dimension N, as seen in Fig. 6, thus allowing fewer simulations to get the same percent error. From the results of Figs. 2 and 6, an increasing reservoir dimension N causes both an increase in decoherence and a decrease in the standard deviation. Fig. 2. (Color online) Plot of steady-state system entropy versus reservoir dimension size N. The parameters are EΩ ¼ 1, Er ¼ 1, Esr ¼ 0:15. Fig. 3. (Color online) Plot of the time in units of 1=γ needed to reach 90% of the steady-state value of the entropy versus dimension N of the reservoir. The parameters used in the plot are EΩ ¼ 1, Er ¼ 1, and Esr ¼ 0:15. Fig. 4. (Color online) Plot of the excited-state population versus time in units of 1=γ for different coupling strengths. From bottom to top, the coupling strengths are Esr ¼ 0:15, 10, 30, 50. Fig. 5. (Color online) Plot of the system excited-state population variance after reaching the steady state versus the number of simula- tions for a reservoir dimension of 12. The parameters used in the plot are Ωs ¼ 1, Er ¼ 1, Esr ¼ 0:15, N ¼ 12. Fig. 6. (Color online) Plot of the system steady-state population variance after averaging 10 runs versus the reservoir dimension N. The mean steady-state value is around 0.5. The parameters used in the plot are Ωs ¼ 1, Er ¼ 1, Esr ¼ 0:15. Jin Wang Vol. 29, No. 1 / January 2012 / J. Opt. Soc. Am. B 77
  • 4. 4. DISCUSSION AND SUMMARY Up to this point this article has compared the RMT approach with the standard master equation in the Markovian regime. The RMT was able to model non-Markovian processes. A few other methods used to simulate non-Markovian include varia- tions of unravelling of the stochastic Schrödinger equation. Some of these methods include using doubled or trippled Hil- bert space methods [21,22], a non-Markovian quantum state distribution method [14,23], and a non-Markovian quantum jump method [24,25]. Each of these methods adds the memory effect of the non-Markovian process in different ways. How- ever, all of these methods based on stochastic simulations make various approximations that may reduce the ability to use them in general quantum system cases. Another advan- tage of RMT is that it solves the Schrödinger wave equation without any approximations besides making the reservoir of finite size. One other method used by Budini involves using random Lindblad operators to model non-Markovian systems [4]. In the paper, the complex reservoir is split into a number of subreservoirs each with a random decay rate in order to observe non-Markovian effects. The ability to simulate the non-Markovian effects comes about in this model because the system evolution depends on the history of which sub- reservoirs are visited on a particular simulation. From the perspective of describing non-Markovian behavior, both ap- proaches are similar; however, the random matrix model is easier to implement and naturally includes decoherence. In summary, this article has used RMT to model a quantum system interacting with a reservoir in both the Markovian and non-Markovian regimes. The results from RMT and the master equation agreeing in the Markovian region. The wave function random matrix theory exhibits non-Markovian dynamics when the atomic system is strongly coupled to the reservoir. The decoherence increases with reservoir dimension and decreases with coupling strength. The standard deviation of the population decreases as the reservoir dimension N and number of simulations increase. ACKNOWLEDGMENTS The author acknowledges valuable discussions with J. Eberly, H. Batelaan, and A. Marino. REFERENCES 1. W. H. Zurek, “Decoherence and the transition from quantum to classical,” Phys. Today 44, 36–44 (1991). 2. C. Kiefer, D. J. W. Giulini, J. Kupsch, I.-O. Stamatescu, E. Joos, and H. D. Zeh, Decoherence and the Appearance of a Classical World in Quantum Theory (Springer, 2010). 3. G. Lindblad, “On the generators of quantum dynamical semi- groups,” Commun. Math. Phys. 48, 119–130 (1976). 4. A. A. Budini, “Random Lindblad equations from complex envir- onments,” Phys. Rev. E 72, 056106 (2005). 5. X. L. Huang, H. Y. Sun, and X. X. Yi, “Non-markovian quantum jump with generalized Lindblad master equation,” Phys. Rev. E 78, 041107 (2008). 6. M. Genkin, E. Waltersson, and E. Lindroth, “Estimation of the spatial decoherence time in circular quantum dots,” Phys. Rev. B 79, 245310 (2009). 7. T. A. Brody, J. Flores, J. B. French, P. A. Mello, A. Pandey, and S. S. M. Wong, “Random-matrix physics: spectrum and strength fluctuations,” Rev. Mod. Phys. 53, 385–479 (1981). 8. J. Gong and P. Brumer, “Coherent control of quantum chaotic diffusion,” Phys. Rev. Lett. 86, 1741–1744 (2001). 9. F. Franchini and V. E. Kravtsov, “Horizon in random matrix the- ory, the hawking radiation, and flow of cold atoms,” Phys. Rev. Lett. 103, 166401 (2009). 10. C. Pineda, T. Gorin, and T. H. Seligman, “Decoherence of two-qubit systems: a random matrix description,” New J. Phys. 9, 1–35 (2007). 11. M. Sadgrove, S. Wimberger, S. Parkins, and R. Leonhardt, “Scaling law and stability for a noisy quantum system,” Phys. Rev. E 78, 025206 (2008). 12. H. J. Carmichael, Statistical Methods in Quantum Optics (Springer, 1996). 13. G. J. Milburn and D. F. Walls, Quantum Optics (Springer, 1994). 14. W. T. Strunz, L. Diósi, and N. Gisin, “Open system dynamics with non-Markovian quantum trajectories,” Phys. Rev. Lett. 82, 1801–1805 (1999). 15. J. Piilo, S. Maniscalco, K. Härkönen, and K.-A. Suominen, “Non- Markovian quantum jumps,” Phys. Rev. Lett. 100, 180402 (2008). 16. R. Vasile, S. Olivares, M. A. Paris, and S. Maniscalco, “Continuous-variable quantum key distribution in non- Markovian channels,” Phys. Rev. A 83, 042321 (2011). 17. K.-L. Liu and H.-S. Goan, “Non-Markovian entanglement dynamics of quantum continuous variable systems in thermal environments,” Phys. Rev. A 76, 022312 (2007). 18. B. Misra and E. C. G. Sudarshan, “The zeno’s paradox in quan- tum theory,” J. Math. Phys. 18, 756–763 (1977). 19. F. M. Spedalieri, H. Lee, M. Florescu, K. T. Kapale, U. Yurtsever, and J. P. Dowling, “Exploiting the quantum zeno effect to beat photon loss in linear optical quantum information processors,” Opt. Commun. 254, 374–379 (2005). 20. S. Mancini and R. Bonifacio, “Quantum zeno-like effect due to competing decoherence mechanisms,” Phys. Rev. A 64, 042111 (2001). 21. H.-P. Breuer, B. Kappler, and F. Petruccione, “Stochastic wave- function method for non-Markovian quantum master equations,” Phys. Rev. A 59, 1633–1643 (1999). 22. H.-P. Breuer, “Genuine quantum trajectories for non-Markovian processes,” Phys. Rev. A 70, 012106 (2004). 23. J. T. Stockburger and H. Grabert, “Exact c-number representa- tion of non-Markovian quantum dissipation,” Phys. Rev. Lett. 88, 170407 (2002). 24. J. Piilo, S. Maniscalco, K. Härkönen, and K.-A. Suominen, “Non- Markovian quantum jumps,” Phys. Rev. Lett. 100, 180402 (2008). 25. J. Piilo, K. Härkönen, S. Maniscalco, and K.-A. Suominen, “Open system dynamics with non-Markovian quantum jumps,” Phys. Rev. A 79, 062112 (2009). 78 J. Opt. Soc. Am. B / Vol. 29, No. 1 / January 2012 Jin Wang