Download Spontaneous emission of an excited two

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

T-symmetry wikipedia , lookup

Coherent states wikipedia , lookup

Density matrix wikipedia , lookup

Copenhagen interpretation wikipedia , lookup

Bohr–Einstein debates wikipedia , lookup

Propagator wikipedia , lookup

Perturbation theory wikipedia , lookup

Hidden variable theory wikipedia , lookup

Double-slit experiment wikipedia , lookup

Quantum state wikipedia , lookup

Scalar field theory wikipedia , lookup

X-ray fluorescence wikipedia , lookup

History of quantum field theory wikipedia , lookup

Matter wave wikipedia , lookup

Probability amplitude wikipedia , lookup

Bohr model wikipedia , lookup

Quantum electrodynamics wikipedia , lookup

Path integral formulation wikipedia , lookup

Symmetry in quantum mechanics wikipedia , lookup

Coupled cluster wikipedia , lookup

Canonical quantization wikipedia , lookup

Electron configuration wikipedia , lookup

Schrödinger equation wikipedia , lookup

Atomic orbital wikipedia , lookup

Renormalization group wikipedia , lookup

Molecular Hamiltonian wikipedia , lookup

Dirac equation wikipedia , lookup

Wave–particle duality wikipedia , lookup

Wave function wikipedia , lookup

Tight binding wikipedia , lookup

Atomic theory wikipedia , lookup

Relativistic quantum mechanics wikipedia , lookup

Hydrogen atom wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Transcript
Eur. Phys. J. D 23, 279–284 (2003)
DOI: 10.1140/epjd/e2003-00033-9
THE EUROPEAN
PHYSICAL JOURNAL D
Spontaneous emission of an excited two-level atom
without both rotating-wave and Markovian approximation
C.-Q. Cao1,a , C.-G. Yu1 , and H. Cao2
1
2
Department of Physics, Peking University, Beijing 100871, P.R. China
Department of Physics and Astronomy, Northwestern University, Evanston, IL 60208-3112, USA
Received 10 July 2002 / Received in final form 12 November 2002
c EDP Sciences, Società Italiana di Fisica, Springer-Verlag 2003
Published online 4 February 2003 – Abstract. The spontaneous emission of an excited atom is analyzed by quantum stochastic trajectory
approach without both rotating-wave approximation and Markovian approximation. The atom finite size
effect is also taken into account. We show by an example that the correction due to the counter-rotating
wave term is rather small, even for the largest atomic number of real nuclei.
PACS. 42.50.Ct Quantum description of interaction of light and matter; related experiments –
42.50.Lc Quantum fluctuations, quantum noise, and quantum jumps
1 Introduction
The spontaneous emission of an excited atom is an old
question of quantum theory. The first important progress
in this respect is the proposition and evaluation of the
Einstein A coefficient γA , which gives the spontaneous
emission rate. It is well known that γA is proportional
to d2 , where d denotes the electric-dipole transition moment. The next important progress is the formulation of
Weisskopf-Wigner theory. It takes γA as the percentage
decay rate of the upper-level population N2 during the
whole emission process, leading to an exponential decay
of N2 and consequently a Lorentzian line profile in the
atomic spectrum. This treatment actually contains two
approximations. One is the neglect of the finite size effect
on the electric-dipole transition along with the omission of
the contribution from all possible higher multipole transitions, caused by the replacement of the plane wave factor eik·x by 1 in the transition matrix element evaluation.
Another is the so-called Markovian approximation on the
reduced equations of the atomic variables derived by eliminating the photon variables. They are differential-integral
equations with correlation functions to denote the memory
effect. The Markovian approximation neglects this memory effect and hence changes them to simple differential
equations.
The Weisskopf-Wigner theory applies well for the emission of outer-shell electron or inner-shell electron of light
nuclei. But in the case of inner-shell electron emission of
heavy nuclei, it may lead to obvious deviation, since (1) ω0
2
is proportional to Zeff
(where Zeff e is the charge of the
atomic core felt by the emitting electron), and the radius a
of the electron cloud is proportional to 1/Zeff , leading to
a
e-mail: [email protected]
a linear increase of k0 a = ω0 a/c with Zeff . This makes
the pointlike electric-dipole approximation (neglecting the
factor eik·x in the integral) poorer for large Zeff . (2) During the emission process, the frequency of the emitted
photon does not need to be ω0 (such photon is the socalled virtual photon) and the finite size correction will
become larger for large ω. As to the Markov approximation, in the usual estimation, the characteristic correlation time is of order 2π/ω0 , while the decay time of
the atomic variable is of order 1/γA . Hence, the validity of the Markovian approximation requires 2π/ω0 much
smaller than 1/γA , namely 2πγA /ω0 1. It can be shown
2
, hence large Zeff disadvanthat γA /ω0 increases as Zeff
tages the Markovian approximation.
Later on, some people considered the correction to the
Weisskopf-Wigner theory [1–4]. But, as far as we know,
all of them applied the Laplace transform to solve the resultant differential-integral equation, and various kinds of
approximations were used in the inverse transform. Some
papers even take the radiating atom as a pointlike electric
dipole, so that the corresponding correlation spectrum diverges linearly with ω, and in the subsequent treatment an
artificial cutoff of frequency is needed. Besides, majority
of them only considers the case of Z = 1.
In a preceding paper [5], two of us with other coworkers restudied this problem by a totally different approach.
The stochastic quantum trajectory formulation [6] is applied for this investigation, and the non-Markovian correction to the decay of upper-level population is taken
into account by introduction of additional fictitious oscillators [7,8]. However, the counter-rotating wave interaction is still not taken into account. Now we will release
this limitation to include the counter-rotating wave term
in the interaction, since its effect also becomes larger for
stronger coupling [9]. In this case there are two correlation
280
The European Physical Journal D
spectra, hence more fictitious oscillators are needed to simulate them. Moreover, in the case of the rotating-wave approximation, the total quanta of the fictitious oscillators
is either one or zero, while the counter-rotating wave interaction puts no limitation on the quanta numbers. Both
of these will increase the amount of numerical calculation.
In Section 2 the relevant differential-integral equation
for the atom operators is given, with the two correlation
spectra explicitly derived. The same hydrogen-like example as that of reference [5] is numerically studied for comparison. In Section 3 the quantum stochastic trajectory
analysis [6] of the enlarged system [7,8] is carried out
to give the time decay of the upper-level population. We
see by this example that the correction due to counterrotating wave term is rather small (except the end stage of
decay) even for the largest atomic number Z of hydrogenlike atom, as conjectured by the usual estimation.
2 The spectra of correlation functions
for spontaneous emission of a two-level atom
The A · P type Ĥint for the spontaneous emission of a
two-level atom is known as
Ĥint (t) = i
k,j
gkj σ̂+ (t)âkj (t) −
field as reservoir, the resultant differential-integral equations are given by
d
σ̂− (t) =
dt
t
[u1 (t − t ) − ū1 (t − t )]σ̂3 (t)σ̂− (t )dt
0
t
+ e2iω0 t [u2 (t − t ) − ū2 (t − t )]σ̂3 (t)σ̂+ (t )dt
0
− σ̂3 (t) Σ̂1 (t) + Σ̂2† (t)
(3a)
t
d
σ̂3 (t) = −2
[u1 (t − t ) − ū1 (t − t )]σ̂+ (t)σ̂− (t )dt
dt
0
t
− e2iω0 t [u2 (t − t ) − ū2 (t − t )]σ̂+ (t)σ̂+ (t )dt
0
+ 2σ̂+ (t) Σ̂1 (t) + Σ̂2† (t) + h.c.
(3b)
where σ̂3 is the atom population-difference operator
and Σ̂j (t)’s are fluctuation operators defined in equation (5). There are totally four correlation functions in
the above differential-integral equations:
u1 (t − t ) =
|gkj |2 e−i(ω−ω0 )(t−t )
kj
∞
∗
gkj
σ̂− (t)â†kj (t)
∗
σ̂− (t)âkj (t)
+ḡkj σ̂+ (t)aˆ† kj (t) − ḡkj
≡
R1 (ω)e−i(ω−ω0 )(t−t ) dω,
(1)
ū1 (t − t ) =
|ḡkj |2 ei(ω+ω0 )(t−t )
kj
∞
with counter-rotating wave terms included, where âkj
(â†kj ) is the photon annihilation (creation) operator of
mode (k, j), σ̂± are atom-level change operators (σ̂+ corresponds to upward change and σ̂− downward change), gkj
and ḡkj are coupling constants for rotating-wave term and
counter-rotating wave term respectively,
gkj
e
=−
m
2π
εkj · Gk ,
V kc
eik·x Ψ2† (x)∇Ψ1 (x)d3 x,
2π
e
εkj · Ḡk ,
=−
m
V
kc
(2a)
≡
Ḡk =
R1 (ω)ei(ω+ω0 )(t−t ) dω,
(4b)
0
u2 (t − t ) =
gkj ḡkj ei(ω−ω0 )(t−t )
kj
∞
≡
ū2 (t − t ) =
0
≡
R2 (ω)ei(ω−ω0 )(t−t ) dω,
(4c)
gkj ḡkj e−i(ω+ω0 )(t−t )
kj
∞
Gk =
ḡkj
(4a)
0
R2 (ω)e−i(ω+ω0 )(t−t ) dω,
(4d)
0
(2b)
e−ik·x Ψ2† (x)∇Ψ1 (x)d3 x
in which εkj is the polarization vector of photon
mode (k, j), Ψ2 (x) and Ψ1 (x) are the upper level and
lower level wave function respectively. They are twocomponent spinor functions. It is seen from equation (2)
that when Ψ2 (x) and Ψ1 (x) have the same parity, Ḡk =
−Gk ; when Ψ2 (x) and Ψ1 (x) have the opposite parity,
Ḡk = Gk ; and similar relations for gkj and ḡkj .
The dynamical equations for both atom and photon
operators are readily deduced from Ĥint . Eliminate the operators âkj and aˆ† kj by taking the electromagnetic (e.m.)
since |gkj |2 = |ḡkj |2 . The correlation spectrum R1 (ω) is
expressed by
R1 (ω) =
V ω2
(2π)3 c3
=
e ω
4π 2 m2 c3
2
dΩk
|gkj |2
j
dΩk |Gk |2 − |nk · Gk |2 . (4e)
Similar expression for R2 (ω) can be written out directly.
The nk in equation (4e) is the unit vector in the direction
of k. We see that there are two spectral functions R1 (ω)
and R2 (ω), not just only one spectral function as in the
references [1–5].
C.-Q. Cao et al.: Spontaneous emission of an excited two-level atom
The two quantum fluctuation forces are expressed by
Σ̂1 (t) =
gkj âkj (0)e−i(ω−ω0 )t ,
(5a)
kj
Σ̂2† (t)
=
kj
ḡkj â†kj (0)ei(ω+ω0 )t .
(5b)
As a first step we calculate the coupling constant Ḡk
and Gk . For simplicity, we omit the spin of the electron.
Substituting the wave function of hydrogen-like atom
Ψ1 (x) = f1 (r)Yl1 m1 (θ ϕ), Ψ2 (x) = f2 (r)Yl2 m2 (θ ϕ) (6)
and also the expansion formula of plane wave by spherical
waves into equation (2a) and making use of the formula
of ∇Ψ1 (x), one gets Gk as in [5]
4π(2l2 + 1) il
m2
√
Gk = (−1)
Ylm (θk ϕk )
2l1 + 1
2l + 1
lmµ
×[− (l1 + 1)(2l1 + 3){Al (ω) − l1 Bl (ω)}
×C(l1 + 1, 1, l1 ; m1 − µ, µ, m1 )
×C(l1 + 1, l2 , l; m1 − µ, −m2 , m)
×C(l1 + 1, l2 , l; 0, 0, 0) + l1 (2l1 − 1){Al (ω)
+(l1 + 1)Bl (ω)}C(l1 − 1, 1, l1 ; m1 − µ, µ, m1 )
×C(l1 − 1, l2 , l; m1 − µ, −m2 , m)
×C(l1 − 1, l2 , l; 0, 0, 0)]nµ
(7)
in which C(l1 + 1, 1, l1 ; m1 − µ, µ, m1 ) etc are ClebaschGordan coefficients, nµ (µ = +1, 0, −1) are spherical bases,
1
n+1 = √ (n1 + in2 ),
2
n0 = n3 ,
1
n−1 = √ (n1 − in2 )
2
and
ω df1 (r)
r f2 (r)
dr,
c
dr
0
∞
ω r f2 (r)f1 (r)dr.
Bl (ω) =
rjl
c
0
Al (ω) =
∞
r2 jl
(9a)
(9b)
The summation in equation (7) actually contains only finite terms because of the angular momentum addition
rule. For the first term in the square brackets, l is restricted in the range |l2 − l1 − 1| ≤ l ≤ l2 + l1 + 1, and for
the second term, in the range |l2 − l1 + 1| ≤ l ≤ l2 + l1 − 1.
The Ḡk can be obtained from equation (7) simply by
the substitution
Ylm (θk ϕk ) → (−1)l Ylm (θk ϕk ).
Substituting Gk and the formula
µ=1
4π Y1µ (θk ϕk )n∗µ
nk =
3 µ=−1
into equation (4e) and carrying out the integration over θk
and ϕk , one will obtain the first spectral function R1 (ω).
The spectrum R2 (ω) can be calculated similarly.
To see explicitly these spectral functions, we consider the same example of hydrogen-like atom studied in
reference [5]: Ψ1 (x) and Ψ2 (x) carry quantum numbers
(n1 = 1, l1 = 0) and (n2 = 2, l2 = 1, m2 = 1) respectively
with the expression
4
1
−r/a1
Ψ1 (x) = √ N1 e
,
N1 =
,
(12)
a31
4π
4
−r/a2
Y11 (θ ϕ), N2 =
·
(13)
Ψ2 (x) = N2 re
3a52
In this simple case, the gradient of Ψ1 (x) may be evaluated
directly. Substituting it into equation (2a) and carrying
out the angular integration, one obtains [5]
4π N1 N2
X0 (ω)Y00 (θk ϕk )n−1
Gk =
3 a1
6
Y2,−2 (θk ϕk )n+1
+ X2 (ω)
5
3
Y2,−1 (θk ϕk )n0
−
5
1
Y2,0 (θk ϕk )n−1
+
(14)
5
where
2 2
ω 2a4 3 − ωc2a
r e−r/a dr = X0 (ω) =
r3 j0
,
2 2 3
c
0
1 + ωc2a
(15)
∞
ω 4 ω 2 a2
8a c2
r e−r/a dr = X2 (ω) =
r3 j2
(16)
2 2 3
c
0
1 + ωc2a
(8)
(10)
with
∞
1
1
1
=
+ ·
a
a1
a2
(17)
In this simple example, both Al (ω) and Bl (ω) are proportional to Xl (ω).
In equation (14) the angular momentum l only takes
the value 0 and 2, hence we know from equation (10) that
in this example Ḡk = Gk , consistent with what we say
below equation (2b). This is in turn implies the equality of
the counter-rotating wave coupling constant ḡkj and the
rotating-wave coupling constant gkj . Thus the correlation
spectrum R2 (ω) is given by
V ω2
R2 (ω) =
(2π)3 c3
(11)
281
e2 ω
=
4π 2 m2 c2
dΩk
g 2kj
j
dΩk G2k − (nk · Gk )2 . (18)
282
The European Physical Journal D
It is easy to show that dΩk G2k = 0. By further utilizing
the coupling rulefor spherical harmonics, nk · Gk may
3
be
expressed as2 l=1 cl Yl,−1 (θk ϕk ), hence we also have
dΩk (nk · Gk ) = 0. Consequently we obtain
R2 (ω) = 0.
(19)
This may be a quite general result [10].
From equations (4e, 11, 14), one gets the result
for R1 (ω) as R(ω) used in reference [5] and also in references [3,4]
R1 (ω) =
ω
γA
,
2πω0 (1 + ω 2 a2 /c2 )4
(20)
in which γA is the Einstein A coefficient, given by γA =
4ω03 |d21 |2 /3c3 and in our example the transition dipole
moment d21 takes its absolute value as
√
64 2
|d21 | =
ea,
(21)
81
with e denoting the magnitude of electron charge.
Actually there is only one parameter in the R1 (ω),
since ω0 may be expressed by a as
ω0 =
1
(E2 − E1 ) =
,
6ma2
(22)
and the two atomic radii a1 and a2 are related to a by
multiplying constants 3/2 and 3 respectively. Since a is
proportional to the atomic number 1/Z, we see from equation (22) that ω0 is proportional to Z 2 .
The relation between γA /ω0 and a is obtained from
equations (21, 22), and may be expressed by
γA
8
=
ω0
3
32
243
2 e2
c
3
a2B
,
a2
64
81
2 e2
c
(23a)
3
Z 2,
(23b)
hence the value γA /ω0 = 10−3 corresponds to Z = 157,
somewhat larger than the upper limit of the real nuclei.
A profile of R1 (ω) is shown in Figure 1 of reference [5],
in which the dimensionless parameter γA /ω0 (instead of a)
in R(ω) is taken as 0.001.
In case the atom is taken as a pointlike particle, the
factor eik·x in the integrand of Gk is omitted, Gk is then
independent of k. Hence, R1 (ω) will become linear in ω
as can be seen from equation (4e), and the coefficient
is just γA /2πω0 . We will call this spectrum as pointlike
dipole spectrum:
R1D (ω) =
γA
ω.
2πω0
∞
1
R1 (ω)
C = πR1 (ω0 ) − i
℘
dω ,
2
ω − ω0
0
∞
1
R1 (ω)
C̄ = i
dω ,
2
ω
+ ω0
0
(25a)
(25b)
where ℘ denotes taking the principal value. The imaginary
parts of C nd C̄ contribute just an level-shift which will
be canceled out in equation (3b) because of the h.c. term.
Taking the expectation value of equation (3b) and noting the upper level population N̂2 (t) = (1 + σ̂3 (t))/2,
one gets immediately
d N̂2 (t) = γ N̂2 (t)
dt
(26a)
with
where aB is the Bohr radius. The dependence of γA /ω0
on the atomic number Z is given by
γA
1
=
ω0
6
Actually in our example only electric dipole emission is
allowed by the selection rule of angular momentum, thus
the factor (1 + ω0 a/c)−4 in R1 (ω) just expresses the finite
size effect on the radiative electric dipole transition.
In this example, both u2 (t − t ) and ū2 (t − t ) are
equal to zero, only two correlation functions u1 (t − t ) and
ū1 (t − t ) remain. The former comes from the rotatingwave interaction and the latter from the counter-rotating
wave interaction. In case we make the Markovian approximation on them, namely approximate them by Cδ(t − t )
and C̄δ(t − t ) respectively to neglect the memory effect,
the constants C and C̄ will be determined by integration
of equations (4a, 4b) respect to t from t to ∞ (note that
u1 and ū1 are defined only in the region t > t ). The results
so obtained is
(24)
It also corresponds to R1 (ω) with ωa/c = 0 in the denominator.
γ = 2πR1 (ω0 ) =
γA
·
(1 + ω02 a2 /c2 )4
(26b)
We see that γ is different from γA by the finite size correction factor (1 + ω02 a2 /c2 )−4 , which in our example is
about 0.729.
We would like to point out that the counter-rotating
d
wave interaction will have no effect on dt
N̂ (t) in the case
Markovian approximation is made, even if R2 (ω) = 0.
In the next section we will study the non-Markovian
approach with the counter-rotating term included.
3 Quantum stochastic trajectory approach
to the spontaneous emission
We have shown by equation (20) that the two-level
atom with large atomic number Z has an evident nonMarkovian reservoir in its spontaneous emission. Hence
we will, as described in references [7,8], introduce N
additional fictitious harmonic oscillators, which interact
with the atom in the same form as photons, to form
an expanded system. This expanded system thus has the
C.-Q. Cao et al.: Spontaneous emission of an excited two-level atom
Hamiltonian
Ĥ =
N
N 1
ωj α̂†j α̂j + i
gj σˆ+ αˆj
ω0 σˆ3 +
2
j=1
j=1
− gj∗ σˆ− αˆj † + ḡj σˆ+ αˆj † − ḡj∗ σˆ− αˆj
where Ĥ is expressed by equation (27). The collapse operators acting on |Ψ (t) are
j = 1, 2, ..., N.
(30)
Ĉj = Γj ∆tα̂j ,
(27)
in which the energies of the two atom-levels are taken
as ± 21 ω0 respectively, α̂j and α̂†j are annihilation and
creation operators of the jth oscillator, ḡj = ±gj for the
relative parity of Ψ2 and Ψ1 equal to ∓1 respectively. Each
of these oscillators, on the other hand, is assumed to interact with its own reservoir with no memory, hence the
expanded system is Markovian. In terms of Langevin equations in Heisenberg picture the dynamics of the above system is expressed by
N d
gj αˆj (t)
σ̂− (t) = −iω0 σ̂− (t) −
dt
j=1
+ ḡj α̂†j (t) σ̂3 (t),
(28a)
N d
gj αˆj (t) + ḡj α̂†j (t) σ̂+ (t)
σ̂3 (t) = 2
dt
j=1
+2
N gj∗ αˆj † (t) + ḡj∗ α̂j (t) σ̂− (t), (28b)
j=1
d
α̂j (t) = −iωj α̂j (t) − gj∗ σ̂− (t) + ḡj σ̂+ (t)
dt
1
− Γj α̂j (t) − F̂j (t)
2
with
283
(28c)
F̂j (t)
= 0,
F̂i (t)F̂j (t )
= 0,
R
R
F̂i (t)F̂j† (t )
= δij Γj δ(t − t ).
(28d)
R
The last two terms of equation (28c) are the usual
Markovian dissipation term and fluctuation term contributed by the reservoir of jth oscillator.
When the formal solution of equation (28c) is used to
eliminate the variables α̂j and α̂†j in equations (28a, 28b),
the reduced equation for σ̂+ (t) and σ̂3 (t) will have nonMarkovian damping terms and non-Markovian fluctuation
terms, since each fictitious oscillator contributes a term
with Lorentzian type spectrum. The next step is to select parameters gj , ḡj , ωj and Γj to simulate the original damping terms and fluctuation terms. Now there are
two spectral functions, not just one spectral function as
in rotating-wave approximation, the simulation task will
become more complicate.
As soon as the simulation task is accomplished, one
may go back to the master-equation formulation and treat
it by quantum stochastic trajectory approach. The nonHermitian Hamiltonian is now taken as
i Ĥnh = Ĥ − Γj α̂†j α̂j ,
(29)
2 j
To proceed with concrete analysis and compare with the
result of rotating-wave approximation, we return to the
example with Ψ2 (x) and Ψ1 (x) described by equation (12).
In reference [5] only two fictitious oscillators are introduced to simulate the spectrum R(ω) ≡ R1 (ω) of this
example, the corresponding simulation spectrum R(s) (ω)
fits pretty well with R(ω) in a large region ω/ω0 ≥ 0.5,
which includes the important region around ω/ω0 = 1.
But the R(s) (ω) drops down less slowly than R(ω) when
ω/ω0 → 0. At the origin, R(s) (ω)/R(s) (ω0 ) is still about
0.36 while R(ω) drops to zero (see Fig. 1 of Ref. [5]).
Now there is an additional spectrum R2 (ω), we have
to use one set of N oscillators to simulate both of them,
namely to make
N
|gj |2 Γj
1 (s)
, (31a)
R1 (ω) ∼
= R1 (ω) ≡
2π j=1 (ω − ωj )2 + 14 Γj2
N
gj g¯j Γj
1 (s)
R2 (ω) ∼
· (31b)
= R2 (ω) ≡
2π j=1 (ω − ωj )2 + 14 Γj2
In general, the simulation is not easy to be carried out. But
in the case R2 (ω) = 0, we may simply take four fictitious
oscillators with following parameters:
1
g1 = √ 0.011ω0,
2
1
g2 = √ 0.018ω0,
2
i
g3 = √ 0.011ω0,
2
i
g4 = √ 0.018ω0,
2
Γ1 = 1.3ω0 ,
ω1 = ω0 ,
(32a)
Γ2 = 2.4ω0 ,
ω1 = 1.85ω0, (32b)
Γ3 = 1.3ω0 ,
ω3 = ω0 ,
Γ4 = 2.4ω0 ,
ω4 = 1.85ω0. (32d)
(32c)
It is easy to check that the above set of parameters
(s)
makes R2 (ω) = 0 no matter the relative parity be(s)
tween Ψ2 and Ψ1 is positive or negative, while R1 (ω)
(s)
remains unchanged from the R (ω) of reference [5].
Since the numerical evolution takes place over discrete
times with a small time step ∆t, the wave function |Ψ (t)
is represented by a sequence |Ψ (tn ) with tn = n∆t. Given
the value of |Ψ (tn ), the next one |Ψ (tn+1 ) is determined
by the following algorithm:
(1) evaluate the four kinds of collapse probabilities during
interval (tn , tn−1 ):
Pj (tn ) = Ψ (tn ) Ĉj† Ĉj Ψ (tn )
= Γj Ψ (tn ) α̂†j α̂j Ψ (tn ) ∆t, j = 1, 2, 3, 4
(33)
in which ∆t should be small enough to make P1 (tn )
and P2 (tn ) much small than 1;
284
The European Physical Journal D
Fig. 1. The spontaneous decay of the upper level population
with parameters (γA /ω0 = 1×10−3 ). The curve I represents the
result without both rotating-wave and Markovian approximation, the curve II represents the result with rotating-wave approximation but without Markov approximation as calculated
in reference [5], the curve III represents the Weisskopf-Wigner
result.
(2) generate four random numbers rj (j = 1, 2, 3, 4) which
have uniform probability distribution among the interval (0, 1);
(3) compare Pj (tn ) with rj and derive |Ψ (tn+1 ) according to the following rule: if Pj (tn ) < rj and for all j,
i
e− Ĥnh ∆t |Ψ (tn )
|Ψ (tn+1 ) = ;
†
i
Ψ (tn )|e (Ĥnh −Ĥnh )∆t |Ψ (tn )
(34)
may neglect this term for stable nuclei without introducing obvious error.
We note that the large difference between curve II and
curve III comes from two respects as mentioned in Section 1: the first one concerns the finite size correction,
corresponding to the replacement of γA by γ with γ =
0.729 γA. This correction is calculated analytically and
should be reliable. The second correction is the memory
effect corresponding to the non-uniform of R(ω), namely
the difference of R(ω) from R(ω0 ). This part is evaluated
numerically and has some errors from simulation and from
the pseudo character of the random numbers generated
by program. Both corrections make the decay slower. As
shown in reference [5], the first correction is much larger
than the second. As to the difference between curves I
and II which comes from the effect of counter-rotating
wave coupling, it is totally due to the non-Markovian contribution since in the present case R2 (ω) = 0, γ remains
unchanged as specified in the end of Section 2.
This work is supported by National Natural Science Foundation of China (project number 19774004), and also by the International Program of National Science Foundation of USA
(ECS-9800068).
References
1.
2.
3.
4.
5.
if Pl (tn ) > rl for some one index l and all other
Pm (tn ) ≤ rm ,
6.
Ĉl |Ψ (tn )
|Ψ (tn+1 ) = ·
Ψ (tn )|Ĉl† Ĉl |Ψ (tn )
7.
8.
(35)
9.
In the practical calculation, ∆t should be taken sufficiently small so that the case with any two (or more)
Pj (tn )’s larger than the corresponding rj in the same
interval (tn , tn + ∆t) will not happen.
By carrying out the
over and over
steps
specified above
1
1
from the initial state
|0, 0, 0, 0 where
is the atom
0
0
state in the upper level and |0, 0, 0, 0 denotes the state of
four oscillators with their number of quanta all being zero,
we get a quantum stochastic trajectory of the Monte Carlo
wave function |Ψ (t). The expectation value of a given
operator respect to |Ψ (t) is then calculated. Finally an
ensemble average of 400 trajectories is carried out to give
the resultant curve.
The population on the upper level N̂2 (t) is shown in
Figure 1. We see that the correction of counter-rotating
wave term to the non-Markovian result obtained in reference [5] is quite small even Z is as large as 157. Hence one
10.
P.L. Knight, P.W. Milonni, Phys. Rev. Lett. 56, 275 (1976)
P. Carrazana, G. Vetri, Nuovo Clim. B 55, 191 (1980)
P.T. Robisco, Phys. Lett. A 100, 407 (1984)
J. Seke, W. Herfort, Phys. Rev. A 38, 833 (1988); Phys.
Rev. A 40, 1926 (1989)
C.-Q. Cao, W. Long, J. Wei, H. Cao, Phys. Rev. A 64,
043810 (2001)
See for example, H.J. Carmichael, An Open System Approach to Quantum Optics (Springer-Verlag, Berlin, 1993),
Lecture 7
A. Imamoglu, Phys. Rev. A 50, 3650 (1994)
P. Stenius, A. Imamoglu, Quant. Semiclass. Opt. 8, 283
(1996)
W.H. Louisell, Coupled mode and parametric electronics
(Wiley, New York, 1960)
We may show thatR2 (ω) = 0, provided m1 = m2 . First,
see the integral
dΩk G2k . Substituting equation (7)
into it and carrying out the integration over the angular
variables, we get a series over (l, m, n), and for each
(l, m, n) there are four terms proportional to
[Cl1 +1,l2 ,l;m1 −µ,−m2 ,m Cl1 +1,l2 ,l;m1 +µ,−m2 ,−m ],
[Cl1 +1,l2 ,l;m1 −µ,−m2 ,m Cl1 −1,l2 ,l;m1 +µ,−m2 ,−m ],
[Cl1 −1,l2 ,l;m1 −µ,−m2 ,m Cl1 +1,l2 ,l;m1 +µ,−m2 ,−m ],
[Cl1 −1,l2 ,l;m1 −µ,−m2 ,m Cl1 −1,l2 ,l;m1 +µ,−m2 ,−m ]
respectively. One may easily check that all of them
unless m1 = m2 . Similarly, the integral
will be zero
(nk · Gk )2 dΩk contains four terms proportional to
[Cl1 +1,l2 ,l;m1 −µ,−m2 ,m Cl1 +1,l2 ,l ;m1 −µ −m2 ,−m−µ−µ ],
[Cl1 +1,l2 ,l;m1 −µ,−m2 ,m Cl1 −1,l2 ,l ;m1 −µ ,−m2 ,−m−µ−µ ],
[Cl1 −1,l2 ,l;m1 −µ,−m2 ,m Cl1 +1,l2 ,l ;m1 −µ ,−m2 ,−m−µ−µ ],
[Cl1 −1,l2 ,l;m1 −µ,−m2 ,m Cl1 −1,l2 ,l ;m1 −µ ,−m2 ,−m−µ−µ ]
for each index-set (l, m, µ, l , µ ), they also will be zero
unless m1 = m2