Casimir-Polder forces: a nonperturbative approach
Stefan Yoshi Buhmann,∗ Ludwig Knöll, and Dirk-Gunnar Welsch
Theoretisch-Physikalisches Institut, Friedrich-Schiller-Universität Jena, Max-Wien-Platz 1, 07743 Jena, Germany
Ho Trung Dung
arXiv:quant-ph/0403128v3 12 Jun 2007
Institute of Physics, National Center for Sciences and Technology,
1 Mac Dinh Chi Street, District 1, Ho Chi Minh city, Vietnam
(Dated: February 1, 2008)
Within the frame of macroscopic QED in linear, causal media, we study the radiation force of
Casimir-Polder type acting on an atom which is positioned near dispersing and absorbing magnetodielectric bodies and initially prepared in an arbitrary electronic state. It is shown that minimal
and multipolar coupling lead to essentially the same lowest-order perturbative result for the force
acting on an atom in an energy eigenstate. To go beyond perturbation theory, the calculations are
based on the exact center-of-mass equation of motion. For a nondriven atom in the weak-coupling
regime, the force as a function of time is a superposition of force components that are related to the
electronic density-matrix elements at a chosen time. Even the force component associated with the
ground state is not derivable from a potential in the ususal way, because of the position dependence
of the atomic polarizability. Further, when the atom is initially prepared in a coherent superposition
of energy eigenstates, then temporally oscillating force components are observed, which are due to
the interaction of the atom with both electric and magnetic fields.
PACS numbers: 12.20.-m, 42.50.Vk, 42.50.Nn, 32.70.Jz
I.
INTRODUCTION
It is well known that in the presence of macroscopic
bodies an atom in the ground state (or in an excited energy eigenstate) is subject to a nonvanishing force — the
Casimir-Polder (CP) force — that results from the vacuum fluctuations of the electromagnetic field. CP forces
play an important role in a variety of processes in physical
chemistry, atom optics, and cavity QED. Moreover, they
hold the key to a number of potential applications in nanotechnology such as the construction of atomic-force microscopes [1] or reflective atom-optical elements [2]. Over
the years, substantial efforts have been made to improve
the understanding of CP forces (for reviews, see Ref. [3]).
Measuring CP forces acting on individual particles is a
challenging task. Since the early observation of the deflection of thermal atomic beams by conducting surfaces
[4], measurement techniques and precision have been improving continuously. More recent experiments have been
performed with atomic beams traversing between parallel plates [5]. Other methods include transmission grating
diffraction of molecular beams [6], atomic quantum reflection [7, 8], evanescent-wave atomic mirror techniques [9],
and indirect measurements via spectroscopic means [10].
Proposals have been made on improvements of monitoring the CP interaction by using atomic interferometry
[11].
The theoretical approaches to the problem of determining the CP force can be roughly divided into two categories. In the first, first-principle approach explicit field
∗ Electronic
address:
[email protected]
quantization is performed and perturbation theory is applied to calculate the body-induced atomic energy shift,
which is regarded as the potential of the force in lowestorder perturbation theory [12, 13, 14, 15, 16, 17, 18, 19].
The calculations have typically been based on macroscopic QED, by beginning with a normal-mode decomposition and including the bodies via the well-known
conditions of continuity at the surfaces of discontinuity.
Since in such a (noncausal) approach the frequency dependence of the bodies’ response to the field cannot be
properly taken into account, material dispersion and absorption are commonly ignored. As has been shown recently [20], the problem does not occur within the frame
of a generalized quantization scheme that properly takes
into account a Kramers-Kronig consistent response of
the bodies to the field. Clearly, the problem can also
be circumvented in microscopic QED, where the bodies are treated on a microscopic level by adopting, e.g.,
harmonic-oscillator models [14]. In the second, semiphenomenological approach, the problem is circumvented by
basing the calculations on linear response theory (LRT),
without explicitly quantizing the electromagnetic field
[21, 22, 23, 24, 25, 26, 27, 28]. In the ansatz for the force,
either the field quantities or both the field and the atomic
entities are expressed in terms of correlation functions,
which in turn are related, via the fluctuation-dissipation
theorem, to response functions.
At first glance one would expect the result obtained
from exploiting LRT to be more generally valid than the
QED result obtained in lowest-order perturbation theory. In fact, this is not the case. In both approaches,
it is not the exact atomic polarizability that enters the
expression for the (ground-state) CP force but the approximate expression which is obtained in lowest-order
2
perturbation theory and which effectively corresponds to
the atomic polarizability in free space. Since the structure
of the electromagnetic field is changed in the presence of
macroscopic bodies, the atomic polarizability is expected
to change as well. It is well known that the atomic level
shifts and broadenings sensitively depend on the material surroundings. In particular, when an atom is situated
very close to a body, the effect can be quite significant
(see, e.g., Refs. [29, 30]), thereby changing the atomic
polarizability. As a result, a position-dependent polarizability is expected to occur, which prevents the CP force
from being derivable from a potential in the usual way.
II.
SKETCH OF THE QUANTIZATION
SCHEME
A.
Minimal coupling
In Coulomb gauge, the minimal-coupling Hamiltonian
of an atomic system (e.g., an atom or a molecule) consisting of nonrelativistic charged particles interacting with
the electromagnetic field in the presence of macroscopic
dispersing and absorbing bodies reads [37]
Z ∞
X Z
Ĥ =
d3 r
dω ~ω f̂λ† (r, ω)f̂λ (r, ω)
0
λ=e,m
+
X
α
A way to derive a more rigorous expression for the CP
force is to base the calculations on the exact quantummechanical center-of-mass equation of motion of the
atom as we shall do in this paper. The calculations are
performed for both minimal and multipolar coupling, and
contact is made with earlier studies of the center-of-mass
motion of an atom in free space, with special emphasis
on the so-called Röntgen interaction term that appears
in the multipolar Hamiltonian [31, 32, 33, 34, 35]. After taking the expectation value with respect to the internal (electronic) quantum state of the atom and the
quantum state of the medium-assisted electromagnetic
field, the resulting force formula can be used to calculate the time-dependent force acting on a nondriven or
driven atom that is initially prepared in an arbitrary (internal) quantum state. In this paper, the force formula
is further evaluated for the case of a nondriven, initially
arbitrarily prepared atom, by assuming weak atom-field
coupling treated in Markovian approximation. It is worth
noting that the theory, being based on the quantized version of the macroscopic Maxwell field, with the bodies
being described in terms of spatially varying, KramersKronig consistent complex permittivities and permeabilities [36, 37], also applies to left-handed materials [38]
where standard quantization runs into difficulties.
The paper is organized as follows. After a brief sketch
of the quantization scheme (Sec. II), in Sec. III attention is focused on the perturbative treatment of the CP
force acting on an atom in an energy eigenstate, and previous results [20] obtained for dielectric surroundings of
the atom are extended to magnetodielectric surroundings, including left-handed materials. In Sec. IV the exact
center-of-mass Heisenberg equation of motion of an atom
and the Lorentz force therein are studied, and Sec. V is
devoted to the calculation of the average force, with special emphasis on a nondriven atom in the weak-coupling
regime. Finally, a summary and some concluding remarks
are given in Sec. VI.
+ 21
Z
1
2mα
h
i2
p̂α − qα Â(r̂α )
3
d r ρ̂A (r)ϕ̂A (r) +
Z
d3 r ρ̂A (r)ϕ̂(r),
(1)
where
X
ρ̂A (r) =
α
qα δ(r − r̂α )
(2)
and
ϕ̂A (r) =
Z
d3 r′
ρ̂A (r′ )
4πε0 |r − r′ |
(3)
are the charge density and scalar potential of the particles, respectively. The particle labeled α has charge qα ,
mass mα , position r̂α , and canonically conjugated momentum p̂α . The fundamental Bosonic fields f̂λ (r, ω) [and
f̂λ† (r, ω)] which can be related to noise polarization (for λ
=e) and noise magnetization (for λ=m), respectively, are
the dynamical variables for describing the system composed of the electromagnetic field and the medium including the dissipative system responsible for absorption,
h
i
fˆλi (r, ω), fˆλ†′ i′ (r′ , ω ′ ) = δλλ′ δii′ δ(r − r′ )δ(ω − ω ′ ), (4)
h
i
(5)
fˆλi (r, ω), fˆλ′ i′ (r′ , ω ′ ) = 0.
Note that the first term on the right-hand side of Eq. (1)
is the energy of that system. Further Â(r) and ϕ̂(r) are
the vector and scalar potentials of the medium-assisted
electromagnetic field, respectively, which in Coulomb
gauge are expressed in terms of the fundamental fields
f̂λ (r, ω) [and f̂λ† (r, ω)] as
Z ∞
Â(r) =
dω (iω)−1 Ê⊥ (r, ω) + H.c.,
(6)
0
−∇ϕ̂(r) =
Z
∞
dω Êk (r, ω) + H.c.,
(7)
d3 r′ Gλ (r, r′ , ω)f̂λ (r′ , ω),
(8)
0
where
Ê(r, ω) =
X Z
λ=e,m
3
ω2
Ge (r, r , ω) = i 2
c
′
r
~
Im ε(r′ , ω) G(r, r′ , ω),
πε0
Z
r
ω
~
Im κ(r′ , ω)
Gm (r, r′ , ω) = −i
−
c
πε0
←
−
× G(r, r′ , ω)×∇r′ ,
(10)
←
−
with G(r, r′ , ω)× ∇r′ ij = ǫjkl ∂l′ Gik (r, r′ , ω) and κ(r, ω)
= µ−1 (r, ω). Here and in the following, transverse and
longitudinal vector fields are denoted by ⊥ and k, respectively, e.g.,
Z
Ê⊥(k) (r, ω) = d3 r′ δ ⊥(k) (r − r′ )Ê(r′ , ω),
(11)
with
k
δij (r) = −∂i ∂j
1
4πr
G(r, r′ , ω) = G⊤ (r′ , r, ω),
(9)
(12)
n
←
−
d3 s Im κ(s, ω) G(r, s, ω) × ∇s ∇s × G∗ (s, r′ , ω)
+
o
ω2
∗
′
Im
ε(s,
ω)
G(r,
s,
ω)G
(s,
r
,
ω)
= Im G(r, r′ , ω).
c2
(18)
Combining Eq. (18) with Eqs. (9) and (10) yields
X Z
d3 s Gλik (r, s, ω)G∗λjk (r′ , s, ω)
λ=e,m
=
⊥
δij
(r)
= δ(r)δij −
(13)
being the longitudinal and transverse dyadic δ functions,
respectively.
In Eqs. (9) and (10), G(r, r′ , ω) is the (classical) Green
tensor, which in the case of magnetodielectric matter
obeys the equation
ω2
∇ × κ(r, ω)∇ × − 2 ε(r, ω) G(r, r′ , ω) = δ(r − r′ )
c
(14)
together with the boundary condition
G(r, r′ , ω) → 0 for |r − r′ | → ∞.
~µ0 2
ω Im Gij (r, r′ , ω).
π
(19)
Note that in Eq. (19) and throughout the remaining part
of this paper, summation over repeated vector indices is
understood.
The total electric field is given by
and
k
δij (r)
(17)
~ˆ
E(r)
= Ê(r) − ∇ϕ̂A (r),
(20)
where
Ê(r) =
∞
Z
dω Ê(r, ω) + H.c.,
(21)
0
with Ê(r, ω) from Eq. (8). Accordingly, the total induction field reads
Z ∞
~ˆ
(22)
dω B̂(r, ω) + H.c.,
B(r)
= B̂(r) =
0
where
B̂(r, ω) = (iω)−1 ∇ × Ê(r, ω).
(15)
(23)
Note that the (relative) permittivity ε(r, ω) and permeability µ(r, ω) of the (inhomogeneous) medium are complex functions of frequency, whose real and imaginary
parts satisfy the Kramers-Kronig relations. Since for absorbing media we have Im ε(r, ω) > 0 and Im µ(r, ω) > 0
⇒ Im κ(r, ω) < 0, the expressions under the square roots
in Eqs. (9) and (10) are positive. It should be pointed out
that the whole space is assumed to be filled with some
(absorbing) media, in which case the aforementioned conditions for Im ε(r, ω) and Im µ(r, ω) ensure that the differential equation (14) together with the boundary condition (15) presents a well-defined problem. However, as
this assumption allows for both ε(r, ω) and µ(r, ω) to be
arbitrarily close to unity (i.e., for arbitrarily dilute matter), it is naturally possible to include vacuum regions in
the theory, by performing the limit ε(r, ω) → 1, µ(r, ω)
→ 1 in these regions after having calculated the desired
expectation values of the relevant quantities as functions
of ε(r, ω) and µ(r, ω).
The Green tensor has the following useful properties
[36]:
Finally, the displacement and magnetic fields are given
by
G∗ (r, r′ , ω) = G(r, r′ , −ω ∗ ),
ˆr̄α = r̂α − r̂A
(16)
~ˆ
D(r)
= D̂(r) − ε0 ∇ϕ̂A (r)
Z ∞
h
i
=
dω D̂(r, ω) + H.c. − ε0 ∇ϕ̂A (r), (24)
0
Z ∞
ˆ
~
dω Ĥ(r, ω) + H.c.,
H(r)
= Ĥ(r) =
(25)
0
where
D̂(r, ω) = ε0 ε(r, ω)Ê(r, ω)
r
~ε0
Im ε(r, ω) f̂e (r, ω),
+i
π
Ĥ(r, ω) = κ0 κ(r, ω)B̂(r, ω)
r
~κ0
Im κ(r, ω) f̂m (r, ω).
− −
π
(26)
(27)
Assuming that the atomic system is sufficiently localized, and introducing shifted particle coordinates
(28)
4
relative to the center of mass
X mα
r̂α
r̂A =
mA
α
terms of the transformed variables] can be given in the
form of (see Appendix A)
(29)
P
(mA = α mα ), we can apply the long-wavelength approximation by expanding the fields Â(r) and ϕ̂(r)
around the center of mass and keeping only the leading nonvanishing terms of the respective field operators.
For a neutral atomic system,
X
qA =
qα = 0,
(30)
α
this is just the familiar electric dipole approximation, and
the Hamiltonian (1) simplifies to
Ĥ = ĤF + ĤA + ĤAF ,
Ĥ =
X Z
Z
3
d r
0
λ=e,m
1
+
2ε0
X
+
α
Z
∞
3
d
dω ~ω f̂λ′† (r, ω)f̂λ′ (r, ω)
r P̂2A (r)
−
Z
d3 r P̂A (r)Ê′ (r)
2
Z
1
′
3
′
p̂ + d r Ξ̂α (r) × B̂ (r) , (38)
2mα α
where
Ξ̂α (r) = qα Θ̂α (r) −
(31)
mα X
mα
qβ Θ̂β (r) +
P̂A (r) (39)
mA
mA
β
and
where
ĤF ≡
ĤA ≡
X Z
3
d r
λ=e,m
X p̂2
α
+
2m
α
α
Z
1
2
∞
0
Z
d3 r ρ̂A (r)ϕ̂A (r),
ĤAF ≡ d̂∇ϕ̂(r)|r=r̂A −
+
dω ~ω f̂λ† (r, ω)f̂λ (r, ω),
(32)
(33)
X qα
p̂α Â(r̂A )
mα
α
X q2
α
Â2 (r̂A ),
2m
α
α
(34)
with
d̂ =
X
qα r̂α =
α
X
qα ˆr̄α
(35)
α
being the total electric dipole moment.
B.
Multipolar coupling
Let us turn to the multipolar coupling scheme widely
used for studying the interaction of electromagnetic fields
with atoms and molecules. Just as in standard QED, so
in the present formalism [36, 37], the multipolar Hamiltonian can be obtained from the minimal-coupling Hamiltonian by means of a Power-Zienau transformation,
Z
i
3
Û = exp
d r P̂A (r)Â(r) ,
(36)
~
where
P̂A (r) =
X
α
qα ˆr̄α
Z
0
Θ̂α (r) = ˆr̄α
Z
0
1
dλ λ δ(r − r̂A − λˆr̄α ).
Note that due to the unitarity of the transformation (36),
the transformed variables of the atomic system r̂′α = r̂α
and p̂′α and the transformed field variables f̂λ′ (r, ω) and
f̂λ′† (r, ω) obey the same commutation relations as the untransformed ones. Needless to say that the transformed
fields Ê′ (r) and B̂′ (r) are related to the transformed fields
f̂λ′ (r, ω) and f̂λ′† (r, ω) according to Eq. (8) and Eqs. (21)–
(23), with primed quantities instead of the unprimed
ones. The Hamiltonian (38) can be regarded as the generalization of the multipolar Hamiltonian obtained earlier
for moving atoms in vacuum [31, 32, 33, 34, 35] to the
case where dispersing and absorbing magnetodielectric
bodies are present. In particular, it can be used to describe effects specifically due to the translational motion
of an atomic system such as Doppler and recoil effects.
Applying the long-wavelength approximation to the
fields Ê′ (r) and B̂′ (r) in Eq. (38), which is equivalent to
approximating δ(r − r̂A − λˆr̄α ) by δ(r − r̂A ) in Eqs. (37)
and (40), respectively, i.e.,
P̂A (r) = d̂ δ(r − r̂A ),
Θ̂α (r) =
1
2
ˆr̄α δ(r − r̂A ),
dλ δ(r − r̂A − λˆr̄α ).
(37)
For a neutral atomic system, the multipolar Hamiltonian
[which is obtained by expressing the Hamiltonian (1) in
(41)
(42)
thus
Ξ̂α (r) = 21 qα ˆr̄α δ(r − r̂A ) +
1
(40)
mα
d̂δ(r − r̂A ),
2mA
(43)
we obtain the multipolar Hamiltonian in long-wavelength
approximation,
′
′
Ĥ = ĤF′ + ĤA
+ ĤAF
,
(44)
5
with
III.
ĤF′ ≡
′
ĤA
X Z
λ=e,m
d3 r
Z
∞
0
dω ~ω f̂λ′† (r, ω)f̂λ′ (r, ω),
Z
X p̂′ 2
1
α
≡
+
d3 r P̂2A (r),
2m
2ε
α
0
α
′
ĤAF
≡ −d̂Ê′ (r̂A ) +
+
+
(45)
(46)
X qα
ˆ ′α ˆr̄α × B̂′ (r̂A )
p̄
2m
α
α
X q2
2
2
3
α ˆ
d̂ × B̂′ (r̂A )
r̄α × B̂′ (r̂A ) +
8m
8m
α
A
α
1 ′
p̂ d̂× B̂′ (r̂A ),
mA A
(47)
where
p̂′A =
X
p̂′α
(48)
VAN DER WAALS POTENTIAL
According to Casimir’s and Polder’s pioneering concept [12], the CP force on an atomic system near macroscopic bodies is commonly regarded as being a conservative force. In particular, it is assumed that for an atom in
an eigenstate |li of the atomic Hamiltonian the positiondependent shift of the corresponding eigenvalue due to
the (electric-dipole) interaction of the atomic system with
the body-assisted electromagnetic field is the potential,
also referred to as van der Waals (vdW) potential, from
which the CP force can be derived, where the calculations
are usually performed within the frame of lowest-order
perturbation theory. In this picture, the center-of-mass
coordinate is a parameter rather than a dynamical variable (r̂A 7→ rA ). Following this line, we first extend previous results [20], and show that minimal and multipolar
coupling schemes yield essentially the same expression for
the force.
α
is the (canonical) momentum of the center of mass, and
ˆ ′α = p̂′α −
p̄
mα ′
p̂
mA A
(49)
denote shifted momenta of the particles relative to the
center of mass. The first two terms on the right-hand side
of Eq. (47) represent electric and magnetic dipole interactions, respectively, the next two terms describe the (generalized) diamagnetic interaction of the charged particles
with the medium-assisted electromagnetic fields, while
the last term describes the Röntgen interaction due to the
translational motion of the center of mass. In particular,
in (generalized) electric dipole approximation, Eq. (47)
reads
′
ĤAF
= −d̂Ê′ (r̂A ) +
1 ′
p̂ d̂× B̂′ (r̂A ).
mA A
(50)
Recall that the transformed medium-assisted electric
field Ê′ (r̂) is related to the physical one, Ê(r̂), according to Eq. (A4).
If the center-of-mass coordinate is treated as a (classical) parameter (r̂A 7→ rA ), then Eq. (39) reduces to
Ξ̂α (r) = qα Θ̂α (r),
(51)
which corresponds to the limit mα /mA → 0. Hence
Eq. (47) becomes
X qα
′
= −d̂Ê′ (rA ) +
ĤAF
p̂′ ˆr̄α × B̂′ (rA )
2mα α
α
+
X q2
2
α ˆ
r̄α × B̂′ (rA ) .
8mα
α
A.
Minimal coupling
We start from the minimal-coupling Hamiltonian in
electric dipole approximation as given by Eqs. (31)–(34)
together with Eq. (35) (r̂A 7→ rA ). Let |ni denote the
eigenstates of the multilevel atomic system and write ĤA
[Eq. (33)] as
X
ĤA =
En |nihn|.
(53)
n
To calculate the leading-order correction to the unperturbed eigenvalue of a state |li|{0}i due to the perturbation Hamiltonian (34) [|{0}i, ground state of the fundamental fields f̂λ (r, ω)], we first note that the first two
terms have no diagonal elements. Thus they start to contribute in second order,
Z ∞
Z
1X X
dω
∆2 El = −
P
d3 r
~
ωkl + ω
0
k λ=e,m
X qα
p̂α Â(rA )
× hl|h{0}|d̂∇ϕ̂(r)|r=rA −
mα
α
× |{1λ (r, ω)}i|ki
If the paramagnetic and diamagnetic terms are omitted,
the interaction Hamiltonian simply reduces to the first
term on the right-hand side of Eq. (52).
(54)
(P, principal part), whereas the third term starts to contribute in first order,
∆1 El = hl|h{0}|
(52)
2
X q2
α
Â2 (rA )|{0}i|li.
2m
α
α
(55)
Here, |{1λ (r, ω)}i ≡ f̂λ† (r, ω)|{0}i denotes single-quantum
Fock states of the fundamental fields, and
ωkl ≡ (Ek − El )/~
(56)
are the atomic transition frequencies. Since ∆1 El and
∆2 El are quadratic in the coupling constant [Eqs. (B9)
6
and (B10) in Appendix B], thus being of the same order of
magnitude, the leading-order correction to the eigenvalue
is given by
∆El = ∆1 El + ∆2 El .
(57)
A straightforward but somewhat lengthy calculation
yields (see Appendix C)
Z ∞
n
µ0 X
dω
dlk ωkl ω
∆El =
P
π
ωkl + ω
0
k
× Im G(rA , rA , ω) − Im k Gk (rA , rA , ω)
o
− ω 2 Im k Gk (rA , rA , ω) dkl ,
(58)
with
dlk = hl|d̂|ki
(59)
being the dipole matrix elements.
Since the atomic system should be located in a freespace region, the Green tensor in this region is a linear
superposition of the (translationally invariant) vacuum
Green tensor G(0) and the scattering Green tensor G(1)
that accounts for the spatial variation of the permittivity
and permeability,
′
G(r, r , ω) = G
(0)
′
(r, r , ω) + G
(1)
′
(r, r , ω).
∆El =
+
(1)
∆El (rA ).
(61)
(0)
∆El
The rA -independent term
associated with the
vacuum Green tensor gives rise to the vacuum Lamb
shift and is not of interest here. The rA -dependent term
(1)
∆El (rA ), associated with the scattering Green tensor,
is just the vdW potential sought,
(1)
(1)
ω 2 (1)
G (r, r, ω) = 0
|ω|→∞ c2
lim
(1)
Ul (rA ) = ∆El (rA ) = ∆1 El (rA ) + ∆2 El (rA ). (62)
Hence from Eq. (58) [G(rA , rA , ω) 7→ G(1) (rA , rA , ω)] we
derive, on recalling Eq. (16) and changing the integration
variable from −ω to ω,
Z ∞
µ0 X
dω
Ul (rA ) =
dlk P
2iπ
ω
kl + ω
0
k
n
× ωkl ω G(1) (rA , rA , ω) − k G(1)k (rA , rA , ω)
Z −∞
o
dω
− ω 2k G(1)k (rA , rA , ω) − P
ωkl − ω
0
n
(1)
k (1)k
× ωkl ω G (rA , rA , ω) − G (rA , rA , ω)
o
(63)
+ ω 2k G(1)k (rA , rA , ω) dkl .
This equation can be greatly simplified by using contour-integral techniques. G(1) (rA , rA , ω) is an analytic
(64)
(cf. Ref. [37]), so we finally arrive at
Ul (rA ) = Ulor (rA ) + Ulr (rA ),
(65)
where
(60)
As a consequence, the eigenvalue correction ∆El can be
decomposed into two parts,
(0)
∆El
function in the upper half of the complex frequency plane,
including the real axis (apart from a possible pole at
ω = 0). Furthermore, knowing the asymptotic behaviour
of the Green tensor in the limit ω → 0 (cf. Ref. [37]), one
can verify that all integrands in Eq. (63) remain finite
in this limit. We may therefore apply Cauchy’s theorem,
and replace the principal value integral over the positive (negative) real half axis by a contour integral along
the positive imaginary half axis (introducing the purely
imaginary coordinate ω = iu) and along a quarter circle
with infinitely large radius in the first (second) quadrant of the complex frequency plane plus, in the case of
ωlk > 0, a contour integral along an infinitesimally small
half circle around ω = ωlk (ω = −ωlk ) in the first (second)
quadrant of the complex frequency plane. The integrals
along the infinitely large quarter circles vanish due to the
asymptotic property
Ulor (rA ) =
Z
ωkl u2
µ0 X ∞
du 2
π
ωkl + u2
0
k
× dlk G(1) (rA , rA , iu)dkl
is the off-resonant part of the vdW potential, and
X
2
Ulr (rA ) = −µ0
Θ(ωlk )ωlk
(66)
k
× dlk Re G(1) (rA , rA , ωlk )dkl
(67)
[Θ(z), unit step function] is the resonant part arising from
the contribution from the residua at the poles. Note that
Ulr (rA ) vanishes when the atomic system is in the ground
state. For an atomic system in an excited state, Ulr (rA )
may dominate Ulor (rA ).
The CP force can be derived from Eq. (65) according
to
Fl (rA ) = −∇A Ul (rA )
(68)
(∇A ≡ ∇rA ). A formula of the type of Eq. (65) together
with Eqs. (66) and (67) was first given in Ref. [23] within
the frame of LRT.
To give Eq. (66) in a more compact form, we introduce
the generalized atomic polarizability tensor
1X
dmk ⊗ dkn
αmn (ω) =
~
ω̃kn − ω − i(Γk + Γm )/2
k
dkn ⊗ dmk
,
(69)
+
ω̃km + ω + i(Γk + Γn )/2
7
where ω̃km are the shifted (renormalized) transition frequencies and Γk are the excited-state widths. Following
Ref. [39], we may regard
αl (ω) = αll (ω)
(70)
as being the ordinary (Kramers-Kronig-consistent) polarizability tensor of an atom in state |li. Hence we may
rewrite Eq. (66) as
Ulor (rA )
~µ0
=
2π
∞
Z
0
(0)
du u2 Tr αl (iu) G(1) (rA , rA , iu) ,
(71)
where
2X
ωkl
2 − ω 2 − iωǫ dlk ⊗ dkl
ǫ→0 ~
ωkl
(0)
αl (ω) = lim
(72)
k
is the polarizability tensor in lowest-order perturbation
theory, which can be obtained from Eq. (70) together
with Eq. (69) by ignoring both the level shifts and broadenings. In particular for an atom in a spherically symmetric state, we have
2 X
ωkl
(0)
(0)
|dlk |2 I
αl (ω) = αl (ω)I = lim
2
ǫ→0 3~
ωkl − ω 2 − iωǫ
k
(73)
(I, unit tensor), so that Eq. (71) reduces to
Ulor (rA )
~µ0
=
2π
Z
∞
0
(0)
du u2 αl (iu) Tr G(1) (rA , rA , iu),
(74)
and Eq. (67) simplifies to
µ0 X
2
Θ(ωlk )ωlk
|dlk |2
3
k
× Tr Re G(1) (rA , rA , ωlk ) .
Ulr (rA ) = −
Note that
(0)
αl (iu) ≃ αl (iu)
(75)
(76)
is typically valid for an atomic system in free space, because of the smallness of the level shifts and broadenings
that result from the interaction of the atomic system with
the vacuum electromagnetic field.
Equation (65) together with Eqs. (66) and (67) can be
regarded as being the natural extension of the QED results obtained on the basis of the familiar normal-mode
formalism, which ignores material absorption. Moreover,
it does not only apply to arbitrary causal dielectric bodies, but, to our knowledge, it first proves applicable to
magnetodielectric matter such as left-handed material,
for which standard quantization concepts run into difficulties. Note that all information about the electric and
magnetic properties of the matter is contained in the
scattering Green tensor.
Finally, let us briefly comment on the ground-state potential as given by Eq. (71) for l = 0. In terms of an integral along the positive frequency axis, it reads
Z
~µ0 ∞
dω ω 2
U0 (rA ) = −
2π 0
n
o
(0)
× Im Tr α0 (ω)G(1) (rA , rA , ω) .
(77)
An expression of this type can also be obtained by using the methods of LRT [23, 25]. It allows for a simple
physical interpretation for the ground-state CP force as
being due to correlations of the fluctuating electromagnetic field with the corresponding induced electric dipole
of the atomic system plus the correlations of the fluctuating electric dipole moment with its induced electric field
[28].
B.
Multipolar coupling
Let us now consider the multipolar Hamiltonian in
long-wavelength approximation as given by Eqs. (44)–
′
(46) together with Eq. (52), and write ĤA
[Eq. (46)] in
the form of Eq. (53). In contrast to the electric dipole approximation considered in the minimal coupling scheme,
the present Hamiltonian also includes magnetic interactions. One might therefore expect that the leading-order
corrections to the unperturbed eigenvalues are given by
the second-order corrections due to the dipole interactions (linear in the field variables) plus the first-order
correction due to the diamagnetic interaction (quadratic
in the field variables), all of these contributions being
quadratic in the coupling constant. However, one can
show [Eqs. (B16)–(B18) in Appendix B] that the secondorder eigenvalue correction due to magnetic dipole interaction is smaller than that due to the electric dipole interaction by a factor of (Zeff α0 )2 , where Zeff is the effective
nucleus charge felt by the electrons giving the main contribution to the energy shift, and α0 is the fine-structure
constant. The current formalism based on Hamiltonian
(1) only treats nonrelativistic atomic systems, which are
characterized by Zeff α0 ≪ 1 [40], so we can safely neglect
the correction arising from the magnetic dipole interaction. Furthermore, the first-order correction arising from
the diamagnetic term can be shown to be smaller than
the second-order correction due to the electric dipole interaction by the same factor (Zeff α0 )2 , so we can disregard it for the same reason.
In summary, the main contribution to the eigenvalue
shift of a state |li|{0′ }i [|{0′ }i, ground state of the
transformed fundamental fields f̂λ′ (r, ω)] is the secondorder correction due to the electric dipole interaction in
8
Summing Eq. (81) over α, recalling definition (29), and
using Eqs. (20) and (22) together with the relationship
X
qα ∇α ϕ̂A (r̂α ) = 0
(82)
Eq. (52), i.e.,
∆El = ∆2 El = −
×
Z
1X X
P
~
k λ=e,m
Z
∞
0
dω
ωkl + ω
d3 r hl|h{0′ }| − d̂Ê′ (rA )|{1′λ (r, ω)}i|ki
α
2
(78)
[|{1′λ (r, ω)}i ≡ f̂λ′† (r, ω)|{0′ }i]. After some algebra it can
be found that (see Appendix C)
Z ∞
ω2
µ0 X
dω
∆El = −
P
π
ωkl + ω
0
k
× dlk Im G(rA , rA , ω)dkl .
(79)
We now apply the same procedure as in Sec. III A, below
Eq. (58). Replacing the Green tensor by its scattering
part and transforming the frequency integral to imaginary frequencies using contour integral techniques, we
arrive at exactly the same form of the vdW potential as
given in Eq. (65) together with Eqs. (66) and (67). It
is worth noting that the two schemes lead to equivalent
results only if in the minimal-coupling scheme the Â2
coupling term is properly taken into account.
IV.
CENTER-OF-MASS MOTION AND
LORENTZ FORCE
Atomic quantities that are related to the atom–field interaction can drastically change when the atomic system
comes close to a macroscopic body, the spontaneous decay thus becoming purely radiationless, with decay rates
and level shifts being inversely proportional to the atomsurface separation to the third power [29]. Clearly, in this
case approximations of the type (76) cannot be made in
general and the perturbative approach to the calculation
of the CP force becomes questionable. Moreover, when
the atomic system is not in the ground state, then dynamical effects can no longer be disregarded. To go beyond perturbation theory, let us first consider the centerof-mass Newtonian equation of motion and the Lorentz
force therein.
(∇α ≡ ∇r̂α ), we derive
mA ¨r̂A = F̂,
where the Lorentz force takes the form
Z
F̂ = d3 r ρ̂A (r)Ê(r) + ĵA (r) × B̂(r) ,
Minimal coupling
As has been shown [37], the Heisenberg equations of
motion governed by the minimal-coupling Hamiltonian
(1),
2 h
i
¨r̂α = 1
r̂α , Ĥ , Ĥ ,
(80)
i~
lead to the well-known Newtonian equations of motion
for the individual charged particles,
i
h
ˆ
ˆ
ˆ
1 ˙
¨
˙
~
~
~
mα r̂α = qα E(rα )+ 2 r̂α × B(rα )− B(rα )× r̂α . (81)
(84)
with charge density ρ̂A (r) and current density ĵA (r) being
defined by Eq. (2) and
i
X h
(85)
qα r̂˙ α δ(r − r̂α ) + δ(r − r̂α )r̂˙ α ,
ĵA (r) = 21
α
respectively. It can be shown [31, 36, 41] that for neutral atoms the atomic charge and current densities can
be expressed in terms of atomic polarization and magnetization according to
ρ̂A (r) = −∇P̂A (r)
(86)
and
˙
ĵA (r) = P̂A (r) + ∇ × M̂A (r) + ∇ × M̂R (r),
(87)
respectively, where
i
X h
M̂A (r) = 12
qα Θ̂α (r) × ˆr̄˙ α − ˆr̄˙ α × Θ̂α (r) , (88)
α
M̂R (r) =
1
2
h
i
P̂A (r) × r̂˙ A − r̂˙ A × P̂A (r) ,
(89)
with P̂A (r) and Θ̂α (r) from Eqs. (37) and (40), respectively. Note that the last term in Eq. (87) represents
the so-called Röntgen current [41, 42], which is a feature of the overall translational motion of any aggregate
of charges.
Inspection of Eqs. (37), (40), (88), and (89) shows that
the relations
∇ ⊗ P̂A (r) = −∇A ⊗ P̂A (r),
A.
(83)
∇ ⊗ M̂A(R) (r) = −∇A ⊗ M̂A(R) (r)
(90)
(91)
(∇A ≡∇r̂A ) are valid. We therefore may write, on recalling
Maxwell’s equations,
Z
− d3 r ∇P̂A (r) Ê(r)
Z
Z
˙
= ∇A d3 r P̂A (r)Ê(r) + d3 r P̂A (r) × B̂(r), (92)
Z
d3 r ∇ × M̂A(R) (r) × B̂(r)
Z
= ∇A d3 r M̂A(R) (r)B̂(r) .
(93)
9
Substituting Eqs. (86) and (87) into Eq. (84) and using Eqs. (92) and (93), we may equivalently express the
Lorentz force as
Z
F̂ = ∇A
d3 r P̂A (r)Ê(r)
Z
3
+ d r M̂A (r) + M̂R (r) B̂(r)
Z
d
+
d3 r P̂A (r) × B̂(r).
(94)
dt
In long-wavelength approximation, Eqs. (88) and (89)
simplify to [recall Eqs. (41) and (42)]
X
qα δ(r − r̂A )ˆr̄α × ˆr̄˙ α
M̂A (r) = 14
α
− ˆr̄˙ α × ˆr̄α δ(r − r̂A )
and
M̂R (r) =
1
2
(95)
δ(r − r̂A )d̂ × r̂˙ A − r̂˙ A × d̂ δ(r − r̂A ) , (96)
respectively, so that the Lorentz force (94) can be written
as
X
qα r̂˙ α B̂(r̂A ) × ˆr̄α
F̂ = ∇A d̂Ê(r̂A ) + 21
α
d
1˙
d̂ × B̂(r̂A ) .
+ 2 r̂A B̂(r̂A ) × d̂ +
dt
(97)
Further, we calculate
i
i
dh
ih
d̂ × B̂(r̂A ) =
Ĥ, d̂ × B̂(r̂A )
dt
~
˙
˙
= d̂ × B̂(r̂A ) + d̂ × B̂(r) r=r̂
A
h
←
− i
1 ˙
+ d̂ × 2 r̂A ∇A ⊗ B̂(r̂A ) + B̂(r̂A ) ⊗ ∇A r̂˙ A .
(98)
Comparing the different terms in Eq. (97), one can show
[Eqs. (B19), (B20), and (B22)–(B24) in Appendix B] that
the second term in curly brackets is typically smaller than
the first one by a factor of v/c + Zeff α0 (v, velocity of the
center of mass), while the third term is smaller than the
first one by a factor of v/c. Similarly, we find [Eqs. (B25)–
(B27) in Appendix B] that the third term in Eq. (98)
is smaller than the first two terms by a factor of v/c.
Thus in the nonrelativistic limit considered throughout
the current work [cf. Hamiltonian (1)] we can set
d
F̂ = ∇ d̂Ê(r) +
d̂ × B̂(r)
.
(99)
dt
r=r̂A
In the absense of magnetodielectric bodies, Eq. (99)
reduces to earlier results derived within the multipolar
coupling scheme for an atom interacting with the electromagnetic field in free space [32, 33]. However, it should
be pointed out that here the electric and magnetic fields
Ê(r) and B̂(r), respectively, are the medium-assisted
fields as defined by Eqs. (21) and (22) [together with
Eqs. (8) and (23)]. Thus Eq. (94) or, in electric dipole
approximation, Eq. (99) determine the force acting on
an atomic system in the very general case of dispersing
and absorbing magnetodielectric bodies being present —
a result that has not yet been derived elsewhere.
B.
Multipolar coupling
Using the multipolar Hamiltonian (38), we obtain, on
recalling that r̂′α = r̂α ,
Z
i
′
˙
mα r̂α = Ĥ, mα r̂α = p̂α + d3 r Ξ̂α (r)×B̂′ (r). (100)
~
Summing Eq. (100) over α and taking into account
Eqs. (29) and (48) yields
Z
′
˙
(101)
mA r̂A = p̂A + d3 r P̂A (r) × B̂′ (r).
Equation (101) leads to
Z
i
mA¨r̂A = F̂ =
Ĥ, p̂′A + d3 r P̂A (r) × B̂′ (r)
~
Z
i
d
= Ĥ, p̂′A +
d3 r P̂A (r) × B̂′ (r). (102)
~
dt
To evaluate the different contributions to the first term
in Eq. (102), we first recall Eq. (90) and note that
Z
Z
1
i 1
3
2
′
d r P̂A (r), p̂A =
d3 r ∇P̂A2 (r) = 0.
~ 2ε0
2ε0
(103)
Further, we derive, on recalling Eq. (100),
"
#
2
Z
i X 1
p̂′α + d3 r Ξ̂α (r) × B̂′ (r) , p̂′A
~ α 2mα
Z
i
Xh
= −∇A d3 r 21
r̂˙ α × Ξ̂α (r) − Ξ̂α (r) × r̂˙ α B̂′ (r).
α
(104)
Substituting Eqs. (103) and (104) into Eq. (102), with Ĥ
as given in Eq. (38), we eventually obtain
Z
F̂ = ∇A
d3 r P̂A (r)Ê′ (r)
Z
i
Xh
Ξ̂α (r) × r̂˙ α − r̂˙ α × Ξ̂α (r) B̂′ (r)
+ 21 d3 r
α
d
+
dt
Z
d3 r P̂A (r) × B̂′ (r).
(105)
It can be shown (see Appendix D) that Eq. (105) is identical to Eq. (94).
It is not difficult to see [recall Eqs. (41) and (43)] that
in long-wavelength approximation Eq. (105) takes the
10
form of Eq. (97), but with Ê′ (r̂A ) and B̂′ (r̂A ) in place
of Ê(r̂A ) and B̂(r̂A ), respectively. The time derivative
d[d̂ × B̂′ (r̂A )]/dt can then be calculated to give an expression of the form of Eq. (98) with B̂(r̂A ) replaced
by B̂′ (r̂A ). Obviously, in the nonrelativistic limit we are
left with an expression similar to Eq. (99). It should
be pointed out that Eqs. (97) and (99) with Ê(r̂A )
and B̂(r̂A ) replaced by Ê′ (r̂A ) and B̂′ (r̂A ), respectively,
yield exactly the same force as the equations with the
unprimed quantities, although the physical meaning of
Ê′ (r̂A ) is different from that of Ê(r̂A ) [recall that B̂′ (r̂A )
= B̂(r̂A )].
It is worth noting that the results of this section can
serve as an example to illustrate that the electric dipole
approximation has to be employed with great care. If
in electric dipole approximation the Röntgen interaction primarily related to the induction field had been
disregarded and Eq. (50) without the second term on
the right-hand side had been used, then in the resulting expression for the force the time-derivative term, i.e.,
the magnetic part of the force, would have been lost.
Note that the pressure exerted by external laser fields on
macroscopic bodies can be dominated by this magnetic
force [43, 44], which contrasts with arguments [32, 45]
that the contribution of this term to the radiation force
on atoms can be neglected.
e.g., Ref. [19]). Note that we are free to choose a convenient operator ordering in Eq. (106), because Ê′ (r, ω)
commutes with d̂.
A.
General case
In order to calculate the average force as a function
of time, we first formally integrate the Heisenberg equations of motion for the fundamental fields f̂λ′ (r, ω, t) to
obtain the source-quantity representation of Ê′ (r, ω, t).
The result reads (see Appendix E)
Ê′ (r, ω, t) = Ê′free (r, ω, t) + Ê′source (r, ω, t),
(109)
where
Ê′free (r, ω, t) = Ê′ (r, ω)e−iωt
(110)
and
Ê′source (r, ω, t)
Z
iµ0 2 t ′ −iω(t−t′ )
=
ω
dt e
Im G[r, r̂A (t′ ), ω]d̂(t′ ). (111)
π
0
Substituting Eq. (109) together with Eqs. (110) and (111)
into Eq. (106), we arrive at
F̂(t) = F̂free (t) + F̂source (t) ,
V.
(112)
where
AVERAGE LORENTZ FORCE
Let us now turn to the problem of determining the
electromagnetic force acting on an atomic system that
is initially prepared in an arbitrary internal (electronic)
quantum state. For convenience, we shall employ the multipolar formalism. On recalling Eqs. (21) and (22) together with Eq. (23), we find that Eq. (99) [with Ê(r̂A )
and B̂(r̂A ) replaced by Ê′ (r̂A ) and B̂′ (r̂A ), respectively]
can be rewritten as
Z ∞
i
h
F̂ =
dω ∇ d̂Ê′ (r, ω)
0
h
i
1 d
′
d̂ × ∇ × Ê (r, ω)
+
+ H.c., (106)
iω dt
r=r̂A
where Ê′ (r, ω) is defined according to Eq. (8). Decomposing F̂ into an average component hF̂i (where the expectation value h. . .i is taken with respect to the internal
atomic motion and the medium-assisted electromagnetic
field only) and a fluctuating component
∆F̂ = F̂ − F̂ ,
(107)
F̂ = F̂ + ∆F̂.
(108)
we may write
In the following, we will only consider the average force
hF̂i (for a discussion of the force fluctuation h∆F̂2 i, see,
Z
∞
dω ∇ d̂(t)Ê′free (r, ω, t)
0
1 d
+ H.c.
d̂(t) × ∇ × Ê′free (r, ω, t)
+
iω dt
r=r̂A (t)
F̂free (t) =
(113)
and
mag
F̂source (t) = F̂el
source (t) + F̂source (t) .
(114)
Here,
F̂el
source (t) =
iµ0
π
Z
∞
dω ω 2
Z
t
′
dt′ e−iω(t−t )
0
0
′
′
×∇ d̂(t)Im G[r, r̂A (t ), ω]d̂(t )
+ H.c.
r=r̂A (t)
(115)
is the electric part of the average force associated with the
source-field part of the medium-assisted electromagnetic
field, and
Z ∞
Z
µ0
d t ′ −iω(t−t′ )
dω
ω
dt e
F̂mag
(t)
=
source
π 0
dt 0
× d̂(t)× ∇×Im G[r, r̂A (t′ ), ω] d̂(t′ )
+ H.c.
r=r̂A (t)
(116)
11
is the respective magnetic part. Equations (112)–(116)
are still general in the sense that they apply to both
driven and nondriven atomic systems and to both weak
and strong atom-field coupling.
B.
Nondriven atom in the weak-coupling regime
. . . Ê′free [r̂A (t), ω, t] = Ê′†
free [r̂A (t), ω, t] . . . = 0,
(117)
then hF̂free (t)i = 0. Consequently, the average force, referred to as CP force, is determined by the source-field
part only,
′
̺ˆ = |{0 }ih{0 }| ⊗ σ̂,
(119)
where the density operator of the internal motion of the
atomic system σ̂ can be written as
X
σ̂ =
(120)
σmn Âmn
m,n
(Âmn = |mihn|, with |ni, |mi being the internal atomic
energy eigenstates). In order to calculate the dipoledipole correlation function appearing in Eqs. (115) and
(116), we make use of the expansion
X
d̂(t) =
(121)
dmn Âmn (t)
m,n
and write
d̂(t) ⊗ d̂(t′ )
X X
=
dmn ⊗ dm′ n′ Âmn (t)Âm′ n′ (t′ ) . (122)
m,n m′ ,n′
In the weak-coupling regime, the Markov approximation can be exploited and the correlation functions
hÂmn (t)Âm′ n′ (t′ )i can be calculated by means of the
quantum regression theorem (see, e.g., Ref. [46]). For
this purpose, the (intra-atomic) master equation has to
be solved for arbitray initial conditions, which in general requires knowledge of the specific level structure of
the atomic system under consideration. Only if the relevant atomic transition frequencies are well separated
from each other, one can go a step forward constructing
a general solution. In this case, the off-diagonal densitymatrix elements can be regarded as being decoupled from
each other and from the diagonal elements. We find (see
Appendix F)
Âmn (t)Âm′ n′ (t′ ) = δnm′ Âmn′ (t′ )
′
× e{iω̃mn (r̂A )−[Γm (r̂A )+Γn (r̂A )]/2}(t−t )
(124)
are the body-induced position-dependent shifted transition frequencies [r̂A = r̂A (t)], where
X
k
δωm (r̂A ) =
δωm
(r̂A ),
(125)
with
k
δωm
(r̂A )
µ0
=
P
π~
(123)
Z
∞
dω ω 2
0
dkm ImG(1) (r̂A , r̂A , ω)dmk
,
ω̃mk (r̂A ) − ω
(126)
and
Γm (r̂A ) =
(118)
Even more specifically, we assume that the density operator of the initial quantum state of the field and the
internal (electronic) motion of the atomic system reads
′
ω̃mn (r̂A ) = ωmn + δωm (r̂A ) − δωn (r̂A )
k
When the atomic system is not driven, i.e.,
F̂(t) = F̂source (t) .
(t ≥ t′ , m 6= n). Here,
X
Γkm (r̂A )
(127)
k
are the position-dependent level widths, with
Γkm (r̂A ) =
2µ0
Θ[ω̃mk (r̂A )][ω̃mk (r̂A )]2
~
×dkm ImG [r̂A , r̂A , ω̃mk (r̂A )]dmk . (128)
One should point out that the position-independent (infinite) Lamb-shift terms resulting from G(0) (r̂A , r̂A , ω)
[recall Eq. (60)] have been thought to be absorbed in
the transitions frequencies ωmn . Equation (126) can be
rewritten by changing to imaginary frequencies [cf. the
discussion below Eq. (63)], resulting in
µ0
Θ[ω̃mk (r̂A )][ω̃mk (r̂A )]2
~
× dkm Re G(1) [r̂A , r̂A , ω̃mk (r̂A )]dmk
Z
dkm G(1) (r̂A , r̂A , iu)dmk
µ0 ∞
.
du u2 ω̃km (r̂A )
+
π~ 0
[ω̃km (r̂A )]2 + u2
(129)
k
δωm
(r̂A ) = −
Recall that in the perturbative treatment the vdW potential of an atomic system in a state |mi is identified
with the energy shift ~δωm , so it is not surprising that
Eq. (125) together with Eq. (129) corresponds to Eq. (65)
together with Eqs. (66) and (67), if in Eq. (129) the ω̃mk
are replaced with ωmk . The calculation of
Âmn (t) = σnm (t)
(130)
[σnm (0) = σnm ] then leads (under the assumptions made)
to
σnm (t) = e{iω̃mn (r̂A )−[Γm (r̂A )+Γn (r̂A )]/2}t σnm
(131)
for m 6= n [cf. Eq. (123)], so the remaining task consists
in solving the balance equations
X
σ̇mm (t) = −Γm (r̂A )σmm (t) +
Γm
n (r̂A )σnn (t). (132)
n
With these preparations at hand, the CP force can
be calculated in the following steps. We first substitute Eq. (122) together with Eqs. (123) and (130) into
12
Eqs. (115) and (116) and perform the time derivative in
Eq. (116). Introducing slowly varying density-matrix elements σ̃nm (t) = eiω̃nm t σnm (t), we then perform the time
integrals in the spirit of the Markov approximation, by
making the replacements σ̃nm (t′ ) 7→ σ̃nm (t) as well as
r̂A (t′ ) 7→ r̂A (t) and letting the upper limit of integration tend to infinity. Recalling Eq. (118) together with
Eq. (114), we derive
X
F̂(t) =
σnm (t)Fmn (r̂A ),
(133)
This result requires two comments. First, in Eqs. (135)
and (136) the replacement G(r, r̂A , ω) 7→ G(1) (r, r̂A , ω)
has again been made, which can be justified by similar arguments as in Sec. III [cf. the discussion preceding
Eq. (62)]. Second, from the derivation of Eqs. (133)–(136)
it is clear that these equations are valid provided that
the center-of-mass motion can be regarded as being sufficiently slow. More precisely, they hold if the condition
G[r, r̂A (t+∆t), ω] ≈ G[r, r̂A (t), ω] for ∆t ≤ Γ−1
(137)
C
m,n
Fmn (r̂A ) =
Fel
mn (r̂A )
+
Fmag
mn (r̂A ),
(134)
where
Fel
mn (r̂A )
=
(
Z
µ0 X ∞
dω ω 2
π
0
k
∇ ⊗ dmk Im G(1) (r, r̂A , ω)dkn
×
ω + ω̃kn (r̂A ) − i[Γk (r̂A ) + Γm (r̂A )]/2
)
+ H.c.,
r=r̂A
(135)
and
(
Z
µ0 X ∞
mag
dω ω ω̃mn (r̂A )
Fmn (r̂A ) =
π
0
k
)
dmk × ∇ × Im G(1) (r, r̂A , ω) dkn
×
ω + ω̃kn (r̂A ) − i[Γk (r̂A ) + Γm (r̂A )]/2
(
~µ0
=−
2π
Fel,r
mn (rA )
=
(
µ0
Z
X
k
∞
0
Ωmnk (rA ) = ω̃nk (rA ) + i[Γm (rA ) + Γk (rA )]/2,
we derive
(138)
el,or
el,r
Fel
mn (rA ) = Fmn (rA ) + Fmn (rA ),
mag,or
mag,r
Fmag
(rA ) + Fmn
(rA ),
mn (rA ) = Fmn
(139)
(140)
+ H.c.
r=r̂A
(136)
Fel,or
mn (rA )
is satisfied, where ΓC is a characteristic intra-atomic decay rate. Under this condition, the internal (electronic)
and external (center-of-mass) motion of the atomic system decouple in the spirit of a Born-Oppenheimer approximation. As a result, r̂A effectively enters the equations as a parameter, so that the caret will be removed
in the following (r̂A 7→ rA ).
We finally rewrite Eqs. (135) and (136), by using contour integration and going over to imaginary frequencies
[cf. the discussion below Eq. (63)]. Recalling the definition of αmn (ω) = αmn (rA , ω) as given in Eq. (69) and
introducing the abbreviating notation
where
)
(1)
du u (αmn )ij (rA , iu) + (αmn )ij (rA , −iu) ∇Gij (r, rA , iu)
2
Θ(ω̃nk )Ω2mnk (rA )∇
,
(141)
r=rA
⊗ dmk G
(1)
[r, rA , Ωmnk (rA )]dkn
)
+ H.c.
(142)
r=rA
and
Fmag,or
(rA )
mn
=
(
~µ0
2π
Fmag,r
mn (rA ) =
(
)
ω̃mn (rA ) ⊤
ω̃mn (rA ) ⊤
(1)
αmn (rA , iu) −
αmn (rA , −iu) × ∇ × G (r, rA , iu)
du u Tr
iu
iu
Z
∞
µ0
X
0
2
,
r=rA
(143)
k
Θ(ω̃nk )ω̃mn (rA )Ωmnk (rA )dmk × ∇ × G(1) [r, rA , Ωmnk (rA )]dkn
[(Tr T )j = Tljl ]. Equation (133) together with Eq. (134)
)
+ H.c.
(144)
r=rA
and Eqs. (139)–(144) is the natural generalization of
13
Eq. (68) together with Eqs. (65), (67), and (71). The
above result is the first nonperturbative expression for
the CP force that incorporates its time dependence in
case of excited atoms and correctly accounts for bodyinduced shifting and broadening of atomic transition
lines.
In the short-time limit, ΓC t ≪ 1, Eq. (133) reads
X
F̂(t) ≃ F̂(0) =
σnm (0)Fmn (rA ),
(145)
m,n
which for σnm (0) = δnl δml reduces to
F̂(t) ≃ F̂(0) = Fel
ll (rA ).
(146)
For the nonrelativistic Hamiltonian (46), we can always
choose real dipole matrix elements (dmn = dnm ), revealing that dmn ⊗ dnm is a symmetric tensor so that, recalling Eq. (17), we may exploit the rule
(1)
(1)
Sij ∇Gij (r, r, ω) = 2Sij ∇s Gij (s, r, ω)|s=r ,
(147)
which is valid for any symmetric tensor S. Hence,
Eqs. (141) and (142) [together with Eq. (70)] lead to
Z
~µ0 ∞
Fel,or
(r
)
=
−
duu2 (αl )ij (rA , iu)
A
ll
4π 0
(1)
(148)
+ (αl )ij (rA , −iu) ∇A Gij (rA , rA , iu)
and
n
µ0 X
Θ(ω̃lk )Ω2lk (rA ) ∇
2
k
o
(1)
+ H.c.
⊗ dlk G [r, r, Ωlk (rA )]dkl
Fel,r
ll (rA ) =
(149)
r=rA
[Ωlk (rA ) ≡ Ωllk (rA )]. Ignoring the position-dependent
shifts and broadenings of the atomic energy levels, i.e.,
disregarding the position dependence of the atomic po(0)
larizability [αl (rA , iu) 7→ αl (iu)], Eqs. (148) and (149)
reduce to the perturbative result in Eq. (68) together
with Eqs. (65), (67), and (71) [Fel
ll (rA ) 7→ Fl (rA )]. Note
that this result can be obtained without choosing real
dipole matrix elements [αl (iu) + αl (−iu) being symmetric in this case]. In the long-time limit, ΓC t≫1, Eq. (133)
obviously reduces to ground-state force
X
F̂(t) ≃
σnm (∞)Fmn (rA ) = Fel,or
(150)
00 (rA )
m,n
[Fel,r
00 (rA ) = 0], because of σnm (∞) = δn0 δm0 .
As already mentioned, the expression for the groundstate CP force F00 (rA ) obtained in lowest-order perturbation theory, Eq. (77), agrees with the expression obtained from LRT. However, its naive extrapolation in
(0)
the sense of the replacement α0 (ω) 7→ α0 (rA , ω) in
Eq. (77) [25] is wrong, because it results in Eq. (148) with
2(α0 )ij (rA , iu) instead of (α0 )ij (rA , iu)+(α0 )ij (rA , −iu).
As a result, a noticeable influence of the level broadening
on the off-resonant part of the CP force is erroneously
predicted in Ref. [25] (cf. Sec. V C), thus demonstrating
that body-induced level broadening is a nonperturbative
effect which lies beyond the scope of the LRT approach
to the problem.
Equation (148) reveals that even the ground-state CP
force cannot be derived from a potential in the usual way,
because of the position dependence of the atomic polarizability. Nevertheless, it is a potential force, provided that
it is an irrotational vector, i.e.,
Z ∞
X
∂
2
∇A × F00 (rA ) =
duu
∇A ω̃k0 (rA )
∂ ω̃k0
0
k
∂
+ ∇A Γk (rA )
(α0 )ij (rA , iu)
∂Γk
(1)
+ (α0 )ij (rA , −iu) × ∇A Gij (rA , rA , iu) = 0. (151)
While for effectively one-dimensional problems (e.g., for
an atom in the presence of planarly, spherically, or cylindrically multilayered media) this condition is satisfied,
there are of course situations where it is violated, implying that Eq. (148) is inaccessible to perturbative methods
in principle.
When the atomic system is initially prepared in a coherent superposition of states such that σnm (0) 6= 0 is
valid for certain values n and m with n 6= m, then —
according to Eq. (133) — the corresponding off-diagonal
force components σnm (t)Fmn (rA ) can also contribute to
the total force acting on the atomic system. Interestingly,
such transient off-diagonal force components contain contributions not only from the electric part of the Lorentz
force but also from the magnetic part, as can be easily seen from inspection of Eqs. √
(143) and (144). Thus
an atomic qubit |ψi = (|0i + |1i)/ 2 (cf., e.g., Ref. [47])
near a body feels, in electric dipole approximation, both
an electric and a magnetic force in general.
Let us briefly comment on atomic systems displaying
(quasi)degeneracies, i.e., systems exhibiting transitions
with ωmn ≃ ωm′ n′ (m 6= m′ and/or n 6= n′ ). In such a case,
the assumption that the (relevant) off-diagonal densitymatrix elements decouple from each other as well as from
the diagonal ones can no longer be made. Let us assume
that the degenerate sublevels are not connected via electric dipole transitions (dmm′ = 0 if ωmm′ ≃ 0). The degeneracy related to the different possible projections of the
angular momentum of an atom (in free space) onto a chosen direction is a typical example. Taking into account
that the degeneracy is removed when the atom is close
to a body, it may be advantageous to change the basis
within each degenerate sublevel accordingly and consider
the master equation in the new basis. An equation of the
form of Eq. (123) is then valid in the new basis. Note that
the new basis will in general depend on the position of
the atom, thus introducing an additional position dependence of the CP force. While Eq. (131) also remains valid
in the new basis for ωmn 6= 0, this is not in general true
for the temporal evolution of the density-matrix elements
14
with ωmm′ ≃ 0 so that, instead of the balance equations
(132), a system of equations has to be solved in which diagonal density-matrix elements and off-diagonal elements
with ωmm′ ≃ 0 are coupled to each other.
To illustrate the effects of body-induced level shifting and broadening, let us consider a two-level atom
with (real) transition dipole matrix element dA ≡ d10 =
dA (cos φ sin θ ex + sin φ sin θ ey + cos θ ez ) (d00 = d11 = 0),
which is situated at position zA very close above (z > 0) a
semi-infinite half space (z < 0) containing a homogeneous
dispersing and absorbing magnetodielectric medium. Let
δω = δω1 − δω0 denote the (position-dependent) shift of
the transition frequency. Using the Green tensor in the
short-distance limit, from Eqs. (124), (125), and (129) we
derive (see Appendix G)
δω(zA ) = δωr (zA ) + δωor (zA ),
δωr (zA ) = −
C |ε[ω̃10 (zA )]|2 − 1
3 |ε[ω̃ (z )] + 1|2 ,
~zA
10 A
2C ω̃10 (zA )
δωor (zA ) =
3
~πzA
Z
0
(152)
δω/ω T
Example: Excited atom near an interface
0
-0.002
1.12
1.13
1.14
0.008
(b)
Γ/ω T
C.
(a)
0.002
0.004
(153)
∞
ε(iu) − 1
du
2 (z ) + u2 ε(iu) + 1 ,
ω̃10
A
(154)
0
1.12
1.13
ω10 /ωT
1.14
where
C=
d2A (1 + cos2 θ)
.
32πε0
(155)
Note that in the short-distance limit the medium effectively acts like a dielectric one. Since the relation ω̃10 =
ω10 + δω is valid, Eq. (152) together with Eqs. (153) and
(154) is a highly transcendental equation for the determination of δω. To solve it, we first note that the offresonant term δωor may be neglected in most practical
situations. For example, for a single-resonance medium
of Drude-Lorentz type,
ε(ω) = 1 +
2
ωP
2 − ω 2 − iγω ,
ωT
(156)
and the parameter values in Fig. 1, one can easily verify
the inequality
2
δωor (zA )
CωP
≤
. 10−4 .
3
2
ω̃10 (zA )
2~zA ωT ω̃10 (zA )
(157)
Thus, keeping only the resonant part of the frequency
shift, we may set
δω(zA ) = −
C |ε[ω̃10 (zA )]|2 − 1
3 |ε[ω̃ (z )] + 1|2 .
~zA
10 A
(158)
For ε(ω̃10 ) from Eq. (156), Eq. (158) is a fifth-order poly-
FIG. 1: (a) Transition frequency shift (solid and dotted lines)
and (b) decay rate (solid and dotted lines) versus bare transition frequency for a two-level atom that is situated at distance
zA from a semi-infinite half space medium of complex permittivity according to Eq. (156) and whose transition dipole moment is perpendicular to the interface [ωP /ωT = 0.75, γ/ωT
2 2
= 0.01; ωT
dA /(3π~ε0 c3 ) = 10−7 ; zA /λT = 0.0075 (solid and
dashed lines), zA /λT = 0.009 (dotted and dot-dashed lines)].
For comparison, the approximate results obtained by using
the bare frequencies in Eqs. (158) and (159) are also displayed
(dashed and dot-dashed lines).
nomial conditional equation for δω, which may be solved
numerically. Having calculated δω, we may calculate the
(position-dependent) decay rate Γ ≡ Γ1 . Neglecting the
small free-space decay rate, we replace the Green tensor
by its scattering part as given by Eq. (G4), hence from
Eqs. (127) and (128) we obtain
Γ(zA ) =
4C Im ε[ω̃10 (zA )]
3 |ε[ω̃ (z )] + 1|2 .
~zA
10 A
(159)
The resonant part of the CP force on the excited atom
in the short-distance limit can be found by taking the
derivative of the scattering part of the Green tensor
[Eq. (G4)] with respect to zA and substituting the re-
15
r
sult into Eq. (149) (l = 1). We derive (Fr11 = F11
ez )
r
F11
(zA ) = −
3C |ε[Ω10 (zA )]|2 − 1
4 |ε[Ω (z )] + 1|2 ,
zA
10 A
(160)
where, according to Eq. (138),
Ω10 (zA ) = ω̃10 (zA ) + iΓ(zA )/2.
(161)
Using Eq. (156), we see that (γ, Γ ≪ ωT )
2
ωP
.
−
− i[Γ(zA ) + γ]ω̃10 (zA )
(162)
Equation (160) differs from the perturbative result in
two respects. First, the bare atomic transition frequency
ω10 is replaced with the (position-dependent) shifted frequency ω̃10 . Second, the absorption parameter γ of the
ε[Ω10 (zA )] = 1 +
CP force
3
2
ωT
2 (z )
ω̃10
A
(a)
α1 (zA , iu) + α1 (zA , −iu) = −
ω̃10 (zA )
2
ω̃10 (zA ) + [u + Γ(zA )/2]2
ω̃ 2 (zA ) + u2 + Γ2 (zA )/4
× 210
ω̃10 (zA ) + [u − Γ(zA )/2]2
1.12
3
4dA ⊗ dA
~
×
0
-3
CP force
medium is replaced with the sum of γ and the (positiondependent) atomic decay rate Γ. The sum γ + Γ obviously
plays the role of the total absorption parameter.
The dependence of δω and Γ on ω10 in the shortdistance limit is shown in Figs. 1(a) and (b), respectively, and Fig. 2 displays the resonant part of the CP
force as a function of ω10 . From Fig. 2 it is seen that in
the vicinity
of the (surface-plasmon induced) frequency
p
2 + ω 2 /2 an enhanced force is observed, which
ωS = ωT
P
is attractive (repulsive) for red (blue) detuned atomic
transition frequencies ω10 < ωS (ω10 > ωS ) — a result already known from perturbation theory (dashed curves in
the figure). However, it is also seen that due to bodyinduced level shifting and broadening the absolute value
of the force can be noticeably reduced (solid curves in
the figure). Interestingly, the positions of the extrema of
the force remain nearly unchanged, because level shifting
and broadening give rise to competing effects that almost
cancel.
In order to calculate the off-resonant part of the CP
force on the excited atom in the short-distance limit, we
first note that, according to Eq. (69),
1.13
1.14
(b)
0
.
(163)
Substituing Eq. (163) into Eq. (148) and making use of
Eq. (G7) [where f (u) is given by u2 times Eq. (163)], we
or
derive (For
11 = F11 ez )
Z ∞
3C
ε(iu) − 1
or
F11
(zA ) =
du
4
πzA 0
ε(iu) + 1
ω̃10 (zA )
× 2
ω̃10 (zA ) + [u + Γ(zA )/2]2
ω̃ 2 (zA ) + u2 + Γ2 (zA )/4
× 210
. (164)
ω̃10 (zA ) + [u − Γ(zA )/2]2
Note that for a two-level atom the relation
or
or
F00
(zA ) = −F11
(zA )
-3
1.12
1.13
ω10 /ω T
1.14
r
FIG. 2: The resonant part of the CP force F11
λ4T ×10−9 /(3C)
on a two-level atom that is situated at distance (a) zA /λT =
0.0075 and (b) zA /λT = 0.009 of a semi-infinite half space
medium of complex permittivity according to Eq. (156) and
whose transition dipole moment is perpendicular to the interface (solid lines). The parameters are the same as in Fig. 1.
For comparison, both the perturbative result (dashed lines)
and the separate effects of level shifting (dotted lines) and
level broadening (dash-dotted lines) are shown.
(165)
is valid.
In Fig. 3, the off-resonant part of the CP force is shown
as a function of the bare atomic transition frequency.
Obviously, the shift of the transition frequency has the
effect of raising and lowering the perturbative values of
the force (dashed curves) for ω10 < ωS and ω10 > ωS , respectively, which is in full agreement with the frequency
response of the frequency shift shown in Fig. 1(a). The
influence of the decay rate on the CP force is extremely
weak, as it can be seen from the insets in the figure.
This may be understood by the fact that in contrast to
the case of the resonant part of the CP force, where the
16
0.0174
−8
6 10
−8
CP force
3 10
0
1.12
0.0173
0.0172
1.13
1.14
VI.
(a)
1.12
1.13
0.00838
1.14
−8
1 10
−9
CP force
5 10
0
1.12
0.00834
0.0083
1.13
1.14
(b)
1.12
by about two orders of magnitude. However, this observation should be considered with great care. While the
two-level atom is a good model for calculating the resonant part of an atom in an excited state, such a simplification is not justified in general when all higher levels can
contribute to the off-resonant force component. However,
provided that the convergence of the corresponding sum
is sufficiently fast, we can still conclude that the resonant
part of the CP force is dominant.
1.13
ω10 /ω T
1.14
r
FIG. 3: The off-resonant part of the CP force F11
λ4T × 10−9
/(3C) on a two-level atom that is situated at distance (a)
zA /λT = 0.0075 and (b) zA /λT = 0.009 of a semi-infinite half
space medium of complex permittivity according to Eq. (156)
and whose transition dipole moment is perpendicular to the
interface (solid lines). The parameters are the same as in
Fig. 1. For comparison, the perturbative result (dashed lines)
is shown. The insets display the difference between the force
with and without consideration of the level broadening (solid
lines). For comparison, we show this difference in the case
where the level shifts are ignored (dashed lines).
decay rate enters directly via the Green tensor, the influence on the off-resonant part is more indirect via the
atomic polarizability. Due to the specific dependence on
the atomic polarizability, the leading-order dependence
is quadratic in Γ and not linear in Γ as erroneously predicted from LRT [25]. Physically, the weak influence of
the level broadening on the off-resonant part of the CP
force may be regarded as being a consequence of the fact
that this part corresponds to energy nonconserving processes (the energy denominators being nonzero), which
implies that they happen on (extremely short) time scales
where real photon emission does not play a role.
Comparing the magnitudes of the resonant and offresonant components of the CP force, we see that the
off-resonant component is smaller than the resonant one
SUMMARY
Basing on electromagnetic-field quantization that allows for the presence of dispersing and absorbing linear media, and starting with the Lorentz force acting
on a neutral atom, we have extended the concept of CP
force beyond the well-known results derived on the basis of normal-mode quantization or LRT in leading order
of pertubation theory to allow for (i) magnetodielectric
bodies, (ii) an atom that is initially prepared in an arbitrary internal (electronic) quantum state, thereby being
subjected to a time-dependent force, (iii) the position dependence of the force via the atomic response, and (iv)
arbitrary strength of the atom-field coupling. The basic
formulas also apply to the calculation of the radiation
forces arising from excited fields such as the force acting
on a driven atom.
For a first analysis, we have restricted our attention to
a nondriven atom in the weak-coupling regime, so that
the internal atomic dynamics can be treated in Markov
approximation. It turns out that the force is a superposition of force components weighted by the time-dependent
intra-atomic density-matrix elements that solve the intraatomic master equation. Each force component is expressed in terms of the Green tensor of the electromagnetic field and the atomic polarizability, which — through
the position-dependent energy level shifts and broadenings — now depends on the position of the atomic system.
In consequence even the force components resulting from
the electric part of the Lorentz force cannot be derived
from potentials in the usual way. Clearly, the position dependence of the atomic polarizability become noticeable
only for very small atom-body separations. In order to
illustrate the effect, we have considered a two-level atom
in the vicinity of a planar semi-infinite medium.
When the atomic system is initially prepared in an
eigenstate of its internal Hamiltonian, then only force
components associated with diagonal density-matrix elements appear. They solely result from the electric part of
the Lorentz force and reduce to the CP forces obtained
in lowest-order perturbation theory if the atomic polarizability is replaced with its position-independent perturbative expression. Force components that are associated
with excited intra-atomic energy levels are of course transient. As in the course of time an initially excited level
is depopulated and lower lying levels are populated, the
force that initially acts on the atomic system in the ex-
17
To transform the momenta of the charged particles, the
identities
cited state changes with time to the force that acts on
the atomic system in the ground state.
The results further show that when the atomic system
is initially prepared in an intra-atomic quantum state
that is a coherent superposition of energy eigenstates,
then additional force components associated with the corresponding off-diagonal density-matrix elements are observed. Thus an atomic qubit would typically feel such
off-diagonal force components. It should be pointed out
that not only the electric but also the magnetic part
of the Lorentz force can contribute to the off-diagonal
force components, with the magnetic contributions being proportional to the transition frequencies. Clearly,
off-diagonal force components are transient.
In contrast to the transient force components that are
associated with excited energy levels, off-diagonal force
components carry an additional harmonic time dependence. Clearly, if the oscillations are too fast, it can be
difficult to detect them experimentally, since they may
effectively average to zero. In this case it may be advisable to assign them to the fluctuating part of the force
rather than to the average force. The situation may be
different in cases where strong atom-field coupling (not
considered here) gives rise to Rabi oscillations.
where Ξ̂α (r) is defined as in Eq. (39). Further, the following quantities remain unchanged under the transformation (A1), because they commute with both Â(r) (cf.
Ref. [37]) and r̂α ,
Acknowledgments
Â′ (r) = Â(r), B̂′ (r) = B̂(r), ϕ̂′ (r) = ϕ̂(r),
(A9)
′
′
′
′
r̂α = r̂α , r̂A = r̂A , ρ̂A (r) = ρ̂A (r), ϕ̂A (r) = ϕ̂A (r), (A10)
S.Y.B. acknowledges valuable discussions with O. P.
Sushkov as well as M.-P. Gorza. This work was supported
by the Deutsche Forschungsgemeinschaft. S.Y.B. is grateful for being granted a Thüringer Landesgraduiertenstipendium.
APPENDIX A: DERIVATION OF THE
MULTIPOLAR HAMILTONIAN (38)
To perform transformations of the type
Ô′ = Û Ô Û † ,
(A1)
with Û being given by Eq. (36) together with Eq. (37),
we apply the operator identity
1
eŜ Ôe−Ŝ = Ô + Ŝ, Ô +
Ŝ, Ŝ, Ô + . . . .
2!
(A2)
Recalling the commutation relations (4) and (5), it is not
difficult to prove that the basic fields f̂ (r, ω) are transformed as
Z
1
′
∗ ′
f̂λ′ (r, ω) = f̂λ (r, ω) +
d3 r′ P̂⊥
A (r )Gλ (r , r, ω). (A3)
~ω
Using Eq. (A2) together with the commutation relation
⊥
(r − r′ ), cf. Ref. [37], we find that
[ε0 Êk (r), Âl (r)] = i~δkl
Ê′ (r) = Ê(r) +
1 ⊥
P̂ (r).
ε0 A
(A4)
∇α δ(r − r̂A − λˆr̄β )
mα
= λ−1
(A5)
− λδαβ ∇δ(r − r̂A − λˆr̄β ),
mA
Z 1
dλ ˆr̄α ∇δ(r−r̂A −λˆr̄α ) = δ(r−r̂A ) − δ(r−r̂α ), (A6)
0
Z
0
1
dλλˆr̄α ∇δ(r − r̂A − λˆr̄α )
= −δ(r − r̂α ) +
Z
0
1
dλ δ(r − r̂A − λˆr̄α )
(A7)
are helpful. They can be proved with the aid of the definitions (28) and (29), and via (partial) integration with
respect to λ. Using Eqs. (A5)–(A7) we derive
Z
′
p̂α = p̂α − qα Â(r̂α ) − d3 rΞ̂α (r) × B̂(r),
(A8)
P̂′A (r) = P̂A (r), Θ̂′α (r) = Θ̂α (r), Ξ̂′α (r) = Ξ̂α (r). (A11)
Applying the transformation rules (A3), (A8), and
(A9)–(A11), we may now express the minimal-coupling
Hamiltonian (1) in terms of the transformed variables.
Recalling Eq. (21) together with Eqs. (8)–(10) and making use of the relations (19) and
Z ∞
π
ω
(A12)
dω 2 Im G(r, r′ , ω) = δ(r − r′ )
c
2
0
(cf. Ref. [36]), we derive
Z ∞
X Z
Ĥ =
d3 r
dω ~ω f̂λ′† (r, ω)f̂λ′ (r, ω)
0
λ=e,m
+
1
2ε0
X
Z
′⊥
d3 r P̂′⊥
A (r)P̂A (r) −
Z
′⊥
d3 r P̂′⊥
A (r)Ê (r)
2
Z
1
′
3
′
′
+
p̂α + d r Ξ̂α (r) × B̂ (r)
2mα
α
Z
Z
+ 12 d3 r ρ̂′A (r)ϕ̂′A (r) + d3 r ρ̂′A (r)ϕ̂′ (r). (A13)
In order to simplify the last two terms of Eq. (A13), we
′k
recall Eq. (86) as well as P̂A (r) = ε0 ∇ϕ̂′A (r) and Ê′k (r)
′
= −∇ϕ̂ (r), obtaining with the aid of partial integration
Z
Z
3
′
′
1
d r ρ̂A (r)ϕ̂A (r) + d3 r ρ̂′A (r)ϕ̂′ (r)
2
Z
Z
= 12 d3 r P̂′A (r)∇ϕ̂′A (r)+ d3 r P̂′A (r)∇ϕ̂′ (r)
Z
Z
1
′k
3
′
d r P̂A (r)P̂A (r)− d3 r P̂′A (r)Ê′k (r). (A14)
=
2ε0
18
Combining Eqs. (A13) and (A14), and noting that integrals containing mixed products of transverse and longitudinal vector fields vanish, we obtain Eq. (38), where
we have made use of Eqs. (A10) and (A11) and hence
dropped the primes of all quantities containing the particle coordinates only.
In the simpler case in which the center-of-mass coordinate is treated as a parameter, the transformation law
(A8) changes to
Z
(A15)
p̂′α = p̂α − qα Â(r̂α ) − d3 r Θ̂α (r) × B̂(r).
Equations (A3), (A4), and (A9)–(A11) remain formally
the same, provided that the replacement r̂A 7→ rA is made.
APPENDIX B: ORDERS OF MAGNITUDE OF
INTERACTION TERMS
To estimate the order of magnitude of atom-field interactions, let us introduce the typical atomic length and
energy scales
~
aB
=
,
Zeff
Zeff α0 me c
Z 2 ~2
2
2
E0 = Zeff
ER = eff 2 ≈ Zeff
13.6eV
2me aB
a0 =
(B1)
(B2)
(aB , Bohr radius; ER , Rydberg energy), where me and
−e are the electron mass and charge, respectively, Zeff e
is the typical effective nucleus charge felt by the electrons
giving the main contributions to the interaction terms to
be calculated, and α0 = e2 /(4πε0 ~c) is the fine-structure
constant. As a rough estimate we can then make the replacements
qα → e,
ˆr̄α → a0 ,
mα → m e ,
r̂˙ A → v,
ωkl → E0 /~,
(B3)
ˆ (′)
p̄
α → p = me E0 a0 /~ (B4)
[for the last replacement, see Eq. (C7)]. With regard to
the length scale of variation of the medium-assisted electromagnetic field we may make the replacements
∇ → λ−1 ∼ ω/c, ∇ϕ̂ → ∇ϕ ∼ ωA,
Ê(′) → E ∼ ωA, B̂(′) → B ∼ (ω/c)A
(B5)
(B6)
(Â(′) → A). Noting that materials typically become
transparent for frequencies that are greater than 20 eV
(cf. Ref. [48]),
ε(r, ω) ≈ 1 ⇒ G(1) (r, r′ , ω) ≈ 0 for ~ω & 20 eV, (B7)
we should require that
~ω . 20 eV
⇒
~ω
. 1.
E0
(B8)
With these approximations at hand, the orders of magnitude of ∆1 E defined by Eq. (55) and ∆2 E defined by
Eq. (54) in Sec. III A can be estimated to be
∆1 E ∼
and
e2 a20 A2
e2 A2
=
E0 = g 2 E0 = O g 2
2me
~2
(B9)
2 2 2
e p A
1
ea0 ∇ϕepA
2
2 2
∇ϕ
+
e
a
+
2
0
E0 + ~ω
m2e
me
"
2 #
~ω
E0
~ω
= g2 1 + 2
+
= O g2 ,
E0
E0
1 + ~ω/E0
∆2 E ∼
(B10)
where the dimensionless coupling constant
g ≡ ea0 A/~
(B11)
has been introduced. Note that in Eq. (B10) we have
approximated p̂α → p, because in Sec. III we treat an
atom at rest, hence relative and absolute momenta are
identical.
In order to give a rough idea of the magnitude of the
coupling constant g, we need to estimate the magnitude
of the field strength A. In the context of the current work
we consider interactions of an atomic system with the
vacuum electromagnetic field, so the relevant quantity is
the vacuum fluctuation of the field strength. Recalling
Eqs. (8) and (21) and making use of the commutation
relations (4) and (5) as well as the integral relation (19),
we find
h[∆Ê(rA )]2 i = h{0}|Ê2 (rA )|{0}i − h{0}|Ê(rA )|{0}i2
Z ∞
~
ω ′2
=
(B12)
dω ′ 2 ImTrG(rA , rA , ω ′ ).
πε0 0
c
When the atomic system is placed sufficiently far away
from all macroscopic bodies, a good estimate for the integral can be given by using the vacuum Green tensor
ImG(0) (r, r, ω) = ω/(6πc)I, leading to
h[∆Ê(rA )]2 i ∼
~ω 4
,
6π 2 ε0 c3
(B13)
where ω is a characteristic frequency contributing to the
interaction, cf. Eq. (B8). Hence making the replacement
s
~ω 2
A∼
(B14)
6π 2 ε0 c3
[cf. Eq. (B6)], we find
r
α0 ~ω
g ∼ Zeff
α0 ∼ 10−2 ,
6π E0
(B15)
depending on the specific atomic system considered and
the characteristic frequencies of the medium. When the
atom is situated close to some macroscopic body, the
19
scattering Green tensor becomes much larger than the
vacuum Green tensor, and the approximation leading to
Eq. (B13) is not valid anymore. The increased value of
the coupling constant g is reflected by the failure of the
perturbative result for small atom-surface separations.
The orders of magnitude of the contributions of the
three terms in Eq. (52) to the eigenvalue shift in Sec. III B
can be estimated according to
X qα
p̂′α × ˆr̄α B̂′ (rA )
2m
α
α
X qα
ˆ α B̂(r̂A ) × ˆr̄α
p̄
2m
α
α
∼
(B16)
∼
~(ωkl + ω)
2
2
E0
ea0 pB
~ω
= (Zeff α0 g)2 41
∼
2me
1 + ~ω/E0
E0
E0
(B17)
= O (Zeff α0 g)2 ,
×
1+~ω/E0
X qα
˙
ˆ α − qα Â(r̂A ) × B̂(r̂A )
p̄
d̂ × B̂(r̂A ) =
m
α
α
2
epB
e AB
ωE0
∼
= O(g),
= g(1 + 2g)
+
me
me
c
(B25)
X q2
2
(ea0 B)2
α
ˆr̄α × B̂′ (rA ) ∼
8mα
8me
α
2
~ω
E0 = O (Zeff α0 g)2 . (B18)
= (Zeff α0 g)2 18
E0
Next, let us estimate the orders of magnitude of
the various contributions to the Lorentz force given in
Sec. IV A. The magnitudes of the first and third terms in
curly brackets in Eq. (97) can be approximated according
to
1
2
r̂˙ A B̂(′) (r̂A ) × d̂
∼ 21 ea0 Bv =
1
2
v
g (~ω) = O(gv/c).
c
e2 a0 AB
= (Zeff α0 g 2 ) 12 (~ω) = O(Zeff α0 g 2 ), (B24)
2me
and Eq. (B20), we see that the magnitude of the second term in curly brackets in Eq. (97) is O(Zeff α0 g +
Zeff α0 g 2 + gv/c) = O[(Zeff α0 + v/c)g]. The magnitudes
of the different contributions to Eq. (98) are
2
d̂Ê(r̂A ) ∼ ea0 E = g (~ω) = O(g),
(B19)
˙
d̂ × B̂(r̂A )
∼ ea0 ωB = g
1
2
~ω
E0
ωE0
c
= O(g),
(B26)
h
←
− i
d̂ × r̂˙ A ∇A ⊗ B̂(r̂A ) + B̂(r̂A ) ⊗ ∇A r̂˙ A
v ~ω ωE
ea0 vωB
0
∼
= O(gv/c).
=
g
c
c
E0
c
(B27)
Finally, let us compare the contributions of the Röntgen interaction to the temporal evolution f̂λ (r, ω, t) with
that from the electric dipole interaction,
(B20)
1
2~ω
In order to estimate the magnitude of the second term,
we make use of the relation
mα r̂˙ α = p̂α − qα Â(r̂α )
eBa0 p
= (Zeff α0 g) 41 (~ω) = O(Zeff α0 g), (B23)
2me
X q2
α ˆ
r̄α B̂(r̂A ) × Â(r̂A )
2m
α
α
2
d̂Ê′ (rA )
E0
∼ (ea0 E)2
~(ωkl + ω)
1 + ~ω/E0
2
E0
~ω
= O g2 ,
= g2
E0
1 + ~ω/E0
Combining this with
i
∗
~ d̂(t)Gλ [r̂A (t), r, ω]
vea . ea
0
0
∼
= O(v/c),
~c
~
(B21)
in order to introduce relative momenta [recall Eq. (49)],
leading to
X qα
X qα
ˆ α B̂(r̂A ) × ˆr̄α
r̂˙ α B̂(rA ) × ˆr̄α =
p̄
2
2m
α
α
α
X q2
α ˆ
r̄α B̂(r̂A ) × Â(r̂A ) + 21 r̂˙ A B̂(r̂A ) × d̂.
+
2m
α
α
(B22)
r̂˙ A (t)d̂(t) × ∇A × G∗λ [rA (t), r, ω]
1
~ωmA d̂(t)
(B28)
× B̂[rA (t), t]d̂(t) × ∇A × G∗λ [rA (t), r, ω]
i
∗
~ d̂(t)Gλ [r̂A (t), r, ω]
me
~ω
e2 a20 B . ea0
2 1
=
(Z
α
)
g
∼
eff 0
2
~mA c
~
E0
mA
2
= O (Zeff α0 ) g .
(B29)
20
APPENDIX C: CALCULATION OF THE
PERTURBATIVE CORRECTIONS (58) AND (79)
Recalling Eq. (6) together with Eqs. (8)–(10), making
use of the commutation relations (4) and (5), and applying Eq. (19), Eq. (55) leads to
X q2 X Z ∞
α
∆1 El =
dω
2m
α
α
λ=e,m 0
Z
1
× d3 r 2 ⊥ Gλ ij (rA , r, ω) ⊥ G∗λ ij (rA , r, ω)
ω
Z ∞
~µ0 X qα2
dωIm ⊥ G⊥ ii (rA , rA , ω), (C1)
=
π α 2mα 0
where we have introduced the notation
⊥(k)
G⊥(k) (r, r′ , ω)
Z
Z
3
≡ d s d3 s′ δ ⊥(k) (r − s)G(s, s′ , ω)δ ⊥(k) (s′ − r′ ).
(C2)
Applying the sum rule
X q2
1 X
α
I=
ωkl (dlk ⊗ dkl + dkl ⊗ dlk ), (C3)
2mα
2~
α
k
we can rewrite Eq. (C1) as
Substituting Eqs. (C5) and (C6) into Eq. (54), we then
derive
Z ∞
Z
1X X
dω
∆2 El = −
P
d3 r(dlk )i
~
ωkl + ω
0
k λ=e,m
h
k
× (dkl )j Gλ in (rA , r, ω) k G∗λ jn (rA , r, ω)
ωkl k
Gλ in (rA , r, ω) ⊥ G∗λ jn (rA , r, ω)
−
ω
ωkl ⊥
−
Gλ in (rA , r, ω) k G∗λ jn (rA , r, ω)
ω
i
2
ωkl
+ 2 ⊥ Gλ in (rA , r, ω) ⊥ G∗λ jn (rA , r, ω)
ω
Z ∞
µ0 X
dω
=
dlk − ω 2 Im k Gk (rA , rA , ω)
P
π
ωkl +ω
0
k
h
i
+ ωkl ω Im k G⊥ (rA , rA , ω) + Im ⊥ Gk (rA , rA , ω)
2
− ωkl
Im ⊥ G⊥ (rA , rA , ω) dkl ,
(C8)
where we have again made use of the identity (19).
Adding Eqs. (C4) and (C8) according to Eq. (57), on
using the identity G = ⊥ G⊥ + ⊥ Gk + k G⊥ + k Gk [which
directly follows from the definition (C2) together with
δ(r) = δ k (r) + δ ⊥ (r)], we eventually arrive at Eq. (58).
The derivation of Eq. (79) is completely analogous. The
relevant matrix elements can be calculated with the aid
of Eq. (21) together with Eqs. (8)–(10) and the commutation relations (4) and (5), cf. the remarks below Eq. (40).
The result is
−hl|h{0′ }|d̂Ê′ (rA )|{1′λ (r, ω)}i|ki = −dlk Gλ (rA , r, ω).
(C9)
Z
µ0 X ∞
∆1 El =
dωωkl dlk Im⊥ G⊥ (rA , rA , ω)dkl .
π
0
k
(C4)
To calculate ∆2 E, as given by Eq. (54), we first calculate
the matrix elements therein. Recalling Eqs. (4)–(10), we
obtain
Substituting Eq. (C9) into Eq. (78) yields
Z ∞
Z
1X X
dω
P
d3 r
~
ωkl + ω
0
k λ=e,m
× (dlk )i (dkl )j Gλ in (rA , r, ω) G∗λ jn (rA , r, ω),
∆2 El = −
(C10)
hl|h{0}|d̂∇ϕ̂(r)r=rA |{1λ (r, ω)}i|ki
k
= −dlk Gλ (rA , r, ω),
(C5)
X qα
−hl|h{0}|
p̂α Â(rA )|{1λ (r, ω)}i|ki
mα
α
ωkl
dlk ⊥ Gλ (rA , r, ω),
(C6)
=
ω
where the second matrix element has been obtained by
means of the identity
X qα
hl|p̂α |ki = −iωkl dlk .
mα
α
(C7)
from which Eq. (79) follows by means of Eq. (19).
APPENDIX D: EQUIVALENCE OF LORENTZ
FORCES (94) AND (105)
To transform the first term in Eq. (105), we apply the
the rule (A4), recall that integrals over mixed products
of transverse and longitudinal vector fields vanish, and
use the identity for the first term in Eq. (A14) as well as
21
the commutation relations (4) and (5), we obtain
Eqs. (82) and (103). We thus derive
Z
3
∇A d r P̂A (r)Ê (r)
Z
Z
1
= ∇A d3 r P̂A (r)Ê(r) + ∇A d3 r P̂A (r)P̂⊥
A (r)
ε0
Z
Z
1
3
= ∇A d r P̂A (r)Ê(r) + ∇A d3 r P̂2A (r)
ε0
Z
− ∇A d3 r ρ̂A (r)ϕ̂A (r)
Z
= ∇A d3 r P̂A (r)Ê(r).
(D1)
In order to simplify the second term in Eq. (105), we use
the definitions (29), (39), (88), and (89) to calculate
Xh
1
2
α
=
X
1
2
−
1
2
which can be integrated to yield [f̂λ′ (r, ω, 0) ≡ f̂λ′ (r, ω)]
h
i
qα Θ̂α (r) × r̂˙ α − r̂˙ α × Θ̂α (r)
X
β
f̂λ′ (r, ω, t) = f̂λ′ free (r, ω, t) + f̂λ′ source (r, ω, t),
h
i
qα Θ̂α (r) × r̂˙ A − r̂˙ A × Θ̂α (r)
i
h
+ 21 P̂A (r) × r̂˙ A − r̂˙ A × P̂A (r)
= M̂A (r) + M̂R (r).
f̂λ′ free (r, ω, t) = e−iωt f̂λ′ (r, ω)
(D2)
Consequently, recalling that B̂ (r) = B̂(r), we may write
i
Xh
Ξ̂α (r) × r̂˙ α − r̂˙ α × Ξ̂α (r) B̂′ (r)
d3 r 12
= ∇A
α
Z
d3 r M̂A (r) + M̂R (r) B̂(r)
(E4)
and
f̂λ′ source (r, ω, t) =
i
~
Z
0
t
′
dt′ e−iω(t−t ) d̂(t′ ) G∗λ [r̂A (t′ ), r, ω].
(E5)
Substituting Eqs. (E3)–(E5) into Eq. (8) [Ê(r, ω, t) 7→
Ê′ (r, ω, t)] and using the identity (19), we arrive at
Eqs. (109)–(111).
(D3)
as well as
Z
d
d3 r P̂A (r) × B̂′ (r) =
d3 r P̂A (r) × B̂(r) .
dt
(D4)
Substituting Eqs. (D1), (D3), and (D4) into Eq. (105),
we see that Eq. (105) is equivalent to Eq. (94).
d
dt
(E3)
where
′
∇A
The third and fourth terms in Eq. (E1), which are due to
the Röntgen interaction, are smaller than the second one
by factors of v/c and g(Zeff α0 )2 , respectively [Eqs. (B28)
and (B29) in Appendix B], so according to the nonrelativistic approximation, Eq. (E1) reduces to
i
˙′
f̂λ (r, ω, t) = −iω f̂λ′ (r, ω, t) + d̂(t)G∗λ [r̂A (t), r, ω], (E2)
~
i
Ξ̂α (r) × r̂˙ α − r̂˙ α × Ξ̂α (r)
α
Z
i
˙′
f̂λ (r, ω, t) = Ĥ, f̂λ′ (r, ω, t)
~
i
′
= −iω f̂λ (r, ω, t) + d̂(t)G∗λ [r̂A (t), r, ω]
~
1 ˙
−
r̂A (t)d̂(t) × ∇A × G∗λ [r̂A (t), r, ω]
2~ω
1 n
−
d̂(t) × B̂′ [r̂A (t), t]d̂(t)
~ωmA
o
× ∇A × G∗λ [r̂A (t), r, ω] . (E1)
′
Z
APPENDIX E: EQUATIONS OF MOTION FOR
f̂λ′ (r, ω, t)
In electric dipole approximation, the temporal evolution of the basic fields f̂λ′ (r, ω, t) is governed by the Hamiltonian given in Eq. (44) together with Eqs. (45), (46), and
(50). Using Eqs. (8) and (21)–(23) (with the unprimed
fields being replaced with the primed ones) and applying
APPENDIX F: INTRA-ATOMIC EQUATIONS OF
MOTION
An estimation similar to that given for the fields
f̂ ′ (r, ω, t) shows that in the nonrelativistic limit the second term in the interaction Hamiltonian in electric dipole
approximation (50) can be disregarded in the calculation
of the temporal evolution of the intra-atomic operators
Âmn (t). By representing the (unperturbed) intra-atomic
Hamiltonian in the form of Eq. (53), recalling Eqs. (8)
and (121), and applying standard commutation relations,
it is not difficult to prove that the Âmn (t) obey the equations of motion
i
˙
Âmn = Ĥ, Âmn = iωmn Âmn
~
Z
∞
iX
+
dnk Âmk −dkm Âkn
dω Ê′ (r̂A , ω)
~
0
k
Z ∞
′†
+
dω Ê (r̂A , ω) dnk Âmk −dkm Âkn . (F1)
0
22
We now substitute the source-quantity representation for
Ê′ (r̂A , ω) = Ê′ [r̂A (t), ω, t] (and its Hermitian conjugate)
according to Eqs. (109)–(111) into Eq. (F1). Carrying
out the time integral in the source-field part in Eq. (F1)
in the Markov approximation, we may set, on regarding
r̂A = r̂A (t) as being slowly varying,
Z ∞
X
gmn (r̂A )Âmn ,
(F2)
dω Ê′source (r̂A , ω) =
0
m,n
where
iµ0
gmn (r̂A ) =
π
Z
∞
i
dnk gkn (r̂A ) = −iδωnk (r̂A ) − 21 Γkn (r̂A ),
~
where δωnk (r̂A ) and Γkn (r̂A ), respectively, are defined according to Eqs. (126) and (128) [with G(1) (r̂A , r̂A , ω) instead of G(r̂A , r̂A , ω) in Eq. (126)], Eqs. (F4) and (F6)
lead to Eqs. (123), (131), and (132).
dω ω 2 Im G(r̂A , r̂A , ω)dmn
× ζ[ω̃nm (r̂A )−ω]
(F3)
n
o
i X
∗
iωmn +
Âmn
dnk gkn − dkm gkm
~
APPENDIX G: HALF SPACE MEDIUM
The equal-position scattering Green tensor for a semiinfinite half space which contains for z < 0 a homogeneous, dispersing, and absorbing magnetodielectric
medium reads for z > 0 [49]
k
+B̂mn + F̂mn ,
(F4)
with
B̂mn =
iX
i X
dnk gkl Âml −
dkm gnl Âkl
~
~
k,l6=n
k,l
i X
iX
∗
∗
dnk gml
Âlk −
dkm gkl
Âln (F5)
+
~
~
k,l
G(1) (r, r, ω) =
1 0
× rs 0 1
0 0
where
Z ∞
q 2iβ0 z
i
e
dq
8π 0
β0
2
−β
0
0
0
2
0
c
0 + rp 2
0 −β02
0 , (G1)
ω
0
0
0 2q 2
k,l6=m
(m 6= n), and
i X
˙
∗
Âmm
dmk gkm − dkm gkm
Âmm =
~
k
i X
∗
−
Âkk
dkm gmk − dmk gmk
~
+Ĉmm + F̂mm ,
(F6)
with
o
i X n
∗
Âlm
dmk gkl Âml − dkm gkl
=
~
k,l6=m
o
i Xn
∗
−
Âlk , (F7)
dkm gml Âkl − dnk gml
~
k,l6=k
where F̂mn denotes contributions from the free-field part
in Eq. (109). Taking expectation values with respect
to the internal atomic motion and the medium-assisted
electromagnetic field, with the density-matrix given by
Eq. (119), we can use the property (117), finding that
the terms F̂mn do not contribute. In the absence of (quasi)degeneracies such that
|ω̃mn − ω̃m′ n′ | ≫ 21 |Γm + Γn − Γm′ − Γn′ |,
rs =
µβ0 − β
,
µβ0 + β
rp =
εβ0 − β
εβ0 + β
(G2)
are the reflection coefficients for s- and p-polarized waves,
2
respectively (β02 = ω 2 /c2 − q 2 with Im β0 > 0, βp
= εµω 2/c2
2
− q with Im β > 0). For q ≫ |ω|/c and q ≫ |εµ||ω|/c,
respectively, the approximations
k
Ĉmm
(F9)
0
[ζ(x) = πδ(x) + iP/x], with ω̃nm (r̂A ) being the shifted transition frequencies. Substituting Eq. (F2) into
Eq. (F1), we obtain
˙
Âmn =
we may disregard couplings between different off-diagonal transitions and between off-diagonal and diagonal
transitions and thus omit the terms B̂mn and Ĉmm , hence
upon using the decomposition
(F8)
β0 ≃ iq,
β ≃ iq
(G3)
can be made. Due to the exponential factor the integration interval is effectively
p limited to values q . 1/z. In the
short-distance limit z |εµ||ω|/c ≪ 1, we therefore introduce a small error, if we extrapolate the approximations
(G3) to the whole integral, resulting in
1 0 0
2
c
ε(ω)
−
1
0 1 0 .
G(1) (r, r, ω) =
32πω 2 z 3 ε(ω) + 1 0 0 2
(G4)
Note that the magnetic properties of the medium represented by the permeability µ begin to contribute via
terms proportional to 1/z. Substitution of Eq. (G4)
into the first term of Eq. (129) for δωnk = δω10 yields
Eq. (153).
In order to obtain Eq. (154), we recall Eq. (G1) to
23
write
Z ∞
du f (u)G(1) (r, r, iu)
0
Z ∞
Z ∞
1
duf (u)
db0 e−2b0 z
=
8π 0
u/c
2
b0 0
0
1 0 0
2
c
0
× rs 0 1 0 − rp 2 0 b20
u
0 0 0
0 0 2b20 − ( uc )2
, (G5)
of b0 . The frequency integral effectively extends up to
frequencies of the order c/z, hence in the short-range
limit zωM/c ≪ 1 (⇒ c/z ≫ ωM ) we introduce only a
small error by extrapolating this approximation to the
whole frequency integral. Performing the b0 integral, retaining only leading-order terms in uz/c (in consistency
with zωM /c ≪ 1) we derive
Z
∞
duf (u)G(1) (r, r, iu)
0
having changed the integration variable to the imaginary
part of β0 (β0 =ib0 ). Let ωM be a characteristic frequency
of the medium such that
(G6)
1
0
0
f (u) ε(iu) − 1
0 1 0 . (G7)
du 2
u ε(iu) + 1 0 0 2
For u > ωM , the approximation β ∼ β0 holds, and consequently the reflection coefficients rs , rp are independent
2
Using Eq. (G7) [with f (u) = u2 /(ω̃A
+ u2 )] together with
Eq. (129), we obtain Eq. (154).
[1] G. Binnig, C. F. Quate, and Ch. Gerber, Phys. Rev. Lett.
56, 930 (1986).
[2] F. Shimizu and J. Fujita, Phys. Rev. Lett. 88, 123201
(2002).
[3] I. E. Dzyaloshinskii, E. M. Lifshitz, and L. P. Pitaevskii,
Adv. Phys. 10, 165 (1961); D. Langbein, Springer Tracts
Mod. Phys. 72, 1 (1974); J. Mahanty and B. W. Ninham, Dispersion Forces (Academic, London, 1976); E.
A. Hinds, in Advances in Atomic, Molecular, and Optical
Physics, edited by D. Bates and B. Bederson (Academic,
New York, 1991), Vol. 28, p. 237; P. W. Milonni, The
Quantum Vacuum: An Introduction to Quantum Electrodynamics (Academic, San Diego, 1994).
[4] D. Raskin and P. Kusch, Phys. Rev. 179, 712 (1969);
A. Shih, D. Raskin, and P. Kusch, Phys. Rev. A 9, 652
(1974); A. Shih and V. A. Parsegian, ibid. 12, 835 (1975).
[5] A. Anderson, S. Haroche, E. A. Hinds, W. Jhe, and D.
Meschede, Phys. Rev. A 37, 3594 (1988); C. I. Sukenik,
M. G. Boshier, D. Cho, V. Sandoghdar, and E. A. Hinds,
Phys. Rev. Lett. 70, 560 (1993).
[6] R. E. Grisenti, W. Schöllkopf, J. P. Toennies, G. C.
Hegerfeldt, and T. Köhler, Phys. Rev. Lett. 83, 1755
(1999).
[7] F. Shimizu, Phys. Rev. Lett. 86, 987 (2001); V. Druzhinina and M. DeKieviet, ibid. 91, 193202 (2003).
[8] H. Friedrich, G. Jacoby, and C. G. Meister, Phys. Rev.
A 65, 032902 (2002).
[9] A. Landragin, J.-Y. Courtois, G. Labeyrie, N. Vansteenkiste, C. I. Westbrook, and A. Aspect, Phys. Rev. Lett.
77, 1464 (1996).
[10] M. Oria, M. Chevrollier, D. Bloch, M. Fichet, and M.
Ducloy, Europhys. Lett. 14, 527 (1991); V. Sandoghdar,
C. I. Sukenik, E. A. Hinds, and S. Haroche, Phys. Rev.
Lett. 68, 3432 (1992); M. Marrocco, M. Weidinger, R. T.
Sang, and H. Walther, ibid. 81, 5784 (1998); H. Failache,
S. Saltiel, M. Fichet, D. Bloch, and M. Ducloy, ibid. 83,
5467 (1999); M. A. Wilson, P. Bushev, J. Eschner, F.
Schmidt-Kaler, C. Becher, R. Blatt, and U. Dorner, ibid.
91, 213602 (2003).
[11] M. Gorlicki, S. Feron, V. Lorent, and M. Ducloy, Phys.
Rev. A 61, 013603 (1999); R. Marani, L. Cognet, V.
Savalli, N. Westbrook, C. I. Westbrook, and A. Aspect,
ibid. 61, 053402 (2000).
H. B. G. Casimir and D. Polder, Phys. Rev. 73, 360
(1948).
R. K. Bullough and B. V. Thompson, J. Phys. C 3, 1780
(1970).
M. J. Renne, Physica (Amsterdam) 53, 193 (1971); 56,
124 (1971).
P. W. Milonni and M.-L. Shih, Phys. Rev. A 45, 4241
(1992); see also J. Schwinger, L. L. DeRaad, and K. A.
Milton, Ann. Phys. (N. Y.) 115, 1 (1978) for a related
treatment.
Y. Tikochinsky and L. Spruch, Phys. Rev. A 48, 4223
(1993); F. Zhou and L. Spruch, ibid. 52, 297 (1995).
M. Boström and B. E. Sernelius, Phys. Rev. A 61, 052703
(2000). Note that the formulas derived in Ref. [16] for
frequency-independent, real permittivities are used for
studying metals, by allowing for complex permittivities,
without any proof.
A. M. Marvin and F. Toigo, Phys. Rev. A 25, 782 (1982).
In fact an energy formula based on a normal-mode expansion is combined with elements of LRT.
C.-H. Wu, C.-I. Kuo, and L. H. Ford, Phys. Rev. A 65,
062102 (2002).
S. Y. Buhmann, Ho Trung Dung, and D.-G. Welsch, J.
Opt. B: Quantum Semiclassical Opt. 6, 127 (2004).
A. D. McLachlan, Proc. R. Soc. London Ser. A 271, 387
(1963); Mol. Phys. 7, 381 (1963).
G. S. Agarwal, Phys. Rev. A 11, 243 (1975).
J. M. Wylie and J. E. Sipe, Phys. Rev. A 30, 1185 (1984);
32, 2030 (1985).
C. Girard, J. Chem. Phys. 85, 6750 (1986); C. Girard
and C. Girardet, ibid. 86, 6531 (1987); C. Girard, S.
Maghezzi, and F. Hache, ibid. 91, 5509 (1989).
S. Kryszewski, Mol. Phys. 78, 5, 1225 (1993).
M. Fichet, F. Schuller, D. Bloch, and M. Ducloy, Phys.
Rev. A 51, 1553 (1995); M.-P. Gorza, S. Saltiel. H.
Failache, and M. Ducloy, Eur. Phys. J. D 15, 113 (2001).
M. Boustimi, J. Baudon, P. Candori, and J. Robert,
ε(iu) − 1 ≪ 1 for u > ωM .
2
=−
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
c
32πz 3
Z
0
∞
24
Phys. Rev. B 65, 155402 (2002).
[28] C. Henkel, K. Joulain, J.-P. Mulet, and J.-J. Greffet, J.
Opt. A, Pure Appl. Opt. 4, 109 (2002).
[29] Ho Trung Dung, L. Knöll, and D.-G. Welsch, Phys. Rev.
A 64, 013804 (2001).
[30] I. V. Bondarev, G. Ya. Slepyan, and S. A. Maksimenko
Phys. Rev. Lett. 89, 115504 (2002).
[31] W. P. Healy, J. Phys. A 10, 279 (1977).
[32] C. Baxter, M. Babiker, and R. Loudon, Phys. Rev. A 47,
1278 (1993).
[33] V. E. Lembessis, M. Babiker, C. Baxter, and R. Loudon,
Phys. Rev. A 48, 1594 (1993).
[34] M. Wilkens, Phys. Rev. A 47, 671 (1993); 49, 570 (1994).
[35] J.-C. Guillot and J. Robert, J. Phys. A 35, 5023 (2002).
[36] L. Knöll, S. Scheel, and D.-G. Welsch, in Coherence and
Statistics of Photons and Atoms, edited by J. Peřina
(Wiley, New York, 2001), p. 1; for an update, see
arXiv:quant-ph/0006121.
[37] Ho Trung Dung, S. Y. Buhmann, L. Knöll, D.-G. Welsch,
S. Scheel, and J. Kästel, Phys. Rev. A 68, 043816 (2003).
[38] V. G. Veselago, Sov. Phys. Usp. 10, 509 (1968).
[39] V. M. Fain and Ya. I. Khanin, Quantum Electronics
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
(MIT Press, Cambridge MA, 1969). Note that here no
distinction is made between bare and shifted transisition frequencies; see also P. W. Milonni and R. W. Boyd,
Phys. Rev. A 69, 023814 (2004).
I. B. Khriplovich and S. K. Lamoureaux, CP Violation
without Strangeness (Springer, Berlin, 1997).
D. P. Craig and T. Thirunamachandran Molecular Quantum Electrodynamics (Academic Press, New York, 1984).
W. Röntgen, Annu. Rev. Phys. Chem. 35, 264 (1888).
J. P. Gordon, Phys. Rev. A 8, 14 (1973).
R. Loudon, Phys. Rev. A 68, 013806 (2003).
S. Stenholm, Rev. Mod. Phys. 58, 699 (1986).
W. Vogel, D.-G. Welsch, and S. Wallentowitz, Quantum
Optics, An Introduction (Wiley-VCH, Berlin, 2001).
L.-M. Duan, J. I. Cirac, and P. Zoller, Science 292, 1695
(2001).
S. Adachi, Optical Constants of Crystalline and Amorphous Semiconductors (Kluwer, Boston, 1999).
W. C. Chew, Waves and Fields in Inhomogeneous Media (IEEE Press, New York, 1995) Secs. 2.1.3, 2.1.4, and
7.4.2.