Introduction To Decoherence Theory
Introduction To Decoherence Theory
Klaus Hornberger
arXiv:quant-ph/0612118v1 14 Dec 2006
Let the quantum state of a the system be described by the density oper-
ator ρ on the Hilbert space H. We take the system to interact with a single
environmental degree of freedom at a time–think of a phonon, a polaron, or a
gas particle scattering off your favorite implementation of a quantum register.
Moreover, let us assume, for the time being, that this environmental “particle”
is in a pure state ρE = |ψin ihψin |E , with |ψin iE ∈ HE . The scattering opera-
tor Stot maps between the in- and out-asymptotes in the total Hilbert space
Htot = H ⊗ HE , and for sufficiently short-ranged interaction potentials we
may identify those with the states before and after the collision. The initially
uncorrelated system and environment turn into a joint state,
Since the Sn are unitary the diagonal elements, or populations, are indeed
unaffected,
Stot and ρE to be arbitrary and carry out the same analysis. P Performing the
trace (5) in the eigenbasis of the environmental state, ρE = ℓ pℓ |ψℓ ihψℓ | E ,
we have
†
X
ρ′ = trE Stot [ρ ⊗ ρE ]Stot = pℓ hψj |Stot |ψℓ iE ρhψℓ |S†tot |ψj iE
j,ℓ
X
= Wk ρWk† , (8)
k
where the hψj |Stot |ψℓ i are operators in H. After subsuming the two indices
j, ℓ into a single one, we get the second line with the Kraus operators Wk
given by
√
Wk = pℓk hψjk |Stot |ψℓk i (9)
Projective measurements
and after the measurement of α the state of the quantum system is given by
the normalized projection
α Pα ρPα
ρ → M (ρ|α) = . (12)
tr (Pα ρ)
The basic requirement that the projectors form a resolution of the identity
operator,
X
Pα = I, (13)
α
hAi = tr (Aρ) .
Generalized measurements
The operators Mα,k appearing in Eq. (16) are called measurement operators,
and they serve to characterize the measurement process completely. The Fα
are sometimes called “effects” or “measurement elements”. Note that different
measurement operators Mα,k can lead to the same measurement element Fα .
A simple class of generalized measurements are unsharp measurements,
where a number of projective operators are lumped together with probabilis-
tic weights in order to account for the finite resolution of a measurement device
or for classical noise in its signal processing. However, generalized measure-
ments schemes may also perform tasks which are seemingly impossible with
a projective measurement, such as the error-free discrimination of two non-
orthogonal states [12, 13].
Efficient measurements
M ρ M†α
M (ρ|α) = α , (18)
tr M†α Mα ρ
implying that pure states are mapped to pure states. In a sense, these are
measurements where no unnecessary, that is no classical uncertainty is in-
troduced during the measurement process, see the following Sect. 1.3(b). By
means of a (left) polar decomposition and the consistency requirement (17)
efficient measurement operators have the form
p
Mα = Uα Fα , (19)
with an unitary operator Uα . This way the state after efficient measurement
can be expressed in a form which decomposes the transformation into a “raw
measurement” described by the Fα and a “measurement back-action” given
by the Uα . √ √
Fα ρ Fα
M (ρ|α) = Uα U†α (20)
|{z} tr (Fα ρ) |{z}
back-action | {z } back-action
“raw measurement”
8 Klaus Hornberger
√
In this transformation the positive operators Fα “squeeze” the state
along the measured property and expand it along the other, complementary
ones, similar to what a projector would do, while the back-action operators
Uα “kick” the state by transforming it in a way that is reversible, in principle,
provided the outcome α is known. Note that the projective measurements (12)
are a sub-class in the set of back-action-free efficient measurements.
This shows that an indirect measurement is efficient (as defined above) if the
probe is initially in a pure state, i.e., if there is no uncertainty introduced
in the measurement process, apart form the one imposed by the uncertainty
relations on ρprobe .
If we know that an indirect measurement has taken place, but do not know
its outcome α we have to resort to a probabilistic (Bayesian) description of the
new system state. It is given by the sum over all possible outcomes weighted
by their respective probabilities,
X X
ρ′ = Prob(α|ρ)M (ρ|α) = Mα,k ρM†α,k . (26)
α α,k
This form is the same as above in Eqs. (8) and (9), where the basic effect of
decoherence has been described. This indicates that the decoherence process
can be legitimately viewed as a consequence of the information transfer from
the system to the environment. The complementarity principle can then be
invoked to understand which particular system properties lose their quantum
behavior, namely those complementary to the ones revealed to the environ-
ment. This “monitoring interpretation” of the decoherence process will help
us below to derive microscopic master equations.
1.4 A few words on nomenclature
Let us take a two-level system, or qubit, described by the Pauli spin operator
σz , and model the environment as a collection of bosonic field modes. In
practice, such fields can yield an appropriate effective description even if the
actual environment looks quite differently, in particular if the environmental
coupling is a sum of many small contributions2 . What is fairly non-generic in
the present model is the type of coupling between system and environment,
which is taken to commute with the system Hamiltonian.
The total Hamiltonian thus reads
~ω X X †
Htot = σz + ~ωk b†k bk +σz gk bk + gk∗ bk (27)
2
k k
| {z } | {z }
H0 Hint
2
A counter-example would be the presence of a degenerate environmental degree
of freedom, such as a bistable fluctuator.
Introduction to decoherence theory 11
with the usual commutation relation for the mode operators of the bosonic
field modes, [bi , b†k ] = δik , and coupling constants gk . The fact that the sys-
tem Hamiltonian commutes with the interaction, guarantees that there is no
energy exchange between system and environment so that we expect pure
dephasing.
By going into the interaction picture one transfers the trivial time evo-
lution generated by H0 to the operators (and indicates this with a tilde). In
particular,
X
e int (t) = eiH0 t/~ Hint e−iH0 t/~ = σz
H (gk eiωk t b†k + gk∗ e−iωk t bk ) , (28)
k
where the second equality is granted by the commutation [σz , Hint ] = 0. The
time evolution due to this Hamiltonian can be formally expressed as a Dyson
series,
Z
e i t ′e ′
U(t) = T← exp − dt Hint (t )
~ 0
X∞ n Z t h i
1 1 e int (t1 ) · · · H
e int (tn ) ,
= dt1 · · · dtn T← H (29)
n=0
n! i~ 0
where T← is the time ordering operator (putting the operators with larger
time arguments to the left). Due to this time ordering requirement the series
usually cannot be evaluated exactly (if it converges at all). However, in the
present case the commutator of H e int at different times is not an operator, but
just a c-number,
X 2
e int (t) , H
[H e int (t′ )] = 2i |gk | sin ( ωk (t′ − t)) . (30)
k
One can now perform the integral over the interaction Hamiltonian to get
!
1 X †
e =e
U(t) iϕ(t)
exp σz ∗
αk (t)bk − αk (t)bk (33)
2
k
1 − eiωk t
αk (t) := 2gk . (34)
~ωk
e
The operator U(t) is diagonal in the eigenbasis of the system, and it describes
how the environmental dynamics depends on the state of the system. In par-
ticular, if the system is initially in the upper level, |ψi = | ↑i, one has
Y
e αk (t)
U(t)| ↑i|ξ0 iE = eiϕ(t) | ↑i Dk |ξ0 i =: eiϕ(t) | ↑i|ξ↑ (t)iE , (35)
2
k
Here we introduced the unitary displacement operator s for the k-th field mode,
which effect a translation of the field state in its attributed phase space. In
particular, the coherent state |αik of the field mode k is obtained from its
ground state |0ik by |αik := Dk (α)|0ik [15]
The equations (35) and (36) show that the collective state of the field
modes gets displaced by the interaction with the system and that the sense
of the displacement is determined by the system state.
„» – «
d iΦ(t) −iΦ(t) 1 e (t) e−iΦ(t) U (t) .
∂t U (t) = e e + eiΦ(t) H
dt i~
The derivative in square brackets has to be evaluated with care since the H e (t) do
not commute at different times. By first showing that [A, ∂t A] = c ∈ C
implies
∂t An = nAn−1 ∂t A − 12 n (n − 1) cAn−2 one finds
»Z t –
d iΦ(t) −1 iΦ(t) e 1 ` ´
e = e H (t) + 2 eiΦ(t) e t′ , H
dt′ H e (t) .
dt i~ 2~ 0
` 2 ´−1 R t ′
Therefore, we have ∂t U (t) = 2~ dt [He (t′ ) , H
e (t)] U (t), which can be inte-
0
grated to yield finally
„ Z t Z t1 h i«
e (t) = exp − 1
U dt 1 dt 2 He (t 1 ) , e
H (t 2 ) e−iΦ(t) .
2~2 0 0
Introduction to decoherence theory 13
Assuming that the states of system and environment are initially uncor-
related, ρtot (0) = ρ ⊗ ρE , the time-evolved system state reads4
e
ρ̃(t) = trE U(t)[ρ e † (t) .
⊗ ρE ]U (38)
It follows from (35) and (36) that the populations are unaffected,
while the coherences are suppressed by a factor which given by the trace over
the displaced initial field state,
!
Y
h↑ |ρ̃(t)| ↓i = h↑ |ρ̃(0)| ↓i trE Dk (αk (t))ρE . (39)
k
| {z }
χ(t)
Incidentally, the complex suppression factor χ (t) is equal to the Wigner char-
acteristic function of the original environmental state at the points αk (t), i.e.,
it is given the Fourier transform of its Wigner function [16].
For times that are short compared to the field dynamics, t ≪ ωk−1 , one ob-
serves a Gaussian decay of the coherences. Modifications to this become rele-
vant at ωk t ∼
2
= 1, provided χ (t) is then still appreciable, i.e., for 4 |gk | /~2 ωk2 ≪
1. Being a sum over periodic functions, χvac (t) is quasi-periodic, that is, it will
come back arbitrarily close to unity after a large period (which increases ex-
ponentially with the number of modes). These somewhat artificial Poincaré
recurrences vanish if we replace the sum over the discrete modes by an integral
over a continuum with mode density µ,
4
In fact, the assumption ρtot (0) = ρ ⊗ ρE is quite unrealistic if the coupling is
strong, as discussed below. Nonetheless, and it certainly represents a valid initial
state.
14 Klaus Hornberger
X Z ∞
f (ωk ) −→ dωµ(ω)f (ω), (41)
k 0
for any function f . This way the coupling constants gk get replaced by the
spectral density,
2
J(ω) = 4µ(ω) |g(ω)| . (42)
This function characterizes the environment by telling how effective the cou-
pling is at a certain frequency.
(b) Thermal state
This factor can be separated into its vacuum component (40) and a ther-
mal component, χ(t) = e−Fvac (t) e−Fth (t) , with the following definitions of the
vacuum and the thermal decay functions:
X 1 − cos (ωk t)
Fvac (t) := 4 |gk |2 (45)
~2 ωk2
k
X
2 1 − cos (ωk t) ~ωk
Fth (t) := 4 |gk | coth −1 (46)
~2 ωk2 2kB T
k
Assuming that the field modes are sufficiently dense we replace their sum by
an integration. Noting (41), (42) we have
Z ∞
1 − cos (ωt)
Fvac (t) −→ dωJ(ω) (47)
~2 ω 2
Z0 ∞
1 − cos(ωt) ~ω
Fth (t) −→ dωJ(ω) 2ω2
coth −1 . (48)
0 ~ 2kBT
5
This` can be found
´ in a small exercise `by using
´ the Baker-Hausdorff relation with
exp αb† − α∗ b = exp(−|α|2 /2) exp αb† exp (−α∗ b), and the fact that coher-
ent states satisfy the eigenvalue equation
√ b|βi = β|βi, have the number rep-
2 n
resentation hn|βi = exp(−|β| /2)β / n!, and form an over-complete set with
I R
= π −1 d2 β|β ihβ|.
Introduction to decoherence theory 15
a t
F (t) ≃ [for ω1−1 ≪ t]. (54)
~ 2 tT
In this thermal regime the decay shows the exponential behavior typical for
the Markovian master equations discussed below. Note that the decay rate
for this long time behavior is determined by the low frequency behavior of
the spectral density, characterized by the damping strength a in (49), and is
proportional to the temperature T .
For d = 3, the case of a “super-Ohmic” bath, the integrals (47), (48) can
be calculated without approximation. We note only the long-time behavior of
the decay.
2
kB T ′ kB T
lim F (t) = 2a ψ 1+ <∞ (55)
t→∞ ~ωc ~ωc
Here ψ (z) stands for the Digamma function, the logarithmic derivative of the
gamma function. Somewhat surprisingly, the coherences do not get completely
reduced as t → ∞, even at a finite temperature. This is due to the suppressed
influence of the important low-frequency contributions to the spectral density
in three dimensions (as compared to lower dimensions). While such a suppres-
sion of decoherence is plausible for intermediate times, the limiting behavior
(55) is clearly a result of our simplified model assumptions. It will be absent if
there is a small anharmonic coupling between the bath modes [18] or if there
is a small admixture of different couplings to Hint .
Let us now discuss the generalization to the case of N qubits, which do not
interact directly among each other. Each qubit may have a different coupling
to the bath modes.
N
X −1
~ωj (j) X
N
X −1 X (j) † (j)
Htot = σz + ~ωk b†k bk + σz(j) gk bk + [gk ]∗ bk (56)
j=0
2 j=0
k k
| {z }
H0
where the displacement of the field modes now depends on the N -qubit state.
As an example, take N = 2 qubits and only a single vacuum mode. For
the states
Introduction to decoherence theory 17
and
we obtain
(1)
e α (t) + α(2) (t) −α(1) (t) − α(2) (t)
U|φi|0iE = c11 | ↑↑i| iE + c00 | ↓↓i| iE
2 2
and similarly
(1)
e α (t) − α(2) (t) −α(1) (t) + α(2) (t)
U|ψi|0iE = c10 | ↑↓i| iE + c01 | ↓↑i| iE ,
2 2
where the α(1) (t) and α(2) (t) are the field displacements (34) due to the first
and the second qubit.
If the couplings to the environment are equal for both qubits, say, because
they are all sitting in the same place and seeing the same field, we have
α(1) (t) = α(2) (t) ≡ α (t). In this case states of the form |φi are decohered
2
once the factor hα (t) | − α (t)iE = exp(−2 |α (t)| ) ≃ 0. States of the form
|ψi, on the other hand, are not affected at all, and one says that the {|ψi}
span a (two-dimensional) decoherence-free subspace. It shows up because the
environment cannot tell the difference between the states | ↑↓i and | ↓↑i if it
couples only to the sum of the excitations.
In general, using the binary notation, e.g., | ↑↓↑i ≡ |1012 i = |5i, one has
where mj ∈ {0, 1} indicates the j-th digit in the binary representation of the
number m.
We can distinguish different limiting cases:
If the separation of the qubits is small compared to the wave lengths of the
(j)
field modes they are effectively interacting with the same reservoir, αk = αk .
One can push the j-summation to the α’s in thisPcase, so that, compared to
the single qubit, one merely has to replace αk by (mj − nj )αk . We find
2
PN −1
χmn (t) = exp − j=0 (mj − nj ) (Fvac (t) + Fth (t)) (60)
with Fvac (t) and Fth (t) given by Eqs. (47) and (48).
18 Klaus Hornberger
Hence, in the worst case, one observes an increase of the decay rate by
N 2 compared to the single qubit rate. This is the case for the coherence
between the states |0i and |2N − 1i, which have the maximum difference in
the number of excitations. On the other hand, the states with an equal number
of excitations form a decoherence-free
subspace in the present model, with a
N
maximal dimension of N/2 .
In the other extreme, the qubits are so far apart from each other that each
field mode couples only to a single qubit. This suggests a re-numbering of the
field modes,
(j)
αk −→ αkj ,
and leads, after transforming the j-summation into a tensor-product, to
NY−1 Y (kj )
χmn (t) = tr Dkj (mj − nj ) αkj (t) ρth
j=0 kj
N
Y −1
2
= exp −|mj − nj | (Fvac (t) + Fth (t))
j=0
| {z }
=|mj −nj |
NX −1
= exp − |mj − nj | (Fvac (t) + Fth (t)) . (61)
j=0
| {z }
Hamming distance
Hence, the decay of coherence is the same for all pairs of states with the
same Hamming distance. In the worst case, we have an increase by a factor
of N compared to the single qubit case, and there are no decoherence-free
subspaces.
An intermediate case is obtained if the coupling depends on the position r j
of the qubits. A reasonable model, corresponding to point scatterings of fields
(j)
with wave vector k, is given by gk = gk exp (ik · rj ), and its implications are
studied in [19].
The model for decoherence discussed in this section is rather exceptional
in that the dynamics of the system can be calculated exactly for some choices
of the environmental spectral density. In general, one has to resort to approx-
imate descriptions for the dynamical effect of the environment, and we turn
to this problem in the following section.
Introduction to decoherence theory 19
Isolated systems evolve, in the Schrödinger picture and for the general case of
mixed states, according to the von Neumann equation,
1
∂t ρ = [H, ρ]. (62)
i~
One would like to have a similar differential equation for the reduced dynamics
of an “open” quantum system, which is in contact with its environment. If we
extend the description to include the entire environment HE and its coupling
to the system, then the total state in Htot = H ⊗ HE evolves unitarily. The
partial trace over HE gives the evolved system state, and its time derivative
reads
d 1
∂t ρ = trE Utot (t)ρtot (0)U†tot (t) = trE ([Htot , ρtot ]) . (63)
dt i~
This exact equation is not closed and therefore not particularly helpful as it
stands. However, it can be used as the starting point to derive approximate
time evolution equations for ρ, in particular, if it is permissible to take the
initial system state to be uncorrelated with the environment.
These equations are often non-local in time, though, in agreement with
causality, the change of the state at each point in time depends only on the
state evolution in the past. In this case, the evolution equation is called a
generalized master equation. It can be specified in terms of superoperator-
functionals, i.e., linear operators which take the density operator ρ with its
past time-evolution until time t and map it to the differential change of the
operator at that time,
∂t ρ = K [{ρτ : τ < t}] . (64)
An interpretation of this dependence on the system’s past is that the environ-
ment has a memory, since it affects the system in a way which depends on
the history of the system-environment interaction. One may hope that on a
coarse-grained time-scale, which is large compared to the inter-environmental
correlation times, these memory effects might become irrelevant. In this case, a
proper master equation might be appropriate, where the infinitesimal change
of ρ depends only on the instantaneous system state, through a Liouville
super-operator L,
∂t ρ = Lρ. (65)
Master equations of this type are also called Markovian, because of their re-
semblance to the differential Chapman-Kolmogorov equation for a classical
Markov process. However, since a number of approximations are involved in
their derivation, it is not clear whether the corresponding time evolution re-
spects important properties of a quantum state, such as its positivity. We will
see that these constraints restrict the possible form of L rather strongly.
20 Klaus Hornberger
The number of the required Kraus operators Wk (t) is limited by the dimension
2
of the system Hilbert space, N 6 dim (H) (and confined to a countable set
in case of an infinite-dimensional, separable Hilbert space), but their choice is
not unique.
This statement is rather strong, and it is certainly violated for truly micro-
scopic times. But it seems not unreasonable on the level of a coarse-grained
time scale, which is long compared to the time it takes for the environment
to “forget” the past interactions with the system due to the dispersion of
correlations into the many environmental degrees of freedom.
For a given dynamical semigroup there exists, under rather weak condi-
tions, a generator, i.e., a superoperator L satisfying
In this case Wt (ρ) is the formal solution of the Markovian master equation
(65).
So far we used the Schrödinger picture, i.e., the notion that the state of an
open quantum system evolves in time, ρt = Wt (ρ0 ). Like in the description of
closed quantum systems, one can also take the Heisenberg point of view, where
the state does not evolve, while the operators A describing observables acquire
a time dependence. The corresponding map Wt♯ : A0 7→ At is called the dual
map, and it is related to Wt by the requirement tr(AWt (ρ)) = tr(ρWt♯ (A)).
In case of a dynamical semigroup, Wt♯ = exp L♯ t , the equation of motion
takes the form ∂t A = L♯ A, with the dual Liouville operator determined by
tr(AL (ρ)) = tr(ρL♯ (A)). From a mathematical point of view, the Heisenberg
picture is much more convenient since the observables form an algebra, and
it is therefore preferred in the mathematical literature.
7
The inverse element required for a group structure is missing for general, irre-
versible CPMs.
22 Klaus Hornberger
We can now derive the general form of the generator of a dynamical semigroup,
taking dim (H) = d < ∞ for simplicity [2, 22]. The bounded operators on H
then form a d2 -dimensional vector space which turns into a Hilbert space, if
equipped with the Hilbert-Schmidt scalar product (A, B) := tr(A† B).
Given an orthonormal basis of operators {Ej : 1 6 j 6 d2 } ⊂ L (H),
We can choose one of the basis operators, say the d2 -th, to be proportional
to the identity operator,
1
Ed2 = √ I (73)
d
so that all other basis elements are traceless,
0 for j = 1, . . . , d2 − 1
tr(Ej ) = √ (74)
d for j = d2 .
Representing the superoperator of the dynamical map (67) in the {Ej } basis
we have
d2
X
Wt (ρ) = cij (t)Ei ρE†j (75)
i,j=1
Wτ (ρ) − ρ
Lρ = lim
τ →0 τ
2 2
1 dX −1 dX−1
d cd d (τ )
2 2 −1 cjd2 (τ ) cd2 j (τ ) †
= lim ρ + lim √ Ej ρ + ρ lim √ Ej
{z τ dτ dτ
τ →0 τ →0 τ →0
| } j=1 j=1
c0 ∈ R | {z } | {z }
B∈L(H) B† ∈L(H)
2
dX −1
cij (τ )
+ lim Ei ρE†j
i,j=1 |
τ →0 τ
{z }
αij ∈ R
2
dX −1
= c0 ρ + Bρ + ρB + †
αij Ei ρE†j
i,j=1
2
dX −1
1 1
= [H, ρ] + (Gρ + ρG) + αij Ei ρE†j (77)
i~ ~ i,j=1
In the last equality the following hermitian operators with the dimension of
an energy were introduced:
~
G= (B + B† + c0 )
2
~
H = (B − B† )
2i
By observing that the conservation of the trace implies tr(Lρ) = 0 one can
relate the operator G to the matrix α = (αij ), since
2
dX −1
2G †
0 = tr(Lρ) = 0 + tr + αij Ej Ei ρ
~ i,j=1
This leads to the first standard form for the generator of a dynamical semi-
group:
dX2
−1
1 1 1
Lρ = [H, ρ] + αij Ei ρE†j − E†j Ei ρ − ρE†j Ei (78)
|i~ {z } i,j=1 2 2
unitary part | {z }
incoherent part
N
X
1 1 1
Lρ = [H, ρ] + γk Lk ρL†k − L†k Lk ρ − ρL†k Lk (79)
i~ 2 2
k=1
Lk → Lk + c k (80)
~ X ∗
H→H+ cj Lj − cj L†j + |c0 |, (81)
2i j
so that the Lk can be chosen traceless. In this case the only remaining freedom
is unitary rotation, X
√ √
γi Li −→ Uij γj Lj . (82)
j
8
It is easy to see that the dual Liouville operator discussed in Sect. 3.1(d) reads,
in Lindblad form,
XN „ «
♯ 1 † 1 † 1 †
L (A) = [A, H] + γk Lk ALk − Lk Lk A − ALk Lk .
i~ k=1
2 2
I I
Note that this implies L♯ ( ) = 0, while L ( ) =
P
k γk [Lk , L†k ] and tr (LX) = 0.
Introduction to decoherence theory 25
∞ n
X t
Wt = e(L0 +S)t = (L0 + S)n
n=0
n!
∞
X ∞
X
P
tn+ j kj
= P Lk0n SLk0n−1 S · · · SLk01 SLk00
n + | {z }
n=0 k0 ,...,kn =0 j j !
k
n times
X∞ Z t Z tn Z t2
= dtn dtn−1 · · · dt1
n=0 0 0 0
∞
X kn kn−1 k0
(t − tn ) (tn − tn−1 ) (t1 − 0)
× ···
kn ! kn−1 ! k0 !
k0 ,...,kn =0
k
×Lk0n SL0n−1 S · · · SLk01 SLk00
∞ Z t
X Z tn Z t2
= eL0 t + dtn dtn−1 · · · dt1
n=1 0 0 0
L0 (t−tn ) L0 (tn −tn−1 )
×e Se S · · · eL0 (t2 −t1 ) SeL0 t1 . (83)
P
The step from the second to the third line, where tn+ j kj /(n + j kj )! is
P
Lk ρ = γk Lk ρL†k , (84)
The latter has a negative imaginary part, Im (HC ) < 0, and can be used to
construct
1
L0 ρ = HC ρ − ρHC†
. (86)
i~
It follows that the sum of these superoperators yields the Liouville operator
(79)
N
X
L = L0 + Lk . (87)
k=1
26 Klaus Hornberger
PN
Of course, neither L0 nor S = k=1 Lk generate a dynamical semigroup.
Nonetheless, they are useful since the interpretation of Eq. (83) can now be
taken one step further. We can take the point of view that the Lk with k > 1
describe elementary transformation events due to the environment (“jumps”),
which occur at random times with a rate γk . A particular realization of n such
events is specified by a sequence of the form
The attributed times satisfy 0 < t1 < . . . . < tn < t, and the kj ∈ {1, . . . , N }
indicate which kind of event “took place”. We call Rnt a record of length n.
The general time evolution Wt can thus be written as an integration over
all possible realizations of the jumps, with the “free” evolution exp (L0 τ ) in
between.
X∞ Z t Z tn Z t2
Wt = eL0 t + dtn dtn−1 · · · dt1 (89)
n=1 0 0 0
X
L0 (t−tn ) L0 (tn −tn−1 )
× e Lkn e Lkn−1 · · · eL0 (t2 −t1 ) Lk1 eL0 t1
{Rn }
| {z }
K Rt
n
As a result of the negative imaginary part in (85) the exp (L0 τ ) are trace
decreasing 9 completely positive maps,
iτ iτ †
eL0 τ ρ = exp − HC ρ exp HC > 0 (90)
~ ~
X N
d
tr eL0 τ ρ = tr L0 eL0 τ ρ = − tr(Lk eL0 τ ρ) < 0 (91)
dτ | {z }
k=1 >0
It is now natural to interpret tr eL0 t ρ as the probability that no jump occurs
during the time interval t,
Prob R0t |ρ := tr eL0 t ρ . (92)
To see that this makes sense, we attribute to each record of length n a n-time
probability density. For a given record Rnt we define
prob Rnt |ρ := tr KRtn ρ , (93)
KRtn := eL0 (t−tn ) Lkn eL0 (tn −tn−1 ) Lkn−1 · · · eL0 (t2 −t1 ) Lk1 eL0 t1 . (94)
This is reasonable since the KRtn are completely positive maps that do not
preserve the trace. Indeed, the probability density for a record is thus de-
termined both by the corresponding jump operators, which involve the rates
9
See the note in Sect. 3.1(b).
Introduction to decoherence theory 27
γk , and by the eL0 τ ρ, which account for the fact that the likelihood for the
absence of a jump decreases with the length of the time interval.
This notion of probabilities is consistent, as can be seen by adding the
probability (92) for no jump to occur during the interval (0; t) to the integral
over the probability densities (93) of all possible jump sequences. As required,
the result is unity,
∞ Z Z Z
X t tn t2 X
Prob R0t |ρ + dtn dtn−1 · · · dt1 prob Rnt |ρ = 1
n=1 0 0 0 {Rtn }
for all ρ and t > 0. This follows immediately from the trace preservation of
the map (89).
It is now natural to normalize the transformation defined by the KRtn .
Formally, this yields the state transformation conditioned to a certain record
Rnt . It is called a quantum trajectory,
KRtn ρ
T ρ|Rnt := . (95)
tr KRtn ρ
and
(0;t) (0;t′ ) (t′ ;t)
T (ρ|Rn+m ) = T (T (ρ|Rn )|Rm ). (99)
Note finally, that the concept of quantum trajectories fits seamlessly into
the framework of generalized measurements discussed in Sect. 1.3(a). In par-
ticular, the conditioned state transformation T (ρ|Rnt ) has the form (18) of an
efficient measurement transformation,
MRtn ρM†Rt
T ρ|Rnt = n
(100)
tr M†Rt MRtn ρ
n
MRtn := e−iHC (t−tn )/~ Lkn · · · Lk2 e−iHC (t2 −t1 )/~ Lk1 e−iHC t1 /~ . (101)
This shows that we can legitimately view the open quantum dynamics gen-
erated by L as due to the continuous monitoring of the system by the en-
vironment. We just have to identify the (aptly named) record Rnt with the
total outcome of a hypothetical, continuous measurement during the interval
(0; t). The jump operators Lk then describe the effects of the corresponding
elementary measurement events10 (“clicks of counter k”). Since the absence
of any click during the “waiting time” τ may also confer information about
the system, this lack of an event constitutes a measurement, as well, which
is described by the non-unitary operators exp (−iHC τ /~). A hypothetical de-
mon, who has the full record Rnt available, would then be able to predict the
final state T (ρ|Rnt ). In the absence of this information we have to resort to
the probabilistic description (96) weighting each quantum trajectory with its
(Bayesian) probability.
We can thus conclude that the dynamics of open quantum dynamics, and
therefore decoherence, can in principle be understood in terms of an infor-
mation transfer to the environment. Apart from this conceptual insight, the
unravelling of a maser equation provides also an efficient stochastic simula-
tion method for its numerical integration. In these quantum jump approaches
[24, 25, 26], which are based on the observation that the quantum trajectory
(95) of a pure state remains pure, one generates a finite ensemble of trajec-
tories such that the ensemble mean approximates the solution of the master
equation.
10
Keep in mind that the Lk are not uniquely specified by a given generator L, see
Eqs. (80)-(82). Different choices of the Lindblad operators lead to different un-
ravellings of the master equation, so that these hypothetical measurement events
must not be viewed as “real”.
Introduction to decoherence theory 29
(a) Dephasing
we are faced with the exceptional fact that the state remains pure during the
Lindblad time evolution. Indeed, the solution of the Lindblad equation reads
11
See also Sect. ?.1.3 in Cord Müller’s contribution for a discussion of the master
equation describing of spin relaxation.
12
As an exception, γ does not have the dimensions of a rate here (to avoid clumsy
notation).
30 Klaus Hornberger
with γ
αt = α0 exp −iωt − t (106)
2
as can be verified easily using (104). It describes how the coherent state spirals
in phase space towards the origin, approaching the ground state as t → ∞.
The rate γ is the dissipation rate since it quantifies the energy loss, as shown
by the time dependence of the energy expectation value,
with a rate γdeco. For macroscopically distinct superpositions, where the phase
space distance of the quantum states is much larger than their uncertainties,
|α0 − β0 | ≫ 1, γdeco can be much greater than the dissipation rate,
γdeco 1 2
= |α0 − β0 | ≫ 1. (113)
γ 2
Introduction to decoherence theory 31
This quadratic increase of the decoherence rate with the separation between
the coherent states has been confirmed experimentally in a series of beautiful
cavity QED experiments in Paris, using field states with an average of 5 – 9
photons [27, 28].
Given this empiric support we can ask about the prediction for a material,
macroscopic oscillator. As an example, we take a pendulum with a mass of
m =100 g and a period of 2π/ω = 1s, and assume that we can prepare it in
a superposition of coherent states with a separation of x =1 cm. The mode
variable α is related to position and momentum by
r
mω p
α= x+i (114)
2~ mω
so that we get the prediction
γdeco ≃ 1030 γ.
This purports that even with an oscillator of enormously low friction corre-
sponding to a dissipation rate of γ =1/year the coherence is lost on a timescale
of 10−22 sec—in which light travels the distance of about a nuclear diameter.
This observation is often evoked to explain the absence of macroscopic su-
perpositions. However, it seems unreasonable to assume that anything physi-
cally relevant takes place on a timescale at which a signal travels at most by
the diameter of an atomic nucleus. Rather, one expects that the decoherence
rate should saturate at a finite value if one increases the phase space distance
between the superposed states.
The three terms in the upper line (with pth from Eq. (124)) constitute the
Caldeira-Leggett master equation. It is a Markovian, but not a completely
positive master equation. In a sense, the last term in (117) adds the mini-
mal modification required to bring the Caldeira-Leggett master equation into
Lindblad form [2].
To see the most important properties of (117) let us take a look at the time
evolution of the relevant observables in the Heisenberg picture. As discussed
in Sect. 3.1(d), the Heisenberg equations of motion are determined by the
dual Liouville operator L♯ . In the present case, it takes the form
1 p2 γ γ p2th
L♯ (A) = [A, + V (x)] − (p [x, A] + [x, A] p) − [x, [x, A]]
i~ 2m i~ 2 ~2
γ 1
− 2 [p, [p, A]] . (118)
2 pth
X∞ n
♯ (−2γt)
pt = e L t p = p = e−2γt p [for V = 0] (120)
n=0
n!
∞
♯ 1 X tn ♯ n−1
xt = eL t x = x + L (p)
m n=1 n!
p − pt
= x+ [for V = 0] (121)
2γm
Introduction to decoherence theory 33
Note that, unlike in closed systems, the Heisenberg operators do not retain
their commutator, [xt , pt ] 6= i~ for t > 0 (since the map Wt♯ = exp L♯ t is
non-unitary). Similarly, (p2 )t 6= (pt )2 for t > 0, so that the kinetic energy
operator T = p2 /2m has to be calculated separately. Noting
p2th
L♯ (T) = γ − 4γT [for V = 0] (122)
2m
we find similarly
p2th p2
Tt = + T − th e−4γt [for V = 0]. (123)
8m 8m
This shows how the kinetic energy approaches a constant determined by the
momentum scale pth . We can now relate pth to a temperature by equating the
stationary expectation value tr (ρT∞ ) = p2th /8m with the the average kinetic
energy in a one-dimensional thermal distribution. This leads to
p
pth = 2 mkB T . (124)
Not always is one able to state the operator evolution in closed form. In
those cases it may be helpful to take a look at the Ehrenfest equations for
their expectation values. For example, given hp2 it = 2mhTit , the other second
moments, hx2 it and hpx + xpit form a closed set of differential equations.
Their solutions, given in [2], yield the time evolution of the position variance
σx2 (t) = hx2 it − hxi2t . It has the asymptotic form
kB T
σx2 (t) ∼ t as t → ∞, (125)
mγ
which shows the diffusive behavior expected of a (classical) Brownian parti-
cle13 .
Let us finally take a closer look at the physical meaning of the third term in
(117), which is dominant if the state is in a superposition of spatially separated
states. Back in the Schrödinger picture we have in position representation,
ρt (x, x′ ) = hx|ρt |x′ i,
γ p2th
∂t ρt (x, x′ ) = − 2
(x − x′ )2 ρt (x, x′ ) + [the other terms]. (126)
2
| ~ {z }
γdeco
The “diagonal elements” ρ (x, x) are unaffected by this term, so that it leaves
the particle density invariant. The coherences in position representation, on
the other hand, get exponentially suppressed,
Again the decoherence rate is determined by the square of the relevant dis-
tance |x − x′ |,
γdeco (x − x′ )2
= 4π . (128)
γ Λ2th
Like in Sect. 3.4(b), the rate γdeco will be much larger than the dissipative
rate, provided the distance is large on the quantum scale, here given by the
thermal de Broglie wavelength,
2π~2
Λ2th = . (129)
mkB T
In particular, one finds γdeco ≫ γ if the separation is truly macroscopic.
Again, it seems unphysical that the decoherence rate does not saturate as
|x − x′ | → ∞, but grows above all bounds. One might conclude from this
that non-Markovian master equation are more appropriate on these short
timescales. However, I will argue that (unless the environment has very special
properties) Markovian master equations are well suited to study decoherence
processes, provided they involve an appropriate description of the microscopic
dynamics.
4 Microscopic derivations
In this section we discuss two important and rather different strategies to
obtain Markovian master equations based on microscopic considerations.
The most widely used form of incorporating the environment is the weak
coupling approach. Here one assumes that the total Hamiltonian is “known”
microscopically, usually in terms of a simplified model,
Htot = H + HE + Hint
and takes the interaction part Hint to be “weak” so that a perturbative treat-
ment of the interaction is permissible.
The main assumption, called the Born approximation, states that Hint
is sufficiently small so that we can take the total state as factorized, both
initially, ρtot (0) = ρ(0) ⊗ ρE , and also at t > 0 in those terms which involve
Hint to second order.
1 e
∂t ρ̃tot = [Hint (t), ρ̃tot (t)]
i~
Z t
1 e 1 e int (t), [H
e int (s), ρ̃tot (s)]].(131)
= [H int (t), ρ̃ tot (0)] + ds [H
i~ (i~)2 0
In the second equation, which is still exact, the von Neumann equation in its
integral equation version was inserted into the differential equation version.
Using a basis of Hilbert-Schmidt operators of the product Hilbert space, see
Sect. 3.2, one can decompose the general H e int into the form
X
e int (t) =
H Ae k (t) ⊗ B e k (t) (132)
k
with Aek = Ae† , B e † . The first approximation is now to replace ρ̃tot (s) by
ek = B
k k
ρ̃(s) ⊗ ρE in the double commutator of (131), where H e int appears to second
order. Performing the trace over the environment one gets
All the relevant properties of the environment are now expressed in terms of
the (complex) bath correlation functions Ckℓ (t − s). They depend only on the
time difference t − s, since [HE , ρE ] = 0,
Ckℓ (τ ) = tr eiHE τ Bk e−iHE τ Bℓ ρE ≡ eiHE τ Bk e−iHE τ Bℓ ρE . (134)
It is called the Redfield equation and it is not Markovian, because the inte-
grated superoperator still depends on time. Since the kernel is appreciable
only at the origin it is reasonable to replace t in the upper integration limit
by ∞.
These steps are summarized by the Born-Markov approximation:
Z t Z ∞
Assumption 2 : ∼
dsK(t − s)ρ̃ (s) = dsK(s)ρ̃ (t) (136)
0 0
we have
X
Ak = Ak (ω). (138)
ω
The time dependence of the operators in the interaction picture is now par-
ticularly simple, X
e k (t) =
A e−iωt Ak (ω) . (139)
ω
Inserting this decomposition we find
XX ′
∂t ρ̃(t) = ei(ω−ω )t Γkℓ (ω ′ ){Aℓ (ω ′ )ρ̃(t)A†k (ω) − A†k (ω)Aℓ (ω ′ )ρ̃(t)}
kℓ ωω ′
+ h.c. (140)
Introduction to decoherence theory 37
with Z ∞
1 e k (s)Bℓ (0)iρ .
Γkℓ (ω) = 2 dseiωs hB E (141)
~ 0
For times t which are large compared to the time scale given by the smallest
system energy spacings it is reasonable to expect that only equal pairs of
frequencies ω, ω ′ contribute appreciably to the sum in (140), since all other
contributions are averaged out by the wildly oscillating phase factor. This
constitutes the rotating wave approximation, our third assumption
X ′ X
Assumption 3 : ei(ω−ω )t f (ω, ω ′ ) ≃ f (ω, ω). (142)
ωω ′ ω
−1
P
with C := ~ ℓ vℓ Bℓ (0). One can now check that due to its particular form the
correlation function
D E
f (t) = eiHE t/~ C† e−iHE t/~ C
ρE
Fig. 1. (a) In the monitoring approach the system is taken to interact at most with
one environmental (quasi-)particle at a time, so that three-body collisions are ex-
cluded. Moreover, in agreement with the Markov assumption, it is assumed that the
environmental particles disperse their correlation with the system before scattering
again. (b) In order to consistently incorporate the state-dependence of the collision
rate into the dynamic description of the scattering process, we imagine that the
system is monitored continuously by a transit detector, which tells at a temporal
resolution ∆t whether a particle is going to scatter off the system, or not.
collision rate, so that such a naive ansatz would yield a nonlinear equation.
To account for this state dependence of the collision rate in a proper way
we will apply the concept of generalized measurements discussed in Sect. 1.3.
Specifically, we shall assume that the system is surrounded by a hypotheti-
cal, minimally invasive detector, which tells at any instant whether a probe
particle has passed by and is going to scatter off the system, see Fig. 1.
The rate of collisions is then described by a positive operator Γ acting in
the the system-probe Hilbert space. Given the uncorrelated state ̺tot = ρ⊗ρE
it determines the probability of a collision to occur in a small time interval
∆t,
Prob (C∆t |ρ ⊗ ρE ) = ∆t tr (Γ [ρ ⊗ ρE ]) . (149)
Here, ρE is the stationary reduced single particle state of the environment.
The microscopic definition of Γ will in general involve the current density
operator of the relative motion and a total scattering cross section, see below.
The important point to note is that the information that a collision is going
to take place changes our knowledge about the state, that is, the description
of the state according to the generalized measurement transformation (16).
At the same time, we have to keep in mind that the detector is not real, but
is introduced here only for enabling us to account for the state dependence
of the collision probability. It is therefore reasonable to take the detection
process as efficient , see Sect. 1.3(a), and minimally-invasive, i.e., Uα = I in
Eq. (20), so that neither unnecessary uncertainty nor a reversible back-action
40 Klaus Hornberger
is introduced. This implies that after a (hypothetical) detector click, but prior
to scattering, the system-probe state will have the form
We can now form the unconditioned system-probe state after time ∆t by tak-
ing into account that the detection outcomes are not really available. The
infinitesimally evolved state is then given by the mixture of the colliding
state transformed by the S-matrix and the untransformed non-colliding one,
weighted with their respective probabilities,
̺′tot (∆t) = Prob (C∆t |̺tot ) SM (̺tot |C∆t ) S† + Prob C∆t |̺tot M ̺tot |C∆t
= SΓ1/2 ̺tot Γ1/2 S† ∆t + ̺tot − Γ1/2 ̺tot Γ1/2 ∆t. (152)
15
Although more general transformations are conceivable, this one is distinguished
by the fact that it introduces no further (time-dependent) operators.
Introduction to decoherence theory 41
d 1 i h i
ρ= [H, ρ] + TrE T + T† , Γ1/2 [ρ ⊗ ρE ] Γ1/2
dt i~ 2
+ TrE TΓ1/2 [ρ ⊗ ρE ] Γ1/2 T†
1
− TrE Γ1/2 T† TΓ1/2 [ρ ⊗ ρE ]
2
1
− TrE [ρ ⊗ ρE ] Γ1/2 T† TΓ1/2 . (155)
2
This general evolution equation generates a dynamical semigroup since the
transformation (152) is completely positive. It incorporates the state depen-
dence of the collision rate in terms of the operators Γ1/2 , while the operators
T describe the individual microscopic interaction process without approxima-
tion. We will see that the second term in (155), which involves a commutator,
accounts for the renormalization of the system energies due to the coupling
to the environment, just like (148).
So far, the discussion was very general. To obtain concrete master equa-
tions one has to specify system and environment, along with the operators
Γ and S describing their interaction. In the following applications, we will
assume the environment to be an ideal Maxwell gas, whose single particle
state
Λ3th p2
ρgas = exp −β (156)
Ω 2m
is characterized by the inverse temperature β, the normalization volume Ω,
and the thermal de Broglie wave length Λth defined in (129).
The reduction factor will be the smaller in magnitude the better the scattered
state of the gas particle can “resolve” between the positions X and X ′ .
In order to obtain the dynamic equation we need to specify the rate oper-
ator. Classically, the collision rate is determined by the product of the current
density j = ngas vrel and the total cross section σ (prel ), and therefore Γ should
be expressed in terms of the corresponding operators. This is particularly sim-
ple in the large mass limit M → ∞, where vrel = |p/m − P /M | → |p| /m, so
that the current density and the cross section depend only on the momentum
of the gas particle, leading to
|p|
Γ = ngas σ (p) . (160)
m
If the gas particle moves in a normalized wave packet heading towards the
origin then the expectation value of this operator will indeed determine the
collision probability. However, this expression depends only on the modulus of
the velocity so that it will yield a finite collision probability even if the particle
is heading away form the origin. Hence, for (155) to make sense either the S-
matrix should keep such a non-colliding state unaffected, or Γ should contain
in addition a projection to the subset of incoming states, see the discussion
below.
In momentum representation, ρ P , P ′ = hP |ρ|P ′ i, equation (155) as-
sumes the general structure16
16
The second term in (155) vanishes in the limit M ≫ m.
Introduction to decoherence theory 43
2
1 P 2 − (P ′ )
∂t ρ P , P ′ = ρ P,P′
i~Z 2M
+ dP 0 dP ′0 ρ P 0 , P ′0 M P , P ′ ; P 0 , P ′0
Z Z
1
− dP 0 ρ P 0 , P ′ dP f M (P f , P f ; P 0 , P )
2
Z Z
1
− dP ′0 ρ P , P ′0 dP f M P f , P f ; P ′ , P ′0 . (161)
2
The dynamics is therefore characterized by a single complex function
M P , P ′ ; P 0 , P ′0 = hP | trgas TΓ1/2 |P 0 ihP ′0 | ⊗ ρgas Γ1/2 T† |P ′ i,
ngas (2π~)3
× |p0 | σ (p0 ) |hp1 |T0 |p0 i|2
m Z Ω
ngas
= δ Q − Q′ dp0 µ (p0 ) |p0 | σ (p0 )
m
3
(2π~)
× |hp0 − Q|T0 |p0 i|2
Ω
=: δ Q − Q′ Min (Q) . (163)
This shows that, apart form the unitary motion, the dynamics is simply char-
acterized by momentum exchanges described in terms of gain and loss terms,
2 Z
′
1 P 2 − (P ′ )
∂t ρ P , P = ρ P , P + dQ ρ P − Q, P ′ − Q Min (Q)
′
i~ 2M
Z
−ρ P , P ′ dQMin (Q) . (164)
We still have to evaluate the function Min (Q), which can be clearly interpreted
as the rate of collisions leading to a momentum gain Q of the Brownian
particle,
Z 3
ngas (2π~)
Min (Q) = dp0 µ (p0 ) |p0 | σ (p0 ) |hp0 − Q|T0 |p0 i|2 . (165)
m Ω
44 Klaus Hornberger
The delta function ensures the conservation of energy during the collision.
At first sight, this leads to an ill-defined expression since the matrix element
(166) appears as a squared modulus in (165), so that the tree-dimensional
integration is over a squared delta function.
The appearance of this problem can be traced back to our disregard of
the projection to the subset of incoming states in the definition (160) of Γ.
When evaluating Min we used the diagonal representation (162) for ρgas in
terms of (improper) momentum eigenstates, which comprise both incoming
and outgoing characteristics if viewed as the limiting form of a wave packet.
One way of implementing the missing projection to incoming states would be
to use a different convex decomposition of ρgas , which admits a separation
into incoming and outgoing contributions [34]. This way, Min can indeed be
calculated properly, albeit in a somewhat lengthy calculation. A shorter route
to the same result sticks to the diagonal representation, but modifies the
definition of S in a formal sense so that it keeps all outgoing state invariant17 .
The conservation of the probability current, which must still be guaranteed
by any such modification, then implies a simple rule how to deal with the
squared matrix element [34],
!
3
(2π~) 2 f (pf , pi )2 p2f p 2
hpf |T0 |pi i −→ δ − i . (167)
Ω pi σ(pi ) 2 2
R 2
Here σ(p) = dΩ ′ |f (pn′ , pn)| is the total elastic cross section. With this
replacement we obtain immediately
Z !
2
ngas 2 p20 (p0 − Q)
Min (Q) = dp0 µ (p0 ) |f (p0 − Q, p0 )| δ − .(168)
m 2 2
Fig. 2. The localization rate (171) describing the loss of wave-like behavior in a
Brownian particle state saturates for large distances at the average collision rate. In
contrast, the Caldeira-Leggett model predicts a quadratic increase beyond all bounds
(dashed line), see Eq. (127). This indicates that linear coupling models should be
taken with care if time scales are involved that differ strongly from the dissipation
time scale.
Here, the unit vectors n1 , n2 are the directions of incoming and outgoing gas
particles associated to the elements of solid angle dΩ1 and dΩ2 and ν (v) is
the velocity distribution in the gas. Clearly, F (x) determines how fast the
spatial coherences corresponding to the distance x decay.
One angular integral in (170) can be performed in the case of isotropic
scattering, f (pf , pi ) = f cos pf , pi ; E = p2i /2m . In this case,
Z Z
∞ 1
m 2
F (x) = dv ν(v) n gas v σ(mv) − 2π d (cos θ) f cos θ; E = v 2
0 −1 2
θ mv |x|
× sinc 2 sin , (171)
2 ~
sinc approaches unity and the angular integral yields the total cross section
σ so that the localization rate vanishes, as required. At very small distances,
a second order expansion in the distance x is permissible and one obtains a
quadratic dependence [36], such as predicted by the Caldeira-Leggett model,
see Eq. (128). However, once the distance is sufficiently large so that the
scattered state can resolve whether the collision took place at position X or
X ′ the sinc function in (171) suppresses the integrand. It follows that in the
limit of large distances the localization rate saturates, at a value given by the
average collision rate F (∞) = hσvngas i, see Figure 2.
Decoherence in this saturated regime of large separations has been ob-
served, in good agreement with this theory, in molecular interference exper-
iments in the presence of various gases [37]. The intermediate regime be-
tween quadratic increase and saturation was also seen in such experiments
on momentum-exchange mediated decoherence, by studying the influence of
the heat radiation emitted by fullerene molecules on the visibility of their
interference pattern [38].
As a conclusion of this section, we see that the scattering approach per-
mits to incorporate realistic microscopic interactions transparently and with-
out approximation in the interaction strength. The results show clearly that
linear coupling models, which imply that decoherence rates grow above all
bounds, have a limited range of validity. They cannot be judged by their suc-
cess in describing dissipative phenomena. Frequent claims of “universality” in
decoherence behavior, which are based on these linear coupling models, are
therefore to be treated with care.
As a second application of the monitoring approach, let us see how the dynam-
ics of an immobile object with discrete internal structure, such as an imple-
mentation of a quantum dot, gets affected by an environment of ideal gas par-
ticles. For simplicity, we take the gas again in the Maxwell state (156), though
different dispersion relations, e.g., in the case of phonon quasi-particles, could
be easily incorporated. The interaction between system and gas will be de-
scribed in terms of the in general inelastic scattering amplitudes determined
by the interaction potential.
In the language of scattering theory the energy eigenstates of the non-
motional degrees of freedom are called channels. In our case of a structureless
gas they form a discrete basis of the system Hilbert space. In the following,
the notation |αi, not to be confused with the coherent states of Sect. 3.4(b),
will be used to indicate the system eigenstates of energy Eα . In this channel
basis, ραβ = hα|ρ|βi, the equation of motion (155) takes on the form of a
general discrete master equation of Lindblad type,
Introduction to decoherence theory 47
Eα + εα − Eβ − εβ X α0 β0
∂t ραβ = ραβ + ρα0 β0 Mαβ
i~
α0 β0
1X X
α0 α 1X X
ββ0
− ρα0 β Mγγ − ραβ0 Mγγ . (172)
2 α γ
2 γ
0 β0
The real energy shifts εα given below describe the coherent modification of the
system energies due to the presence of the environment. They are due to the
second term in (155) and are the analogue of the Lamb shift (148) encountered
in the weak coupling calculation. The incoherent effect of the environment,
on the other hand, is described by the set of complex rate coefficients
α0 β0
Mαβ = hα| TrE TΓ1/2 [|α0 ihβ0 | ⊗ ρgas ] Γ1/2 T† |βi. (173)
In order to calculate these quantities we need again to specify the rate operator
Γ. In the present case, it is naturally given in terms of the current density
operator j = ngas p/m of the impinging gas particles multiplied by the channel-
specific total scattering cross sections σ (p, α),
X |p|
Γ= |αihα| ⊗ ngas σ (p, α) . (174)
α
m
Like in Sect. 4.3, this operator should in principle contain a projection to the
subset of incoming states of the gas particle. Again, this can be accounted
for in two different ways in the calculation of the rates (173). By using a
non-diagonal decomposition of ρgas , which permits to disregard the outgoing
states, one obtains18
Z
α0 β0 α0 β0 ngas
Mαβ = χαβ dp dp0 µ (p0 ) fαα0 (p, p0 )
m2
2
∗ p − p20
×fββ0 (p, p0 ) δ + Eα − Eα0 , (175)
2m
with the Kronecker-like factor
1 if Eα − Eα0 = Eβ − Eβ0
χααβ
0 β0
:= (176)
0 otherwise.
The energy shifts are determined the real parts of the forward scattering
amplitude,
Z
2 ngas
εα = −2π~ dp0 µ (p0 ) Re [fαα (p0 , p0 )] . (177)
m
18
For the special case of factorizing interactions, Hint = A ⊗ BE , and for times
large compared to all system time scales this result can be obtained rigorously
in a standard approach [39], by means of the “low density limit” scaling method
[22, 2].
48 Klaus Hornberger
Some details of this calculation can be found in [32]. Rather than repeating
them here we note that the result (175) can be obtained directly by using the
diagonal representation (162) of ρgas and the multichannel-generalization of
the replacement rule (167),
3 α0 β0 ∗
(2π~) χαβ fαα0 (p, p0 ) fββ (p, p0 )
hαp1 |T0 |α0 p0 ihβ0 p0 |T†0 |βp1 i → p 0
Ω p0 m σ (p0 , α0 ) σ (p0 , β0 )
2
p − p20
×δ + Eα − Eα0 .(178)
2m
The expression for the complex rates simplifies further if the scattering
am-
plitudes are rotationally invariant, fαα0 cos (p, p0 ) ; E = p20 /2m . In this case
we have
Z ∞ Z 1
α0 β0 α0 β0
Mαβ = χαβ dvν (v) ngas vout (v) 2π d (cos θ)
0 −1
m m 2
×fαα0 cos θ; E = v 2 fββ ∗
cos θ; E = v . (179)
2 0
2
with ν (v)the velocity distribution like in (171) and
r
2
vout (v) = v 2 − (Eα − Eα0 ) (180)
m
the velocity of a gas particle after a possibly inelastic collision.
This shows that limiting cases of (172) display the expected dynamics. For
the populations R ραα it reduces to a rate2 equation, where the cross sections
σαα0 (E) = 2π d (cos θ) |fαα0 (cos θ; E)| for scattering from channel α0 to α
determine the transition rates,
Z m
α0 α0
Mαα = dvν (v) ngas vout (v) σαα0 v2 . (181)
2
α0 β0
In the case of purely elastic scattering, on the other hand, i.e., for Mαβ =
αβ
Mαβ δαα0 δββ0 , the coherences are found to decay exponentially,
elastic
∂t |ραβ | = −γαβ |ραβ | . (182)
As one expects in this case, the better the scattering environment can dis-
tinguish between system states |αi and |βi the more coherence is lost in this
elastic process.
In the general case, the decay of off-diagonal elements will be due to a
combination of elastic and inelastic processes. Although little can be said
without specifying the interaction, it is clear that the integral over |fαα − fββ |2
in (183), a “decoherence cross section” without classical interpretation, is not
related to the inelastic cross sections characterizing the population transfer,
and may be much larger. In this case, the resulting decoherence will be again
much faster than the corresponding relaxation time scales.
We seek a nonlinear time evolution equation for robust pure states Pt which,
on the one hand side, preserves their purity, and on the other, keeps them as
close as possible to the evolved state following the master equation.
The possibly simplest nonlinear equation keeping a pure state pure is a
nonlinear extension of the Heisenberg form for the infinitesimal time step,
1
Pt+δt = Pt + δt [At , Pt ] + [Pt , [Pt , Bt ]] , (187)
i
where A and B are hermitian operators. In fact, the unitary part can by
absorbed into the nonlinear part by introducing the hermitian operator
Xt = −i[At , Pt ] + Bt . It “generates” the infinitesimal time translation of the
projectors (and may be a function of Pt ),
With this choice one confirms easily that the evolved operator has indeed the
properties of a projector, to leading order in δt,
P†t+δt = Pt+δt (189)
and
2
(Pt+δt ) = Pt+δt + O(δt2 ). (190)
The corresponding differential equation reads
Pt+δt − Pt
∂t Pt = = [Pt , [Pt , Xt ]] . (191)
δt
To determine the operator Xt one minimizes the distance between the time
derivatives of the truly evolved state and the projector. If we visualize the set
of all pure states as the extremal points in the convex set of states, then the
trajectory starting from a pure state will in general dive into the interior set
of mixed states, under the time evolution generated by L. With the help of
the minimization the operator Xt is chosen such that Pt remains extremal,
but as close as possible to the truly evolved state.
The (Hilbert-Schmidt) distance between the time derivatives can be cal-
culated as
h i
2
k L(Pt ) −∂t Pt k2HS = tr (Z − [Pt , [Pt , Xt ]])
| {z }
≡Z
2
= tr Z2 − 2(Z2 Pt − (ZPt ) )
+2 tr (Z − X)Pt − ((Z − X) Pt )2 . (192)
We note that the first term is independent of X, whereas the second one is
minimal for (Z − Xt )Pt = λPt . With the obvious choice Xt = Z ≡ L(Pt ) one
gets a nonlinear evolution equation for robust states Pt , which is trace and
purity preserving [43],
∂t Pt = [Pt , [Pt , L(Pt )]]. (193)
It is useful to write down the equation in terms of the vectors |ξi which
correspond to the pure state Pt = |ξihξ|,
∂t |ξi = [L(|ξihξ|) − hξ|L(|ξihξ|)|ξi]|ξi. (194)
| {z }
“decay rate”
5.2 Applications
Let us start with the damped harmonic oscillator discussed in Sect. 3.4(b).
By setting H = ~ωa† a and L = a Eq. (195) turns into
† † 1 † †
∂t |ξi = −iωa a|ξi + γ ha iξ (a − haiξ ) − a a − ha aiξ |ξi. (196)
2
Note that the first term of the non-unitary part vanishes if |ξi is a coherent
state, i.e, an eigenstate of a. This suggests the ansatz |ξi = |αi which leads to
h γ † γ 2i
∂t |αi = −iω − αa + |α| |αi. (197)
2 2
It is easy to convince oneself that this equation is solved by
2
/2 αt a†
|αt i = |α0 e−iωt−γt/2 i = e−|αt | e |0i, (198)
with αt = α0 exp (−iωt − γt/2). It shows that the predicted robust states
are indeed given by the slowly decaying coherent states encountered in
Sect. 3.4(b).
p2 4π
∂t |ξi = |ξi − γ 2 [(x − hxiξ )2 − h(x − hxiξ )2 iξ ]|ξi. (200)
2mi~ Λth | {z }
σξ2 (x)
This localizing effect is countered by the first term in (200) which causes the
dispersive broadening of the wave function. Since both effects compete we
expect stationary, soliton-like solutions of the equation.
Indeed, a Gaussian ansatz for |ξi with ballistic motion, i.e., hpiξ = p0 ,
hxiξ = x0 + p0 t/m,
and a fixed width σξ (x) = σ0 solves (200) provided
s 1/2
2 1 kB T 2 ~3
σ0 = Λ = , (201)
4π ~γ th 4γm2 kB T
Acknowledgments
Many thanks to Álvaro Tejero Cantero who provided me with his notes,
typed with the lovely TEXmacs program during the lecture. The present text
is based on his valuable input. I am also grateful to Marc Busse and Bassano
Vacchini for helpful comments on the manuscript.
This work was supported by the Emmy Noether program of the DFG.
References
1. E. Joos, H. D. Zeh, C. Kiefer, D. Giulini, J. Kupsch, and I.-O. Stamatescu, Deco-
herence and the Appearance of a Classical World in Quantum Theory, Springer,
Berlin, 2nd edition, 2003.
2. H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems, Oxford
University Press, Oxford, 2002.
3. W. T. Strunz, Decoherence in Quantum Physics, in Coherent Evolution in Noisy
Environments, edited by A. Buchleitner and K. Hornberger, Lecture Notes in
Physics 611, Berlin, 2002, Springer.
4. G. Bacciagaluppi, The Role of Decoherence in Quantum Mechanics, in The
Stanford Encyclopedia of Philosophy, edited by E. N. Zalta, Stanford University,
2005, http://plato.stanford.edu.
5. M. Schlosshauer, Decoherence, the measurement problem, and interpretations
of quantum mechanics, Rev. Mod. Phys. 76, 1267–1305 (2004).
6. W. H. Zurek, Decoherence, Einselection, and the Quantum Origins of the Clas-
sical, Rev. Mod. Phys. 75, 715–775 (2003).
7. J. P. Paz and W. H. Zurek, Environment-Induced Decoherence and the Transi-
tion From Quantum to Classical, in Les Houches Summer School Series, edited
by R. Kaiser, C. Westbrook, and F. David, volume 72, page 533, Springer-Verlag,
2001.
8. A. Bassi and G. Ghirardi, Dynamical Reduction Models, Phys. Rev. 379, 257
(2003).
54 Klaus Hornberger
35. K. Hornberger, Master equation for a quantum particle in a gas, Phys. Rev.
Lett. 97, 060601 (2006).
36. E. Joos and H. D. Zeh, The Emergence of Classical Properties Through Inter-
action with the Environment, Z. Phys. B: Condens. Matter 59, 223–243 (1985).
37. K. Hornberger, S. Uttenthaler, B. Brezger, L. Hackermüller, M. Arndt, and
A. Zeilinger, Collisional Decoherence Observed in Matter Wave Interferometry,
Phys. Rev. Lett. 90, 160401 (2003).
38. L. Hackermüller, K. Hornberger, B. Brezger, A. Zeilinger, and M. Arndt, De-
coherence of Matter Waves by Thermal Emission of Radiation, Nature 427,
711–714 (2004).
39. R. Dümcke, The low density limit for an N-level system interacting with a free
bose or fermi gas, Commun. Math. Phys. 97, 331–359 (1985).
40. W. H. Zurek, Pointer Basis of Quantum Apparatus: Into What Mixture Does
the Wave Packet Collapse?, Phys. Rev. D 24, 1516–1525 (1981).
41. L. Diósi and C. Kiefer, Robustness and Diffusion of Pointer States, Phys. Rev.
Lett. 85, 3552–3555 (2000).
42. W. H. Zurek, S. Habib, and J. P. Paz, Coherent states via decoherence, Phys.
Rev. Lett. 70, 1187–1190 (1993).
43. N. Gisin and M. Rigo, Relevant and irrelevant nonlinear Schrödinger equations,
J. Phys. A: Math. Gen. 28, 7375–7390 (1995).
Contents