Preskill Linblad Ops Chap3
Preskill Linblad Ops Chap3
Preskill Linblad Ops Chap3
1
2 CHAPTER 3. MEASUREMENT AND EVOLUTION
from its tunable coupling to the quantum system being measured. Since we
intend to measure the position of the pointer, it should be prepared initially
in a wavepacket state that is narrow in position space — but not too narrow,
because a vary narrow wave packet will spread too rapidly. If the initial
width of the wave packet is ∆x, then the uncertainty in it velocity will be
of order ∆v = ∆p/m ∼ ~/m∆x, so that after a time t, the wavepacket will
spread to a width
~t
∆x(t) ∼ ∆x + , (3.1)
m∆x
which is minimized for [∆x(t)]2 ∼ [∆x]2 ∼ ~t/m. Therefore, if the experi-
ment takes a time t, the resolution we can achieve for the final position of
the pointer is limited by
s
~t
∆x >
∼(∆x)SQL ∼ , (3.2)
m
the “standard quantum limit.” We will choose our pointer to be sufficiently
heavy that this limitation is not serious.
The Hamiltonian describing the coupling of the quantum system to the
pointer has the form
1 2
H = H0 + P + λMP, (3.3)
2m
where P2 /2m is the Hamiltonian of the free pointer particle (which we will
henceforth ignore on the grounds that the pointer is so heavy that spreading
of its wavepacket may be neglected), H0 is the unperturbed Hamiltonian of
the system to be measured, and λ is a coupling constant that we are able to
turn on and off as desired. The observable to be measured, M, is coupled to
the momentum P of the pointer.
If M does not commute with H0, then we have to worry about how the
observable evolves during the course of the measurement. To simplify the
analysis, let us suppose that either [M, H0 ] = 0, or else the measurement
is carried out quickly enough that the free evolution of the system can be
neglected during the measurement procedure. Then the Hamiltonian can be
approximated as H ' λMP (where of course [M, P] = 0 because M is an
observable of the system and P is an observable of the pointer), and the time
evolution operator is
U(t) ' exp[−iλtMP]. (3.4)
3.1. ORTHOGONAL MEASUREMENT AND BEYOND 3
we express U(t) as
X
U(t) = |ai exp[−iλtMaP]ha|. (3.6)
a
Now we recall that P generates a translation of the position of the pointer:
d d
P = −i dx in the position representation, so that e−ixo P = exp −xo dx , and
by Taylor expanding,
X
= αa |ai ⊗ |ψ(x − λtMa )i; (3.8)
a
the position of the pointer is now correlated with the value of the observable
M. If the pointer wavepacket is narrow enough for us to resolve all values of
the Ma that occur (∆x ∼ <λt∆Ma ), then when we observe the position of the
pointer (never mind how!) we will prepare an eigenstate of the observable.
With probability |αa|2 , we will detect that the pointer has shifted its position
by λtMa , in which case we will have prepared the M eigenstate |ai. In the
end, then, we conclude that the initial state |ϕi or the quantum system is
projected to |ai with probability |ha|ϕi|2. This is Von Neumann’s model of
orthogonal measurement.
The classic example is the Stern–Gerlach apparatus. To measure σ 3 for a
spin- 12 object, we allow the object to pass through a region of inhomogeneous
magnetic field
B3 = λz. (3.9)
4 CHAPTER 3. MEASUREMENT AND EVOLUTION
we can carry out a measurement procedure that will take a pure state |ψihψ|
to
Ea|ψihψ|Ea
(3.12)
hψ|Ea |ψi
with probability
Prob(a) = hψ|Ea|ψi. (3.13)
The measurement outcomes can be described by a density matrix obtained
by summing over all possible outcomes weighted by the probability of that
outcome (rather than by choosing one particular outcome) in which case the
measurement modifies the initial pure state according to
X
|ψihψ| → Ea |ψihψ|Ea. (3.14)
a
3.1. ORTHOGONAL MEASUREMENT AND BEYOND 5
H = HA ⊕ H⊥
A. (3.16)
Our observers who “live” in HA have access only to observables with support
in HA , observables MA such that
MA |ψ ⊥ i = 0 = hψ ⊥ |MA , (3.17)
for any |ψ ⊥ i ∈ H⊥
A . For example, in a two-qubit world, we might imagine
that our observables have support only when the second qubit is in the state
|0i2 . Then HA = H1 ⊗ |0i2 and H⊥ A = H1 ⊗ |1i2 , where H1 is the Hilbert
space of qubit 1. (This situation may seem a bit artificial, which is what I
meant in saying that the direct sum decomposition is not so well motivated.)
Anyway, when we perform orthogonal measurement in H, preparing one of
a set of mutually orthogonal states, our observer will know only about the
component of that state in his space HA . Since these components are not
6 CHAPTER 3. MEASUREMENT AND EVOLUTION
How does a general POVM affect the quantum state? There is not any
succinct general answer to this question that is particularly useful, but in
the case of a POVM by one-dimensional operators (as just discussed), where
the outcome |ψaihψa | occurs with probability tr(F aρ), summing over the
outcomes yields
ρ → ρ0 =
X
|ψaihψa |(λa hψa |ρ|ψai)
a
X q q
= λa |ψa ihψa | ρ λa |ψaihψa |
a
Xq q
= F aρ F a, (3.22)
a
where the λa ’s are positive real numbers, 0 < λa < 1, such that λa = 1.
P
a
Let
F a = λa (1 + n̂a · σ
~ ) = 2λa E(n̂a ), (3.24)
(where E(n̂a) is the projection | ↑n̂a ih↑n̂a |). Then
X X X
Fa = ( λa )1 + ( λa n̂a) · σ
~ = 1; (3.25)
a a a
F a = |ψ̃aihψ̃a |, (3.27)
where the vector |ψ̃a i is not normalized. Writing out the matrix elements
explicitly, the property a F a = 1 becomes
P
n n
∗
X X
(Fa)ij = ψ̃ai ψ̃aj = δij . (3.28)
a=1 a=1
Now let’s change our perspective on eq. (3.28). Interpret the (ψa )i ’s not as
n ≥ N vectors in an N-dimensional space, but rather an N ≤ n vectors
(ψiT )a in an n-dimensional space. Then eq. (3.28) becomes the statement
that these N vectors form an orthonormal set. Naturally, it is possible to
extend these vectors to an orthonormal basis for an n-dimensional space. In
other words, there is an n × n matrix uai , with uai = ψ̃ai for i = 1, 2, . . . , N,
such that
u∗aiuaj = δij ,
X
(3.29)
a
for any vector |ψi, and (at least for finite-dimensional matrices) the range
of U is the whole n-dimension space. Returning to the component notation,
we have
uaj u∗bj = δab ,
X
(3.31)
j
(3.37)
q q
1/3 − 1/3 1/3
and we see (as the theorem also assured us) that the columns (the |uai’s) are
then orthonormal as well. If we perform an orthogonal measurement onto
the |ua i basis, an observer cognizant of only the two-dimensional subspace
will conclude that we have performed the POVM {F 1, F 2 , F 3 }. We have
shown that if our qubit is secretly two components of a qutrit, the POVM
may be realized as orthogonal measurement of the qutrit.
where all E a ’s are mutually orthogonal projectors. Let us imagine that the
initial system of the quantum system is an “uncorrelated” tensor product
state
ρAB = ρA ⊗ ρB . (3.39)
To an observer who has access only to system A, the new density matrix for
that system is given by the partial trace of the above, or
trB [E a (ρA ⊗ ρB )E a ]
ρ0A (a) = . (3.42)
trAB [E a (ρA ⊗ ρB )]
The expression eq. (3.40) for the probability of outcome a can also be written
or
X
(Fa )ji = (Ea )jν,iµ (ρB )µν . (3.45)
µν
µν
X
= (Ea )jµ,iν (ρB )νµ = Fji
µν
(2) Positivity:
In the basis that diagonalizes ρB = µ pµ |µiB B hµ|, A hψ|F a |ψiA =
P
≥ 0 (because E a is positive).
(3) Completeness:
X X X
Fa = pµ B hµ| E a |µiB = 1A
a µ a
X
(because E a = 1AB and tr ρB = 1).
a
12 CHAPTER 3. MEASUREMENT AND EVOLUTION
Now consider, to start with, the special case n = rN, where r is a positive
integer. Then it is convenient to decompose |ψ̃a⊥ i as a direct sum of r − 1
N-component vectors:
|ψ̃a⊥ i = |ψ̃1,a
⊥ ⊥
i ⊕ |ψ̃2,a ⊥
i ⊕ · · · ⊕ |ψ̃r−1,a i; (3.48)
⊥
Here |ψ̃1,a i denotes the first N components of |ψ̃a⊥ i, |ψ̃2,a
⊥
i denotes the next
N components, etc. Then the orthonormality of the |ua i’s implies that
r−1
⊥ ⊥
X
δab = hua |ub i = hψ̃a |ψ̃b i + hψ̃µ,a |ψ̃µ,b i. (3.49)
µ=1
3
If there are more E a ’s than F a ’s, all but n outcomes have probability zero.
3.1. ORTHOGONAL MEASUREMENT AND BEYOND 13
{|µiB }, µ = 0, 1, 2, . . . , r − 1. (3.50)
and that we perform an orthogonal projection onto the basis {|Φa iAB } in
HA ⊗ HB . Then, since B h0|µiB = 0 for µ 6= 0, the outcome |Φa iAB occurs
with probability
and thus,
which isn’t quite the same thing as what we obtained in our “direct sum”
construction. In any case, there are many possible ways to realize a POVM
by orthogonal measurement and eq. (3.56) applies only to the particular
construction we have chosen here.
Nevertheless, this construction really is perfectly adequate for realizing
the POVM in which the state |ψa iA A hψa | is prepared in the event that
outcome a occurs. The hard part of implementing a POVM is assuring that
outcome a arises with the desired probability. It is then easy to arrange that
the result in the event of outcome a is the state |ψaiA A hψa |; if we like, once
the measurement is performed and outcome a is found, we can simply throw
ρA away and proceed to prepare the desired state! In fact, in the case of the
projection onto the basis |Φa iAB , we can complete the construction of the
POVM by projecting system B onto the {|µiB } basis, and communicating
the result to system A. If the outcome is |0iB , then no action need be taken.
⊥
If the outcome is |µiB , µ > 0, then the state |ψ̃µ,a iA has been prepared,
which can then be rotated to |ψa iA .
So far, we have discussed only the special case n = rN. But if actually
n = rN − c, 0 < c < N, then we need only choose the final c components of
⊥
|ψ̃r−1,a iA to be zero, and the states |ΦiAB will still be mutually orthogonal.
To complete the orthonormal basis, we may add the c states
2
F a = | ↑n̂a iA A h↑n̂a |, a = 1, 2, 3. (3.58)
3
where |β̃µ iB is the result of orthogonally projecting |βµ iB onto the support
of ρB . We may now perform the POVM on the support of ρB with F µ =
|β̃µ iB B hβ̃µ |, and thus prepare the state |ψµ iA with probability qµ .
3.2 Superoperators
3.2.1 The operator-sum representation
We now proceed to the next step of our program of understanding the be-
havior of one part of a bipartite quantum system. We have seen that a pure
state of the bipartite system may behave like a mixed state when we observe
subsystem A alone, and that an orthogonal measurement of the bipartite
system may be a (nonorthogonal) POVM on A alone. Next we ask, if a state
of the bipartite system undergoes unitary evolution, how do we describe the
evolution of A alone?
Suppose that the initial density matrix of the bipartite system is a tensor
product state of the form
Now we perform the partial trace over HB to find the final density matrix of
system A,
ρ0A = trB UAB (ρA ⊗ |0iB B h0|) U†AB
X
= B hµ|UAB |0iB ρA B h0|UAB |µiB , (3.66)
µ
If we denote
It follows from the unitarity of UAB that the Mµ ’s satisfy the property
†
M†µ Mµ =
X X
B h0|UAB |µiB B hµ|UAB |0iB
µ µ
Eq. (3.69) defines a linear map $ that takes linear operators to linear
operators. Such a map, if the property in eq. (3.70) is satisfied, is called a
superoperator, and eq. (3.69) is called the operator sum representation (or
Kraus representation) of the superoperator. A superoperator can be regarded
as a linear map that takes density operators to density operators, because it
follows from eq. (3.69) and eq. (3.70) that ρ0A is a density matrix if ρA is:
We showed that the operator sum representation in eq. (3.69) follows from
the “unitary representation” in eq. (3.66). But furthermore, given the oper-
ator sum representation of a superoperator, it is always possible to construct
a corresponding unitary representation. We choose HB to be a Hilbert space
whose dimension is at least as large as the number of terms in the operator
sum. If {|ϕA } is any vector in HA , the {|µiB } are orthonormal states in HB ,
and |0iB is some normalized state in HB , define the action of UAB by
X
UAB (|ϕiA ⊗ |0iB ) = Mµ |ϕiA ⊗ |µiB . (3.71)
µ
18 CHAPTER 3. MEASUREMENT AND EVOLUTION
Nν ρA N†ν ,
X
$(ρA ) = (3.74)
ν
where Nν = Uνµ Mµ . We will see shortly that any two operator-sum repre-
sentations of the same superoperator are always related this way.
Superoperators are important because they provide us with a formalism
for discussing the general theory of decoherence, the evolution of pure states
into mixed states. Unitary evolution of ρA is the special case in which there
is only one term in the operator sum. If there are two or more terms, then
there are pure initial states of HA that become entangled with HB under
evolution governed by UAB . That is, if the operators M1 and M2 appearing
in the operator sum are linearly independent, then there is a vector |ϕiA such
that |ϕ̃1iA = M1 |ϕiA and |ϕ̃2iA = M2 |ϕiA are linearly independent, so that
the state |ϕ̃1 iA |1iB + |ϕ̃2iA |2iB + · · · has Schmidt number greater than one.
Therefore, the pure state |ϕiA A hϕ| evolves to the mixed final state ρ0A .
Two superoperators $1 and $2 can be composed to obtain another super-
operator $2 ◦ $1 ; if $1 describes evolution from yesterday to today, and $2
3.2. SUPEROPERATORS 19
3.2.2 Linearity
Now we will broaden our viewpoint a bit and consider the essential properties
that should be satisfied by any “reasonable” time evolution law for density
matrices. We will see that any such law admits an operator-sum representa-
tion, so in a sense the dynamical behavior we extracted by considering part
of a bipartite system is actually the most general possible.
A mapping $ : ρ → ρ0 that takes an initial density matrix ρ to a final
density matrix ρ0 is a mapping of operators to operators that satisfies
(0) $ is linear.
While (1), (2), and (3) really are necessary if ρ0 is to be a density matrix,
(0) is more open to question. Why linearity?
One possible answer is that nonlinear evolution of the density matrix
would be hard to reconcile with any ensemble interpretation. If
One can easily check that $ is positive and trace-preserving. Suppose that
the initial density matrix is ρ = 12 1, realized as the ensemble
1 1
ρ = | ↑z ih↑z | + | ↓z ih↓z |. (3.77)
2 2
Since tr(σ 1 ρ) = 0, the evolution of ρ is trivial, and both representatives of
the ensemble are unchanged. If the spin was prepared as | ↑z i, it remains in
the state | ↑z i.
But now imagine that, immediately after preparing the ensemble, we do
nothing if the state has been prepared as | ↑z i, but we rotate it to | ↑x i if it
has been prepared as | ↓z i. The density matrix is now
1 1
ρ0 = | ↑z ih↑z | + | ↑x i| ↑x i, (3.78)
2 2
so that trρ0 σ 1 = 12 . Under evolution governed by $, this becomes $(ρ0 ) =
σ 1 ρ0σ 1 . In this case then, if the spin was prepared as | ↑z i, it evolves to the
orthogonal state | ↓z i.
The state initially prepared as | ↑z i evolves differently under these two
scenarios. But what is the difference between the two cases? The difference
was that if the spin was initially prepared as | ↓z i, we took different actions:
doing nothing in case (1) but rotating the spin in case (2). Yet we have found
that the spin behaves differently in the two cases, even if it was initially
prepared as | ↑z i!
We are accustomed to saying that ρ describes two (or more) different
alternative pure state preparations, only one of which is actually realized
each time we prepare a qubit. But we have found that what happens if we
prepare | ↑z i actually depends on what we would have done if we had prepared
| ↓xi instead. It is no longer sensible, apparently, to regard the two possible
preparations as mutually exclusive alternatives. Evolution of the alternatives
actually depends on the other alternatives that supposedly were not realized.
3.2. SUPEROPERATORS 21
Joe Polchinski has called this phenomenon the “Everett phone,” because the
different “branches of the wave function” seem to be able to “communicate”
with one another.
Nonlinear evolution of the density matrix, then, can have strange, perhaps
even absurd, consequences. Even so, the argument that nonlinear evolution
should be excluded is not completely compelling. Indeed Jim Hartle has
argued that there are versions of “generalized quantum mechanics” in which
nonlinear evolution is permitted, yet a consistent probability interpretation
can be salvaged. Nevertheless, we will follow tradition here and demand that
$ be linear.
or
Nρ0 (|ϕiA ⊗ |ψiB ) = |ψiA ⊗ |ϕiB . (3.83)
Hence Nρ0 is a swap operator (which squares to the identity). The eigenstates
of Nρ0 are states symmetric under the interchange A ↔ B, with eigenvalue 1,
and antisymmetric states with eigenvalue −1. Since ρ0 has negative eigenval-
ues, it is not positive, and (since ρ is certainly positive), therefore, TA ⊗ IB
does not preserve positivity. We conclude that TA , while positive, is not
completely positive.
then the measurement in B that projects onto the {|µiE } basis has outcome
µ with probability
†
Prob(µ) = A hϕ|M µ M µ |ϕiA . (3.85)
Expressing ρA as an ensemble of pure states, we find the probability
Prob(µ) = tr(F µ ρA ), F µ = M †µ M µ , (3.86)
for outcome µ; evidently F µ is positive, and µ F µ = 1 follows from the
P
where {|iiA } and {|i0iB } are orthonormal bases of HA and HB . (We have
√ |ψ̃iAB so that AB hψ̃|ψ̃iAB = N; this saves us from writing
chosen to normalize
various factors of N in the formulas below.) Note that any vector
X
|ϕiA = ai|iiA , (3.93)
i
We say that |ϕiA is the “relative state” of the “index state” |ϕ∗iB . The map
|ϕiA → |ϕ∗ iB , (3.96)
is evidently antilinear, and it is in fact an antiunitary map from HA to a
subspace of HB . The operator M A ⊗ 1B acting on |ψ̃iAB gives
M A |iiA ⊗ |i0iB .
X
(M A ⊗ 1B )|ψ̃iAB = (3.97)
i
5
The argument given here follows B. Schumacher, quant-ph/9604023 (see Appendix A
of that paper.).
6
We say that the state |ψiAB is maximally entangled if trB (|ψiAB AB hψ|) ∝ 1A .
3.3. THE KRAUS REPRESENTATION THEOREM 25
then
(M µ |iiA |i0iB )(A hj|M †µ B hj 0 |)
X
($A ⊗ IB )(|ψ̃iAB AB hψ̃|) =
i,j
X
= qµ |Φ̃µ iAB AB hΦ̃µ |, (3.104)
µ
3.3. THE KRAUS REPRESENTATION THEOREM 27
where
√
M µ |iiA |i0iB . )
X
qµ |Φ̃µ iAB = (3.105)
i
Now consider two such ensembles (or correspondingly two operator-sum rep-
√ √
resentations of $A ), { qµ|Φ̃µ iAB } and { pa |Υ̃a iAB }. For each ensemble,
there is a corresponding “purification” in HAB ⊗ HC :
X√
qµ |Φ̃µ iAB |αµ iC
µ
X√
pa |Υ̃aiAB |βaiC , (3.106)
a
where {(αµ iC } and {|βaiC } are two different orthonormal sets in Hc . The
GHJW theorem asserts that these two purifications are related by 1AB ⊗ U0C ,
a unitary transformation on HC . Therefore,
X√
pa |Υ̃aiAB |βa iC
a
X√
= qµ |Φ̃µ iAB U0C |αµ iC
µ
X√
= qµ |Φ̃µ iAB Uµa |βaiC , (3.107)
µ,a
where, to establish the second equality we note that the orthonormal bases
{|αµ iC } and {|βaiC } are related by a unitary transformation, and that a
product of unitary transformations is unitary. We conclude that
√ X√
pa |Υ̃a iAB = qµ |Φ̃µ iAB Uµa , (3.108)
µ
error is equally likely. If {|0i, |1i} is an orthonormal basis for the qubit, the
three types of errors can be characterized as:
|0i→|1i 0 1
1. Bit flip error: |1i→|0i
or |ψi → σ 1|ψi, σ 1 = 1 0
,
|0i→|0i 1 0
2. Phase flip error: |1i→−|1i
or |ψi → σ 3|ψi, σ 3 = 0 −1
,
|0i→+i|1i 0 −i
3. Both: |1i→−i|0i
or |ψi → σ 2 |ψi, σ2 = i 0
.
If an error occurs, then |ψi evolves to an ensemble of the three states σ 1 |ψi, σ2 |ψi, σ 3|ψi,
all occuring with equal likelihood.
Unitary representation
The depolarizing channel can be represented by a unitary operator acting on
HA ⊗ HE , where HE has dimension 4. (I am calling it HE here to encour-
age you to think of the auxiliary system as the environment.) The unitary
operator UAE acts as
Kraus representation
To obtain an operator-sum representation of the channel, we evaluate the
partial trace over the environment in the {|µiE } basis. Then
so that
p p p
q r r r
M0 = 1 − p 1, M , = σ1, M 2 = σ2, M 3 = σ3.
3 3 3
(3.113)
ρ → ρ0 = (1 − p)ρ+
p
(σ 1 ρσ 1 + σ 2 ρσ 2 + σ 3 ρσ 3) . (3.115)
3
where we are summing over the four (in principle distinguishable) ways that
the environment could evolve.
Relative-state representation
We can also characterize the channel by describing how a maximally-entangled
state of two qubits evolves, when the channel acts only on the first qubit.
There are four mutually orthogonal maximally entangled states, which may
be denoted
1
|φ+ iAB = √ (|00iAB + |11iAB ),
2
1
|φ− iAB = √ (|00iAB − |11iAB ),
2
1
|ψ +iAB = √ (|01iAB + |10iAB ),
2
1
|ψ −iAB = √ (|01iAB − |10iAB ). (3.116)
2
If the initial state is |φ+ iAB , then when the depolarizing channel acts on the
first qubit, the entangled state evolves as
The “worst possible” quantum channel has p = 3/4 for in that case the
initial entangled state evolves as
1
|φ+ ihφ+ | → |φ+ ihφ+ | + |φ− ihφ− |
4
1
+|ψ + ihψ + | + |ψ − ihψ − | = 1AB ; (3.118)
4
it becomes the totally random density matrix on HA ⊗ HB . By the relative-
state method, then, we see that a pure state |ϕiA of qubit A evolves as
1 1
|ϕiA A hϕ| → B hϕ |2 1AB |ϕ∗ iB = 1A ;
∗
(3.119)
4 2
it becomes the random density matrix on HA , irrespective of the value of the
initial state |ϕiA . It is as though the channel threw away the initial quantum
state, and replaced it by completely random junk.
An alternative way to express the evolution of the maximally entangled
state is
4 4 1
|φ+ ihφ+ | → 1 − p |φ+ ihφ+ | + p 1AB . (3.120)
3 3 4
Thus instead of saying that an error occurs with probability p, with errors of
three types all equally likely, we could instead say that an error occurs with
probability 4/3p, where the error completely “randomizes” the state (at least
we can say that for p ≤ 3/4). The existence of two natural ways to define
an “error probability” for this channel can sometimes cause confusion and
misunderstanding.
One useful measure of how well the channel preserves the original quan-
tum information is called the “entanglement fidelity” Fe . It quantifies how
“close” the final density matrix is to the original maximally entangled state
|φ+ i:
Fe = hφ+ |ρ0 |φ+ i. (3.121)
For the depolarizing channel, we have Fe = 1 − p, and we can interpret Fe
as the probability that no error occured.
32 CHAPTER 3. MEASUREMENT AND EVOLUTION
Block-sphere representation
It is also instructive to see how the depolarizing channel acts on the Bloch
sphere. An arbitrary density matrix for a single qubit can be written as
1
ρ= 1 + P~ · σ
~ , (3.122)
2
where P~ is the “spin polarization” of the qubit. Suppose we rotate our axes
so that P~ = P3 ê3 and ρ = 12 (1 + P3 σ 3 ). Then, since σ 3 σ 3σ 3 = σ 3 and
σ 1 σ 3 σ 1 = −σ 3 = σ 2 σ 3 σ 2, we find
p 1 2p 1
ρ0 = 1 − p + (1 + P3 σ 3 ) + (1 − P3 σ 3 ),
3 2 3 2
(3.123)
or P30 = 1 − 43 p P3 . From the rotational symmetry, we see that
4
P~ 0 = 1 − p P~ , (3.124)
3
irrespective of the direction in which P points. Hence, the Bloch sphere
contracts uniformly under the action of the channel; the spin polarization
is reduced by the factor 1 − 34 p (which is why we call it the depolarizing
channel). This result was to be expected in view of the observation above
that the spin is totally “randomized” with probability 43 p.
Invertibility?
Why do we say that the superoperator is not invertible? Evidently we can
reverse a uniform contraction of the sphere with a uniform inflation. But
the trouble is that the inflation of the Bloch sphere is not a superoperator,
because it is not positive. Inflation will take values of P~ with |P~ | ≤ 1 to
values with |P~ | > 1, and so will take a density operator to an operator
with a negative eigenvalue. Decoherence can shrink the ball, but no physical
process can blow it up again! A superoperator running backwards in time is
not a superoperator.
Unitary representation
A unitary representation of the channel is
q √
|0iA |0iE → 1 − p|0iA |0iE + p|0iA |1iE ,
q √
|1iA |0iE → 1 − p|1iA |0iE + p|1iA |2iE . (3.125)
In this case, unlike the depolarizing channel, qubit A does not make any
transitions. Instead, the environment “scatters” off of the qubit occasionally
(with probability p) being kicked into the state |1iE if A is in the state |0iA
and into the state |2iE if A is in the state |1iA . Furthermore, also unlike the
depolarizing channel, the channel picks out a preferred basis for qubit A; the
basis {|0iA , |1iA } is the only basis in which bit flips never occur.
Kraus operators
Evaluating the partial trace over HE in the {|0iE , |1iE , |2iE }basis, we obtain
the Kraus operators
√ 10 √ 00
q
M0 = 1 − p1, M 1 = p , M2 = p .
00 01
(3.126)
$(ρ) = M 0 ρM 0 + M 1 ρM 1 + M 2ρM 2
! !
ρ00 0 ρ00 (1 − p) ρ01
= (1 − p)ρ + p = ;
0 ρ11 (1 − p)ρ10 ρ11
(3.127)
thus the on-diagonal terms in ρ remain invariant while the off-diagonal terms
decay.
Now suppose that the probability of a scattering event per unit time is
Γ, so that p = Γ∆t 1 when time ∆t elapses. The evolution over a time
34 CHAPTER 3. MEASUREMENT AND EVOLUTION
Bloch-sphere representation
This will be worked out in a homework exercise.
Interpretation
We might interpret the phase-damping channel as describing a heavy “clas-
sical” particle (e.g., an interstellar dust grain) interacting with a background
gas of light particles (e.g., the 30 K microwave photons). We can imagine
that the dust is initially prepared in a superposition of position eigenstates
|ψi = √12 (|xi + | − xi) (or more generally a superposition of position-space
wavepackets with little overlap). We might be able to monitor the behavior
of the dust particle, but it is hopeless to keep track of the quantum state of
all the photons that scatter from the particle; for our purposes, the quantum
state of the particle is described by the density matrix ρ obtained by tracing
over the photon degrees of freedom.
Our analysis of the phase damping channel indicates that if photons are
scattered by the dust particle at a rate Γ, then the off-diagonal terms in
ρ decay like exp(−Γt), and so become completely negligible for t Γ−1 .
At that point, the coherence of the superposition of position eigenstates is
completely lost – there is no chance that we can recombine the wavepackets
and induce them to interfere. (If we attempt to do a double-slit interference
pattern with dust grains, we will not see any interference pattern if it takes
a time t Γ−1 for the grain to travel from the source to the screen.)
The dust grain is heavy. Because of its large inertia, its state of motion is
little affected by the scattered photons. Thus, there are two disparate time
scales relevant to its dynamics. On the one hand, there is a damping time
scale, the time for a significant amount of the particle’s momentum to be
transfered to the photons; this is a long time if the particle is heavy. On the
other hand, there is the decoherence time scale. In this model, the time scale
for decoherence is of order Γ, the time for a single photon to be scattered
by the dust grain, which is far shorter than the damping time scale. For a
3.4. THREE QUANTUM CHANNELS 35
hγ + |γ−i = 1 − ε, ε 1. (3.128)
The phase-damping channel would still describe this situation, but with p
replaced by pε (if p is still the probability of a scattering event). Thus, the
decoherence rate would become Γdec = εΓscat , where Γscat is the scattering
rate (see the homework).
The intuition we distill from this simple model applies to a vast variety
of physical situations. A coherent superposition of macroscopically distin-
guishable states of a “heavy” object decoheres very rapidly compared to its
damping rate. The spatial locality of the interactions of the system with its
environment gives rise to a preferred “local” basis for decoherence. Presum-
ably, the same principles would apply to the decoherence of a “cat state”
√1 (| deadi + | alivei), since “deadness” and “aliveness” can be distinguished
2
by localized probes.
Unitary representation
We denote the atomic ground state by |0iA and the excited state of interest
by |1iA . The “environment” is the electromagnetic field, assumed initially to
be in its vacuum state |0iE . After we wait a while, there is a probability p
36 CHAPTER 3. MEASUREMENT AND EVOLUTION
that the excited state has decayed to the ground state and a photon has been
emitted, so that the environment has made a transition from the state |0iE
(“no photon”) to the state |1iE (“one photon”). This evolution is described
by a unitary transformation that acts on atom and environment according
to
(Of course, if the atom starts out in its ground state, and the environment
is at zero temperature, then there is no transition.)
Kraus operators
By evaluating the partial trace over the environment in the basis {|0iE , |1iE },
we find the kraus operators
! √ !
1 √ 0 0 p
M0 = , M1 = , (3.130)
0 1−p 0 0
The operator M 1 induces a “quantum jump” – the decay from |1iA to |0iA ,
and M 0 describes how the state evolves if no jump occurs. The density
matrix evolves as
ρ → $(ρ) = M 0 ρM †0 + M 1 ρM †1
√ ! !
ρ00 1 − pρ01 pρ11 0
= √ +
1 − pρ10 (1 − p)ρ11 0 0
√ !
ρ00 + pρ11 1 − pρ01
= √ . (3.132)
1 − pρ10 (1 − p)ρ11
If we apply the channel n times in succession, the ρ11 matrix element decays
as
The atom always winds up in its ground state. This example shows that it
is sometimes possible for a superoperator to take a mixed initial state, e.g.,
!
ρ00 0
ρ= , (3.135)
0 ρ11
If we detect the photon (and so project out the state |1iE of the environment),
then we have prepared the state |0iA of the atom. In fact, we have prepared
a state in which we know with certainty that the initial atomic state was the
excited state |1iA – the ground state could not have decayed.
On the other hand, if we detect no photon, and our photon detector has
perfect efficiency, then we have projected out the state |0iE of the environ-
ment, and so have prepared the atomic state
q
a|0iA + b 1 − p|1iA . (3.137)
38 CHAPTER 3. MEASUREMENT AND EVOLUTION
The atomic state has evolved due to our failure to detect a photon – it has
become more likely that the initial atomic state was the ground state!
As noted previously, a unitary transformation that entangles A with E,
followed by an orthogonal measurement of E, can be described as a POVM
in A. If |ϕiA evolves as
X
|ϕiA |0iE → M µ |ϕiA |µiE , (3.138)
µ
This condition often applies in practice, for example in atomic physics, where
(∆t)res ∼ ~/kT ∼ 10−14 s (T is the temperature) is orders of magnitude larger
than the typical lifetime of an excited atomic state.
An instructive example to study is the case where the system A is a
single harmonic oscillator (HA = ωa† a), and the reservoir R consists of many
oscillators (H R = i ωi b†i bi , weakly coupled to the system by a perturbation
P
H0 = λi (ab†i + a†bi ).
X
(3.143)
i
and that $t=0 = I. If the elapsed time is the infinitesimal interval dt, and
where H and K are both hermitian and Lµ , H, and K are all zeroth order
in dt. In fact, we can determine K by invoking the Kraus normalization
condition:
M †µ M µ = 1 + dt(2K + L†µ Lµ ),
X X
1= (3.152)
µ µ>0
or
1X †
K=− L Lµ . (3.153)
2 µ>0 µ
Substituting into eq. (3.149), expressing ρ(dt) = ρ(0) + dtρ̇(0), and equating
terms of order dt, we obtain Lindblad’s equation:
X 1 1
ρ̇ ≡ L[ρ] = −i[H, ρ] + Lµ ρL†µ − L†µ Lµ ρ − ρL†µ Lµ .
2 2
µ>0 (3.154)
The first term in L[ρ] is the usual Schrodinger term that generates unitary
evolution. The other terms describe the possible transitions that the system
may undergo due to interactions with the reservoir. The operators Lµ are
called Lindblad operators or quantum jump operators. Each Lµ ρL†µ term in-
duces one of the possible quantum jumps, while the −1/2L†µ Lµ ρ−1/2ρL†µ Lµ
terms are needed to normalize properly the case in which no jumps occur.
Lindblad’s eq (3.154) is what we were seeking – the general form of (com-
pletely positive) Markovian evolution of a density matrix: that is, the master
equation. It follows from the Kraus representation that we started with that
Lindblad’s equation preserves density matrices: ρ(t + dt) is a density matrix
3.5. MASTER EQUATION 43
if ρ(t) is. Indeed, we can readily check, using eq. (3.154), that ρ̇ is Hermitian
and trρ̇ = 0. That L[ρ] preserves positivity is somewhat less manifest but,
as already noted, follows from the Kraus representation.
If we recall the connection between the Kraus representation and the uni-
tary representation of a superoperator, we clarify the interpretation of the
master equation. We may imagine that we are continuously monitoring the
reservoir, projecting it in each instant of time onto the |µiR basis. With
probability 1 − 0(dt), the reservoir remains in the state |0iR , but with prob-
ability of order dt, the reservoir makes a quantum jump to one of the states
|µiR , µ > 0. When we say that the reservoir has “forgotten” the information
it acquired from the system (so that the Markovian approximation applies),
we mean that these transitions occur with probabilities that increase linearly
with time. Recall that this is not automatic in time-dependent perturbation
theory. At a small time t the probability of a particular transition is propor-
tional to t2; we obtain a rate (in the derivation of “Fermi’s golden rule”) only
by summing over a continuum of possible final states. Because the number
of accessible states actually decreases like 1/t, the probability of a transition,
summed over final states, is proportional to t. By using a Markovian de-
scription of dynamics, we have implicitly assumed that our (∆t)coarse is long
enough so that we can assign rates to the various possible transitions that
might be detected when we monitor the environment. In practice, this is
where the requirement (∆t)coarse (∆t)res comes from.
Let us also suppose that the reservoir is at zero temperature; then the ex-
citation level of the oscillator can cascade down by successive emission of
photons, but no absorption of photons will occur. Hence, there is only one
jump operator:
√
L1 = Γa. (3.156)
Here Γ is the rate for the oscillator to decay from the first excited (n = 1)
state to the ground (n = 0) state; because of the form of H, the rate for
44 CHAPTER 3. MEASUREMENT AND EVOLUTION
the decay from level n to n − I is nΓ.9 The master equation in the Lindblad
form becomes
1 1
ρ̇ = −i[H 0 , ρ] + Γ(aρa† − a† aρ − ρa† a). (3.157)
2 2
where H 0 = ωa†a is the Hamiltonian of the oscillator. This is the same
equation obtained by Carmichael from a more elaborate analysis. (The only
thing we have missed is the Lamb shift, a radiative renormalization of the
frequency of the oscillator that is of the same order as the jump terms in
L[ρ].)
The jump terms in the master equation describe the damping of the os-
cillator due to photon emission.10 To study the effect of the jumps, it is
convenient to adopt the interaction picture; we define interaction picture
operators ρI and aI by
so that
1 1
ρ̇I = Γ(aI ρI a†I − a†I aI ρ − ρI a†I aI ). (3.159)
2 2
where in fact aI (t) = ae−iωt so we can replace aI by a on the right-hand
side. The variable ã = e−iH0 t ae+iH0 t = eiωt a remains constant in the absence
of damping. With damping, ã decays according to
d d
hãi = tr(aρI ) = traρ̇ , (3.160)
dt dt
and from eq. (3.159) we have
1 1
traρ̇ = Γtr a ρI a − aa†aρI − aρI a† a
2 †
2 2
9
The nth level of excitation of the oscillator may be interpreted as a state of n nonin-
teracting particles; the rate is nΓ because any one of the n particles can decay.
10
This model extends our discussion of the amplitude-damping channel to a damped
oscillator rather than a damped qubit.
3.5. MASTER EQUATION 45
1 † Γ Γ
= Γtr [a , a]aρI = − tr(aρI ) = − hãi. (3.161)
2 2 2
Integrating this equation, we obtain
1 1
= Γtr a† aaρI a† − a† aa† aρI − a† aρI a†a
2 2
† † †
= Γtra [a , a]aρI = −Γtra aρI = −Γhni, (3.163)
which integrates to
Thus Γ is the damping rate of the oscillator. We can interpret the nth
excitation state of the oscillator as a state of n noninteracting particles,
each with a decay probability Γ per unit time; hence eq. (3.164) is just the
exponential law satisfied by the population of decaying particles.
More interesting is what the master equation tells us about decoherence.
The details of that analysis will be a homework exercise. But we will analyze
here a simpler problem – an oscillator undergoing phase damping.
Thus, there is just one Lindblad operator, and the master equation in the
interaction picture is.
1 1
ρ̇I = Γ a† aρI a† a − (a† a)2 ρI − ρI (a†a)2 . (3.166)
2 2
46 CHAPTER 3. MEASUREMENT AND EVOLUTION
Here Γ can be interpreted as the rate at which reservoir photons are scattered
when the oscillator is singly occupied. If the occupation number is n then
the scattering rate becomes Γn2 . The reason for the factor of n2 is that
the contributions to the scattering amplitude due to each of n oscillator
“particles” all add coherently; the amplitude is proportional to n and the
rate to n2 .
It is easy to solve for ρ̇I in the occupation number basis. Expanding
X
ρI = ρnm |nihm|, (3.167)
n,m
Of course, we have seen that not all superoperators are unitary. The point
of the hypothesis is that nonunitary evolution in an open system, including
50 CHAPTER 3. MEASUREMENT AND EVOLUTION
the collapse that occurs in the measurement process, always arises from dis-
regarding some of the degrees of freedom of a larger system. This is the view
promulgated by Hugh Everett, in 1957. According to this view, the evolution
of the quantum state of “the universe” is actually deterministic!
But even if we accept that collapse is explained by decoherence in a system
that is truly deterministic, we have not escaped all the puzzles of quantum
theory. For the wave function of the universe is in fact a superposition of a
state in which the cat is dead and a state in which the cat is alive. Yet each
time I look at a cat, it is always either dead or alive. Both outcomes are
possible, but only one is realized in fact. Why is that?
Your answer to this question may depend on what you think quantum
theory is about. There are (at least) two reasonable schools of thought.
1
|No decayiatom |Aliveicat |Know it0s Aliveime Prob =
2
(3.178)
But how do I infer that pdead and palive actually are probabilities that I (in
my Bayesian posture) may assign to my future perceptions? I still need
a rule to translate this density operator into probabilities assigned to the
alternatives. It seems contrary to the Everett philosophy to assume such a
rule; we could prefer to say that the only rule needed to define the theory
is the Schrödinger equation (and perhaps a prescription to specify the initial
wave function). Postulating a probability formula comes perilously close to
allowing that there is a nondeterministic measurement process after all. So
here is the issue regarding the foundations of theory for which I do not know
a fully satisfying resolution.
Since we have not been able to remove all discomfiture concerning the
origin of probability in quantum theory, it may be helpful to comment on an
interesting suggestion due to Hartle. To implement his suggestion, we must
return (perhaps with regret) to the frequency interpretation of probability.
Hartle’s insight is that we need not assume the probability interpretation as
part of the measurement postulate. It is really sufficient to make a weaker
assumption:
|ψi = a| ↑z i + b| ↓z i. (3.182)
and we imagine measuring σ 3 for each of the copies. Formally, the case of an
infinite number of trials can be formulated as the N → ∞ limit of N trials.
Hartle’s idea is to consider an “average spin” operator
N
1 X (i)
σ¯3 = lim σ3 , (3.184)
N →∞ N
i=1
We can compute
where the constant may depend on the particular computer used or on the
particular sequence, but not on N.
Coleman and Lesniewski consider an orthogonal projection operator E random
(i)
that, acting on a state |ψi that is an eigenstate of each σ 3 , satisfies
E random|ψi = 0, (3.189)
if the sequence is not random. This property alone is not sufficient to de-
termine how E random acts on all of (H2 )∞ , but with an additional technical
assumption, they find that E random exists, is unique, and has the property
physics is compatible with the hypothesis that all physical processes (includ-
ing measurements) can be accurately modeled by the unitary evolution of
a wave function (or density matrix). When a microscopic quantum system
interacts with a macroscopic apparatus, decoherence drives the “collapse” of
the wave function “for all practical purposes.”
If we eschew measurement as a mystical primitive process, and we accept
the wave function as a description of physical reality, then we are led to the
Everett or “many-worlds” interpretation of quantum theory. In this view,
all possible outcomes of any “measurement” are regarded as “real” — but I
perceive only a specific outcome because the state of my brain (a part of the
quantum system) is strongly correlated with the outcome.
Although the evolution of the wave function in the Everett interpretation
is deterministic, I am unable to predict with certainty the outcome of an
experiment to be performed in the future – I don’t know what branch of the
wavefunction I will end up on, so I am unable to predict my future state of
mind. Thus, while the “global” picture of the universe is in a sense deter-
ministic, from my own local perspective from within the system, I perceive
quantum mechanical randomness.
My own view is that the Everett interpretation of quantum theory pro-
vides a satisfying explanation of measurement and of the origin of random-
ness, but does not yet fully explain the quantum mechanical rules for com-
puting probabilities. A full explanation should go beyond the frequency
interpretation of probability — ideally it would place the Bayesian view of
probability on a secure objective foundation.
3.7 Summary
POVM. If we restrict our attention to a subspace of a larger Hilbert space,
then an orthogonal (Von Neumann) measurement performed on the larger
space cannot in general be described as an orthogonal measurement on the
subspace. Rather, it is a generalized measurement or POVM – the outcome
a occurs with a probability
Prob(a) = tr (F a ρ) , (3.191)
where ρ is the density matrix of the subsystem, each F a is a positive hermi-
tian operator, and the F a ’s satisfy
X
Fa = 1 . (3.192)
a
58 CHAPTER 3. MEASUREMENT AND EVOLUTION
M µ ρM †µ ,
X
$ : ρ → $(ρ) = (3.193)
µ
where
M †µ M µ = 1 .
X
(3.194)
µ
Here each Lindblad operator (or quantum jump operator) represents a “quan-
tum jump” that could in principle be detected if we monitored the envi-
ronment faithfully. By solving the master equation, we can compute the
decoherence rate of an open system.
3.8. EXERCISES 59
3.8 Exercises
3.1 Realization of a POVM
Consider the POVM defined by the four positive operators
1 1
P1 = | ↑z ih↑z | , P2 = | ↓z ih↓z |
2 2
1 1
P3 = | ↑xih↑x | , P4 = | ↓x ih↓x | .
2 2
(3.196)
Mµ |ψihψ|M†µ ,
X
M(|ψihψ|) = (3.197)
µ
inverse of M if N ◦ M = I, or
|hψ|Na Mµ |ψi|2 = 1.
X
(3.199)
µ,a
Na Mµ = λaµ 1, (3.200)
for each a and µ; i.e., that each NaMµ is a multiple of the identity.
b) Use the result of (a) to show that M†ν Mµ is proportional to the
identity for each µ and ν.
60 CHAPTER 3. MEASUREMENT AND EVOLUTION
$ : ρ → ρ0 , (3.201)
a) zero temperature.
b) room temperature.
√ 1
M2 = p (1 − σ 3 ). (3.203)
2
3.8. EXERCISES 61
Mµ = Uµa Na . (3.204)
q √
|1iA |0iE → 1 − p |1iA |0iE + p |1iA |γ1 iE ,
(3.205)
Show that this is again the phase-damping channel, and find its
operator-sum representation with two Kraus operators.
d) Suppose that the channel in (c) describes what happens to the qubit
when a single photon scatters from it. Find the decoherence rate
Γdecoh in terms of the scattering rate Γscatt .
eαa |0i.
2 /2 †
|αi = e−|α| (3.214)
Use the result of (a) to infer the density matrix at a later time
t. Assuming Γt 1, at what rate do the off-diagonal terms in ρ
decay (in this coherent state basis)?