Quantum metrology with full and fast quantum control
Pavel Sekatski
1
, Michalis Skotiniotis
1 2
, Jan Kołodyński
3
, and Wolfgang Dür
1
1
Institut für Theoretische Physik, Universität Innsbruck, Technikerstr. 21a, A-6020 Innsbruck, Austria
2
Física Teòrica: Informació i Fenòmens Quàntics, Departament de Física, Universitat Autònoma de Barcelona, 08193 Bellatera (Barcelona)
Spain
3
ICFO-Institut de Ciencies Fotoniques, The Barcelona Institute of Science and Technology, 08860 Castelldefels (Barcelona), Spain
September 6, 2017
We establish general limits on how precise
a parameter, e.g., frequency or the strength
of a magnetic field, can be estimated with the
aid of full and fast quantum control. We con-
sider uncorrelated noisy evolutions of N qubits
and show that fast control allows to fully re-
store the Heisenberg scaling ( 1/N
2
) for all
rank-one Pauli noise except dephasing. For
all other types of noise the asymptotic quan-
tum enhancement is unavoidably limited to
a constant-factor improvement over the stan-
dard quantum limit (1/N) even when allow-
ing for the full power of fast control. The lat-
ter holds both in the single-shot and infinitely-
many repetitions scenarios. However, even
in this case allowing for fast quantum control
helps to improve the asymptotic constant fac-
tor. Furthermore, for frequency estimation
with finite resource we show how a parallel
scheme utilizing any fixed number of entangled
qubits but no fast quantum control can be out-
performed by a simple, easily implementable,
sequential scheme which only requires entan-
glement between one sensing and one auxiliary
qubit.
1 Introduction
Precision measurements play a fundamental role in
physics and beyond, as they constitute the main in-
gredient for many state-of-the-art applications and ex-
periments [1]. When investigating the limits of known
or speculative theories experimentally one quickly en-
ters into regimes where quantities and parameters
need to be measured with unprecedented precision.
In this context, it is of utmost importance to know
the ultimate limits nature sets on how precise any
given quantity can be determined and how to achieve
this.
These questions are at the focus of quantum metrol-
ogy [2]. Quantum mechanics is a probabilistic the-
ory and the intrinsically stochastic nature of mea-
surements ultimately limits the achievable precision.
When considering classical probes (or particles) inde-
pendently sensing a physical parameter, such as phase
or frequency, the maximum attainable precision (as
quantified by the Mean Squared Error—MSE) follows
the standard scaling, 1/N , where N is the number of
probes [3]. In turn it was shown that quantum entan-
glement allows one to achieve the so-called Heisen-
berg scaling (HS) in precision, 1/N
2
, a quadratic im-
provement as compared to classical approaches [4, 5].
These precision limits apply to both single-shot pro-
tocols [613] as well as protocols utilizing many rep-
etitions [3, 5, 14]. Still, it remains unclear to what
extent such an improvement can be harnessed in prac-
tice under non-idealized conditions, i.e., when taking
unavoidable noise and imperfections into account [15
17].
Due to the difficulty of obtaining exact precision
limits in the presence of noise, several asymptotic
lower bounds have been established for particular
noise models [1825]. These lower bounds are of-
ten not only cumbersome to evaluate but also hard
to optimize, relying on educated guesses, numerical
methods employing semi-definite programming or a
combination of the two. Nevertheless, these bounds
show that for typical uncorrelated noise processes the
possible gain due to the usage of quantum resources
is limited to a constant-factor improvement over the
standard scaling, as opposed to a different scaling.
On the other hand, it was shown that for some types
of noise—namely noise perpendicular to the Hamilto-
nian that encodes the parameter of interest—the re-
striction to standard scaling can be circumvented, and
HS can be fully restored, by allowing for additional re-
sources, such as perfectly protected auxiliary particles
used to perform quantum error correction [2632].
In this work we develop a general framework that
provides us with analytic results for all types of noise
processes described by a time-homogeneous master
equation. In order to establish general limits, we ac-
count for the possibility of ancillary resources that do
not take part in the sensing process together with full
and fast quantum control (FFQC) of the system and
the ancillae
1
. Such FFQC allows one to effectively
modify the Lindblad superoperator that describes the
noise process, a possibility which has hitherto not
1
We remark that we do not allow for additional control over
the environment as has been recently considered in [33, 34].
Accepted in Quantum 2017-09-04, click title to verify 1
arXiv:1603.08944v5 [quant-ph] 5 Sep 2017
been considered
2
. Our approach allows us to identify
all types of qubit noise that can be fully corrected, and
thus restore HS in precision. In addition, for all other
types of uncorrelated noise we are able to provide fully
analytic bounds applicable to the entire hierarchy of
metrology schemes. In contrast to the available meth-
ods [19, 21], our bounds can be generally determined
without need to explicitly solve the system dynamics.
Our main results can be summarized as follows:
(i) FFQC allows to restore HS by completely elimi-
nating any Pauli rank-one noise (that is not par-
allel to the Hamiltonian) at the cost of slowing
down the unitary evolution by a constant factor;
(ii) All other noise processes unavoidably limit the
quantum gain to a constant factor improvement
over standard scaling despite full quantum con-
trol. We obtain analytic bounds for the achiev-
able improvement factor.
(iii) For standard scaling-limited noise processes
FFQC may yet allow for significant improvement
of precision in case of limited resources. We pro-
vide explicit examples demonstrating the advan-
tage of FFQC-assisted schemes over ones without
intermediate control for both continuous and dis-
crete processes.
Let us already stress that our results do not per-
tain only to the frequentist approach to parameter
estimation, which assumes an infinite number of pro-
tocol repetitions (i.e., sufficiently large statistics), but
also apply to the Bayesian approach within which one
considers a finite statistical data or, in the extreme
case, even a single experimental run. Moreover, the
most general protocol involving FFQC lends itself nat-
urally to adaptive strategies in which one is allowed
to modify the protocol (measurements, control oper-
ations etc.) "on-the-fly"—basing on the record of the
measurements already collected.
The paper is organized as follows. In Sec. 2 we recall
some recent results in quantum metrology, and com-
pare them through a hierarchy of quantum metrol-
ogy schemes accounting for various levels of control.
We show that the FFQC scheme, that tops the hi-
erarchy, is the most powerful one allowed by quan-
tum mechanics. Sec. 3 introduces the model of noisy
quantum processes described by a time-homogeneous
master equation, and reviews the concept of quantum
Fisher information (QFI). In Sec. 4, we review upper
bounds on the QFI for metrology schemes, account-
ing for intermediate control, and generalize them to
incorporate FFQC (Sec. 4.2.1) demonstrating (i) and
(ii). Our extension of the bounds to the limit of short
evolution times is crucial to show (ii), and is the key
to obtain analytic bounds (as it removes the necessity
2
To the best of our knowledge, quantum control has been
used to modify the overall system-environment dynamics but
not the dissipative evolution itself [29, 35].
Preparation
...
...
...
...
Control #1
...
...
(a)
(c)
(d)
(b)
(c)
(d)
Preparation
...
...
ancillae
(b)
(a)
ancillae
ancillae
Control #2
Control #k
Control #1
Control #2
Control #m
Figure 1: Hierarchy of quantum metrology protocols for sens-
ing a frequency-like parameter ω that is encoded on each
probe during its noisy evolution E
ω,t
. Within the standard
schemes of type (a), N probes independently sense the pa-
rameter for a time t. In (b), additionally an unlimited number
of ancillae is allowed to perform a more general measurement
strategy that may include a single error-correcting step at the
end of the protocol. In (c), the sensing process is interspersed
at time intervals, δt, with quantum control operations act-
ing on both the sensing probes and ancillae. Finally in (d),
control operations are further allowed to be of infinitesimally
small duration, i.e, δt dt, so that it is enough to con-
sider a single probe that sequentially senses the parameter
for the elongated total time t
0
= Nt. As full and fast quan-
tum control (FFQC), applied frequently on the global state,
may constitute swap operations any protocol of type (c) can
be simulated. Hence, the schemes form the following hi-
erarchy: (a)(b)(c)(d), when ordered in terms of their
ultimate power.
to solve system dynamics). In Sec. 5 we discuss the
performance of FFQC-assisted metrology both in the
single shot regime as well as the regime of asymptoti-
cally many repetitions. Sec. 6 considers protocols with
limited resources, where we show that FFQC-assisted
schemes outperform parallel schemes without inter-
mediate control (see (iii)). This is demonstrated for
noisy frequency estimation in Sec. 6.1, and for noisy
phase estimation in Sec. 6.2. We summarize and con-
clude our results in Sec. 7. A reader who is familiar
with quantum metrology, or who is primarily inter-
ested in the results, can directly proceed to sections
4.1 and 4.2.4 where the first two of our main results
((i) and (ii)) are presented.
Accepted in Quantum 2017-09-04, click title to verify 2
2 Quantum metrology protocols
In a standard quantum metrology protocol, a system
consisting of N probes (photons, atoms etc.) is care-
fully engineered, so that each probe may be used to
independently sense a parameter of interest ω. As de-
picted in Fig. 1(a), the experimentalist prepares then
a suitably entangled state of the N probes, which un-
dergo the sensing process in parallel for a time t, dur-
ing which the unknown parameter ω is imprinted on
the state of the probes independently. The final state
of the system is measured in order to most precisely
retrieve information regarding ω.
If the sensing process is noiseless, the strategy (a)
of Fig. 1 is known to optimally achieve the HS [5].
However, various uncorrelated noise types have been
shown to constrain the precision to follow the stan-
dard scaling in the asymptotic N limit, even if one op-
timally prepares the probes in an entangled but noise
robust state [15, 18, 19, 21].
A more powerful metrological protocol is depicted
in Fig. 1(b). Here, in addition to the N sensing
probes, the experimenter is equipped with ancillary
particles that do not take part in the sensing pro-
cess, and may be used to implement single-step error
correction at the final measurement stage [22]. No-
tice that the protocol of Fig. 1(b) is more general
than that of Fig. 1(a); indeed one recovers the lat-
ter by choosing not to entangle the sensing probes
with the ancillae. Consequently any bound on pre-
cision valid for scheme (b) also applies to scenario
(a). Methods have been recently proposed that al-
low to derive bounds on the precision for protocols
(b) from a Kraus representation of a single chan-
nel E
ω,t
[21, 22]. Although these methods allowed
to prove the asymptotic standard scaling for various
noise types, the derived bounds cannot be guaranteed
to be tight (even in the asymptotic N limit). Never-
theless, in case of dephasing and particle-loss noise
types, the corresponding bounds have been shown to
be asymptotically achievable already within scheme
(a) [24, 25, 36]. For amplitude-damping noise, how-
ever, scheme (b) has been shown to give a strictly
better scaling than scheme Fig. 1(a) [37].
An interesting situation is the case of the X-noise: a
noise similar to dephasing but with the generator per-
fectly transversal to the Hamiltonian encoding the pa-
rameter. For a fixed sensing time t the bounds [19, 21]
impose asymptotic standard scaling. However, in fre-
quency estimation one strongly benefits from decreas-
ing the sensing time while increasing the number of
probes N. Indeed, by optimizing the sensing time
of the protocols (a) and (b) for each given N, it was
shown that one may asymptotically beat the standard
scaling and achieve the 1/N
5/3
precision scaling [38].
Moreover, it has been shown that one can even re-
store the HS of precision for such strictly transversal
noise after allowing for fast possibly multi-step error
correction [2730]. This is a particular case of the
metrological scenario depicted in Fig. 1(c). Here, in
addition to employing ancillary particles, the exper-
imenter is capable of freely interjecting the sensing
process with m control pulses that may act on both
probes and ancillae representing, e.g. error correction
steps [2730], dynamical-decoupling pulses [3944] or
any general adaptive feedback scheme [45, 46]. No-
tice that protocol (c) is more general than Fig. 1(b)
and one can obtain the latter from the former sim-
ply by allowing the intermediate operations to be the
identity.
The most general and powerful metrology protocol,
and the main focus of the current work, is the one de-
picted in Fig. 1(d). Here, the experimenter prepares
a suitably entangled state between a single probe and
many ancillae, and is capable of frequently interject-
ing the evolution with FFQC—an arbitrary number,
k, of most general intermediate control operations act-
ing on the overall state of probe-plus-ancillae. More-
over, by choosing k sufficiently large one can ensure
the sensing time spent by the probe in between suc-
cessive FFQC steps to be infinitesimal. It is this ad-
ditional power of frequently interjecting the sensing
process with fast quantum control that we exploit
throughout the remainder of this work to establish
the ultimate bounds on precision. This intermediate
control allows one in fact to modify the noise process,
a possibility that has not been considered in previous
approaches.
To see that this is indeed the most general strategy,
we note that for any protocol (c), with N parallel
probes and total sensing time t, there exists a protocol
of type (d) with total sensing time t
0
= Nt whose
metrological performance is just as good as that of
(c) or better. We note in passing that protocols (c)
and (d) are not equivalent; not only are the timesteps
in protocol (d) assumed to be infinitesimally small,
but protocol (d) allows us to use the same resources
(namely N and t) in a "sequentialized" fashion.
The above metrological schemes apply equally to
the case of Bayesian parameter estimation, i.e., when
considering only a finite number of experimental rep-
etitions and a prior probability p(ω) representing our
knowledge about the parameter ω. Examples of
such Bayesian scenarios include reference frame align-
ment [912], adaptive Bayesian estimation [7, 13, 47],
and single-shot estimation [48, 49].
2.0.1 A remark on error correction
Let us briefly remark that the most general
error correction step (encoding+syndrome read-
out+correction) only requires the dimension of the
space describing the ancillae to be the same as the
system one. Indeed, given that the system containing
the sensing probes has an overall dimension d that is
smaller than the dimension of the ancillary space, any
bipartite pure state of the two can be written using the
Accepted in Quantum 2017-09-04, click title to verify 3
Schmidt decomposition as
P
d
i=1
λ
i
|ii
S
|ψ
i
i
A
. Here,
{|ii}
d
i=1
denotes a given fixed basis for the system,
while {|ψ
i
i}
d
i=1
is a set of orthonormal states of ancil-
lae that spans a d-dimensional subspace. Thus, since
any error correction protocol may be implemented by
restricting to states of the above form, it suffices to
choose the overall space describing the ancillae to be
of same dimension as the system.
3 Background
In order to assess the performance of any of the proto-
cols introduced in Sec. 2, we review below the crucial
tool of quantum metrology—the QFI. In addition, in
Sec. 3.2 we describe the general time-homogeneous
qubit noise processes we consider throughout this
work.
Notation. Before doing so, let us first introduce the
notation we utilize. In what follows, we denote the
derivative with respect to the estimated parameter ω
(whenever applied to operators, vectors or matrices;
differentiating adequately the entries) as
˙
:=
d
dω
; and
shorten the trigonometric functions to c
θ
:=cos(θ) and
s
θ
:= sin(θ). ||A|| := max
ψ
|hψ|A |ψi| stands for the
operator norm, whereas and I for the identity oper-
ator and identity linear map respectively. We use bold
face to denote vectors, so that r := (x, y, z)
T
repre-
sents a 3-dimensional complex vector, while x, y and
z the Cartesian unit vectors. We denote the standard
Pauli 4- and 3-vectors as ς := (σ
0
, σ
1
, σ
2
, σ
3
)
T
and
σ := (σ
1
, σ
2
, σ
3
)
T
respectively (σ
0
:= ). The scalar
and outer products of vectors are then defined in the
usual way irrespectively of the type of vector entries,
i.e., v
v
0
:=
P
i
v
i
v
0
i
and [v
0
v
]
ij
:= v
0
i
v
j
. We write
σ
n
:= n
T
σ to denote a Pauli operator in the spa-
tial direction n = (s
θ
c
φ
, s
θ
s
φ
, c
θ
)
T
. We reserve the
special font, M, for matrices whose entries, M
ij
, are
(complex) numbers, so that the multiplication of all
types of vectors can then be defined in the standard
way: Mv =(
P
j
M
1j
v
j
,
P
j
M
2j
v
j
, . . . )
T
. For example,
any linear qubit map, L, can be unambiguously repre-
sented by constructing its corresponding matrix L (in
the Pauli basis) such that L =ς
Lς =
P
3
µ,ν=0
σ
µ
L
µν
σ
ν
.
3.1 QFI and its role in quantum metrology
In all metrology schemes discussed in Sec. 2 the in-
formation about the estimated parameter of interest
is encoded in the final state of the system consisting
generally of both probes and ancillae. Thus, we de-
note the final state of the system in any of the schemes
(a,b,c,d) of Fig. 1 as %
ω
. It is then %
ω
that is mea-
sured, in order to most precisely construct an estimate
of the parameter of interest, ω.
How best to quantify the precision of this estimate
depends crucially on how the information about ω has
been collected. If we are freely allowed to repeat the
protocol sufficiently many times and collect statistics,
then the best precision of estimation, as quantified by
the mean squared error (MSE), is determined by the
well-known Cramér-Rao bound (CRB) [50, 51]
δ
2
ω
1
νF(%
ω
)
, (1)
where F(%
ω
) is the QFI of the state and ν is the num-
ber of repetitions. On the other hand, if the informa-
tion about ω is inferred from the outcome of just a
single repetition of the protocol, then the single-shot
MSE has to be averaged over the prior distribution
representing the knowledge we possess about ω prior
to its estimation. Then, the average mean squared
error (AvMSE ) can be lower-bounded by either the
Bayesian Cramér-Rao bound (BCRB) [49] (for well
behaved prior distributions) or the Ziv-Zakai bound
(ZZB) [48].
Whether considering the CRB, BCRB, or
ZZB, the crucial quantity of interest is the
QFI: F(%
ω
). Given the spectral decomposition
%
ω
=
P
i
p
i
(ω) |ψ
i
(ω)ihψ
i
(ω)|, the QFI of the state %
ω
evaluated with respect to the estimated parameter ω
generally reads
F(%
ω
) := 2
X
i,j
p
j
+p
j
6=0
1
p
i
(ω) + p
j
(ω)
|hψ
i
(ω)| ˙%
ω
|ψ
j
(ω)i|
2
.
(2)
On the other hand, the QFI may be equivalently de-
fined by using its relation with the fidelity between
quantum states [52], F (ρ, σ):=tr
p
σρ
σ:
F(%
ω
) := 8 lim
dω0
1 F (%
ω
, %
ω+dω
)
dω
2
. (3)
The QFI satisfies the following important proper-
ties. It is additive, so that if the final state of the
system is a product state, i.e., %
ω
= ρ
N
ω
, then its
QFI satisfies F(ρ
N
ω
)= NF(ρ
ω
) and thus gives rise to
standard scaling [3]. If the final state of the probes is
pure, %
ω
=|ψ
ω
ihψ
ω
|, then
F(%
ω
) = 4
˙
ψ
ω
|
˙
ψ
ω
˙
ψ
ω
|ψ
ω
2
. (4)
Moreover, if the parameter is also encoded via
a unitary, so that |ψ
ω
i = e
iωH
|ψi where H is
some Hamiltonian, then F(%
ω
) = 4(hψ|H
2
|ψi
hψ|H |ψi
2
) =: 4 Var
|ψi
(H). In this case the QFI
is maximized by preparing the probes in the state
|ψi =
1
2
(|λ
max
i+|λ
min
i), where
λ
max/min
are the
eigenstates corresponding to the maximum/minimum
eigenvalues of H. Furthermore, if the N probes are
subjected to a unitary evolution generated by a lo-
cal Hamiltonian, e.g., H =
1
2
P
N
n=0
σ
(n)
3
, then such a
state corresponds to the GHZ state and attains the
HS by virtue of the CRB (1) [3]. However, note that
the QFI is important not only for metrology but also
for entanglement detection [5355]; defines a natural
Accepted in Quantum 2017-09-04, click title to verify 4
geometric distance between quantum states [56], thus
serving as a tool to derive speed limits on quantum
evolution [56, 57], as well as constitutes a measure of
macroscopicity [58]. The QFI is the primary focus of
this work and we further discuss its role in quantify-
ing the precision of estimation in metrology protocols
in more detail in Sec. 5.
3.2 Time-homogeneous qubit evolution
The interaction dynamics via which a probe senses
the parameter is generally described by a quantum
channel [59] (see Fig. 1), E
ω,t
, that encodes the pa-
rameter ω onto the state of the probe over the inter-
rogation time t: ρ
ω
(t)=E
ω,t
(ρ). The channel consists
of a unitary evolution part, encoding the parameter
of interest, and a non-unitary (noise) part describ-
ing additional system-environment interactions. In
this work we shall consider that the non-unitary part
arises due to uncorrelated noise processes described
by a time-homogeneous master equation of Lindblad
form—often referred to as the semigroup dynamics
[60]—which provides an appropriate description for
most physically relevant noise processes. In particu-
lar, in such a typical setting the environment is as-
sumed to be sufficiently large such that it disturbs
the system in the same fashion independently of the
system state and the time instance [61]. Thus, the
time-homogeneity of evolution allows us to unambigu-
ously apply the FFQC techniques as any operation
applied on the system does not affect the environ-
ment. Moreover, one may then freely swap the sensing
particles at any time with the ancillary ones, which
experience afterwards the same fixed noisy dynam-
ics. In contrast, noise processes described by time-
inhomogeneous master equations have been recently
considered within the context of quantum metrology
in [6265], where time-inhomogeneity has been shown
to be beneficial at short time-scales— the so-called
Zeno regime—where the bath can no longer be as-
sumed to be uncorrelated from the system. Let us
stress that such a regime, however, does not allow for
the general control operations considered here to be
unambiguously applied without explicitly modelling
the environment [35, 66].
We describe the evolution of a single qubit probe
by following time-homogeneous master equation
dρ
ω
(t)
dt
= i
ω
2
[σ
3
, ρ
ω
(t)] + L(ρ
ω
(t)), (5)
where ω is the parameter to be estimated and the
noise is given by the Lindblad super-operator, L. The
Liouvillian is then generally defined as
L(ρ) :=
1
2
3
X
µ,ν=0
L
µν
([σ
µ
ρ, σ
ν
] + [σ
µ
, ρσ
ν
]) , (6)
where L is the matrix representation of the Lindblad
superoperator – an Hermitian positive semi-definite
matrix whose entries are independent of t and ω.
We note that all the noise terms L
0i
and L
i0
can be
straightforwardly corrected within FFQC as they lead
to an additional Hamiltonian evolution term in the
Eq. (5), sometimes referred to as Lamb-shift. This
ω-independent Hamiltonian term can be cancelled by
continuously applying the inverse unitary rotation. In
what follows we assume that the control operations
already incorporate such unitary and will, thus, be
concerned with the restriction of L to the subspace
spanned by {σ
1
, σ
2
, σ
3
}, which we denote as
¯
L.
We can group all the relevant noise processes into
three important families:
Rank-one Pauli noise of strength γ
L
1P
n
(ρ) :=
γ
2
(σ
n
ρσ
n
ρ), (7)
with the particular case of dephasing noise L
1P
z
for
which σ
z
= σ
3
.
General rank-one noise L
1G
r
defined by the matrix
¯
L
1G
r
:= r r
, (8)
where r = (x, y, z)
T
and the strength of the noise
is denoted γ/2 = |r|
2
. The special case Re(r) ×
Im(r) = 0 corresponds to the rank-one Pauli noise
of Eq. (7), whereas whenever |Re(r)|= |Im(r)| and
Re(r)
T
Im(r) = 0 an amplitude damping channel is
recovered, which represents spontaneous emission
along some particular direction.
Rank-two Pauli noise L
2P
defined by the matrix
¯
L
2P
:=
1
2
R
T
γ
1
γ
2
0
!
R
, (9)
where SO(3) and R
= R
z
(ϕ)R
y
(θ)R
z
(ξ) is
its matrix representation written in the Euler form
with angles (ϕ, θ, ξ). The special case of R
= cor-
responds to asymmetric X-Y noise, which we con-
veniently parametrize with γ
1
=γ p and γ
2
=γ(1p)
for 0p1 to define
L
X-Y
(ρ) :=
γ
2
(p σ
1
ρ σ
1
+ (1 p)σ
2
ρ σ
2
ρ).(10)
In order to explicitly determine the form of the
quantum channel describing the qubit evolution, one
has to integrate Eq. (5) [67]. The resulting dynamics
can then be expressed, e.g, with help of the corre-
sponding dynamical matrix, S (specified in the Pauli
operator basis), or via a Kraus representation as fol-
lows:
E
ω,t
(ρ) =
3
X
µ,ν=0
S
µν
(ω, t) σ
µ
ρ σ
ν
(11)
=
X
i
K
i
(ω, t) ρ K
i
(ω, t)
. (12)
Note that the Kraus representation of Eq. (12) is not
unique as starting from a set of r linearly independent
Accepted in Quantum 2017-09-04, click title to verify 5
Kraus operators, K = (K
1
, . . . , K
r
)
T
, we may simply
construct another valid set K
0
=uK choosing u to be
any (potentially ω-,t- dependent) r×r unitary matrix.
4 Noisy metrology with full and fast
control
In this section we derive the main results of our work.
In Sec. 4.1 we outline the optimal FFQC proto-
col suitable for correcting rank-one Pauli noise that
is not parallel to the Hamiltonian (7). In Sec. 4.2
we introduce the necessary tools required to bound
the QFI for FFQC (Fig. 1(d)), and show that any
other noise unavoidably leads to a linear scaling of the
QFI. We exemplify our findings by deriving analyti-
cal bounds for the QFI for several physically relevant
noise-types. Finally, in Sec. 4.3 we discuss different
aspects in which FFQC allows to outperform scenar-
ios without fast control, using the cases of X-Y noise
(7) and transversal rank-one Pauli noise (40c) as ex-
amples.
4.1 Removing rank-one Pauli noise
Here we construct a general FFQC strategy that al-
lows one to correct for any rank-one Pauli noise L
1P
n
introduced in Eq. (7) with n 6= z. For convenience,
we perform a change of basis so that the noise is gen-
erated by σ
1
(n=x), whereas ω is encoded via a ro-
tated Hamiltonian σ
θ
:= s
θ
σ
3
+ c
θ
σ
1
. We consider
the qubit probe to be aided by an ancillary qubit
and define the two-qubit code space as a subset of
H
S
H
A
H
C
:= span{|00i, |11i}, with correspond-
ing projector Π
C
:= |00ih00| + |11ih11|, and the error
space, H
E
:=span{|01i, |10i}, with Π
E
:= Π
C
.
Let the probe-plus-ancilla be prepared in a pure
state % = |ψihψ| with |ψi= α |00i + β |11i H
C
. The
dynamics on the probe-plus-ancilla is then described
by Eq. (7) with only the probe system evolving
d%
dt
= i
ω
2
[σ
θ
, %] + L
1P
x
I
A
(%). (13)
Integrating Eq. (13) over an elementary timestep dt,
we may write the probe-plus-ancilla state, up to first
order in dt, as
%(dt) = %(1
γ
2
dt) i
ω
2
[σ
θ
, %]dt (14)
+
γ
2
(σ
1
)%(σ
1
)dt + O(dt
2
).
Projecting %(dt) onto the code and error subspaces
yields the unnormalized states %
C
(dt) = Π
C
%(dt
C
and %
E
(dt)= Π
E
%(dt
E
given by
%
C
(dt) = %(1
γ
2
dt) i s
θ
ω
2
[σ
3
, %]dt + O(dt
2
),
%
E
(dt) =
γ
2
(σ
1
)%(σ
1
) dt + O(dt
2
) (15)
respectively.
If an error is detected, we simply apply σ
1
on the
sensing probe (so after correction ¯%
E
(dt) := (σ
1
)%
E
(dt)(σ
1
)), and otherwise do nothing. Hence,
the state of probe-plus-ancilla after an infinitesimal-
timestep followed by fast error correction is the mix-
ture
¯%(dt) := %
C
(dt) + ¯%
E
(dt)
= % i s
θ
ω
2
[σ
3
, %] dt + O(dt
2
), (16)
which, to first order in dt, is equivalent to a unitary
evolution under the projected Hamiltonian
ω
2
s
θ
σ
3
. As
the measurement may always be adjusted to compen-
sate for a known rotation of the Hamiltonian encoding
the parameter, the above strategy perfectly corrects
rank-one Pauli noise (7) at the price of slowing the
evolution down by a factor s
θ
=
p
1 (n
T
z)
2
.
We summarize the above result in the following ob-
servation:
Result 1. A general FFQC strategy allows one to
completely eliminate the impact of any rank-one Pauli
noise L
1P
n
(7) that is not exactly parallel to the
parameter-encoding Hamiltonian. The resulting dy-
namics of the system are then described by a noise-
less evolution with a rotated Hamiltonian and the es-
timated parameter being rescaled to
p
1 (n
T
z)
2
ω,
which still yields HS in precision.
Note that in the above derivation nothing forbids us
from replacing the scalar parameter ω with any opera-
tor B acting on an additional system described by the
Hilbert space H
B
. Consequently, one can effectively
modify the noisy dynamics of any %
BS
B(H
B
H
S
):
d%
BS
dt
= i
1
2
[B σ
θ
, %
BS
] + I
B
L
1P
x
(%
BS
), (17)
by adding an ancillary qubit and implementing the
FFQC strategy described above. Then, the dynamics
of the error-corrected ¯%
BS
= tr
A
{¯%
BSA
} is governed
by
d¯%
BS
dt
= i
p
1 (n
T
z)
2
2
[B σ
3
, ¯%
BS
], (18)
with the rank-one Pauli noise removed. Hence, our
strategy can be directly used for the implementation
of unitary gates U = e
ig Bσ
3
in the presence of rank-
one Pauli noise on the qubit.
4.2 Other noise-types: unavoidable linear scal-
ing
In this subsection we prove our main no-go result,
F
L
4 α
L
t
0
, showing that even with FFQC the QFI
is bound to a linear scaling for all uncorrelated noise-
types other than rank-one Pauli noise, and to min-
imize the constant α
L
appearing in this bound for
several practically relevant noise-types introduced in
Accepted in Quantum 2017-09-04, click title to verify 6
Sec. 3.2. These results are presented in Sec. 4.2.4
and Sec. 4.2.5 respectively. The preceding sections
are devoted to elaborate the method that allows us to
derive these results. In Sec. 4.2.1 and Sec. 4.2.2 we
describe the channel extension (CE) method used to
upper bound the QFI [18, 21, 22, 37] and then review
the bounds it yields for strategies (a), (b) and (c) of
Fig. 1. However, the CE method fails to yield a mean-
ingful bound when considering the limit in which the
sensing process lasts for an infinitesimal time-step. In
Sec. 4.2.3 we resolve this issue by amending the CE
method in order to obtain an upper bound on the QFI
for strategies that employ FFQC.
We stress that we are able to derive analytic bounds
by extending the CE method to the limit dt 0,
in which the problem actually simplifies: one does
not have to solve the master equation and determine
the full dynamics! This is even more remarkable as
these bounds apply to any metrological scheme (see
the hierarchy in Fig. 1)
4.2.1 Upper bounds on the QFI
We would like to compare the maximum achievable
QFI in the various metrology schemes introduced in
Sec. 2 and, in particular, quantify the improvement (if
any) between protocols (b), (c) and (d). To this end
we use variations of the CE method [18, 21, 22, 37]
that allow one to upper-bound the QFI depending
on the scenario considered. Crucially, they constrain
the QFI using solely the properties of the quantum
channel responsible for encoding the parameter.
CE bound. The QFI of a state representing the
output of a quantum channel encoding the parameter
may always be upper-bounded purely by the channel
properties after performing a channel extension (CE).
Given a probe in a state ρ and a quantum channel E
ω,t
one may always construct the CE bound on the QFI
of the output state E
ω,t
(ρ) [18]:
F(E
ω,t
(ρ)) max
%
F(E
ω,t
I
A
(%)) 4 ||α(t)||,
(19)
where %B(H
S
H
A
) represents a joined (extended)
state of the probe and an ancilla, and the operator
α(t) :=
X
i
˙
K
i
˙
K
i
=
˙
K
˙
K (20)
for any of the Kraus representation Eq. (12) of the
probe channel. Here,
˙
K
i
are the derivatives of the
Kraus operators with respect to ω (in what follows,
we write explicit dependencies of the Kraus operators
on t and ω only if necessary)
Although the first inequality in Eq. (19) may not
be tight—entangling the probe and ancilla may in-
crease the achievable QFI—the second is. A valid
Kraus representation of E
ω,t
for which the second in-
equality is saturated is guaranteed to exist as long
as dim(H
A
) dim(H
S
) [18]. In consequence, the CE
bound (19) directly determines the maximal attain-
able QFI for the ancilla-assisted parallel scheme de-
picted in Fig. 1(b) with N =1 [22].
Parallel CE bound. The CE bound (19) may be
directly applied to upper-bound the ultimate attain-
able QFI of any parallel scheme of Fig. 1(b) that em-
ploys N probes and ancillae. In scenario (b), the final
state generally reads %
(b)
N
(t)= E
N
ω,t
I
A
(%
(b)
N
). Hence,
applying the CE bound (19) to channel E
N
ω,t
I
A
that describes the overall evolution and fixing a ten-
sor product Kraus representation such that
E
N
ω,t
(ρ
N
) =
X
i
1
,...,i
N
K
i
1
...K
i
N
ρ
N
K
i
1
...K
i
N
, (21)
one obtains the parallel CE bound [18, 21, 22]:
F
%
(b)
N
(t)
max
%
(b)
N
F
E
N
ω,t
I
A
(%
(b)
N
)
4N||α(t)|| + 4N(N 1)||β(t)||
2
, (22)
where %
(b)
N
B(H
N
S
H
A
) describes the initial
probes-plus-ancillae state and
β(t) := i
X
i
˙
K
i
K
i
= i
˙
K
K (23)
which depends solely on Kraus operators and their
derivatives. In order to obtain Eq. (22) from Eq. (19)
note that due to the tensor product structure the term
˙
K
i
˙
K
i
=
d
dω
(K
i
1
... K
i
N
)
d
dω
(K
i
1
... K
i
N
) in
Eq. (19) contains N terms for which in both parenthe-
sis the derivative applies on the Kraus operator acting
on the same probe and N(N 1) terms where it ap-
plies on the Kraus operators acting on two different
probes [18]. Yet, in contrast to Eq. (19), the par-
allel CE bound (22) is then not generally attainable,
even after optimizing Eq. (22) over all Kraus represen-
tations of the single-probe channel. This is because
the Kraus decomposition for the map E
N
ω,t
I
A
that
makes Eq. (19) saturable might not be of a tensor-
product form, as assumed in Eq. (21).
Sequential CE bound. Now, we would like to ap-
ply the CE methods to the two strategies involving
control, i.e., (c) and (d) of Fig. 1. However, as argued
in Sec. 2, any scheme of type (c) employing N probes
with m control steps (each lasting δt) can always be
mimicked by a scheme Fig. 1(d) employing a single
probe but lasting N-times longer, i.e., with t
0
= Nt
and involving more k = N m steps each still of dura-
tion δt. Hence, we may always upper-bound the QFI
of any scheme depicted in Fig. 1(c) as
F
%
(c)
N,m
(t)
max
schemes (d)
k=N m, t
0
=Nt
F
%
(d)
k
(t
0
)
(24)
Accepted in Quantum 2017-09-04, click title to verify 7
where the maximization is over all sequential schemes
of type (d) with step duration δt = N/m. Thus, in
what follows, we focus on applying the CE methods
to the scenario of Fig. 1(d). Note that by decreasing
δt further, or equivalently by raising the number of
control steps k, we may only increase the right hand
side of Eq. (24). Hence, the optimal FFQC scheme
yielding the maximum in Eq. (24) must correspond
to the limit of k , or equivalently, δt 0. It
is so, as any FFQC scheme of step duration δt can
be mimicked by a protocol with shorter δt
0
= δt/k
(and any k 2) and setting some of the intermediate
control operations to be trivial, i.e., the identity.
For sequential strategies employing intermediate
control a CE-based upper bound on the correspond-
ing QFI has recently been derived in Ref. [37]. Let the
output state for the sequential protocol of Fig. 1(d)
be given as
%
(d)
k
(t
0
) =
X
i
(k)
K
i
(k)
(t
0
) %
(d)
K
i
(k)
(t
0
) (25)
with %
(d)
B(H
S
H
A
) being the initial state. How-
ever, in contrast to Eq. (21), the overall Kraus oper-
ators
K
i
(k)
(t):= (K
i
k
(δt)
A
)U
k
. . . (K
i
1
(δt)
A
)U
1
(26)
have a composition rather than tensor-product struc-
ture, where U
`
stands for the a control operation act-
ing on both probe and ancillae at the `th step , while
K
i
`
are the Kraus operators of the channel E
ω,δt
act-
ing solely on the probe at the `th step. An application
of the general CE bound (19) to this case was derived
in Ref. [37], leading to the sequential CE bound on
the QFI attainable using any sequential scheme with
FFQC (with timestep δt):
F
%
(d)
k
(t
0
)
4k||α(δt)||+ (27)
+ 4k(k 1)||β(δt)||(||α(δt)|| + ||β(δt)|| + 1),
which applies to any protocol (d) of Fig. 1 with time-
steps of duration δt, and hence also to all schemes (c)
of Fig. 1 upon substituting k =N m.
Notice that, just as in the case of the parallel CE
bound (22), there may not exist a Kraus representa-
tion of the single-probe channel such that the sequen-
tial CE bound of Eq. (27) is guaranteed to be tight.
This holds also for the optimal sequential strategy (d)
that maximizes the corresponding QFI (i.e., the one
maximizing the right hand side of Eq. (24)).
4.2.2 Optimization of the CE bounds
All the CE bounds presented this far: stan-
dard (Eq. (19)), parallel (Eq. (22)) and sequential
(Eq. (27)); rely solely on the structure of the single-
probe channel E
ω,t
encoding the parameter and its
particular Kraus representation (12). Thus, in order
to obtain the tightest versions of these bounds one
should minimize them over all equivalent Kraus rep-
resentations, K
0
= uK (with u
u = ), describing the
single-probe dynamics. Although such a gauge free-
dom is generally parameter dependent, i.e., u = u(ω),
the operators α (20) and β (23) depend solely on
h := i u
˙
u. Hence, it always suffices to search only
through Kraus representations satisfying K
0
=K and
˙
K
0
=
˙
KihK or, in other words, any CE-based bound
considered may always be optimized by simply per-
forming its minimization over all Hermitian matrices
h [18].
In contrast to the sequential CE bound (27),
the standard (19) and parallel (22) CE bounds are
quadratic (convex) in h, and thus may be minimized
numerically in a systematic manner by means of semi-
definite programming (SDP) given some K [21, 22].
On the other hand, when considering t (or δt in case
of Eq. (27)) to be fixed and the asymptotic limit of
N (22) (N m (24), or k (27)), it is always optimal, if
possible (contrary to the unitary case), to set β(t)=0
in the CE bounds, so that they asymptotically be-
come linear in N (Nm, or k respectively). In such a
regime, as the constraint β(t) = 0 is linear in h, one
may always apply the SDP methods to minimize the
remaining coefficient 4||α(t)||, which is then emergent
in all the bounds [21, 22].
Equivalently, the same argumentation can be made
starting from the dynamical matrix S of the channel
E
ω,t
, defined in Eq. (11), instead of its Kraus represen-
tation. The two are directly related (see also App. A),
as any valid vector of Kraus operators K can always
be expressed in a complete operator basis, e.g. the
Pauli basis in case of a single-qubit maps, via a ma-
trix M satisfying
K = Mς and M
M = S
T
. (28)
As a result, one may rewrite the operators α (20)
and β (23) with help of the matrix M and a general
Hermitian matrix h to be optimized as
α(t) = ς
(
˙
M
+ iM
h) (
˙
M ihM) ς , (29)
β(t) = ς
(
˙
M
+ iM
h) M ς . (30)
In what follows, we always choose M = M
=
S
T
and thus unambiguously fix the starting Kraus rep-
resentation K in Eq. (28). We then make use of
Eqs. (29) and (30) in order to analytically optimize
over h the infinitesimal-timestep version of the sequen-
tial CE bound (27) as we now explain.
4.2.3 Infinitesimal-timestep CE bound
In the most powerful FFQC setting of Fig. 1(d), one
allows for the control operations to be arbitrarily fast
and thus assumes the steps of the protocol to last an
infinitesimal duration dt. Hence, one deals then with
a sequential strategy of infinitesimal step duration, for
which the sequential bound of Eq. (27) can be directly
Accepted in Quantum 2017-09-04, click title to verify 8
applied after setting k
t
0
dt
and taking the limit δt
dt0 (see App. B). In order for the bound to remain
meaningful, we adjust the derivation of Ref. [37] in
App. B, so that it now reads:
F
%
(d)
t
0
/dt
(t
0
)
4
t
0
dt
||α(dt)|| (31)
+4
t
0
dt
2
||β(dt)||
1
dt
||α(dt)||+||β(dt)||+
dt
.
We refer to Eq. (31) as the infinitesimal-timestep CE
bound.
In order to derive the ultimate precision bounds
valid in the presence of noise and FFQC, we compute
the infinitesimal-timestep CE bound (31) after substi-
tuting the expressions for α and β given by Eqs. (29)
and (30). We explicitly minimize Eq. (31) over Kraus
representations as described in Sec. 4.2.2, by expand-
ing the Kraus operators, K(dt), and the Hermitian
matrix, h(dt), up to small orders of
dt. As we are
interested in the regime of asymptotic resources, i.e.,
the limit t
0
dt (or equivalently N tδt in case of the
scheme (c)), we may assume without loss of general-
ity that it is always optimal to make the second term
in Eq. (31) vanish by setting β(dt) = 0. We outline
the methods we use below, with detailed calculations
deferred to App. C.
Consider the dynamical matrix S, defined in
Eq. (11), of the infinitesimal time channel E
ω,dt
. Ex-
panding it in dt around t = 0 gives S(dt) := S
(0)
+
S
(1)
dt+O(dt
2
). The evolution of the probe after an
infinitesimally short time is then given by
E
ω,dt
(ρ) =
3
X
µ,ν=0
S
µν
(dt) σ
µ
ρ σ
ν
. (32)
=
3
X
µ,ν=0
(S
(0)
µν
+ S
(1)
µν
dt) σ
µ
ρ σ
ν
+ O(dt
2
).
The above expansion is unambiguously specified up
to O(dt
2
) by the master equation in Eq. (5)
E
ω,dt
(ρ) = ρ+
i
ω
2
[σ
3
, ρ] + L(ρ)
dt+ O(dt
2
), (33)
so we may directly relate the short-time expansion
of S to the Liouvillian of Eq. (6) without explicitly
integrating Eq. (5):
S
(0)
= diag(1, 0, 0, 0), (34)
S
(1)
=
P
3
i=1
¯
L
ii
Im
¯
L
23
Im
¯
L
31
Im
¯
L
12
+ i
ω
2
Im
¯
L
23
Im
¯
L
31
¯
L
Im
¯
L
12
i
ω
2
,
(35)
where
¯
L is just the restriction of the Liouvillian L (6)
to the subspace spanned by the Pauli operators. We
remark that in the limit dt 0 these are the only
meaningful terms, as higher orders in the expansion
do not affect the final state (which is precisely the
reason why the master equation formalism is valid).
From Eqs. (34) and (35) we directly obtain the the
expansion of M in
dt, using the convenient choice
M= M
=
S
T
.
Next, we expand the infinitesimal-timestep CE
bound (31) in orders of
dt by defining the ex-
pansions for operators α(dt) :=
P
`0
α
(`)
(dt)
`/2
and
β(dt) (similarly). The definitions Eqs. (29) and (30)
of α(dt) and β(dt) allow one to relate their expansions
to the ones of the matrix M (fixed by the Liouvillian
via Eqs. (34) and (35)) and the Hermitian matrix h
(which is still free) order by order. In the last step one
optimizes the matrix h in order to obtain the tight-
est infinitesimal-timestep CE bound (31) in the limit
dt0.
We include the full details in App. B, yet let us
summarize here that a non-trivial bound in Eq. (31)
can only be obtained if the expansion terms satisfy
α
(0)
= α
(1)
= β
(0)
= β
(1)
= 0, these condition are al-
ways satisfied if the low-order terms in the expansion
of the matrix h are constrained. If it is possible to
choose h such that β
(2)
= β
(3)
= 0 also hold then the
infinitesimal-timestep CE bound (31) becomes
F
%
(d)
t
0
/dt
(t
0
)
4||α
(2)
||t
0
+ (t
0
)
2
O(
dt). (36)
bounding the QFI to a linear scaling in t
0
in the limit
dt 0 with k = t
0
/dt. Note that β
(3)
can always be
set to zero without affecting lower order terms, while
the existence of a solution for β
(2)
= 0 depends on the
noise, in particular it is impossible for rank-one Pauli
noise as implied by the result of Sec. 4.1. If there
exists a matrix h such that α
(0)
= α
(1)
= β
(0)
= β
(1)
=
β
(2)
=0 we obtain
F
%
(d)
t
0
/dt
(t
0
)
dt0
4 α
L
t
0
with (37)
α
L
:= min
h such that α
(i)
=β
(j)
=0
||α
(2)
||, (38)
for i 1 and j 3.
Note that as the sequential scheme with FFQC is
the most powerful one (see Sec. 2), Eq. (37) holds for
all protocols depicted in Fig. 1 after accordingly set-
ting t
0
= Nt. In particular, Eq. (37) also applies in
case of control with finite time-step δt (scheme (c))
and in the absence of quantum control when δt = t
(schemes (a) and (b)). Finally, let us remark that
although in the case of parallel strategies with ancil-
lae of Fig. 1(b) it is the parallel CE bound (22) that
must yield tighter limits on precision for finite t, one
may show (following exact argumentation as in the
previous paragraphs) that the linearly scaling parallel
CE bound must always converge to its infinitesimal-
timestep equivalent of Eq. (37) when fixing t
0
= Nt
and considering the t 0 limit, in which the parallel
protocol becomes infinitesimally short but involves an
infinite number of probes.
Accepted in Quantum 2017-09-04, click title to verify 9
4.2.4 Universal asymptotic linear bound
Thanks to the infinitesimal-timestep CE bound (31)
yielding Eq. (37), we may formulate a general obser-
vation about the scaling of the QFI applicable to any
of the schemes depicted in Fig. 1:
Result 2. For all noise processes described by Eq. (6)
except rank-one Pauli noise-types that are not parallel
to the parameter-encoding Hamiltonian, the QFI in
any scheme of Fig. 1 is upper-bounded by
F
L
4 α
L
t
0
, (39)
where α
L
is a constant that solely depends on the par-
ticular form of the Liouvillian L defined in Eq. (6),
while t
0
is the effective protocol time as defined via the
most powerful scheme (d) with FFQC in Fig. 1.
We explicitly prove Res. 2 in App. C, but outline
the derivation here. In App. C.1, for any qubit Liou-
villian L of Eq. (6), except the rank-one Pauli noise of
(7) with σ
n
6= σ
3
, we give the form of the hermitian
matrix h for which α
(0)
= α
(1)
= β
(0)
= β
(1)
= β
(2)
=
β
(3)
=0. Hence, there exists some finite α
L
for which
(37) holds. We first show that dephasing noise, as
well as any rank-one noise which is not of Pauli type,
imposes a linear scaling of the QFI due to Eq. (37).
Next, we consider noise processes described by Liou-
villians whose rank is strictly greater than one. Any
such noise process corresponds to a matrix represen-
tation of the Lindblad superoperator (6) that may be
written as a direct sum of matrices of lower rank. We
show that the QFI of noise processes of rank greater
than one is always smaller than the QFI of any of
its orthogonal components. Thus, the only rank-two
noise processes that, in principle, could still allow for
quadratic scaling of the QFI are the ones correspond-
ing to a direct sum of two orthogonal rank-one Pauli
noises. For this case, however, we explicitly show that
the QFI scales linearly with t
0
. Moreover, this also set-
tles the case of the rank-three noise types as any such
noise can be expressed as a direct sum that contains
either a rank-one non-Pauli or a rank-two Pauli noise,
or both. Finally, in App. C.2 we derive explicit forms
of α
L
for the exemplary noise-types discussed below.
4.2.5 Exemplary noise-types
Although we have shown the validity of Eq. (39) for
most types of noise processes, it still remains to com-
pute α
L
for a general Liouvillian (see Eq. (6)) by ade-
quately minimizing Eq. (38). In App. C.2 we perform
such minimization in a completely analytic manner.
In particular, we minimize ||α
(2)
|| in Eq. (37) over
Hermitian matrices h (under the constraint of β
(2)
=0
and α
(`)
= β
(`)
= 0 for ` 1) for all the exemplary Li-
ouvillians stated in Sec. 3.2. The corresponding CE
bounds for dephasing, general rank-one, and rank-two
Pauli noise types respectively read
F
1P
z
t
0
2γ
(40a)
F
1G
r
t
0
4
max{|x + iy|, |x iy|}
2
(|Re(r)|
2
|Im(r)|
2
(Re(r)
T
Im(r))
2
)
(40b)
F
2P
t
0
2γ
1
γ
2
c
2
θ
(γ
1
+ γ
2
) + s
2
θ
(γ
1
s
2
ϕ
+ γ
2
c
2
ϕ
)
(40c)
= F
X-Y
t
0
2γ p(1 p)
(40d)
We note that in the case of general rank-one noise,
yielding F
1G
r
in Eq. (40b), we analytically find the
optimal h for the cases where one of the components
of r (x, y, or z) vanishes (see App. C.2). This allows
us to use the structure of the optimal h as an ansatz
for the case of general rank-one noise L
1G
. However,
although the minimization of ||α
(2)
|| subject to the
constraint β
(2)
= 0 can be done for any fixed Liouvil-
lian, it generally contains a large number of param-
eters to be optimized. Hence, we derive a valid, but
not provably tightest, analytical bound for the case of
general noise.
4.3 The gain allowed by fast control
We now discuss the impact of fast control, in particu-
lar the gap in the achievable QFI between the scheme
(b) and the schemes with FFQC, i.e., (c) and (d) (with
δt 0 throughout this section) in Fig. 1. To do the
comparison we focus on the example of the X-Y noise
L
X-Y
in Eq. (10), and the rank-one transversal Pauli
noise L
1P
x
in Eq. (7) with σ
n
= σ
1
, which is also the
limiting case of X-Y noise for p = 1. For X-Y noise
the master equation (5) can be analytically solved,
hence the parallel CE bound Eq. (22) allows one to
upper-bound the QFI of the parallel scheme F
(b)
(t)
(see App. D).
4.3.1 Transversal rank-one Pauli noise
We begin with the rank-one transversal Pauli noise
L
1P
x
= L
X-Y
|
p=1
. From Res. 1 and from previous re-
sults [2730], we know that this noise can be com-
pletely removed by FFQC without harming the evolu-
tion. Consequently, the strategies with FFQC attain
the QFI given by F
(d)
(t
0
) = (t
0
)
2
and F
(c)
(t) = (N t)
2
respectively. Remarkably, in order to attain such QFI
within (d) one only requires one qubit and one ancilla
in parallel: this is enough to implement the FFQC
strategy of Sec. 4.1 and get rid of the noise, and it
is known that in the noiseless case a sequential single
qubit strategy is optimal [5]. For the parallel strat-
egy (b) the CE bound reads F
(b)
4N α
(b)
ω,p=1
(t) (see
App. D). We plot 4 α
(b)
ω,p
for p = ω = γ = 1 in Fig. 2
Accepted in Quantum 2017-09-04, click title to verify 10
(thin solid line). This clearly shows the advantage of-
fered by FFQC: for any fixed t in strategy (b) the at-
tainable F
(b)
is bound to a linear scaling in N. More-
over, for any γ and ω as t increases the bound even-
tually starts to decrease (actually it has been shown
that F
(b)
(t)/t is maximized for t 1/N
1
3
[38]). Inter-
estingly, the bound α
ω,0
diverges when t approaches
zero. This is to be expected, as in this limit a super-
linear scaling in N is known to be possible already
within scenario (a) [38, 68].
4.3.2 X-Y noise
Now let us turn to X-Y noise L
X-Y
(with 0 < p <
1). For the parallel strategy (b), the same bound
F
(b)
4Nα
(b)
ω,p
(t) applies and the function α
(b)
ω,p
(t)
is now well-behaved (it starts at zero for t = 0). We
plot α
(b)
ω,p
(t) for ω = γ = 1 and p = 0.1 in Fig. 2
(thick solid line). The situation for FFQC schemes
is a bit more subtle, and allows to nicely illustrate
different aspects in which fast control is helpful. First
there is the ultimate bound of Eq. (40c) which holds
for any FFQC strategy and reads F
(d)
X-Y
t
0
2γp(1p)
(F
(c)
X-Y
Nt
2γp(1p)
for scheme (c)) but might not be
attainable (dashed line in Fig. 2).
Now let us consider a particular FFQC strategy,
which is described in details in Sec. D.1. Using the
error correction code of Sec. 4.1 one can detect if an
error happens at a given timestep. However, as there
is no way to tell if the error corresponds to a σ
1
or
σ
2
the noise cannot be fully corrected. Still, one can
chose to correct the most probable error term, say p
1/2 and it is σ
2
. Hence, if the guess was right the error
is corrected σ
2
σ
2
= , while the other case leads to a
z-error σ
2
σ
1
= iσ
3
. This strategy modifies the X-Y
noise process to an announced dephasing noise (one
knows how many correction steps were performed),
for which the same bound Eq. (40d) holds as all our
control operations are unitaries.
Next, consider a suboptimal variant of the above
scheme (we will show that this is strictly subopti-
mal in Sec. 6) where we forget the error register, i.e.,
the knowledge about the number of correction steps.
In this case the resulting effective noise process is
the usual dephasing (L
1P
z
of Eq. (7)) with strength
γ
z
= p γ, and the bounds F
(d)
t
0
2γp
and F
(c)
tN
2γp
of Eq. (40a) apply (dotted line in Fig. 2). Notice that
these bounds are attainable; it is known that the CE
bound for dephasing may be asymptotically achieved
already within strategy (a) when considering the limit
of short dynamics t 0 and large number of probes
N [24, 25, 36]. Crucially, when considering the
more powerful strategies (c) and (d), such a regime
can always be mimicked by employing an unbounded
number of ancillary qubits. These are then continu-
ously swapped with the sensing probes and measured
so that the required protocol of type (a) is indeed re-
0 20 40 60 80
500
1000
1500
2 4
60
100
200
300
t
Figure 2: X-Y noise: Scaling of the QFI as function of t for
various protocols and ω = 1, γ = 0.1 and p = 0.1. The solid
line corresponds to the CE bound on F
(b)
(t)/N for strat-
egy (b) without FFQC (the original X-Y noise). The dashed
straight is the ultimate upper-bound
t
2γp(1p)
Eq. (40c). The
dotted straight line corresponds to the attainable bound for
the particular FFQC strategy of Sec. 4.3.2 where one cor-
rects the dominant noise but discards the error record, it
corresponds to
t
2γp
from Eq. (40a). The dot-dashed line is
the attainable bound for the same strategy but with fixed
number of ancillae equal to N and no intermediate measure-
ments. X-noise: Finally, the thin line is the CE bound on
F
(b)
(t)/N for the X-noise with p = γ = ω = 1, note that it
diverges as t 0 since strategy (b) allows to super-classical
scaling in N for vanishing t. The inset is a zoom on the
region close to the origin.
covered.
A natural question to ask then is what happens if
we keep the number of ancillary qubits constant, and
equal to N in (c) or (d). In this case, the FFQC strat-
egy described above allows one to mimic the scheme
(a) with dephasing noise L
1P
z
of strength γ
z
= p γ
and a fixed number of probes N. For such an effec-
tive noise-type, the parallel CE bound is known to
yield F
(a)
(t) N
t
2
η
2
(1η
2
)
with η = e
t
[22], which is
attainable when N is large (dot-dashed line in Fig. 2).
Hence, it is also attainable by strategy (c) with the
number of ancillary qubits equal the number of probes
N, or by strategy (d) with N + 1 qubits in total (only
one qubit is needed as an ancilla to run the error cor-
rection protocol, while the others can be used for con-
secutive preparation of the required initial entangled
states).
4.3.3 Advantages of FFQC
In summary, we see that FFCQ offers several advan-
tages. First, it allows to modify the noise process par-
ticularly rank-one transversal Pauli noise which can
be completely removed, while X-Y noise can be trans-
formed to a weaker dephasing noise (dot-dashed line
in Fig. 2). In addition to this, if the number of avail-
able ancillary qubits is large FFQC offers the possibil-
ity to prolong the short-time dynamics by “paralleliz-
Accepted in Quantum 2017-09-04, click title to verify 11
ing the evolution” (dotted line in Fig. 2)—mimicking
a parallel evolution with short running time but large
N and the modified noise. Finally, it also gives the
possibility to keep the register of errors that happened
during the evolution (dashed line in Fig. 2). To get
a feeling of how the latter improves the QFI, think
of the frequentist scenario where for the construction
of the final estimator one values more the runs where
less or no errors happened (we explore this in more
depth in Sec. 6). In any case, FFQC undoubtedly out-
performs the parallel scheme (b) (solid line in Fig. 2)
by a large amount.
We showed how the introduction of FFQC leads
to an improved scaling of the QFI. Yet an interest-
ing question is if there exists a gap between the two
strategies (c) and (d). If the frequency with which one
applies the controls in (c) is limited, i.e. the duration
of each step δt is fixed, it is easy to see that there
exists a gap. The bound derived in the App. D also
applies to this case and yields F
(c)
4Nm α
(b)
ω,p
(δt),
and we just demonstrated that it can be outperformed
by FFQC. However, it is not clear if there is a gap
between (c) and (d) when δt 0 and t = t
0
/N. Fur-
thermore, it is not clear these strategies outperform
scheme (b) for t 0 with N =
t
0
t
, as all the CE
bounds coincide in this limit. On the other hand, the
fact that the CE bounds coincide does not disprove
the existence of a gap, as none of the bounds is nec-
essarily tight. Finally, we note that protocol (d) has
following important advantage: one can read the er-
ror syndrome at any time-step before deciding on the
optimal strategy for the future, whereas for protocols
(b) and (c) some decisions have to be made before-
hand. We will come back to this particular point,
and show for a simple phase estimation example how
this difference allows to boost the attainable QFI in
Sec. 6.2.
5 Implications on attainable precision
in metrology schemes
Hitherto our analysis focused on the scaling of the
QFI evaluated on the final system state of the metrol-
ogy protocols depicted in Fig. 1 in presence of general
time-homogeneous Liouvillian noise. We now discuss
the implications of FFQC on the precision with which
one can estimate the parameter ω. As already men-
tioned in Sec. 3.1, how to best quantify the estimation
precision depends crucially on how information about
ω is obtained. In Sec. 5.1 we consider the implications
of FFQC on precision for the case of asymptotically
many repetitions, whereas Sec. 5.2 deals with single-
shot estimation. Regardless of how one chooses to
quantify the precision, whenever a noiseless estima-
tion scenario exhibits a quantum improvement in pre-
cision scaling then this improvement may always be
maintained in the presence of rank-one Pauli noise by
employing the FFQC protocol described in Sec. 4.1.
We note again that the ability of our scheme to ame-
liorate for any rank-one Pauli noise finds applications
beyond metrology, i.e., in the design of high delity
gates, or indeed any other protocol where such noise
terms may appear.
The achievable precision in estimating ω in any
metrological protocol is dictated by the resources at
hand. In what follows we shall consider two differ-
ent ways of quantifying the resources of any metro-
logical protocol: time-particles and number of probes.
In the time-particles approach the total resource per
experimental run is defined as the product between
the number of probes N and the protocol duration
t. Thus, for all the types of protocols depicted Fig. 1
the time-particle resource is equal to t
0
= Nt. In
the number of probes approach one is primarily inter-
ested in how estimation precision scales with N only.
In particular, this is how resources are defined in fre-
quency estimation [15, 19, 22, 38], where the time for
each experimental run, t, is bounded from above so
that one can consistently consider the limit of many
experimental repetitions by letting the overall exper-
iment last sufficiently long. In all cases we show how
our results yield noise dependent upper bounds on
the attainable precision and restrict it to a classical
scaling in terms of the resource.
5.1 Precision in the presence of free repetitions
We now consider precision bounds for estimating pa-
rameter ω using FFQC for the case where the protocol
is repeated asymptotically many times so that CRB
is applicable. We first consider the time-particles pic-
ture before moving onto the scenario in which the
number of probes N is the resource.
5.1.1 Time-particles
Given ν repetitions of the protocol and using the
CRB, we bound the ultimate MSE attainable by the
most general FFQC strategy in the presence of rank-
one Pauli noise as
δω
2
ν
1
(1 (n
T
z)
2
)
1
t
0
2
. (41)
On the other hand, stemming from Eq. (39) the MSE
for all other noise processes may be lower-bounded as
follows
δω
2
ν
1
4α
L
1
t
0
. (42)
Yet, in contrast to Eq. (41), Eq. (42) is not guaranteed
to be achievable in the limit ν .
5.1.2 Frequency estimation
We now turn to the setting of frequency estimation in
which the resource of interest is the number of probes
Accepted in Quantum 2017-09-04, click title to verify 12
N. Notice that the FFQC protocol in this case is the
one depicted in Fig. 1(c). Yet, its QFI can always be
upper-bounded by that of scheme (d) after setting t
0
=
Nt. In order to ensure a large number of repetitions
we fix the time of a single run t and demand that the
total time T := νt t. Hence, we can again use the
CRBs of Eqs. (41) and (42) and obtain the precision
bounds:
δω
2
T
1
(1 (n
T
z)
2
)
1
N
2
t
, (43)
δω
2
T
1
4α
L
1
N
. (44)
Indeed, Eq. (43) proves that HS is achievable for rank-
one Pauli noises. On the other hand, for all other
noise processes, thanks to Eq. (39) being linear in t
0
,
the bound of Eq. (44) is independent of the single-
run duration t. This suggests that FFQC allows us to
indefinitely maintain the ultimate precision normally
exhibited only at very short time scales (see Sec. 4.3).
5.2 Single-shot precision bounds
We now consider precision bounds applicable in the
single-shot scenario, i.e., when ν = 1. As a result
we are forced to consider the AvMSE as the figure of
merit, where the average is with respect to the prob-
ability distribution p
0
(ω) describing our prior knowl-
edge about the parameter. The AvMSE can then be
lower-bounded with the help of BCRB and ZZB (see
Sec. 3.1). However, note that the FFQC-assisted se-
quential strategies of Fig. 1(d) with sufficiently large t
0
also incorporate all many repetition protocols. This
is because one can always set the intermediate con-
trol gates to implement measurements and state re-
preparations. Hence, we expect all the single-shot
bounds to reproduce the ones derived in the regime
of many repetitions when considering t
0
.
5.2.1 Pauli rank-one noise
As by repeating we may only improve the precision,
whenever the precision is forced to follow the stan-
dard scaling in the many-repetitions regime, it must
be standard scaling-bounded in case of the single-shot
estimation. Thus, it is most important to ask whether
the HS can still be attained in presence of rank-one
Pauli noise within strategies (c) and (d) with only a
single run. In particular any protocol that achieves
the maximum attainable precision in the absence of
noise can maintain this precision in the presence of
rank-one Pauli noise by utilizing FFQC. For example,
by incorporating FFQC in the protocol in [47] one
can maintain its optimal performance for any rank-
one Pauli noise process; the number of measurements
N in [47] correspond to the control-steps k in our pro-
tocol (d). In App. F, we provide an explicit strategy
which for a uniform prior—a prior distribution that
is constant over a finite interval, and zero elsewhere—
leads to the AvMSE satisfying
hδω
2
i
FFQC
π
4
κ
4(1 (n
T
z)
2
)
1
t
02
(45)
for (d) and (c) (after substituting t
0
= N t), where
κ = 6.74. This shows that, at least for this prior,
quadratic scaling of precision in the total resource t
0
2
(or (Nt)
2
) is attainable by a single shot strategy. In
the context of frequency estimation of Sec. 5.1, this
demonstrates that if one can decide whether to divide
the total time T in many runs of a fixed duration
t or perform a single shot strategy with FFQC, one
should choose the latter as it achieves precision that
scales quadratically hδω
2
i 1/(N T )
2
in N and T ,
and not only in N , as in Eq. (43).
5.2.2 Bayesian Cramér-Rao bound
The BCRB is defined as [49]
hδω
2
i
1
hF
L
i + F
p
0
(ω)
, (46)
where F
p
0
(ω)
= h(
ω
log p
0
(ω))
2
i is the classical
Fisher information of the prior distribution and h. . . i
denotes averaging with respect to the prior distribu-
tion p
0
(ω). Hence, using Eq. (39), we obtain the fol-
lowing lower bound on the AvMSE :
hδω
2
i
1
4α
L
t
0
+ F
p
0
(ω)
. (47)
Notice that if F
p
0
(ω)
is finite, by substituting t
0
νt
0
above we adequately recover the many-repetition
bound of Eq. (42).
5.2.3 Ziv-Zakai bound
A more general bound that also allows for irregular
priors is the ZZB introduced in [48]. In App. D we
generalize the bound of [48] so that it can be applied
also to mixed states. As a result we can lower bound
the AvMSE for large enough t
0
by
hδω
2
i
1
12α
L
t
0
(48)
for any prior that fulfills some very mild regularity
conditions (see App. D). Again, by substituting t
0
νt
0
we adequately recover the many-repetition bound
of Eq. (42).
6 Importance of quantum control for
metrology with limited resources
In this section we further analyse the benefits of con-
tinuous quantum control for metrology in presence of
Accepted in Quantum 2017-09-04, click title to verify 13
noise that can not be completely removed. In addi-
tion, we adopt a practical perspective where the re-
sources available for the implementation of the scheme
are severely limited. This allows us to keep the analy-
sis simple, but also makes the proposed protocols eas-
ily implementable in real experiments. In Sec. 6.1 to
cope with an unbalanced X-Y noise we propose a sim-
ple FFQC scheme that only requires one sensing and
one ancillary qubit, but can still outperform a paral-
lel scheme with a large number of entangled probes
and ancillae (but no continuous quantum control). In
Sec. 6.2 we show that also in the case of phase esti-
mation the possibility to perform intermediate con-
trol (applied between the two successive applications
of the channel on the probe qubits as shown in Fig. 6)
leads to an improvement in the attainable QFI.
6.1 Frequency estimation with X-Y noise
We study the setting of frequency estimation in which
the probes sensing the parameter are also affected by
the X-Y noise specified in Eq. (10). In accordance to
Eq. (40d, 44) the attainable precision for this noise
is asymptotically constrained to a constant improve-
ment over standard scaling, even when considering
the most powerful FFQC scheme of Fig. 1(d) (with
the exception of the limiting cases p = 0 or p = 1,
where one of the components vanishes and one recov-
ers the correctable perpendicular dephasing noise [27
30, 38]). Also in these cases strategies with FFQC can
outperform those without yielding a large improve-
ment factor as we showed in Sec. 4.3. In this sec-
tion we strengthen this point by demonstrating how
the most general parallel scheme of Fig. 1(b) with N
qubits and N ancillae is outperformed by a simple and
experimentally tangible strategy, which only requires
entanglement between one sensing and one ancillary
qubit but employs fast control.
More precisely, we compare the most general paral-
lel strategy Fig. 3(b) with simple sequential strategies
Fig. 3(s1) and Fig. 3(sN), where one is only allowed to
entangle probes and ancillae two by two rather then
manipulate a global entangled state as required by
(b). The two strategies (s1) and (sN) are actually the
same, but correspond to two different ways of treat-
ing the total resources. In (s1) the experimentalist
only manipulates one qubit and one ancilla and does
the experiment for a total duration T . This protocol
has the same duration as (b) but uses N times less
“calls of the master equation”, i.e., one effectively ap-
plies the single-qubit evolution for a time that is N
times shorter as compared to (b). The scheme (sN)
is equivalent to (s1) upon replacing T with NT , such
that it uses the same amount of time-particles as (b).
Equivalently (sN) corresponds to running N proto-
cols (s1) simultaneously. In this respect it has the
same number of qubits and the same time duration
as (b). However, in contrast to (b), the scheme (sN)
is granted with fast control on the one hand, but only
requires two-qubit entanglement on the other. We
demonstrate that, even with such restrictions, for any
fixed N both strategies (sN) and (s1) outperform any
parallel schemes (b) given that the asymmetry in the
X-Y noise is high enough.
Our results show that the use of control and error
correction techniques allow one to attain resolutions
with a two-qubit (probe-plus-ancillae) setup which
outperform ones reached when considering systems
containing large-scale entanglement. We believe that
our results may support current state-of-art quantum
metrology experiments with nitrogen-vacancy (NV)
centres, in which the error correction protocols have
already been implemented with great success [69, 70].
In particular, as the dominant noise is such systems
has been argued to be nearly transversal [28, 29] and,
hence, highly asymmetric [68], our work proves that
such systems are indeed capable of attaining resolu-
tions unreachable by the entanglement-based schemes
employing comparable resources.
6.1.1 Parallel strategy with N qubits and N ancillae
A general parallel frequency estimation scenario is de-
picted in Fig. 1(b) and Fig. 3(b). For a regular prior
knowledge and sufficiently large T (see Sec. 5.2.2), or
with the explicit assumption of sufficiently many rep-
etitions (T /t 1), the ultimate attainable precision
is determined by the CRB:
T δ
2
ω min
t
t
F
%
(b)
N
(t)
. (49)
The right hand side in Eq. (49) is the inverse of the
maximal QFI rate for a given scheme of type (b) em-
ploying N sensing qubits. It is the optimized for the
single-run duration that we dub t
(b)
opt
. The optimal
time in general depends on the system size (the probe
number N) and the form of the noise (the Liouvil-
lian in Eq. (6))
3
. Following the methods described
in Sec. 4.2.1 (see also App. D), we can upper-bound
the maximal QFI rate by employing the parallel CE
bound (22):
F
%
(b)
N
(t
(b)
opt
)
t
(b)
opt
N f
(b)
N
, where
f
(b)
N
:= 4 max
t
min
h(t)
||α(t)|| + (N 1)||β(t)||
2
t
(50)
is the so-obtained upper bound on the maximal QFI
rate per probe. We adequately label then as t
(b)
opt
the
optimal t maximising Eq. (50). As a result, we may
generally lower-bound the precision dictated by the
3
Note that in the absence of noise it is optimal to set t = T
as large as possible, which makes the CRB (49) not applicable.
Accepted in Quantum 2017-09-04, click title to verify 14
...
...
N
2N
T
...
...
...
...
T
(b)
...
T
...
(s1)
... ... ...
... ...
(sN)
NT
... ... ...
...
...
Figure 3: (s1): A control-assisted sequential protocol that only involves one probe and one ancillary qubit. (b): The most
general parallel protocol of Fig. 1(b) that requires to manipulate entangled states of N probe and N ancillary qubits. (sN): A
control-assisted sequential protocol that employs N independent probe-plus-ancilla entangled pairs running in parallel.
CRB (49) as follows:
(b): T δ
2
ω
1
Nf
(b)
N
, (51)
at the price of the saturability that cannot be guar-
anteed any more, even in the T/t
(b)
opt
1 limit.
In order to perform the minimization in (50), we re-
sort to the SDP-based methods of Ref. [22] reviewed in
Sec. 4.2.2. To this end we solve the master Eq. (10) for
the X-Y noise, and compute the Kraus representation
of the resulting channel, E
X-Y
ω,t
, given in App. D. For
any fixed N and t this allows us to perform the mini-
mization over the Hamiltonian h(t) numerically, while
for N = 1
4
and N we obtain the corresponding
expressions analytically for all times. Finally, to ob-
tain f
(b)
N
we numerically search for the optimal t
(b)
opt
in Eq. (50). The detailed analysis on the solution
of the master equation and derivation of the bounds
may be found in App. D. Note that the balanced X-Y
noise (p = 1/2) is phase-covariant, i.e., it commutes
with the Hamiltonian, so that for this particular case
the bound (50) can be obtained analytically [65].
6.1.2 Simple sequential strategy
For the sequential strategy, all the probe plus ancilla
pairs behave independently, and since the QFI is addi-
tive in this case in order to establish the global perfor-
mance it is sufficient to consider a single pair up to the
first measurement in Fig. 3(s). We denote the state of
a probe-plus-ancilla at any time by %(t) B(C
2
C
2
).
The maximal QFI rate per probe then reads
f
(s)
:=max
t
F(%(t))
t
=
F(%(t
(s)
opt
))
t
(s)
opt
(52)
with t
(s)
opt
being now determined solely by the noise.
The sequential FFQC strategy was already
sketched in Sec. 4.3.2 and is formally described in
4
For which the CE bound (50) is guaranteed to be tight, see
Eq. (19).
Sec. D.1. It consists of continuously checking if the
probe+ancilla state is in the code H
C
or the error
subspace H
E
. At each time step the probability of de-
tecting an error is
γdt
2
, and if an error is detected the
most probable error term, say σ
1
, is canceled by ap-
plying the same unitary on the probe σ
1
σ
1
= . This
maps the state back into the code space but slightly
degrades it with a residual dephasing coming from the
other noise term σ
1
σ
2
= iσ
3
.
As one keeps track of the number of errors m that
happened during such an FFQC assisted evolution,
but disregards the exact times the errors occurred, the
final state at time t possesses a direct-sum structure
¯%(t)=
M
m=0
p(m; t)¯%
m
(t). (53)
The probability of errors follows a Poissonian distri-
bution p(m; t) = e
γt/2
(γt/2)
m
/m!, while the state
¯%
m
(t) conditional to the detection of m errors is given
in (112).
The corresponding QFI may be straightforwardly
evaluated:
F(¯%(t)) =
X
m
p(m; t)F
¯%
m
(t)
= t
2
e
γt 2p(1p)
,
(54)
and yields the following maximal QFI rate per probe
(52):
f
(s)
=
1
2γe
1
p(1 p)
, (55)
which is attained after optimally setting t
(s)
opt
=
1/(2γp(1 p)).
In the regime of long experimental duration
T/t
(s)
opt
1 that we are interested in, the CRBs (49)
for the two strategies Fig. 4(s1) and Fig. 4(sN) respec-
tively read:
(s1): T δ
2
ω
1
f
(s)
(56)
(sN): T δ
2
ω
1
Nf
(s)
, (57)
Accepted in Quantum 2017-09-04, click title to verify 15
Figure 4: Maximal QFI rate per probe f
(s)
attainable by the
sequential strategies (dashed lines) of Fig. 3(s1) and (sN),
as compared to the upper bound, f
(b)
N
(solid curves), valid
for all parallel protocols of Fig. 3(b) in presence of X-Y noise
(ω =1, γ =0.05). For p =0.25 (black) the curves indicate the
sequential strategy (sN) to be superior for N < 5, whereas
for p=0.05 (red) up to N < 316 (while (s1) is superior up to
N < 4dotted red curve). The inset depicts dependence on
the probe number of t
(b)
opt
that maximizes Eq. (50). Despite
not being smooth at finite N , t
(b)
opt
asymptotically follows
1/N
1/3
scaling (dashed line).
and are guaranteed to be attainable.
Before we proceed to the comparison of strategies
(s) and (b), let us briefly comment on the role of
the error register, and how it allows to boost the
attainable performance. As we argued in Sec. 4.3.2
discarding the error register is equivalent to effec-
tively modifying the X-Y noise to a dephasing noise
of strength p γ. Concretely, if the error register is
discarded, the state at time t is a mere mixture
¯%(t) =
P
m=0
p(m; t)¯%
m
(t) rather than a direct sum
as in Eq. (53). The corresponding maximal QFI rate
per probe is then given by f
(s)
=1/(2γep), which is re-
duced by a factor 1/(1p) as compared to the strategy
including the error register and described by Eq. (55).
The intuition behind this gap is rather simple. When
inferring the value of ω after a certain number of runs
it is helpful to know how noisy each run was. This
information, contained in the error register, allows to
properly ponder the data obtained in each run in or-
der to optimally construct the global estimator.
6.1.3 Comparison of the strategies
In Fig. 4, we explicitly present the maximal QFI rate
per probe attained by the sequential strategy f
(s)
of
Eq. (55), and compare it to the bound limiting the
performance of any parallel scheme f
(b)
N
of Eq. (50),
which we plot as a function of the number of probes
N. This corresponds to a comparison of strategies
(sN) and (b) described in Fig. 3. In particular, after
fixing ω = 1, γ = 0.05, we show that when the noise
generated by σ
2
is three times more dominant than
the one generated by σ
1
, i.e., p = 1/4, the sequential
Figure 5: Threshold numbers of probes, N
(s1)
th
and N
(sN)
th
,
below which the two sequential strategies (s1) and (sN) are
superior to all parallel protocols (b) in Fig. 3. as a function of
the X-Y noise asymmetry p for ω = 1 and γ = {0.1, 0.2, 0.3}
(top to bottom). Red lines are plotted assuming sequential
protocol performance to be evaluated per probe (sN), while
blue lines for the sequential protocol employing only a single
probe-plus-ancilla pair (s1). In the former case, N
(sN)
th
2
with equality always occurring strictly at p = 1/2, while in
the single-pair scenario there exists a range of 0.13 /p 0.5
for which trivially N
(s1)
th
=1. In both cases, N
th
diverges with
noise asymmetry, as p0, and eventually follows the p
3/2
,
p
3/5
scalings respectively (dashed lines).
strategy is guaranteed to be superior for N less than
5. However, with the increase of noise asymmetry
the superiority of the sequential strategy drastically
improves, so that for p = 0.05 only if the probe num-
ber N is greater than 316 the parallel schemes (b)
can potentially beat the sequential one (sN). In this
case, in order to also compare (b) to the strategy (s1),
which employs only a single (rather than N) probe-
plus-ancilla qubit pair for the same duration T , we
also plot f
(s)
/N. Strikingly, the parallel strategies (b)
are still outperformed by (s1) as long as N <4.
To compare the strategies more directly we explic-
itly present in Fig. 5 the threshold probe numbers
below which the sequential strategies are guaranteed
to be superior over all parallel schemes. Concretely,
to compare (sN) with (b) we plot N
(sN)
th
such that
f
(s)
> f
(b)
N
for all N < N
(sN)
th
(the top curves), and
to compare (s1) with (b) we plot N
(s1)
th
such that
f
(s)
/N > f
(b)
N
for all N < N
(s1)
th
(the bottom curves).
Note that N
(sN)
th
2 for any value of p, as in the
worst case of balanced noise (p = 1/2) one may ex-
plicitly show that f
(b)
N=2
=f
(s)
, see App. D. Regarding
the other threshold, although N
(s1)
th
= 1 in the range
0.13 / p 0.5, meaning that the sequential strategy
(s1) only outperforms parallel strategies with a single
probe-plus-ancilla pair; N
(s1)
th
still diverges to infinity
when p 0 and the noise approaches the perpendic-
ular dephasing. We observe that the thresholds scale
as N
(sN)
th
p
3/2
and N
(s1)
th
p
3/5
. This proves
that both sequential strategies (s1) and (sN) can out-
Accepted in Quantum 2017-09-04, click title to verify 16
(i)
(ii)
Figure 6: Two different strategies for phase estimation in-
volving two applications of the noisy channel E
θ
presented in
Sec. 6.2. In strategy (i) the two probe and ancillary qubits
are prepared in a suitably entangled state and the channels
are applied to the sensing systems in parallel before they are
finally measured. This strategy intrinsically includes error-
correction but only at the final step (see strategy (iii) in [37]).
(ii) The sequential strategy from Sec. 6.2, after the applica-
tion of the first channel one reads out the error syndrome
and decides on the optimal strategy for the next step (see
strategy (iv) in [37]).
perform the most general parallel one of type (b) for
any N, given sufficient noise asymmetry.
6.2 Phase estimation with bit- and bit-phase-
flip noise
We now return to the canonical phase estimation sce-
nario [5], which may yet be interpreted as the fre-
quency estimation considered before but with the evo-
lution time t being fixed, such that the parameter of
interest is θ = ωt. In this case the sensing interac-
tion of a single probe of duration t constitutes the
elementary block, drawn as a square box in Fig. 6.
Each box corresponds to a CPTP map E
θ
incorpo-
rating the noise [37]. In this context one is free to
use ancillary particles to benefit from error correction
techniques, but also to combine the boxes in differ-
ent ways. The two schemes depicted in Fig. 6(i) and
Fig. 6(ii) are the elementary phase estimation equiva-
lents of the schemes depicted in Fig. 1(b) and Fig. 1(d)
respectively.
We provide a simple example demonstrating that
a sequential strategy outperforms the parallel one.
Specifically, consider the situation where each line in
Fig. 6 corresponds to a qubit and the map E
θ
is a
combination of the unitary evolution given by U
θ
=
e
i
θ
2
σ
3
and of the noise which consists of a random ap-
plication of either a bit flip (σ
1
) or bit-phase flip (σ
2
)
with equal probability
1p
2
. Accordingly, E
θ
has a
Kraus representation of the form {K
0
=
p U
θ
, K
1
=
1 p |1ih0|, K
2
=
1 p |0ih1|}. Note that this
noise channel corresponds to a Pauli X Y noise
channel which, however, differs from the solution of
the master equation for the balanced X Y noise
described in Sec. 6.1.
In the parallel strategy of Fig. 6(i) one starts with
a suitably entangled state of two sensing qubits and
two ancillae. The channel is applied on both qubits
in parallel, and the final state is measured (this in-
cludes error-correction at the final stage see strat-
egy (iii) in [37]). Without loss of generality the in-
put two-qubit plus two-ancilla state is given by |ψi =
P
1
i,j=0
a
ij
|i, ji
Q
|i, ji
A
with a
ij
R. This encoding
is optimal as it allows us to distinguish between all
the different branches of the noisy evolution. More-
over, within every branch the state remains pure, as
the channel corresponds to either leaving the sensing
qubit untouched—with probability p—or projecting
it onto the {|0i, |1i} basis (followed by an irrelevant
flip of the qubit) which can be detected by our error
correction code. As all the branches can be detected,
and within every branch the state remains pure, the
QFI of the final state is given by the mean variance
of H =
1
2
(σ
3
+ σ
3
) over all the branches. The
mean variance can be straightforwardly maximized by
a brute force numerical optimization over the three di-
mensional manifold spanned by the coefficients a
ij
. It
turns out that the optimal states are symmetric un-
der permutations of the two qubits (a
01
= a
10
) and,
moreover, satisfy a
00
= a
11
. Using these conditions at
the outset one can analytically determine the optimal
state to be
|ψi
opt
=
q
x
opt
2
(|00i
Q
|00i
S
+ |11i
Q
|11i
S
)
+
q
1x
opt
2
(|01i
Q
|01i
S
+ |10i
Q
|10i
S
), (58)
with x
opt
= min(1,
2p
4(1p)
). The corresponding maxi-
mal QFI is
F
(i)
=
(
(p2)
2
p
2(p1)
p
2
3
4p
2
otherwise.
(59)
Now consider the following sequential strategy with
intermediate error correction involving two sensing
qubits and a single ancilla, see Figure 6(ii). The first
sensing qubit and the ancilla are prepared in the state
1
2
|00i
Q
1
A
+ |11i
Q
1
A
. After the application of the
first box on the first sensing qubit but before the sec-
ond one we perform our error detection scheme. If
an error is detected the state contains no information
about the parameter, hence one discard Q
1
and pre-
pare the second qubit and the ancilla in the same state
1
2
|00i
Q
2
A
+ |11i
Q
2
A
. If no error is detected after
the first application of the channel the second sensing
qubit is introduced and entangled with the first qubit
such that the overall state is given by
r
1
2
e
i
θ
2
|0i
Q1
(
y |00i
Q2,A
+
p
1 y |11i
Q2,A
)
+e
i
θ
2
|1i
Q1
(
y |11i
Q2,A
+
p
1 y |00i
Q2,A
)
.
(60)
One easily checks that the state that optimizes the
final QFI (after the application of the second box) is
the one with y = y
opt
= min(1,
1
2(1p)
). Finally the
overall QFI of this scheme is given by
F
(ii)
=
(
p(3p + 1) p
1
2
p
(
p
2
2p+2
)
1p
otherwise.
(61)
Accepted in Quantum 2017-09-04, click title to verify 17
0.2
0.4
0.6
0.8
1.0
0.05
0.10
0.15
0.20
p
Figure 7: Comparison of the QFI’s for strategies (i) and (ii).
The curve corresponds to the difference of QFI between the
strategies (ii) in Eq. (61) and (i) Eq. (59).
In Figure 7, we plot the difference between QFI’s of
the strategies (ii) in Eq. (61) and (i) in Eq. (59). One
clearly sees that the strategy with intermediate con-
trol Fig. 6(ii) performs better than the parallel strat-
egy Fig. 6(i). Moreover there is a gap for all values of
p except the trivial cases: the noiseless case p = 1 and
the case where no information about θ is contained in
the channel p = 0.
This simple example shows that intermediate
control (FFQC) allows for an improvement in the
achievable accuracy. Notice that such an advantage
can be maintained for any finite number of blocks
as one can choose to introduce new sensing qubits,
appropriately entangled with the already existing
ones, only after one has obtained information about
the errors from the intermediate error correcting
steps. In the parallel strategy however, all the qubits
have to be initialized in an entangled state, which
is highly fragile to subsequent errors. However the
question whether this improvement vanishes in the
asymptotic case—as conjectured in [37]—or persists,
as one might suspect from the finite N results shown
here, remains open.
7 Summary and Outlook
We have considered the general limits of quantum
metrology where one is, in principle, equipped with
a full-scale quantum computer to assist in the sensing
process. We have shown that one can use techniques
from quantum error correction to detect or correct for
certain kinds of errors while maintaining the sensing
capabilities of the system. In particular, we discover
that the use of FFQC allows to restore the Heisen-
berg scaling of precision for all all rank-one Pauli noise
processes, except the Pauli noise that is identical to
the Hamiltonian, at the cost of a slowing down the
evolution by a constant factor. For all other noise-
types, we have shown that the QFI is limited to a
linear scaling with the number of resources. As we
demonstrated, this result forbids the Heisenberg scal-
ing of the estimation precision and provides noise de-
pendent upper bounds for all metrological scenarios,
local or Bayesian. Remarkably, considering the ulti-
mate metrological scheme (FFQC) actually lead to a
simplification of the problem, as in the limit of short
evolution steps dt 0 the CE method does not re-
quire to solve the master equation, this allowed us to
obtain fully analytic bounds on the achievable preci-
sion for relevant noise processes.
However, even in the cases where FFQC does not
allow for full restoration of the Heisenberg scaling,
one can achieve a significant improvement over par-
allel strategies that operate on limited resources, and
do not utilize FFQC. We have demonstrated this for
the example of asymmetric X-Y noise where the im-
provement over parallel schemes employing the same
number of resources but do not use FFQC can be
significant. Moreover, we have put forward simple
sequential protocols that make use of only a single
sensing and auxiliary system, but that nevertheless
outperform parallel entanglement-based strategies for
limited number of sensing systems. From a practical
perspective, the existence of simple practical schemes
that operate with only a single sensing and auxiliary
system, are of high relevance, and may open the way
for practical noisy metrology with a significant quan-
tum advantage. This demonstrates that fast quan-
tum control and error correction are powerful tools in
quantum metrology and may enable practical sensing
with a significant quantum enhancement even in the
presence of noise.
An important question one may ask is how our re-
sults generalize to higher dimensional systems. Note
that in higher dimensions the parameter encoding
Hamiltonian may have support only on a subspace
of the Hilbert space of the probe system, which may
allow for more involved error correcting codes.
Note added. After making this work available on-
line, the error correction protocol presented in Sec. 4.1
capable of removing any rank-one Pauli noise has been
successfully implemented in an NV-centre experiment
to compensate for transversal dephasing noise [71].
On the other hand, a dissipation-based scheme has
been proposed to correct for such noise in ultra-cold
ion sensing experiments [72]. Lastly, the general re-
sults presented in Sec. 4.1 and Sec. 4.2.4 have been
very recently generalized [73, 74] beyond the case of
qubit probes, considered in this paper.
Acknowledgements
This work has been supported by Austrian Science
Fund (FWF: P24273-N16, P28000-N27), Swiss Na-
tional Science Foundation Grant (P2GEP2_151964),
Spanish Ministry National plans FOQUS and
Accepted in Quantum 2017-09-04, click title to verify 18
MINECO (Severo Ochoa Grant No. SEV-2015-
0522), Spanish MINECO FIS2013-40627-P, Gener-
alitat de Catalunya CIRIT 2014 SGR 966 Fun-
dació Privada Cellex, Generalitat de Catalunya Grant
(No. SGR875), as well as received funding from the
European Union’s Horizon 2020 research and innova-
tion programme under the Marie Skłodowska-Curie
Grant: Q-METAPP (No. 655161).
References
[1] Jonathan P. Dowling and Kaushik P. Se-
shadreesan. Quantum Optical Technologies for
Metrology, Sensing, and Imaging. J. Lightwave
Technol., 33(12):2359–2370, June 2015. ISSN
0733-8724. DOI: 10.1109/JLT.2014.2386795.
[2] Vittorio Giovannetti, Seth Lloyd, and Lorenzo
Maccone. Advances in quantum metrol-
ogy. Nature Photon., 5:222–229, 2011. DOI:
10.1038/nphoton.2011.35.
[3] Vittorio Giovannetti, Seth Lloyd, and Lorenzo
Maccone. Quantum-enhanced measurements:
Beating the standard quantum limit. Science,
306(5700):1330–1336, 2004. DOI: 10.1126/sci-
ence.1104149.
[4] V. Buzek, R. Derka, and S. Massar. Optimal
quantum clocks. Phys. Rev. Lett., 82:2207–2210,
Mar 1999. DOI: 10.1103/PhysRevLett.82.2207.
[5] Vittorio Giovannetti, Seth Lloyd, and Lorenzo
Maccone. Quantum metrology. Phys. Rev.
Lett., 96:010401, 2006. DOI: 10.1103/Phys-
RevLett.96.010401.
[6] B. C. Sanders and G. J. Milburn. Optimal
quantum measurements for phase estimation.
Phys. Rev. Lett., 75:2944–2947, Oct 1995. DOI:
10.1103/PhysRevLett.75.2944.
[7] D. W. Berry and H. M. Wiseman. Optimal states
and almost optimal adaptive measurements for
quantum interferometry. Phys. Rev. Lett., 85:
5098–5101, Dec 2000. DOI: 10.1103/Phys-
RevLett.85.5098.
[8] Asher Peres and Petra F. Scudo. Entangled
quantum states as direction indicators. Phys.
Rev. Lett., 86:4160–4162, Apr 2001. DOI:
10.1103/PhysRevLett.86.4160.
[9] E. Bagan, M. Baig, and R. Muñoz Tapia.
Quantum reverse engineering and reference-
frame alignment without nonlocal correlations.
Phys. Rev. A, 70:030301, Sep 2004. DOI:
10.1103/PhysRevA.70.030301.
[10] G. Chiribella, G. M. D’Ariano, P. Perinotti,
and M. F. Sacchi. Efficient use of quantum re-
sources for the transmission of a reference frame.
Phys. Rev. Lett., 93:180503, Oct 2004. DOI:
10.1103/PhysRevLett.93.180503.
[11] Giulio Chiribella, Giacomo Mauro D’Ariano,
Paolo Perinotti, and Massimiliano F. Sacchi. Co-
variant quantum measurements that maximize
the likelihood. Phys. Rev. A, 70:062105, Dec
2004. DOI: 10.1103/PhysRevA.70.062105.
[12] G. Chiribella, G. M. D’Ariano, and M. F. Sacchi.
Optimal estimation of group transformations us-
ing entanglement. Phys. Rev. A, 72:042338, Oct
2005. DOI: 10.1103/PhysRevA.72.042338.
[13] B. L. Higgins, D. W. Berry, S. D. Bartlett,
H. M. Wiseman, and G. J. Pryde. Entanglement-
free heisenberg-limited phase estimation. Nature,
450:393, 2007. DOI: 10.1038/nature06257.
[14] B. Yurke. Input states for enhancement
of fermion interferometer sensitivity. Phys.
Rev. Lett., 56:1515–1517, Apr 1986. DOI:
10.1103/PhysRevLett.56.1515.
[15] S. F. Huelga, C. Macchiavello, T. Pellizzari, A. K.
Ekert, M. B. Plenio, and J. I. Cirac. Improve-
ment of frequency standards with quantum en-
tanglement. Phys. Rev. Lett., 79:3865–3868, Nov
1997. DOI: 10.1103/PhysRevLett.79.3865.
[16] Konrad Banaszek, Rafał Demkowicz-Dobrzański,
and Ian A. Walmsley. Quantum states made to
measure. Nature Photon., 3:673–676, 2009. DOI:
10.1038/nphoton.2009.223.
[17] Lorenzo Maccone and Vittorio Giovannetti.
Quantum metrology: Beauty and the noisy
beast. Nature Phys., 7:376–377, 2011. DOI:
doi:10.1038/nphys1976.
[18] Akio Fujiwara and Hiroshi Imai. A fibre bundle
over manifolds of quantum channels and its appli-
cation to quantum statistics. Journal of Physics
A: Mathematical and Theoretical, 41(25):255304,
2008. DOI: 10.1088/1751-8113/41/25/255304.
[19] B. M. Escher, R. L. de Matos Filho, and
L. Davidovich. General framework for estimating
the ultimate precision limit in noisy quantum-
enhanced metrology. Nat. Phys., 7:406, March
2011. DOI: 10.1038/nphys1958.
[20] B. M. Escher, L. Davidovich, N. Zagury, and
R. L. de Matos Filho. Quantum metrological
limits via a variational approach. Phys. Rev.
Lett., 109:190404, Nov 2012. DOI: 10.1103/Phys-
RevLett.109.190404.
[21] R. Demkowicz-Dobrzański, J. Kołodyński, and
M. Guţă. The elusive heisenberg limit in
quantum-enhanced metrology. Nat. Commun.,
3:1063, 2012. DOI: 10.1038/ncomms2067.
[22] J. Kołodyński and R. Demkowicz-Dobrzański.
Efficient tools for quantum metrology with un-
correlated noise. New J. Phys., 15(7):073043,
2013. DOI: 10.1088/1367-2630/15/7/073043.
[23] S. Alipour, M. Mehboudi, and A. T. Reza-
khani. Quantum metrology in open systems:
Dissipative cramér-rao bound. Phys. Rev. Lett.,
112:120405, Mar 2014. DOI: 10.1103/Phys-
RevLett.112.120405.
[24] Sergey Knysh, Vadim N. Smelyanskiy, and
Gabriel A. Durkin. Scaling laws for precision
in quantum interferometry and the bifurcation
Accepted in Quantum 2017-09-04, click title to verify 19
landscape of the optimal state. Phys. Rev.
A, 83:021804, Feb 2011. DOI: 10.1103/Phys-
RevA.83.021804.
[25] Sergey I Knysh, Edward H Chen, and Gabriel A
Durkin. True limits to precision via unique quan-
tum probe. preprint, arXiv:1402.0495[quant–
ph], 2014. URL https://arxiv.org/abs/1402.
0495.
[26] John Preskill. Quantum clock synchronization
and quantum error correction. preprint, arXiv:
0010098[quant–ph], 2000. URL http://arxiv.
org/abs/quant-ph/0010098.
[27] W. Dür, M. Skotiniotis, F. Fröwis, and B. Kraus.
Improved quantum metrology using quantum er-
ror correction. Phys. Rev. Lett., 112:080801, Feb
2014. DOI: 10.1103/PhysRevLett.112.080801.
[28] E. M. Kessler, I. Lovchinsky, A. O. Sushkov,
and M. D. Lukin. Quantum error correction for
metrology. Phys. Rev. Lett., 112:150802, Apr
2014. DOI: 10.1103/PhysRevLett.112.150802.
[29] G. Arrad, Y. Vinkler, D. Aharonov, and A. Ret-
zker. Increasing sensing resolution with error cor-
rection. Phys. Rev. Lett., 112:150801, Apr 2014.
DOI: 10.1103/PhysRevLett.112.150801.
[30] Roee Ozeri. Heisenberg limited metrology us-
ing quantum error-correction codes. preprint,
arxiv:1310.3432[quant–ph], 2013. URL https:
//arxiv.org/abs/1310.3432.
[31] Xiao-Ming Lu, Sixia Yu, and CH Oh. Robust
quantum metrological schemes based on protec-
tion of quantum fisher information. Nat. Com-
mun., 6:7282, 2015. DOI: 10.1038/ncomms8282.
[32] David A. Herrera-Martí, Tuvia Gefen, Dorit
Aharonov, Nadav Katz, and Alex Retzker.
Quantum error-correction-enhanced magnetome-
ter overcoming the limit imposed by relaxation.
Phys. Rev. Lett., 115:200501, Nov 2015. DOI:
10.1103/PhysRevLett.115.200501.
[33] Tuvia Gefen, David A. Herrera-Martí, and Alex
Retzker. Parameter estimation with efficient
photodetectors. Phys. Rev. A, 93:032133, Mar
2016. DOI: 10.1103/PhysRevA.93.032133.
[34] Martin B. Plenio and Susana F. Huelga. Sens-
ing in the presence of an observed environ-
ment. Phys. Rev. A, 93:032123, Mar 2016. DOI:
10.1103/PhysRevA.93.032123.
[35] P Sekatski, M Skotiniotis, and W Dür. Dynami-
cal decoupling leads to improved scaling in noisy
quantum metrology. New J. Phys., 18(7):073034,
2016. DOI: 10.1088/1367-2630/18/7/073034.
[36] Duger Ulam-Orgikh and Masahiro Kitagawa.
Spin squeezing and decoherence limit in ramsey
spectroscopy. Phys. Rev. A, 64:052106, Oct 2001.
DOI: 10.1103/PhysRevA.64.052106.
[37] Rafal Demkowicz-Dobrzański and Lorenzo Mac-
cone. Using entanglement against noise in quan-
tum metrology. Phys. Rev. Lett., 113:250801, Dec
2014. DOI: 10.1103/PhysRevLett.113.250801.
[38] R. Chaves, J. B. Brask, M. Markiewicz,
J. Kołodyński, and A. Acín. Noisy metrology
beyond the standard quantum limit. Phys. Rev.
Lett., 111:120401, Sep 2013. DOI: 10.1103/Phys-
RevLett.111.120401.
[39] Lorenza Viola and Seth Lloyd. Dynamical sup-
pression of decoherence in two-state quantum
systems. Phys. Rev. A, 58:2733–2744, Oct 1998.
DOI: 10.1103/PhysRevA.58.2733.
[40] Lorenza Viola, Emanuel Knill, and Seth Lloyd.
Dynamical decoupling of open quantum systems.
Phys. Rev. Lett., 82:2417–2421, Mar 1999. DOI:
10.1103/PhysRevLett.82.2417.
[41] Lorenza Viola and Emanuel Knill. Robust dy-
namical decoupling of quantum systems with
bounded controls. Phys. Rev. Lett., 90:037901,
Jan 2003. DOI: 10.1103/PhysRevLett.90.037901.
[42] Kaveh Khodjasteh and Lorenza Viola. Dynami-
cally error-corrected gates for universal quantum
computation. Phys. Rev. Lett., 102:080501, Feb
2009. DOI: 10.1103/PhysRevLett.102.080501.
[43] Kaveh Khodjasteh, Daniel A. Lidar, and Lorenza
Viola. Arbitrarily accurate dynamical control
in open quantum systems. Phys. Rev. Lett.,
104:090501, Mar 2010. DOI: 10.1103/Phys-
RevLett.104.090501.
[44] Jacob R. West, Daniel A. Lidar, Bryan H. Fong,
and Mark F. Gyure. High fidelity quantum
gates via dynamical decoupling. Phys. Rev.
Lett., 105:230503, Dec 2010. DOI: 10.1103/Phys-
RevLett.105.230503.
[45] Howard M Wiseman and Gerard J Milburn.
Quantum Measurement and Control. Cambridge
University Press, 2009. ISBN 0521804426. DOI:
10.1017/CBO9780511813948.
[46] Giulio Chiribella. Optimal networks for quantum
metrology: semidefinite programs and product
rules. New Journal of Physics, 14(12):125008,
2012. DOI: 10.1088/1367-2630/14/12/125008.
[47] Alexandr Sergeevich, Anushya Chandran, Joshua
Combes, Stephen D. Bartlett, and Howard M.
Wiseman. Characterization of a qubit hamilto-
nian using adaptive measurements in a fixed ba-
sis. Phys. Rev. A, 84:052315, Nov 2011. DOI:
10.1103/PhysRevA.84.052315.
[48] Mankei Tsang. Ziv-zakai error bounds for quan-
tum parameter estimation. Phys. Rev. Lett.,
108:230401, Jun 2012. DOI: 10.1103/Phys-
RevLett.108.230401.
[49] Richard D Gill and Boris Y Levit. Applications
of the van Trees inequality: a Bayesian Cramér-
Rao bound. Bernoulli, 1(1/2):59–79, 1995. DOI:
10.2307/3318681.
[50] C. W. Helstrom. Quantum Detection and Es-
timation Theory. Academic Press, 1976. ISBN
0123400503.
[51] A. S. Holevo. Probabilistic and Statistical As-
pects of Quantum Theory. North-Holland Se-
Accepted in Quantum 2017-09-04, click title to verify 20
ries in Statistics and Probability, 1980. DOI:
10.1007/978-88-7642-378-9.
[52] Samuel L. Braunstein and Carlton M. Caves.
Statistical distance and the geometry of quan-
tum states. Phys. Rev. Lett., 72:3439–3443, May
1994. DOI: 10.1103/PhysRevLett.72.3439.
[53] Luca Pezzé and Augusto Smerzi. Entanglement,
nonlinear dynamics, and the Heisenberg limit.
Phys. Rev. Lett., 102:100401, Mar 2009. DOI:
10.1103/PhysRevLett.102.100401.
[54] Bernd Lücke, Jan Peise, Giuseppe Vitagliano,
Jan Arlt, Luis Santos, Géza Tóth, and Carsten
Klempt. Detecting multiparticle entanglement of
dicke states. Phys. Rev. Lett., 112:155304, Apr
2014. DOI: 10.1103/PhysRevLett.112.155304.
[55] Helmut Strobel, Wolfgang Muessel, Daniel Lin-
nemann, Tilman Zibold, David B. Hume,
Luca Pezzè, Augusto Smerzi, and Markus K.
Oberthaler. Fisher information and entangle-
ment of non-gaussian spin states. Science,
345(6195):424–427, 2014. DOI: 10.1126/sci-
ence.1250147.
[56] Diego Paiva Pires, Marco Cianciaruso, Lucas C.
Céleri, Gerardo Adesso, and Diogo O. Soares-
Pinto. Generalized geometric quantum speed
limits. Phys. Rev. X, 6:021031, Jun 2016. DOI:
10.1103/PhysRevX.6.021031.
[57] M. M. Taddei, B. M. Escher, L. Davidovich,
and R. L. de Matos Filho. Quantum speed
limit for physical processes. Phys. Rev. Lett.,
110:050402, Jan 2013. DOI: 10.1103/Phys-
RevLett.110.050402.
[58] Florian Fröwis and Wolfgang Dür. Measures of
macroscopicity for quantum spin systems. New
J. Phys., 14(9):093039, 2012. DOI: 10.1088/1367-
2630/14/9/093039.
[59] M. A. Nielsen and I. L. Chuang. Quan-
tum computation and quantum information.
Cambridge university press, 2010. DOI:
10.1017/CBO9780511976667.
[60] Robert Alicki and Karl Lendi. Quantum Dynami-
cal Semigroups and Applications. Springer, 1987.
DOI: 10.1007/3-540-18276-4.
[61] Heinz-Peter Breuer and Francesco Petruc-
cione. The Theory of Open Quantum Sys-
tems. Oxford University Press, 2002. DOI:
10.1093/acprof:oso/9780199213900.001.0001.
[62] Yuichiro Matsuzaki, Simon C. Benjamin, and
Joseph Fitzsimons. Magnetic field sensing be-
yond the standard quantum limit under the ef-
fect of decoherence. Phys. Rev. A, 84:012103, Jul
2011. DOI: 10.1103/PhysRevA.84.012103.
[63] Alex W. Chin, Susana F. Huelga, and Martin B.
Plenio. Quantum metrology in non-markovian
environments. Phys. Rev. Lett., 109:233601, Dec
2012. DOI: 10.1103/PhysRevLett.109.233601.
[64] Katarzyna Macieszczak. Zeno limit in fre-
quency estimation with non-markovian environ-
ments. Phys. Rev. A, 92:010102, Jul 2015. DOI:
10.1103/PhysRevA.92.010102.
[65] Andrea Smirne, Jan Kołodyński, Susana F.
Huelga, and Rafał Demkowicz-Dobrzański. Ul-
timate precision limits for noisy frequency esti-
mation. Phys. Rev. Lett., 116:120801, Mar 2016.
DOI: 10.1103/PhysRevLett.116.120801.
[66] Carole Addis, Elsi-Mari Laine, Clemens Gneit-
ing, and Sabrina Maniscalco. Problem of coher-
ent control in non-Markovian open quantum sys-
tems. Phys. Rev. A, 94:052117, Nov 2016. DOI:
10.1103/PhysRevA.94.052117.
[67] E. Andersson, J. D. Cresser, and M. J. W.
Hall. Finding the Kraus decomposition from
a master equation and vice versa. J.
Mod. Opt., 54(12):1695–1716, 2007. DOI:
10.1080/09500340701352581.
[68] J. B. Brask, R. Chaves, and J. Kołodyński. Im-
proved quantum magnetometry beyond the stan-
dard quantum limit. Phys. Rev. X, 5:031010, Jul
2015. DOI: 10.1103/PhysRevX.5.031010.
[69] T. H. Taminiau, J. Cramer, T. van der Sar, V. V.
Dobrovitski, and R. Hanson. Universal control
and error correction in multi-qubit spin registers
in diamond. Nat. Nanotechnol., 9(3):171–176,
March 2014. DOI: 10.1038/nnano.2014.2.
[70] G. Waldherr, Y. Wang, S. Zaiser, M. Jamali,
T. Schulte-Herbruggen, H. Abe, T. Ohshima,
J. Isoya, J. F. Du, P. Neumann, and
J. Wrachtrup. Quantum error correction in a
solid-state hybrid spin register. Nature, 506
(7487):204–207, February 2014. ISSN 0028-0836.
DOI: 10.1038/nature12919.
[71] Thomas Unden, Priya Balasubramanian, Daniel
Louzon, Yuval Vinkler, Martin B. Plenio,
Matthew Markham, Daniel Twitchen, Alastair
Stacey, Igor Lovchinsky, Alexander O. Sushkov,
Mikhail D. Lukin, Alex Retzker, Boris Nayde-
nov, Liam P. McGuinness, and Fedor Jelezko.
Quantum metrology enhanced by repetitive
quantum error correction. Phys. Rev. Lett.,
116:230502, Jun 2016. DOI: 10.1103/Phys-
RevLett.116.230502.
[72] F. Reiter, A. S. Sørensen, P. Zoller, and C. A.
Muschik. Autonomous Quantum Error Correc-
tion and Application to Quantum Sensing with
Trapped Ions. preprint, arXiv:1702.08673[quant–
ph], 2017. URL http://arxiv.org/abs/1702.
08673.
[73] R. Demkowicz-Dobrzański, J. Czajkowski, and
P. Sekatski. Adaptive quantum metrology un-
der general Markovian noise. preprint, arXiv:
1704.06280[quant–ph], 2017. URL http://
arxiv.org/abs/1704.06280.
[74] Sisi Zhou, Mengzhen Zhang, John Preskill, and
Liang Jiang. Achieving the Heisenberg limit
in quantum metrology using quantum error
correction. preprint, arXiv:1706.02445[quant–
Accepted in Quantum 2017-09-04, click title to verify 21
ph], 2017. URL http://arxiv.org/abs/1706.
02445.
[75] Ingemar Bengtsson and Karol Życzkowski. Ge-
ometry of Quantum States: An Introduction to
Quantum Entanglement. Cambridge University
Press, 2006. DOI: 10.1017/CBO9780511535048.
Accepted in Quantum 2017-09-04, click title to verify 22
A Relation between the Kraus representations and the dynamical matrix
Any Kraus operator, K
i
=
P
k
M
ik
σ
k
, can be expressed in the operator basis given by the four Pauli operators
in the qubit case, which we write for the vector of Kraus operators as
K = Mς . (62)
The action of the channel on the density matrix is expressed in the Kraus representation as
E(ρ) =
X
i
K
i
ρK
i
=
X
i
(
X
k
M
ik
σ
k
)ρ(
X
`
M
i`
σ
`
) (63)
=
X
k`
(
X
i
M
`i
M
ik
)σ
k
ρσ
`
=
X
k`
(M
M)
`k
σ
k
ρσ
`
.
Imposing the equality with the dynamical matrix representation (11) one arrives at
M
M = S
T
. (64)
From this expression it follows that all Kraus representation satisfying M
0
= uM (with a unitary u) lead to the
same dynamical matrix and therefore correspond to the same channel. Remark that in general M may not be
a square but a rectangular n × 4 matrix, where n is the number of Kraus operators in a given representation.
However, any M admits a singular value decomposition of the form
M = u D
n×4
v
, (65)
with u a unitary n × n matrix , v a unitary 4 × 4 matrix and D
n×4
a rectangular n × 4 diagonal matrix.
Now M
0
= D
n×4
v
is also a valid Kraus representation, and so is the 4 × 4 matrix M
00
= D
4×4
v
(where to get
to D
4×4
from D
n×4
we either remove all the n 4 zero lines in the case n > 4, or add 4 n zero lines in the
case n < 4). Finally, we chose the Hermitian 4 × 4 matrix
M
000
= v D
4×4
v
(66)
to be our canonical Kraus representation for the channel.
B Infinitesimal-timestep CE bound
Here, we prove the infinitesimal-timestep CE bound of Eq. (31) that may be applied also to schemes that
incorporate FFQC, i.e., the ones of type (d) in Fig. 1. We essentially follow the derivation contained in the
Supplementary Material of Ref. [37] except the last step, at which we slightly generalize the bound presented
therein.
In particular, one should follow the reasoning up to Eq. (S15) of Ref. [37], at which a more general upper
bound (for any real
x) may be derived via
||
X
k
˙
K
k
iAK
k
+ h.c.|| = ||i
X
k
(
x
˙
K
k
)A(
1
x
K
k
) (
1
x
K
k
)A(
x
˙
K
k
)|| (67)
= ||i
X
k
(
x
˙
K
k
+ i
1
x
K
k
A(
x
˙
K
k
+ i
1
x
K
k
) λ
˙
K
k
A
˙
K
k
1
x
K
k
AK
k
|| (68)
Using the triangle inequality and the one stated in Eq. (S13) of Ref. [37], we arrive at a generalized version of
Eq. (S15) therein:
||
X
k
˙
K
k
iAK
k
+ h.c.|| 2||A||
x||
X
k
˙
K
k
˙
K
k
|| + ||
X
k
˙
K
k
K
k
|| +
1
x
, (69)
which in turn results in the generalization of Eq. (S19):
F 4 k||α|| + 4 k(k 1)||β||
x||α|| + ||β|| +
1
x
. (70)
By taking the limit of infinitesimal timesteps k = t
0
/dt and writing explicitly the time-dependences, we obtain
the bound:
F 4
t
0
dt
||α(dt)|| + 4
(t
0
)
2
dt
2
||β(dt)||
x||α(dt)|| + ||β(dt)|| +
1
x
, (71)
Accepted in Quantum 2017-09-04, click title to verify 23
whose tightest form we would like determine in the dt 0 limit. Hence, as otherwise Eq. (71) must diverge, we
can restrict to Kraus representations for which α(dt) = α
(2)
dt+O(dt
3/2
) and β(dt) = β
(2)
dt+β
(3)
dt
3/2
+O(dt
2
).
As we are willing to show a linear scaling of the QFI in t
0
due to vanishing of the second term in Eq. (71), we
set x = 1/
dt and compute the Taylor expansion:
F 4 t
0
||α
(2)
|| + O(dt)
+ 4 t
02
β
(2)
dt
+
β
(3)
dt
+ O(1)
h
||α
(2)
|| + 1
dt + O(dt)
i
. (72)
Thus, if we are able to find a Kraus representation for which both β
(2)
= 0 and β
(3)
= 0, the above expression
does not diverge in the dt0 limit but rather provides the desired upper bound—Eq. (37) in the main text:
F 4 t
0
||α
(2)
|| with β = O(dt
2
), (73)
which importantly imposes the SQL-like scaling due to its linearity in t
0
.
C Asymptotic FFQC-valid CE bounds for qubit Liouvillians
C.1 Existence of asymptotic bound for any Liouvillian but the non-parallel rank-one Pauli noise
In what follows, we perform analytic minimization of the infinitesimal-timestep CE bound introduced in
Sec. 4.2.3 that lead to the Eq. (37). For ease of notation we denote the matrix S
(1)
in Eq. (35) by
S
(1)
=
s
(1)
s
(1)
s
(1)
¯
S
(1)
. (74)
Recall that our goal is to minimize over all Kraus decompositions K = Mς of the channel E
ω,dt
up to first
order in dt. Expanding the matrix M up to first order in dt, M = M
(0)
+
dt M
(1)
+ dt M
(2)
, and using Eq. (28)
the matrices M
(`)
satisfy
M
(0)
M
(0)
= S
(0)
M
(0)
M
(1)
+ M
(1)
M
(0)
= 0
M
(1)
M
(1)
+ M
(0)
M
(2)
+ M
(2)
M
(0)
= S
(1)
. (75)
These constraints enforce the following structure on the matrices M
(`)
M
(0)
= diag(1, 0, 0, 0),
M
(1)
=
0 0
T
0
¯
M
(1)
=
¯
L
, M
(2)
=
1
2
s
(1)
s
(1)
s
(1)
¯
M
(2)
, (76)
where
¯
M
(2)
does not contribute to the first order expansion of the channel and can thus be chosen at will.
By performing a power series expansion of the matrices α and β in terms of dt
α = α
(0)
+
dt α
(1)
+ dt α
(2)
+ O(dt
3/2
)
β = β
(0)
+
dt β
(1)
+ d
(2)
+ dt
3/2
β
(3)
+ O(dt
2
), (77)
and using Eq. (29) and Eq. (30) with h = h
(0)
+ h
(1)
dt + h
(2)
dt the minimization over all equivalent Kraus
operators is equivalent to searching over Hermitian matrices h
(`)
that minimize the bound of Eq. (37).
To yield a non-trivial bound a Kraus decomposition has to satisfy α
(0)
= α
(1)
= β
(0)
= β
(1)
= 0. Expanding
Eq. (29) and Eq. (30) in powers of dt and minimizing order by order imposes the following structure for h
(0)
, h
(1)
h
(0)
=
0 0
0 H
(0)
, h
(1)
=
0 h
(1)
h
(1)
H
(1)
, (78)
and the coefficients α
(2)
and β
(2)
are given by
α
(2)
= ς
h
(1)
h
(1)
h
(1)
H
(0)
¯
M
(1)
¯
M
(1)
H
(0)
h
(1)
¯
M
(1)
H
(0)
H
(0)
¯
M
(1)
ς (79)
β
(2)
= i
1
2
σ
3
+ ς
h
(2)
00
h
(1)
¯
M
(1)
¯
M
(1)
h
(1)
¯
M
(1)
H
(0)
¯
M
(1)
ς
. (80)
Accepted in Quantum 2017-09-04, click title to verify 24
Moreover, the coefficient β
(3)
in Eq. (77) reads
β
(3)
= iς

1
2
h
(3)
00
+ s
(1)
h
(1)
h
(1)
¯
M
(2)
+ (s
(1)
H
(0)
+ h
(2)
)
¯
M
(1)
0
1
2
¯
M
(1)
H
(1)
¯
M
(1)
+
¯
M
(2)
H
(0)
¯
M
(1)
+ h.c.
ς + iς
0 (0 0
1
2
)
¯
M
(1)
0 0
ς ,
(81)
where h.c. stands for the Hermitian conjugate. As none of the parameters h
(3)
00
, H
(1)
,
¯
M
(2)
, and h
(2)
appearing
in Eq. (81) enter into the expressions for α
(2)
and β
(2)
the norm of β
(3)
can be set to zero independently of all
lower orders of coefficients for α and β. Indeed, the following choice of the free parameters
h
(3)
00
= (s
(1)
h
(1)
+ h
(1)
· s
(1)
)
H
(1)
=2 H
(0)
¯
M
(2)
=
¯
M
(1)
Re(h
(2)
) =Re
h
(1)
H
(0)
s
(1)
1
4
0
0
1
(82)
sets β
(3)
to zero.
C.1.1 Rank-one noise
We start with a general rank-one Liouvillian
¯
L
1G
r
defined in Eq. (8)
¯
L
1G
r
= r r
=
γ
2
v v
(83)
with r = (x, y, z)
T
,
γ
2
= |x|
2
+ |y|
2
+ |z|
2
and the vector v has unit length such that v v
is a rank-1 projector.
Equation (76) implies
¯
M
(1)
=
p
γ
2
v v
. Expressing h
(1)
= (h
r
+ ih
i
)v + h
v
in a basis containing v allows to
re-write Eq. (80) as
β
(2)
= i(
σ
3
2
+ ς
h
(2)
00
(h
r
ih
i
)r
h.c. H
(0)
11
r r
ς ), (84)
where H
(0)
11
= v
H
(0)
v. We decompose the vector r = r
r
+ i r
i
, with r
r
= Re(r) and r
i
= Im(r). A little bit of
algebra allows one to rewrite Eq. (84) as
β
(2)
= iσ
0
(h
(2)
00
+
γ
2
H
(0)
11
)
+ i
σ
3
2
+ 2 σ
(h
r
r
r
h
i
r
i
+ H
(0)
11
r
r
× r
i
)
, (85)
where we used the property [r
i
σ, r
r
σ] = 2i(r
i
× r
r
)
σ. We wish to set the operator β
(2)
to zero. The σ
0
term can be trivially put to zero by choosing h
(00)
2
=
γ
2
H
(0)
11
. The remaining three terms can be put to zero
if the following conditions are fulfilled. Either r
r
and r
i
are linearly independent—in which case the vectors
r
r
, r
i
, r
r
× r
i
span the whole vector space; or r
r
and r
i
are parallel to each other and to the vector z—which
corresponds to the dephasing noise.
Hence, β
(2)
can be set to zero, so that Eq. (37) holds for any rank-one noise except for Pauli noise which is
not parallel to σ
3
. For this later case we have outlined a FFQC strategy in Sec. 4 that effectively removes such
noise at the cost of slowing down the evolution by a constant factor.
C.1.2 Liouvillians of higher rank and CE bound
Let us now consider a general Lindbladian, it can always be written in the diagonal form
¯
L = r
1
r
1
M
r
2
r
2
M
r
3
r
3
(86)
with r
i
r
j
= 0 for i 6= j (as suggested by the direct sum). Again by v
j
=
r
j
|r
j
|
we denote the normalized vecors.
Now we show that if at least one Liouvillian
¯
L
j
= r
j
r
j
(e.g., j = 1) implies SQL scaling, then also
¯
L leads to
SQL scaling. To see this simply pick h
(1)
= (h
r
+ ih
i
)v
1
and H
(0)
= H
(0)
11
v
1
v
1
, Eq. (80) implies
β
(2)
= i
σ
3
2
+ ς
h
(2)
00
(h
r
ih
i
)r
1
h.c. H
(0)
11
r
1
r
1
ς
. (87)
Accepted in Quantum 2017-09-04, click title to verify 25
This expression is exactly equal to Eq. (84), consequently it can be set to zero if the Liouvillian
¯
L
1
leads to
standard scaling. Accordingly any higher rank Liouvillian that contains a non-Pauli rank-1 noise in its diagonal
decomposition leads to standard scaling.
C.1.3 Rank-two noise
Now consider a general Lindbadian of rank-two. Because of what we just showed in the previous section, the
only case which can potentially lead to Heisenberg scaling is the rank-two Pauli noise in Eq. (40c)
¯
L
2P
=
1
2
R
T
γ
1
γ
2
0
R
= r
1
r
1
M
r
2
r
2
,
(88)
with orthogonal real vecrtors r
1
=
p
γ
1
2
v
1
and r
2
=
p
γ
2
2
v
2
. The choice h
(2)
00
= 0, h
(1)
= h
1
v
1
+ h
2
v
2
(with
h
1
, h
2
R), and
H
(0)
=
0 ic 0
ic 0 0
0 0 0
(89)
expressed in the basis {v
1
, v
2
, v
3
= v
1
× v
2
} yields in Eq. (80)
β
(2)
= i
σ
3
2
+ 2 σ
(h
1
r
1
+ h
2
r
2
+ c r
1
× r
2
)
. (90)
As the three vectors {r
1
, r
2
, r
1
× r
2
} form a basis, the expression above can always be set to zero. Hence,
Eq. (37) holds for any rank two noise process as well.
C.1.4 Rank-three noise
Finally let us consider the case of rank-three noise. In this case the matrix
¯
M
(1)
is invertible such that β
(2)
in
Eq. (80) can be trivially set to zero by choosing h
(1)
=
¯
M
(1)1
(0, 0, 1/4)
T
, h
(2)
00
= 0 and
¯
H
0
= 0.
Thus, we have shown that even in the most general situation where the experimentalist has full quantum
control any Liouvillian with the exception of the rank-one Pauli noise (not parallel to the generator of the
evolution) yields a QFI that is upper bounded by, F 4t||α
(2)
(h)||, i.e., the standard scaling.
C.2 Optimization of the bound for the exemplary noise-types
We derive below for exemplary noise types the general bounds on QFI presented Eq. (40) that encapsulate also
the FFQC-assisted schemes and show their optimality.
C.2.1 Rank-one noise
Let’s go back to the general rank-one Liouvillian in Eq. (83)
¯
L
1G
r
= r r
=
γ
2
v
1
v
1
. (91)
We call r
R
= Re(r) = (x
R
, y
R
, z
R
)
T
and r
I
= Im(r) = (x
I
, y
I
, z
I
)
T
, and choose the two other unit vectors v
2
and
v
3
such that the triplet {v
1
, v
2
, v
3
} forms an orthonormal basis. Expressing the component of the Hamiltonian
matrix in this basis, we may restrict to writing:
h
(1)
=
h
R
1
+ ih
I
1
h
R
2
+ ih
I
2
h
R
3
+ ih
I
3
H
(0)
=
H
(0)
11
H
(0)
12
H
(0)
13
H
(0)
12
H
(0)
13
,
(92)
as the missing terms do not enter in α
(2)
in Eq. (79), and thus do not influence the final bound. Setting β
(2)
= 0
in Eq. (85) uniquely specifies the parameters h
R
1
, h
I
1
and H
(0)
11
h
R
1
=
(x
2
I
+y
2
I
)z
R
(x
I
x
R
+y
I
y
R
)z
I
4(|r
R
|
2
|r
I
|
2
(r
R
r
R
)
2
)
h
I
1
=
(x
I
x
R
+y
I
y
R
)z
R
(x
2
R
+y
2
R
)z
I
4(|r
R
|
2
|r
I
|
2
(r
R
r
R
)
2
)
H
(0)
11
=
x
R
y
I
x
I
y
R
4(|r
R
|
2
|r
I
|
2
(r
R
r
R
)
2
)
.
(93)
Accepted in Quantum 2017-09-04, click title to verify 26
The rest of the parameters are free and have to be chosen such that ||α
(2)
|| is minimal. To this end we rewrite
Eq. (79)
α
(2)
= (h
(1)
h
(1)
+
γ
2
v
1
H
(0)
H
(0)
v
1
) σ
0
+ (h
(1)
H
(0)
v
1
) r
σ + (v
1
H
(0)
h
(1)
) σ
r
+ (2v
1
H
(0)
H
(0)
v
1
) σ
(r
R
× r
I
). (94)
A direct way to obtain an upper-bound is to set all the free parameters to zero, this yields
α|
(2)
0
= σ
0
(|h
R
1
|
2
+ |h
I
2
|
2
+ |r|
2
(H
(0)
11
)
2
) (95)
+H
(0)
11
σ
(h
R
1
+ ih
I
1
)r + (h
R
1
ih
I
1
)r
+ 2H
(0)
11
r
R
× r
I
and
||α|
(2)
0
|| =
|x|
2
+ |y|
2
+ 2|x
I
y
R
x
R
y
I
|
16(|r
R
|
2
|r
I
|
2
(r
R
· r
I
)
2
)
=
max{|x + iy|, |x iy|}
2
16(|r
R
|
2
|r
I
|
2
(r
R
· r
I
)
2
)
. (96)
This holds as an upper bound for any r, but might not be optimal in general. Now consider the deviation
from this bound in case where the free parameters are not set to zero, and denote h
1
= h
R
1
+ ih
I
1
, h =
(h
R
2
+ ih
I
2
, h
R
3
+ ih
I
3
)
T
and H = (H
(0)
12
, H
(0)
13
)
T
. One easily obtains
||α
(2)
|| ||α|
(2)
0
|| = |h|
2
+ |H|
2
|r|
2
+ 2|r
R
× r
I
|
+
(H
h + H
(0)
11
h
1
) r + (h
H + H
(0)
11
h
1
) r
(H
(0)
11
h
1
) r + (H
(0)
11
h
1
) r
. (97)
In the case where one of the coefficients of r is zero the only negative term in the expression above vanishes
making ||α|
(2)
0
|| the optimal bound. As follows from Eq. (93), if x = 0 (up to a rotation around z) this happens
because H
(0)
11
= 0, while the case z = 0 implies h
1
= 0. In general, the bound in Eq. (96) could be potentially
improved.
C.2.2 Pauli noise in the z-direction
This case corresponds to the much studied case of dephasing noise [15, 19, 2123]. As the noise commutes with
the evolution our FFQC can not detect it and therefore our FFQC strategy cannot improve the metrological
performance. Nevertheless we treat it here for the sake of completeness.
The Liouvillian is given by
¯
L
1P
z
=
0
0
γ
2
= r r
, (98)
with r = (0, 0,
p
γ/2). It is easy to see from Eq. (85) that β
(2)
= 0 implies 4
p
γ/2 h
r
= 1. Whereas the norm
of α
(2)
in Eq. (79) is given by
||α
(2)
|| = |h
(1)
|
2
+ r
H
(0)
H
(0)
r
+|h
(1)
H
(0)
r + r H
(0)
h
(1)
|. (99)
Which is minimized by
||α
(2)
|| = h
2
r
=
1
8γ
(100)
for h
(1)
= h
r
r
|r|
and H
(0)
= 0.
Accepted in Quantum 2017-09-04, click title to verify 27
C.2.3 Rank-two Pauli noise
Finally, let us consider the case of rank two Liouvillian noise. For convenience we perform a basis rotation so
that the Liouvillian is given by X-Y Pauli noise
¯
L
2P
=
1
2
γ
1
γ
2
0
, (101)
whereas the evolution is given by
ω
2
(s
θ
c
ϕ
, s
θ
s
ϕ
, c
θ
)
T
σ. In the rotated basis, the relevant elements in the series
expansion of the Hermitian operator h read:
h
(1)
=
h
r
1
+ ih
i
1
h
r
2
+ ih
i
2
h
r
3
+ ih
i
3
, (102)
H
(0)
=
H
(0)
11
R
(0)
12
iI
(0)
12
H
(0)
13
R
(0)
12
+ iI
(0)
12
H
(0)
22
H
(0)
23
H
(0)
13
H
(0)
23
H
(0)
33
. (103)
The requirement that β
(2)
= 0 imposes the following conditions
h
r
1
=
s
θ
c
ϕ
4
γ
1
/2
h
r
2
=
s
θ
s
ϕ
4
γ
2
/2
I
(0)
12
=
c
θ
2
γ
1
γ
2
.
(104)
For α
(2)
one gets a cumbersome expression, whose exact form is not important. It is sufficient to notice that
it can be written as
α
(2)
=
σ
0
2
2h
(1)
h
(1)
+ γ
1
(|H
(0)
11
|
2
+ R
(0)2
12
+ I
(0)2
12
+ |H
(0)
13
|
2
)
+γ
2
(|H
(0)
22
|
2
+ R
(0)2
12
+ I
(0)2
12
+ |H
(0)
23
|
2
)
+ f(h
(1)
, H
(0)
) σ
ˆ
n,
(105)
where f(h
(1)
,
¯
H
(0)
) = 0 when all the parameters except h
r
1
, h
r
2
and I
(0)
12
are set to zero. Since this is also the
choice that minimizes the pre-factor of σ
0
in the expression above, it is the optimal choice for the norm of the
whole operator. This implies the optimal bound
||α
(2)
|| = (h
r
1
)
2
+ (h
r
2
)
2
+
γ
1
+ γ
2
2
I
(0)2
12
=
c
2
θ
(γ
1
+ γ
2
) + s
2
θ
(γ
1
s
2
ϕ
+ γ
2
c
2
ϕ
)
8γ
1
γ
2
. (106)
D Analysis of the X-Y noise
The full master equation incorporating the X-Y noise introduced in Eq. (10) reads
dρ
ω
(t)
dt
= i
ω
2
[σ
3
, ρ
ω
(t)] +
γ
2
[p σ
1
ρ
ω
(t)σ
1
+ (1 p)σ
2
ρ
ω
(t)σ
2
ρ
ω
(t)] (107)
and can be integrated analytically, e.g., by methods described in [67], to explicitly derive the probe dynamics.
D.1 Simple FFQC strategy
Let us carefully explain the sequential FFQC strategy used in Sec. 4.3.2 and Sec. 6.1.2, which is inspired
by the strategy of Sec. 4.1 for perpendicular dephasing [28]. Recall that neglecting the higher orders in dt
the probe-plus-ancilla state after an evolution of infinitesimal duration can follow two possible branches, see
Eq. (14):
%(t + dt) = %
C
(t + dt) + %
E
(t + dt). (108)
Accepted in Quantum 2017-09-04, click title to verify 28
Either no error happened %
C
(t+ dt) = (1
γ
2
dt)%(t)i
ω
2
dt[σ
3
, %(t)] and the parameter was imprinted on the
state or with probability
γ
2
dt an error occurred %
E
(t+dt) =
γ
2
dt
p (σ
1
)%(t)(σ
1
)+(1p)(σ
2
)%(t)(σ
2
)
.
If in addition the state at time t belongs to the code space % B(H
C
) with H
C
= span(|00i, |11i) and
H
E
= span(|01i, |10i), the two branches of the infinitesimal evolution can be distinguished in a non-demolition
manner by projecting onto the code the error spaces via a parity measurement (since %
C
(t + dt) B(H
C
) and
%
E
(t + dt) B(H
E
)). If the state happens to be in the correct branch we simply leave it alone, wheres if an
error is detected our strategy consists of mapping the error branch back into the code space by correcting for
the most probable error term that we assume to be σ
2
(0 < p 1/2) without loss of generality. Thus, after
the correction step the error branch reads ¯%
E
(t + dt) = (σ
2
)%
E
(t + dt)(σ
2
) with
¯%
E
(t + dt) =
γ
2
dt
(1 p)%(t) + p(σ
3
)%(t)(σ
3
)
+ O(dt
2
). (109)
Note that here we assume that the control steps are performed fast enough, so that Eq. (109) holds up to first
order in dt
5
. Importantly, after the correction the final state belongs to the code space again ¯%(t + dt) B(H
C
),
and we can continue running the strategy. Another important point is that we keep track of the number of errors
that have been detected before a time t in a classical error register. This can be made explicit by introducing an
error register mode R initialized in the state |0i
R
for each probe qubit, and a Hermitian operator that increases
its count by one R
+
|mi
R
= |m + 1i
R
. For example the register can be modelled by a bosonic mode a with
R
+
=
1
a
a
a
. Such a FFQC strategy leads to an effective evolution of the probe+register system (the ancilla
does not play a role at this stage) given by the master equation
d
dt
%
ω
(t) = i
ω
2
[σ
3
R
, %
ω
(t)]+
γ
2
%
ω
(t) +
γ
2
( R
+
)
(1 p)%
ω
(t) + p(σ
3
R
)%
ω
(t)(σ
3
R
)
( R
+
)
.
(110)
For a single probe initialized in a state |+i =
1
2
1
1
(or
|00i+|11i
2
if we remember the ancillary qubit) the final
state resulting from such an effective evolution after time t reads
%
ω
(t) =
X
m=0
p(m; t) ¯%
m
(t) |mihm|
R
, (111)
where the total number of errors m detected during time t follows the Poissonian distribution p(m; t) =
e
γt/2
(γt/2)
m
/m!, and the state of the probe conditional on the occurrence of m errors is
¯%
m
(t) =
1
2
1 (1 2p)
m
e
iωt
(1 2p)
m
e
iωt
1
!
. (112)
D.2 Description of the probe channel
In particular, one may derive the Choi-Jamiolkowski (CJ) matrix [75] representing the probe map E
X-Y
ω,t
, which
is defined then as P(ω, t) :=E
X-Y
ω,t
I[I] with |Ii=|00i + |11i and reads
P(ω, t) = 2
C
γ
2
0 0 C
i˜ωS
0 S
γ
2
˜γS
0
0 ˜γS
S
γ
2
0
C
+ i˜ωS
0 0 C
γ
2
, (113)
where
C
x
:=
1
2
e
γt
2
cosh(xt), S
x
:=
1
2
e
γt
2
sinh(xt),
µ :=
γ(1 2p)
2
, Ω :=
p
µ
2
ω
2
,
˜γ :=
µ
, ˜ω :=
ω
. (114)
For t >0 the CJ-matrix (113) is of rank four, so the probe channel, E
X-Y
ω,t
, is full-rank always possessing four Kraus
operators in Eq. (12). By performing the spectral decomposition of the CJ-matrix, P(ω, t) =
P
4
i=1
λ
i
|e
i
ihe
i
|,
5
Note that the control steps need not be ultra fast, just sufficiently fast so that the first order approximation holds. The duration
of these control-pulses depends on the strength of the decoherence process.
Accepted in Quantum 2017-09-04, click title to verify 29
we obtain the CJ-canonical Kraus representation of channel E
X-Y
ω,t
after identifying
λ
i
|e
i
i= K
i
⊗ |Ii for all
i, where
K
1
=
q
S
γ
2
˜γS
σ
1
, K
2
= i
q
S
γ
2
+ ˜γS
σ
2
, (115)
K
3
=
r
C
γ
2
q
e
γt
4
+ (˜γS
)
2
p
e
γt
4
+(˜γS
)
2
C
+i˜ωS
0
0 1
,
K
4
=
r
C
γ
2
+
q
e
γt
4
+ (˜γS
)
2
p
e
γt
4
+(˜γS
)
2
C
+i˜ωS
0
0 1
.
On the other hand, we may rewrite the action of the probe channel, E
X-Y
ω,t
, in the Pauli basis and thus derive
its corresponding dynamical matrix S defined in Eq. (11):
S(ω, t) =
C
γ
2
+ C
0 0 i˜ωS
0 S
γ
2
˜γS
0 0
0 0 S
γ
2
+ ˜γS
0
i˜ωS
0 0 C
γ
2
C
. (116)
In the special case of balanced X-Y noise, for which p = 1/2, all the above expressions dramatically simplify
(µ = ˜γ = 0, = iω, ˜ω = i), as the Liouvillian part of Eq. (107) yields a noisy channel that commutes with
parameter encoding, or in other words, the channel E
X-Y
ω,t
becomes then phase-coviariant with respect to rotations
generated by σ
3
[65].
D.3 Parallel CE bound
We follow the methods of Ref. [22] summarised in Sec. 4.2.2, in order to compute the parallel CE bound (22)
for the X-Y noise and the scenario depicted Fig. 3(b):
F
(b)
X-Y
(N, t) := 4 min
h(t)
N||α(t)|| + N(N 1)||β(t)||
2
, (117)
which then allows to determine an upper bound on the maximal QFI rate per probe, defined in Eq. (51):
f
(b)
N
:= max
t
F
(b)
X-Y
(N, t)
N t
. (118)
For given γ, ω, N and t, we substitute into Eq. (117) the canonical Kraus operators of Eq. (115), K(t) =
(K
1
, . . . , K
4
)
T
, and their shifted derivatives,
˙
K(t) ih(t)K(t), so that the minimisation can be efficiently per-
formed by means of semi-definite programming (SDP) [22]. In order to compute f
(b)
N
, we repeat such procedure
for various t, while numerically optimising Eq. (118) over t that then yields t
(b)
opt
.
However, in the special single-qubit (N = 1) and asymptotic cases (N ), we are able to determine the
analytic expressions for F
(b)
(t) in Eq. (117) after correctly choosing an analytic ansatz form (motivated by
the numerical SDPs) for the Hermitian matrix h(t) and explicitly performing the minimisation in Eq. (117).
Although such expressions cannot be claimed to be derived in a fully analytic manner, their correctness can be
verified numerically to arbitrary precision thanks to the SDP formulation.
Nevertheless, due to their cumbersome form we present below only their simplified versions for the special
case of balanced X-Y noise. On the other hand, in case of the asymptotic (N ) regime, we provide a
compact derivation of slightly loosened CE bound (117), which possesses much simpler form but nonetheless
approximates Eq. (117) to negligible precision, what we have numerically verified.
D.4 Parallel CE bound for p= 1/2
In the case of balanced X-Y noise we explicitly compute the CE bound (117) thanks to the recently derived
analytic generalisation of the SDP-formulation of Eq. (117) for phase-covariant, unital qubit maps [65]:
F
(b)
X-Y
(N, t)
p=
1
2
=
N
2
t
2
1 + 2N e
γt
S
γ
2
, (119)
Accepted in Quantum 2017-09-04, click title to verify 30
which is obtained after choosing as the channel Kraus representation the canonical one of Eq. (115), and
optimally setting in Eq. (117):
h
opt
(t)|
p=
1
2
=
t
2κ
0 1 N 0 0
1 N 0 0 0
0 0 κ κ 1
0 0 κ 1 κ
(120)
with κ=1 + 2N e
γt
S
γ
2
.
Thanks to analytic form of Eq. (119), we explicitly compute the corresponding upper bound on maximal QFI
rate (118):
f
(b)
N
p=
1
2
=
2
γ
N
1 + W
2N
eN

2
1 e
1+W
(
2N
eN
)
N
(121)
which occurs at t
(b)
opt
(N) =
1+W
(
2N
eN
)
γ
with W (x) representing the Lambert function.
Note that for N = 2, f
(b)
N=2
p=
1
2
=2/(γe), so that the bound (121) coincides with the QFI rate attained by our
proposed sequential strategy of Fig. 3(a2), i.e., f
(s)
in Eq. (55). As for any smaller p (more noise asymmetry),
the sequential strategy can perform only better in comparison to the parallel schemes (b) of Fig. 1—see also
Fig. 5—this proves that for any p there always exist a non-trivial range of N > 1 for which the sequential
protocol of Fig. 3(a2) outperforms all the parallel schemes.
D.5 Asymptotic parallel CE bound
In the asymptotic regime of N and fixed t, it is always optimal to set β(t) = 0 in Eq. (117) if possible, so
that the parallel CE bound (117) then reads
F
(b)
X-Y
(N , t) = 4N min
h(t) s.t. β(t)=0
||α(t)||. (122)
We determine an upper-bound on the r.h.s. of Eq. (122) making use of the M-matrix formulation introduced
in Sec. 4.2.2 and the alternative expressions for α and β of Eqs. (29) and (30).
Basing on the dynamical matrix S stated Eq. (116), we derive the M matrix following Eq. (28) and choosing
M= M
=
S
T
:
M(ω, t) =
ε + ζ 0 0 iχ
0 0 0
0 0 Γ 0
iχ 0 0 ε ζ
, (123)
with
ε
2
=
Θ
+
2
, ζ
2
=
C
2
, χ
2
=
˜ωS
2
, (124)
2
= S
γ
2
˜γS
, Γ = S
γ
2
+ ˜γS
, (125)
and Θ
±
=C
γ
2
±
q
C
γ
2
2
C
2
˜ωS
2
.
We inspect Eq. (30) and, in order to satisfy the asymptotic condition β(t) = 0, we set following an educated
guess:
h =
˙
ζχ ζ ˙χ
Γ
0 0 0 0
0
0
σ
2
0
0
0 0 0 0
, (126)
which then after evaluating α
(b)
ω,p
(t):= ||α(t)|| according to Eq. (29) yields an upper bound on Eq. (122):
F
(b)
X-Y
(N , t) 4 N α
(b)
ω,p
(t), (127)
Accepted in Quantum 2017-09-04, click title to verify 31
where
α
(b)
ω,p
(t) = 2( ˙ε
2
+
˙
ζ
2
+ ˙χ
2
) +
+
˙
2
+
˙
Γ
2
+
(∆
2
+ Γ
2
)(ζ ˙χ
˙
ζχ)
2
2
Γ
2
(128)
is of complex but importantly of fully analytic form.
E Ziv-Zakai precision bound and the QFI
Consider the parameter ω with prior distribution p
0
(ω). The state at time t is denoted as ρ
ω,t
. The Ziv-Zakai
bound on the AvMSE of the parameter reads [48]
hδω
2
i
Z
0
τ
Z
−∞
min[p
0
(ω), p
0
(ω + τ )]Pr
e
(ω, ω + τ ),
where Pr
e
(ω, ω+τ) is minimum error probability of the binary hypothesis testing between ω and ω+τ . Quantum
mechanically it is given by the trace distance [48]
Pr
e
(ω, ω + τ )
1
2
(1 D[ρ
ω,t
, ρ
ω+τ,t
])
:=
1
2
(1
1
2
||ρ
ω,t
ρ
ω+τ,t
||
1
). (129)
We use the triangle inequality to decompose the trace distance in elementary steps
D[ρ
T
ω
, ρ
T
ω+τ
] min
1,
Z
ω+τ
ω
D[ρ
τ
0
,t
, ρ
τ
0
+
0
,t
]
0
, (130)
and the local expansion of the trace distance D[ρ
τ
0
,t
, ρ
τ
0
+
0
,t
]
0
F(ρ
τ
0
,t
)
2
we rewrite the error probability
as
Pr
e
(ω, ω + τ )
1
2
(1 min
1,
Z
ω+τ
ω
0
p
F(ρ
τ
0
,t
)
2
(131)
Now we are in position to plug in the bound for the QFI, as we showed for the SQL Liouvillians the QFI of the
state at time t satisfies F
L
4α
L
t, such that the error probability fulfills
Pr
e
(ω, ω + τ )
1
2
max(0, 1 τ
α
L
t). (132)
And the Ziv-Zakai bound reads
hδω
2
i
1
2
Z
1
α
L
t
0
τ (1 τ
α
L
t)
Z
−∞
min[p
0
(ω), p
0
(ω + τ )]
Which in the case t 1 leads to
hδω
2
i
1
12α
L
t
, (133)
for any prior which respects the regularity condition lim
τ0
R
−∞
dω min[p
0
(ω), p
0
(ω + τ )] 1.
F Strategy for noiseless frequency estimation
In the context of phase alignment it is known [7] that the state minimizing the cost-function sin
2
(
ϕ
2
) is given by
|Ψi =
r
2
M + 2
N
X
n=0
sin(
π(n + 1)
N + 2
) |ni, (134)
where |ni is the symmetric eigenstate of the total spin J
z
=
1
2
P
N
k=1
σ
(k)
3
corresponding to the eigenvalue
2nN
2
.
We call the rotated state |Ψ
ϕ
i = e
iϕJ
z
|Ψi. For the measurement given by the covariant POVM {|θihθ|} with
|θi = (N + 1)
1/2
P
e
iθn
|ni and the state |Ψ
ϕ
i we denote the probability to observe an outcome θ by
p(θ ϕ) = p(θ|ϕ) := |hθ|Ψ
ϕ
i|
2
. (135)
Accepted in Quantum 2017-09-04, click title to verify 32
It has been shown in [7] that the probability p(δ) in Eq. (135) yields a cost
C =
Z
π
π
sin
2
(
δ
2
)p(δ) =
π
2
4N
2
(136)
in the regime of large N . Note that the cost-function sin
2
(
δ
2
) is appropriate in the context of phase estimation
where the estimated parameter is defined a circle ϕ [π, π). But in the context of frequency estimation the
parameter of interest is defined on the whole line ω (−∞, ) (similar for the resulting phase ϕ = ωt), and the
appropriate cost-function is the variance (MSE). We show below how it can be lower bounded from the value
of C.
Using the fact that δ
2
sin
2
(
δ
2
)π
2
on the interval [π, π] we get a bound on the second moment.
V =
Z
π
π
δ
2
p(δ) π
2
C =
π
4
4N
2
. (137)
On the other hand, one can also upper-bound the tail probabilities from the cost C
T (r) =
Z
π
p(δ)
C
2 sin
2
(rπ/2)
(138)
The average mean square error over all possible outcomes of the covariant POVM is given by
hδϕ
2
i =
Z
π
π
p(θ)
Z
(ϕ ˜ϕ
θ
)
2
p(ϕ|θ) (139)
=
Z
π
π
Z
(ϕ ˜ϕ
θ
)
2
p(θ ϕ)p
0
(ϕ), (140)
where p
0
(ϕ) is the prior knowledge and ˜ϕ
θ
is the optimal estimator of the phase given the measurement outcome
θ. Choosing an explicit estimator ˜ϕ
θ
= θ (which might not be optimal for the MSE) and assume
p
0
(ϕ) =
(
1
2πr
|ϕ| rπ
0 |ϕ| > rπ.
(141)
This implies
hδϕ
2
i =
Z
π
π
Z
(ϕ θ)
2
p(θ ϕ)p
0
(ϕ) (142)
=
1
2πr
Z
π
π
Z
(ϕ θ)
2
p(θ ϕ). (143)
Now we can change the integration variables to Σ = ϕ + θ and = ϕ θ. The domain of integration then is
a rotated rectangle, but since the integrand is positive we can enlarge the domain to the minimal square that
contains it without lowering the integral. This implies
hδϕ
2
i
1
2πr
R
(1+r)π
(1+r)π
R
(1+r)π
(1+r)π
dΣd
2
2
p(∆)
=
r+1
2r
R
(1+r)π
(1+r)π
2
p(∆)d
=
r+1
2r
V + 2
R
π(1+r)
π
2
p(∆)d
. (144)
Finally, because p(∆) is periodic, the second term in brackets in the last expression is upper-bounded by
Z
π(1+r)
π
2
p(∆)d π
2
(1 + r)
2
Z
π
(1r)π
p(∆) d∆ = π
2
(1 + r)
2
T (1 r) (145)
C
π
2
(1 + r)
2
2 sin
2
((1 r)π/2)
, (146)
which yields
hδϕ
2
i C
(r + 1)π
2
2r
1 +
(1 + r)
2
sin
2
((1 r)π/2)
. (147)
Accepted in Quantum 2017-09-04, click title to verify 33
Now, minimizing the above bound over r and substituting for the cost C according to Eq. (136), we obtain
hδϕ
2
i C κ π
2
π
4
κ
4
1
N
2
, (148)
which is attained for r
opt
0.33 that gives κ 6.74.
It remains to show how the above bound applies within the frequency estimation scenario. Let us assume the
prior distribution of the frequency parameter to also have an interval form:
p
0
(ω) =
(
1
2ω
0
if |ω| ω
0
0 if |ω| > ω
0
.
(149)
Remark that if the interval is not centred at zero, it can be always shifted be means of FFQC. Given a total
resource t
0
= N t, we can construct a circuit that is equivalent to running N =
t
0
ω
0
πr
opt
parallel qubits prepared
in the state (134) for a time t =
r
opt
π
ω
0
. This choice maps the frequency estimation task on the estimation of
the phase ϕ = ωt = ω
r
opt
π
ω
0
for the prior defined in Eq. (141) with r = r
opt
, which we have solved above.
Consequently, the MSE for the frequency reads
hδω
2
i =
hδϕ
2
i
t
2
π
4
κ
4
1
t
2
N
2
=
π
4
κ
4
1
(t
0
)
2
. (150)
However, in contract to the case of phase estimation, note that if the evolution is slowed down—as it is for the
rank-one Pauli noise—the time t (and, hence, t
0
) has to be set longer, what affects the frequency MSE above.
Accepted in Quantum 2017-09-04, click title to verify 34