Distribution of Quantum Entanglement: Principles and Applications
Distribution of Quantum Entanglement: Principles and Applications
2020
space
2020
Abstract
1
Dedication
I would like to give a gazillion thanks to my supervisor A/Prof. Paterek for all his
thoughts and guidance. I really appreciate the time we spent discussing and arguing
throughout my PhD programme. I also thank my co-supervisor Dr. Zuppardo for
helping me to get started in this research. Special thanks to Prof. Paternostro for
his advice and collaboration for most parts of our projects.
I thank my current supervisor Prof. Dumke and also Prof. Brukner for taking
interest in our projects and approving my transitions.
Many thanks to our colleagues Mr. Ganardi, Dr. Lee, Prof. Kim, Dr. Mar-
letto, Prof. Vedral, Dr. Noh, A/Prof. Streltsov, Prof. Bandyopadhyay, Dr. Banerjee,
Dr. Deb, Dr. Halder, Dr. Modi, Prof. Sengupta, Mr. Pal, Ms. Batra, and Prof. Ma-
hesh for their collaboration.
I thank Prof. Winter, Dr. Coles, and Prof. Gröblacher for stimulating discus-
sions, which improved our articles.
I wish to express my gratitude to Prof. Laskowski for hospitality at the Univer-
sity of Gdańsk, Dr. Dutta for hospitality and collaboration at the Korea Institute
for Advanced Study, and Prof. Liew for hospitality at Nanyang Technological Uni-
versity.
I had the opportunity to co-supervise undergraduate students who were keen in
doing research. I would like to thank Guo Yao Tham, Wen Yu Kon, Koo Sui Ho
Edmund, Parth Patel, Swetha Sridhar, Jeremy Goh Swee Kang, Chee Mun Yin, and
Tan Xue Yi for their contributions to our projects.
I thank my colleagues Dr. Miller, Dr. Lake, Minh Tran, Kai Sheng, Zhao Zhuo,
Liu Zheng, Lamia Varawala, and Ankit Kumar for the time we spent together.
Thank you to my sister, my mother, my stepfather, and my friends Andhita,
David, Hendra, Wiswa, and Zhonglin for their company. I also thank Mary and
Steven for their hospitality. Towards the end of my PhD, there is a special person
who added new colours in my life. I would like to express my gratefulness to my
girlfriend Qiannan Zhang.
I wish to acknowledge the funding support for our projects by the National
Research Foundation Singapore and Singapore Ministry of Education Academic Re-
search Fund Tier 2 Project No. MOE2015-T2-2-034.
3
Contents
Abstract 1
Acknowledgements 3
Table of contents 5
Abbreviations 9
List of symbols 11
1 Introduction 17
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.1.1 Direct interactions . . . . . . . . . . . . . . . . . . . . . . . . 18
1.1.2 Indirect interactions . . . . . . . . . . . . . . . . . . . . . . . 19
1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.3 Thesis organisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5
3 Revealing non-classicality of inaccessible objects 37
3.1 Motivation and objectives . . . . . . . . . . . . . . . . . . . . . . . . 38
3.2 Instrumental discord . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3 Quantum discord is a resource for entanglement gain . . . . . . . . . 42
3.3.1 The role of correlated mediator . . . . . . . . . . . . . . . . . 46
3.4 Entanglement localisation via classical mediators . . . . . . . . . . . . 47
3.5 Non-classicality detection protocols . . . . . . . . . . . . . . . . . . . 49
3.5.1 Via entanglement breaking channel . . . . . . . . . . . . . . . 49
3.5.2 Via initial entropies . . . . . . . . . . . . . . . . . . . . . . . . 50
3.6 Minimalistic assumptions and applications . . . . . . . . . . . . . . . 51
3.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6
6.3.4 Noisy dynamics: Numerical simulation . . . . . . . . . . . . . 88
6.3.5 Details of entanglement dynamics . . . . . . . . . . . . . . . . 89
6.4 Dynamics of gravitationally induced entanglement: Released masses . 90
6.4.1 Langevin equations and covariance matrix . . . . . . . . . . . 90
6.4.2 Noiseless dynamics: Analytical solution . . . . . . . . . . . . . 91
6.4.3 Quantum and classical trajectories . . . . . . . . . . . . . . . 92
6.5 Other shapes of masses . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.6 Casimir interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.7 Standard decoherence . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.8 Comparison with recent experimental achievements . . . . . . . . . . 97
6.9 Gravity: Direct or mediated interaction? . . . . . . . . . . . . . . . . 97
6.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
7
9.2.1 Entanglement bound for indirect continuous interactions . . . 131
9.2.2 The strength of non-decomposability and correlation gain . . . 132
9.2.3 Protocols witnessing the presence or absence of quantum objects134
9.2.4 Complete link between initial correlations with mediators and
entangling time limit . . . . . . . . . . . . . . . . . . . . . . . 135
9.2.5 Gravity-mediated entanglement: Full treatment . . . . . . . . 135
Bibliography 141
8
Abbreviations
am Air molecules
BChl Bacteriochlorophyll
cg Casimir-gravity
CM Covariance matrix
CV Continuous variable
det Determinant
di Direct interaction
FWHM Full-width at half-maximum
gnd Ground
inf Infimum
int Interaction
LIGO Laser Interferometer Gravitational-Wave Observatory
loc Local
LOCC Local operations and classical communication
max Maximum
min Minimum
ph Photon
PPT Positive under partial trasposition
QC, qc Quantum-classical
QSL Quantum speed limit
RED Relative entropy of discord
REE Relative entropy of entanglement
RHS Right hand side
SEC System-environment correlation
sep Separable
SDE Standard deviation of energy
sq Squeezed
sup Supremum
th Thermal
tr Trace
9
List of notations
A list of general symbols used throughout this thesis is provided below. Some related
notations are defined in groups for better clarity. Special usage in some Chapters
will be mentioned. Note that trivial notations are not included.
11
j
H, HX , HX General Hamiltonian, Hamiltonian applies only on system X,
and its jth variant. A lowercase letter can also indicate variant,
e.g., Hj . Also, Hg is used to indicate gravitational term.
hHi, ∆H Expectation value of energy with respect to the ground state
level, i.e., tr(ρH) − Eg and energy variance. The dimensionless
quantities are written as hM i and ∆M , respectively.
IX:Y Mutual information between systems X and Y . Not to be
confused with IX:Y , see below.
j, j † Annihilation and creation operator, respectively,
for a general d-dimensional quantum system.
jin Used to denote input noise in Chapter 8. The corresponding
noise quadratures are denoted by xin,J and yin,J .
K A drift matrix. Note also the notation W± (t) = exp (±Kt) or
M (t) = exp (Kt).
LX:Y Logarithmic negativity between systems X and Y . Not to be
confused with L which denotes distance.
Lρ, LX ρ Lindblad operation on a quantum state ρ and with the operation
only on system X.
m, mX Mass of an object. Not to be confused with dummy index.
nr Refractive index.
n(t) Noise vector used in Chapter 8.
n̄, N̄ Mean excitation numbers. For example, n̄ = (eβ − 1)−1 is
mean thermal phonon with β = ~ω/kB T .
NX:Y Negativity between systems X and Y .
P Used to indicate pressure. Not to be confused with power P .
Q, QX , QjX A quantum operation, operation applied only on system X,
and its jth variant.
QX:Y General correlation quantifier between systems X and Y .
r1 , r2 , r3 Interaction rates for different configurations in Chapter 6.
rcg The ratio between Casimir and gravitational energy terms,
which are relevant for entanglement gain in Chapter 6.
R, RX Radius, radius of object X. Not to be confused with reflectivity
of mirrors, R1,2 , in Chapter 7.
S, S † Squeezing operators with squeezing strength s. Subscripts can
be used to indicate the application on a quantum system,
e.g., sA , sB .
12
S(ρ), S(ρX ) von Neumann entropy of state ρ and its subsystem ρX ,
also can be simply written as SX .
SX|Y Conditional von Neumann entropy.
S(ρ||σ) Relative entropy distance between states ρ and σ.
T Dimensionless time, T = ωt or T = Ωt, where ω and Ω
being frequencies. Not to be confused with temperature. As a
power, it denotes transposition. Also TX denotes partial
transposition with respect to system X.
u Used to define quadrature vector, e.g., u = (XA , PA , XB , PB )T .
Not to be confused with dummy index.
U, UX Unitary operator in general and only on system X.
V, Vij Covariance matrix and its elements. The elements can also be
denoted by 2 × 2 matrices. For example, for a two-mode system,
the components are IA , IB , L, and LT . A useful quantity is
defined in this regard, i.e., Σ = detIA + detIB − 2 detL. After
partial transposition, a covariance matrix is written as Ṽ .
Vm Mode volume, only defined in Chapter 7.
x, p, X, P Position and momentum. The bold uppercase notations are used
to indicate the corresponding dimensionless quantities (or
quadratures). Also, subscripts are used to indicate their systems,
e.g., xA , pB , XA . For field operators, quadratures are denoted
by X and Y .
X, Y Generally used to indicate quantum systems, but they can
indicate parties each consisting of one or more quantum systems.
X:Y Used to indicate symmetrical partition of systems X and Y .
X|Y Indicates asymmetrical partition of systems X and Y . In
particular, it denotes a situation given the knowledge of system
Y , e.g., measurement outcomes.
C A non-convex set containing quantum-classical states.
Ec Casimir energy between two objects.
F(ρ, σ) Fidelity between quantum states ρ and σ.
Fi Finesse.
IX:Y A distance from a state to the closest product state of the form
σX ⊗ σY as quantified by a general correlation quantifier D(·, ·).
L, LX General map, applies only on system X. It is used in association
with Lindblad operators.
M A non-convex set containing product states.
P(ρ) Purity of a quantum state ρ.
13
Q Mechanical quality factor.
S A convex set containing separable states. Also used to indicate
a set in general for general illustration.
1, 1X Identity operator, identity operator applied only on system X.
|ψi, |ψiX A pure quantum state, state of system X.
A lowercase subscript indicates its variant, e.g., |ψij or |ψj i.
Λ, ΛX A general completely positive trace-preserving map,
applies only on system X. Not to be confused with frequency of
lasers Λm in Chapter 7.
Λph , Λam Decoherence coefficients as a result of scattering photons and air
molecules.
ηg Dimensionless figure of merit for gravity-mediated
entanglement gain between two oscillators. σ is the
corresponding figure of merit of entanglement between two
released masses.
Θ(ρ, σ) Bures angle defined as Θ(ρ, σ) = arccos(F(ρ, σ)).
ΞX Charge of quantum battery for system X.
ω, Ω Used to denote frequencies. In Chapter 8 Section 8.3, we also
use g.
ΩN Symplectic form composed of N CV modes. Not to be confused
with the frequency Ω.
λ Wavelength. Not to be confused with linear mass density or
complex coefficients.
∆0J Cavity-laser detuning. Note also its effective detuning ∆j .
δmn Kronecker delta.
δ(t − t0 ) Direct delta.
κν , υ(t), l(t) Constant vector, noise vector, and a vector defined as
l(t) = υ(t) + κν .
ξ, ξX Brownian noise, affecting system X. Also used to denote a
combined annihilation or creation operators in Chapter 8.
Section 8.3.
γ, γX Decay rate, affecting system X. Note that a notation κ is also
used to indicate decay in Chapter 7 and 8.
τe , τph , τam Entangling time, coherence time due to photon scattering, and
due to air molecules.
14
τ (ρ, σ) The time it takes for a state ρ to evolve to another state σ.
Its dimensionless quantity is denoted as Γ(ρ, σ) or simply Γ.
Note, for example, Γdi stands for dimensionless time bound for
direct interactions scenario.
Γn FWHM of bacterial spectrum in Chapter 7.
µ Dipole moment of two-level transition. Not to be confused with
dummy index.
νk , ν̃k The kth symplectic eigenvalue of a covariance matrix,
and its value after partial transposition. Not to be confused
with the frequency constant ν in Chapter 6
ρ, ρX , ρjX General density matrix, density matrix of system X, and its jth
variant. A lowercase subscript also denotes its variant, e.g., ρj .
Not to be confused with mass density in Chapter 6.
Note that σ is also often used to denote a density matrix.
j
σ j , σX j-Pauli matrix (j = x, y, z). The subscript indicates its
application on system X in the context of multiple two-level
systems.
+(−)
σ +(−) , σX Two-level system raising (lowering) operator in general and for
system X.
ΠjX The jth projection operator of system X.
15
Chapter 1
Introduction
This chapter presents the motivation, objectives, and organisation of this thesis.1 It
begins with the importance of quantum entanglement, leading to the need for under-
standing the process of its creation. In particular, we focus on mediated interaction
scenarios where the aim is to create or distribute entanglement between two quantum
objects by utilising an ancillary system. This sets the stage for the thesis. Finally,
I will describe the objectives of our study and how the thesis is organised.
1
Note that this chapter serves as a brief introduction. Some terminologies used here will be
explained more thoroughly in Chapter 2.
17
1.1 Motivation
Entanglement is a form of quantum correlation first recognised by Einstein, Podol-
sky, and Rosen [1], and Schrödinger [2] in 1935. This correlation has proven crucial
and resulted in key proposals for enabling tasks that are not possible in the classical
regime. This includes, to name a few, quantum cryptography [3], quantum dense
coding [4], and quantum teleportation [5]. The rapid utilisation and manipulation
of quantum entanglement for the purpose of quantum information tasks have made
it a crucial resource, some argued, as real as energy [6]. Therefore, the study of its
creation (or distribution) is a necessity. In this section, we will briefly review proto-
cols for creating quantum entanglement between two principal objects, here denoted
as, A and B. In particular, we describe a scenario where A interacts directly with
B and the one where the interactions are mediated by an ancillary system C. For
the latter case, the nature of the interactions can be discrete or continuous, which
will be the main focus of this thesis.
subscripts indicate the corresponding objects. Note also that the Hamiltonian com-
ponents can be local, i.e., HA ⊗ 1B or 1A ⊗ HB , which we will simply write as HA
and HB hereafter. As an example, let us consider two qubits that are interacting
with Hamiltonian H = ~ω σAx ⊗ σBx , where ~ω is the energy unit and σA(B) x
is the
x-Pauli matrix with the subscript denoting the corresponding object. For the initial
condition, let us consider the state |ψ(0)i = |0iA |0iB , where |0iA(B) is the eigen-
state (ground state) of the z-Pauli matrix. One can see that the dynamics from the
Hamiltonian will create excitations in both objects A and B, i.e., the state at time t
is given by |ψ(t)i = α(t) |0iA |0iB + β(t) |1iA |1iB , where α(t) and β(t) are complex
coefficients. This is a form of state where A is said to be entangled with B, as the
state of one object cannot be written separately from the other.2
Figure 1.1: Direct interaction setting. Two quantum particles A and B are inter-
acting with Hamiltonian HAB .
2
A detailed analysis will be given later in this thesis.
18
1.1.2 Indirect interactions
In some cases, the principal objects are not situated together, e.g., A is in Alice’s
lab and B is in Bob’s lab (see Fig. 1.2). Therefore, they require an ancillary system
to mediate interactions between them. We will refer to this as indirect interaction
setting. Two scenarios can be applied here based on the nature of the interactions.
A discrete interaction scenario (Fig. 1.2a) involves performing operations (or discrete
interactions) on objects A and C in Alice’s lab, then C is sent to Bob’s lab, after
which Bob can perform operations on B and C. On the other hand, a continuous
interaction scenario (Fig. 1.2b) considers continuous (in time) interactions between
A and C as well as between B and C. We note that both scenarios have been
presented by Cubitt et al. [7].
Figure 1.2: Indirect interaction setting. Quantum particles A and B are in different
locations. In discrete interaction scenario (a), an ancillary system C first interacts
with A, then it is sent to Bob, after which it also interacts with B. In scenario (b),
the ancillary system is interacting continuously with A and B. Note that in both
scenarios, A and B are not interacting directly, i.e., there is no Hamiltonian HAB .
Discrete interactions
19
via separable mediators has also been verified experimentally [15–18]. Additionally,
Refs. [19, 20] have shown that the entanglement gain in this scenario is bounded by
another form of non-classical correlation, known as quantum discord, between the
principal particles AB and the mediator C. It demonstrates that the change of quan-
tum entanglement as a result of relocating the object C from Alice’s lab to Bob’s lab
is bounded by the quantum discord carried by C. This way, some form of quantum
correlation is required to distribute entanglement. We also note the distribution in
the presence of noise with the use of different entanglement quantifiers [21, 22].
Continuous interactions
On the other hand, the continuous interaction scenario assumes that the total energy
of the objects is described by Hamiltonian of the form H = HAC + HBC . We note
A and B do not interact directly, and therefore there is no HAB component. More
explicitly, the total Hamiltonian can be written as H = j HAj ⊗HCj 1 + k HBk ⊗HCk 2
P P
(local Hamiltonian terms are included). Note that the subscripts 1 and 2 are used
to indicate that the terms HCj 1 and HCk 2 can be different in general. For most parts
of this thesis, we allow each object to interact with its own local environment, with
the corresponding dynamics following the Lindblad master equation. This way, the
continuous interaction scenario corresponds to ample situations in nature, some of
which will be considered in this thesis. Yet, this entanglement distribution method
is not as well understood as its counterpart, the discrete interaction scenario.
1.2 Objectives
As first shown by Cubitt et al., the continuous scenario allows entanglement distri-
bution via a separable mediator [7]. Investigating further questions in this direction
is therefore the goal of this thesis. In particular, one seeks answers to the following:
In addition to answering these questions, we will also present the implications of our
study and explore applications that can benefit from it.
20
proceed to Chapters 3–8. Chapter 3 presents our results on the resource necessary
for distribution of quantum entanglement via continuous interactions. An impor-
tant and general application will be constructed that is aimed at indirect revelation
of non-classical properties of unknown objects. Another resource for entanglement
distribution will be presented in Chapter 4. More specifically, we link the amount
of distributed correlations to one of the most elementary non-classical features of
quantum observables, that is non-commutativity. Chapter 5 considers the speed of
entanglement distribution for direct and indirect interaction cases. We present con-
ditions for the indirect interaction setting to saturate the maximum entangling speed
that is derived from the quantum speed limit. Major applications of our theorems
will be presented in Chapter 6 and 7. Chapter 6 addresses one of the most unsettled
quests in physics, that is showcasing quantum nature of gravitational interactions.
In particular, we propose experimental setups where gravity can mediate observable
quantum entanglement between two otherwise non-interacting masses. Chapter 7
covers the problem of detecting quantum features in living organisms by proposing
an experimental setup where one can probe non-classical properties of photosyn-
thetic organisms, such as green sulphur bacteria. Finally, other applications will be
presented in Chapter 8.
21
Chapter 2
This chapter provides preliminaries that will be useful for the remainder of this
thesis. First, I describe quantum objects and their states, both for discrete systems
and continuous variable systems. Additionally, quantum systems evolve in time due
to interactions between them or simply as a result of local energy. In this regard,
I provide the dynamics of quantum objects both for closed and open (interactions
with environment) systems. Finally, I proceed to review important properties that
quantum objects can possess. This includes quantities such as von Neumann entropy,
purity, and fidelity, as well as correlations such as quantum entanglement, quantum
discord, and mutual information. More in-depth analysis can be found in Refs. [6,
23–25]. I will also introduce the notion of classicality and non-classicality that will
be used throughout the thesis.
23
2.1 Quantum objects
Within the framework of quantum mechanics, a physical object is fully described
by its wave function (or, more abstractly a state |ψi). Knowing the state of the
object, one can then calculate expectation values of observables such as position,
momentum, and others. In this section, we focus on describing the state of objects
for two cases. One is when an object “lives” in a finite dimensional Hilbert space,
such as the spin state of an electron, while the other is when the description requires
infinite dimension, such as an atom in a harmonic potential. Indeed, the latter has
infinitely many energy levels. Finally, I will also cover the evolution of these objects
both where they are isolated and having interactions with environment.
where, e.g., {|ni} correspond to discrete energy levels,1 d is the dimension of the
object’s Hilbert space, and λn is a complex coefficient. As a simple example, consider
a two-dimensional system such as the spin state of an electron |ψs i = λ0 |0i + λ1 |1i,
where |0i (|1i) denotes spin up (down). Note that normalisation of the state requires
|λ0 |2 + |λ1 |2 = 1, which can be interpreted as the electron having spin up with
probability |λ0 |2 and spin down with probability |λ1 |2 .
A quantum system can also be in a mixture of states (simply, mixed state or
density matrix)
X
ρ= pm |ψm i hψm | , (2.2)
m
P
where |ψm i is any pure state. The normalisation condition requires m pm = 1
and pm ≥ 0 gives the probability of observing the state |ψm i. This requirement
is referred to as positive semidefiniteness, i.e., a physical state ρ has to have real
non-negative eigenvalues with the sum equal to unity. It turns out that any such ρ
has a spectral decomposition
d−1
X
ρ= pn |nihn|, (2.3)
n=0
where {|ni} form a basis in which ρ is diagonal. Note that in a matrix representation,
the mixed state in Eq. (2.3) has dimension d×d, while the pure state |ψi in Eq. (2.1)
1
We will also call them eigenstates and they form an orthonormal basis, i.e., hn|n0 i = δnn0 .
24
is a vector with dimension d. In this thesis, we will use |ψi or ρ to describe the
state of quantum objects, from which one can calculate important quantities. Some
examples will be given later in this chapter.
Two-level systems
A special case where d = 2 (simply, a qubit) will be used often in this thesis. Now
we provide useful notations for the states and operators.
! ! ! !
1 0 1 1 1 1
|0i = , |1i = , |±i = √ , |y± i = √ . (2.4)
0 1 2 ±1 2 ±i
Note that {|0i , |1i}, {|±i}, and {|y± i} each form an eigenbasis of the Pauli matrix
! ! !
1 0 0 1 0 −i
σz = , σx = , σy = , (2.5)
0 −1 1 0 i 0
respectively. Additionally, we also use the raising and lowering operators, respec-
tively written as σ + = |1i h0| and σ − = |0i h1|.
hui uj + uj ui i
Vij = − hui ihuj i, (2.6)
2
a + a† a − a†
X= √ , P = √ , (2.7)
2 2i
2
Note that we adopt the term quadrature vector from literature. However, one should note that
u is not a vector in the normal algebraic sense. For instance, the cross product u × u 6= 0, which
is a consequence of the non-commuting relation between X and P .
3
In this example, V is a single-mode covariance matrix. For an N -mode CM, one writes the
quadrature vector u = (X1 , P1 , X2 , P2 , · · · , XN , PN )T .
25
where a (a† ) is the annihilation (creation) operator of the corresponding infinite
dimensional object. We note that an N -mode covariance matrix has N symplectic
eigenvalues {νk }N
k=1 , which can be calculated from the spectrum of a matrix |iΩN V |,
where !
N
M 0 1
ΩN = , (2.8)
k=1
−1 0
is known as symplectic form. A physical covariance matrix has
νk ≥ 1/2 (2.9)
where the mean phonon number n̄ = (exp(β) − 1)−1 with β = ~ω/kB T , ~ω the
energy unit of the oscillator, and T the temperature. Also, |ji denotes the jth
eigenstate of the operator a† a. The single-mode squeezing operator is given by
η∗
η † †
S = exp aa − a a , (2.11)
2 2
where η = s exp (iθ) is a complex number with s and θ being the squeezing strength
and angle respectively. The corresponding CM is obtained by computing the expec-
tation value of observables in Eq. (2.6) with respect to the state of the object. One
can confirm
!
2s
1 2n̄ + 1 2n̄ + 1 e 0
Vα = 1, Vth = 1, Vsq = −2s
, (2.12)
2 2 2 0 e
where 1 here denotes a 2 × 2 identity matrix and we have assumed the case θ = 0.5
Moreover, the symplectic eigenvalues are given by 1/2, (2n̄ + 1)/2, and (2n̄ + 1)/2
respectively. Note that all the covariance matrices above are physical, i.e., they
satisfy the condition in Eq. (2.9).
4
√
Note that some literature uses the relations in (2.7) without the factor 2, and therefore the
corresponding condition in (2.9) becomes νk ≥ 1.
5
This corresponds to squeezing the momentum (position) quadrature for s > 0 (s < 0).
26
2.1.3 Quantum dynamics
Here we describe the evolution of quantum objects, i.e., how their state |ψi (or, in
general ρ) changes in time.
Closed systems
for pure and mixed states respectively, where U = exp(−iHt/~) is a unitary operator
and H is the Hamiltonian that represents the corresponding total energy of the
object.6 In general, the Hamiltonian can be composed of many terms, i.e., H =
†
P
j Hj . Note also that during the evolution, we have tr(ρ(t)) = tr(U ρ(0)U ) =
tr(ρ(0)U † U ) = tr(ρ(0)) = 1, where we used the cyclic property of trace and unitary
property U † U = U U † = 1.
Let us now provide a tool for the expression of the unitary operator that will be
crucial for some parts of this thesis. In particular, we consider a unitary operator
with Hamiltonian H = H1 + H2 . We utilise the Trotter expansion (see Appendix A
for a simple derivation) and write the operator as
n
i(H1 + H2 )t iH1 ∆t iH2 ∆t
U = exp − = lim exp − exp − , (2.14)
~ n→∞ ~ ~
where ∆t = t/n. Hence, one can think of the evolution as if it consisted of sequences
of two unitary operations with Hamiltonian H2 and H1 (or, in reverse), each for a
time ∆t. For a special case, in which the Hamiltonians commute, [H1 , H2 ] = 0, the
corresponding operator reads
iH1 t iH2 t
U = exp − exp − . (2.15)
~ ~
Open systems
For quantum systems that are open, i.e., having interactions with their environment,
we will consider the evolution of the state following the Lindblad equation
dρ(t) i
= − [H, ρ(t)] + Lρ(t), (2.16)
dt ~
6
We have also assumed that the Hamiltonian here is time-independent which may not apply
in general. In this thesis, we always try to move into a frame where the Hamiltonian is time-
independent (e.g., by changing to a rotating frame) and, wherever convenient, solve the problem
in Heisenberg picture.
27
where Lρ(t) ≡ k Qk ρ(t)Qk† − 12 {Qk† Qk , ρ(t)}, H is the Hamiltonian of the object,
P
and Qk describes the operation on the object as a result of interactions with its
environment.7 Note that, for simplicity, the strength of the interaction is absorbed
in Qk such that the term Lρ(t) has the unit frequency. For the case where the
strength goes to zero, the 2nd term on the RHS of Eq. (2.16) vanishes and the
dynamics becomes unitary as in Eq. (2.13). Also, by using the cyclic property of
trace, it is clear that tr([H, ρ(t)]) and tr(Lρ(t)) both equal zero, and consequently,
the Lindblad equation is trace-preserving.
For a general mixed state, e.g., the one in Eq. (2.3), the von Neumann entropy
P
(simply, entropy) is given by − n pn log2 (pn ). As {pn } form a probability distribu-
tion, the entropy of a quantum object is non-negative. Some other useful properties
of von Neumann entropy are listed below.
28
2.2.2 Purity
Some quantum states are more mixed than the others. In order to quantify the
purity of a state ρ, we use
P(ρ) = tr(ρ2 ). (2.18)
with the help of the cyclic property of trace and the unitary property. Therefore,
unitary dynamics is purity-preserving. While the purity of the state ρ is invariant,
the purity of its subsystems can change. We will see below some examples showing
this case.
As the von Neumann entropy characterises the disorder in the system, one can
also quantify the purity as 1 − S(ρ)/ log2 (d). In this case, its value ranges from a
minimum of 0 to a maximum of unity.
2.2.3 Fidelity
In order to assess how close a quantum state is from another, one of the quantifiers
we use is the Uhlmann root fidelity
q
√ √
F(ρ, σ) = tr ρσ ρ , (2.20)
where ρ and σ are two density matrices having the same dimension [26, 27]. In
a way, the fidelity above is related to a probability of identifying one state as the
other. Some properties of fidelity are listed below
3. For pure states ρ = |ψi hψ| and σ = |φi hφ|, F(ρ, σ) = |hψ|φi|.
One might also consider a measure introduced by Helstrom that shows the best
success probability to distinguish two quantum states, i.e.,
1
pbest = (1 + Dtr (ρ, σ)) , (2.21)
2
29
where Dtr (ρ, σ) = ||ρ − σ||1 /2 is known as the trace distance and || · ||1 is trace
norm [28]. The probability in Eq. (2.21) is then related to the fidelity in Eq. (2.20)
through the following relation [23]
p
1 − F(ρ, σ) ≤ Dtr (ρ, σ) ≤ (1 − F(ρ, σ)2 ). (2.22)
where the subscripts denote the corresponding party. Therefore, if the description of
the whole system XY does not follow Eq. (2.23), the parties X and Y are entangled.
For mixed states, separability is defined as a mixture of product states
X
ρsep = pj ρjX ⊗ ρjY , (2.24)
j
where {pj } form a probability distribution. States which are not of this form there-
fore possess quantum entanglement. We note an important property of quantum
entanglement – monotonicity, i.e., it cannot be created or increased on average un-
der local operations and classical communication (LOCC). It is clear that both the
states in (2.23) and (2.24) can be prepared via LOCC.
As examples, we list below the Bell states (maximally entangled two-qubit states)
1 1
|ψ± i = √ (|01i ± |10i) , |φ± i = √ (|00i ± |11i) . (2.25)
2 2
Note the simplified notation, e.g., |0iX ⊗ |1iY = |01i, will be used throughout this
9
In general, each party can consist of more than one physical object.
30
thesis. In general, some states are more entangled than the others, and therefore are
more useful for performing tasks. Below we provide some exemplary entanglement
quantifiers.
Entropy of entanglement
For pure states, one can quantify the amount of entanglement using the entropy of
entanglement, which is defined below. Schmidt decomposition allows one to write a
pure state |ψi of system XY as
X
|ψi = λj |xj i |yj i , (2.26)
j
where {|xj i} and {|yj i} are orthonormal on system X and Y respectively. The state
of each party is given by partially tracing out the other. For example,
X
ρX = trY (|ψi hψ|) = |λj |2 |xj i hxj | ,
j
X
ρY = trX (|ψi hψ|) = |λj |2 |yj i hyj | (2.27)
j
The information regarding the correlation between X and Y (λj ) is carried in the
states ρX and ρY . The entropy of entanglement is defined as
By using the so-called distance approach based on relative entropy, one can quantify
the amount of entanglement in a state as its distance from the closest separable state.
In particular, we have the relative entropy of entanglement (REE)
where the relative entropy S(ρ||σ) = −tr(ρ log(σ)) − S(ρ) and S is a convex set
which contains separable states of Eq. (2.24) [29]. Note that for pure states, REE
reduces to the entropy of entanglement. In general, REE is hard to compute as it
involves optimisation.
For separable states in Eq. (2.24), it has been shown that the eigenvalues are pos-
itive after partial transposition with respect to one party [30]. This is commonly
31
referred to as the PPT criterion. If a state is not PPT, one argues that the corre-
sponding negative eigenvalues quantify the amount of entanglement in the system.
These negative eigenvalues show that the corresponding density matrix after partial
transposition is not physical. In particular, one defines a computable quantifier of
entanglement known as negativity
||ρTX ||1 − 1
NX:Y = , (2.30)
2
where ||ρTX ||1 is the trace norm of the state after partial transposition with respect
to party X [31–34].10
A logarithmic quantifier, which stems from negativity, can be defined as
and is known as logarithmic negativity. In this thesis, we will mostly used the
logarithmic negativity to quantify quantum entanglement between CV systems. In
this case, the expression is given by
LX:Y = max 0, − log2 (2ν̃min ) , (2.32)
where ν̃min is the minimum symplectic eigenvalue after partial transposition of the
state (or covariance matrix) with respect to one party [34, 35]. Similar to the PPT
criterion, entanglement is shown by the violation of the physicality requirement of
the covariance matrix, i.e., ν̃min < 1/2.
where {|yj i} form an orthonormal basis. We note that there exists a von Neu-
mann measurement on the side of Y that does not perturb the whole system, i.e.,
P j j j P j 11
j ΠY ρΠY = ρ, where ΠY = |yj i hyj | is a projector and j ΠY = 1Y . Other states
will be perturbed by any von Neumann measurement, characterising the quantum-
ness in the correlation. Quantum discord is a non-classical correlation of states that
10
Note that the partial transposition can be done with respect to party Y , i.e., it is symmetrical.
11
To clarify, this is a non-selective von Neumann measurement and will be referred to as simply
von Neumann measurement henceforward.
32
cannot be written as in Eq. (2.33) [36, 37]. We note that entanglement is a stronger
type of quantum correlation in the following sense. As any quantum-classical state
in Eq. (2.33) is a type of separable state in Eq. (2.24), all entangled states contain
quantum discord, but the reverse is not always true. As an example, consider the
state
1
ρ= |0i h0| ⊗ |0i h0| + |1i h1| ⊗ |+i h+| . (2.34)
2
This state is clearly separable, but is discorded since it cannot be written as (2.33)
(with the orthonormal basis on the side of system Y ). One can see that by exchang-
ing the system X ↔ Y in Eq. (2.34), the resulting state is of the form (2.33) and
therefore has zero discord. This emphasises that discord is not symmetrical.
Based on two expressions of mutual information that are equal in the classical regime
but give different values in the quantum regime, a quantifier of quantum discord was
first presented as the difference [37]. In this thesis, however, we will use another form
of quantifier using the distance approach based on relative entropy. In particular,
quantum discord is a distance from a given state ρ to the closest quantum-classical
state of Eq. (2.33), i.e.,
DX|Y = inf S(ρ||σ), (2.35)
σ∈C
Here we note a useful property for the states of Eq. (2.33). Let us consider three
quantum objects A, B, and C with the state written as
X
ρ= pj ρjAB ⊗ |cj i hcj | . (2.36)
j
33
The entanglement of this quantum-classical state follows
X
EA:BC (ρ) = pj EA:B (ρjAB ). (2.37)
j
This equation is known as the flags condition [40]. Furthermore, if one uses the REE
as entanglement quantifier, the closest separable state to ρ reads
X j
σ= pj σAB ⊗ |cj i hcj | , (2.38)
j
j
where σAB is the closest separable state to ρjAB [40].
34
2.3.3 Mutual information
In addition to non-classical correlations mentioned above, quantum systems can
be classically correlated. As a measure of total correlation, we use the mutual
information
IX:Y = S(ρX ) + S(ρY ) − S(ρ), (2.39)
where ρX and ρY are marginals (reduced states) of ρ [43]. One can also phrase the
mutual information using the distance approach. In particular, one writes IX:Y =
inf σ∈M S(ρ||σ), where M is a non-convex set containing product states. As the
closest product state to ρ is simply given by its marginals, the expression simplifies
to Eq. (2.39) [38]. Also note that the mutual information is equal to twice the
entropy of entanglement for pure states.
Finally, all considered correlations are summarised in Fig. 2.1. In this unified
view, one can clearly see that EX:Y ≤ DX|Y ≤ IX:Y . This is a consequence of the
relation for the sets M ∈ C ∈ S.
We should also mention that the presence of quantum discord, e.g., DX|Y 6= 0,
means that the total correlations between systems X and Y , i.e., mutual information,
cannot all be accessed by measuring system Y alone [37]. In the context of this thesis,
we will show how to reveal quantum discord without performing measurements on
system Y , which is one of the main results in Chapter 3.
Conditional entropy
Conditional entropy is a useful quantity that we will utilise later in this thesis. It is
defined as
SX|Y = S(ρ) − S(ρY ). (2.40)
35
We note that for separable states in Eq. (2.24), the conditional entropy follows
SX|Y , SY |X ≥ 0 [44]. Consequently, the negative value, SX|Y , SY |X < 0, implies
quantum entanglement between party X and party Y . This way, one notices a
difference between the classical regime where the conditional entropy is always non-
negative and the quantum regime where it can be negative. For example, any pair
of maximally entangled qubits in Eq. (2.25) gives SX|Y = SY |X = −1.
36
Chapter 3
Revealing non-classicality of
inaccessible objects
In this chapter, the resource for entanglement gain in a scenario where two principal
objects are continuously interacting via a mediating system will be investigated.1 I
begin with an application-related motivation of showcasing quantum features of un-
known objects that are not accessible by direct experimentation and a construction
of a general scenario that will be considered throughout this chapter. Next, by tak-
ing non-correlated systems as the initial condition, I will show that quantum discord
between the mediator and the principal objects is a necessary requirement. On the
other hand, an interesting scenario will be presented where the entanglement is dis-
tributed via a classical mediator, i.e., zero discord. This scheme, however, needs
some correlation already present in the initial state. By utilising this knowledge, I
will devise a detection method capable of revealing a quantum property of an inac-
cessible object, as quantified by the quantum discord. Our protocols use minimalistic
assumptions: no information is required regarding the dimensionality of the objects
involved, the initial state, and the explicit form of the interactions. Also, all the
objects can be open systems and no measurements are applied on the inaccessible
object.
1
Parts of this chapter are reproduced from our published article of Ref. [45], c [2019] American
Physical Society. Where applicable, changes made will be indicated.
37
3.1 Motivation and objectives
What should be known about an inaccessible object to conclude that it is “not
classical”? In this thesis, inspired by quantum communication scenarios, I show the
verification that such object can be used to increase quantum entanglement between
remote probing particles (that individually interact with it) is a sufficient indication.
Specifically, we prove that such gain in quantum entanglement is only possible
if, during its evolution, the object shares with the probes quantum correlations in
the form of quantum discord [24, 36, 37, 46, 47]. In turn, the presence of quantum
discord between the probes and the object entails a non-classical feature of the
object itself. According to the definition of discord, two or more subsystems share
quantum correlations if there is no von Neumann measurement on one of them
that keeps the total state unchanged. This can only happen when non-orthogonal
(indistinguishable) states are involved in the description of the physical configuration
of the measured subsystem. This indistinguishability is the non-classical feature that
we aim to detect. We formulate analytical criteria revealing such non-classicality
based on operations performed only on the probing objects, and without any detailed
modelling of the inaccessible object in question.
We emphasise that the non-classicality is revealed under a set of minimal as-
sumptions. Namely: (i) The object may remain inaccessible at all times, i.e., it
needs not be directly measured. In particular its quantum state and Hilbert space
dimension can remain unknown throughout the whole assessment. Our method is
thus valid when the object is an elementary system or an arbitrarily complex one;
(ii) The details of the interactions between the object and the probes may also re-
main unspecified; (iii) Every party can be open to its own local environment. These
properties make our method applicable to a large number of experimentally relevant
situations. We demonstrate the revealing power of our criteria in Chapters 6, 7, and
8 for the detection of quantum property of gravitational interaction, photosynthetic
organisms, optomechanical mirror in the membrane-in-the-middle setting, and more.
The general scenario we consider in this chapter is depicted in Fig. 3.1. System
C is assumed to be the inaccessible object and to mediate the interaction between
two remote probes, labeled A and B. Therefore, here, we refer to system C as the
mediator. It is essential for our method that the probes are not directly coupled
and only interact via the mediator. This means that the Hamiltonian for the pro-
cess under scrutiny can be written as HAC + HBC , with HJC being the interaction
Hamiltonian between the mediator C and probe J = A, B.
38
Figure 3.1: General setting for revealing non-classicality of inaccessible objects. An
inaccessible object C is mediating interactions between otherwise non-interacting
probing objects A and B. In general, the state of C can be unknown and no
measurement can be conducted on it. In the text, we present conditions for which
entanglement gain between A and B implies non-classicality of C, in the form of
quantum discord DAB|C . Note that our methods do not assume the dimensionality
of any object and the explicit form of the interacting Hamiltonians HAC + HBC .
Also, all the objects can be open to their own local environment (grey shaded area).
Here EX:Y is the relative entropy of entanglement in the partition X : Y [29], and
DX|Y is the relative entropy of discord [38], also known as the one-way quantum
deficit [39]. Note that relative entropy of discord is in general not symmetric, i.e.,
DX|Y 6= DY |X . Eq. (3.1) shows that the change in entanglement due to the relocation
of C is bounded by the quantum discord carried by it.
Let us start from the simple case where the overall probes-mediator system is
closed (which allows us to ignore for now the grey-colored shadows in Fig. 3.1). If
the interaction Hamiltonians HJC satisfy [HAC , HBC ] = 0, the evolution operator
from the initial time t = 0 to some finite time τ is just U = UBC UAC , where
UJC = exp (−iHJC τ /~). This situation is equivalent to first interacting C with A
and then C with B (or in reversed order). However, note that the density matrix
†
ρ0 = UAC ρ0 UAC obtained by “evolving” the initial state through UAC only does not
describe the state of the system at τ . Nevertheless, we now show the relevance of
the properties of state ρ0 for entanglement gain.
Consider the following forms of Eq. (3.1) written for the initial state ρ0 and the
39
instrumental state ρ0 , respectively
0
Note that EAC:B (0) = EAC:B , because UAC is a local unitary operator in this parti-
†
tion. The state at time τ is given by ρτ = UBC ρ0 UBC 0
, and thus EA:BC (τ ) = EA:BC ,
this time owing to UBC being local. Summing the above inequalities we obtain a
bound on the entanglement gain
0
EA:BC (τ ) − EA:BC (0) ≤ DAB|C (0) + DAB|C . (3.3)
1
ρ0 = 2
|011ih011| + 12 |100ih100|, (3.5)
where, e.g., σ z |0i = |0i. One can now readily check that the relative entropy of
entanglement EA:BC grows from 0 to 1 in the timespan from T = 0 to T = π/4,2
whereas discord DAB|C remains zero at these two times. The gain is indeed due to
non-classical correlations of the instrumental state: applying only UAC for a time
0
T = π/4 produces discord DAB|C = 1. A summary of this dynamics is presented in
Fig. 3.2.
For completeness, I provide the detailed calculations below. First, we note that
the Hamiltonian has commuting components, i.e., [HAC , HBC ] = 0 and therefore we
write the evolution operator as U = UBC UAC . Next, entanglement in the partition
2
For simplicity, we define a dimensionless time variable T = ωt.
40
Figure 3.2: Entanglement distribution utilising instrumental discord. The Hamil-
tonian is taken to be H = (σAx ⊗ 1 ⊗ σCx + 1 ⊗ σBx ⊗ σCx ) ~ω with initial state
1
2
|011ih011| + 12 |100ih100|. Note the maximum quantum discord is DAB|C ≈ 0.81 at
ωt = π/8. This shows that the entanglement gain is not bounded by the maximum
discord.
A : BC is calculated as follows
0
EA:BC (T ) = EA:BC
1 † 1 †
= EA:C (UAC |01i h01| UAC ) + EA:C (UAC |10i h10| UAC ), (3.6)
2 2
0 †
where now EA:BC = EA:BC (UAC ρ0 UAC ) for a time T and the steps are justified
as follows. In the first line, we use the fact that UBC is local in the partition
A : BC. Next, by applying UAC operator to the initial state, the state of B is still
classical, i.e., represented by an orthogonal basis {|0i , |1i}. Therefore, we arrive at
the second line with the help of the flags condition [40]. We note that the states
in the second line are pure and one can equate the entanglement EA:C with the
corresponding von Neumann entropy of object A or C. As an example, by using
UAC = cos(T )1 − i sin(T )σAx σCx , one has for the first term
†
trB UAC |01i h01| UAC = cos2 (T ) |0i h0| + sin2 (T ) |1i h1| , (3.7)
giving an entropy − sin2 (T ) log2 (sin2 (T )) − cos2 (T ) log2 (cos2 (T )). The calculation
for the second term gives the same result, and therefore the entanglement dynamics
is simply given by
41
0
Now I will show that the evolution of DAB|C is equal to EA:BC (T ) as follows.
0 0
First, note that EA:BC (T ) = EA:BC = EAB:C , where the second equality is from
the flags condition as a result of classical B. By showing that the state ρ0 has
0 0
the closest separable state σAB:C that is classical on C, we confirm that DAB|C =
0
EAB:C = EA:BC (T ). In order to do this, we recall that the flags condition also states
that the closest separable state of a quantum–classical density matrix ρ0 (classical
B) is given by
0 1 01 1 02
σAB:C = σA:C ⊗ |1iB h1| + σA:C ⊗ |0iB h0| , (3.9)
2 2
01 02 †
where σA:C and σA:C are the closest separable states respectively to UAC |01i h01| UAC
†
and UAC |10i h10| UAC . Next, we use the fact that the closest separable state of a
pure state is given by the density matrix that is de-phased in the local Schmidt basis,
01
to arrive at σA:C 02
= cos2 (T ) |01i h01|+sin2 (T ) |10i h10| and σA:C = cos2 (T ) |10i h10|+
sin2 (T ) |01i h01|. By putting the expressions together, one realises that σAB:C0
is also
classical on the side of C, hence proving the claimed statement above.
Finally, the dynamics of DAB|C is computed numerically. We note that the von
Neumann measurement on C minimising the discord is given by the basis {|0i , |1i}
for T = [0, π/8] and {|y+ i , |y− i} (the eigenbasis of σ y ) for T = [π/8, π/4].
For general non-commuting interaction Hamiltonians, one can pursue a similar
analysis with the help of the Suzuki-Trotter expansion (see Appendix A for a simple
proof). The evolution operator U is now discretised into short-time interactions of
C with A and then B (or the reversed order) as
n
U = lim e−iHBC ∆t/~ e−iHAC ∆t/~ , (3.10)
n→∞
42
(might be in the form of classical correlation). We begin with the following theorem.
Theorem 3.3.1. Consider two quantum objects A and B that are interacting via a
mediator C, but not with each other. The entanglement follows
Proof. For simplicity, the Hamiltonian of the system is taken as H = HAC + HBC ,
where HAC = HA ⊗ HC1 and HBC = HB ⊗ HC2 .3 We note that it is straightforward
to extend the proof for more general Hamiltonians HAC = µ HAµ ⊗HCµ1 and HBC =
P
P ν ν
ν HB ⊗ HC2 .
As we assume, at all times, quantum–classical state in the partition AB : C, one
writes the initial density matrix of the system as
X
ρ0 = pc ρAB|c ⊗ |ci hc| , (3.12)
c
where {|ci} and {|φc i} both form orthonormal bases of the mediating system C.
Now we incorporate the local environment of all the objects under scrutiny. In
particular, the dynamics of the whole system assumes the following coarse-grained
master equation in Lindblad form
ρ∆t − ρ0 i X
= − [H, ρ0 ] + LX ρ0 , (3.14)
∆t ~ X=A,B,C
where the first term on the RHS is the coherent evolution while the second de-
scribes incoherent interactions with the local environments. We recall that LX ρ0 ≡
P k k† 1 k† k k
k QX ρ0 QX − 2 {QX QX , ρ0 }, with the operator QX only acting on object X. Note
that the strength of the interactions with environments have been absorbed in the
operator QkX .
By utilising the master equation (3.14), one obtains the following conditional
3
Note that the term HC1 can be different from HC2 in general.
43
state after a short time ∆t
We note that the continuous dynamics follows by taking the limit ∆t → 0 and
applying the short time evolution successively. In this proof, terms of the order
O(∆t2 ) will not be taken into account, and we use “'” to denote equality where the
O(∆t2 ) terms are irrelevant and ignored. In this short time, we write the change
in the basis of C, up to ∆t order, as |ci → |φc i = αc |ci + βc |c⊥ i, where |c⊥ i and
|ci are orthogonal, and βc ∝ ∆t. As a result, one gets |hc|φj i|2 ' δcj , making
hφj | ρ0 |φj i ' pj ρAB|j .
The coherent part of the dynamics in Eq. (3.15) is proportional to ∆t, and
therefore, we render ∆t order terms in hφj | [H, ρ0 ] |φj i irrelevant. As a result, we
have
hφj | [H, ρ0 ] |φj i ' pj [ECj 1 HA + ECj 2 HB , ρAB|j ], (3.16)
where we define ECj 1 (C2 ) ≡ hj| HC1 (C2 ) |ji as the expectation value of energy.4 Note
that, in this way, the coherent component is simply equivalent to an effective dy-
namics that are local in A and B.
On the other hand, the incoherent part in Eq. (3.15) consists of three terms. The
first two in the summation are given by
and they are local operations with respect to object A and object B. The last one
reads
XX
hφj | LC ρ0 |φj i ' pc | hj| QkC |ci |2 ρAB|c
c k
X
− pj hj| Qk† k
C QC |ji ρAB|j . (3.18)
k
Plugging all these findings to Eq. (3.15), we write the conditional state explicitly
4
For clarity, we note that one can make these expectation values dimensionless such that the
energy unit is absorbed in the terms HA and HB respectively.
44
as
hj| Qk† k
P
pj (1 − k C QC |ji ∆t)
ρAB|j (∆t) ' ρ̃AB|j
pj (∆t)
X pc X
+ | hj| QkC |ci |2 ∆t ρAB|c , (3.19)
c
pj (∆t) k
where we define
∆t
ρ̃AB|j ≡ ρAB|j − i[ECj 1 HA + ECj 2 HB , ρAB|j ]
~
+ (LA + LB )ρAB|j ∆t. (3.20)
Therefore, the density matrix ρ̃AB|j gives the evolution of ρAB|j through an effective
Lindblad master equation, where the interactions are local on the side of A and B.
By using tr(ρAB|j (∆t)) = 1, one obtains the probability
!
X
pj (∆t) = pj 1− hj| Qk† k
C QC |ji ∆t
k
X X
+ pc | hj| QkC |ci |2 ∆t, (3.21)
c k
where we have utilised the cyclic property of trace. Note that the factors for the
states ρ̃AB|j and ρAB|c in Eq. (3.19) are all real, non-negative, and sum up to unity.
Finally, the entanglement in the partition A : BC reads
X
EA:BC (∆t) = pj (∆t) EA:B (ρAB|j (∆t)) (3.22)
j
X
≤ pj EA:B (ρAB|j )
j
X X
− pj hj| Qk† k
C QC |ji ∆t EA:B (ρAB|j )
j k
XX X
+ pc | hj| QkC |ci |2 ∆t EA:B (ρAB|c ), (3.23)
j c k
X
= pj EA:B (ρAB|j ) = EA:BC (0), (3.24)
j
where we explain the steps taken above as follows. The flags condition [40] is applied
in the first line to the quantum–classical state in the partition AB : C. Next, the
inequality is obtained by using the convexity property of entanglement on the state
ρAB|j (∆t) in Eq. (3.19), and EA:B (ρ̃AB|j ) ≤ EA:B (ρAB|j ) from the monotonicity of
entanglement under local interactions. We note that the second and third lines in
Eq. (3.23) cancel out. This is apparent by inserting c |ci hc| = 1 between Qk†
P
C and
k
QC in the second line and exchanging the dummy indices c ↔ j. The last equality
45
is obtained from the flags condition applied to the initial state. The conclusion of
the theorem follows by applying the arguments above successively from t = 0 to τ
with the condition DAB|C (t) = 0 at all times.
Theorem 3.3.1 above has an important special case, which is shown by the corol-
lary below.
Corollary 3.3.1.1. For the premise of Theorem 3.3.1, where all the objects are
closed systems, we have
EA:BC (τ ) = EA:BC (0). (3.25)
Proof. In this case, the dynamics is simply given by a unitary operator with Hamil-
tonian H = HAC + HBC . The conditional state of Eq. (3.19) simplifies to
pj
ρAB|j (∆t) = ρ̃AB|j , (3.26)
pj (∆t)
where ρ̃AB|j now reads ρAB|j −i[ECj 1 HA +ECj 2 HB , ρAB|j ]∆t/~. One arrives at pj (∆t) =
pj and ρjAB (∆t) = ρ̃jAB by taking the trace of Eq. (3.26). Therefore, ρAB|j (∆t) is
simply an evolution from the initial state ρAB|j via an effective unitary operator that
is local in both A and B, making the corresponding entanglement in the partition
A : B invariant. Finally, from the flags condition applied to the initial state and the
state after ∆t one gets EA:BC (∆t) = EA:BC (0). The Corollary follows by considering
successive applications of the arguments above for a time τ .
We emphasise the generality of Theorem 3.3.1, where both the mediator and
probing objects are open to their own local environments. This matches a large
number of experimentally relevant situations – some of the important ones will be
addressed later in this thesis. It is also worth stressing that this Theorem extends the
monotonicity of entanglement under local operations and classical communication
(LOCC) [48] to the case of continuous interactions. In general, zero-discord states
are good models for classical communication as they allow for continuous projective
measurements on C that do not disturb the whole multipartite state.
Lemma 3.3.2. Consider the premise of Theorem 3.3.1. It follows that the gain
of entanglement in the partition A : B implies the existence of correlation in the
partition AB : C.
46
Proof. This is proof by contradiction. Let us write the initial state with an uncor-
related mediator C as ρ(0) = ρAB ⊗ ρC . This gives us EA:BC (0) = EA:B (0) from the
monotonicity of entanglement under tracing out the uncorrelated object C. Fur-
thermore, if one has quantum discord satisfying DAB|C (t) = 0 for t ∈ [0, τ ], one can
utilise the conclusion of Theorem 3.3.1 and obtain
where the first inequality is also a consequence of the monotonicity property. This
shows that the observation of entanglement gain EA:B (τ ) > EA:B (0) implies that
either the initial state ρ(0) 6= ρAB ⊗ ρC or DAB|C (t) 6= 0 at some time during the
dynamics. In either case, the entanglement gain detects the presence of correlation
between AB and C.
Let us note that in many experimental settings, one usually assumes decoupled
objects as the initial condition, i.e., ρ0 = ρA ⊗ ρB ⊗ ρC . As this type of state is a
special case of ρAB ⊗ ρC considered in Lemma 3.3.2, one concludes that, for initial
decoupled objects, quantum discord is a necessary resource for entanglement gain
between the principal objects.
where σ x |±i = ±|±i, and |ψ+ i = √12 (|01i + |10i) and |φ+ i = √12 (|00i + |11i)
are two Bell states between subsystems AB. As the initial state in Eq. (3.28)
contains the eigenstates of HC , the system remains classical, as measured on C,
at all times. Furthermore, the classical basis is the same at all times. Yet, one
can verify that the relative entropy of entanglement between the probes is given
by EA:B (T ) = 1 − SAB (T ), where SAB (T ) is the von Neumann entropy of the AB
state at time T , and oscillates between 0 and 1. Note this means, in general, that
entanglement gain in the partition A : B does not signify the non-classicality of C
(nonzero DAB|C ). The corresponding dynamics is illustrated in Fig. 3.3. One can
47
also see from Fig. 3.3 that there is entanglement already present initially in the bigger
partition A : BC. The subsequent evolution only localises such entanglement to the
A : B partition. Let us therefore call this phenomenon entanglement localisation.
Figure 3.3: Entanglement distribution via classical ancillary system. The Hamil-
tonian is taken to be H = (σAx ⊗ 1 ⊗ σCx + 1 ⊗ σBx ⊗ σCx ) ~ω with initial state
1
2
|ψ+ i hψ+ | ⊗ |+i h+| + 12 |φ+ i hφ+ | ⊗ |−i h−|. The dynamics shows that the entan-
glement between A and B is increasing while the quantum discord DAB|C is zero at
all times.
For completeness I will provide the calculations for the dynamics in Fig. 3.3. Let
us start with the state of AB at time T
† †
ρAB (T ) = trC UBC UAC ρ0 UAC UBC
1 1
= ρα + ρβ , (3.29)
2 2
where we define ρα as the density matrix for a pure state |αi = cos(2T ) |ψ+ i −
i sin(2T ) |φ+ i and similarly, ρβ for |βi = cos(2T ) |φ+ i + i sin(2T ) |ψ+ i. As ρα and
ρβ are both pure states, one can calculate the corresponding entanglement in the
partition A : B by von Neumann entropy of either A or B. It is straightforward to
obtain that EA:B (ρα ) = EA:B (ρβ ) = 1, i.e., both states are maximally entangled at
all times. From the definition of REE:
In this case, EA:B (ρα ) = 1 and S(ρα ) = 0. Therefore, we have −tr (ρα log2 (σα )) ≥ 1
(*), and similarly, −tr (ρβ log2 (σβ )) ≥ 1 (**). Now we show that there exists a single
48
separable state that achieves the minima in both (*) and (**). By choosing
1 1
σ= |ψ+ i hψ+ | + |φ+ i hφ+ | , (3.31)
2 2
one directly verifies that −tr (ρα log2 (σ)) = −tr (ρβ log2 (σ)) = 1. Now, we calculate
the entanglement as follows
1 1
EA:B (T ) = min −tr ρα + ρβ ||σ − SAB (T )
σ 2 2
1 1
= min (−tr (ρα ||σ)) + (−tr (ρβ ||σ)) − SAB (T )
σ 2 2
= 1 − SAB (T ), (3.32)
where we have used (3.31) as the minimising separable state. See Appendix B for
the dynamics of EA:BC and DAB|C , and for a general prescription of entanglement
localisation. We note that our entanglement localisation proposal has been realised
experimentally [49].
We note that similar considerations have been presented in Ref. [50] to provide
a counter-example to the impossibility of entanglement gain via LOCC. However,
as mentioned above, the partition A : BC is entangled already from the beginning
(in our example we have EA:BC (0) = 1). Therefore, the entanglement localisation
emphasises that the ancillary systems within the framework of LOCC, here C, are
not allowed to be initially correlated with the principal objects A and B, even if the
correlations are classical.
49
Now, if we ensure by operating on the probes only that the initial entanglements
coincide, i.e. EA:BC (0) = EA:B (0), entanglement gain in system AB is only possi-
ble due to nonzero discord DAB|C . As we are interested in observing entanglement
gain, it is natural to start with as small entanglement as possible. This leads us
to propose the application of an entanglement-breaking channel to one of the prin-
cipal systems, at time t = 0. Indeed, after application of the channel on A, we
have EA:B (0) = EA:BC (0) = 0. In a more concrete example, the channel is a von
Neumann measurement. An arbitrary measurement is allowed and experimentalist
should choose the one having potential for biggest entanglement gain. Note that
the measurement results need not be known. Our method is symmetrical, i.e., the
entanglement-breaking channel can be applied on B. This is a consequence of the
symmetry in Theorem 3.3.1, leading to EA:B (τ ) ≤ EB:AC (τ ) ≤ EB:AC (0). Together
with the condition after the channel, EA:B (0) = EB:AC (0) = 0, they give the same
conclusion. The non-classicality detection method is illustrated and summarised
in Fig. 3.4a. Note that entanglement estimation in step (iii) can be realised with
entanglement witnesses [6, 51], rendering state tomography unnecessary.
Lemma 3.5.1. Provided the premise of Theorem 3.3.1, the following inequality holds
where the entanglement quantifier here is REE and SX denotes the von Neumann
entropy of object X.
Proof. Let us begin here with Eq. (3.33), that is EA:B (τ ) ≤ EA:BC (τ ) ≤ EA:BC (0).
Now we will provide an upper bound to the initial entanglement EA:BC (0) in terms
of initial von Neumann entropies of the accessible objects A and B. First, note that
REE is bounded by mutual information [38], which in our case reads EA:BC ≤ IA:BC .
We also have IA:BC ≤ SA + SB − SAB|C from the sub-additivity of von Neumann
entropy. As the state of object C is classical, the conditional entropy follows SAB|C ≥
0, and therefore
EA:B (τ ) ≤ EA:BC (0) ≤ SA (0) + SB (0). (3.35)
50
Figure 3.4: Summary of protocols for revelation of non-classicality of an inaccessible
object C. (a) Protocol via entanglement breaking channel involves the following
steps: (i) entanglement breaking channel on object A (or object B), e.g., von Neu-
mann measurement; (ii) dynamics of the whole system; (iii) estimation of entan-
glement between A and B. In this scheme, nonzero entanglement implies positive
discord DAB|C . (b) Protocol via initial entropies has similar steps with the excep-
tion of measurement of initial entropies for objects A and B. This scheme reveals
positive DAB|C if the entanglement between A and B is larger than the sum of initial
entropies, i.e., EA:B > SA (0) + SB (0). Note that for both protocols we have mini-
malistic assumptions. All objects can be open systems and no assumption is made
regarding the initial state nor the explicit form of the Hamiltonian HAC + HBC .
51
3.7 Summary
In this study we considered two principal objects that are only indirectly interacting,
via a mediator. We have shown that quantum discord (a form of quantum property
of the mediator) is a resource for entanglement gain. This prompted us to propose
schemes for assessing quantumness of an inaccessible object (the mediator). The
schemes are based on monitoring entanglement dynamics between the principal ob-
jects, which serve as probes over which we have control. Our method is robust and
experimentally friendly as it allows the controlled objects and the inaccessible one
to be open systems, makes no assumptions about the initial state, dimensionality of
involved Hilbert spaces, and details of the interaction Hamiltonian.
52
Chapter 4
Non-decomposability of evolution
& extreme gain of correlations
In this chapter, the amount of distributed correlations will be studied.1 I show that
non-commutativity of interaction Hamiltonians (or, in a more general setting, non-
decomposability of time evolution) is a necessary resource for having high gain of
correlations between two objects. From the extreme gain of correlations one can
then detect the non-commutativity, which shows another level of quantum property
that is present in the system. This chapter starts with a motivation and a construc-
tion of a general setting that will be considered herein. I will then proceed to show
that for decomposable dynamics the correlations between two principal objects are
bounded by a correlation capacity of a mediating system. This applies for a plethora
of correlation quantifiers, some of which will be presented in this chapter. Next, I
will discuss a special scenario where all the objects involved are open to their local
environments. Finally, the origin of the possible extreme gain of correlations will be
covered.
1
Parts of this chapter are reproduced from our published article of Ref. [55], c [2019] American
Physical Society. Where applicable, changes made will be indicated.
53
4.1 Motivation and objectives
All classical observables are functions of positions and momenta. Since there is
no fundamental limit on the precision of position and momentum measurement in
classical physics, all classical observables are, in principle, measurable simultane-
ously. Quite differently, the Heisenberg uncertainty principle forbids simultaneous
exact knowledge of quantum observables corresponding to position and momentum.
The underlying non-classical feature is their non-commutativity: Any pair of non-
commuting observables cannot be simultaneously measured to arbitrary precision,
as first demonstrated by Robertson in his famous uncertainty relation [56]. Other
examples of non-classical phenomena with underlying non-commutativity of observ-
ables include violations of Bell inequalities [57, 58] or, more generally, non-contextual
inequalities; e.g., see [59]. Here we describe a method to detect non-commutativity of
interaction Hamiltonians, and generally non-decomposability of temporal evolution,
from the dynamics of correlations.
Consider the situation depicted in Fig. 4.1, where the probing systems A and
B do not interact directly but only via the mediator C; i.e., there is no Hamil-
tonian term HAB . In general, we allow all objects to be open systems and study
whether the evolution operator cannot be represented by a sequence of operations
between each probe and the mediator, i.e., ΛABC = ΛBC ΛAC or in reverse order.
For the special case in which all the systems are closed, non-decomposability im-
plies non-commutativity of interaction Hamiltonians, i.e., [HAC , HBC ] 6= 0. Indeed,
for commuting Hamiltonians, the unitary evolution operator is decomposable into
UBC UAC , where, for example, UAC = exp(−itHAC /~). We show that for decom-
posable evolution, correlations between A and B are bounded. We also show with
concrete dynamics generated by non-commuting Hamiltonians that these bounds
can be violated. The bounds derived depend solely on the dimensionality of C
and not on the actual form of the evolution operators. Hence, these operators can
remain unknown throughout the assessment. This is a desired feature, as experi-
menters usually do not reconstruct the evolution operators via process tomography.
It also allows applications of the method to situations where the physics is not un-
derstood to the extent that reasonable Hamiltonians or Lindblad operators can be
written down. Furthermore, the assessment does not depend on the initial state
of the tripartite system and does not require any operations on the mediator. It
is therefore applicable to a variety of experimental situations; Refs. [52, 53, 60, 61]
provide concrete examples.
Below I shall begin by presenting the general bounds on the amount of corre-
lations one can establish if the evolution is decomposable. It is shown that these
bounds are generic and hold for a large number of correlation quantifiers. We then
54
Figure 4.1: General setting for detecting non-decomposability of time evolution.
High dimensional objects A and B are interacting with a low dimensional object C,
but not with each other. In general, we allow all objects to be open to their local
environments. The interaction Hamiltonians are given by HAC and HBC , which
describe the coherent part of the dynamics. In the text, we show that large gain
of correlation between A and B, exceeding certain thresholds (functions of the di-
mensionality of C), implies non-decomposability of time evolution, which makes the
latter crucial for substantial correlation distribution. Note that a lot of correlation
quantifiers are useful in our method. This setup is different from the one considered
in Fig. 3.1 of Chapter 3. For non-decomposability detection, here the dimension of
C is assumed to be known and, as we will show in the text, small value of such
dimension is preferable.
55
for varied initial states and more correlation quantifiers.
Proof. Let us suppose that one performs an operation on party Y . Based on the
monotonicity property, QX:Y cannot increase in value. Furthermore, the value also
cannot decrease for a reversible operation, because this means QX:Y can increase
by reversing the process. Therefore, QX:Y can only be invariant. As tracing-out
uncorrelated objects is a reversible process, the conclusion follows.
It follows that for a decomposable dynamical operator, i.e., ΛABC = ΛBC ΛAC , one
has
QA:B (t) ≤ IAC:B (0) + sup QA:C , (4.1)
|ψi
where we define here IAC:B (0) ≡ inf σAC ⊗σB D(ρ, σAC ⊗ σB ), use ρ as the initial
density matrix, and the supremum in the last term is over pure states {|ψi} of AC.
56
Proof. First, we show that QX:Y is monotone under local operations ΛY . This is
done as follows
0
where σXY is a state in S closest to ρXY . The second line in (4.2) is due to property
0
(i) and that ΛY [σXY ] might not be the the closest to ΛY [ρXY ], while the third line
follows from property (ii). This allows us to use the property in Lemma 4.2.1.
Next, we have the following arguments
where the steps are explained as follows. We utilise that QX:Y is monotone under
local maps on Y (in this case, tracing out C) in line (4.3). Similarly, in (4.4), QA:BC
is monotone under the operation ΛBC . The inequality (4.5) utilises the definition of
QA:BC as a distance-based quantifier and holds for any state µ ∈ SA:BC . The triangle
inequality (iii) confirms line (4.6). We note that the first term in (4.6) is independent
0
of µ and one can choose any state σAC and σB0 at this point. Property (ii) is utilised
0
in equality (4.7). For line (4.8), the state σAC ⊗ σB0 is chosen as the closest product
0
state to ρ, while µ ∈ SA:BC as the closest state to ΛAC [σAC ] ⊗ σB0 . The equality
(4.9) is obtained as QA:BC is invariant under the partial trace of the decoupled state
σB0 . The final line follows from a property that a bipartite correlation quantifier is
maximal on pure states if it is non-increasing under local operations on at least one
party [62].
Note that although the relative entropy does not satisfy (iii) it still follows The-
57
orem 4.2.2. This is the result of the following lemma.
Lemma 4.2.3. For a distance measure based on the relative entropy the conclusion
in Theorem 4.2.2 follows.
where ρX and ρY are the reduced states of ρ. We have also used, e.g., the relation
tr(ρ log σX ⊗ σY ) = tr(ρX log σX ) + tr(ρY log σY ).
While relative entropy is known to follow property (ii) [63], it does not satisfy
property (iii). In this case, one starts with line (4.4) in Theorem 4.2.2 and obtains
where ρ0AC and ρ0B are the reduced states of ΛAC [ρ]. We justify the steps above as
follows. In line (4.13), we use a state of the form µAC ⊗ µB that is in the set of
SA:BC . The identity of (4.11) confirms line (4.14). From the definition of mutual
information as the distance from a state to its marginals [38], one arrives at equality
(4.15). Note that µAC ∈ SA:C has been chosen as the closest to ρ0AC and µB = ρ0B .
The final line is obtained from the monotonicity of mutual information under the
local map ΛAC and the supremum of QA:C is attained over pure states.
Correlations between probe A and probe B are therefore bounded by the maximal
achievable correlation with the mediator, sup|ψi QA:C . The additional term IAC:B (0)
in Eq. (4.1) reduces to the usual mutual information IAC:B (0) if D(ρXY , σXY ) is the
relative entropy distance [38] and characterizes the amount of total initial correla-
tions between one of the probes and the rest of the system. Note that the bound
is independent of time. This can be seen as a result of the effective description of
58
such dynamics given by ΛBC ΛAC . The particle C is exchanged between A and B
only once, independently of the duration of the dynamics.
In a typical experimental situation the initial state can be prepared as completely
uncorrelated ρ = ρA ⊗ρB ⊗ρC , in which case Theorem 4.2.2 simplifies and the bound
is given solely in terms of the “correlation capacity” of the mediator:
Clearly, the same bound holds for initial states of the form ρ = ρAC ⊗ ρB . In
Section 4.2.2 we show that, with this initial state, Eq. (4.17) holds for any correlation
quantifier that is monotonic under local operations ΛBC , not necessarily based on
the distance approach, e.g., any entanglement monotone.
Theorem 4.2.4. Consider the initial density matrix of the system being in the form
ρ = ρAC ⊗ ρB . For a decomposable dynamical operator ΛABC = ΛBC ΛAC , it follows
that
QA:B (t) ≤ sup QA:C , (4.18)
|ψi
where Q is any correlation quantifier that is monotone under local maps ΛBC .
where the steps are justified as follows. The first inequality is obtained as tracing
out party C (which might be correlated with AB in general) is a local operation on
the side of BC. The monotonicity of the quantifier under ΛBC confirms the second
argument. Party B stays uncorrelated after the application of ΛAC on the initial
density matrix ρAC ⊗ ρB . As a result, the equality follows by utilising Lemma 4.2.1.
The quantifier QA:C is maximised on pure states, giving the final inequality.
We note that for initial states that are close to ρ = ρAC ⊗ ρB one can utilize the
continuity of the von Neumann entropy [64] and see that IAC:B (0) in Eq. (4.1) is
indeed small. We can also ensure that the initial state is of the form ρ = ρAC ⊗ ρB
by performing a correlation breaking channel on B first. One example of such a
59
channel is a measurement in the computational basis followed by a measurement in
some complementary (say Fourier) basis. This implements the correlation breaking
channel (11AC ⊗ ΛB )(ρABC ) = ρAC ⊗ d11B .2 In this way, our method does not require
any knowledge of the initial state and any operations on the mediator, similar in
spirit to the detection of quantum discord of inaccessible objects in Ref. [45]. We
now move to concrete correlation quantifiers and their correlation capacities.
dC
1 X
|Ψi = √ |aj i|cj i, (4.20)
dC j=1
where |aj i and |cj i form orthonormal bases. One finds sup|ψi IA:C = 2 log2 (dC ).
An interesting quantifier in the context of non-classicality detection is the classi-
cal correlation in a quantum state. It is defined as mutual information of the classical
state obtained by performing the best local von Neumann measurements on the orig-
inal state ρ [66], i.e., CX:Y = supΠX ⊗ΠY IX:Y (ΠX ⊗ ΠY (ρ)), where ΠX ⊗ ΠY (ρ) =
P
xy |xyi hxy| ρ |xyi hxy|, and |xi, |yi form orthonormal bases. The supremum of
mutual information over classical states of AC is log2 (dC ).
Quantum discord is a form of purely quantum correlations that contain quantum
entanglement. It can be phrased as a distance-based measure. In particular, we
consider the relative entropy of discord [38], also known as the one-way deficit [39].
It is an asymmetric quantity defined as DX|Y = inf ΠY S(ΠY (ρ)) − S(ρ), where ΠY is
a von Neumann measurement conducted on subsystem Y . The relative entropy of
discord is maximized by the state (4.20), for which we have sup|ψi DA|C = log2 (dC ).
Our last example is negativity, a computable entanglement monotone [31–34].
For a bipartite system negativity is defined as NX:Y = (||ρTX ||1 − 1)/2, where ||.||1
2
In this context one might ask whether there exists a channel such that (11AC ⊗ ΛB )(ρABC ) =
ρAC ⊗ ρB , where ρAC and ρB are reduced density matrices. Such a channel does not exist, e.g., it
would be nonlinear. See also Ref. [65].
60
denotes the trace norm and ρTX is a matrix obtained by partial transposition of ρ
with respect to X. Negativity is maximized by the state (4.20), and the supremum
reads sup|ψi NA:C = (dC − 1)/2.
Clearly, many other correlation quantifiers are suitable for our detection method
because the assumptions behind Eqs. (4.1) and (4.17) are not demanding. For
those presented here, see Table 4.1 for a summary. In fact, one may wonder which
correlations do not qualify for our method. A concrete example is the geometric
quantum discord based on p-Schatten norms with p > 1, as it may increase under
local operations on BC [67, 68].
i X
ρ̇ = − [HAC + HBC , ρ] + LX ρ, (4.21)
~ X=A,B,C
X 1
LX ρ = QkX ρQk† k† k
X − {QX QX , ρ},
k
2
where the last term in (4.21) is the incoherent part of the evolution and LX de-
scribes interactions of system X with its local environment, i.e., the operators
QkX act on system X only. We denote LAC = −(i/~)[HAC , ·] + LA + LC and
LBC = −(i/~)[HBC , ·] + LB . One readily verifies that if [HAC , HBC ] = 0 and
[LC , HBC /~] = 0, we have commuting Lindblad operators [LAC , LBC ] = 0. Note
that, if one includes LC in LBC instead, the second condition for commuting Lind-
blad operators now reads [LC , HAC /~] = 0. For commuting Lindbladians, the cor-
responding evolution decomposes as ΛBC ΛAC , or in reverse order. Therefore, our
bounds apply accordingly. Their violation implies that either the Hamiltonians do
61
not commute or the operators describing dissipative channels on C do not commute
with HAC and HBC . In particular, if C is kept isolated so that its noise can be
ignored, the violation of our bounds is solely the result of the non-commutativity of
the interaction Hamiltonians.
4.6 Summary
One of the most elementary non-classical traits of quantum observables is their
non-commutativity. In this chapter, we have linked non-commutativity of interac-
tion Hamiltonians (more generally, non-decomposability of time evolution) to the
amount of correlations that can be created in the associated dynamics. This led us
to a method for detection of non-decomposability of evolution in a scenario where a
mediator C mediates interactions between two probing objects A and B (all these
objects can interact with their local environments). The Hamiltonians or Lindblad
operators can remain unknown throughout the assessment, we only require knowl-
edge of the dimension of the mediator. Furthermore, no operation on C is necessary
at any time, which makes this strategy experimentally friendly. Technically, under
the assumption of decomposable evolution, we derived upper bounds on correlations
62
between the probes. Non-decomposability is detected by observing violation of cer-
tain bounds on AB correlations. A plethora of correlation quantifiers are helpful
in our method, e.g., quantum entanglement, discord, mutual information, and even
classical correlation. We also provided an intuitive explanation in terms of multiple
exchanges of a virtual particle which lead to the possible excessive accumulation of
correlations.
63
Chapter 5
This chapter investigates the rate at which quantum entanglement is generated. I will
begin with a motivation comprising the well-known quantum speed limit and general
scenarios that will be considered here. Entangling speed limit for the direct inter-
action setting will be examined. The following section then formulates the limit for
its counterpart, the indirect interaction setting, with exemplary dynamics saturating
the speed. Next, I will revisit the entanglement localisation introduced in Chapter 3
and provide an upper bound for entanglement gain in terms of initial correlations
present in the initial state. I will also present the implications of the results and a
simple application to charging quantum batteries.
65
5.1 Motivation and objectives
A quantum state requires finite time to evolve into a state distinguishable from its
initial form. This limitation on the shortest time is widely known as the quantum
speed limit (QSL). The lower bound on time required for the evolution was first
obtained by Mandelstam and Tamm [69]. Since then, advances for the time bound
include unitary evolution of pure states [70–72] and for mixed states [73–76]. These
results have many applications, e.g., to the investigation of the rate of entropy
increase [77], the limit in quantum metrology [78], the limit on computation [79,
80], and the bound on charging power of quantum batteries [81–83]. Furthermore,
previous studies have also shown the application of QSL to classical domain [84, 85].
We will use the time bound that is given by the unified quantum speed limit
[86, 87], which, for an evolution of a state ρ to another state σ, reads
Θ(ρ, σ)
τ (ρ, σ) = ~ , (5.1)
min{hHi, ∆H}
p√ √
where Θ(ρ, σ) = arccos(F(ρ, σ)) is the Bures angle, F(ρ, σ) = tr ρσ ρ is the
Uhlmann root fidelity [26, 27], hHi = tr(Hρ)−Eg is the average energy with respect
p
to the ground state energy of H, and ∆H = tr[H 2 ρ] − tr[Hρ]2 is the standard
deviation of energy (SDE). Note that for pure states, the Bures angle is given by
the Fubini-Study distance, i.e., Θ(|ψi hψ| , |φi hφ|) = arccos |hψ|φi| [88–90].
In this chapter, we study the quantum speed of system AB with direct and
indirect (through a mediator) continuous interactions, see Fig. 1.1 and Fig. 1.2b.
The direct interaction case considers the objects A and B with Hamiltonian HAB
(local Hamiltonians HA and HB included). On the other hand, the interactions
between A and B are mediated by an object C, such that the total Hamiltonian is
of the form HAC + HBC (again, local terms HA , HB , and HC included). During the
evolution, we study properties of the principle objects A and B, such as quantum
entanglement between them and charging power. The latter has been defined as
the speed to evolve the quantum state of AB from |00i to |11i (fully charged).
We compare the entanglement creation and the charging power between the direct
and indirect interaction scenarios in terms of speed. In particular, we investigate
whether one setup has an advantage over the other and the factors that are relevant
for optimal speed.
5.2 Preliminaries
It can be seen from the QSL of Eq. (5.1) that the first relevant quantity is the
fidelity between the initial state ρAB (0) and the target state ρAB (t). The second
66
quantity is min{hHi, ∆H}, which involves the average energy or SDE. It is clear
that the change in Hamiltonian H → kH (e.g., having more energy) results in
hHi → khHi and ∆H → k∆H for a constant k. This means one can always speed
up the evolution of the quantum state by, for example, providing a higher amount
of energy. Therefore, to put all processes on equal footing, we fix min{hHi, ∆H} to
be ~Ω. For ease of computation, let us rewrite Eq. (5.1) as
Θ(ρ, σ)
Γ(ρ, σ) = , (5.2)
min{hM i, ∆M }
where |xj i and |yj i form orthonormal bases. We will also use the relative entropy
of entanglement (REE) [91], denoted as EX:Y . The maximum of REE is given by
log2 (d). We consider the relative entropy of discord [38], which is also referred to as
the one-way quantum deficit [39]. It is an asymmetric quantity denoted as ∆X|Y .
Similar to REE, maximum discord is given by log2 (d). To quantify the amount
of total correlation, we use the mutual information IX:Y [43]. Finally, on some
occasions, we use conditional entropy SX|Y .
67
where σ x is the x-Pauli matrix and Ω denotes the frequency unit. It is easy to see
that the state at time t takes the form |ψ(t)i = cos(Ωt) |00i − i sin(Ωt) |11i, and
therefore the state of AB is oscillating between the disentangled states |00i or |11i
√
and a maximally entangled state (|00i − i |11i)/ 2.
We note that in small time, the component |11i grows ∝ ∆t, which means that
entanglement, as quantified by negativity, NA:B increases immediately. In other
words, the rate of entanglement follows ṄA:B (0) > 0. It is also easy to confirm that
min{hM i, ∆M } = 1 in this example, meaning that the time bound is simply given by
the Bures angle, i.e., Γ(ρAB (0), ρAB (T )) = Θ(ρAB (0), ρAB (T )).2 The corresponding
dynamics of entanglement and Bures angle is illustrated in Fig. 5.1. One can see
that this dynamics saturates the time bound, i.e., the time equals the time bound
Θ (dashed line in Fig. 5.1).
Figure 5.1: Exemplary dynamics showing maximum entangling speed between two
qubits. The Hamiltonian is given by HAB = ~Ω σAx ⊗ σBx with initial state |00i.
Maximum negativity (solid curve) is achieved in t = π/4Ω.
From the simple example above, it is apparent to see the generalisation of the
bound for the case where the objects under consideration are d-dimensional. In
particular, by having the resource equality, the direct bound reads
1
Γdi = arccos |h00|Ψmax i| = arccos √ . (5.5)
d
For the actual dynamics, one has to figure out the Hamiltonian that both gives the
resource equality and evolves the initial state to a state with given fidelity in the
shortest possible time.
2
We use T = Ωt as a dimensionless time variable.
68
5.4 Entangling speed limit: Indirect interactions
It is now reasonable to ask whether the dynamics in Fig. 5.1 can be sped up or at
least replicated through indirect interactions. Although it is definitely possible to
have entanglement growth from zero to maximum, we will show that it is impossible
to beat the speed set by the direct interaction scenario with initial state |00i for
system AB.
Let us begin with the following theorem, which investigates the initial rate of
entanglement gain in the partition A : B.
Theorem 5.4.1. Consider the indirect interaction scenario, where an ancilla C is
mediating interactions between objects A and B. We allow all the objects to be open
to their own local environments. If the initial state of the tripartite system is of the
form ρ0 = ρAB ⊗ρC , then the rate of any entanglement monotone reads ĖA:B (0) ≤ 0.
Proof. Let us start with the evolution of the tripartite state following the Lindblad
form
ρ∆t − ρ0 i X
= − [H, ρ0 ] + LX ρ0 , (5.6)
∆t ~ X=A,B,C
X 1 k† k
LX ρ 0 = QkX ρ0 Qk†
X − {QX QX , ρ0 },
k
2
where the first term in (5.6) is the coherent part of the evolution, whereas the second
term is the incoherent part and LX describes interactions of system X with its local
environment, i.e., the operators QkX act on system X only. Without loss of generality
we assume the Hamiltonians of the form H = HA ⊗ HC1 + HB ⊗ HC2 , where HC2 can
be different from HC1 . Note that the generalisation of the proof to Hamiltonians
H = µ HAµ ⊗ HCµ1 + ν HBν ⊗ HCν 2 is straightforward.
P P
where we have used ρ0 = ρAB ⊗ ρC , and EC1 = tr(HC1 ρC ) and EC2 = tr(HC2 ρC )
are the initial mean energies.3 We have also utilised the cylic property of trace such
that trC (QkC ρC Qk† 1 k† k
C − 2 {QC QC , ρC }) = 0.
3
We will assume that these constants are dimensionless and the energy unit is transferred
entirely to the terms HA and HB , respectively.
69
One can see from Eq. (5.7) that the state of AB at time ∆t is simply as a
result of effective local Hamiltonians HA EC1 + HB EC2 and local interactions with
environments. For any entanglement monotone, i.e., non-increasing under local
operations and classical communication (LOCC), we have EA:B (∆t) ≤ EA:B (0) and
therefore ĖA:B (0) ≤ 0.
For example, if we consider unitary dynamics and the initial pure state |00i
of AB, it requires C to be in a decoupled form, i.e., the whole system assumes
|00i h00| ⊗ ρC . From Eq. (5.7) of Theorem 5.4.1, it is easy to see that EA:B (∆t) =
EA:B (0), i.e., ĖA:B (0) = 0. This is due to the fact that unitary operations resulting
from local Hamiltonians leave the entanglement invariant. This implies that, in
order to have positive (or negative) initial entanglement rate, there has to be initial
correlation between the mediator C and the principal objects A and B.
Note also that if C is uncorrelated with AB at all times, i.e., IAB:C (t) = 0, one
can apply the method in Theorem 5.4.1 successively and conclude that entanglement
distribution is impossible as EA:B (t) ≤ EA:B (0). This has been proven similarly in
Chapter 3. This way, entanglement gain between the principal objects can be used
as a witness of correlation with the mediator C, i.e., IAB:C > 0 at some time during
the evolution.
Theorem 5.4.1 is not enough to conclude the impossibility of beating the direct
bound as it does not explain what is in play during the dynamics. It might happen
that although the initial rate of production is non-positive, the rate during the dy-
namics is high such that after some finite time T < Γdi the maximum entanglement
is reached. We will show that this scheme is impossible by utilising the following
ultimate bound.
Theorem 5.4.2. Consider any coherent dynamics of three objects A, B, and C with
the total Hamiltonian given by H. This may include the direct interaction scenario
between A and B as well as the indirect interaction where C is the mediator. Starting
with initial density matrix of the form ρ(0) = ρABC , where the reduced state ρAB is
separable, we show that the time it takes to maximally entangle A and B is lower
bounded as
1
T ≥ arccos √ , (5.8)
d
where T = Ωt and min{hHi, ∆H} = ~Ω.
Proof. First let us note that maximum entanglement between A and B implies
that the state of AB is pure, and that the state of C is decoupled, i.e., ρ(T ) =
70
|Ψmax i hΨmax | ⊗ ρ0C . Taking the fidelity of the initial and final states gives
Let us now discuss a special case of Theorem 5.4.2 above. In particular, consider
the initially decoupled mediator, i.e., ρ(0) = ρAB ⊗ ρC . Consequently, we have
q
√ √
F(ρ(0), ρ(T )) = tr ρAB ρC |Ψmax i hΨmax | ρ0C ρAB ρC
q
√ √ √ 0 √
= tr ρAB |Ψmax i hΨmax | ρAB ρC ρC ρC
q q
√ √ √ 0 √
= tr ρAB |Ψmax i hΨmax | ρAB tr ρC ρC ρC
Note that as the fidelity 0 ≤ F(ρC , ρ0C ) ≤ 1, we recover the second line of Eq. (5.9).
Now let us discuss the consequence of the initial state ρ(0) = ρAB ⊗ ρC . From
Eq. (5.11), we notice that if ρAB is strictly mixed, then for a unitary dynamics
(purity-preserving), ρ0C will be more mixed than ρC , i.e., ρC 6= ρ0C . This means
F(ρC , ρ0C ) < 1, making the corresponding first inequality in (5.9) strict, and resulting
in a strict time bound T > Γdi . Therefore, in order to saturate the direct bound,
the state ρAB has to be pure and F(ρC , ρ0C ) = 1. A trivial dynamics (via direct
interactions in a tripartite setting) is given by the example presented in Section 5.3
with the addition of a decoupled mediator ρC and a local Hamiltonian HC .
71
5.4.1 Saturating the limit
Now we present examples of indirect interactions that saturate the direct time
bound. As initial states of the form ρAB ⊗ ρC cannot be used for maximum entan-
glement speed, one would normally think of utilising entanglement in the partition
AB : C. In particular, consider the following initial state and Hamiltonian:
1
|ψ(0)i = √ (|000i + |111i),
2
~Ω z
H = √ (σA ⊗ HC1 + σBz ⊗ HC2 ), (5.12)
2 2
where HC1 = −(1 + σCx + σCy + σCz ) and HC2 = 1 − σCx − σCy + σCz . One finds that this
example has min{hM i, ∆M } = 1 and therefore the bound is also given by the Bures
angle Θ(ρAB (0), ρAB (T )), where now ρAB (T ) = trC (ρABC (T )). Furthermore, the
resulting dynamics is the same as that in Fig. 5.1. From this example we note that
the initial state has some purely quantum properties: entanglement NAB:C (0) = 0.5
and mutual information IAB:C (0) = 2.
However, is it necessary to have entanglement with the mediator C to saturate
the entangling speed of direct interactions? We now show that this is not the case.
In particular, it is enough to have classical correlation with C initially. For this
purpose, we take the initial state and Hamiltonian
1 1
ρ(0) = |ψ+ i hψ+ | ⊗ |+i h+| + |φ+ i hφ+ | ⊗ |−i h−| ,
2 2
~Ω x
H = (σ ⊗ σCx + σBx ⊗ σCx ), (5.13)
2 A
√ √
where σ x |±i = ± |±i, and |ψ+ i = (1/ 2)(|01i+|10i) and |φ+ i = (1/ 2)(|00i+|11i)
are two Bell states of party AB.4 One can see that the initial state in Eq. (5.13)
is disentangled in the partition AB : C and is correlated with mutual information
IAB:C (0) = 1. This example also has the resource equality satisfied and the dynamics
of that in Fig. 5.1. Furthermore, the dynamics from Eq. (5.13) has been shown to
have zero quantum discord DAB|C at all times (see Chapter 3), showing that the
correlation between C and AB is purely classical. One can extend further and see
that initial states of the form p |ψ+ i hψ+ | ⊗ |+i h+| + (1 − p) |φ+ i hφ+ | ⊗ |−i h−|,
where p stands for probability, can be made maximally entangled by Hamiltonian
in Eq. (5.13). This maximum entanglement is achieved in T = π/4 for 0 < p < 1,
see Fig. 5.2. Independent of p, the Bures angle dynamics is given by the dashed line
in Fig. 5.1, which indicates saturation.
Let us compare the three examples shown above. First of all, it is clear that
4
Note that this example is similar to that already presented in Chapter 3 to demonstrate
entanglement localisation.
72
Figure 5.2: Maximum entangling speed for various initial states. The Hamiltonian
is taken to be H = ~Ω2
(σAx ⊗ σCx + σBx ⊗ σCx ) with initial state p |ψ+ i hψ+ | ⊗ |+i h+| +
(1−p) |φ+ i hφ+ |⊗|−i h−|. Entanglement is plotted for p = 0.1 (red curve), 0.3 (blue
curve), and 0.5 (black curve).
starting with pure states, e.g., |00i, it is possible to get maximum entanglement via
direct interactions. Also note that during the evolution, the purity of the state is
preserved. This is because unitary operations do not change purity. Therefore, if
one were to start with a mixed initial state ρAB , the maximum entanglement can
never be achieved. This is true as maximum entanglement is given by pure states (as
measured by any entanglement monotone) [62]. In the indirect interaction scenario,
the evolution of the state of AB is not unitary. Therefore, the dynamics can change
the purity of the state ρAB . For example, in the last two examples, ρAB changes from
having zero entanglement with purity 0.5 to maximum entanglment with purity 1.
Hence, for mixed initial states, one would want to utilise mediators to distribute
maximum entanglement between the principal objects A and B.
Finally, one might wonder, for an initial mixed state, is it possible to achieve
maximum entanglement with direct interaction setup, this time allowing the prin-
cipal objects to be open to their local environments? This question is sensible as
interactions with environments can change the purity of the principal objects, and
therefore allowing them to reach maximum entanglement. To illustrate this, let us
assume a scenario where one of the principal objects, say A, is interacting with a
single-qubit environment C. Furthermore, take the initial state of Eq. (5.13) and
Hamiltonian of the form H = ~Ω σAx ⊗ σCx . One can confirm that the resulting A : B
entanglement dynamics is identical as that in Fig. 5.1.
73
5.5 Initial correlation with mediators bounds lo-
calised entanglement
Note that the example in (5.13), the entanglement localisation, realises maximum
entanglement with dynamics saturating the direct bound by having initial state with
only classical correlation in the partition AB : C. It is then natural to ask what
bounds the entanglement gain. In particular, whether the gain is related to the
initial mutual information IAB:C (0). In order to answer this question, let us begin
by presenting the following Lemma.
Lemma 5.5.1. For a three-party system ABC, where the subsystem C is only clas-
P
sically correlated, i.e., ρ = c pc ρAB|c ⊗ |ci hc| with {|ci} forming an orthonormal
basis, the relative entropy of entanglement follows the bound
Proof. By definition of the REE we have EA:BC (ρ) = −tr(ρ log2 σ) − S(ρ), where σ
is the closest separable state to ρ [91]. According to the flags condition EA:BC =
P
c pc EA:B (ρAB|c ), where EA:B (ρAB|c ) = −tr(ρAB|c log2 σAB|c ) − S(ρAB|c ) [40]. Also,
P
it has been shown that σ = c pc σAB|c ⊗ |ci hc|, where σAB|c is a separable state
closest to ρAB|c [40]. Next, we note for the first term in the definition of EA:BC :
!
X X
−tr pc ρAB|c ⊗ |ci hc| log2 ( pj σAB|j ⊗ |ji hj|)
c j
!
X X
= −tr pc ρAB|c ⊗ |ci hc| log2 (pj σAB|j ) ⊗ |ji hj|
c j
!
X
= −tr pc ρAB|c log2 (pc σAB|c )
c
X
= − pc (log2 (pc ) + tr( ρAB|c log2 (σAB|c )))
c
X
= SC + pc (−tr( ρAB|c log2 (σAB|c ))), (5.15)
c
where each term in the sum is minimised as σAB|c is closest to ρAB|c . On the other
P
hand, we have EA:B = −tr( c pc ρAB|c log2 σAB ) − SAB , where σAB is the closest
P
separable state to c pc ρAB|c . Putting things together we have EA:BC − EA:B =
74
IAB:C + η, where IAB:C = SAB + SC − SABC and
X
η = pc [−tr( ρAB|c log2 (σAB|c ))]
c
X
− pc [−tr( ρAB|c log2 (σAB ))]. (5.16)
c
One can see that η is negative, since −tr( ρAB|c log2 (σAB|c )) is minimised, confirming
the Lemma.
Now, for the dynamics, let us consider the indirect interaction scenario where all
our objects are allowed to be open to their own local environment. The entanglement
bound is proven in the following Theorem.
Theorem 5.5.2. Consider the indirect interaction scenario where all the objects A,
B, and the mediator C are open to their local environment. If C is classical at all
times, i.e., DAB|C (t) = 0 for t ∈ [0, τ ], we have
where the steps are justified as follows. The inequality (5.18) is due to the mono-
tonicity of entanglement under local operations (tracing out the mediator C), i.e.,
EA:B (τ ) ≤ EA:BC (τ ). Line (5.19) follows from Theorem 3.3.1, stating that entan-
glement in the partition A : BC cannot grow via classical C, that is EA:BC (τ ) ≤
EA:BC (0). Finally, Lemma 5.5.1 confirms line (5.20).
The bound in Eq. (5.17) can be made simpler at the expense of tightness. Let us
note that for states separable in the partition AB : C (including states where C is
P
classical, i.e., ρ = c pc ρAB|c ⊗ |ci hc|), the conditional entropy {SAB|C , SC|AB } ≥ 0.
This means that the mutual information can be bounded as IAB:C = SAB − SAB|C ≤
SAB or IAB:C = SC − SC|AB ≤ SC . Therefore, entanglement gain via classical
mediators cannot be larger than log2 (dC ) as it is the maximum entropy for system
C. We would also like to mention that if one observes EA:B (t) − EA:B (0) > SAB (0)
then there had to be non-zero quantum discord DAB|C during the evolution. This
is a witness of non-classicality of C by observing only the objects A and B, similar
to the detection of quantum discord in Chapter 3.
75
As mentioned in Preliminaries, the evolution of a quantum state can be sped up
simply by providing more energy to the system. From the results presented in this
chapter, one can also speed up the creation of entanglement between two objects
via a mediating system that is initially correlated with them. Indeed, for initially
uncorrelated systems, it would take longer time to reach maximum entanglement.
In a way, this puts the two quantities, energy and correlations, on the same footing.
Both can be seen as a resource for fast distribution of quantum entanglement.
|ψ(0)i = |0i⊗N ,
H = ~Ω (σ1x ⊗ σ2x ⊗ · · · ⊗ σN
x
). (5.21)
76
the Hint . On the other hand, free Hamiltonians of the form HA + HB would evolve
the subsystems separately, such that the state at time ∆t reads |ψAB (∆t)i = (1 −
i∆tHA /~) |0i⊗(1−i∆tHB /~) |0i. One can easily see that, in this case, ΞAB (∆t) = 0.
From these observations, we note that positive charging rate is linked with the
presence of entanglement between objects A and B. This is because entanglement
is nonzero for the state |ψAB (∆t)i with |11i component.
Figure 5.3: Dynamics of charge Ξ (solid curves) and Bures angle (dashed lines) for
interacting (black) and non-interacting (blue) qubits. In both cases, the resource
has been fixed to the same amount, i.e., min{hM i, ∆M } = 1.
For indirect interactions, with system C as the mediator, the initial state is of the
form |00i h00| ⊗ ρC . We show with the following Lemma that an initially correlated
mediator is also a factor for having positive charging rate. This is similar to the
entanglement distribution task.
Lemma 5.6.1. For indirect interactions, where all objects can be open to their local
environment, the charging rate follows Ξ̇AB (0) = 0.
Proof. First, let us express the charge as
p
ΞAB (t) = F(|11i h11| , ρAB (t)) = h11|ρAB (t)|11i. (5.22)
∆t
ρAB (∆t) = |00i h00| − i [HA EC + HB Eγ , |00i h00|]
~
+∆t(LA + LB ) |00i h00| , (5.23)
where we have used ρ0 = |00i h00| ⊗ ρC . One can see that ΞAB (∆t) = 0 from the
observation that h11|ρAB (∆t)|11i = 0, which proves the Lemma.
77
5.7 Summary
Quantum speed limit sets the minimum time required for evolving a quantum state
and therefore also its properties. In this chapter we have presented the time bound
for the aim of distributing quantum entanglement between two objects both via
direct as well as indirect interactions. We have shown that the indirect interac-
tion setting cannot beat the direct interaction setting in terms of entangling speed.
Furthermore, the correlation between the mediator and the principal objects is re-
quired for optimal distribution. From the perspective of speeding up the creation
of correlations, we discussed that both energy and correlated mediators play similar
roles. We also presented briefly a simple application of the quantum speed limit
to charging of quantum batteries and compared the direct and indirect interaction
settings.
78
Chapter 6
1
Parts of this chapter are reproduced from our published article of Ref. [93],
which is licensed under the Creative Commons Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/). Where applicable, changes made will be indi-
cated.
79
6.1 Motivation and objectives
The successful unification of electromagnetic, weak, and strong interactions within
the quantum framework strongly suggests that gravity should also be quantised.
Up to date, however, there is no experimental evidence of quantum features of grav-
ity. In numerous experiments gravity is key to the interpretation of the observed
data, but it is often sufficient to use Newtonian theory (quantum particle moving
in a background classical field) or general relativity (quantum particle moving in
a fixed spacetime) to gather a meaningful understanding of such data. Milestone
experiments described within Newtonian framework include gravity-induced quan-
tum phase shift in a vertical neutron interferometer [94], precise measurement of
gravitational acceleration by dropping atoms [95], or quantum bound states of neu-
trons in a confining potential created by the gravitational field and a horizontal
mirror [96]. Quantum experiments that require general relativity include gravita-
tional redshift of electromagnetic radiation [97] or time dilation of atomic clocks at
different heights [98].
A number of theoretical proposals discussed scenarios capable of revealing quan-
tumness of gravity. For example Refs. [99–107] proposed the observation of a probe
mass interacting with the gravitational field generated by another mass. More recent
proposals put gravity in a role of mediator of quantum correlations and are based
on the fact that quantum entanglement between otherwise non-interacting objects
can only increase via a quantum mediator [45, 108, 109], see Chapter 3. Motivated
by these proposals and by advances in optomechanics [110], in particular the cooling
of massive mechanical (macroscopic) oscillators close to their quantum ground state
[111–113] and the measurement of quantum entanglement of a two-mode system
[114–116], we study two nearby cooled masses interacting gravitationally.
We propose two scenarios capable of increasing gravitational entanglement be-
tween the masses. In the first scenario, we consider the masses trapped at all times
in 1D harmonic potentials (optomechanics). In the second one, the masses are re-
leased from the optical traps. For both settings, we derive an analytic figure of merit
characterising the amount of gravitationally induced entanglement and the time it
takes to observe it. The derivation includes various initial states and shows that the
objects have to be cooled down very close to their ground states and that squeezing
of their initial state significantly enhances the amount of generated entanglement.
We then formulate a numerical approach, which accounts for all the relevant sources
of noise affecting the settings that we propose, to identify a set of parameters re-
quired for the observation of such entanglement. Finally, we discuss the conclusions
that can be drawn from this experiment with emphasis on the need for indepen-
dent verification that the gravitational interaction between nearby objects is indeed
80
mediated.
p2A 1 2 2 p2B 1
H0 = + mω xA + + mω 2 x2B (6.1)
2m 2 2m 2
and Hg describes the gravitational term. If the harmonic traps are removed the
corresponding Hamiltonian simplifies to H0 = (p2A + p2B )/2m. Before we proceed
with detailed calculations, we shall discuss generic features of the gravitational term
and the conditions required for the creation of entanglement.
Figure 6.1: Thought experiment for the generation of entanglement via gravitational
interactions. Two particles are a distance L apart and trapped with unidimensional
harmonic potentials. It is assumed that both particles are cooled down near their
ground state. We consider two scenarios in which the particles are either trapped
at all times or released from the traps. In the text, we investigate entanglement
dynamics for both scenarios. Note that the generated entanglement can be probed
with weak lights. Apart from the dominant gravitational coupling, our analysis
takes noise, damping, decoherence, and Casimir forces into account.
81
cooled down close to their ground state, one gets
The first term is a rigid energy offset, while the second is a bi-local term and cannot
thus give rise to quantum entanglement. The third term, which is proportional to
(xA − xB )2 , is the first that couples the masses. When written in second quan-
tisation, it becomes apparent that this term includes contributions responsible for
the correlated creation of excitations in both oscillators. In the quantum optics
language, this is commonly referred to as a “two-mode squeezing” operation, which
can in principle entangle the masses provided a sufficient strength of their mutual
coupling.
82
is reached in a time tmax
sq = π/2ηg ω; (iii) Weaker entanglement is generated with
increasing temperature of the masses or coupling to the environment.
We will begin by constructing Langevin equations of the masses in Heisenberg
picture. As we will be dealing with Gaussian states, we provide the dynamics of the
covariance matrix, which completely describes the state of the whole system. From
this dynamics, we will calculate important quantities such as quantum entanglement
between the masses, both analytically for special cases and numerically for general
cases.
Ẋj = ω Pj (j = A, B),
ṖA = −ω (1 − ηg ) XA − ωηg XB − γ PA + ξA + ν, (6.4)
ṖB = −ω (1 − ηg ) XB − ωηg XA − γ PB + ξB − ν,
√
where we have introduced the constant frequency ν = Gm2 / ~mωL4 and the di-
p √
mensionless quadratures Xj = mω/~ xj and Pj = pj / ~mω. These equations
incorporate Brownian-like noise – described by the noise operators ξj – and damp-
ing (at rate γ) affecting the dynamics of the mechanical oscillators, due to their
interactions with their respective environment. We assume the (high mechanical
quality) conditions Q = ω/γ 1, as it is the case experimentally, so that the
Brownian noise operators can de facto be treated as uncolored noise and we can
write hξj (t)ξj (t0 ) + ξj (t0 )ξj (t)i/2 ' γ(2n̄ + 1)δ(t − t0 ) for j = A, B [118, 119]. Here,
n̄ = (eβ − 1)−1 is the thermal phonon number with β = ~ω/kB T and T the temper-
ature of the environment with which the oscillators are in contact.
The linearity of Eqs. (6.4) and the Gaussian nature of the noise make the theory
of continuous variable Gaussian systems very well suited to the description of the
dynamics and properties of the oscillators under scrutiny. In this respect, the key
tool to use is embodied by the covariance matrix V (t) associated with the state
of the system, whose elements Vij (t) = hui (t)uj (t) + uj (t)ui (t)i/2 − hui (t)ihuj (t)i
encompass the variances and correlations of the elements of the quadrature vector
3
p Note that the ratio between any two consecutive terms in Eq. (6.2) is given by (xA − xB )/L ∼
~/mωL2 . For instance, taking m = 100 µg, ω = 100 kHz, and L = 0.1 mm gives this ratio
∼ 10−12 , and for macroscopic values m = 1 kg, ω = 0.1 Hz, and L = 1 cm the ratio is ∼ 10−15 .
83
u(t) = (XA (t), PA (t), XB (t), PB (t))T . The temporal behaviour of physically rele-
vant quantities for our system of mechanical oscillators can be drawn from V (t)
by making use of the approach for the solution of the dynamics that is illustrated
below.
We split the last term in the matrix equation into two parts, representing the noise
and constant term respectively, i.e., l(t) = υ(t)+κν , where υ(t) = (0, ξA (t), 0, ξB (t))T
√
and the constant κν = ν(0, 1, 0, −1)T with ν = Gm2 / ~mωL4 .
The solution to the Langevin equations is given by
Z t
u(t) = W+ (t)u(0) + W+ (t) dt0 W− (t0 )l(t0 ), (6.6)
0
where W± (t) = exp (±Kt). This allows one to calculate the expectation value of
the ith quadrature hui (t)i numerically, which is given by the ith element of
Z t
W+ (t)hu(0)i + W+ (t) dt0 W− (t0 )κν , (6.7)
0
where we have used the fact that the noises have zero mean, i.e., hυi (t)i = 0 and
that hκν i = tr(κν ρ) = κν . From Eq. (6.6), one can also calculate other important
quantities via the covariance matrix as shown below.
Covariance matrix of our system is defined as Vij (t) ≡ h{∆ui (t), ∆uj (t)}i/2 =
hui (t)uj (t)+uj (t)ui (t)i/2−hui (t)ihuj (t)i where we have used ∆ui (t) = ui (t)−hui (t)i.
This means that κν does not contribute to ∆ui (t) (and hence the covariance matrix)
since hκν i = κν . We can then construct the covariance matrix at time t from Eq.
84
(6.6) without considering κν as follows
Vij (t) = hui (t)uj (t) + uj (t)ui (t)i/2 − hui (t)ihuj (t)i
V (t) = W+ (t)V (0)W+T (t)
Z t
+W+ (t) dt0 W− (t0 )DW−T (t0 ) W+T (t), (6.8)
0
where D = Diag[0, γ(2n̄ + 1), 0, γ(2n̄ + 1)] and we have assumed that the initial
quadratures are not correlated with the noise quadratures such that the mean values
of the cross terms are zero. A more explicit solution of the covariance matrix, after
integration in Eq. (6.8), is given by
85
the minimum symplectic eigenvalue ν̃min < 1/2. The explicit expression is given by
√ √
ν̃min = (Σ − Σ2 − 4 detV )1/2 / 2, where Σ = detIA + detIB − 2 detL. Entangle-
ment between mode A and mode B is then quantified by logarithmic negativity as
E = max 0, − log2 (2ν̃min ) [34, 35]. Note that the separability condition, when
V (t)TB has ν̃min ≥ 1/2, is sufficient and necessary for two-mode systems [120].
Figure 6.2: Entanglement dynamics with initial ground state for each mass. The
figure of merit is varied as ηg = 0.01 (red curve), 0.05 (blue curve), and 0.1 (black
curve). The maximum entanglement approximately follows ηg / ln 2.
4
We use ηg 1 for all analytical derivations in this chapter.
86
Squeezed thermal initial state
An analytic solution is also possible for the case of mechanical systems initially pre-
pared in squeezed thermal states, a situation that can be arranged by suitable optical
driving [121, 122]. Each mass is prepared in a state Sρth S † , where ρth is a thermal
state and S = exp (−i s(X 2 − P 2 )/2) is the squeezing operator with strength s
(assumed to be real). This operator corresponds to anti-squeezing (squeezing) the
position quadrature for s > 0 (s < 0). By writing individual-oscillator squeezing as
sj and assuming sj ηg , the entanglement is again observed to oscillate, but with
amplitude |sA + sB |/ ln 2 − log2 (2n̄ + 1). We present the entanglement dynamics
in Fig. 6.3 for varying values of the squeezing strength. Note that it is irrelevant
whether quadratures of both masses are squeezed or anti-squeezed. We provide ex-
planation in the Details of entanglement dynamics (Section 6.3.5). Therefore, only
the degree of pre-available single-oscillator squeezing and the environmental tem-
perature set a limit to the amount of entanglement that can be generated between
the mechanical systems through the gravitational interaction. In the low temper-
max
ature limit, where Esq ≈ |sA + sB |/ ln 2, which is in principle arbitrarily larger
than the case without squeezing, a time tmaxsq = π/(2ηg ω) tth
max
would be required
for such entanglement to accumulate. Needless to say, long accumulation times are
far from the possibilities offered by state-of-the-art optomechamical experiments,
which prompts an assessment that includes ab initio the effects of environmental
interactions.
Figure 6.3: Entanglement dynamics with squeezed initial ground state for each
mass. The figure of merit is taken as ηg = 10−4 and the squeezing strength is varied
as sA,B = 0.1 (red curve), 0.5 (blue curve), and 1 (black curve). The maximum
entanglement approximately follows 2sA,B / ln 2.
87
6.3.4 Noisy dynamics: Numerical simulation
In the case of noisy dynamics, however, an analytical solution is no longer available
and we have to resort to a numerical analysis. Let us therefore consider the figure
of merit ηg in order to set the parameters for numerical investigation. We consider
two oscillators of spherical shape with uniform density ρ and radius R, which are
separated by a distance L = 2.1R. This might be a situation matching current
experiments in levitated optomechanics [123, 124], which are rapidly evolving to-
wards the possibility of trapping multiple dielectric nano-spheres in common optical
traps and controlling their relative positions [125]. However, low-frequency oscilla-
tors, which are favourable for the figure of merit and typically associated with large
masses, are unsuited to such platforms and would require a different arrangement,
such as LIGO-like ones [111].
In terms of the density ρ, we have ηg = 8πGρ/3(2.1)3 ω 2 , which does not depend
on the dimensions of the oscillators nor their mass. As the density of materials
currently available for such experiments varies within a range of only two orders of
magnitude, the linear dependence on ρ sets a considerable restriction on the values
that ηg can take. The densest naturally available material is Osmium, which has ρ =
22.59 g/cm3 and, in order to provide an upper bound to the generated entanglement
that would be attainable using other materials, we shall use this density in our
numerical simulations. Accordingly, ηg = 1.36 × 10−6 /ω 2 , where ω is in Hz.
Fig. 6.4 shows exemplary entanglement dynamics for different values of the ther-
mal phonon number n̄ and mechanical quality factor Q.5 The frequency has been
5
The oscillations of entanglement for unsqueezed initial state are still present in this dynamics,
88
fixed to ω = 0.1 Hz (cf. Section 6.8). As expected, higher damping (lower Q) results
in the decay of entanglement, and the higher the temperature of the mirror (higher
n̄) the higher the mechanical quality factor needed to maintain entanglement. The
setup allows for high entanglement, even with low coupling strength ηg ∼ 10−4 .
However, this comes at the expense of the time for which the dynamics of the oscil-
lators should be kept coherent. In the same figure one can see the time required for
accumulation of entanglement. It shows that cooling down the masses close to their
ground state, n̄ ≈ 0, is crucial for the reduction of the required coherence time.
Figure 6.5: Exemplary dynamics of entanglement and width of two coupled oscil-
lators. Each oscillator is made of Osmium and has a spherical shape with mass
m = 1 kg. The equilibrium distance is L = 2.1R and the frequency of the trapping
potential is ω = 0.1 Hz. (a) Short dynamics within 100 s for squeezing parameters
sA,B = ±1.73. (b) Higher gain of entanglement requires longer time. Note that
the dynamics in (b) is approximately the same for positive and negative squeezing
parameters.
Fig. 6.5a shows the entanglement dynamics with both sA,B = 1.73 and −1.73,
corresponding to initial states of the masses that are both anti-squeezed or squeezed,
respectively, in position quadrature. During the evolution, the width of each mass
oscillates as illustrated in the inset of Fig. 6.5a. It is clear that the oscillation of the
width matches the oscillation of entanglement, for both squeezing parameters. For
longer time, i.e., t ∼ 104 s, the oscillation in width leads to an accumulation of high
entanglement as can be seen in Fig. 6.5b. Note that the oscillation of the position
showing repeating pattern with a period of π/[(1 − ηg )ω] ≈ 31 s.
89
variance is very rapid on this timescale and the envelope of these oscillations is the
same for both positive and negative s, see the inset of Fig. 6.5b. As a result, the
entanglement dynamics is approximately equal for squeezed and anti-squeezed cases.
This confirms our analytical result regarding the maximum entanglement gain being
|sA + sB |/ ln 2.
90
6.4.2 Noiseless dynamics: Analytical solution
One can obtain the covariance matrix V (t) from Eqs. (6.12) and consequently derive
the entanglement dynamics using the approach described above. After imposing the
√
limits ηg 1 and ηg ωt 1, which apply in typical experimental situations, one
obtains the analytical expression for the entanglement dynamics as follows
Eth (t) = max 0, Egnd (t) − log2 (2n̄ + 1) , (6.14)
q p
Egnd (t) = − log2 1 + 2σ(t) − 2 σ(t)2 + σ(t) ,
where Egnd (t) is the entanglement with initial ground state for each mass and σ(t) =
4G2 m2 ω 2 t6 /9L6 . Since entanglement is an increasing function of σ(t), the latter is
a figure of merit for entanglement gain relevant in the case of released masses.
We present exemplary entanglement dynamics in Fig. 6.6 for which entanglement
∼ 10−2 is achieved within seconds. The parameters used here are m = 100 µg,
ω = 100 kHz, and L = 3R. We will show later that with these values gravity is the
dominant interaction and coherence times are much longer than 1 s. Note that this
setup does not require any squeezing.
Figure 6.6: Entanglement dynamics between two released spherical particles. Each
particle with mass 100 µg is initially trapped in 1D harmonic potential (ω =
100 kHz) and having phonon number n̄. This corresponds to the initial state of
each mass being Gaussian thermal state. We take the equilibrium distance to be
L = 3R ≈ 0.1 mm, making gravitational coupling more dominant than the Casimir
interactions. We note that detectable entanglement, E ∼ 0.01, is generated in sec-
onds: 0.8, 4.5, and 7.5 s for n̄ = 0, 1, and 5 respectively. It is shown in the text that
this accumulation is within the coherence times of the particles.
These improvements over the scheme with trapped masses are the result of un-
limited expansion of the wave functions. For example, for initial ground state, the
p √
evolution of the width of each sphere closely follows ∆x(t) ≈ ~/2mω 1 + ω 2 t2 ,
which is an exact solution to a free non-interacting mass. The effect of gravity is
91
stronger attraction of parts of the spatial superposition that are closer, and hence
generation of position and momentum correlations, leading to the growth of quan-
tum entanglement, see inset in Fig. 6.6.
In order to understand the effect of squeezing in this setup, let us suppose, for
simplicity, the squeezing strengths sA,B = s. It is as if one initially prepared each
mass in a Gaussian state with a new initial spread ∆x0 (0) = ∆x(0) exp(s). One can
then calculate the entanglement dynamics using Eq. (6.14) with a new frequency
ω 0 = ω exp(−2s). This means that anti-squeezing the initial position quadrature
(s > 0) would decrease entanglement gain, a situation opposite to the oscillators
setup. This is because a Gaussian state with smaller ∆x(0) spreads faster, such that
during the majority of the evolution, the width is larger than that if one started with
larger ∆x(0). In principle, one obtains higher entanglement gain by squeezing the
position quadrature (s < 0). However, this will result in higher final width, making
it more susceptible to decoherence by environmental particles (see calculations in
Section 6.7).
From numerical simulations, one confirms that, within t = [0, 10]s, the displace-
ments of the two masses follow xA − xB L. Furthermore, the trajectories coin-
cide for both quantum treatment with truncated gravitational energy and classical
treatment with full Hg (see Section 6.4.3). This justifies the approximations used. A
summary of comparison between our two proposed setups can be seen in Table 6.1.
Table 6.1: Summary of parameters for the proposed experimental setups: Oscillators
and released masses. Note that for the dynamics (shaded), we assume initial ground
state for each mass.
Parameters Oscillators Released masses
Mass mA,B 1 kg 100 µg
Radius R 2.2 cm 0.1 mm
Equilibrium distance L 4.6 cm 0.3 mm
Trap frequency ω 0.1 Hz 100 kHz
Squeezing strength sA,B 1.73 0
−17
Average width ∆x 8.3 × 10 m 2.8 × 10−12 m
Entangling time τe 31.7 s 0.8 s
92
and obtain the following equation for xt :
r
2Gm p L π −1
t = xt (L − 2xt ) + √ − tan (θ(xt )) , (6.15)
L 2 2 2
p
where θ(xt ) = (L − 4xt )/ 8xt (L − 2xt ) and t is the time taken for the left mass
to move a distance xt . The trajectory of the right mass is simply −xt . One can
confirm that the classical trajectories indeed coincide with the quantum ones. This
justifies the truncation of Hg in our calculations. Note also that, within 10 s, the
displacement (xA − xB ) ∼ 10−9 m, which is much smaller than the initial distance
between the masses L ≈ 0.3 mm. This validates the use of the limit xA − xB L.
Figure 6.7: Expectation value of the displacement of the masses. Two Osmium
spheres, each with mass 100 µg, are separated by L = 3R. The dynamics shown is
independent of the temperature and the frequency of the initial trapping potential
ω.
93
have used L = 2.1RA . We have also assumed spring-like scaling for the frequency,
p
i.e., ωB = ωA mA /mB . This result is intuitive as one expects stronger gravitational
interaction by making the second sphere larger, i.e., larger α.
More intriguing is the setting in Fig. 6.8b, in which a rod with length d, mass
λB d, and radius RB is interacting with a sphere of mass mA and radius RA . For
simplicity, we assume both objects have a single mechanical frequency ωA and ωB
respectively, and that the rod is thin such that its radius is much smaller than L.
The gravitational interaction reads
p !
d/2 + (d/2)2 + (L0 )2
Hg = −2GmA λB ln , (6.16)
L0
and we define ς = 2L/d. We have also taken L = 1.1RA , RB = 0.1RA , and the
same spring-like scaling as in the two-sphere configuration. Although the strength
of the gravitational energy (6.16) increases monotonically with d, it is not the case
for f (ς), and hence r3 , which peaks at ς ≈ 1.14, i.e., d ≈ 1.75L ≈ 1.93RA . The
maximum fmax = 1.07 gives maximum interaction rate 2.33 × 10−7 /ωA . For higher
d the rate r3 decreases, implying that r3 is always weaker than r1 .
We also note that both gravitational field and field gradient are necessary (but
not sufficient) for producing entanglement. This is clear from consideration of yet
another configuration – an infinite plane and a point mass separated by a distance
L. One immediately observes that there is no field gradient here and that the
gravitational energy of this configuration is proportional to L − (xA − xB ), which
does not couple the masses and therefore does not contribute to entanglement.
94
Figure 6.8: Other configurations and notation for the two interacting masses. A
mass with spherical shape is situated next to another spherical mass with a different
radius (a) or a thin rod (b).
typically xA −xB L−2R, we can expand such expression to find a quadratic term
in xA − xB that can produce entanglement between the masses. The strength of this
term, however, is much weaker than the strength of the corresponding entangling
term of gravitational origin: for Osmium oscillators with mass ∼ 1 kg separated
by L = 2.1R, the ratio between the Casimir and gravitational term is rcg ∼ 10−12 .
Similar calculations made for released masses of the same material with m = 100 µg
and L = 3R, give rcg ∼ 10−2 . It is thus legitimate to ignore Casimir interaction in
both schemes.
where R is the radius of the sphere and T is the temperature of the environment.
Note that all variables are in SI units.
The decoherence due to interactions with other scattering particles, e.g., air
molecules, gives
8 N√
Λam = 2 2πmair R2 (kB T )3/2 , (6.19)
3~ V
where N/V is the density of air molecules with mass mair . We take mair ≈ 0.5 ×
10−25 kg. For ultrahigh vacuum, pressure ∼ 10−10 Pa, the density is ∼ 1012
particles/m3 . One could also consider performing these experiments in space. In
this case, by taking the pressure ∼ 10−15 Pa, the density can be as low as ∼
107 particles/m3 .
We take the average width of the wave function as an estimate for the super-
95
position that is subjected to decoherence. For oscillators made of Osmium, we use
m ∼ 1 kg and frequency ω ∼ 0.1 Hz. Taking L = 2.1R and starting with ground
state give ∆x ≈ 8 × 10−17 m. From interactions with thermal photons at environ-
mental temperature of 4 K (liquid Helium), the coherence time for the oscillators is
τph ∼ 5s. The coherence time due to collisions with air molecules can be improved
by evacuating the chamber with the oscillators – for about 1012 molecules/m3 (ul-
trahigh vacuum), the coherence time is also τam ∼ 5s. For the space experiment, by
taking the temperature as 2.7 K (cosmic microwave background) and assuming 107
molecules/m3 , we obtain τph ∼ 170s and τam ∼ 106 s.
By making similar calculations for released masses, with parameters considered
in Fig. 6.6, one obtains τph ∼ 105 s and τam ∼ 10−4 s for the experiment on Earth
with liquid Helium temperature and ultrahigh vacuum. For the space experiment
the coherence times improve to τph ∼ 107 s and τam ∼ 41s respectively. We present
in Table 6.2 a summary of coherence times.
Table 6.2: Coherence times for oscillators (m = 1 kg, ω = 0.1 Hz) and released
masses (m = 100 µg, ω = 100 kHz) both for experiments on Earth and in space. We
use the pressure P ∼ 10−10 Pa and the temperature T = 4 K for Earth experiment,
while P ∼ 10−15 Pa and T = 2.7 K for space experiment.
Coherence Oscillators Released masses
time Earth Space Earth Space
τph 5s 170 s 4 × 10 s 2 × 107 s
5
Other schemes have been proposed for gravitationally induced entanglement [108,
109]. They are based on Newtonian interaction between spatially superposed mi-
crospheres with embedded spins. In those proposals, entanglement is reached faster
and the small size of the experiment is the main advantage. However, in order
to separate gravitational and Casimir contributions in that setup, each diamond
sphere with mass m = 10−14 kg has to be superposed across 250 µm. Decoherence
due to scattering of molecules then becomes the main limiting factor. The schemes
we discussed here are complementary in a sense that vibrations of each oscillator
are minute (no macroscopic superposition) but larger mass, 100 µg, is required for
observable entanglement.
96
6.8 Comparison with recent experimental achieve-
ments
We have shown that two nearby masses – both trapped and released – can be-
come entangled via gravitational interaction. Let us now discuss the experimental
conditions required to observe this entanglement in light of recent experimental
achievements.
Logarithmic negativity in the order 10−2 has already been observed experimen-
tally between mechanical motion and microwave cavity field [114]. Extrapolating
the same entanglement resolution to the case of two massive oscillators sets the
required frequency to ω ∼ 10−2 Hz, see Eq. (6.3) and its expression in terms of
ω. Interestingly, kilogram-scale mirrors of similar frequency (ω ∼ 10−1 Hz) were
recently cooled down near their quantum ground state [111]. Furthermore, recent
experiments on squeezed light have reported high squeezing strength [129, 130] (see
also a review in this context [131]), up to 15 dB, which corresponds to s ≈ 1.73.
Advances in the state transfer between light and optomechanical mirrors [110] make
this high squeezing promising also for mechanical systems.
For released masses, the experimental requirements are more relaxed. Their
mass can be considerably smaller while the frequency for initial trapping consider-
ably higher, which is close to common experimental parameters used for optome-
chanical system [110, 112, 113]. Note that higher frequency (lower ∆x(0)) actually
improves entanglement gain, unlike in the oscillators case where small ω is prefer-
able. However, one has to be cautious of decoherence mechanisms as a result of faster
spreading rate of the wave functions. For future experiments, an improvement in
the sensitivity of entanglement detection will also be beneficial.
97
potential for generating entanglement.
The objection to the conservative approach is instantaneity of gravitational in-
teraction: Newtonian potential directly couples masses independently of their sepa-
ration. On the other hand, it has been shown that gravitational waves travel with
finite speed [132]. For nearby masses this retardation is hardly measurable and
Newtonian potential is dominant and expected to correctly describe the amount
of generated entanglement. A more consistent option in our opinion, motivated
by quantum formalism and comparison with other fundamental interactions, is to
treat gravitational field as a separate physical object. In this picture the masses are
not directly coupled, but each of them individually interacts with the field. It has
been argued within this mass-field-mass setting that entanglement gain between the
masses implies non-classical features of the field [45, 108, 109], see Chapter 3.
This discussion shows that it would be useful to provide methods for independent
verification of the presence or absence of a physical object mediating the interaction.
We finish with a toy example of a condition capable of revealing that there was no
mediator. To this end we consider two scenarios: (i) evolving a bipartite system
described at time t by a density matrix ρ12 (t); (ii) two objects interacting via a
mediator M , i.e., with Hamiltonian H1M + HM 2 , described by a tripartite state
ρ1M 2 (t). We ask whether there exists bipartite quantum dynamics ρ12 (t) that cannot
be obtained by tracing out the mediator in scenario (ii). Indeed, if ρ12 (t) is a pure
state at all times and entanglement increases, the dynamics could not have been
mediated. The purity assumption requires the mediator to be uncorrelated from
ρ12 (t), and uncorrelated mediator is not capable of entangling the principal system,
composed of particles 1 and 2 [45]. It would be valuable to generalise this argument
to mixed states measured at finite number of time instances.
For example, in an experimental situation, the state of particles 1 and 2 might
only be close to a pure state (with purity 1−, where is a small positive parameter)
and therefore they could be weakly entangled to the mediator M (with entanglement
∼ O()). One would naturally expect that entanglement gain between particles 1
and 2 is bounded, e.g., as a function of . An observation of higher entanglement
gain therefore excludes the possibility of mediated dynamics.
6.10 Summary
No experiment to date verified that gravity possesses quantum nature. A possible
way towards providing such evidence requires the generation of quantum entangle-
ment between massive objects. In this chapter, we have provided systematic study
of entanglement dynamics between two masses that are coupled gravitationally. We
put forward two proposals where the masses are either trapped in 1D harmonic po-
98
tentials or released from such traps. We derived figures of merit that are useful for
estimating experimental parameters needed to produce entanglement. Our schemes
have the advantage that no macroscopic superpositions develop during the dynam-
ics, resulting in observable entanglement generated within the coherence times of the
masses. Finally, we concluded with a discussion on the nature of the gravitational
interactions that can be infered from our proposals.
99
Chapter 7
1
Parts of this chapter are reproduced from our published article of Ref. [133],
which is licensed under the Creative Commons Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/). Where applicable, changes made will be indi-
cated.
101
7.1 Motivation and objectives
There is no a priori limit on the complexity, size or mass of objects to which quantum
theory is applicable. Yet, whether or not the physical configuration of macroscopic
systems could showcase quantum coherences has been the subject of a long-standing
debate. The pioneers of quantum theory, such as Schrödinger [134] and Bohr [135],
wondered whether there might be limitations to living systems obeying the laws of
quantum theory. Wigner even claimed that their behaviour violates unitarity [136].
A striking way to counter such claims on the implausibility of macroscopic quan-
tum coherence would be the successful preparation of quantum superposition states
of living objects. A direct route towards such goal is provided by matter-wave inter-
ferometers, which have already been instrumental in observing quantum interference
from complex molecules [137], and are believed to hold the potential to successfully
show similar results for objects as large as viruses in the near future.
However, other possibilities exist that do not make use of interferometric ap-
proaches. An instance of such alternatives is to interact a living object with a
quantum system in order to generate quantum correlations. Should such correla-
tions be as strong as entanglement, measuring the quantum system in a suitable
basis could project the living object into a quantum superposition. Furthermore,
requesting the establishment of entanglement is, in general, not necessary as the
presence of quantum discord, that is a weaker form of quantum correlations, would
already provide evidence that the Hilbert space spanned by the living object must
contain quantum superposition states [24, 36, 37, 46, 47]. For example, by operating
on the quantum system alone one could remotely prepare quantum coherence in the
living object [42].
A promising step in this direction, demonstrating strong coupling between living
bacteria and optical fields and suggesting the existence of entanglement between
them [138], has recently been realised [139]. See also Refs. [140–150] for a broader
picture of quantum effects in photosynthetic organisms. However, the experimental
results reported in Ref. [139] can as well be explained by a fully classical model [138,
139, 151, 152], which calls loud for the design of a protocol with more conclusive
interpretation.
In this chapter we make a proposal in such a direction by designing a thought
experiment in which the bacteria are mediating interactions between otherwise un-
coupled light modes. This scheme fits into the general framework of Ref. [45], which
shows in the present context that quantum entanglement between the light modes
can only be created if the bacteria are non-classically correlated with them during
the process. It is important to realise that in this way we bypass the need of exact
modelling of the living organisms and their interactions with external world. In-
102
deed, experimenters are never asked to directly operate on the bacteria, it is solely
sufficient to observe the light modes. A positive result of this experiment, i.e. obser-
vation of quantum entanglement between the light modes, provides an unambiguous
witness of quantum correlations, in the form of quantum discord, between the light
and bacteria.
In order to demonstrate that there should be observable entanglement in the
experiment we then propose a plausible model of light-bacteria interactions and
noises in the experiment. We focus on the optical response of the bacteria and
model their light-sensitive part by a collection of two-level atoms with transition
frequencies matching observed bacterial spectrum [139]. All processes responsible
for keeping the organisms alive are thus effectively put into the environment of these
atoms. We argue that standard Langevin approach gives a sensible treatment of this
environment due to its quasi-thermal character, low energies compared to optical
transitions and no evidence for finite-size effects. Within this model we find scenarios
with non-zero steady-state entanglement between the light modes which is always
accompanied by light-bacteria entanglement (in addition to quantum discord), which
is in turn empowered by the ultra-strong coupling2 between such systems.
(ii) All environments are local, i.e., they do not interact with each other.
103
We now propose a concrete scheme for revealing non-classicality of the bacteria
and argue how it meets the conditions above. Consider the arrangement in Fig. 7.1.
The bacteria are inside a driven single-sided multimode Fabry-Perot cavity where
they interact independently with a few cavity modes. The cavity modes are divided
into two sets which play the role of systems A and B in the general framework. The
bacteria are mediating the interaction between the modes and hence they represent
system C. Condition (i) above can be realised in practice in at least two ways.
An experimenter could utilise the polarisation of electromagnetic waves and group
optical modes polarised along one direction to system A and those polarised orthog-
onally to system B. Another option, which we will study in detail via a concrete
model below, is to choose different frequency modes and arbitrarily group them into
systems A and B. Condition (ii) holds under typical experimental circumstances
where the environment of the cavity modes is outside the cavity whereas that of
the bacteria is inside the cavity or even part of bacteria themselves. The electro-
magnetic environment outside the cavity is a large system giving rise to the decay
of cavity modes but having no back-action on them. Therefore each cavity mode
decays independently and cannot get entangled via interactions with the electro-
magnetic environment. Finally, condition (iii) is satisfied right before placing the
bacteria into the cavity, because at this time all three systems A, B, and C are in
a completely uncorrelated state ρA ⊗ ρB ⊗ ρC .
Figure 7.1: Proposed experimental setup for probing quantum features of photosyn-
thetic organisms. In particular, green sulphur bacteria are mediating interactions
between a few cavity modes in a driven multimode Fabry-Perot cavity, which are
otherwise non-interacting. The cavity has an input mirror with reflectivity R1 and
an end mirror with perfect reflectivity, i.e., R2 ≈ 1. As a result of interactions with
the environment, the mth cavity mode has a decay rate κm , while the bacterial sys-
tem dissipates at a rate γn . In the text, we treat cavity fields and bacterial modes
as open systems.
We note again that this discussion is generic with almost no modelling of the
involved systems. In particular, nothing has been assumed regarding the physics of
the bacteria and their interactions with light and the external world. This makes
104
our proposal experimentally attractive.
In order to make concrete predictions about the amount of intermodal entangle-
ment EA:B we now study a specific model for the energy of the discussed system.
This additional assumption about the overall Hamiltonian will allow us to demon-
strate that the entanglement EA:B is accompanied by light-bacteria entanglement
EAB:C . This independently confirms the presence of light-bacteria discord as entan-
glement is a stronger form of quantum correlations than discord [24, 46, 47]. In the
remainder of this chapter we will therefore only calculate entanglement.
Here, m = 1, . . . , M is the label for the mth cavity mode, whose annihilation (cre-
ation) operator is denoted by am (a†m ) and having frequency ωm . Moreover, both
105
harmonic oscillators describing the bacteria are labelled by n = I, II with bn (b†n ) de-
noting the corresponding bosonic annihilation (creation) operator. Each oscillator
is coupled to the mth cavity field at a rate Gmn . The collective form of the coupling
√ p
allows us to write Gmn = gmn N with gmn = µn ωm /2~r 0 Vm , where µn is the
dipole moment of the nth two-level transition, r relative permitivity of medium,
and Vm the mth mode volume [152], see also [156, 157] for similar treatments. The
cavity is driven by a multimode laser, each mode having frequency Λm , amplitude
p
Em = 2Pm κm /~Λm , power Pm , and amplitude decay rate of the corresponding
cavity mode κm . It is important to notice that in Eq. (7.1) we have not invoked the
rotating-wave approximation but actually retained the counter-rotating terms am bn
and a†m b†n . These cannot be ignored in the regime of ultra-strong coupling and we
will show that they actually play a crucial role in our proposal.
106
where γn is the amplitude decay rate of the bacterial system. Fm and Qn are op-
erators describing independent zero-mean Gaussian noise affecting the mth cavity
field and the nth bacterial mode respectively. The only nonzero correlation func-
tions between these noises are hFm (t)Fm† 0 (t0 )i = δmm0 δ(t − t0 ) and hQn (t)Q†n0 (t0 )i =
δnn0 δ(t − t0 ) [158, 159]. We note that in this model the light-sensitive part of the
bacteria is treated collectively, i.e. all its two-level atoms are indistinguishable. This
assumption is standardly made in present-day literature, see e.g. [139, 154] where
modelling of the bacteria / chlorosomes as a harmonic oscillator fits observed exper-
imental results. But it should be stressed that this assumption deserves an in-depth
experimental assessment.
We express the Langevin equations in terms of mode quadratures. In particular,
√ √
we use Xm ≡ (am + a†m )/ 2 and Ym ≡ (am − a†m )/i 2. This allows us to write
the Langevin equations for the quadratures in a simple matrix equation u̇(t) =
Ku(t) + l(t), with the vector u = (X1 , Y1 , · · · , XM , YM , XI , YI , XII , YII )T and
I1 0 ··· 0 L1I L1II
0 I2 ··· 0 L2I L2II
.. .. ... .. .. ..
. . . . .
K= , (7.3)
0
0 · · · IM LM I LM II
L1I L2I · · · LM I II 0
L1II L2II · · · LM II 0 III
107
i.e. l(t) = η(t) + p(t) where
√
κ1 X1 (t) E1 cos Λ1 t
√
κ1 Y1 (t) −E1 sin Λ1 t
.. ..
√ .
.
κ X (t)
p(t) EM cos ΛM t
M M
η(t)
√ = √κM YM (t)
, √ = −EM sin ΛM t . (7.5)
2 √ 2
γI XI (t) 0
√
√ γI YI (t) 0
γ X (t) 0
II II
√
γII YII (t) 0
√
We have also used quadratures for the noise terms, i.e. through Fm = (Xm +iYm )/ 2
√
and Qn = (Xn + iYn )/ 2.
The solution to the Langevin equations is given by
Z t
u(t) = W+ (t)u(0) + W+ (t) dt0 W− (t0 )l(t0 ), (7.6)
0
where W± (t) = exp (±Kt). This allows numerical calculation of expectation value
of the quadratures as a function of time, i.e. hui (t)i is given by the ith element of
Z t
W+ (t)hu(0)i + W+ (t) dt0 W− (t0 )p(t0 ), (7.7)
0
Vij (t) = hui (t)uj (t) + uj (t)ui (t)i/2 − hui (t)ihuj (t)i
V (t) = W+ (t)V (0)W+T (t)
Z t
+W+ (t) dt0 W− (t0 )DW−T (t0 ) W+T (t), (7.8)
0
108
where D = Diag[κ1 , κ1 , · · · , κM , κM , γI , γI , γII , γII ] and we have assumed that the
initial quadratures are not correlated with the noise quadratures such that the mean
of the cross terms are zero. A more explicit solution of the covariance matrix, after
integration in Eq. (7.8), is given by
where Z is the total number of modes, which is M + 2 in our case. The block
component, here denoted as Bjk , is a 2 × 2 matrix describing local mode correlation
when j = k and intermodal correlation when j 6= k. A Z-mode covariance matrix
has symplectic eigenvalues {νk }Zk=1 that can be computed from the spectrum of
109
matrix |iΩZ V | [25] where !
Z
M 0 1
ΩZ = . (7.12)
k=1
−1 0
For a physical covariance matrix 2νk ≥ 1.
Entanglement is calculated as follows. For example, the calculation in the par-
tition 12 : 34 only requires the covariance matrix of modes 1, 2, 3, and 4:
B11 B12 B13 B14
T
B12 B22 B23 B24
V = T T
, (7.13)
B13 B23 B33 B34
T T T
B14 B24 B34 B44
that can be obtained from Eq. (7.11). If the covariance matrix Ṽ , after partial trans-
position with respect to mode 3 and 4 (this is equivalent to flipping the sign of the
operator Y3 and Y4 in V ) is not physical, then our system is entangled. This unphys-
ical Ṽ is shown by its minimum symplectic eigenvalue ν̃min < 1/2. Entanglement is
then quantified by logarithmic negativity as follows E12:34 = max[0, − log2 (2ν̃min )]
[34, 35].3 Note that the separability condition, when ν̃min ≥ 1/2, is sufficient and
necessary when one considers bipartitions with one mode on one side [120], e.g.,
partition between bacterial modes I : II.
Simulation parameters
The parameters used in our simulations are taken, wherever possible, from the ex-
periments of Ref. [139]. We place the bacteria in a single-sided Fabry-Perot cavity
of length L = 518 nm (cf. Fig. 7.1). The refractive index due to aqueous bacterial
√
solution embedded in the cavity is nr = r ≈ 1.33, which gives the frequency of
the mth cavity mode ωm = mπc/nr L ≈ 1.37m × 1015 Hz. The reflectivities of the
mirrors are engineered such that R2 = 100% and R1 = 50%. We assume the reflec-
tivities are the same for all the optical modes, giving κm ≈ 7.5 × 1013 Hz through
the finesse Fi = −2π/ ln (R1 R2 ) = πc/2κm nr L. The decay rate of the excitons can
be calculated as γn = 1/2τn where τn = 2h/Γn is the coherence time with Γn be-
ing the full-width at half-maximum (FWHM) of the bacterial spectrum [160]. We
approximate the spectrum in Fig. 1b of Ref. [139] as a sum of two Lorentzian
functions centred at ΩI and ΩII having FWHM of (ΓI , ΓII ) = (130, 600) meV, giving
(γI , γII ) ≈ (0.78, 3.63)×1013 Hz respectively. Note that the decay rate solely depends
on the coherence time, i.e., we assume only homogenous broadening of the spectral
lines.
3
Note that we have used logarithm with base 2, instead of natural logarithm as in Ref. [133],
to maintain consistencies in this thesis.
110
All the spectral components of the driving laser are assumed to have the same
power Pm = 50mW and frequency Λm = ωm . By using the mode volume Vm =
2πL3 /m(1−R1 ) [161], we can express the interaction strength as Gmn = mG̃n , where
p
we define G̃n ≡ µn c(1 − R1 )N/4~n3r 0 L4 . This quantity is a rate that characterises
the base collective interaction strength of the cavity mode and the nth bacterial
mode. Instead of fixing the value of G̃n , we vary this quantity G̃n = [0, 0.2] 1015 Hz,
which is within experimentally achievable regime (cf. Refs. [138, 139]).
Entanglement dynamics
Let us consider as initial the time right before the bacteria are inserted into the
cavity. Then all the cavity modes and the bacteria are completely uncorrelated and
do not interact. The dynamics is then started by placing the bacteria in the cavity.
In what follows, as an example of the dynamics we start with vacuum state for the
cavity modes and ground state for the bacteria. The initial state of the bacteria is
justified by the fact that ~Ωn kB T , even at room temperature.
Fig. 7.2 shows the resulting entanglement dynamics. Panel (a) displays exis-
tence of steady-state entanglement between cavity modes 1, 2 and 3, 4, which is not
altered heavily if the calculations take into account five and six cavity modes in
total. Therefore, we consider 4 cavity modes in all other calculations. In recent
experiments, the rate G̃I was shown to be 3.9 × 1013 Hz [139] and the corresponding
G̃II = 6 × 1013 Hz. In our calculations we vary the latter rate as in panels (b) and
(c) (also see Fig. 7.3). As expected the higher the rate the more entanglement gets
generated. It is also apparent that entanglement between the cavity modes and
bacteria E1234:I II grows faster than entanglement between the cavity modes. More
precisely, nonzero E12:34 implies nonzero E1234:I II .
We consider four cavity modes as the addition of higher modes shows negligible
effects to the steady state entanglement. In the steady state regime, we calculate
entanglement between the cavity modes E12:34 , between the cavity modes and bac-
teria E1234:I II , and between the bacterial modes EI:II , cf. Fig. 7.3. This steady
state regime is reached in ∼ 100 fs (see Fig. 7.2), which is faster than relaxation
processes (∼ ps) occuring within green sulphur bacteria [154]. Our results show
that the steady state entanglement E12:34 is always accompanied by E1234:I II , i.e.,
the bacteria are non-classically correlated with the cavity modes. This is in agree-
ment with the general detection method of Ref. [45] as entanglement is a stronger
type of quantum correlation than discord, i.e., nonzero E1234:I II implies nonzero cav-
ity modes-bacteria discord D1234|I II . Our results also show that the entanglement
111
Figure 7.2: Entanglement dynamics in the field-bacteria-field system. (a) Entangle-
ment E12:34 between the cavity modes {1, 2} and {3, 4}. The entanglement is shown
up to 4, 5, and 6 cavity modes in the cavity. Higher modes do not contribute to the
steady state entanglement due to higher detuning from the bacterial system. We
therefore only consider 4 cavity modes for the rest of the dynamics. (b) Entangle-
ment between the cavity modes for varied interaction strength with the bacterial
system. (c) Cavity modes–bacteria entanglement. The correlation from this parti-
tion is higher and showing faster initial growth compared to the others. The coupling
parameters used are G̃I = 3.9 × 1013 Hz for all graphs and G̃II = 6 × 1013 Hz for (a).
Note that the steady state of all dynamics is reached within ∼ 100 fs and we have
used the logarithmic negativity as entanglement quantifier.
dynamics of E12:34 is dominated by modes 2 and 3 since other modes are further
off resonance with the bacterial modes. Moreover, there is entanglement generated
within the bacteria. This requires both G̃I and G̃II to be nonzero and relatively
high. We see that the bacteria can be strongly entangled with the cavity modes,
much stronger than entanglement between the cavity modes. While the latter is in
the order of 10−2 − 10−3 , we note that entanglement in the range 10−2 has already
been observed experimentally between mechanical motion and microwave cavity
field [114]. We have also indicated, as black dots in Fig. 7.3, the coupling strengths
112
G̃I = 3.9 × 1013 Hz from Ref. [139] and the corresponding G̃II = 6 × 1013 Hz, which
R
is estimated as follows. From the relation µ2n ∝ f (ω)dω/ωn [162], where f is the
extinction coefficient, one can obtain the ratio G̃II /G̃I = µII /µI ≈ 1.53.
Figure 7.3: Steady state entanglement in field-bacteria-field system for varied inter-
action strengths. In all panels, the horizontal axis is G̃I = [0, 0.2] 1015 Hz, while G̃II
is varied in 1015 Hz as: 0 (red curves), 0.05 (green curves), 0.1 (blue curves), 0.15
(magenta curves), and 0.2 (black curves). Black dots in the panels show the values
of entanglement for recently realised coupling strengths G̃I = 3.9 × 1013 Hz by Coles
et al. [139] and the corresponding G̃II = 6 × 1013 Hz. We note from our simulations
that the entanglement E1234:I II is nonzero whenever E12:34 is present. Furthermore,
entanglement in the bacterial system, panel (c), also persists for stronger couplings.
113
a single cavity mode A and a mechanical mirror B the coupling is proportional to
a† aXB , which is a third-order operator [164]. This results in the effective coupling
strength being proportional to the classical cavity field intensity α after linearisation
of the Langevin equations. This classical signal enters the covariance matrix via the
effective coupling strength and introduces the dependence on the driving power.
In order to justify the low atomic excitation limit we first note that the number
of steady-state photons for the mth cavity mode without the presence of the bacteria
2
is given by Em /κ2m ∝ Pm . When one considers the bacteria in the cavity having the
base interaction strength G̃n and a decay rate γn in the same order as the cavity
decay rate, the number of excitations of the bacterial modes would also be in the
2
order of Em /κ2m , which in our case is 103 . With ∼ 108 actively coupled dipoles in
the cavity [139], this gives ∼ 10−3 % excitation, which justifies the low-excitation
approximation. We also plotted the evolution of excitation numbers of the bacterial
modes (together with the number of photons in different cavity modes) within our
model in Fig. 7.4. It shows that excitation numbers are oscillating in the “steady
state”. The oscillations are caused by the combination of interactions between the
light and bacteria (Rabi-like oscillations) and the time-dependent driving laser. Set-
ting the interactions Gmn = 0 or the driving off (Pm = 0) indeed produces constant
steady-state value. We observe that the excitation number of the bacterial system
is always below 2000, which is in agreement with the statement above.
The excitation number of the cavity modes and bacteria as a function of time
can be calculated from hui (t)i and Vii (t). For example, the mean excitation number
for the first cavity mode is given by
1
N̄1 (t) = ha†1 (t)a1 (t)i = (V11 (t) + V22 (t) + hu1 (t)i2 + hu2 (t)i2 − 1). (7.14)
2
We present the evolution of photon number of the cavity modes and excitation
of the bacterial modes in Fig. 7.4. Note that photon number of the third cavity
mode (solid magenta line) is showing oscillations well below its “off-interaction”
value (dashed magenta line). This is because ω3 is almost in resonance with the
frequency of the atomic transition ΩII .
114
Figure 7.4: Dynamics of the bacterial excitation numbers and photon number of
the cavity modes. Solid curves represent the interacting system: N̄1 , N̄2 , N̄3 , N̄4
for the cavity modes and N̄I , N̄II for the bacterial system. The couplings are taken
as G̃I , G̃II = (3.9, 6) × 1013 Hz. The corresponding dynamics for non-interacting
system, i.e., Gmn = 0, are given by the dashed curves.
state and on the power of the driving lasers, we might start with all atoms in the
ground state, vacuum for the light modes, and no driving. Under such circumstances
there is no interaction between bacterial modes and light modes as every term in the
interaction Hamiltonian contains an annihilation operator. In physical terms, since
we begin with the lowest energy state and the interaction Hamiltonian preserves
energy, the ground state will be the state of affairs at any time. Therefore, nonzero
entanglement observed in experiments will provide evidence of the existence of the
counterrotating terms, showing a signature of the ultra-strong coupling regime.
7.6 Summary
Recent experimental studies have shown that the regime of strong coupling between
photosynthetic organisms such as green sulphur bacteria and light is accessible. In
this chapter, we put forward a proposal for probing a quantum property of the bac-
teria (as characterised by quantum correlations) without directly measuring them.
Our general proposal bypasses the need of modelling both the bacteria and the in-
teractions they have with external world. The proposed scheme utilises the bacteria
as mediators between two groups of non-interacting cavity light modes. In order to
show feasibility, we modelled the light-sensitive part of the bacteria and their interac-
tions with the cavity modes and environments using recent experimental parameters.
Within this model, our simulations show that, in most cases, quantum entanglement
endures the environmental noises and remains present in the steady state. Our sim-
ulations also confirmed that entanglement between the cavity light modes reveals a
quantum property of the bacteria, i.e., quantum entanglement between the bacterial
modes and cavity modes, and for stronger interactions, even quantum entanglement
115
within the bacterial modes. We note that the presence of entanglement also provides
independent evidence of the ultra-strong coupling regime.
116
Chapter 8
Other applications
In this chapter, I will present other applications that result from the principles of
distribution of correlations between two objects whose interactions are mediated by
an ancillary system – investigated in Chapters 3 to 5.1 First, I will apply our
non-classicality detection method to an experimentally relevant platform of optome-
chanics, in particular, the membrane-in-the-middle setting. Our protocol indirectly
detects the quantum property of the mechanical membrane by observing the dynamics
of entanglement between probing light modes. Next, I present a protocol to reveal
the correlation between a system and its environment in a scenario of open system
dynamics. Then, high gain of correlations will be demonstrated in a dynamics of
two cavity fields interacting with a two-level atom. This gain will be shown to imply
a quantum trait – non-commutativity of interaction Hamiltonians. Finally, a brief
discussion will be presented that shows an application of our method to estimate
dimensionality of a quantum object.
1
Parts of this chapter are reproduced from our published articles of Refs. [45, 55], c [2019]
American Physical Society. Where applicable, changes made will be indicated.
117
8.1 Membrane-in-the-middle optomechanics
In this section, we address the practical implication of our criteria from Chapter 3 to
experiments of cavity optomechanics [110]. This is a paradigmatic open mesoscopic
quantum system for which the non-classicality detection method in Chapter 3 holds
the potential to be practically significant. In fact, one of the goals of optomechanics
is to infer the non-classicality of the state of a massive mechanical system, in similar
spirit as “certification” in Refs. [165, 166], without affecting its (in general fragile)
state. A possible setting for such a task is given by the so-called membrane-in-
the-middle configuration, where a mechanical oscillator (a membrane) is suspended
at the centre of a two-sided optical cavity [163]. By driving the cavity with laser
fields from both its input mirrors, respectively, we realise a situation completely
analogous to that in Fig. 3.1 (cf. Fig. 8.1). We now show that our scheme detects
non-classicality of the membrane without measuring it.
~ωC 2
Hloc = ~∆0A a† a + ~∆0B b† b + (PC + XC2 )
2
+i~EA (a† − a) + i~EB (b† − b) (8.1)
and
Hint = −~G0A a† a XC + ~G0B b† b XC , (8.2)
118
Figure 8.1: Membrane-in-the-middle optomechanical setup. Cavity field A is inter-
acting indirectly with cavity field B via a perfectly reflecting mechanical membrane
C. The interactions of the movable mechanical mirror with its thermal environment
result in the Brownian noise. The cavities are driven by lasers and experience energy
dissipation through the fixed input mirrors.
The dynamics of the operators, adding into account noise and damping terms (also
local), can be well written by a set of Langevin equations in Heisenberg picture
√
ȧ = −(κA + i∆0A )a + iG0A aXC + EA + 2κA ain
√
ḃ = −(κB + i∆0B )b − iG0B bXC + EB + 2κB bin
X˙ C = ωC PC
P˙C = −ωC XC + G0A a† a − G0B b† b − γC PC + ξ (8.3)
where γC is damping rate of the membrane. Also jin is input noise of field J associ-
ated with cavity-input mirror interface and has only correlation function
†
hjin (t)kin (t0 )i = δjk δ(t − t0 ) [159], whereas ξ is Brownian noise of the membrane and
has correlation function hξ(t)ξ(t0 ) + ξ(t0 )ξ(t)i/2 ≈ γC (2n̄ + 1)δ(t − t0 ) in the limit of
interest that is large mechanical quality of the membrane, i.e., ωC /γC 1 [118, 119].
The mean phonon number of the membrane reads n̄ = 1/(exp (~ωC /kB T ) − 1).
The linearised Langevin equations can be obtained by splitting the operators into
steady state values and fluctuating terms. In particular we write XC = XCs + δXC ,
PC = PCs + δXC , and j = αJs + δj. By inserting these into Eq. (8.3) and ignoring
nonlinear terms δj † δj and δj δXC one gets a set of linear Langevin equations for
119
the fluctuations of the quadratures
√
δ X˙ A = −κA δXA + ∆A δYA + 2κA xin,A
√
δ Y˙A = −κA δYA − ∆A δXA + GA δXC + 2κA yin,A
√
δ X˙ B = −κB δXB + ∆B δYB + 2κB xin,B
√
δ Y˙B = −κB δYB − ∆B δXB − GB δXC + 2κB yin,B
δ X˙ C = ωC δPC
δ P˙C = −ωC δXC − γC δPC + GA δXA − GB δXB + ξ (8.4)
where effective detuning ∆A ≡ ∆0A − G0A XCs , ∆B ≡ ∆0B + G0B XCs , and effective
√
coupling GJ ≡ 2G0J αJs . The steady state values are given by PCs = 0, XCs =
p
(G0A |αAs |2 − G0B |αBs |2 )/ωC , and αJs = |EJ |/ κ2J + ∆2J . The quadratures of the
√
field XJ and YJ are related to the field operator j through j = (XJ + iYJ )/ 2.
√
This relation also applies for the input noise, i.e., jin = (xin,J + iyin,J )/ 2.
For simplicity one can re-write Eqs. (8.4) as a single matrix equation u̇(t) =
Ku(t) + n(t) where the vector u(t) = (δXA , δYA , δXB , δYB , δXC , δPC )T ,
√ √ √ √
n(t) = ( 2κA xin,A , 2κA yin,A , 2κB xin,B , 2κB yin,B , 0, ξ)T , (8.5)
and
−κA ∆A 0 0 0 0
−∆A −κA 0 0 GA 0
0 0 −κB ∆B 0 0
K= . (8.6)
0 0 −∆B −κB −GB 0
0 0 0 0 0 ωC
GA 0 −GB 0 −ωC −γC
The solution to this linearised Langevin equation is then given by u(t) = M (t)u(0)+
Rt
0
dsM (s)n(t − s) where M (t) = exp (Kt).
120
matrix, after integration, is given by
which is linear and can easily be solved numerically. For our simulations of the
dynamics below, we take the initial state to be thermal state for c and coherent
state for field j, this gives V (0) = Diag[1, 1, 1, 1, 2n̄ + 1, 2n̄ + 1]/2. If one is only
interested in steady state solution, it is guaranteed when all real parts of eigenvalues
of K are negative, giving M (∞) = 0 such that the steady state covariance matrix
can be calculated from a simpler equation KV (ts ) + V (ts )K T = −D.
N
!
M 0 1
ΩN = . (8.9)
k=1
−1 0
For a physical covariance matrix 2νk ≥ 1. For an entangled system, e.g., in the par-
tition AB : C, the covariance matrix will not be physical after partial transposition
with respect to mode C (this is equivalent to flipping the sign of the membrane’s
momentum fluctuation operator δPC in V ). For our system, this unphysical V TC
is shown by one of its three symplectic eigenvalues ν̃min < 1/2. Entanglement is
then quantified by logarithmic negativity as follows EAB:C = max[0, − log2 (2ν̃min )]
[34, 35].2 Note that the separability condition, when V TC has ν̃min ≥ 1/2, is suffi-
cient and necessary for 1 : N mode partition [120]. Entanglement EA:B is calculated
2
Note that we have used logarithm with base 2, instead of natural logarithm as in Ref. [45], to
maintain consistencies in this thesis.
121
in similar manner by only considering system AB where the covariance matrix is
now !
LAA LAB
VAB = . (8.10)
LTAB LBB
In order to independently confirm the non-classicality of the membrane and
demonstrate that there is considerable entanglement to be detected we calculate the
ensuing entanglement dynamics. As mentioned above, we start with the experimen-
tally natural state where C is in a thermal state and A and B are coherent states,
and calculate the dynamics of entanglement EA:B and EAB:C . Since initially there
is no entanglement, the first step in Fig. 3.4a can be omitted.
The parameters used in our simulations all adhere to present-day technology
[167]. This includes mC = 145 ng, T = 300 mK, lJ = 25 mm, and (ωC , ωlJ , γC ) =
2π(947 × 103 , 2.8 × 1014 , 140) Hz. Finesse of each cavity is 1.4 × 104 . The results of
our analysis are presented in Fig. 8.2 for varying power of the right laser. We see
that non-zero EA:B (τ ) is always accompanied by non-zero EAB:C at some time (0, τ ).
Note that entanglement is a stronger type of quantum correlations than discord. We
have also performed similar calculations by varying the power of the left laser as well
as the frequencies of the lasers within experimentally accessible ranges and observed
consistent results.
If one is interested only in the steady state regime, Fig. 8.3 shows the corre-
122
sponding entanglement EA:B while EAB:C is zero in this range (not shown). Note
that red colour has been used in the plots for parameters that do not correspond to
steady state solution.
Figure 8.3: Exemplary steady state entanglement between the cavity fields in the
membrane-in-the-middle optomechanical setup. The power of the right laser is var-
ied as PB = [0, 100] mW and the detuning ∆B = [−1.05, −0.85] ωC . Other param-
eters are the same as those stated in Fig. 8.2. Red colour region corresponds to
non-steady state regime.
123
8.2.2 Our protocol
In order to apply our method in this scenario, let us consider again a closed-system
dynamics and, in line with the assumed inaccessibility of the mediator, focus the at-
tention to the probes only (see Fig. 3.1). We could thus think of C as an environment
in contact with the open system AB, as seen in Fig. 8.4.
Now we use our scheme of Fig. 3.4a to reveal SECs, with the advantage that
the applications of an entanglement breaking channel and state tomography are
not necessary. This is a consequence of Lemma 3.3.2 that, if rephrased in terms
of the aim in this section, states that quantum entanglement in the open system
between two objects A and B cannot increase if the environment is not correlated
with them at all times. Note that the amount of entanglement can be quantified by
any entanglement monotone.
Recall also that our detection protocol of Fig. 3.4a is used to infer non-classicality
of correlations, i.e., quantum discord, between the environment and the open system.
We note that previous schemes detect the non-classicality of the system [179, 180],
i.e., presence of DC|AB , whereas our schemes ascertain the non-classicality of the
environment, DAB|C , which is perhaps a prime example of an inaccessible object.
124
8.3.1 Witnessing non-commutativity and discord
Consider a two-level atom C, i.e., dC = 2, mediating interactions between two cavity
fields A and B as illustrated in Fig. 8.5. A similar scenario has been considered and
implemented, for example, in Refs. [60, 61, 184, 185]. The interaction between the
atom and each cavity field is taken to follow the Jaynes-Cummings model,
where a (b) is the annihilation operator of field A (B), while σ + (σ − ) is the raising
(lowering) operator of the two-level atom. For simplicity, we have assumed that the
interaction strengths between the two-level atom and the fields are the same. Note
that H is of the form HAC + HBC with non-commuting components.
Figure 8.5: Two orthogonally polarised cavity fields interact via a two-level atom,
but not with each other. In the text, we show that correlation gain between A and
B is linked with non-commutativity of interaction Hamiltonians, or more generally,
non-decomposability of time evolution.
The resulting correlation dynamics are plotted in Fig. 8.6. Mutual informa-
tion and negativity were calculated directly, whereas for the classical correlation
and the relative entropy of discord, we provide the lower bounds C̃A:B and −SA|B ,
respectively. C̃A:B is calculated as the mutual information of the state resulting
from projective local measurements in the Fock basis (no optimization over mea-
surements performed). The negative conditional entropy −SA|B is a lower bound
on the distillable entanglement [186], which in turn is a lower bound on the rela-
tive entropy of entanglement EA:B [187]. Therefore, we note the chain of inequal-
ities −SA|B ≤ EA:B ≤ DA|B ≤ IA:B , where the last two inequalities follow from
[38]. Already these lower bounds can beat the limit set by decomposable evolu-
tion, and therefore, all mentioned correlations can detect non-decomposability of
the evolution. Since we consider closed systems, this infers non-commutativity of
the Jaynes-Cummings couplings. We also note another non-classical feature of the
studied dynamics: since Fig. 8.6 shows entanglement gain, according to Chapter 3
there must be quantum discord DAB|C during the evolution.
125
It is apparent that the detection is easier (faster and with more pronounced vio-
lation) with a higher number of photons in the initial states of the cavity fields.
We offer an intuitive explanation. Consider, for example, |mn0i as the initial
√
state of ABC. By defining ξ = (a + b)/ 2, the Hamiltonian of Eq. (8.11) be-
√
comes 2~g(ξσ + + ξ † σ − ) and it is straightforward to obtain the unitary evolu-
tion [188]. One finds that the quantum state of the fields oscillates incoherently
between m+n
P Pm+n−1
j=0 cj (t)|jiA |m + n − jiB and j=0 dj (t)|jiA |m + n − 1 − jiB . Both of
these states are superpositions of essentially m + n bi-orthogonal terms giving rise
to high entanglement and, therefore, also other forms of correlations.
Figure 8.6 illustrates that different correlation quantifiers have different detection
capabilities and it is not clear at this stage whether there is a universal measure with
which non-commutativity is detected, e.g., the fastest. For most initial states we
studied mutual information detected non-commutativity the most rapidly, but there
are exceptions, as shown by the black curve corresponding to the initial state |101i.
With this initial state the mutual information never violates its bound, but the
negativity does.
Figure 8.6: Correlation dynamics in the field-atom-field system. The solid curves
represent the Jaynes-Cummings coupling while the dashed lines show the corre-
sponding correlation bound for decomposable dynamics. In the panels, we show
correlations between field A and field B: (a) the mutual information, (b) lower
bound on the classical correlation, (c) lower bound on the relative entropy of dis-
cord, and (d) the entanglement as quantified by negativity. The varied parameter
is the initial state of the tripartite system ABC: |110i (red curves), |101i (black
curves), |210i (green curves), and |220i (blue curves). Note that the horizontal axis
is given by the dimensionless time gt.
126
8.3.2 Strong interactions and bounded entanglement
Our results imply that the non-commutativity (non-decomposability in general) is a
desired feature of interactions in the task of correlation distribution, which is impor-
tant for quantum information applications. As a contrasting physical illustration,
we consider the strong dipole-dipole interactions in our field-atom-field example.
The Hamiltonian reads
with commuting components, i.e., [HAC , HBC ] = 0. One can verify that the results
of this model are in agreement with all the bounds we derived in Chapter 4. Further-
more, we prove below that, with this coupling, the state of AB at time t is effectively
given by a two-qubit separable state. This makes NA:B (t) = 0 and IA:B (t) ≤ 1. Note
the counter-intuitive result that strong interactions produce bounded correlations
between the probes, while weak interactions (Jaynes-Cummings coupling) can in-
crease the correlations above the bounds.
√
Let us define ξ = (a + b)/ 2. The dipole-dipole Hamiltonian, Eq. (8.12), is
√
reformulated as H 0 = 2~g(ξ + ξ † )σ x , where σ x = σ + + σ − and [ξ, ξ † ] = 1. The
unitary evolution operator is given by
iH 0 t
Ut = e− ~ (8.13)
1 √ †
√ †
= [(1 − σ x )ei 2gt(ξ+ξ ) + (1 + σ x )e−i 2gt(ξ+ξ ) ]
2
1
= [(1 − σ x )Da (α)Db (α) + (1 + σ x )Da (−α)Db (−α)],
2
where α = igt and, e.g., Da (α) = exp(αa† − α∗ a). Given an initial state |mn0i, the
state at time t reads
where
q
(mn)
d±± = 2 [1 ± e−2|α|2 Lm (4|α|2 )][1 ± e−2|α|2 Ln (4|α|2 )],
(n) 1
|D± i = q [D(α) ± D(−α)]|ni.
(nn)
d±±
127
the atomic mode C, the state of the fields is effectively given by a two-qubit state,
(mn) (mn) (mn)
(d++ )2 0 0 d++ d−−
(mn) 2 (mn) (mn)
1
0 (d+− ) d+− d−+ 0
,
(mn) (mn) (mn)
16
0 d+− d−+ (d−+ )2 0
(mn) (mn) (mn) 2
d++ d−− 0 0 (d−− )
(8.14)
which is positive under partial transposition and, hence, separable [30, 51]. The
same result follows for initial state |mn1i.
128
Chapter 9
In this chapter, I present the conclusion of the work that has been reported in this
thesis. Additionally, some immediate questions arise, which have not yet been set-
tled. I will discuss these key questions with preliminary results, wherever possible.
In particular, this includes a natural question on the general entanglement bound
for the indirect interaction scenario, a quantitative bound on the amount of non-
commutativity of Hamiltonians (in general, non-decomposability of time evolution
operator), a request on the protocol capable of witnessing the presence of quantum
objects (i.e., the ones that can be given a description within the quantum framework
such as living in a Hilbert space), and a strict bound on entangling time for initial
states where the mediator is decoupled.
129
9.1 Conclusion
We have studied the creation of correlations, mostly quantum entanglement, be-
tween two principal quantum objects that are continuously interacting via a medi-
ator – the indirect interaction setting. The research in this thesis has been devoted
to understanding the factors that are crucial in this scenario. It includes the re-
quired property of the mediator, the amount of correlation gain, and the speed at
which entanglement is created. Our work also resulted in key applications ranging
from the current technologically available platform detecting non-classicality of an
optomechanical mirror to a proposal on revealing a quantum nature of gravity.
In particular, Chapter 3 has shown that, for initially uncorrelated objects, quan-
tum discord between the mediator and the principal objects is a necessary condi-
tion for distribution of quantum entanglement. The contrapositive of our theorem
is therefore the revelation of quantum discord (a form of non-classical feature of
the mediator) from the observation of entanglement gain between the principal ob-
jects. Next, in Chapter 4, we have made a connection between the amount of
distributed correlations and non-commutativity of Hamiltonians or, more generally,
non-decomposability of the dynamical operator (another form of non-classicality).
The speed of entanglement creation has been investigated in Chapter 5. We have
presented a lower bound on entangling time for both direct and indirect interaction
settings. We proved that entanglement cannot be indirectly distributed faster than
the direct time bound. However, some examples, all of which require an initially
correlated mediator, have been shown to saturate the bound. We note that in Chap-
ters 3–5, wherever possible, we have used minimalistic assumptions in order to make
our theorems as general as possible. This includes the relaxation of knowledge of
the dimensionality of all the objects under scrutiny, the initial state of the system,
and details of interactions. All the objects can also be open systems.
Applications resulted from this work have been reported in three chapters. First,
Chapter 6 focussed on our proposal on the observation of quantum entanglement
between two masses. This has been argued as a potential route towards the detec-
tion of a quantum nature of gravitational interactions. Next, Chapter 7 has shown
another proposal on probing of non-classicality of photosynthetic organisms, in par-
ticular, the green sulphur bacteria, without measuring them directly. From our
model of the bacterial modes and their interactions, we have confirmed our probing
scheme. Finally, other applications were presented in Chapter 8. This includes,
among others, the revelation of a non-classical property of a mechanical mirror that
is mediating interactions between two otherwise non-interacting cavity light modes.
130
9.2 Future work
9.2.1 Entanglement bound for indirect continuous interac-
tions
In Chapter 3, we considered the indirect interaction setting where the interactions
are continuous. An illustration can be seen in Fig. 3.1 where an ancillary object C
is mediating interactions between two principal objects A and B. In this case, we
have proven that classical mediator (zero discord DAB|C at all times) cannot increase
quantum entanglement, i.e., EA:BC (τ ) ≤ EA:BC (0). See Theorem 3.3.1 for details.
However, we note that having positive discord is not sufficient for entanglement gain.
Moreover, we often deal with imperfections in real situations. It has been shown
that states with zero discord are impossible to prepare in experiments, i.e., small
perturbations will drive a zero-discord state into a state having positive, albeit small,
discord [192]. As an example, let us consider a state of the form
where p stands for probability, {ρAB , ρ0AB } are states of system AB, and {|αi , |α0 i}
are coherent states of C. Here we note that the states |αi and |α0 i are non-
orthogonal. In the limit where the expectation values of position, hα|x|αi and
hα0 |x|α0 i are far apart, the states will be effectively orthogonal. Theoretically, how-
ever, quantum discord DAB|C of the state in (9.1) will never be exactly zero. Yet, one
expects that not much entanglement can be gained with the effectively orthogonal
states. This calls for a bound on entanglement gain in terms of quantum discord
DAB|C , or other related quantities.1 Indeed, an observation of finite entanglement
gain would certify that useful quantum discord (i.e., finite) formed during the dy-
namics.
We have performed preliminary investigation towards this direction. First, we
note from the example in Fig. 3.2 that entanglement increment in the partition
A : BC within ωt = π/4 (EA:BC = 1 at this time), is not bounded by the maximum
discord (≈ 0.81) or the average discord (≈ 0.38). For a special class of initial states
and Hamiltonian, we have obtained a simple bound. In particular, the gain of
entanglement (as quantified by REE) is always bounded by entanglement with the
mediator, i.e.,
|EA:BC (τ ) − EA:BC (0)| ≤ EAB:C (τ ). (9.2)
131
[HAC , HBC ] = 0, and pure decoupled states as the initial condition [44]. The present
aim is therefore to generalise this result, taking into account more general states and
Hamiltonians with non-commuting components. One could also make it experimen-
tally friendly by considering incoherent interactions with environments.
where η is the to-be-proven bound that is independent of initial state. One can then
think of the Trotterized process, as depicted in Fig. 9.1. For a general evolution,
suppose one writes the dynamical operator as
where the superscript denotes the number of sequences where C first interacts with
A and then with B. Next, one can apply the bound of Eq. (9.3) to the sequences
above and obtain
where τ = n∆t. In real situation, the observation of correlation gain that is ex-
ceeding certain n, would imply that the corresponding evolution operator cannot
be modelled by n sequences of decomposable operator of the form ΛBC ΛAC . We
propose this as characterisation of the strength of non-decomposability of the actual
132
dynamical operator Λ. In particular, the higher the observed correlation QA:B (τ )
the stronger the non-decomposability in the dynamics.
Figure 9.1: Illustration of Trotterized dynamics. The evolution of the whole system
consists of sequences of interactions between the mediator object C with A, followed
by C with B, each for a time ∆t.
We now provide a simple example of the bound in Eq. (9.3). Consider the mutual
information as the correlation quantifier. The corresponding bound is given by the
following Lemma.
Lemma 9.2.1. For a tripartite system with a dynamical operator having the form
ΛBC ΛAC , the mutual information follows
where dX is the dimension of object X and we have assumed dA , dB > dC for sim-
plicity.
Proof. Let us begin with the monotonicity of mutual information under local oper-
ations to arrive at
IA:B (t) ≤ IA:BC (t) = IA:BC (ΛBC ΛAC [ρ(0)]) ≤ IA:BC (ΛAC [ρ(0)]). (9.8)
0
IA:B (t) − IA:B (0) ≤ IA:BC − IA:B (0)
0 0 0
= IAC:B − IB:C + IA:C − IA:B (0)
0 0
≤ IAC:B (0) − IA:B (0) − IB:C + IA:C
0
≤ IB:C|A (0) + IA:C
≤ 4 log2 (dC ), (9.9)
133
where the prime symbol denotes the state ΛAC [ρ(0)]. We justify the steps as follows.
Eq. (9.8) has been used in the first line. The second line is apparent by utilising the
definition of mutual information between two parties, i.e., IX:Y = SX + SY − SXY ,
where for example SX is the von Neumann entropy of object X. The third line uses
the monotonicity property under the operation ΛAC that is local in the partition
AC : B. We have used the definition of conditional entropy and the positivity of
mutual information in the fourth line. The mutual information between B and C
(conditioned on A) is bounded by 2 log2 (dC ). Note that we assume dB > dC . This is
also true for mutual information between A and C, which completes the proof.
To demonstrate violations of the bound for certain n, one can make use of the setup
described in Section 8.3, where two cavity fields are interacting via a two-level atom.
Our simulations show that high correlations can be obtained by having pure states
with more photons as the initial condition. Future work is aimed at reducing the
bound in Eq. (9.9) as our simulations show the gain of mutual information being
bounded by only 2 log2 (dC ). Also, one might consider other correlation quantifiers.
134
entanglement. This protocol is in general hard to implement due to the requirement
of maximum entanglement. We aim at making this scheme also experimentally
friendly. A potential direction would be to consider an entanglement bound for a
given purity. If the entanglement EA:B reaches a higher value, this would imply the
presence of a quantum mediator.
π
T ≥ . (9.11)
2
Therefore, in this scenario, it follows that T ≥ 2Γdi for the case of three qubits. A
simple example saturating this limit is given as follows
|ψ(0)i = |110i ,
~Ω
H = √ σA+ ⊗ σC− + σA− ⊗ σC+ + σB+ ⊗ σC− + σB− ⊗ σC+ ,
(9.12)
2
+(−)
where σX is the raising (lowering) operator of object X. Indeed, maximum en-
tanglement (negativity NA:B = 0.5) is reached in T = Ωt = π/2.
135
potential. The gravitational energy is taken as
Gm2
Hg = − (9.13)
(L + xB − xA )
Gm2 (xA − xB ) (xA − xB )2
' − 1+ + , (9.14)
L L L2
136
Appendix A
Trotter expansion
A simple derivation of the Trotter formula [193] is provided in Theorem A.0.1 below.
Theorem A.0.1. For arbitrary square matrices A and B, the following holds
A B
n
A+B
e = lim e e
n n . (A.1)
n→∞
1 1 1
ln eA eB = A + B + [A, B] + [A, [A, B]] + [B, [B, A]] + · · ·
(A.2)
2 12 12
137
Appendix B
Entanglement localisation:
Prescription
where [HC1 , HC2 ] = 0, i.e., these components share the same eigenbasis. Now we
take a quantum–classical initial state in the partition AB : C,
X
ρ(0) = pc ρAB|c ⊗ |ci hc| , (B.2)
c
where {|ci} form the common eigenbasis of HC1 and HC2 . Let us also use, as the
eigenvalues (assumed dimensionless), E1c and E2c respectively.
The state after a short time ∆t reads
i∆t i∆t
ρ(∆t) = 1− H ρ(0) 1 + H
~ ~
X i∆t c i∆t c
= pc ρAB|c − E1 HA ρAB|c − E HB ρAB|c
c
~ ~ 2
i∆t c i∆t c
+ ρAB|c E1 HA + ρAB|c E2 HB ⊗ |ci hc|
~ ~
X i∆t
[(E1c HA + E2c HB ), ρAB|c ] ⊗ |ci hc|
= pc ρAB|c −
c
~
X
= pc ρ∆t
AB|c ⊗ |ci hc| , (B.3)
c
139
where we have used ρ∆t AB|c to denote the short coherent evolution of ρAB|c with
weighted local Hamiltonians E1c HA + E2c HB . Note that we have ignored terms pro-
portional to (∆t)2 . The evolution for a time t simply requires successive applications
of the above process, i.e., ρ(t) = c pc ρtAB|c ⊗ |ci hc|.
P
where we have used the flags condition [40] on the quantum–classical state in the
first and last equality, and the monotonicity of entanglement under local unitary
operations in the second equality.
Now let us consider entanglement between A and B. In particular we have
X X X
EA:B (t) = EA:B pc ρtAB|c ≤ pc EA:B (ρtAB|c ) = pc EA:B (ρAB|c ), (B.5)
c c c
where we have used the convexity of entanglement for the inequality and the mono-
tonicity for the last equality. One immediately realises that, in order to have entan-
glement gain in the partition A : B with this method, there has to be entanglement
already present initially, i.e., in the state ρAB|c .
Note that the necessary requirements above are satisfied in the example given in
Chapter 3, where H = (σAx ⊗ 1 ⊗ σCx + 1 ⊗ σBx ⊗ σCx )~ω, and the initial state reads
1
2
|ψ+ i hψ+ | ⊗ |+i h+| + 12 |φ+ i hφ+ | ⊗ |−i h−|. In particular, the classical basis of sys-
tem C is the eigenbasis of the Pauli matrix σx and there is maximum entanglement
in the initial states {|ψ+ i , |φ+ i}. Furthermore, the evolution of the states for a time
τ = π/8ω follows
where ρmax is, up to a universal phase factor, a density matrix of a pure maximally
entangled state
1
√ (|0i |y+ i + |1i |y− i) , (B.7)
2
with {|y+ i , |y− i} being the eigenbasis of the σy Pauli matrix. Therefore, one can see
that EA:B (0) = 0 and EA:B (τ ) = 1, where we have used REE as the entanglement
quantifier.
140
Bibliography
141
[12] L. Mišta Jr. and N. Korolkova. Improving continuous-variable entanglement
distribution by separable states. Phys. Rev. A, 80:032310, 2009.
[13] L. Mišta Jr. Entanglement sharing with separable states. Phys. Rev. A, 87:
062326, 2013.
[15] X.-D. Yang, A.-M. Wang, X.-S. Ma, F. Xu, H. You, and W.-Q. Niu. Experi-
mental Creation of Entanglement Using Separable States. Chin. Phys. Lett.,
22:279, 2005.
142
[24] K. Modi, A. Brodutch, H. Cable, T. Paterek, and V. Vedral. The classical-
quantum boundary for correlations: Discord and related measures. Rev. Mod.
Phys., 84:1655, 2012.
[26] W. K. Wootters. Statistical distance and Hilbert space. Phys. Rev. D, 23:357,
1981.
[27] A. Uhlmann. The Metric of Bures and the Geometric Phase, in Groups and
Related Topics (Springer, Netherlands, Dordrecht, 1992), pp. 267-274.
[28] C. Helstrom. Quantum detection and estimation theory. J. Stat. Phys., 1:231,
1969.
[30] A. Peres. Separability Criterion for Density Matrices. Phys. Rev. Lett., 77:
1413, 1996.
[32] J. Lee and M. S. Kim. Entanglement Teleportation via Werner States. Phys.
Rev. Lett., 84:4236, 2000.
143
[39] M. Horodecki, P. Horodecki, R. Horodecki, J. Oppenheim, A. Sen, U. Sen, and
B. Synak-Radtke. Local versus nonlocal information in quantum-information
theory: Formalism and phenomena. Phys. Rev. A, 71:062307, 2005.
144
[52] T. A. Baart, T. Fujita, C. Reichl, W. Wegscheider, and L. M. K. Vandersypen.
Coherent spin-exchange via a quantum mediator. Nat. Nanotechnol., 12:26,
2017.
[58] A. Peres. Quantum Theory: Concepts and Methods. Kluwer Academic, Dor-
drecht, Netherlands, 2002.
[62] A. Streltsov, G. Adesso, M. Piani, and D. Bruß. Are General Quantum Cor-
relations Monogamous? Phys. Rev. Lett., 109:050503, 2012.
[64] M. Fannes. A continuity property of the entropy density for spin lattice sys-
tems. Commun. Math. Phys., 31:291, 1973.
145
[65] X. Yuan, S. M. Assad, J. Thompson, J. Y. Haw, V. Vedral, T. C. Ralph, P. K.
Lam, C. Weedbrook, and M. Gu. Replicating the benefits of Deutschian closed
timelike curves without breaking causality. npj Quantum Inf., 1:15007, 2015.
[67] M. Piani. Problem with geometric discord. Phys. Rev. A, 86:034101, 2012.
[71] J. Uffink. The rate of evolution of a quantum state. Am. J. Phys., 61:935,
1993.
[74] S. Deffner and E. Lutz. Energy-time uncertainty relation for driven quantum
systems. J. Phys. A, 46:335302, 2013.
[75] Y.-J. Zhang, W. Han, Y.-J. Xia, J.-P. Cao, and H. Fan. Quantum speed limit
for arbitrary initial states. Sci. Rep., 4:4890, 2014.
[76] D. Mondal, C. Datta, and S. Sazim. Quantum coherence sets the quantum
speed limit for mixed states. Phys. Lett. A, 380:689, 2016.
[80] S. Lloyd. Computational Capacity of the Universe. Phys. Rev. Lett., 88:
237901, 2002.
146
[81] F. C. Binder, S. Vinjanampathy, K. Modi, and J. Goold. Quantacell: powerful
charging of quantum batteries. New J. Phys., 17:075015, 2015.
[82] F. C. Binder. Work, Heat, and Power of Quantum Processes. PhD The-
sis, University of Oxford, 2016.
[85] M. Okuyama and M. Ohzeki. Quantum Speed Limit is Not Quantum. Phys.
Rev. Lett., 120:070402, 2018.
[90] E. Study. Kürzeste Wege im komplexen Gebiet. Math. Ann., 60:321, 1905.
147
[96] V. V. Nesvizhevsky, H. G. Börner, A. K. Petukhov, H. Abele, S. Baeßler,
F. J. Rueß, T. Stöferle, A. Westphal, A. M. Gagarski, G. A. Petrov, and
A. V. Strelkov. Quantum states of neutrons in the Earth’s gravitational field.
Nature, 415:297, 2002.
[97] R. V. Pound and G. A. Rebka Jr. Apparent Weight of Photons. Phys. Rev.
Lett., 4:337, 1960.
[99] D. N. Page and C. D. Geilker. Indirect Evidence for Quantum Gravity. Phys.
Rev. Lett., 47:979, 1981.
[102] C. Anastopoulos and B.-L. Hu. Probing a gravitational cat state. Class.
Quantum Grav., 32:165022, 2015.
[105] C. Marletto and V. Vedral. Witness gravity’s quantum side in the lab. Nature,
547:156, 2017.
[106] A. Belenchia et al. Quantum superposition of massive objects and the quan-
tization of gravity. Phys. Rev. D, 98:126009, 2018.
148
[110] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt. Cavity optomechanics.
Rev. Mod. Phys., 86:1391, 2014.
[118] R. Benguria and M. Kac. Quantum Langevin Equation. Phys. Rev. Lett., 46:
1, 1981.
[120] R. F. Werner and M. M. Wolf. Bound Entangled Gaussian States. Phys. Rev.
Lett., 86:3658, 2001.
149
[123] N. Kiesel, F. Blaser, U. Delić, D. Grass, R. Kaltenbaek, and M. Aspelmeyer.
Cavity cooling of an optically levitated submicron particle. Proc. Natl. Acad.
Sci. U.S.A., 110:14180, 2013.
[131] R. Schnabel. Squeezed states of light and their applications in laser interfer-
ometers. Phys. Rep., 684:1, 2017.
150
[137] S. Gerlich, S. Eibenberger, M. Tomandl, S. Nimmrichter, K. Hornberger, P. J.
Fagan, J. Tüxen, M. Mayor, and M. Arndt. Quantum interference of large
organic molecules. Nat. Commun., 2:263, 2011.
[147] C.-M. Li, N. Lambert, Y.-N. Chen, G.-Y. Chen, and F. Nori. Witnessing
Quantum Coherence: from solid-state to biological systems. Sci. Rep., 2:885,
2012.
151
[148] R. Hildner, D. Brinks, J. B. Nieder, R. J. Cogdell, and N. F. Hulst. Quantum
Coherent Energy Transfer over Varying Pathways in Single Light-Harvesting
Complexes. Science, 340:1448, 2013.
[149] N. Lambert, Y.-N. Chen, Y.-C. Cheng, C.-M. Li, G.-Y. Chen, and F. Nori.
Quantum biology. Nat. Phys., 9:10, 2013.
[157] C.-H. Bai, D.-Y. Wang, H.-F. Wang, A.-D. Zhu, and S. Zhang. Robust en-
tanglement between a movable mirror and atomic ensemble and entanglement
transfer in coupled optomechanical system. Sci. Rep., 6:33404, 2016.
[158] C. W. Gardiner and P. Zoller. Quantum noise. Springer Science & Business
Media, Berlin, 2004.
[159] D. F. Walls and G. J. Milburn. Quantum optics. Springer Science & Business
Media, Berlin, 2007.
152
[160] D. Bajoni. Polariton lasers. Hybrid light–matter lasers without inversion. J.
Phys. D: Appl. Phys., 45:313001, 2012.
[162] C. Houssier and K. Sauer. Circular dichroism and magnetic circular dichroism
of the chlorophyll and protochlorophyll pigments. J. Am. Chem. Soc., 92:779,
1970.
[168] H.-P. Breuer and F. Petruccione. The theory of open quantum systems. Oxford
University Press, 2002.
[169] E.-M. Laine, J. Piilo, and H.-P. Breuer. Witness for initial system-environment
correlations in open-system dynamics. Europhys. Lett., 92:60010, 2011.
[170] A. Smirne, H.-P. Breuer, J. Piilo, and B. Vacchini. Initial correlations in open-
systems dynamics: The Jaynes-Cummings model. Phys. Rev. A, 82:062114,
2010.
[171] J. Dajka and J. Luczka. Distance growth of quantum states due to initial
system-environment correlations. Phys. Rev. A, 82:012341, 2010.
153
[172] J. Dajka, J. Luczka, and P. Hänggi. Distance between quantum states in the
presence of initial qubit-environment correlations: A comparative study. Phys.
Rev. A, 84:032120, 2011.
[177] C.-F. Li, J.-S. Tang, Y.-L. Li, and G.-C. Guo. Experimentally witnessing
the initial correlation between an open quantum system and its environment.
Phys. Rev. A, 83:064102, 2011.
[180] M. Gessner and H.-P. Breuer. Local witness for bipartite quantum discord.
Phys. Rev. A, 87:042107, 2013.
[182] J.-S. Tang, Y.-T. Wang, G. Chen, Y. Zou, C.-F. Li, G.-C. Guo, Y. Yu, M.-F.
Li, G.-W. Zha, H.-Q. Ni, Z.-C. Niu, M. Gressner, and H.-P. Breuer. Experi-
mental detection of polarization-frequency quantum correlations in a photonic
quantum channel by local operations. Optica, 2:1014, 2015.
154
[183] S. Cialdi, A. Smirne, M. G. A. Paris, S. Olivares, and B. Vacchini. Two-step
procedure to discriminate discordant from classical correlated or factorized
states. Phys. Rev. A, 90:050301, 2014.
[184] A. Messina. A single atom-based generation of Bell states of two cavities. Eur.
Phys. J. D, 18:379, 2002.
[186] I. Devetak and A. Winter. Distillation of secret key and entanglement from
quantum states. Proc. R. Soc. A, 461:207, 2005.
155