Macroscopic quantum states: measures, fragility and implementations

Florian Fröwis1,∗ Pavel Sekatski2,∗ Wolfgang Dür2,† Nicolas Gisin1, and Nicolas Sangouard3 1Group of Applied Physics - Université de Genève - CH-1211 Geneva - Switzerland 2Institut für Theoretische Physik - Universität Innsbruck - Technikerstraße 21a - A-6020 Innsbruck - Austria 3Quantum Optics Theory Group - Department of Physics - University of Basel - CH-4056 Basel - Switzerland

(Dated: June 20, 2018)

Large-scale quantum effects have always played an important role in the foundations of quantum theory. With recent experimental progress and the aspiration for quantum en- hanced applications, the interest in macroscopic quantum effects has been reinforced. In this review, we critically analyze and discuss measures aiming to quantify various aspects of macroscopic quantumness. We survey recent results on the difficulties and prospects to create, maintain and detect macroscopic quantum states. The role of macroscopic quantum states in foundational questions as well as practical applications is outlined. Finally, we present past and on-going experimental advances aiming to generate and observe macroscopic quantum states.

CONTENTS 2. Linking measures for photons and spins 28 F. Connection to multipartite entanglement 28 I. Introduction 2 G. Connection to resource theory of 28 A. Defining macroscopic quantumness: not an easy task 2 1. Resource theory for macroscopic coherence 29 B. Motivation 4 2. Free operations in proposed measures 29 C. Terminology 5 H. How to determine the effective size in experiments 30 D. Physical systems 5 1. Macroscopic distinctness 30 1. Spin ensemble 5 2. Macroscopic coherence 30 2. Photonic systems 7 3. Measuring loss of coherence 31 3. Massive systems 7 4. Correlations for index q 31 4. Superconducting systems 8 I. Summary and conclusion 31 E. Structure of the review and reading guide 8 III. Limits for observing quantum properties in macroscopic II. Measures for macroscopic superpositions and quantum states 32 states 8 A. Maintaining macroscopic quantum states 32 A. Summary of measures 8 1. Decoherence 32 1. Leggett (1980, 2002) 9 2. Fragility of macroscopic quantum superpositions 33 2. Dür et al. (2002) 10 3. Sensitivity of macroscopic quantum states 34 3. Shimizu and Miyadera (2002) and followups 10 B. Measuring and detecting macroscopic quantum states 35 4. Björk and Mana (2004) 11 1. Coarse-graining and control of measurements 35 5. Cavalcanti and Reid (2006, 2008) 12 2. Reference frames and the size of measurement 6. Korsbakken et al. (2007) 12 apparatus 36 7. Marquardt et al. (2008) 13 C. Preparation of macroscopic quantum states 36 8. Lee and Jeong (2011) and Park et al. (2016) 13 D. Further limitations and counter-strategies 37 9. Fröwis and Dür (2012b) and followups 14 1. Certifiability of large-scale quantum states 37 10. Nimmrichter and Hornberger (2013) 15 2. Counter-strategies against noise 37 11. Sekatski et al. (2014c, 2017b) 15 3. Counter-strategies against coarse-graining 38 12. Laghaout et al. (2015) 16 4. Encoded macroscopic quantum states 38 13. Yadin and Vedral (2015) 17 E. Summary 39 14. Kwon et al. (2017) 17 B. Examples 18 IV. Potentials of macroscopic quantum states 39 1. Spin ensemble 18 A. Probing the limits of quantum theory 39 2. Photonic systems 21 1. Macrorealism and Leggett-Garg-like inequalities 39 C. Classifications of measures 23 2. Validity of at large scales - 1. Mechanisms to break unitary equivalence 23 Collapse models 40 2. Goals of the measures 23 B. Quantum metrology 41 D. Structure of applied states 24 C. Quantum computing 41 arXiv:1706.06173v2 [quant-ph] 19 Jun 2018 1. Macroscopic superpositions 25 1. MBQC, entanglement and macroscopicity 41 2. Macroscopic quantum states 25 2. States occurring in quantum computation and 3. Connections between some measures 26 metrology 42 E. Application to various physical setups 27 3. Quantum phase transitions 43 1. Implications of large variance 27 D. Summary 43

V. Implementations 43 A. Photonic experiments 43 1. Optical photons 43 ∗ These authors contributed equally. 2. Microwave photons 46 † Electronic address: [email protected] B. Spin experiments 47 2

1. Spins with individual addressing 47 degrees of freedom. In all cases, a reduction of complex- 2. Spins with collective addressing 47 ity, often to a single degree of freedom, is considered in C. Experiments with massive systems 47 order to keep essential properties while making it theo- 1. Matter interferometry 48 2. Quantum optomechanics 49 retically tractable and bringing it closer to experimental D. Superconducting quantum interference devices 50 reality. While on an abstract level states might be iso- E. Comparing the size of states describing photonic, spins morphic, there is no consensus whether a superposition and massive systems 50 state with different spin values or different positions of 1. Discussions on the size of flux states in SQUID systems 51 the wave packet can equally well be called a macroscopic 2. Comparing the size of observed states 52 superposition. The latter for instance is affected by grav- F. Summary 53 itational collapse and allows to test proposed modifica- VI. Discussion and outlook 53 tions of quantum mechanics, while the former is not. An- other issue is how number, distance and mass Acknowledgments. 54 enter in the assessment. How can we compare the spa- References 54 tial superposition of a single atom being 1 m apart to a Bose-Einstein condensate with one million atoms where the center-of-mass is separated by 1 µm? To address I. INTRODUCTION these issues, one needs to formalize the observation that “dead and alive” are more than two orthogonal states in With recent progresses in experimental physics, it is a Hilbert space but are somehow “macroscopically dis- nowadays possible to investigate quantum effects such tinct”. Furthermore, it is unclear how to take into ac- as interference and entanglement in larger and larger count loss of coherence (i.e., purity). Is there a way to systems. Experimentalists bring mechanical oscillators deal with reduced visibility when scaling-up the system to the quantum regime, set interferometers with single size? Finally, the quest does not end with macroscopic giant , produce superposition states with many superpositions of two states. Generalizations to arbitrary atoms, photons and high superconducting currents, or quantum states including mixed states are important to reveal entanglement in many-body systems.1 Starting further abstract the problem and to apply the theoretical with Leggett (1980), many physicists came up with concepts directly to experiments. measures to compare these experiments, that is, to In this review we summarize and discuss proposals to quantify how macroscopic and quantum a state is. Such measure quantum states (or entire experiments) concern- measures allow one to characterize sets of states and ing some aspect of macroscopic quantumness. We do not to study systematically the requirements to observe only aim to give a technical summary of the measures, the quantum features of macroscopic states. From a but also to facilitate a discussion of the motivation and fundamental point of view, this helps to gain insight into intuition behind them, as well as relations between the the quantum-to-classical transition and to investigate different proposals. We then discuss fundamental diffi- the limits of quantum theory. From a more applied per- culties and limitations to prepare, maintain and certify spective, this is useful to reveal general mechanisms for macroscopic quantum states, and briefly mention poten- quantum enhancement in applications such as quantum tial applications in foundations of quantum mechanics as computing and metrology. well as quantum metrology and quantum computation. Finally, we evaluate the current status of experimental It should be emphasized, though, that identifying key progress by reviewing experiments with different systems, features of macroscopic quantumness is highly contro- and applying different measures and proposals to assess versial. Intuitively, any approach should distinguish a their level of macroscopic quantumness. In the remainder genuine macroscopic quantum effect from accumulated of the introduction, we specify more precisely the scope microscopic effects. However, already the precise mean- of the review, clarify the motivation and the terminology ing of these and similar words is unclear and disputed, and mention the structure of this paper. also because they are heavily loaded with emotions and prejudice. The example from Schrödinger (1935) of a cat in superposition of being dead and alive is a paradigmatic A. Defining macroscopic quantumness: not an easy task starting point for many considerations and experiments with different physical systems. But we face many open It is often claimed that quantum mechanics is one of questions. The first issue involves the role of the physical the most successful theories in physics. The basis of system (atoms, , photons, etc.) and the different this assertion is its passing of all experimental tests so far. This is certainly true in the microscopic realm. Here, we are interested in large systems, that is, ex- periments involving many atoms, photons or electrons. 1 See Sec. V for references and further details. There, the experimental evidence and its interpretation 3 are less clear. As mentioned by several physicists such cause the unease of Schrödinger and many others? We as Leggett (1980), many well-established large-scale ex- now list some aspects frequently appearing in the litera- periments can be seen as a macroscopic accumulation of ture that is discussed in this review. microscopic quantum effects. As an example, the genuine (1) The superposition principle is one of the most quantum effect of Cooper pair formation in the BCS the- straightforward illustrations of the drastic difference be- ory of superconductivity is a two- problem and tween classical and quantum physics. For any two pos- many-body correlations are not necessary to observe su- sible quantum states |Ai and |Di, the superposition perconductivity on human scales (Leggett, 1980). Hence, |Ai ± |Di is also a valid quantum state. This is a well as opposed to microscopic quantum effect, one could de- accepted fact for microscopic systems or even for macro- fine a macroscopic quantum effect as a situation in which scopic systems if |Ai and |Di are “hardly” distinguishable. the experimental evidence excludes a model based on an However, the superposition principle appears paradoxical accumulated microscopic quantum effect (regarding ter- when |Ai and |Di represent states that are drastically dif- minology, see Sec. I.C). However, this definition does not ferent or irreconcilable in classical physics, like an object seem to be restrictive enough, as it can be fulfilled in being here and there or a cat being dead and alive. This is drastically different situations, some of which defy our a key feature in Schrödinger’s example, as the two super- intuition. Let us illustrate this point with the two fol- posed states |Alivei and |Deadi are not only orthogonal lowing examples. but also macroscopically distinct, using a term coined by Consider a large atomic ensemble that was prepared in Leggett (1980). the ground state and coherently absorbed a single pho- (2) The number of biological cells in the cat is in ton. The resulting atomic state is in a so-called Dicke the order of trillions; the number of atoms in the or- state or W state (Dicke, 1954; Dür et al., 2000), that is, der of 26. But it is not only the bare number of con- a coherent superposition of all states where exactly one 10 stituents. In the biological cat many degrees of freedom atom has absorbed the photon. This state is genuinely (or, modes) are “acitve”, that is, are accessible via inter- multipartite entangled, that is, nonseparable for any bi- actions between them. This leads to an enormous com- partition of the ensemble. According to the previous def- plexity within an unthinkably large Hilbert space. Some inition, genuine multipartite entanglement between all physicists see the presence of this large number of degrees atoms constitutes a genuine macroscopic quantum effect. of freedom (and not only a large number of ) as In particular, this has direct observable consequences if a necessary perquisite for quantum systems to be called we consider the spontaneous reemission of the photon. macroscopically quantum (Shimizu and Miyadera, 2002). The coherent phase relation between the atoms leads to Others tend to drop this aspect when going to realistic a temporally and directionally well-defined emission of systems like Bose-Einstein condensates or superconduct- the photon (Dicke, 1954; Duan et al., 2001; Scully et al., ing devices (see Sec. V), where only one or few modes 2006). On the other side, if one divides the ensemble effectively exist. into several groups without a shared phase relation (and thus losing the entanglement between the groups), the (3) Finally, an aspect that is often emphasized is the spatial emission pattern progressively becomes isotropic. micro-macro entanglement between the atom and the cat. Yet this quantum effect has a completely different char- It is a priori not clear what happens if –in a bipartite acter than the following example. scenario– the system size of one party is increased. Sev- In his seminal paper, Schrödinger (1935) sketched the eral experiments in quantum optics and optomechanics famous cat paradox, where the total system consists of aim to realize this aspect of Schrödinger’s cat (see dis- an atom with two levels (ground |gi and excited |ei) and cussion later). Also some of the proposals presented in a cat, which is coupled to the atom via a mechanism this review reflect this idea. U that kills the cat whenever the excited atom decays In this review, we will focus on recent contributions to the ground state. By applying the linearity of quan- that aim for a more systematic approach to the topic of tum mechanics even to macroscopic scales, Schrödinger macroscopic quantum mechanics. In many papers, the argued that the coherent superposition of the atom be- Schrödinger-cat paradox serves as a starting point to for- ing excited and already decayed results in a micro-macro malize concepts such as macroscopic quantum effects or entanglement between atom and cat macroscopic quantum states. There is a broad agreement in the literature that coherence between macroscopically U |Alivei ⊗ (|ei + |gi) → (1) distinct states (1) has to be necessarily present in exper- |ei ⊗ |Alivei + |gi ⊗ |Deadi . (2) iments that aim to mimic the Schrödinger’s cat example. In contrast, high complexity in terms of many accessi- This paradox challenges our world view much more than ble degrees of freedom (2) is, for many but not for all the example of the absorbed single photon, even though authors, neither necessary nor sufficient. Likewise, the the size of the atomic ensemble can be truly macroscopic. micro-macro entanglement (3) plays only a minor role in But what are the essential features of this example that the papers discussed here. 4

B. Motivation of Schrödinger cat states, such that quantum laws have to be supplemented with an explicit collapse mechanism. The theoretical and experimental study of macroscopic This is commonly done by introducing a stochastic ex- quantum systems is motivated by a wide range of inter- tension of the Schrödinger equation. This extension ba- ests. Many open questions in the foundations of quan- sically does not affect any microscopic quantum system tum mechanics touch on quantumness on large scales, composed of a few atoms. Large masses and distances, the quantum-to-classical transition and the measurement however, lead to an efficient collapse whenever the wave problem. This includes comparisons of standard quan- function is widely spread over a characteristic amount tum mechanics against potential modifications relevant of time (see Arndt and Hornberger (2014); Bassi et al. on large scales but also a better understanding of quan- (2013) and references therein). Collapse models are a way tum mechanics itself. The latter is expected to have a to stay within an extended quantum theory and at the cross fertilization with applications of quantum mechan- same time avoiding paradoxes à la Schrödinger. Hence, ics that are particularly interesting when performed with the experimental verification of superposing two macro- large quantum systems (e.g., for quantum computation scopically distinct states (in position space) is a typical and metrology). Let us discuss some of these points in way to falsify such modifications (however, see the dis- more depth. cussion in Sec. II.C.2 about “more suited” states to test Historically, Schrödinger’s motivation for his para- collapse models under realistic conditions). dox was to demonstrate the interpretational problems Along the same lines an additional motivation to study of what he calls the “blurriness” of the wave function macroscopically distinct states stems from the measure- (Schrödinger, 1935). He argued that at a scale of a ra- ment problem, that is, the appearance of a single mea- dioactive nucleus one might be able to accept that the surement outcome irreversibly chosen among all possible state of a quantum system cannot be described by a outcomes (Schlosshauer, 2005). In particular, in stan- well-defined collection of properties like position, mo- dard quantum mechanics (Copenhagen interpretation) mentum, excitation level etc. The speculative reason is the problem is explicitly solved by introducing the mea- that we anyway cannot directly access these small scales surement postulate – the ability of the measurement ap- and everyday intuition breaks down. However –assuming paratus to collapse the wave function. However, this pos- full validity of quantum mechanics also at large scales– tulate also has provoked controversial discussions, as it one can easily construct examples where the initial mi- is not a priori clear what qualifies to be a measurement croscopic blurriness is translated to human scales. In apparatus. It is supposed to be a macroscopic device Schrödinger’s example, it is the coherent superposition obeying the laws of classical physics, but down the line of a cat that is dead and alive correlated with a radioac- any such device is made of atoms – quantum systems tive substance being decayed and excited. We can easily with reversible unitary dynamics. Where should this determine the basic vital function of a cat, which prevents boundary between microscopic and macroscopic (quan- us to accept a “blurred model” as an accurate picture of tum and classical), postulated by Copenhagen, be found? reality. On the opposite side, the many-world interpretation de- Very generally, there is a natural motivation to formal- prives the measurement process from such a special role ize this intuition and essential aspects of Schrödinger’s (Everett, 1957; Wheeler, 1957). In this theory, a mea- example into a mathematically solid and abstract tool. surement is just a particular case of unitary dynam- This would allow us to benchmark experimental progress, ics in which the measurement apparatus entangles with tell us at which scales we did verify quantum laws, and the system. Hence, all possible outcomes, in fact, oc- maybe get closer to the original question: Are quantum cur, as the state of the system and the apparatus after laws, such as the superposition principle, valid or at least the coupling is a superposition of all possible outcomes. observable at all scales? In fact, one might hear different Nevertheless, we, conscious beings, have the impression answers to this question, which all provide motivation to to live in a world where only a single outcome occurs. study macroscopic quantumness. This discrepancy also provoked many discussions, sum- The common intuition is that, while nature might al- marized in the famous Wigner’s friend paradox (Wigner, low for quantum effects on macroscopic scale, it makes 1961). Schrödinger’s thought experiment is of course at them practically impossible to observe. This is due to the heart of this debate (Leggett, 2002; Schlosshauer, technical limitations that forbid one to perfectly isolate 2005). a system from its environment and to perform measure- Besides the interest in the transition from quantum ments with unlimited precision. This leads to an effec- mechanics to our classical world, effective or not, it is tive quantum-to-classical transition, which can be ideally important to understand the structure of quantum me- derived from the quantum laws themselves (Joos, 2003; chanics itself when applied to large systems. In entangle- Zurek, 2003b). ment theory, the structure of multipartite entanglement One might also take a more radical attitude on this becomes richer and more complex as the number of par- question, saying that nature prohibits even the existence ties increases. A similar behavior is expected for other as- 5 pects of quantum mechanics when brought to large scales. macroscopic quantumness is the goal of most of the con- Although we here review papers that are often in the tributions discussed in this review. This cannot be cap- vicinity of Schrödinger’s cat example, this can be seen as tured by a single characteristic trait but the problem is just one out of many interesting aspects of macroscopic expected to be multidimensional. Thus, a peaceful coex- quantumness. In general, the systematic study leads to istence of different ideas is likely to be possible. Very gen- new insight on quantum effects, new proposals for exper- erally, different concepts could be referred to as macro- iments and constraints on the experimental requirements scopic quantumness. In this review however, macroscopic to prepare, maintain and observe macroscopic quantum- quantumness means quantum coherence between macro- ness. scopically distinct states inspired by point (1) in Sec. I.A. Such a broad account might give general insight also We emphasize that the two aspects –macroscopicity and useful for applications of quantum mechanics. In quan- quantumness– are sometimes separately studied, some- tum computation and quantum metrology, for example, times they are combined to a single concept. Being one the performance of algorithms and protocols is often mea- aspect of macroscopic quantumness, quantum coherence sured as a function of the system size . Quantum advan- between far distant parts in the spectrum of a given ob- tage is particularly striking for large system sizes. Con- servable is called macroscopic coherence. nections made between foundations and applications can In many papers, the authors are interested in finding lead to a new point of view and additional understanding a function f that assigns a nonnegative number f(ρ) ≥ 0 of the mechanisms (e.g., see Sec. IV.C.2). to a quantum state ρ. This number should ideally reflect the degree of macroscopic quantumness of ρ. In this re- view, we call f a measure (of macroscopic quantumness) C. Terminology and f(ρ) is called effective size (or simply size) of ρ. For a unified notation, we name the measures by the authors A primary obstacle when discussing the present topic who first proposed them, which is also used to title the is the used terminology. First, as already mentioned be- subsections in Sec. II.A. We generally keep the mathe- fore, the topic is strongly filled with emotions and pre- matical symbols for the measures as introduced in the conceptions. Second, different authors use words like original papers (see table I). macroscopic or large in different ways. Hence, on the one hand we prefer to avoid loaded terminology such as Schrödinger-cat states, but on the other we cannot com- D. Physical systems pletely ignore the common terminology. Let us clarify how we here use some frequently used words. Macroscopic quantumness is not restricted to a single In this review, macroscopic2 is a synonym for large. All physical realization but can be studied for many differ- physical systems considered here necessarily consist of a ent degrees of freedom (see Caldeira (2014); Chou et al. large number of microscopic constituents (e.g., atoms, (2011); Leggett (2002) for an introduction to the physics electrons or photons). This is referred to as the system of macroscopic quantum phenomena). In this review, we size or number of constituents. However, the system size treat few “canonical” systems for which we introduce the is not necessarily in the order of 1023 or even close to it. terminology and the notation in this section. The only While it could be misleading to use the word macroscopic “global” conventions we state here are that ~ = 1 and even for “mesoscopic” system sizes, it frequently appears (∆A)2 = hA2i − hAi2 for the variance of an operator A. in many of the discussed papers. Note that macroscopic Furthermore, the components |Ai and |Di of a superposi- can refer to a scaling or to a single number, depending on tion |Ai + |Di are referred to as branches or components. the context and the intention of the reviewed papers. On Sometimes, a pure state as an argument of a function f an abstract level, one might prefer to discuss the prop- is abbreviated f(|ψihψ|) ≡ f(ψ). erties of a state family and the behavior as a function of the system size. Given a specific situation like real data from an experiment, one is probably more interested in 1. Spin ensemble extracting the bare numbers than the hypothetical scal- ing. When we consider many microscopic constituents each As stated before, we are interested in macroscopic carrying an identical, discrete and finite degree of free- quantumness, which is more than a quantum state of a dom we speak of a spin ensemble. Whether this degree of macroscopic system. Determining the precise meaning of freedom is a physical spin or a pseudo-spin is of little rel- evance here. Alternatively, some authors call this system atomic ensemble. The constituents are called particles in the following. If not stated otherwise, each particle is 2 There are even papers on the term macroscopic in the founda- considered to be a two-level system (spin 1/2 particles); tions of quantum mechanics (Jaeger, 2014). hence the Hilbert space is H = C2⊗N . In this case, one 6 might call the system a qubit ensemble. The system size conveniently parametrized by the Bloch angles |φi = is the number of constituents, N. cos ϑ/2 |0i + eiϕ sin ϑ/2 |1i. An important class of operators are local operators A quantum state that is often discussed in the present context is the multipartite Greenberger-Horne-Zeilinger N X (l) (GHZ) state (Greenberger et al., 1989) A = al , (3) l=1 ± 1  ⊗N ⊗N  GHZN = √ |0i ± |1i , (4) (l) 1⊗l−1 1⊗N−l 2 where al ≡ ⊗ al ⊗ is a single-particle op- erator al acting the lth spin. For convenience and with- which is considered to be macroscopically quantum by (l) (l) many physicists. This state is a limiting case of the “gen- out loss of generality, we set kal k = 1 and Tral = 0. For two-level systems, al can be uniquely decomposed eralized GHZ state” into Pauli operators σ , σ and σ . Local operators x y z | i ∝ | i⊗N | i⊗N (5) are sometimes called linear operators. Nonlinear oper- Φ 0 +  , ators describe interactions between the particles and are where |i = cos  |0i + sin  |1i. typically written as polynomials of local operators. As We are also interested in the symmetric superposition an extension of local operators, we also consider quasi- states local operators. These observables are sums of opera- X ⊗N−k ⊗k (l) | i ∝ | i ⊗ | i (6) tors where each addend al has a nontrivial support on N, k π 0 1 , group l consisting of O(1) particles. Here, we limit the π:perm (l) al to act on nonoverlapping groups only and again we where π are particle permutations. These quantum states (l) set kal k = 1. In this way, one can see quasi-local op- are often called Dicke states (Dicke, 1954). For k = 0,N, erators as local operators of quasi-particles (“molecules the states are product states, while all other Dicke states composed of atoms”) each living in a high-dimensional are genuinely multipartite entangled. An important in- space. stance is |N, 1i, which was called W state in Sec. I.A (Dür Local operators for which all single-particle terms are et al., 2000). It is typically cited as a counterexample to (l) (l) identical (i.e., al ≡ a ) are called collective operators a macroscopic quantum state, despite its widespread en- or extensive observables. For spin-1/2 particles, every col- tanglement. P (l) Another important state class are spin-squeezed states lective operator is hence a weighted sum of Sx = l σx , P (l) P (l) as defined by Kitagawa and Ueda (1993). There are S = σy and S = σz , which fulfill the SU(2) y l z l various ways to generate spin-squeezing. A well-known commutation relation [Sx,Sy] = 2iSz (and permuta- tions). The unusual factor 2 comes from the choice method is the one-axis twisting, 2 ⊗N of normalization. The ladder operators are defined as −iνSx −iµSz 1 |Sµi = e e |+i , (7) S± = 2 (Sx ± iSy). An important class of operations are local operations where the rotation generated by Sx is just to align the and classical communication (LOCC). This implies ac- squeezed axis to z (thus, ν = ν(N, µ)). The optimal cess to single particles, arbitrary operations on them and squeezing is achieved for µ = 241/6(N/2)−2/3 (Kita- processing possible measurement results for future oper- gawa and Ueda, 1993). A common way to characterize ations. squeezed states without reference to its generation is the Note that here we are not much concerned of whether squeezing parameter the particles are distinguishable (by an additional degree N(∆S )2 of freedom such as position) or whether they are in a 2 ≡ n1 (8) ξ 2 2 , bosonic mode and hence symmetrized. The difference is hSn2 i + hSn3 i that we only deal with collective operators in the latter which is strictly smaller one, , for squeezed states case. ξ < 1 (Sørensen et al., 2001). Here, are three or- There are several important state families for spin (n1, n2, n3) thogonal orientations of collective spin operators. This 1/2 particles. The computational basis is denoted by means that the states has to exhibit a large polarization {|0i, |1i}, that is, the eigenbasis of σ . We denote |±i as z in the n − n plane and simultaneously a small (i.e., the eigenstates of σ . 2 3 x squeezed) variance in . Among the symmetric states, the spin-coherent states n1 |φi⊗N are typically seen as the pure states with the “most classical” properties3. The single-qubit state is ∆S/kSk ≤ O(N −1/2) for all collective operators S. In addi- tion, the Heisenberg uncertainty relation ∆Sx∆Sy ≥ |hSzi| is tight in case |φi⊗N is polarized somewhere in the x − z or in the 3 Spin-coherent states have a vanishing relative uncertainty y − z plane. 7

The last class of spin states introduced here are the positions of coherent states (SCS).5 A typical instance cluster states (Briegel and Raussendorf, 2001). Clus- is ter states are special instances of graph states (Hein et al., 2006), which are constructed by applying con- |SCSi ∝ |αi + |−αi , (10) (i,j) (i) (i) (j) trolled phase gates UC = |0ih0| + |1ih1| σz on a set of spin pairs (i, j). If the UC are applied to near- whenever α is large. The overlap between the two coher- est neighbors in a certain geometry, the graph state is ent states, hα| − α i = exp(−2|α|2), vanishes for α  1. ⊗N called cluster state |Cli = UC |+i ; for example, a one- The equivalent state in photon number space is |0i+|Ni, dimensional arrangement gives with N  1. However, since |Ni is considered to be highly nonclassical, |SCSi and the superposition of Fock N−1 Y i,i states have a different characteristics. U ( +1) (9) C = UC . Another important state class are squeezed states. i=1 The most important instance of squeezing is the one of squeezed coherent states

2. Photonic systems (1) |ζ, αi = Sζ |αi , (11) We consider well-defined temporal, spatial, frequency and polarization modes. For every mode, we define the (1) 1 †2 where the squeezing operator Sζ = exp(− 2 (ζa − usual quadrature operators X and P with the canoni- ∗ 2 ζ a√)) reduces the variance of the quadrature cal commutation relation [X,P ] = i and the decompo- i arg(ζ) † −i arg(ζ) −2|ζ| 1/ 2(e a + e √a) by a factor e and sition into dimensionless√ creation and √ oper- increases the variance of i/ 2(ei arg(ζ)a† − e−i arg(ζ)a) by † † ators X = 1/ 2(a + a) and P = i/ 2(a − a). The e2|ζ|. notion of a linear operator refers to operators that are All these examples have extensions to multi-mode sys- † linear in a and a . The system size is the mean photon tems. For SCS, this includes states like |α, βi+|−α, −βi. † number ha ai. A famous two-mode superposition of Fock states is the Sometimes we discuss multi-mode scenarios, which NOON state |N, 0i+|0,Ni. Finally, two-mode squeezing have some parallels with spin ensembles in case of many of vacuum with the operator modes (except that there are infinitely many levels per mode). The equivalence of a local operator is a sum of (2) S = exp(−ζa†b† + ζ∗ab) (12) single-mode operators linear in X(l) and P (l). The sys- ζ tem size is the total mean photon number. plays an important role in spontaneous parametric down- Several quantum states are repeatedly discussed in the conversion. remainder of the paper. A useful countable basis in a single mode consists of Fock states (or photon number states) a†a |ni = n |ni for integer numbers including the vacuum |0i. 3. Massive systems The pure state that behaves “most classically”4 is the coherent state |αi defined via a |αi = α |αi for all α ∈ C. Massive systems are clearly of high importance for the It can equally be seen as a displacement of the vacuum topic of macroscopic quantumness. Mathematically, they |αi = D(α) |0i ≡ exp(αa† − α∗a) |0i. The mean number are similarly treated as a photonic mode, except that X of photons is |α|2 with a spread of |α| in the photon and P now become position and momentum. In particu- number spectrum. The variance of the quadratures is lar for a massive particle in a harmonic trap, the creation independent of α. and annihilation operators play the same role as for pho- Among other representations, the Wigner function is tons. In this case, state classes like coherent states, SCS a well-established way of representing quantum states in and squeezed states are also important here. Unlike pho- phase space. The Wigner function is a quasi-probability tons, however, the role of parameters, in particular, the distribution, whose marginals gives the statistics of the mass m, plays a crucial role in massive systems and is quadrature operators (Scully and Zubairy, 1997; Wigner, typically considered to be the system size. In addition, 1932). also in the case of “free” particles, the role of the distance Well-studied states in the present context are super- in a spatial superposition is not obvious.

4 Regarding classicality, a similar comment as footnote 3 applies, 5 In the literature, SCS often stands for Schrödinger-cat state, a where collective operators are replaced by quadrature operators. frequent name for state Eq. (10). 8

4. Superconducting systems macroscopic quantum states using different setups. We summarize and conclude in Sec. VI. There are many different degrees of freedom that po- tentially show macroscopic quantum behavior. Notably, superconducting circuits such as superconducting quan- II. MEASURES FOR MACROSCOPIC SUPERPOSITIONS tum interference devices (SQUIDs) play an important AND QUANTUM STATES role in the present context. Often, one works with col- After motivating and specifying macroscopic quantum- lective degrees of freedom such as the total flux of the ness in Sec. I.A, we now have a closer look at theoreti- system, Φ. Mathematically, this is equivalent to a mas- cal proposals that aim at formalizing intuitive ideas in sive particle moving in one-dimensional potential. The a mathematical framework. The terminology and nota- system size in this case might be defined as the total tion used in this section is defined in Secs. I.C and I.D. number of electrons that are involved in the experiment. We start with a detailed summary of several measures in Sec. II.A focusing on the motivation, the mathemat- ical formulation and some basic properties. Inside this E. Structure of the review and reading guide subsection, we stay close to the original language and notation of the reviewed papers. In addition every en- The purpose of the review is two-fold. On the one try is opened with an encapsulated text, where we try to hand, we provide an introduction to different aspects convey the idea behind each measure as simply as possi- of macroscopic quantum states for interested non-expert ble. This is meant to help readers novice to the topic, but readers. On the other hand, we also give a comprehensive comes at the price of giving our own interpretation which overview and detailed discussion of various approaches to might not exactly match the authors’ original motivation. quantum macroscopicity. This involves detailed discus- A reader familiar with the literature might skip this part. sions and comparison between approaches, some of which In Sec. II.B, we apply the measures to several examples: might only be relevant to experts in the field as techni- standard situations discussed in many papers as well as calities, subtilities and particular aspects are discussed. specialized examples to see similarities and differences In order to make the article accessible for both, experts between the proposals. The discussion is continued in and non-experts, we provide here a readers guide. Secs. II.C–II.H, in which we elaborate on several details Section I provides the basis of this review: the pre- and comparisons (see table I for an overview). We sum- cise scope and motivation of the review, the most im- marize in Sec. II.I. This section is meant to complement portant terminology, the mathematical notation and the and extend previous contributions6. discussed physical systems. In Sec. II we give an overview of different measures, and provide detailed discussions on their relations and differences. Some of these discussions A. Summary of measures are rather technical and might be hard to understand for a non-expert reader. Such a reader has to keep this Although the goal of this section is to give a summary in mind, and do not get stack or intimidated with the of the relevant literature as neutral and objective as pos- technicalities. To facilitate this, we provide –for each of sible, we would like to draw the reader’s attention to the measures reviewed in section II– a boxed text that some basic observations. While it is true that different conveys the general idea behind each measure in non- authors partially use different concepts to define macro- technical terms. We suggest to only read the encapsu- scopic quantumness, it is obvious that a common goal lated text first, and continue with sections II.C and II.D. is to distinguish “interesting” from “uninteresting” states At that point, the reader is now equipped with the neces- and to find an ordering between states. However, since sary background information to proceed reading the rest all pure quantum states are connected via unitary oper- of the article following his or her interest. ations, one has to find a way to break this unitary equiv- Instructive examples and an in depth discussion of alence. This is similar to entanglement theory, where the measures are provided in Secs. II.B–II.I. Section III the partition of a large Hilbert space into subspaces is contains a detailed discussion on limitations to prepare, the structure necessary to define separable states (i.e., maintain and measure macroscopic quantum states, and “uninteresting” states; Horodecki et al. (2009)). For the might be of particular relevance for readers interested present topic, this prestructuring is far less obvious. We in fundamental questions regarding macroscopic quan- invite the reader to observe which mechanisms for break- tum states. In turn, Sec. IV deals with potentials of ing the unitary equivalence has been chosen in the follow- macroscopic quantum states and contains a brief discus- ing summary (and refer to table I and Secs. II.C and II.D sion of possible applications in the context of probing the limits of quantum theory, in quantum metrology and in quantum computation. Section V provides an overview 6 For example, Farrow and Vedral (2015); Fröwis and Dür (2012b); of implementations and previous attempts to generate Fröwis et al. (2015); Jeong et al. (2015). 9

(2007)

(2008)

(2015)

(2014c) (2017b)

et al. (2017)

et al. (2016)

(2002)

et al.

et al. et al.

et al.

et al.

et al.

Leggett (1980, 2002) Dür Shimizu and Miyadera (2002) Björk and Mana (2004) Cavalcanti and Reid (2006, 2008) Korsbakken Marquardt Lee and Jeong (2011) Park Fröwis and Dür (2012b) Nimmrichter and Hornberger (2013) Sekatski Sekatski Laghaout Yadin and Vedral (2015) Kwon Preferred observable x o x x o o x x x x Partition in subsystems x x x x x x x x Basic approach Preferred representation x x Preferred decoherence o x x Macroscopic coherence x x x x o o x x x o o x x Effective particle number x x x x o x Ease to distinguish x o x x x Motivation/goal Relative improvement x o Collapse models/ fragility o x o o x o x Nonclassicality x x x x x |Ai + |Di x x x x x State structure General state x x x x x x x x x x x Spin ensemble x x x x x x x x x x x x x Photons x x o o x x x o x o Physical system Mass o x SQUID x x x x x ∗ Mathematical symbol for measure D, Λ p, q M SCδ D¯ I IS Neff µ Size MIC N , D N Mσ

Table I Properties of measures as discussed in Secs. II.C–II.E. Each column corresponds to a measure that is reviewed in Sec. II.A in the order of appearance. Four different categories are compared: the basic approach of the measure to break the unitary equivalence (Sec. II.C.1); the motivation or goal of the measure (Sec. II.C.2); the state structure for which the measure is formalized (Sec. II.D); and the physical system to which the measure can be applied (Sec. II.E). The symbol “x” means general applicability; the symbol “o” stands for partial applicability (e.g., “macroscopic coherence” is not the primary motivation but a consequence; or some measures are applied to photonic states only for specific instances). The last line summarizes the mathematical symbols used by the authors for their measures (if any). As some papers leave room for slightly different view points, we understand this table not as an ultimate judgment but as a starting point for further discussions. for further discussion). Some authors partition the sys- When discussing the measures in details, it is worth keep- tem into subsystems like in entanglement theory. Others ing the implications of each step (intuition, basic frame- choose an observable to specify a basis and a spectrum work and fine-tuning) in mind. (i.e., a set of eigenvalues), which is close to attempts in coherence theory. Alternatively, one can focus on a specific state representation like the Wigner function in 1. Leggett (1980, 2002) phase space. To give another example for a difference in the struc- For a superposition state |Ai+|Di of a system com- ture of the proposals, the measures vary in their range posed of a large number of particles, the extensive of applicability. Some discuss specific examples, others difference Λ is the difference between the expec- work with pure states decomposed into a structure of tation values of some extensive observable of the “dead and alive”, |Ai + |Di, while some proposals are states |Ai and |Di. The disconnectivity D mea- defined for general mixed states. sures the quantumness (nonseparability and purity) of the total superposition state. A macroscopic Let us finally summarize the typical way a measure quantum state is required to have both D and Λ is constructed. It starts with an intuition or examples, large. which is then formalized. This normally consists of the basic framework and a sort of fine-tuning, for example, Motivated by the question “What experimental evi- by choosing the right observable or fixing a parameter. dence do we have that quantum mechanics is valid at 10 the macroscopic level?”, Leggett (1980) was the first to 2. Dür et al. (2002) point out a qualitative difference between quantum ef- fects on microscopic scales amplified to large scales and Two ways to quantify the macroscopic quantumness genuine large-scale quantum signatures. As Leggett ar- of a many-body state |Φi of type Eq. (5) are in- gues, the common feature of macroscopic quantum states troduced, both via a comparison with GHZ states. is their long-range coherence. This cannot be revealed First, the effective size of |Φi is identified with the with single local measurements. Only simultaneous mea- size of GHZ that has the same decoherence rate surements of a large number of particles allow to distin- under local noise. Second, it is identified with the guish this from an incoherent mixture. largest GHZ state, that can be obtained from |Ψi with local operations and classical communication. In order to quantify this insight, Leggett (1980, 2002) Both lead to the same result. introduces two concepts for a large system consisting of N of particles. First, consider a superposition of |Ai+|Di Motivated by several proposals and experiments, Dür and calculate the difference in the expectation value for et al. (2002) study the generalized GHZ state |Φi, a particular extensive variable (e.g., total charge, total Eq. (5), with   1, that is, | h0|  i |2 ≈ 1−2. While the , or total momentum). This number di- two branches are orthogonal even for  close to zero (as vided by a characteristic microscopic unit (e.g., the Bohr long as N2 = O(1)), Dür et al. argue that Eq. (5) ef- magneton for the magnetic moment) is called extensive fectively corresponds to the ideal state |GHZni, Eq. (4), difference Λ. Second, to characterize the quantumness of with n  N. To show this, one generally chooses a the state (here, the entanglement between the particles), key property (or a set of key properties) for which |Φi Leggett defines the disconnectivity D as follows. For sim- and |GHZni coincide. As argued by the authors, prop- plicity, we consider symmetric states ρ of spin ensembles. erties connected to the nonclassicality of the states are clearly of importance. They choose two methods: (i) the The von Neumann entropy SM = −TrρM log ρM of the rate with which the coherence decays is compared; (ii) reduced density matrix ρM = TrM+1,...,N ρ is a measure of the correlation between two parts with M and N − M the average size n one can achieve by going from |Φi to particles, respectively. The quantity |GHZni with LOCC. For both methods, they find that n ≈ N2 in the limit N  1 and   1. This example, which was not directly expanded to a general measure, was later used as a test bed for proposals to measure SM δM = (13) arbitrary macroscopic quantum states. minn(Sn + SM−n)

3. Shimizu and Miyadera (2002) and followups compares SM with the minimal correlations from smaller partitions to the rest of the system. If numerator and de- The macroscopic quantumness of pure states – the nominator are zero, δ is set to one and δ := 0. Leggett M 1 index p is identified via the maximal variance of now defines to be the largest integer for which D M δM the state with respect to all extensive observables is smaller than some predetermined small fraction. In A. The extension to mixed state is formally intro- words, a large disconnectivity is found for states that are duced via an additional maximization to uncover rather pure when considering a large part of the system widespread coherence. (i.e., SM  1 for M = O(N)) but all smaller parts of size n still exhibit large entropy indicating entanglement Shimizu and Miyadera (2002) study the stability of fi- to the rest of the system.7 nite macroscopic quantum states under weak noise and local measurements. The systems they consider can be How to combine the two measures D, Λ is ultimately decomposed into subsystems like the spin ensembles in- an open question. In the examples discussed by Leggett, troduced in Sec. I.D.1. They find a difference for the both should be somehow large in order to call a state decay rate Γ of the purity depending on whether or not a macroscopically quantum. quantum state has the so-called cluster property, whose absence implies long-range correlations in a many-body system and wide-spread entanglement for pure states. The noise and decoherence is supposed to be generated by some operator A. From a physical viewpoint, it is clear that A has to be a local operator. While the detailed results regarding fragility are discussed in Sec. III.A.3.a, 7 Note, however, that Leggett himself does not attach much im- portance to the precise mathematical formulation of this idea, we emphasize here that the authors are able to connect which “could almost certainly be substantially improved” Leggett Γ with the variance (∆A)2. (2002). A series of papers was devoted to formalize this ba- 11 sic insight (Morimae et al., 2005; Shimizu and Morimae, which reflects the initial motivation as discussed in the 2005; Ukena and Shimizu, 2004). Shimizu and Morimae beginning of this subsection. For pure states, it is shown (2005) remark that two pure states of N subsystems can that q = 1 ⇒ p = 1, p = 1 ⇒ q ≤ 1.5 and q = 2 ⇔ p = 2 be reasonably called macroscopically distinct if there ex- (which are the corrected statements from Shimizu and ist some local operator A such that the difference of its Morimae (2016); Tatsuta and Shimizu (2017)). expectation value between the two states is O(N) (with the normalization of Sec. I.D.1). A general state ρ is more macroscopically quantum the more coherence between 4. Björk and Mana (2004) macroscopically distinct states it contains. Defining the eigenstates A |ak, νi = ak |ak, νi (ν accounts for possible The macroscopic quantumness M of a superposi- degeneracy) of a given local observable A, the amount of tion state is identified with the advantage it offers macroscopic coherence is quantified by the total weight for interferometry as compared to the individual 0 of all terms | hak, ν| ρ |al, ν i | with |ak − al| = O(N). The components |Ai or |Di. The interforometric use- authors propose to measure this coherence in the follow- fulness is defined by how fast the state becomes ing way. orthogonal to itself when subject to a unitary evo- iθA For a pure state ρ = |ψihψ|, the spread of coherence is lution e for some fixed observable A. quantified by the index p (with 1 ≤ p ≤ 2) of the state Björk and Mana (2004) motivate their contribution by (Ukena and Shimizu, 2004), defined as the best scaling the need for an operational meaning of a measure rather of the variance of the state with respect to all normalized than focusing on particle and mode numbers. In their local observables opinion, a genuinely macroscopic quantum state should 2 p give some advantage over states that lack this feature in max (∆A)ψ = O(N ). (14) A:local some practical application. Björk and Mana consider the interferometry as such a task. This general idea is for- Expressing A that maximizes Eq. (14) with single-spin (l) malized for a state of the form |Ai + |Di in the following operators al , one finds that the variance reads way. (i) The starting point is to identify a preferred ob-

N N servable A which results from the experimental context 0 0 2 X (l) 2 X (l) (l ) (l) (l ) as the most useful one for a particular interferometric (∆A) = (∆al ) + hal al0 i − hal ihal0 i. l=1 l=6 l0=1 application. (ii) One identifies some semi-classical states (15) |cai, that are pure and have a smooth but rather narrow The scaling p = 1 hence implies that the quantum corre- distribution c(A − a) in the eigenvalues of A centered at lations expressed in the second term in Eq. (15) do not a. (iii) One imagines the superposition state to undergo i θA play a significant role. For every spin l, there exist only an evolution e generated by the observable A. In such O(1) “neighbors” with which the spin shares correlations. an interferometric scenario, large oscillation frequencies Shimizu and Miyadera (2002) call this the cluster prop- between the initial state and the finial state indicate a erty. In contrast, p > 1 and in particular p = 2 means large separation between coherent components of a su- that as growing number of pairs share nonvanishing quan- perposition state, which do not occur for mixtures. This tum correlations.8 last observation is formalized as follows. Consider a su- However, for mixed states this is not sufficient since perposition state the variance of a state with respect to A only depends on 0 |ψi ∝ |ca i + |ca i , (17) the diagonal terms | hak, ν| ρ |ak, ν i |. To overcome this 1 2 problem, Shimizu and Morimae (2005) introduce the so- where the two components have a negligible overlap. The called index defined as (in the formulation of Morimae q overlap of the evolved state with the original one is given (2010)) by   q Z max N, max k[A, [A, ρ]]k1 = O(N ). (16) iθA θ(a1 − a2) iθA A:local | hψ| e |ψi | = 2 cos e c(A)dA , (18) 2 | {z } While the “outer” maximization is just to guarantee c˜(θ) 2 1 ≤ q ≤ 2, a large trace norm k[A, [A, ρ]]k1 = O(N ) for local operators A is only possible with significant contri- with two contributions to the oscillation of this overlap. 2 0 2 butions from elements (ak−al) hak, ν| ρ |al, ν i = O(N ), The first one comes from the shape of underlying classi- cal states c˜(θ), while the second and more rapid effect is  θ(a1−a2)  due to the superposition cos 2 . Hence, one com- 8 Even though the name “index p” was first introduced by Ukena pares the minimal θ for which the superposition evolves and Shimizu (2004), we refer to Shimizu and Miyadera (2002) to an orthogonal state, θsup ≈ π/(a1 − a2), to the cor- for measure (14). responding θ for a single classical state, θc ≈ π/(∆A)c 12

(with (∆A)c the width of the distribution c(A)). Hence the dimensionless ratio between the two “orthogonaliza- tion times” P(x) −1 0 +1

θc P (x) P (x) M = (19) − + θsup quantifies the “interferometric macroscopicity” of the su- perposition state.

5. Cavalcanti and Reid (2006, 2008)

For a given observable A, the macroscopic quan- −S/2 S/2 x tumness of a state is identified with the maximum spectral range S on which the state exibits quan- Figure 1 Division of the spectrum of the observable X into tum coherence. That is, the density matrix con- three intervals. The curve is a schematic example of the prob- tains some coherence terms between eigenstates of ability distribution of the considered state ρ. Cavalcanti and A with eigenvalues different by at least S. Reid (2006) aim to verify coherence between the left (-1) and the right interval (+1). From Cavalcanti and Reid (2006). Cavalcanti and Reid (2006) are interested in witness- ing coherent superpositions on the macroscopic scale. Following their argument, macroscopic coherence contra- where ρL completely lies in the intervals I− and I0 and dicts a general notion of macroscopic reality, in which an ρR in the intervals I0 and I+. Let us define the nor- object is necessarily in one out of several states. To be malized distributions p±(x) = p(x|x ∈ I±) and the R weights π± = p±(x)dx and π = 1 − π − π−. more specific, let us consider a preselected observable X I± 0 + on the real line which is divided into three intervals called From this, we define the mean µ± and the variance 2 2 2 2 left I−, central I0 and right I+ (see Fig. 1). The interval (∆X)± of p±(x), (∆X)ave = π+(∆X)+ + π−(∆X)− and 2 2 2 2 I0 has length S. A general pure state |ψi can be writ- δ = (µ+ + S/2) + (∆X)+ + (µ− − S/2) + (∆X)− + S/2. ten as a superposition of states located in the respective Then, Cavalcanti and Reid (2006) show that for all states intervals (21) it holds that 2 2 ≥ (22) |ψi = c− |ψ−i + c0 |ψ0i + c+ |ψ+i . (20) [(∆X)ave + π0δ](∆P ) 1.

The goal is now to guarantee that a state prepared in an Violating Eq. (22) proves coherence in X between I+ and experiment exhibits coherence between the left and the I−. right interval. Working with general mixed states ρ, this This inequality can be generalized or modified in sev- means that states of the form Eq. (20) with contributions eral ways (Cavalcanti and Reid, 2008). First, one can from the left and the right interval are present in every use other uncertainty relations (e.g., involving sums of pure state decomposition of ρ. variances rather than products); one can derive them To derive a witness, Cavalcanti and Reid point out for bipartite scenarios; and one can drop the predeter- mined binning into three intervals and only verify coher- that a state with nonvanishing c− and c+ must have a minimal spread in the spectrum of X. From the Heisen- ence with a minimal distance S. The latter is gives a berg uncertainty relation it follows that states with lower simpler expression: The violation of the bound variance in X necessarily have a larger variance in a con- 2 (∆P )2 ≥ (23) jugate observable P . This argument can be extended to S mixed states. The authors derive such general bounds for the canonical commutation relation9 [X,P ] = 2i. If implies that the corresponding states exhibits coherence ρ does not contain a state like Eq. (20), it can only be with a spread of at least S. written in the form 6. Korsbakken et al. (2007) ρ = PLρL + PRρR, (21)

The macroscopic quantumness Cδ of a many-body superposition |Ai + |Di is related to the minimum 9 In this subsection, we adapt to the convention of Cavalcanti and number of subsytems that have to be measured in Reid. X and P , as introduced in Sec. I.D.2, would give [X,P ] = order to learn the “which-path” information (i.e., i. This only changes some constants in Eqs. (22) and (23) and later in example 7. |Ai or |Di). 13

Let us consider a system that can be divided into N the other. More precisely, they introduce a hierarchy of subsystems, called particles. The quantum state is sup- Hilbert spaces Hd for d ≥ 0, where H0 = span{|Ai} and posed to be of the form |Ai + |Di. The basic intuition of each new subspace Hd+1 is constructed from the pre- Korsbakken et al. (2010) is that the (macroscopic) dis- vious spaces in the following way. One takes the span tinctness of the components |Ai and |Di can be quanti- H˜d+1 of all states that can be obtained by applying a fied by asking how many particles have to be measured single elementary operation to all states in Hd. Then, in order to distinguish the two states, or equivalently to H˜d+1 is made orthogonal to all subspaces H0,..., Hd by collapse the superposition |Ai + |Di to just one branch. subtracting them, which is called Hd+1. In this way all This is formalized in the following way. Given the de- subspaces of the hierarchy are orthogonal (hence the di- sired guessing probability Pg = 1 − δ, Korsbakken et al. rect sum notation is justified) and, for a properly chosen L∞ consider the minimal number nmin of particles one has to set of elementary operations, one finds H = d=0 Hd. measure on average in order to distinguish |Ai from |Di Consequently, the state |Di admits a unique decomposi- with probability Pg (subsystems are drawn at random). tion Then, the size of the superposition |Ψi ∝ |Ai + |Di is ∞ defined by X |Di = λd |νdi , (26) N d=0 Cδ(Ψ) ≡ . (24) nmin where each |vdi ∈ Hd is the projection of the state on The optimal guessing probability between two states ρ the corresponding subspace. The distinctness between and σ is given by the states |Ai and |Di is then defined as 1 1 P [ρ, σ] = + tr|ρ − σ|. (25) ∞ 2 4 X 2 D¯ = |λd| d (27) Hence, one can compute the average guessing probabil- d=0 N ity with n subsystems given all the n reduced density (n) (n) quantifying the average number of elementary operations matrices ρA = TrN−n |AihA| and ρD = TrN−n |DihD|. The complexity of the minimization is significantly re- that are needed to go from |Ai to |Di. The authors note duced if the initial state is permutationally invariant. that the measure is in general not symmetric under ex- The authors discuss several aspects of their measures, change of |Ai and |Di. Volkoff and Whaley (2014b) pro- including the dependency on the choice of splitting into pose an extension of this measure to multi-mode photonic |Ai and |Di , cases for which the two branches do not systems by discussing a specific example. have equal amplitudes and slight variations of the defini- tion of Cδ(Ψ). In particular, the author emphasize that P should be chosen to be close to one for a reasonable g 8. Lee and Jeong (2011) and Park et al. (2016) notion of macroscopic distinctness. Volkoff and Whaley (2014b) discuss a potential formulation for photonic sys- tems (see also Sec. II.E.2). The macroscopic quantumness I of the state is definied by quantifying the non-classical features of its phase-space representation (i.e., the frequency 7. Marquardt et al. (2008) and amplitude of Wigner function oscillations).

Several measures discussed so far have been defined The macroscopic quantumness D¯ of a superposition with respect to an observable A, which gives rise to a |Ai + |Di is defined by the number of elementary preferred basis. Alternatively, one can also consider rep- operations that have to be applied in order to map resentations of quantum states that are not just expan- |Ai to |Di. sions in a given basis. In particular, the Wigner function If the system admits a partition into N subsystems or representation is commonly used in phase space to visu- consists of N particles, one can also define a set of all ele- alize quantum states. Lee and Jeong (2011) observe that mentary operations. For example, such a set can include the Wigner function of states that are intuitively consid- all operations that only affect one subsystem, or only ered to be macroscopically quantum (e.g., |αi + |−αi) modify the state of one particle. Marquardt et al. (2008) exhibit two or more distinct peaks with some oscillating focus on the example of N fermions, where an elemen- pattern between the peaks. In contrast, classical states tary operation is naturally given by the exchange of one are known to have a positive and smooth Wigner func- † fermion cjck. Their idea is to quantify the distinctness tion. Following this intuition Lee and Jeong propose to of the states |Ai and |Di by counting how many elemen- quantify the size of a state by the “frequency” of oscil- tary operations one has to apply to go from one state to lations of its Wigner function. Formally, the size of a 14 quantum state ρ for M bosonic modes is defined as 9. Fröwis and Dür (2012b) and followups

The macroscopic quantumness Neff of a many-body M Z M  2  π 2 X ∂ state is related to its maximal quantum Fisher in- I(ρ) = dα W (α) − ∗ − 1 W (α), 2 ∂αm∂α formation with respect to all extensive observables m=1 m (28) A. The QFI can be read as a signature of inter- which takes a simpler form when expressed in terms of ferometric improvement offered by the state over P product states, or as an extension of the variance the characteristic function χ(ξ) = tr ρ exp( m ξmam + ∗ † to mixed states. ξmam). As the authors show, the quantity I(ρ) can equally be expressed as the susceptibility of the state We consider a spin ensemble with N particles. Fröwis to lose purity when all the modes are subject to photon and Dür (2012b) argue that a genuinely macroscopic loss, that is, quantum state of the joint system should exhibit some quantum effect that does not reduce to an accumulation of microscopic quantum effects displayed by its individual 1 d P(ρt) I(ρ) = − , (29) constituents. In this sense the effective size of a state can 2 dt t=0 be intuitively thought of as the minimal irreducible num- ber of constituents. To formalize this guideline, Fröwis with the purity of a state P(ρ) ≡ trρ2. The decoherence and Dür use the quantum Fisher information (Braunstein process is specified by and Caves, 1994; Helstrom, 1976; Holevo, 2011)

2 X (πk − πl) 2   F(ρ, A) = 2 |hψk| A |ψli| , (32) X † 1 † πk + πl ρ˙t = amρtam − {ρt, amam} . (30) k,l m 2 with the spectral decomposition of the state, ρ = P k πk |ψkihψk|. The quantum Fisher information is a Lee and Jeong emphasize that I(ρ) simultaneously measure of the susceptibility of ρ to small influences gen- quantifies the quantumness and the macroscopicity of a erated by A (see Sec. II.H.2). state. In particular there is no need to assume that the The effective size of the state is defined as state is pure. One simply expects that the oscillations of 1 Neff(ρ) = max F(ρ, A). (33) the Wigner function are smoothed with noise, as it hap- 4N A:local pens for the examples discussed in the paper. For later 1 2 2 Note that the quantum Fisher information reduces to comparison, we note that I(Ψ) = 2 [(∆X) + (∆P ) − 1] for pure states. Gong (2011) remarks that I(ρ) can be- four times the variance for pure states. Furthermore, it come negative, which can be simply fixed by adding 1/2 is the convex roof of the variance (see Yu (2013) and to the definition of I(ρ) (Jeong et al., 2011). Sec. II.D.2). The normalization factor 1/(4N) is chosen such that all pure separable states have size 1, while the This measure for the Wigner function was later gen- maximal possible size is N (attained by the GHZ state). eralized to spin ensembles with N particles with each This measure has an operational aspect as, on the one spin S (Park et al., 2016). A similar reasoning as be- hand, a large quantum Fisher information can be wit- fore leads to measure the frequency components of the nessed via fast unitary time evolution generated by A so-called Stratonovich-Weyl distribution, which is the (Fröwis, 2012). On the other hand, it has an applied spin-equivalent of the Wigner function (Agarwal, 1998; aspect as Neff (ρ) tells us how much better ρ can be in Klimov and Chumakov, 2009). In contrast to the phase a potential parameter estimation scenario compared to space treatment, Park et al. choose to maximize the fre- the best separable states (see Sec. IV.B). The definition quency measure over all quantization axes (using collec- Eq. (33) might be extended to quasi-local observables A tive spin operators A). In addition, they add the purity (i.e., sums of few-particle operators; see also example 4 in the denominator and define in Sec. II.B.1). In order to explicitly deal with a “dead and alive” struc-

1  2 2  ture, the authors define the so-called relative Fisher in- IS(ρ) = max Tr A ρ − ρAρA . (31) NSTrρ2 A:collective formation for |Ψi ∝ |Ai + |Di, N (Ψ) rel eff (34) The measure is particularly compared to the quan- Neff (Ψ) = 1 1 . Neff (A) + Neff (D) tum Fisher information (see Sec. II.A.9). In addition, 2 2 its role in quantum phase transitions is discussed (see Later, the proposal Eq. (33) was extended to photonic Sec. IV.C.3). systems. As argued by Fröwis et al. (2015); Oudot et al. 15

(2015), the equivalent of local operators in spin ensem- by several physical requirements imposed on the model, bles are quadrature operators in phase space. The fac- like the invariance under Galilean transformation, sym- tor 1/(4N) is generally replaced by the quantum Fisher metry under particle exchange and others (Nimmrichter information of the “most classical” state Neff (Ψclassical) and Hornberger, 2013). At this stage, it is fully specified (see also Fröwis (2017)). For spin ensemble, this is by the coherence time parameter τe and the distribution chosen to be product states. In phase space, coher- ge(s, q). (The reference mass parameter me can be ab- ent states are selected as the most-classical pure states. sorbed in the previous two.) The authors further assume To handle multi-mode situations, one considers sums of the distribution to be a product of Gaussians for s and M (m) (m) quadrature operators P with q, with the corresponding width σs and σq. Hence, the Xθ = m=1 Xθm Xθ = cos(θ)X(m) + sin(θ)P (m) and maximizes over the angles dynamical modification of quantum mechanics is just de- scribed by three parameters τe, σs and σq. From this, the θ = (θ1, . . . , θM ), that is, macroscopicity measure µ of an experiment is defined as 1 a recalibration Neff (ρ) = max F(ρ, Xθ). (35) 2M θ  τ  µ ≡ log e (38) 10 1s

10. Nimmrichter and Hornberger (2013) of the greatest time parameter τe excluded by the exper- iment. For this, one optimizes over the other two param- The macroscopic quantumness µ of an experimen- eters with σs ≤ 10 pm and ~/σq ≥ 10 fm, which are ar- tal setup is defined via the range of unconven- gued to be the limiting value for which a non-relativistic tional mass-induced decoherence models (as poten- treatment is still valid. Concretely, the modified master tial modifications of standard quantum mechnics) equation Eq. (36) predicts an evolution where branches of some type that are ruled out by the experiment. that correspond to different phase space configurations of massive particles progressively decohere. This effect The proposal presented in this paragraph differs from constrains the results of any measurement that register most other ideas, as Nimmrichter and Hornberger (2013) the interference between different branches (e.g., inter- do not consider the macroscopic quantumness of an iso- ference visibility in a matter-wave interferometer). Re- lated state (including the structure of additional oper- ciprocally any experimental data obtained from such a ators or partitions). They attribute a size to a whole measurement puts a limit on the modification term. The experiment, from the preparation step to the time evo- stringency of this limit is quantified by the measure µ. lution and to the observation of measurement results. This is motivated by their goal to evaluate how well ex- periments exclude slightest variations of standard quan- 11. Sekatski et al. (2014c, 2017b) tum mechanics. In particular, Nimmrichter and Horn- berger focus on dynamical modifications also known as The macroscopic quantumness –Size for |Ai + |Di collapse models for spatial superpositions of massive sys- superpositons and MIC for multicomponent super- tems (Bassi et al., 2013). position states– is identified by how much one can More precisely, the authors consider a model in which learn about the state with a classical detector that the master equation of the system composed of N parti- lacks microscopic resolution. cles ρ˙ = [H, ρ]/(i~) + LN (ρ) is modified by the addition of a dissipative term Sekatski et al. (2014c) define macroscopic distinctness for the two components of a pure state |Ai + |Di by ask- Z 1  † ing how well the two states can be distinguished with a LN (ρ) = d(s, q) WN (s, q)ρWN (s, q) τe coarse-grained measurement of some observable A. Sim-  1 † ilarly as in other proposals, the observable is chosen de- − {W (s, q)WN (s, q), ρ} (36) 2 N pending on the experimental context. This measurement is said to be performed with classical detectors. If the with distributions of the states |Ai and |Di with respect to the eigenbasis of the observable A are given by p0 (a) N   A X mn i me and p0 (a) respectively, the distributions observed with a WN (s, q) = exp ( s · Pn − q · Xn) , D me mn coarse-grained measurement are given by their convolu- n=1 ~ (37) tion where Xn, Pn and mn are the position operator, the X pσ (λ) = n (a|λ)p0 (a), (39) momentum operator and the mass of the n-th particle, A σ A 3 3 a respectively, while d(s, q) = ge(s, q)d sd q is a measure with an isotropic phase-space distribution ge(s, q). The with the coarse-graining nσ(a|λ) = nσ(a − λ) typically particular form of the modification term is motivated chosen as a Gaussian distribution of width σ and mean a 16

(and similar for |Di). Hence, the probability to correctly U = exp(−iA⊗p). The disturbance of the post-measured 0 † distinguish the two states with such a measurement is state ρ = TrEUρ ⊗ |ξ∆ihξ∆| U is measured via given by 0 Z C∆(ρ) = S(ρ ) − S(ρ), (44) σ 1 1 σ σ P [ |Ai , |Di] = + |pA(λ) − pD(λ)|dλ. (40) 2 4 where S is the von Neumann entropy. Sekatski et al. A measure of macroscopicity is constructed from this (2017b) show that this is a sensitive measure for macro- probability provided a choice of a target guessing prob- scopic coherence similar to Eq. (42). Like before, inver- ability Pg and a calibration set of states. This set is a sion of Eq. (44) leads to a measure that scales (at most) range of superposition states |ΨN i with a naturally de- with the spectral radius and that is applicable to general fined size N. To do so, one first computes the maximal al- quantum states, but without the need of a convex-roof σ lowed coarse-graining σmax = sup{σ|P [ |Ai , |Di] ≥ Pg} construction. that still allows one to distinguish |Ai and |Di with a probability of at least Pg. Then, one identifies the size of the superposition |Ψi = |Ai+|Di with the smallest state 12. Laghaout et al. (2015) from the calibration set that attains the same guessing probability Pg under the same amount of coarse-raining The measure consists of two parts. The objective σmax part N quantifies the quantum fluctuations of the state in phases space. The subjective part D quan- Size(Ψ) = inf{N|P σmax [|Ψ i] ≥ P }. (41) N g tifies the average distinguishability of each compo-

A natural choice of the calibration states |ΨN i are nent of in the superposition from all the others. the superpositions of two eigenstates of A separated The product N × D is proposed as a measure for σmax macroscopic quantumness. by N. With√ such a choice one has P [|ΨN i] = (1 + Erf(N/(2 2σ ))) and the size can be directly ex- max Laghaout et al. (2015) are interested in characterizing pressed as a function of σ and P . max g systems that are large, quantum and are composed of The idea of using classical detectors with limited reso- macroscopically distinct branches in at least some of its lution to define macroscopic distinctness (Sekatski et al., subsystems. The authors exclusively treat pure states 2014c) was extended to general states without a prede- and divide their attempt into two parts, which they call termined |Ai + |Di structure. Oudot et al. (2015) adapt “objective” and “subjective” macroscopicity. the work of Sekatski et al. (2014c) to argue that two- For the objective macroscopicity, they note that the mode squeezed states can be considered as macroscopi- spread of phase space distribution accounts for the quan- cally quantum (see Sec. II.B, example 7). tum fluctuations (i.e., coherence). They hence define Later, Sekatski et al. (2017b) rephrased the question of how well |Ai and |Di can be distinguished by asking 1 how much information can be learned from a state by N = [(∆X)2 + (∆P )2 − 1] (45) measuring it with a classical device. Similar as before, 2 consider a pure state |Ψi with the probability distribution as the quantum fluctuations that go beyond the spread of 0 σ pΨ(a) and the convolution pΨ(λ). The information the a coherent state (cf. Lee and Jeong (2011) in Sec. II.A.8). classical detector can learn about |Ψi can be quantified The subjective part of the measure builds on previ- by the mutual information ous work on characterizing the macroscopic distinctness Z using classical detectors with limited resolution (in par- 0 σ Iσ = H(pΨ(a)) − pΨ(λ)H(nσ(a|λ)), (42) ticular, Sekatski et al. (2014c)). The attribute “subjec- tive” comes from the choice that has to be done for the P where H(p(x)) = − x p(x) log p(x) is the Shannon en- measurement. This reduces the problem to a distinction tropy. Sekatski et al. (2017b) define the size of |Ψi as the of probability distributions P (λ) with measurement out- largest σ for which Iσ gives at least b bits of information comes λ. Unlike previous works, Laghaout et al. develop a formalism that allows to define macroscopic distinct- { | ≥ } (43) MICb(Ψ) = max σ Iσ b . ness for more than two states and for different weights. To be more precise, Eq. (43) defines a family of measures Suppose that a pure state |ψi is written as a superposi- parametrized by b. A natural way of extending Eq. (43) tion of preselected states |bki with probability amplitudes . Then, one calculates the distance between any | i to mixed states, MIC[ (ρ), is done via the convex roof ck bk b and the mixture of all other {| i} construction (see Sec. II.D.2.b). bl l=6 k This idea can be rephrased by using von Neumann’s P 2 |cl| |blihbl| pointer model in which the system is coupled to an aux- l=6 k (46) ρ˜k = P 2 . iliary system E in state |ξ∆i with initial spread ∆ via l=6 k |cl| 17

The specific formulation depends on the chosen distance 14. Kwon et al. (2017) measure. The authors discuss one based on the Bhat- tacharrya coefficient (Bhattacharyya, 1946) The macroscopic quantumness Mσ of a state is quantified by the sensitivity of the state to dephas- Z BC X 2 p ing noise generated by an observable A. A large D = 1 − |ck| P (λ|bk)P (λ|ρ˜k)dλ (47) sensitivity is argued to reveal the presence of coher- k ence between spectrally distant eigenstates of A in and another one based on the Kolmogorov distance the state, i.e. eigenstates with very different eigen- (Fuchs and Van De Graaf, 1999) values. Z The approach of Kwon et al. (2017) is motivated by KD 1 X 2 D = |ck| |P (λ|bk) − P (λ|ρ˜k)|dλ. (48) finding an operational meaning of quantum macroscop- 2 k icity in terms of coherence (in the sense of Yadin and Ve- dral (2016), see Sec. II.G). After fixing a spectrum and The subjectivity of D becomes clear by noting that, for P a basis by choosing an observable A = k ak |kihk|, the orthogonal |bki and optimizing over all measurements, authors define a dephasing channel with Kraus operators the distinguishability can always be maximal, that is, σ P p σ σ Qx = k qk (x) |kihk|, where qk (x) is a Gaussian func- D = 1. tion with mean value ak and spread σ. The physical real- The combined proposal is a product of both measures ization of this dephasing channel can come from environ- of macroscopicity mentally induced decoherence or from a low-resolution measurement for which σ is the resolved scale. Then, a M = N × D. (49) state of interest ρ is compared with the same state after P σ σ† dephasing, Φσ(ρ) = x QxρQx . The measure is defined as 13. Yadin and Vedral (2015) Mσ(ρ) = D(ρ, Φσ(ρ)), (51) ∗ The macroscopic quantumness N of a many-body where D is a distance-like function. The authors men- state |ψi is defined as the maximal average size of tion the Bures distance and the quantum relative entropy the GHZ state that can be prepared from |ψi with but characterize the general properties of D such that local operation and classical communication. Eq. (51) fulfill the criteria of Yadin and Vedral (2016) The work of Yadin and Vedral (2015) is devoted to view (see Sec. II.G) for all σ > 0. macroscopic quantumness as a statement about long- As noted by Kwon et al., Mσ(ρ) may lead to surprising range quantum correlations. In the doctrine of multipar- results for small σ. For example, the spin-coherent state tite entanglement, LOCC cannot create nonclassicality. gives higher values than the GHZ state contradicting all Using LOCC protocols one can only reveal entanglement other proposed measures and our intuition. To fix this properties that have been present before. From this point issue, the authors argue that σ has be sufficiently large of view, it is natural not only to consider, for example, in order to faithfully represent macroscopic distinctness. the GHZ state as macroscopically quantum but also all In particular with the interpretation as the scale of mea- other quantum states that can be brought to a GHZ state surement resolution, σ should be in the “classical” regime. by means of LOCC. Several heuristic arguments√ show that a measurement To formalize this insight, Yadin and Vedral propose with resolution σ & O( N) for collective spin operators and for quadrature operators in phase space to consider generalized local measurements M that map σ & O(1) a can be reasonably called classical. This is further sup- |ψi to |GHZ i, n ≤ N, with a probability p (cf. Dür na a a ported by the inequality et al. (2002); the remaining N − na qubits are no longer ∗ of interest). Then, the effective size of |ψi, N (ψ), is  IW (ρ,A)  − 2 defined as Mσ(ρ) ≤ 2 1 − e 4σ , (52) √ ∗ X where − 1 2 is the Wigner-Yanase- N (ψ) = max pana. (50) IW (ρ, A) = 2 Tr[ ρ, A] M a a Dyson skew information, which reduces to IW (Ψ,A) = (∆A)2/2 in the case of pure state |Ψi. Hence, large There exist quantum states detected as macroscop- variance as a measure for macroscopic quantumness pro- ically quantum by Yadin and Vedral (2015) that are posed by Fröwis and Dür (2012b); Lee and Jeong (2011); not identified by any other proposal (see example 4 in Shimizu and Miyadera (2002) is necessary√ for large val- Sec. II.B.1). The reason is that the LOCC paradigm – ues of Mσ(Ψ) in the case of σ = O( N). For mixed which forms the basis of this proposal– stems from entan- states, the relation to the quantum Fisher information glement theory and is not considered in other proposals. 4IW (ρ, A) ≤ F(ρ, A) ≤ 8IW (ρ, A) (Kwon et al., 2017) 18

makes further connections to the measure of Fröwis and Scaling Connection Measures Dür (2012b). to width N2 (∆A)2/N Dür et al. (2002), Korsbakken et al. (2007), Marquardt et al. (2008), B. Examples Fröwis and Dür (2012b), Park et al. (2016), Yadin and Vedral (2015), a We now present a selection of quantum states for spin Kwon et al. (2017) N ∆A Leggett (1980) (Λ), Kwon et al. ensembles and photonic systems, to which we apply the (2017)b, Sekatski et al. (2017b) measures presented in the previous section. The states √ √ N (∆A)/ N Björk and Mana (2004) and operators appearing in the examples are defined in N Independent Leggett (2002) (D)c, Shimizu and Sec. I.D and the measures are defined in Sec. II.A. Note Miyadera (2002)d that massive systems and superconducting states are dis- √ a cussed in Sec. V.D, as they are more specific examples of In case of fixed σ/ N. b In case of fixed σ. real experiments. c For a large range of , see text. Not all measures are originally conceived to be applica- d For all  = O(1). ble to both spins and photons. Measures not mentioned in an example implies that it is not clear how to apply it Table II Evaluation of generalized GHZ state, example 1, with several measures. Disregarding prefactors, we find four to this specific case. The reason might be the measure is different scaling classes for the effective size, which can be con- not applicable to the physical system (spins or photons) nected to the width of the state maximized over all collective or that it is difficult to apply it to the specific instance. observables A. In the following, approximate expressions are valid in the large N regime. We refer to further examples discussed by Volkoff and Whaley (2014a,b) and by the papers re- In addition, note that the measure of Kwon et al. (2017) N is connected to N via M (ρ) 2 N (ρ) in the limit viewed in Sec. II.A. √ eff σ . 2σ eff σ  N (see paragraph around Eq. (52)). Similarly, the measure of Sekatski et al. (2017b) has a relation to Neff q 1. Spin ensemble Neff (ρ)N in the limit of small b via MIC[b(ρ) ≈ 2b log 2 (Sekatski While connections and differences between the mea- et al., 2017b). sures are discussed later in Secs. II.C–II.E, we would like Example 1. [Generalized GHZ state] to point out similarities between some measures when This state, Eq. (5), is discussed in two different applied to pure spin states |Ψi. Fröwis and Dür (2012b); regimes. For  = π/2, one recovers the GHZ state, Park et al. (2016); Shimizu and Miyadera (2002) optimize Eq. (4), which is considered to be macroscopically quan- the variance of the state over all local operators10. This tum by all measures applicable to spin ensembles. In means that Fröwis and Dür (2012b); Park et al. (2016) the other regime,   1, Dür et al. (2002) show that will result in the same values I . Shimizu (Ψ) = Neff (Ψ) the state behaves like a GHZ state with reduced system and Miyadera (2002) are merely interested in the scal- size N2 (see Sec. II.A.2 and Fig. 2). This example was ing of the variance in , that is, I p−1 . Note N = O(N ) used by many subsequent proposals to test and scale the that there is an efficient method to calculate the max- measures. In particular, as shown in the respective pa- imal variance (Gühne et al., 2007; Ukena and Shimizu, pers, many measures12 have the same effective size in 2004). For the measure of Björk and Mana (2004) there is this parameter regime. The framework of Shimizu and some room for variations, but following some of their spin Miyadera (2002) leads to p = 2 whenever  = O(1) > 0. examples, one way of defining their measure M is to look √ The measure of Björk and Mana (2004) gives M ≈ N. at the maximal standard deviation over all collective op- Measures for which one optimizes over local operators are erators divided by the maximal spread of a spin-coherent √ maximal for A = cos S + sin S . state (i.e., ). It follows that this exactly corresponds opt x z N The behavior of the disconnectivity from Leggett the square root of the variance-based measures (Fröwis (1980) is rather different for this example. For instance and Dür, 2012b; Park et al., 2016). In summary, for pure 2 11 in the case 1  N  N, δM drops to zero only for spin ensemble states |Ψi we find that 2 Mmax & N − c/ with some characteristic constant c. 2 p−1 2 M(Ψ) = I(Ψ) = Neff (Ψ) = O(N ). (53) Hence, the disconnectivity can be much larger than N .

10 Strictly speaking, Park et al. (2016) optimize over collective op- erators. Since in all examples the states are symmetric, this leads 12 Fröwis and Dür (2012b); Korsbakken et al. (2007); Marquardt to the same results. et al. (2008); Park et al. (2016); Yadin and Vedral (2015); the 11 See table I for an overview of the mathematical symbols used for measure of Korsbakken et al. (2007) assigns the value Cδ ≈ the measures. N2/ log(1/δ) for , δ  1. 19

(a) (b)

Figure 2 The probability distribution of the generalized GHZ state, example 1, in the basis of Aopt, which is the local opera- tor maximizing the difference between the expectation values. In many papers, one considers lowest-order approximations of .

Figure 3 The measure of Kwon et al. (2017) applied to The difference becomes clear when noting that is large δM the generalized GHZ state, example 1, for different N and when the bipartite splitting M :(N − M) is entangled  ∈ {0.01, 0.02,..., 0.1} (from top to bottom). (a) N is cho- (by roughly one ebit), which is the case up to Mmax. sen such that 2N sin  = 20. For large σ, Mσ is the same This suggests to see Leggett’s extensive difference as a for all pairs (N, ), implying that these states have the same measure for macroscopic distinctness (here, Λ = N sin ) macroscopic quantumness according to Kwon et al. (2017). 2 2 and interpret the disconnectivity as a measure of quan- (b) N is chosen such that N sin  + cos  = 20 (which is the tumness (see also the discussion about separate consid- exact value of the variance divided by N). In√ order to find eration of macroscopic distinctness and quantumness in similar values of Mσ, one has to rescale σ to σ/ N. Sec. II.D.1). In order to calculate the measure of Sekatski et al. state. We consider a damping of the coherence terms of (2014c), one chooses the optimal measurement√ basis the GHZ state (i.e.,  = π/2), Aopt. In the regime in which N  N, the actual width of the two components does not play a significant |0ih1|⊗N → p |0ih1|⊗N (54) role. The effective size of the generalized GHZ state sim- ply scales as the distance between the two peaks, that is, and similar for the conjugate term. The quantum state is ± 1 Size ∝ 2N. The same result in scaling is found by apply- hence a mixture of |GHZN i with weights µ± = 2 (1 ± p), ing the framework of Sekatski et al. (2017b) in the regime respectively. In the following, all measures are evaluated b < 1, which is the relevant regime when discussing the for A = Sz, which is optimal for p & 1/N. We give the macroscopic distinctness between two peaks. effective sizes relative to the pure state p = 1. Regarding Even though the inter-peak distance and the standard the index q of Shimizu and Morimae (2005), we calculate deviation of the total state are basically the same (up Q(p) = k[Sz, [Sz, ρ]]k1. Simple calculations lead to to a factor of two), the measures of Björk and Mana Q(p)/Q(1) = p (2004) and Sekatski et al. (2014c, 2017b) differ in scaling √ 2 of . The reason is that Björk and Mana (2004) in- 2p N IS(p)/IS(1) = 1 + p2 troduce a√ normalization with “classical states” exhibiting 2 (55) a width N. A similar renormalization is –implicitly or Neff (p)/Neff (1) = p explicitly– done by all measures that find an effective size C∆(p)/C∆(1) → 1 + µ+ log µ+ + µ− log µ− of N2. p 2 The issue of renormalization also appears when apply- Mσ(p)/Mσ(1) → 1 − 1 − p ing the measure of Kwon et al. (2017) to the general- for the measures of Shimizu and Morimae (2005), Park ized GHZ state. Any measure should provide an answer et al. (2016), Fröwis and Dür (2012b), Sekatski et al. to, for example, how much N has to be increased when (2017b) and Kwon et al. (2017), respectively (see Fig. 4). decreases in order to keep the macroscopic quantum-  The last two expressions are valid in the limit ∆/N → ness constant. For this example, we find that, in the 0 and σ/N → 0, respectively. Note that p can be N- “classical regime” of the measurement resolution (i.e., √ σ dependent in common dephasing models (e.g., p = e−λN σ N), the measure M gives the same values for 2 & σ or p = e−λN ); in this case, Q(p) =6 O(N 2) in general. (N, ) if 2N sin  = const (see Fig. 3 (a)) and is hence comparable with Sekatski et al. (2014c, 2017b). If we Example 2 (Dicke states). Dicke states |N, ki, Eq. (6), want to compare Mσ with measures that find an effec- do not offer an obvious splitting |Ai + |Di and are hence 2 tive size of ≈ N√ , one has to rescale the measurement primarily addressed by measures that do not impose such resolution to σ/ N (see Fig. 3 (b)). a structure (however, see discussion later in this exam- Finally, we discuss the simplest example of a mixed ple). Since |N, ki have the same properties as |N,N − ki 20

1.0 0.08 0.8 k = 0 0.07 k = 1 k = N/2 0.6 0.06 0.05 0.4 0.04

0.2 0.03 Probabilities

p 0.02 0.2 0.4 0.6 0.8 1.0 0.01 Figure 4 Relative effective size for the noisy GHZ state, 0 -100 -50 0 50 100 Eq. (54), for measures listed in Eq. (55) (from top to bot- Spectrum tom in the order of Eq. (55)). All measures except the index 2 q scale with p for small p. For p close to one (i.e., close to Figure 5 Three examples of probability distributions for Sx purity), the first derivative lies between 1 and infinity. The and Dicke states |N, k√i with√ N = 100 and√ k ∈ {0, 1,N/2}. only nonconvex function is IS (p)/IS (1). The spread is roughly N, 3N and N/ 2, respectively.

we restrict the following discussion to 0 ≤ k ≤ N/2. the state into |Ai+|Di. Indeed, one could ask whether an Morimae et al. (2005) evaluated in index p for this state optimal cloning device could amplify the components of class in detail. They find that p = 2 for Dicke state with |0i ∝ |+i + |−i to create a superposition of two macro- k = O(N). If k = O(1), one has p = 1 and hence these scopically distinct states. Considering 1 → N cloning, states are not considered to be macroscopically quantum one finds |+i → |Ai ∝ |N, (N − 1)/2i + |N, (N + 1)/2i despite being genuinely multipartite entangled. More and |−i → |Di ∝ |N, (N − 1)/2i − |N, (N + 1)/2i for quantitatively, Fröwis and Dür (2012b); Park et al. (2016) odd N and similarly for even N (Bruß et al., 2000; find that Neff = I ≈ 2k(N − k) + 1 which confirms the D’Ariano and Macchiavello, 2003). While the total findings of Morimae et al. (2005). The optimal observ- state |Ai + |Di ∝ |Ψi = |N, (N − 1)/2i is macroscopi- able is any collective operator on the equator. We work cally quantum for all previously mentioned measures, the with Sx in the following without loss of generality. measures of Björk and Mana (2004); Korsbakken et al. Kwon et al. (2017) agree with the previous measures (2007); Marquardt et al. (2008) were shown to not classify qualitatively. Similar conclusions can also drawn apply- |Ai + |Di as a superposition of macroscopically distinct ing the measure of Sekatski et al. (2017b). states (Fröwis and Dür, 2012a).14 The disconnectivity of Leggett (1980) again shows a A similar analysis using the measure of Sekatski et al. different behavior. Except for k = 0 (which are product (2014c) gives comparable results (see Fig. 6). One finds states), all Dicke states exhibits large δmax = O(N). The that Size(Ψ) = O(N), whenever Pg < P (∆ = 1) ≈ reason is that one can divide the ensemble in ≈ k + 1 0.8183; however, Size(Ψ) . 1 for larger Pg. This example groups such that each group is entangled with the rest of clearly illustrates the influence of the success probability the spins with about one ebit. This means that δM is far Pg that also plays a role in the approach of Korsbakken from being zero up to M ≈ Nk/(k + 1). The extensive et al. (2007). The authors of the latter argue that Pg difference, Λ, is not rigorously enough defined by Leggett should be “very close” to one and hence the total state (2002) to be directly applicable to Dicke states. If it is is not considered to be a superposition of macroscopi- understood to be the spread of the probability distribu- cally distinct states (Fröwis and Dür, 2012a). This is in tion, then only for k = O(N) the spread is broad enough contrast to Sekatski et al. (2014c) who consider success to have a large Λ. probabilities in the order of 2/3. The example of |N,N/2i is interesting because it also Second, we look for a splitting that increases the effec- highlights the impact of the choice of the splitting in tive size for the measure at hand. An obvious choice for cases where |Ai and |Di are not two well-separated wave |N, (N − 1)/2i is to define |Ai as the superposition of all functions but |Ai and |Di themselves are widely spread eigenstates of Sx with positive eigenvalues (with weights 13 over the spectrum . We now discuss two possible split- coming from |N, (N − 1)/2i) and likewise |Di for the neg- tings. First, the appearance of high-k Dicke states in ative part of the spectrum. Then, the distinguishability so-called optimal covariant cloning (Bruß et al., 2000; in the approach of Sekatski et al. (2014c) is close to one D’Ariano and Macchiavello, 2003) might be used to split

14 For the measure of Björk and Mana (2004), Fröwis and Dür 13 The following analysis can be repeated for the example discussed (2012a) choose the reference state to be |Ai , which differs from by Korsbakken et al. (2007), Sec. III A, or the SQUID example the “classical reference” mentioned in the beginning of the sec- by Marquardt et al. (2008) tion. 21 also for ∆  1 (see Fig. 6). Even for Pg = 0.99, we not have a large spread in any local observable, see also numerically find a linear scaling of the size. The same Sec. IV.C.1). splitting leads to similar results for Korsbakken et al. However, a superposition of cluster states is an inter- (2007), while the framework of Marquardt et al. (2008) esting case study to understand the difference between is more difficult to apply for this splitting. some measures. Consider the operator UC , Eq. (9), that maps a product state |+i⊗N to a one-dimensional cluster Example 3 (Spin-squeezing). The discussion for spin- ⊗N ⊗N squeezed states is similar as for Dicke states, as it is state. Then, the state |Cl-GHZi ∝ UC (|+i + |−i ) a general quantum state with a broad distribution in is a state that has a similar structure as a GHZ state. some collective observable without obvious splitting into The measures of Korsbakken et al. (2007); Marquardt |Ai + |Di. Hence we only mention some specific points. et al. (2008) indeed assign an effective size of N/3 (Fröwis Spin-squeezing means that the spread of the state in one and Dür, 2012b). On the other side, every measure that is build upon local operators will still qualify this quantization axis, say S1, is reduced while keeping the state as not being macroscopically quantum because it polarization in axis S2 large. From the Heisenberg uncer- tainty relation, it follows that the state is widely spread does not exhibit a large spread in any local operator. For this reason, Fröwis and Dür (2012b) extended their over the axis S3. One-axis twisted spin-squeezed states, Eq. (7), exhibit measure to quasi-local operators. The state | Cl-GHZi † 2 5/3 has a large variance for the operator U S U , which a scaling of (∆S3) = O(N ) for the optimal squeez- C x C (i−1) (i) (i+1) ing (Kitagawa and Ueda, 1993); hence I(Ψ) = Neff = is a sum of three-body interactions σz σz σz for O(N 2/3). This means that large effective sizes can be i = 2,...,N − 1 plus two-body boundary terms. In the achieved, even if the scaling is less than for some Dicke spectrum of this operator, | Cl-GHZi is maximally spread states or the GHZ state. Thus, one has p = 5/3 and and every observable-based measure using this operator the state is not macroscopically quantum according to will find a large effective size. Shimizu and Miyadera (2002). The two-axis twisting leads to a maximal variance in S3 (Kitagawa and Ueda, 1993) and hence I(Ψ) = Neff = O(N) and p = 2. 2. Photonic systems Example 4 (Cluster states). Despite being multi-partite Similar as for spin examples, we first note a num- entangled states, cluster states are seldom considered by ber of mathematical connections between some measures. most of the discussed papers. An exception are Yadin Laghaout et al. (2015) have two elements that they com- and Vedral (2015), who propose cluster states as macro- bine to single measure. For pure, single-mode states, the scopically quantum because one can distill GHZ states element N = 1 [(∆X)2 +(∆P )2 −1] = ha†ai −|hai |2 is |GHZ i with n = O(N) out of cluster states. How- 2 Ψ Ψ n identical to the measure of Lee and Jeong (2011), I, if ap- ever, all other presented measures do not assess this state plied to pure states. In the photonic extension of Fröwis as macroscopically quantum (essentially because it does and Dür (2012b), one optimizes the quantum Fisher in- formation over all quadrature operators, which reduces to four times the variance for pure states (Oudot et al., 2015). In summary, one finds for all single-mode pure states |Ψi that

N (Ψ) = I(Ψ) 1 1 1 (56) N (Ψ) < I(Ψ) + ≤ N (Ψ). 4 eff 2 2 eff Björk and Mana (2004) use either quadrature or num- ber operators to calculate M in their example. The authors calculate the standard deviation and rescale it σ Figure 6 Success probability P as a function of the coarse- with the standard deviation of a coherent state. When graining σ for |N, (N − 1)/2i with N ∈ {101, 401, 701, 1001} using quadrature operators, this implies that M(Ψ)2 = (from left to right). Inverting this functions lead to effective size in the proposal of Sekatski et al. (2014c). The success Neff (Ψ). Sekatski et al. (2014c) generally work with pho- probability depends on the splitting |N, (N − 1)/2i ∝ |Ai + ton number measurements to calculate Size(Ψ). |Di. The “natural” splitting based the cloner scenario (solid lines, see text) leads to a reduced P σ and hence the effective Example 5 (Superposition of coherent states). An σ size is O(N) only if P . 0.8183. The splitting based on a cut archetypal instance of a photonic state with macroscopic of the spectrum into two parts (dashed lines, see text) allows quantumness is the superposition of coherent states for high distinguishability also for ∆  1. As a consequence, |SCSi (see Eq. (10) and Fig. 7 (a) for a plot of the Wigner Size(Ψ) = O(N) even for Pg close to one. function for a specific α). Without loss of generality, we 22

assume α ∈ R in the following. One easily finds15 the second example in depth for several measures and Volkoff (2015) for so-called hierarchical photonic su- † 2 2 N (SCS) = I(SCS) = ha aiSCS = α tanh α perposition states). The additivity of the measure 2 of Lee and Jeong (2011) results in a similar effective Neff (SCS) ≈ 4α + 1 (57) M 2 2 size for both states I(SCS1 ) = Mα tanh(α ) and M(SCS) ≈ 2α M 2 2 I(SCS2 ) = Mα tanh(Mα ), while the approach of Size(SCS) ≈ 4α2 M Fröwis and Dür (2012b) yields Neff (SCS1 ) = Neff (SCS) but N (SCSM ) ≈ 4Mα2 + 1. The latter two expressions are valid for large α. Note eff 2 that Size(SCS) is evaluated for |0i + |2αi. The second Example 6 (Fock states). Photon number states defined element of Laghaout et al. (2015), the distinguishability via a†a |ni = n |ni, also called Fock states, are not of- between |αi and | − αi, quickly goes to one for α & 1 ten discussed in the context of macroscopic quantumness. using homodyne detection. One reason might be the lack of a natural decomposition The SCS has macroscopic coherence also in the spirit of into |Ai+|Di. Note, however, that |ni can be written√ as a Cavalcanti and Reid (2006, 2008), in the order of S = 4α. superposition of coherent states with large (i.e., O( n)) However, the witness Eq. (22) is more suited for squeezed difference between the amplitudes α. For the measures states (see example 7) and less to detect large coherence of Fröwis and Dür (2012b); Lee and Jeong (2011), the in SCS. Consequently, the achievable lower bound is only Fock state has an effective size of Neff (n) = 2n + 1 Smax ≈ 2.5 for α = 0.5. This value is just above S = 2 and I(n) = n, respectively. Hence, Fock states become which is found for coherent states (Cavalcanti and Reid, macroscopically quantum with increasing n. 2008). Example 7 (Squeezed states). Similar to Fock states, squeezed states, Eq. (11), do not exhibit a clear |Ai+|Di structure. Measures that are not focusing on a binary di- vision typically recognize these states as macroscopically quantum for large squeezing despite having a positive Wigner function. This is traced back to the large spread in phase space in the “anti-squeezed” direction. In the following, we consider ζ ∈ R without loss of generality. This implies squeezing of the X quadrature and anti- squeezing of P . It directly follows that I(ζ) = sinh2(ζ) 2ζ (Lee and Jeong, 2011) and Neff (ζ) = e (Fröwis and Dür, 2012b), which scales linearly with the mean photon number h † i 2 . Figure 7 Plot of the Wigner function of |αi + |−αi (a) for a a = sinh (ζ) α = 2.3 and (b) for α = 4.96 with reduced visibility Γ ≈ 0.464 Squeezed states are the ideal states to verify macro- (see text). The two states exhibit the same effective size I ≈ scopic coherence in the framework of Cavalcanti and Reid 5.29 and Neff ≈ 22.2 for the measures of Lee and Jeong (2011) (2006, 2008). By measuring ∆X and by using Eq. (22), and Fröwis and Dür (2012b), respectively. From Jeong et al. 1 one can show coherence up to S = 2 ∆P . However, under (2014a). experimentally realistic conditions, a finite S can only be verified if ∆X∆P . 1.6 (Cavalcanti and Reid, 2006). Consider the case in which the visibility is nonunit, The witness Eq. (23), in which the binning of I−, I0 and that is, |αih−α| → Γ |αih−α| (see Fig. 7 (b)). The I+ is dropped (Cavalcanti and Reid, 2008), is much more measures of Fröwis and Dür (2012b); Lee and Jeong robust. More specifically, coherence up to S = 2∆P (2011) can be straightforwardly applied to this mixed independent of the value ∆X∆P can be shown. Note states. In particular, one finds I(Γ) ≈ Γ2I(SCS) and the similarity between Eq. (23) and using the Heisenberg 2 2 Neff (Γ) ≈ 4α Γ + 1. One has to go to more complex uncertainty relation, Eq. (63), to witness large quantum examples to see a difference between I and Neff (Yadin Fisher information to lower bound Neff (Fröwis and Dür and Vedral, 2016). (2012b); see Sec. II.H for details). There exist two interesting multimode extension of Similar considerations can be done for two-mode M ⊗M |SCSi: M identical copies SCS1 = |SCSi or squeezed states, Eq. (12). Cavalcanti and Reid (2006, M ⊗M ⊗M 2008); Fröwis and Dür (2012b); Lee and Jeong (2011) give the entangled version |SCS2 i ∝ |αi + |−αi (see also Volkoff and Whaley (2014b), who studied basically the same results as for single-mode squeezed states. In addition, the framework of Sekatski et al. (2014c) can be modified to be applicable to two-mode states (Oudot et al., 2015). The idea is to rephrase the 15 (Björk and Mana, 2004; Fröwis and Dür, 2012b; Laghaout et al., original formulation by having a bipartite, entangled sys- 2015; Lee and Jeong, 2011; Sekatski et al., 2014c) tem, where one party “Alice” prepares the state of the 23

other party “Bob” by means of local measurement (i.e., a unitary transformations. In any case, this choice defines kind of steering scenario). Then, Bob has to guess which a preferred basis |ki and a notion of spectral distance 0 state Alice measured using only measurements with fi- ak − ak0 between any two eigenstates |ki and |k i. Thus, nite resolution. The quantum state used for this protocol a measure based on a preferred observable is a function is said to be a superposition of macroscopically distinct of the spectral distribution and the coherence properties states if Bob is able to distinguish different preparation of the considered state. with very low resolution. Using coarse-grained√ quadra- (ii) In some cases the authors start by identifying a ture operators, this results in Size = O( N), where N is specific partition of the system into subsystems17, for ex- the mean photon number of the states. Note that this is ample, H = H1 ⊗ H2 ... . Often, like for spin systems, the optimal scaling in this scenario. the subsystems correspond to physical particles18, but can be also identified with a locality argument (Shimizu and Miyadera, 2002) for systems with a spatial extent. C. Classifications of measures Though the starting point here is different from (i), in some cases the final measure can be equally well un- In the previous sections we reviewed several proposals derstood from the perspective of a preferred operator to quantify macroscopic quantumness and discussed rel- (Fröwis and Dür, 2012b; Park et al., 2016; Shimizu and evant examples. In this and the following sections, we Miyadera, 2002). This is because the measures are com- discuss and compare different contributions (see table I). puted via optimizing some property of the state over a set However, the measures do not all share the same mo- of local operators. This corresponds to the identification tivation or goals, neither they have the same range of of a preferred observable A. applicability. Hence, one observes a family resemblance (iii) The third approach is to look at a preferred rep- situation, where objects in a set are not connected by a resentation of the considered state; here, it is the phase unique common feature, but rather they are connected by space for position and momentum of a particle or conju- a series of features, each shared by some but not all of the gated quadratures of a bosonic mode. This is the starting objects. The goal of the following subsections is there- point of Laghaout et al. (2015); Lee and Jeong (2011), fore to discuss several recurring aspects of the measures who focus on the the Wigner function of a given state. and to identify common features and differences within The quantification of high-frequency components leads subsets of measures. Some original proposals leave room to an equivalent formulation of their measure based on for interpretation or small modifications that influence a decoherence model. Despite the different motivation, the final measure. Hence we understand this comparison Nimmrichter et al. (2011) is directly related to Lee and not as an ultimate judgment but as a starting point for Jeong (2011) (Yadin and Vedral, 2016)19. This is be- further discussions. cause the modification of quantum mechanics considered in Nimmrichter et al. (2011) also corresponds to a con- tinuous spontaneous localization in phase space of the 1. Mechanisms to break unitary equivalence particles. Alternatively, one may say that the starting point here is the choice of a preferred Lindbladian su- In the introduction of Sec. II.A, we drew the reader’s P † 1 † peroperation L(%) = j(Lj%Lj − 2 {LjLj,%}), which is attention to the structure that breaks the unitary equiva- used to model dissipative evolution in terms of a master lence between quantum states. This is necessary in order equation. Interestingly, the observables that appear as to define a measure of macroscopic quantumness and to Lindblad operators in the master equation of the deco- introduce a hierarchy of macroscopic quantumness. On herence are linear in the quadrature operators and fre- a formal level this requires to identify some additional quently appear in the proposals of group (i). structure over the Hilbert space H associated with the physical system. After reviewing the literature, we are now in the position to comment more on the mechanisms 2. Goals of the measures to break the unitary equivalence between all pure states. We identify at least three basic approaches. All proposals aim for measuring a kind of macroscopic (i) One can choose a preferred hermitian operator A = quantumness, but one can identify different variations in P R 16 k ak |kihk| (or dk) as a starting point . This operator can be an observable that is measured or a generator of

17 Dür et al. (2002); Fröwis and Dür (2012b); Korsbakken et al. (2007); Leggett (1980); Marquardt et al. (2008); Park et al. (2016); Shimizu and Miyadera (2002); Yadin and Vedral (2015) 16 Björk and Mana (2004); Cavalcanti and Reid (2006, 2008); Kwon 18 Dür et al. (2002); Fröwis and Dür (2012b); Korsbakken et al. et al. (2017); Laghaout et al. (2015); Leggett (2002); Sekatski (2007); Leggett (1980); Yadin and Vedral (2015) et al. (2014c, 2017b); the “extensive difference” part of Leggett 19 The connection is valid if only the ideal quantum state plus the (2002) and “subjective” part of Laghaout et al. (2015). canonically introduced modification is considered. 24

formulating the precise goal. (iv) Relative improvement. Björk and Mana (2004); (i) Macroscopic coherence. For a fixed operator A = Fröwis and Dür (2012b) motivated their proposals with P k ak |kihk|, the macroscopic quantumness of a state is situations where the superposition of (quasi-)classical related to its coherence spread20, that is, the amount of states overcome the performance of its single branches. coherence between basis state |ki, |k0i that are far apart Although not explicitly mentioned, it seems that “macro- 0 in the spectrum |ak − ak|  1. The superposition of far- scopic quantum effects” like the single excitation men- distant parts of the spectrum is seen as a macroscopic tioned in Sec. I.A should be excluded. Instead, the quantum effect if the system size is large. In the case authors focus on the speed of unitary evolution or on P of a pure state |Ψi = ψk |ki, one often considers the quantum-enhanced sensing. While the motivation might 2 variance of the state (∆A)|Ψi, or the total amount of differ from (i), the choice of the figure of merit leads to P coherence with distance ≥ S, that is, 0 |ψkψk0 |. measures based on the variance for pure states (Fröwis |ak−ak|>S (ii) Effective particle number. Some measures 21 ex- and Dür, 2012b) or at least is closely connected to it in plicitly aim to quantify how many particles effectively relevant situations (Björk and Mana, 2004). participate to a macroscopic superposition. As an ex- (v) Falsification of collapse models. The proposal by ample (Korsbakken et al., 2009), consider a system of n Nimmrichter and Hornberger (2013) is explicitly moti- electrons in a pure quantum state building up the mag- vated by evaluating quantum states (more precisely, en- netic flux in a superconducting device (see Sec. V.D). tire experiments) that potentially show small deviations Suppose that the induced magnetic moment from a su- from standard quantum mechanics. Surely there are perposition of right |Ai and left |Di circulating currents connections to (i), since typical collapse models predict larger collapse rates for larger spatial spread of the mas- has a spread ∆µ = nµB, where µB denotes the Bohr mag- neton. For Korsbakken et al. (2009), it matters whether sive particle. Likewise, heavier objects generally trigger a single electron makes the difference between the |Ai the hypothetical collapse much faster. However, if one is and |Di or whether all electrons behave differently in the tempted to analyze the entire experiment including “un- two branches. In contrast, representatives from (i) might avoidable” imperfections and ask how much it is capa- care less because a single particle in a superposition of ble to show modifications of quantum mechanics, larger macroscopically different momenta could be seen as one quantum systems (e.g., larger masses) are not necessarily with a very large effective mass. From this example, the more useful (i.e., give a larger number) for such tests as concept of effective particle number seems to be more environmentally induced decoherence is more disturbing restrictive than (i) as the large spread for some exten- (Nimmrichter et al., 2014). sive quantity necessarily appears whenever the effective (vi) Amount of nonclassicality. Lee and Jeong (2011) particle number is large (see Sec. II.D). identify the total amplitude of wiggles in the Wigner (iii) Ease to distinguish. Macroscopic quantumness is function for the state of one or several optical modes. identified with the easy to distinguish between the super- As coherent states correspond to classical fields and their posed components in the state (Korsbakken et al., 2007; Wigner function is smooth, such wiggles are considered Laghaout et al., 2015; Sekatski et al., 2014c, 2017b). The as a trace of nonclassicality of the state. Similarly, Dür ease can be quantified by measuring how invasive a mea- et al. (2002); Laghaout et al. (2015); Park et al. (2016) are surement apparatus should be in order to extract the de- concerned with the nonclassical aspects of macroscopic sired information about the superposition state, e.g., dis- quantum states. tinguish |Ai and |Di with the desired probability. Such a non-invasive measurement apparatus can be modeled D. Structure of applied states as a weak measurement of a fixed operator A (Sekatski et al., 2014c, 2017b), or as a general measurement that The presented literature is rather diverse regarding the only acts on a limited number of subsystems (Korsbakken application to quantum states. While Dür et al. (2002) et al., 2007). In the case of weak measurement of a fixed basically discuss a class of examples, others22 derive a observable, easily distinguishable states clearly exhibit a measure assuming an explicit decomposition of the state large macroscopic coherence (i), as the superposed states into |Ai + |Di. In the following, we call states that are have to be far enough in the spectrum of A. macroscopically distinct for a fixed partition |Ai + |Di

20 Björk and Mana (2004); Cavalcanti and Reid (2006, 2008); Fröwis and Dür (2012b); Kwon et al. (2017); Laghaout et al. (2015); Lee 22 Björk and Mana (2004); Korsbakken et al. (2007); Laghaout et al. and Jeong (2011); Leggett (2002); Park et al. (2016); Shimizu and (2015); Marquardt et al. (2008); Sekatski et al. (2014c); Björk Morimae (2005); the “extensive difference” of Leggett (2002), the and Mana (2004) phrase their idea for general states, but the “objective” part of Laghaout et al. (2015) and the large-σ regime mathematical formulation for Eq. (19) is derived for the |Ai+|Di of Kwon et al. (2017). structure. Laghaout et al. (2015) extend the |Ai + |Di structure 21 Dür et al. (2002); Korsbakken et al. (2007); Leggett (1980); Mar- to more than two states, but still need to provide the substruc- quardt et al. (2008); Yadin and Vedral (2015) ture. 25

macroscopic superpositions. The index p from Shimizu local observable that is particularly suited to distinguish and Miyadera (2002) is applicable to pure states with- |Ai and |Di. In other words, the most obvious decompo- out determining any substructure. Finally, there exist sition does not have to lead to the maximal size in these several proposals23 that are defined for arbitrary mixed cases. This ambiguity is particularly relevant when real states. Large-size states beyond the |Ai + |Di structure experiments are discussed (e.g., Marquardt et al. (2008) are here called macroscopic quantum states. Let us fur- discussing the experiment of van der Wal et al. (2000)). ther elaborate on these differences. The ambiguity is lifted (though not fully) when the state is given in the micro-macro entangled form

1. Macroscopic superpositions |Ψi = |↑i |Ai + |↓i |Di , (58)

a. Separate treatment of macroscopic distinctness and quan- where the macroscopic system is entangled with an tumness. In most of the cases, measures designed for a “atom” like in Schrödinger’s thought experiment. The |Ai + |Di structure aim for quantifying the macroscopic entanglement in Eq. (58) reduces the complexity of distinctness between the two components, but they do the problem and it suffices to find the optimal distin- not measure the coherence between the components. A guishability in a two-dimensional subspace spanned by mixed state |AihA| + |DihD| has equally distinct compo- span{|Ai , |Di}. This is a feasible task in many exam- nents, but it has no coherence and hence cannot be called ples. a macroscopic quantum state. Therefore, the distinct- ness of a macroscopic superposition and its quantumness appear as two independent values. Consequently, show- 2. Macroscopic quantum states ing that a state contains distant components only reveals its macroscopic distinctness. This is not sufficient if the Measures for macroscopic quantum states quantify the state can not be assumed to be pure. In practice, an macroscopic quantumness of a general state ρ, which does additional measurement is required to certify the quan- not need to be pure or have a required structure. The tumness of the superposition state, for example, through macroscopic distinctness and the coherence are hence the purity of the state, some entanglement to an aux- measured with a single number. For reasons of clarity, it iliary system or some interferometric visibility between is useful to first consider the application of the measures |Ai and |Di. This contrasts measures that aim to give a to pure states, for which the interpretation is more di- single number quantifying macroscopic quantumness. rect. After that, we discuss how they are lifted to mixed states.

b. Freedom for choosing the branching. Measures for macroscopic superpositions often provide an intuitive ac- a. Variance as recurring measure for pure macroscopic quan- count for characterizing macroscopic distinctness. Hav- tum states. In most of the cases, the formalization of ing the Schrödinger-cat example in mind, it is easy to measures for macroscopic quantum states amounts to a connect the concept of macroscopic distinctness with the study of coherences of the state in the basis of a given op- thought experiment. Macroscopic distinctness can mean, erator A. For pure states, once A is fixed, the variance of for instance, that measuring only one biological cell is the state with respect to A is a natural choice that allows enough to determine the vitality of the cat (Korsbakken to quantify the coherent spread of the state. Indeed, de- 24 et al., 2007); that it is necessary to modify the state all spite the diversity, several proposals directly have the N cells from “alive” to “dead” (Marquardt et al., 2008); variance of the state (for given A) at the core of their or that a quick (i.e. coarse-grained) look suffices to de- measures when applied to pure state. Leggett (2002) de- cide whether the cat is dead or alive (Sekatski et al., fines the extensive difference as the spectral distance of 2014c). On the other side, given just a pure state |Ψi, |Ai and |Di for a well-chosen extensive observable, which there are infinitely many ways to split it in two branches is a special case of the variance. Moreover, while Björk |Ψi = |Ai+|Di. As in example 2 in Sec. II.B, the decom- and Mana (2004) do not explicitly use the variance in position can become problematic when the state does not their approach, the connection between the variance and show two well-isolated distributions in the spectrum of a improved sensitivity is well established for pure states (see Secs. II.H and IV.B). In addition, there are measures that do not reduce to the variance but still are closely related. Cavalcanti 23 Cavalcanti and Reid (2006, 2008); Fröwis and Dür (2012b); Kwon et al. (2017); Leggett (1980); Sekatski et al. (2017b); Shimizu and Morimae (2005); Yadin and Vedral (2015); the basic idea of Cavalcanti and Reid (2006, 2008) is for general states, but the derived bounds for the detection basically work only for squeezed, 24 Fröwis and Dür (2012b); Lee and Jeong (2011); Park et al. not necessarily pure, states. (2016); Shimizu and Miyadera (2002) 26

and Reid (2006, 2008) are interested in the coherence be- macroscopic” states. Specifically, there exist examples tween distant parts of the spectrum | hk| ρ |ji | > 0, with for pure states with p = 1 (i.e., the variance is O(N)), |ak −aj| ≥ S, exhibited by the state. For pure states it is but q = 1.5 (Shimizu and Morimae, 2016). This suggests easy to see that any state with variance (∆A)2 ≥ V nec- that the intuition for the index q does not entirely corre- essarily exhibits coherence between√ spectral parts with spond to coherent spread of pure states as measured by a distance of at least S = 4 V .25 Hence, a large vari- the variance. In particular, there is no general relation ance is sufficient for the presence of large coherence S in between the quantum Fisher information and the index the state. However, it is not necessary, as a state with q. arbitrarily large S might have an arbitrarily small V . The measure I of Lee and Jeong (2011) is similar to the In addition, the measure of Sekatski et al. (2017b) gives quantum Fisher information for low-rank instances (see a lower bound to an expression proportional to the vari- example 5). However, they differ in general (Yadin and ance (and in general to the quantum Fisher information). Vedral, 2016). Furthermore, Yadin and Vedral (2016) For general b, this bound can be arbitrarily loose but it pointed out a problematic relation of I with the two- becomes tight in the limit of low information gain b. norm (see Sec. II.G). Finally, the variance is fully sufficient to capture the spread of a probability distribution in the limit of many copies of the system (Yadin and Vedral, 2016). All this 3. Connections between some measures shows an important role for the variance as quantifier of a specific aspect of macroscopic quantumness. In this section, we review connections between some measures. In particular, one finds for some partition- based measures that macroscopic quantum states in- b. Extensions of the variance for mixed states. Since the cludes the concept of macroscopic superposition (Fröwis presence or absence of coherence in the basis of A does and Dür (2012b); see Fig. 8). Let us consider quantum not influence the variance (∆A)2, it can only be used as a states |Ai + |Di for spin ensembles. Fröwis and Dür measure of macroscopic quantumness if the state is pure. (2012b) show that Cδ = O(N) (see Eq. (24)) if and only ¯ The generalization of the variance for mixed states is not if D = O(N) (see Eq. (27)), which suggests a connec- unique and the proposals of Fröwis and Dür (2012b); Lee tion between the measures of Korsbakken et al. (2007) and Jeong (2011); Park et al. (2016); Shimizu and Mori- and Marquardt et al. (2008). Next, if Cδ = O(N) then mae (2005) all go in different directions. there exists a quasi-local operator A (see Sec. I.D.1) for 2 2 In principle, any measure f for pure states can be ex- which the variance scales quadratically (∆A) = O(N ). tended to a measure fˆ for mixed states via the convex Given that one accepts the step from local to quasi-local roof construction (e.g., see Sec. II.A.11). To this end, operators in the formulation of measures for macroscopic consider all possible pure state decompositions (PSD) of quantum states, this implies a large effective size for all P measures based on the variance (see example 4). the state ρ = k πk |ψkihψk| and minimize the average size

ˆ X f(ρ) = min πkf(ψk). (59) Macroscopic superpositions PSD k

By definition, the convex roof is the “worst-case” aver- Macroscopic quantum states age f among all possible pure-state decompositions of ρ. At the same time, fˆ is the largest convex function Figure 8 Hierarchy between states that are macroscopically that reduces to f when ρ is pure (Tóth and Petz, 2013). quantum according to some partition-based measures. Macro- The quantum Fisher information, which appears in the scopic superpositions according to Korsbakken et al. (2007); measure of Fröwis and Dür (2012b), is (four times) the Marquardt et al. (2008) are also macroscopic quantum states convex roof of the variance (Yu, 2013). according to variance-based measures if quasi-local operators are considered (Fröwis and Dür, 2012b). The index q was introduced as a generalization of the index p for mixed states (see Sec. II.A.3). Shimizu and Further connections can be found between general Morimae (2005) are mainly interested in identifying “fully measures26 (see Secs. II.B and II.E.2). macroscopic” quantum states, for which the two measures Independent of these connections, the measure of match p = q = 2. But this is no longer true for “less Sekatski et al. (2017b) applies ideas for macroscopic su- perpositions (such as Korsbakken et al. (2007); Sekatski

25 The same argument holds for mixed states if the variance is re- placed by the quantum Fisher information, since the latter is the 26 Cavalcanti and Reid (2006); Fröwis and Dür (2012b); Kwon et al. convex roof of the variance (Yu, 2013) (2017); Sekatski et al. (2017b) 27 et al. (2014c)) to general quantum states. This can be number operator A = a†a for measuring the macroscopic seen as a way to bridge measures for macroscopic super- spread. A large variance in photon number implies that positions and macroscopic quantum states. large Fock states are involved in the superposition. How- Note that one can also spot significant differences be- ever, it does not generally imply large mean photon num- tween the measures, in particular, when the effective size ber as examples with a fixed average photon number and is not O(N). An interesting example are the SQUID ex- arbitrarily large variance show.27 In addition, note that periments as discussed in Sec. V.D, for which diametrical coherent states |αi have a variance that increases linearly results are obtained. with the mean number of photons |α|2 for the photon number, but remains constant for any quadrature. The simultaneous presence of quantumness and large E. Application to various physical setups system size for large-variance states in spin and photonic systems may encourage us to use it in other systems, The proposed measures sometimes attempt to be ap- such as superconducting devices or massive systems. Af- plicable to all physical systems (such as spin ensembles, ter all, photonic states, superconducting circuits and mo- photons, superconducting devices and massive systems). tional degrees of freedom of massive systems are math- Nevertheless, they are typically motivated with a specific ematically connected by using the same algebraic struc- setup in mind and one has to analyze the implications ture with the canonical commutation relation [a†, a] = 1. when applied to other systems. In the following, we focus However, the interpretation of the variance for photons on some general issues regarding the application of mea- is not trivially transferable to other systems. For exam- sures for macroscopic quantum states to various physical ple, the SQUID experiments from Friedman et al. (2000); systems. This analysis is done for the variance. How- Hime et al. (2006); van der Wal et al. (2000) lead to ever, it can be partially repeated for other, more specific a controversial discussion of whether a macroscopic su- measures. perposition of clockwise and anti-clockwise currents has been generated. While this is reviewed in more detail in Sec. V.D, we just summarize here the analysis of Björk 1. Implications of large variance and Mana (2004). Effectively, they calculate the variance of the ideal target state and find it to be roughly 1000 Let us start with spin ensembles, for which the oper- times larger than the variance of the ground state, which ator at hand is typically chosen to be a local operator is assumed to be the most classical state.28 However, a PN (l) (l) A = l=1 al . Since ||al || = O(1), the spectral radius clear interpretation of this number, for example, as an of the total operator is proportional to the number of effective size of the electronic state is missing so far. spins, N. A state is said to be macroscopically quan- For massive systems, the situation is even more puz- tum if the standard deviation (i.e., the square root of zling. For a fixed mass, it seems that the spread of the the variance) is large compared to the spectral radius. A wave function in the spatial degree of freedom is a natu- large number of spins is necessary for a large variance. ral choice to measure the macroscopic quantumness of the Thus, this automatically leads to a connection between system. However, the connection between large variance the large-scale character (i.e., the spread in the spectrum) in position and large system size is lost. Moreover for par- with quantumness (through the coherence between the tition based measures (Korsbakken et al., 2009; Leggett, eigenstates). The connection to multipartite entangle- 2002) one does not change the macroscopic quantum ment (in particular for the quantum Fisher information, character of, for example, a nucleus by increasing the dis- see Sec. II.F) further supports the use of the variance to tance of a superposition of being “here and there” from 1 measure macroscopic quantumness of pure spin-ensemble cm to 1 m, as the states of individual and neu- states. trons in the two branches are equally orthogonal in both For photons, one encounters a very similar situation. cases. Similarly, it is an open question how the parti- For simplicity, we discuss single-mode states in the fol- tion in subsystems should be made, for example it can lowing. By taking quadrature operators A = Xθ = be done equally well on the level of atoms or protons √1 eiθa + e−iθa† as the canonical choice for the vari- and neutrons. For the same reason the role of the mass 2 ance, a large variance necessarily comes from many- of the individual constituents, and how it interacts with † 1 2 2 the spread in position, is open. photon states since ha ai ≥ 2 [(∆X) + (∆P ) − 1]. A large variance also implies more nonclassicality of Note that if we go away from macroscopic distinct- the photonic state in the sense that k-partite mode- ness and ask about quantum states/experiments that ex- entanglement can be created by splitting the photonic mode into many spatial modes in a beamsplitter net- √ work. The same effect is achieved by letting the photons 27 For instance, |ψi = p1 − 1/k |0i + 1/ k |ki for k  1. be absorbed by a sufficiently large atomic ensemble (see 28 Note that Björk and Mana (2004) work with the standard devi- Sec. II.E.2). Alternatively one can choose the photon ation rather than the variance. 28

clude small deviations from quantum mechanics, we in- with LOCC. This yields a direct connection to entangle- deed care about the distance between the two possible ment theory, in which LOCC is the set of free operations. positions of the neutron. This is because typical col- For other partition-based measures, LOCC are not free lapse models are more effective for larger distances, and in general and might change the macroscopic quantum are have a precise mass dependence. From this point character of a state (e.g., the variance can increase un- of view, the mathematical connection between (Nimm- der LOCC). Nevertheless, it is very natural to look for richter and Hornberger, 2013) and (Lee and Jeong, 2011) a relation between entanglement properties and macro- (as mentioned in Sec. II.C.2) seems to be loose since the scopic quantumness of a state, given that both have the interpretation and meaning of the respective measures partition of the system in subsystem as a starting point. are different. The measure of Fröwis and Dür (2012b) based on the quantum Fisher information is related to multipartite en- tanglement in the following sense. An effective size of 2. Linking measures for photons and spins Neff & k implies genuine entanglement within groups of Recently, Fröwis et al. (2015) compared measures that at least k spins (this statement can be made fully rigor- are primarily defined for spin ensembles and photonic ous including O(1) correction terms (Hyllus et al., 2012; systems. The authors consider an ideal -matter in- Tóth, 2012)). Furthermore, large Neff is not only a wit- teraction between a spin ensemble in the ground state ness for k-partite entanglement in the sense of Sørensen and an incoming photonic field. Under the assumption and Mølmer (2001), but Neff & k also means that the that the number of spins is much larger than the num- two-body correlations within these entangled groups have ber of photons, a photon is linearly mapped to an atomic a certain minimal strength (Tóth, 2012). Note also that excitation. In technical words, the first-order approxima- since a high value for the measures of Korsbakken et al. tion of the Holstein-Primakoff transformation (Holstein (2007) and Marquardt et al. (2008) implies a large quan- √ tum Fisher information, these measures are also sufficient and Primakoff, 1940) allows one to identify S−/ N ↔ a. This enables one to compare quantum states from the two of a large entanglement depth. Furthermore, Morimae physical systems and eventually entire measures defined (2010) proves that a state with q = 2, Eq. (16), contains for spins or photons. For example, it turns out that the multipartite entanglement measured with the distance to measure of Korsbakken et al. (2007) could be formulated the set of separable states. for a single photonic mode by asking how well |Ai and On the other hand there is no measure for which a large |Di can be distinguished with highly inefficient detectors entanglement depth on its own is sufficient for macro- (i.e., ideal detectors with preceding photon loss). This scopic quantumness, as follows from the example of the resembles the ideas of Sekatski et al. (2014c), who consid- state |W i = |N, 1i discussed in the sec. I.A. This state is ered coarse-graining instead of loss. Even though the dif- is fully non-separable (exhibits N-partite entanglement), ferent realization of the detector’s imperfection changes but is not recognized as a macroscopic quantum state by the measure qualitatively, a strong conceptual connec- any of measures listed above. tion of two or more measures leads to confidence in the principal idea. As another example, the connections be- tween several works29 are further reinforced. Not only the measures are based on the variance (or related to it), G. Connection to resource theory of coherence the canonical choice of the operators (local operators for spin ensembles and quadratures in the photonic case) are In recent years, there has been a trend to formalize intimately connected via this ideal light-matter interac- certain aspects of quantum mechanics such as entangle- tion. ment (Horodecki et al., 2009), athermality (Goold et al., 2016; Gour et al., 2015), asymmetry (Gour and Spekkens, F. Connection to multipartite entanglement 2008) and coherence (Baumgratz et al., 2014; Marvian and Spekkens, 2016; Yadin et al., 2016) in the framework Many authors do not consider multipartite entangle- of resource theories. The basic idea is to assume certain ment per se when constructing a measure of macroscopic constraints on the generation and manipulation of quan- quantumness. An exception is the measure of Yadin and tum states. For this, one defines free states and free oper- Vedral (2015), who identify the macroscopic quantum- ations. Quantum states that are not producible from free ness for the size of the maximally distillable GHZ state states and free operations constitute a resource for a task with typically some quantum advantage. The most well- known resource theory is entanglement in multipartite 29 Cavalcanti and Reid (2006, 2008); Fröwis and Dür (2012b); Lee settings, in which free operations are LOCC. Free states and Jeong (2011); Shimizu and Miyadera (2002); Shimizu and are states that can be generated starting with product Morimae (2005) states and using LOCC. 29

1. Resource theory for macroscopic coherence D(ρ, Φ(ρ)) is a valid measure in the sense of Yadin and Vedral (2016). However, as pointed out by Kwon et al., The concept of macroscopic distinctness is naturally the measure behaves in a surprising way if σ is small (i.e., connected to superposing states from different parts of a the distance between the original state and a strongly given spectrum. This suggests to work with a resource dephased stated is measured, see Sec. II.A.14). If the de- theory of coherence where not only a basis is chosen but phasing is strong (i.e., σ is small), the coherence between√ a spectrum is associated to it. This is the framework even nearby basis states is lost. In the case σ = O( N), of asymmetry or “unspeakable coherence” (Marvian and the measure coincides with the intuition for basic exam- Spekkens, 2016). There, quantum states that are not in- ples. variant under translations in the spectrum are a resource to detect the presence of processes that are generated by these translations. Yadin and Vedral (2016) realized 2. Free operations in proposed measures the similarity between asymmetry and the attempts to define macroscopic quantum states. Like in asymme- The arguments of Kwon et al. (2017) suggest that while try, they define free operations as completely positive the axioms of Yadin and Vedral (2016) might be a good maps that cannot increase coherence between parts of starting point for a resource theory of macroscopic quan- a spectrum with a certain distance. This is in contrast tumness, they seem to be not sufficient to unambigu- to many variants of “speakable coherence” (Marvian and ously identify relevant measures for macroscopic quan- Spekkens, 2016), in which coherence between neighbor- tum states. A further noteworthy point is that fully in- ing basis states can be freely transferred to coherence coherent states in one basis generally exhibit large coher- between distant states. The motivation for this choice of ence in an other basis. Since in many examples for pho- free operations is obvious as it makes the creation of co- tonic systems or spin ensembles it is not possible to de- herence a “difficult” task. Note, however, that with this termine a single preferred observable but we have a set of choice of free operations it is not more difficult to cre- observables to choose from, further improvements for the ate far-distant coherence than coherence between nearby choice of free operations could lead to a resource theory states. This is achieved by additionally requiring that that captures our intuition of macroscopic distinctness any measure of quantum macroscopicity –in addition to even better. Finally, the right choice of free operations is be nonincreasing under free operations– should assign nontrivial in general (Marvian and Spekkens, 2016). For larger values to superpositions of two basis states with example, LOCC in entanglement theory is not the most increasing spectral distance. general class of operations that do not increase entangle- It turns out that the variance is a valid measure for ment in the system. Yet it is the preferred set from a this resource theory of quantum macroscopicity for pure physical point of view. states. So the present resource theory agrees with all We emphasize that many measures discussed in measures of general macroscopic quantum states that use Sec. II.A implicitly or explicitly choose a set of free opera- the variance for pure states30. The quantum Fisher infor- tions. Let us briefly comment on this choice for measures mation, which is used by Fröwis and Dür (2012b) as an designed for spin systems. First, even a collective local ⊗N extension for mixed states, is the convex roof of variance rotation U can formally affect the macroscopic quan- (Yu, 2013) and hence a valid measure in the framework of tumness of the state for a measure that is defined for a Yadin and Vedral (2016). In contrast, the measure of Lee fixed operator (e.g., a collective operator like A = Sz). 31 and Jeong (2011) generally increase under the present In practice , however, one typically chooses an opti- 0 ~ free operations because of a problematic connection to mal direction of the collective observable, A = ~n · S. 0 †⊗N ⊗N the Hilbert-Schmidt norm, which is known to be not con- This leads to the choice A = U SzU and hence ⊗N tractive under trace-preserving operations (Ozawa, 2000; collective rotations U are implicitly considered to be Piani, 2012; Yadin and Vedral, 2016). Whether the index free. The next level is to consider all local unitary (LU) q (Shimizu and Morimae, 2005) fulfills the requirements transformations (i.e., U1 ⊗ · · · ⊗ UN ) as free. This is im- is open (see Sec. II.D.3). plicitly done in Fröwis and Dür (2012b); Shimizu and In this context, the work of Kwon et al. (2017) gives Morimae (2005) as one chooses the optimal local opera- further insight. Consider a distance function D(ρ, τ) that tor that maximizes the variance of the state. Next, one is (D1) positive semidefinite, (D2) invariant under joint could consider a single round of local operations (i.e., unitary rotations of ρ and τ, (D3) contractive under phys- measurements) followed by local unitaries (LO+LU). In 2 ical maps and (D4) jointly convex. Using the dephasing fact, the state |GHZN/2i, with variance (N/2) , can be prepared from a one-dimensional cluster state composed map Φσ(ρ) defined in Sec. II.A.14, the authors show that

30 Fröwis and Dür (2012b); Lee and Jeong (2011); Park et al. (2016); 31 For example, Björk and Mana (2004); Kwon et al. (2017); Park Shimizu and Miyadera (2002) et al. (2016); Sekatski et al. (2017b) 30

of N qubits with variance N + O(1), by measuring ev- operators. Implicitly, this idea is present in the following ery second qubit in the σx with subsequent local rota- findings. tions on the remaining qubits depending on the mea- The way Björk and Mana (2004) formalize the idea of surement results (Fröwis and Dür, 2012b; Yadin and Ve- interference utility is precisely in the spirit of Leggett dral, 2015). Hence, under the assumption that LO+LU (1980). A broad distribution in the spectrum of the are free, cluster states become macroscopically quantum. Hamiltonian H is directly connected to the maximal Finally, Yadin and Vedral (2015) consider all LOCC as “speed” of evolution via the inequality (Fleming, 1973; being free operations. Mandelstam and Tamm, 1945)

| hψ| e−iHt |ψi |2 ≥ cos2 (∆Ht) (60) H. How to determine the effective size in experiments for all ∆H|t| ≤ π/2. This inequality tells us that wit- An important point is the applicability of the measures nessing a fast change of the state implies wide-spread to real experiments. In addition of being convincing on coherence spread in the spectrum of H. Equation (60) a theoretical level, the effective size of a measure should can be directly generalized to mixed states (Fröwis et al., ideally be extractable from experimental data. In this 2008), but the inequality generally becomes loose for low section, we examine some proposed protocols where this purity. is possible without a full state tomography. Luckily, the intimate connection between a sensitive notation of statistical distance measured by the Fubini- Study metric and the variance of the generator has a 1. Macroscopic distinctness well-behaving extension to mixed states (Braunstein and Caves, 1994; Uhlmann, 1991; Wootters, 1981). There, Macroscopic distinctness as defined by Korsbakken the equivalent metric is the Bures metric dsB and the et al. (2007); Sekatski et al. (2014c) can theoretically be variance is replaced by the quantum Fisher information. measured in the lab. For the effective number of parti- One then finds that cles (Korsbakken et al., 2007), one needs to have access 1p to single particles or at least small groups of them. The dsB = F(ρ, H)dt, (61) distinguishability is given by the optimal measurement. 2 If this is not available in the lab, any measurement will where dsB measures the infinitesimal distance between give a lower bound on the effective size. In the case of ρ(t) and ρ(t + dt) generated by H. This relation has classical detectors (Sekatski et al., 2014c), it is neces- far-reaching consequences for the foundations and appli- sary to have a detector that can be modeled in the way cations of quantum mechanics. Here, we are interested presented in Sec. II.A.11, ideally with a tunable resolu- in measuring lower bounds on the quantum Fisher infor- tion. For both approaches, one has to generate either mation using Eq. (61). This can be done in several ways. |Ai or |Di and measure the probability distribution as a First, one can directly tighten Eq. (60). One replaces the √ √ function of number of measured particles or detector res- left-hand side by the fidelity, F (ρ, σ) = [Trp ρσ ρ]2, olution. From this distribution, one can calculate the ef- between the initial and final state and, in addition,√ fective size. Similarly, the “subjective” part of Laghaout the standard deviation on the right-hand side by F/2 et al. (2015) can be measured by preparing the single (Fröwis, 2012). After choosing a fixed measurement with components |bki, but a special access to subsystems or outcomes {x}, the probability distribution before, p(x), tunable detectors is not necessary. The quantumness, as and after, q(x), the evolution of duration t can be used discussed in Sec. II.D.1, is measured independently. to bound the quantum Fisher information (Fröwis et al., 2016)

2. Macroscopic coherence 4 X p F(ρ, A) ≥ arccos2 p(x)q(x). (62) t2 Already Leggett (1980) noted that quantum states x with large-scale quantum correlations are only distin- Second, one can derive a tighter version of the Heisenberg guishable from mixtures if O(N) particle correlations are uncertainty relation33 measured.32 He argued that a way out is the unitary time evolution even for local or two-body Hamiltonians hi[A, B]i2 F(ρ, A) ≥ . (63) as the expansion of exp(−iHt) contains O(N) correlation (∆B)2

32 An extreme example is the GHZ state, Eq. (4), for which omitting 33 Fröwis et al. (2015); Hotta and Ozawa (2004); Kholevo (1974); a single particle is enough to hide all essential quantum features. Pezzé and Smerzi (2009) 31

By measuring the variance of B and the expectation value 4. Correlations for index q of i[A, B] one is able to find lower bounds on F(ρ, A). Note the similarity between Eq. (63) and Eq. (23), The index q from Shimizu and Morimae (2005), whose conceptual closeness becomes even more evident Eq. (16), is in principle measurable without full state when studying the proofs of the bounds. This highlights tomography. We consider a scenario in which one has the connection between the ideas of Cavalcanti and Reid access to single particle measurements. Then, the goal is (2006, 2008) and Fröwis and Dür (2012b). to measure two-body correlations in the spirit of Eq. (15), but in several conjugate bases. This is a way to generalize the concept of Bell inequalities or entanglement witnesses to multipartite settings. By verifying O(N 2) pairs with 3. Measuring loss of coherence O(1) correlations, one can conclude that the present state exhibits q = 2 (Shimizu and Morimae, 2005). The measure of Nimmrichter and Hornberger (2013) is specifically designed to be experimentally accessible. Ev- ery experiment that excludes a certain parameter regime I. Summary and conclusion of a collapse model has a value in the framework of Many different aspects of macroscopic quantumness Sec. II.A.10, which is directly measured by the amount are covered by a growing number of proposals. In view of of time over which quantum coherence is maintained. our initial starting point –to formalize the idea of macro- The contribution of Lee and Jeong (2011) is also scopic distinctness in Schrödinger’s cat example– a pos- strongly connected to the susceptibility under decoher- sible conclusion is the following. ence. The larger the loss of purity by applying a canon- Measures based on an explicit |Ai + |Di structure ical noise channel, the more macroscopically quantum clearly work well whenever |Ai and |Di themselves are the state is according to the measure I. Jeong et al. localized in a given spectrum in which the two states are (2014a) showed that I can be experimentally estimated maximally distinguishable (i.e., |Ai and |Di behave “clas- without the need of a full state tomography. The basic sically”, see example 1). In this regime, a macroscopic idea is to extract the purity from an overlap measure- superposition is a special case of a macroscopic quan- ment of two identically prepared states (see Fig. 9). This tum states as characterized by more general measures. is done by comparing the purity of the state before and However, it is open whether the original intention of an after a short application by the decoherence channel de- effective particle number is maintained for general states, scribed by Eq. (30). An overlap measurement realizing in particular if |Ai and |Di themselves are nonclassical 2 Trρ can be implemented with a SWAP operation be- (see example 2). tween the two systems, followed by a suitable measure- To measure macroscopic distinctness, it seems appro- ment of both modes. For a single-mode photonic state, priate to find functions that evaluate the spread of prob- this can be realized using a beamsplitter and a photon- ability distribution in the spectrum of a certain opera- number resolving detector after one of the modes. Sim- tor (or a set of possible operators). If the underlying ilarly, one can combine the two copies of the state with quantum state is pure, the coherence between far dis- auxiliary systems and use controlled SWAP operations tant parts is then a signature of macroscopic quantum- (Jeong et al., 2014a). Furthermore, the scheme is adapt- ness. The variance is proposed by many authors and is a able to the measure of Park et al. (2016) for spin ensem- good choice to classify the global structure of the prob- bles. ability distribution. More sophisticated measures (e.g., entropies and distant functions in combination with ad- ditional parameters) are able to resolve finer aspects. It is important to have a proper scaling of the func- tion that measures macroscopic distinctness for a given observable. For example, the variance of a product state in spin ensembles increases linearly for collective opera- Figure 9 Basic scheme of Jeong et al. (2014a) to experi- mentally access the measures of Lee and Jeong (2011); Park tors. However, this should not be seen as macroscopic et al. (2016). Two identical copies of a state ρ are gener- quantum effect but as an accumulation of coherence on ated. Then, either both copies are subject to decoherence, the single-particle level. In this example, a pure product Eq. (30), or not. In both cases, the purity of ρ is measured state is then considered as a state carrying “one unit of via an overlap measurement to determine Trρ2 (which is not quantumness” and hence a measure has to be appropri- the fidelity between the two states). Comparing the purity in ately rescaled. the presence and absence of decoherence allows to estimate The presented measures are applied to various physical I(ρ). From Jeong et al. (2014a). systems. While the (rescaled) variance arguably makes sense for spins and photons, its applicability when deal- 32 ing with spatial or superconducting degrees of freedom not necessarily macroscopically quantum in a sense as is open. On the other hand, for massive systems and discussed in Sec. II, but only on a coarse-grained level. superconducting devices, measuring the (potential) falsi- We summarize in Sec. III.E. fication of collapse models has a clear operational mean- ing. The extension to mixed states should satisfy basic re- A. Maintaining macroscopic quantum states quirements from information theory. The convex-roof construction for measures defined for pure states gener- Let us start by discussing the limitations arising from ally fulfills these conditions, but it is not the only option. the impossibility to maintain the state of the system by In the case of the variance, the convex-roof extension shielding it perfectly from the environment. leads to the quantum Fisher information with the addi- tional benefit of having tight and accessible lower bounds. Currently, the quest of a well-motivated set of free op- 1. Decoherence erations for macroscopic quantumness will further help to classify and understand macroscopic quantumness. On The decoherence program (Zurek, 2003a,b) can be the more practical side, there is a trend to make the pre- viewed as a general explanation why quantum features sented measures applicable to experimental data. do not prevail at a macroscopic scale. Every quantum system interacts with its environment and becomes en- tangled. As the environment cannot be controlled, it has III. LIMITS FOR OBSERVING QUANTUM PROPERTIES to be traced out and the system-environment entangle- IN MACROSCOPIC STATES ment manifests itself in decoherence of the system, i.e., a certain mixedness due to the lack of knowledge on the en- The measures that were presented in Sec. II provide vironment leading to a reduction of quantum coherences. various ways to characterize sets of macroscopic states As we show below a very simple instance of the and hence to study the sensitivity of these states to dif- generic effective quantum-to-classical transition for states ferent noise in a systematic way. This reveals certain of bosonic fields follows from the seminal works of Husimi limits and inherent difficulties to prepare, maintain and (1940), Glauber (1963), Sudarshan (1963) and others. observe macroscopic states. In the following we will in- Any state of a bosonic field ρ can be faithfully repre- vestigate these limitations, and summarize a number of sented by quasi-probability distributions P, W and Q results that were obtained in this context. In a certain (Vogel and Welsch, 2006). In particular, the Glauber- sense, they suggest that the very same features that make Sudarshan P-representation is an expansion of the state a state macroscopic also make it susceptible to noise, dif- in the over-complete basis of coherent states ficult to prepare and maintain and almost impossible to Z measure. ρ = P(α) |αihα| d2α. (64) We start by discussing how quantum states are af- fected by noise and decoherence in Sec. III.A. To this If P(α) ≥ 0 is positive the state ρ is a mere mixture of aim we first consider some general results on decoher- coherent states and is said to be classical. This terminol- ence that are applicable to all quantum states, and then ogy arises from the fact that the coherent states can be state more specifically how macroscopic quantum super- thought of as the most classical subset of the set of all position states and general macroscopic quantum states possible state of a mode of a quantum field: they saturate are affected. In Sec. III.B, we then turn to the mea- the uncertainty relations for all pairs of quadratures, and surement of such states, and show how limited detector also phases and number of photons. Furthermore, they efficiency and resolution hinders the detection of macro- are eigenstates of the forward part of the field operator, scopic quantumness. We also discuss the closely related and they remain within the set of coherent states under but slightly less demanding task of certifying quantum passive operations. For instance coherence states are the states (prior to the action of noise), which however re- only pure states of the field that do not generate entan- quires (exponentially) growing resources for macroscopic glement when sent on a beam splitter. Negative values of superpositions. We shortly touch the issue of preparing the P-function indicate intrinsic quantum features. The states in Sec. III.C. Finally, in Sec. III.D, we present a P-function is related to the Husimi Q-function via a con- number of methods and approaches to circumvent some volution with a Gaussian, of the mentioned problems. Most of these methods are however only applicable in a limited sense, in particular Z 2 Q P −|β−α| 2 (65) for certain, specific kinds of noise. The only exception is (α) = (β)e d β. quantum error correction, where we argue that encoded 1 quantum superposition states can indeed be prepared, However, Q(α) = π trρ |αihα| is a probability distribu- maintained and measured. Theses states are however tion and hence is always positive. A direct consequence 33 of these two observations is that, regardless of the ini- effective size of a macroscopic superposition (Dür et al., tial state ρ, a decoherence process E that acts on the 2002) took the fragility of the coherences, i.e., the in- P-representation of a state as a convolution renders the trinsic quantum features of a superposition, as a starting state classical – the P-function of the state after the de- point. Hence, any state that is macroscopic in this sense coherence ρ0 = E(ρ) is positive. It is easy to see that such is by definition also fragile to noise. Later attempts to a decoherence process is given by thermal noise, i.e., the provide measures for macroscopic quantumness do not diffusion of the state in phase-space generated by contain a direct reference to the behavior of the state under noise and decoherence. Nevertheless, one can still Z −i(λ1X+λ2P ) i(λ1X+λ2P ) 2 relate the features responsible for macroscopic quantum- E(ρ) = e ρ e gσ(λ)d λ, (66) ness to the behavior of the states under certain noise processes as we now discuss. 1 −|λ|2/(2σ2) with a Gaussian gσ(λ) = 2πσ2 e , the random variable λ = λ1 with σ = 1 and two conjugate quadra- λ2 tures X and P . A similar observation holds for spins. 2. Fragility of macroscopic quantum superpositions In this case quasi-probability distributions P(n), the ex- pansion in spin-coherent states, and Q(n) are defined on We first consider macroscopic quantum superpositions. the sphere (Agarwal, 1998; Arecchi et al., 1972), and P In order to confirm that this is indeed a macroscopic is mapped to Q by spherical smoothing (Agarwal, 1998; quantum superposition, one needs to show that |Ai and Schmied, 2017). The noise that maps P to Q in this case |Di are macroscopically distinct, and that the state is corresponds to a random rotation of the state generated indeed quantum, i.e., a coherent superposition and not i λ·S by e with the total spin operator S and a “spherical an incoherent classical mixture. Since these two aspects Gaussian” random variable λ. are independent we can formulate our question as fol- Decoherence is a generic mechanism that concerns lows: how does the decay of the quantumness under quantum features in general, and i.e., it is not specific noise is affected by the macroscopic distinctness of the to macroscopic quantum states or macroscopic quantum state? Let us consider the original situation imagined by superpositions, discussed in this article. Nevertheless, it Schrödinger where the macroscopic system M is entan- is commonly believed that macroscopic quantum super- gled with a microscopic atom A as |↑iA |AiM +|↓iA |DiM , positions are particularly affected by decoherence (Zurek, the quantumness of the state is then identified with the 2003b), and therefore are particularly fragile and hard to entanglement between the systems M and the A.34 We observe. Such beliefs are supported by a growing mani- are interested in the decay of entanglement in the state fold of examples, but also generic statements, which we after the action of some noise channel on the macroscopic will discuss in this section. Let us mention a few early system, which can always be represented by a unitary in- results. Caldeira and Leggett (1985) discussed the in- teraction of the system with the environment. stability of superpositions of Gaussian wave packets in The example of the N-particle GHZ state, with |Ai a harmonic potential weakly coupled to a bath. Yurke = |0i⊗N and |Di = |1i⊗N , mentioned earlier is a good and Stoler (1986) emphasized the same effect for the su- starting point. It is enough for the environment to mea- perposition of coherent states under loss. Milburn and sure a single particle in order to extract the which-branch Holmes (1986) demonstrated that quantum signatures of information and collapse the superposition to a mix- an initial coherent state evolving in an anharmonic po- ture losing all entanglement. Importantly, the proba- tential and coupled to a bath disappear faster when the bility that no particle is measured by the environment of the initial state is larger. This list can be easily decreases exponentially with N. This makes the GHZ extended. For instance, consider the effect of the ther- state such fragile. The crucial property is the ease to mal noise of Eq. (64) on a superposition of coherent states distinguish the two branches, which increases with N. |αi+|−αi. From the expansion of the coherent state state √ As one could expect, such a link between fragility and −1/4 (x∓ 2α)2/2 in the X-quadrature basis h±α|xi = π e macroscopic distinctness can be generally made for any (for real α) it directly follows that for large α the coher- measure based on the ease to distinguish between the two −α2σ2 ence E(|αih−α|) is damped by a factor ∝ e . Another branches (cf. Sec. II.C.2 and Sekatski et al. (2014b)). prominent example is the GHZ state, for which it easy to The proposal of Korsbakken et al. (2007) (Sec. II.A.6) see that by interaction with independent local environ- implies that a macroscopic superposition is fragile with ments the coherences (|0ih1|)⊗N also decay exponentially fast ∝ e−γNt, as single-qubit coherences |0ih1| decay as ∝ e−γt. See (Aolita et al., 2010; Cavalcanti et al., 2009; Simon and Kempe, 2002) for detailed studies of the ef- 34 The presence of the atom is very instructive but by no means fect of noise on the entanglement in GHZ and other graph necessary. In its absence the quantumness can be identified as states. the distance of the state from the mixture, or as the amplitude Interestingly, one of the first attempts to provide an of the coherence term |AihD| in the density matrix. 34 respect to projectively measuring small parts of the sys- state. Given a pure initial state |ΨihΨ|, Shimizu and tem or simply losing these particles to the environment. Miyadera (2002) considered the effect of weak local noise This follows from the observation that if the two branches on the purity of the state by analyzing its decay rate can be distinguished by measuring only a small number of 1 d 2 subsystems, then even a tiny amount of loss leaks enough Γ = − ln Tr[ρ(t) ]|τct1/Γ, (67) particle to the environment to allow it to fully collapse 2 dt the state to one of the two branches. where τc is a correlation time of the noise process. The Similarly, a superposition state that is macroscopic re- local noise process is modeled as a local classical noise P garding classical measurements of an operator A (see Sec. described by the Hamiltonian H = λ x f(x)ˆa(x) with II.A.11), i.e., measurements that have in general some a random variable f(x) and a local operator aˆ(x) acting finite resolution and only disturb the system weakly, is on the system at position x, or as local environmental P ˆ fragile with respect to a dephasing noise generated by the induced decoherence described by H = λ x f(x)ˆa(x) same operator. This follows from the observation that where fˆ(x) is a local operator at position x of the envi- such a noise channel can be represented as a weak mea- ronment. Shimizu and Miyadera found that, quite gener- surement of the observable A by the environment. Again ically for both kinds of weak disturbances, for a superposition of macroscopically distinct states even 2 X † a tiny amount of noise allow the environment to obtain Γ ≈ λ g(k) hΨ| δAkδAk |Ψi , (68) full which-branch information and destroy the quantum- k ness of the state. where λ is the small interaction strength, g(k) = O(1) This implies that a coherent superposition of two P −ikx is the spectral intensity and δAk = x e (ˆa(x) − macroscopically distinct quantum states very quickly be- hΨ| aˆ(x) |Ψi) is a collective operator. Hence, the rela- comes a mixture under the effect of decoherence. There- tive decoherence rate Γ/N is always constant if the ini- fore the very same feature that defines macroscopic dis- tial state has an O(N) variance with respect to all lo- tinctness makes the maintenance of the quantumness of cal operators Ak. On the other hand, Γ/N is system the superposition state extremely challenging. size dependent if there exists an Ak ≡ A for which We remark that a similar result was shown (Sekatski (∆A)2 = O(N 1+) with  > 0 (cf. the effective size et al., 2017b) for states with a more general structure max (∆A)2/N, Sec. II.A.9). P √ A:local n pn |niA |AniM , for which the entanglement of for- Similar results were found for the phase space when mation is also increasingly fragile with respect to the the decoherence is generated by quadrature operators. measure of Sec. II.A.11. Lee and Jeong (2011) showed that their measure can be equivalently defined as the susceptibility of the state to the noise process given by loss of photons I(Ψ) ≈ Γ (see 3. Sensitivity of macroscopic quantum states Sec. II.A.8).

A separation between macroscopicity and quantum- ness for a given state is typically not made for general b. Susceptibility regarding change of state. Shimizu and macroscopic quantum states. Therefore, the argument Miyadera (2002) define the stability of the state under lo- from the previous section cannot be easily extended to cal measurements through the correlations observed be- arbitrary states. Here, we present several ways to tackle tween two local observables a(x) and b(y) at two dif- the question about the fragility of macroscopic quantum ferent locations. More precisely they say that a state states: (a) First, one can quantify the effect of noise on a ρ is stable under local measurements if for any ε > 0 state by measuring how fast the state becomes mixed un- the correlation between the observable is low enough der the effect of noise. (b) Second, the effect of noise on |P (b|a) − P (a)| ≤ ε for sufficiently large distance |x − y|. a state can be quantified by asking how much the state Shimizu and Miyadera (2002) show that for this defini- itself is susceptible or stable with respect to noise, i.e., tion a quantum state is insensitive to local measurements how far does the state get from itself after the action of if and only if it has the cluster property (i.e., states with the noise. (c) Finally, one can directly analyze how the p = 1 in Eq. (14)). Otherwise, the measurement of a macroscopic quantumness of the state is affected by the single spin l might significantly change the probability noise. In other words one can bound the maximal size distribution for another spin l0. This becomes evident of the state that is the output of some noise process, or when inspecting Eq. (15): A large variance implies the derive requirements on the noise that have to be satisfied presence some two-body correlations between different 0 in order to prepare states of a desired size. spins, hence the measurement results of a(l) and a(l ) are correlated. The disturbance-based measure Mσ(ρ) proposed by a. Susceptibility regarding purity. The purity of a state, Kwon et al. (2017) (see Sec. II.A.14) by definition quan- Trρ2, measures how close a quantum state ρ is to a pure tifies how far the state gets from itself after the action 35

of the map Φσ, which corresponds to the coarse-grained B. Measuring and detecting macroscopic quantum states measurement of an operator A. Since the map Φσ can be equivalently seen as a dephasing noise generated by the We will now discuss limitations to observe macroscopic same operator A, it follows that macroscopic quantum quantum states with imperfect or size-limited measure- states are particularly sensitive to such noise. ment devices.

1. Coarse-graining and control of measurements

c. Fragility of macroscopic quantum states under noise. Another reason which is sometimes invoked to explain Let us start with the measure by Fröwis and Dür the apparent absence of quantum effects on macroscopic (2012b) based on the quantum Fisher information (see scale is the limited resolution of measurements apparatus. Sec. II.A.9). In quantum metrology, the question about The intuition here is that the detection of quantum effects the maximally attainable quantum Fisher information in such as quantum superpositions or entanglement on a the presence of noise is central, as the quantum Fisher macroscopic scale require high measurement precision. information is related to the precision of the estimation There is a long list of examples confirming this intuition. protocol. Hence, as summarized in Sec. IV.B, many tools Early works by Mermin (1980) and Peres (2002) ana- were developed to upper-bound the quantum Fisher in- lyzed the task of witnessing the entanglement of a singlet formation of a protocol given a Hamiltonian and a noise state (total spin zero) of two large spins by measuring 35 36 process. Using these tools it can be shown that the the spin operators Sn on the two spins. Mermin showed quantum Fisher information of any state after the ac- that when the size of the spins increases one requires a tion of generic local noise processes for any operator better and better control on the angle of the measure- 37 A can only scale as O(N). This limitation is rather ments, the direction of the measured spin n, in order to 38 severe. For instance, for the depolarizing noise with demonstrate entanglement. Peres showed that the reso- only 1% error probability per particle, one is limited lution of the measurement 1/σ required to demonstrate to F/(4N) ≤ 10. For the measure of Fröwis and Dür entanglement also becomes more stringent with the size (2012b) this means that the effective size can not exceed of a spin. A lack of resolution of the measurement of ten. In other words the same quantum Fisher informa- P an operator A = k ak |kihk| refers to coarse-graining of tion can be attained by a state where spins are only en- the measurement outcomes. Formally, to account for a tangled within groups of size ten. finite coarse-graining σ, the POVM elements of an ideal Similarly, Park and Jeong (2016) showed that thermal- measurement Ek = |kihk| are modified as ization suffices to destroy macroscopic quantum states. σ X 0 0 To this end, the authors argued that after thermalization Ek = |kihk| → Ek = nσ(ak − ak0 ) |k ihk | , (69) the variance of all local operators behaves extensively, k0 (∆A)2 ∼ O(N) (as it is assumed in ). with some distribution nσ(x) of width σ (often taken to Carlisle et al. (2015) used the measure of Lee and Jeong be a Gaussian distribution or a square function). Similar (2011) (see Sec. II.A.8) to assess the achievable size of results for coarse-grained measurements of Stokes oper- macroscopic quantum superpositions in optomechanical ators39 have been obtained by Simon and Bouwmeester set-ups. They showed that it is in general very hard to (2003) for multi-photon singlet states, and by Raeisi et al. obtain macroscopic superpositions, and in fact require (2011) for states where micro-macro entanglement is gen- a large single-photon optomechanical coupling strength erated via parametric amplification of one photon from and postselection. an entangled pair (De Martini and Sciarrino, 2012; Fröwis and Dür, 2012a; Sekatski et al., 2009). In the case of bosonic fields, coarse-grained measure- ments of quadratures Xθ are mathematically equivalent to decoherence as discussed in Sec. III.A.1. Gaussian 35 Demkowicz-Dobrzanski et al. (2012); Escher et al. (2011); Fuji- coarse-graining of quadrature measurements wara and Imai (2008); Sekatski et al. (2017a) 36 2 A. Lopez Incera et al., in preparation. (Xθ −x) 37 1 − In the case of qubits the only exception is the Pauli noise given δ(Xθ − x) → √ e 2σ2 (70) by E(ρ) = pρ + (1 − p)σnρ σn, under which the quantum Fisher 2πσ information can still scale quadratically. As an example, recall that a Dicke state is an eigenstate of such noise process for σn = σZ . 38 Local depolarizing noise is described by a completely positive 39 The cases of photonic states discussed here is very similar to the (a) trace preserving map acting on qubit a, E (p)ρ = pρ + case of large spins, as the Stokes operators for two bosonic modes 1−p P3 (a) (a) 1 † † 1 † † −i † † 4 j=0 σj ρσj , where 1 − p is called the error probabil- Jz = 2 (a a − b b), Jx = 2 (a b + a b ) and Jy = 2 (a − a b ) ity. form the Schwinger representation of the spin operators. 36

q can be modeled by inserting a noise channel, Eq.(64), ~ δθ ∼ cRM . Hence, any measurement suffers from some that acts on the state just before the measurement. This intrinsic amount of coarse graining due to its finite size. is exactly the noise channel that diffuses the state in Along the same lines it was recently shown in Skotini- phase space. Hence coarse-graining of quadratures is suf- otis et al. (2017) that the limited size of measurement ficient to make all observable measurement statistics re- devices forbids the observation of superposition states producible by “classical” states (those having a positive |Ai |Di that are macroscopically distinct in the center P-function). + of mass position, total spin or energy. The authors adopt In the case of spins it is a lack of control on the mea- a formalism in which the total system consisting of the surement angle n which is in direct correspondence to superposition state and the measurement device (refer- decoherence. Indeed the lack of control over the angle ence frame) is closed and has to abide to the fundamental can be modeled by applying a random rotation on the symmetries of nature given by the Galilean group. To ex- state prior to the measurement. Again this is exactly plicitly account for the lack of any additional reference the noise process that maps the P-function of the spin to frame, a twirling map is applied on the total system. Un- its Q-function, demonstrating that a lack of control on der the assumption that the state of the reference frame the angle is sufficient to wash out non-classicality from is classical, it is then shown that in order to distinguish the observed statistics.40 In addition, the strength of the superposition from the mixture the size of the mea- the angular noise (i.e., the width of the “spherical Gaus- surement device has to be quadratically bigger than the sian” that is convoluted with the P-function) decreases size of the superposition. On the other hand it is also with the size of the spin (Schmied, 2017)41. Kofler and shown that superpositions in relative degrees of freedom Brukner (2008) showed that a coarse-graining of the spin do not suffer from such limitations. measurement approximately correspond to a lack of con- While these fundamental limitations are not a prob- trol on the angle, hence if the amount of coarse-graining√ σ is larger then the square root of the total spin size N lem for microscopic or even mesoscopic experiments, it the observed measurement statistics can be explained by becomes highly relevant when considering true macro- classical states. This, however, does not imply that no scopic superpositions. For example, Skotiniotis et al. quantum features can ever be observed with measure- (2017) show within a simple model that in order to ob- ment devices with limited stability or/and precision, as serve a superposition state of the size of a cat one requires the system can be manipulated prior to the measurement a reference frame of the size of the Earth. or in-between subsequent measurements. We will come back to this in Sec. III.D.3.

C. Preparation of macroscopic quantum states 2. Reference frames and the size of measurement apparatus Let us now focus on the difficulties to prepare macro- Precision and stability of measurement devices is a scopic quantum states. Several works have considered technical problem that can be in principle overcome by the question whether macroscopic superposition states technological efforts. On the other hand there are lim- or macroscopic quantum states can be ground states of itations that can not be overcome, for example the size “physical” Hamiltonians. If this would be the case, one of the measurement device, that are intrinsically limited could generate these states simply by means of cooling. (by the size of the in the most optimistic case). However, this is not the case for all states. For example, it It turns out that such limitation become relevant when was shown that GHZ state, Eq. (4), cannot be the unique one considers macroscopic quantum effects. ground states of a quasi-local Hamiltonian, but requires Using Heisenberg uncertainty relations and relativistic at least one global interaction term in the Hamiltonian causality Kofler and Brukner (2010) demonstrated that a that affects all particles (Van den Nest et al., 2008). A measurement device of mass M and size R can only lead similar result was found in Dakić and Radonjić (2017) for to a spin measurement that has an angular resolution of general macroscopic superposition states. In particular, it was shown that local Hamiltonians that have macro- scopic superposition states as unique ground states have a vanishing energy gap, which in turn requires the presence 40 Note this does not mean that one can not distinguish classical of a long-ranged interaction term in the Hamiltonian. and non-classical states with such measurement. If the noise in the measurement is well characterized, it can be in principle On the other hand, there exist unique ground states of deconvoluted from the observed statistics (cf. Sec. III.D.1). two-body Hamiltonians that have a variance of 2 , 41 O(N ) This can be intuitively understood from the observation that the i.e., these states are macroscopically quantum according overlap between two spin coherent states at different angles de- creases with the size of the spin | hn|⊗N |n0i⊗N | = |n · n0|N . to several measures. One such example is an N-qubit Therefore non-classical features of the state (regions with a neg- Dicke state with N/2 excitations (see example 2). This ative P-function) get more and more narrow with increasing N. state is the unique ground state of two-body Hamilto- 37

nians42 and might be efficiently prepared by means of tween macroscopic distinctness of |Ai and |Di with re- cooling, or even occur naturally. Notice, however, that spect to coarse-grained measurements of an operator A for Dicke states the corresponding Hamiltonian is not in Sec. II.A.11 and the incertifiability of their superpo- local in the sense of arranging qubits on a lattice, but sition under dephasing noise generated by A has been contains long-range two-body interaction terms. discussed by Sekatski et al. (2014b). Conceptually, the incertifiability of macroscopic quan- tum superposition is quite a strong statement. It shows D. Further limitations and counter-strategies that in addition to be hard to maintain, even detecting traces of the superposition in the final state is extremely Even though there are severe limitations and restric- hard in practice, as it requires to increase the total du- tions to prepare, maintain and measure macroscopic ration of the experiment exponentially in the size of the quantum states, there exist some counter-strategies and superposition state. bypasses that allow one to circumvent these limitations to a certain extend. On the other hand, (Fröwis et al., 2013) also show that a quantum state is certifiable in presence of depolarizing noise if it is a unique ground state of a gapped quasi-local 1. Certifiability of large-scale quantum states Hamiltonian. Certifiable here means that one can distin- guish the initial state from all orthogonal states with only One of the main goals of Sec. III.A.1 was to understand polynomially many repetitions. While any state after how much of macroscopic quantumness can survive in a the action of local depolarizing noise has linear quantum state after the action of some noise process. We saw that Fisher information (see Sec. III.A.3.c), one can show that quite generically macroscopic quantum states are fragile, for the Dicke N/2 state, for instance, a quadratic Fisher and thus very hard to observe in practice. Nevertheless, information of the initial state can be certified with any one can be less demanding and ask whether the macro- desired statistical confidence (P-value) with only O(N 4) scopic character of the state prior to the noise can be repetitions of the measurement. This makes the Dicke certified by measuring the final state and assuming that state a promising candidate for the experimental detec- the noise process is perfectly known. If the noise process tion of a macroscopic quantum state, however performing E is not too strong (more precisely, if its action on the O(N 4) repetitions can still be quite challenging for large vector space of operators B(H) has a trivial kernel), then N. the state of the system prior to the noise can always be re- constructed to any desired precision by collecting enough measurement statistics on the final state. But the crucial 2. Counter-strategies against noise question then is how many times one has to repeat the measurements in order to collect enough statistics. Fröwis et al. (2013) showed that any superposition For particular noise channels, passive or active strate- gies might be available to maintain macroscopic quan- state state |Ai + |Di that is macroscopic distinct by the criterion of Korsbakken et al. (2010) (see Sec. II.A.6) tum states or macroscopic quantum superpositions. Such is incertifiable whenever a (tiny) amount of noise local strategies were particularly discussed in the context of depolarization noise acts on the state. By incertifiabil- quantum metrology, where states with a large quantum ity the authors mean that the number of repetitions that Fisher information are required to maintain a quantum allows to distinguish the initial macroscopic superposi- scaling advantage. In Landini et al. (2014) the usage of particular states in a decoherence free subspace are dis- tion state |Ai + |Di from the orthogonal state |Ai − |Di cussed to protect the system against collective dephasing. (and hence also from the mixture |AihA| + |DihD|) in- creases exponentially with its size. The reason for this Active quantum error correction is used in Arrad et al. is that the coherence between macroscopically distinct (2014); Dür et al. (2014); Kessler et al. (2014) to pro- tect the system against a specific noise process, namely states E(|AihD|) is exponentially damped by local depo- larizing noise, such that one needs exponentially many rank-one noise that is orthogonal to the sensing field. copies in order to distinguish between the macroscopic Fast quantum control is used in Sekatski et al. (2017a) superposition and the corresponding mixture with a con- to maintain the usability in quantum metrology for any stant probability. This observation is directly related to rank-one noise. The methods are similarly applicable the fragility of the macroscopic superposition states dis- to actively maintain a large quantum Fisher information cussed in Sec. III.A.2. In particular, the connection be- under certain noise processes. However, we emphasize that only some very specific noise processes can be dealt with in this way. The ulti- mate bounds for generic noise processes reported in Sec. 42 As a simple exercise one can construct this Hamiltonian using III.A.3.c still apply, making a quadratic quantum Fisher 2 2 the total spin operators S , Sz and Sz (Fröwis et al., 2013). information generically inaccessible. 38

3. Counter-strategies against coarse-graining trarily small. On the other hand, quantum error cor- rection or fault-tolerant quantum computation (Nielsen The results discussed in Sec. III.B.1 always considered and Chuang, 2000) can be used to actively protect quan- a restricted set of operators that can be measured with tum information. Hence, this gives us the tools to pre- coarse-grained detectors: the spin in some direction Sn pare, maintain and certify encoded macroscopic quantum for spin systems and the quadratures Xθ for photons. states. For this, one has to accept a slight shift of the def- Without any restriction, the impact of coarse-graining inition of macroscopic quantum states, which we discuss would be weak in general. To see this, imagine an ob- in this section. servable with the same eigenbasis as before, but with a In quantum error correction, quantum information is clever rearrangement of the eigenvalues. In this way, one actively protected against the influence of noise and de- can easily get rid of the difficulty to distinguish neigh- coherence by encoding a logical two-level system into sev- boring eigenstates of the original operator. While such eral physical two level systems, i.e., into a larger space. a rearrangement of eigenvalues might seem rather ab- Each qubit of a quantum state |ψi is thereby replaced stract at first glance, something similar can be physically by a logical qubit that consists of several physical qubits. done by applying a unitary U on the system just before For example, the state |φi = α|0i + β|1i is replaced by it is measured (Kofler and Brukner, 2008). Later Jeong |φLi = α|0Li + β|1Li, where now the quantum informa- et al. (2009) showed that a simple Kerr-nonlinear inter- tion, i.e., the parameters α and β are protected. Codes action Hamiltonian H is sufficient to generate unitaries can be designed to protect quantum information against U = e−itH that allow to observe quantum features (vio- different kinds of noise, most notable noise that is local lation of macrorealism, see Sec. IV.A.1 for more details) in some sense, i.e., acting jointly in a correlated way only with extremely coarse-grained detectors. on a bounded, localized number of qubits. However, this strategy might add additional con- Note that operations for encoding and error correc- straints. Wang et al. (2013) analyzed the superposition tion might themselves be noisy. The theory of fault- of two coherent states |αi+|−αi in the setting of coarse- tolerant quantum computation (Nielsen and Chuang, grained quadrature measurements and the possibility to 2000) tells us, however, that if noise in elementary single- perform a Kerr-nonlinearity. The authors showed that and two-qubit operations is below a certain threshold while by using such a Kerr interaction it is possible to value, then one can perform error correction in a fault- distinguish the superposition from a mixture even with tolerant way. In fact, any quantum computation can a low resolution, the requirement on the control preci- be realized fault-tolerantly. In particular, an encoded sion of the interaction (the exact value of the interaction macroscopic quantum√ superposition state | L-GHZN i = ⊗N ⊗N time t) also increases with the size of the superposition (|0Li +|1Li )/ 2 (Fröwis and Dür, 2011) can be cre- |α|2. It is open whether these findings can be extended ated, maintained and measured using noisy elementary to general macroscopic quantum states. operations, and under the influence of (quasi-)local deco- A similar setting was analyzed for the task of observ- herence. The same can be done for any other macroscopic ing macroscopic quantum states. In Fröwis et al. (2016), quantum state via the mapping |ii → |iLi , i = 0, 1. Note a method to detect large quantum Fisher information us- that these encoded states are still susceptible against ing detectors with limited resolution was presented. The noise at the logical level, which is however exponentially main idea is to re-use the same operation as to prepare suppressed by error correction if it results from noise on the state (e.g., a squeezing operation) in order to real- individual qubits. ize a modified measurement process. While the initial We emphasis that mapping a macroscopic quantum measurement device is coarse-grained, the squeezing op- state to its encoded version might drastically change its eration allows one to increase the relevant resolution, and effective size according to some measures presented in to detect a quantum Fisher information that, for exist- Sec. II. For instance, measures that maximize over all lo- ing set-ups, could be two orders of magnitude larger. A cal operators and do not consider an extension to quasi- similar approach was used in Davis et al. (2016). In both local operators generally do not assign a large effective cases, the requirements on the stability of the squeezing size to encoded states. The logical GHZ state | L-GHZN i, operation was analyzed. for example, does not have a large variance with respect to any local operator (which is necessary to ensure ro- bustness), but it exhibits maximal variance for the sum of 4. Encoded macroscopic quantum states logical Pauli operators σz,L = |0Lih0L|−|1Lih1L|. There- fore, it depends on the precise definition of the measure The results discussed in previous sections show that (and hence on us) whether we call this state macroscopi- generic noise processes, in particular independent cou- cally quantum or not. Note that other measures (e.g., the pling of system particles to the environment, destroy measures of Korsbakken et al. (2007); Marquardt et al. macroscopic superpositions and macroscopic quantum (2008)) do not decrease under the encoding (cf. example 4 states, even if each of the local noise processes is arbi- in Sec. II.B.1). Considering | L-GHZN i as a toy example 39 of a Schrödinger cat in a quantum superposition state be- quantum entanglement and its classification and quan- tween two macroscopically distinct states, one could say tification (Horodecki et al., 2009). We are rather con- that the distinguishability between different constituents cerned with large (macroscopic) quantum systems and of the cat is here not at the level of the individual atoms, their potential applications. Naturally, this is interlinked but at the level of molecules. In this sense, one may still with certain issues of entanglement. We start however call this a macroscopic quantum superposition. by considering the role of macroscopic quantum states in a more fundamental issue, namely for probing the limits of quantum theory. E. Summary

In this section, we have reviewed many different find- A. Probing the limits of quantum theory ings that show that macroscopic quantum states are gen- erally difficult to observe. First, such states are difficult Quantum mechanics provides an accurate description to maintain because even weak perturbations leak enough of the microscopic world and is in fact the most accu- information to the environment in order to lose the coher- rate description of nature we have come up with so far. ence between macroscopically distinct states. Second, a However at macroscopic scales, quantum effects typically lack of resolution of the detectors or simply the finite size can not be observed. So the question remains if quantum of the measurement devices also seems to make the ob- mechanics is indeed valid on all scales. Here we discuss servations of quantum features on the macroscopic scale two approaches that are related to this issue. difficult. Third, we have argued that the preparation of macroscopic quantum states might be difficult in the first place. For some problems such as measurement precision or 1. Macrorealism and Leggett-Garg-like inequalities collective noise, counter-strategies have been conceived which promise significant improvement. However, they In quantum mechanics, the superposition principle typically work only under certain constraints on the ex- makes it impossible to assign definite properties to a sys- perimental control, and, at least for some examples, these tem. As long as this principle holds only on microscopic constraints become hard to fulfill for macroscopic quan- scales (where it has been thoroughly confirmed), it does tum states. Finally, we have seen that when consider- not contradict our classical perception of the macroscopic ing encoded macroscopic quantum states, which are only world. Leggett and Garg (1985) formalize consequences macroscopically quantum on the logical level, it is possi- from the basic assumption that a macroscopic object has ble to overcome many no-go results. well-defined properties independent of any observer. To do so, they introduce a second premise, namely, that there exist in principle measurements that do not disturb IV. POTENTIALS OF MACROSCOPIC QUANTUM STATES the measured system. With these two assumptions, the authors derive a bound –called Leggett-Garg inequality In this section we review some selected applications (LGI)– on correlations between measurements performed and potentials of macroscopic quantum states. We will at different times. Any experimental violation of an LGI concentrate on applications in quantum information pro- implies the absence of one of the assumptions in nature. cessing and quantum metrology, but also touch upon fun- The so-called clumsiness loophole (i.e., the violation of damental issues such as probing the limits of quantum the noninvasiveness of the measurement through lack of theory. We will base our considerations on the character- experimental control), can be in principle avoided with ization of macroscopic quantum states discussed in Sec. ideal negative measurements and/or carefully designed II. control measurements (Wilde and Mizel, 2012). Lately, While the previous discussion was mainly concerned an LGI could be experimentally violated using Cesium with the issue of how to classify macroscopic quantum atoms (Robens et al., 2015) and superconducting devices states and how to determine their effective size, here we (Knee et al., 2016); see Emary et al. (2014) for a com- are concerned with more practical issues. With emergent prehensive review. quantum technologies the focus has shifted from fun- Recently, alternative formulations of the same intuition damental considerations, e.g., on entanglement or non- were proposed (Devi et al., 2013; Kofler and Brukner, locality, to practical applications of quantum theory in 2013; Saha et al., 2015). In the following, we are par- different tasks, ranging from quantum communication ticularly interested in the “no-signaling in time” (NSIT) and quantum computation to high-precision measure- condition by Kofler and Brukner (2013), which simply ments in quantum metrology. Entanglement is often said states that the probability distribution for a measure- to play a key role in these applications. While this is ment performed at time t = t3 should not depend on certainly true, we do not aim for providing a review on whether or not a measurement at an earlier time t2 was 40

done (see Fig. 10)43. This condition can be even further states that gives the largest LGI-violations. Second, mea- relaxed. The operation at time t2 does not have to be a suring a large quantum Fisher information F(ρ, A) based measurement, but can be any sort of (small) disturbance on Eq. (62) is directly connected to the NSIT protocol (Fröwis et al., 2016; Knee et al., 2016). as in Fig. 10, where O is an application of exp(−iθA) with θ small enough44 to disturb only states with large F (Fröwis et al., 2016). Last, macroscopic quantum co- herence in the spirit of Cavalcanti and Reid (2006) seems to be the kind of coherence necessary to violate an LGI with coarse-grained measurements as discussed by Lam- bert et al. (2016). Note that tests of macrorealism can also be done in bipartite scenarios in which measurements are space-like separated rather than time-like separated for a single sys- tem (Reid, 2016). As a result, a Bell-like inequality can be derived. The locality assumption replaces the second Figure 10 Basic scheme of violating an LGI or the NSIT premise of noninvasiveness for the derivation of an LGI, condition. At time t1, the system is initialized in state |gi which can be seen as a conceptual advantage. However, it and evolves under S1 until time t2. The free evolution is is still necessary to require low-resolution measurements ideally designed to increase the spread of coherence in the (“macroscopic degree of fuzziness”) in order to guarantee measurement basis. Then, for half of the runs, a measurement that a potential violation comes from entanglement (i.e., is done (O, upper line); nothing is done for the other half correlated coherence) between macroscopically distinct (lower line). A subsequent free evolution S2 is followed by a states (Reid, 2016). final measurement at time t3. The protocol can be further relaxed by letting O be any operation. Adapted from Knee et al. (2016). 2. Validity of quantum mechanics at large scales - Collapse models Kofler and Brukner (2007) show that any pure quan- tum state potentially violates an LGI and the NSIT con- One of the main motivations to study macroscopic dition for the right choice of measurements. This is not quantum states, and in particular to try to generate them surprising as projective measurements are highly inva- in the lab, is to show the validity of quantum mechanics sive. On the other hand, limited measurement resolution at large scales - or the need for some alternative the- prevents the violation of LGIs, given that the free time ory. While many, if not most, researchers believe that evolution between the measurements is linear (Kofler and quantum mechanics is indeed valid on all scales, mod- Brukner, 2007). The reason of an LGI-violation is coher- ifications of quantum theory have been suggested that ence of the quantum state in the measurement basis. The lead to the absence of quantum effects such as superposi- effect of the limited resolution is to disregard coherence tions of states on a macroscopic scale. Most prominently, between basis states with a spectral distance less than collapse models (Bassi et al., 2013) including gravitation- the scale of resolution. Therefore, violating an LGI with ally induced collapse (Diósi, 1989; Penrose, 1996) or con- low measurement resolution implies coherence between tinuous spontaneous localization (Ghirardi et al., 1990, far distant parts. 1986; Gisin, 1989) have been suggested. In these models, This directly links macroscopic quantumness to such the time evolution is no longer given by the Schrödinger tests of macrorealism, which becomes particularly evi- equation, but replaced by a master equation including dent by inspecting the measures of Cavalcanti and Reid a diffusion term that prevents massive systems to be in (2006); Fröwis and Dür (2012b); Kwon et al. (2017). spatial superposition states. As discussed in Sec. II.A.10, First, Kwon et al. (2017) directly define macroscopic these models are at the core of the measure proposed by quantumness as being highly susceptible to small influ- Nimmrichter and Hornberger (2013). ences. A low-resolution measurement O in the basis of A Collapse models have been extensively discussed in is precisely covered by the map Φσ, Eq. (51), when σ is Bassi et al. (2013), and a thorough review on the lim- large (i.e, σ & O(N)). Therefore, macroscopic quantum its of quantum superpositions is provided in Arndt and states in the sense of Kwon et al. (2017) are exactly those Hornberger (2014). We will hence only briefly com- ment on these aspects in the following. In Bose et al.

43 The precise differences between the LGI and the NSIT are sub- ject to ongoing investigations (see, e.g., Halliwell (2017); Kumari 44 A small disturbance is often necessary in case one wishes to ex- and Pan (2017)). Here, we are more interested in the global clude the clumsiness loophole with control measurements demon- concept rather than in differences between them. strating (almost) no impact on “semiclassical” states. 41

(1999), Romero-Isart (2011), Nimmrichter et al. (2011) via the Cramér-Rao bound (Cramér, 1945; Radhakr- and Diósi (2015) for example, the requirements to test ishna Rao, 1945), where the precision scales inversely different collapse models using superpositions of massive proportional to the quantum Fisher information. Cer- objects are investigated using quantum optomechanical tain quantum states have a quantum Fisher information systems, levitating nanospheres, matter-wave interferom- that scales as O(N 2), while classical states are limited to etry and classical mechanical oscillator respectively. The a quantum Fisher information of O(N). This establishes observation of quantum interference effects or sponta- a quadratic advantage of quantum strategies over clas- neous thermalization on a large scale allows one to put sical ones. Sometimes, the ratio of the quantum Fisher bounds on parameters in different collapse models (e.g., information of a state |ψi over the optimal classical state the strength/rate of the collapse), thereby allowing one is called the metrological gain, which can be up to N to confirm or rule out these models in these parameter (Pezzè et al., 2016). regimes. Note that in Sekatski et al. (2014a), a pro- The quantum Fisher information is the basis of the posal is discussed to test collapse models by mapping measure by Fröwis and Dür (2012b) (see Sec. II.A.9), the macro component of a photonic micro-macro state which is hence directly linked to the usefulness in metro- to an optomechanical system. This opens a way to con- logical tasks. For pure state, the quantum Fisher in- nect measures for macroscopicity for photonic systems to formation simplifies to four times the variance. Hence, measures for massive objects. much more measures are implicitly connected to pure- state quantum metrology (see Sec. II.D.2). The reason is the intimate connection between the coherent spread B. Quantum metrology of state in the spectrum of an observable A and the state’s sensitivity to small changes induced by exp(−iϑA) We now turn to quantum metrology (Giovannetti (cf. Eq. (61)). et al., 2011; Pezzè et al., 2016; Tóth and Apellaniz, 2014) as a potential application of large-scale quantum systems. In quantum metrology, the goal is to determine an un- C. Quantum computing known parameter, e.g., the strength of a magnetic field, a frequency, a phase or a force, as accurately as possi- Quantum computation is perhaps the holy grail of ble with the given resources. To this aim, a quantum quantum information processing, and provides a long- state of a certain number N of systems is prepared. The term perspective with its applications in solving cer- state then undergoes an evolution that is governed by a tain problems with an (possibly exponential) quantum Hamiltonian which depends on the unknown parameter speedup. Here we concentrate on one particular model ϑ (or possibly several parameters), and is then subse- for quantum computation, the so-called measurement- quently measured. The experiment is repeated ν times, based quantum computation (MBQC) (Briegel et al., and from the gathered measurement data an estimate for 2009), with the one-way model as the most prominent the unknown parameter ϑ is determined. One can dis- representative (Raussendorf and Briegel, 2001). tinguish between phase estimation and frequency estima- tion, where in the latter one has control over the evolu- tion time where in the former this is not the case. There 1. MBQC, entanglement and macroscopicity is a distinction between local metrology scenarios, where the value of the parameter is (almost) known and should In MBQC an entangled state serves as a resource, and be determined with increased accuracy, and Bayesian sce- is manipulated by single qubit measurements only. For narios where the initial knowledge is expressed as a prob- a universal resource such as the 2D cluster state (Briegel ability distribution which is updated. The local scenario and Raussendorf, 2001), by definition any target state deals with many repetitions, while the Bayesian scenario can be generated. An efficient generation (with poly- is a single-shot one. Depending on the concrete problem nomial overhead in auxiliary particles) is possible for all and scenario, the number of systems N, the evolution states that can be prepared using a quantum circuit with time t and number of repetitions of the experiment ν are polynomially many single- and two-qubit gates. It fol- counted as resources. Bounds on the achievable accuracy lows that a universal resource for MBQC must contain all can be found, and in many relevant cases optimal strate- types of entanglement to an arbitrary amount (Van den gies (i.e., initial states, evolution time and measurements) Nest et al., 2007). In order to create a target state, local can be determined (Giovannetti et al., 2011; Pezzè et al., measurements transform and concentrate the entangle- 2016; Tóth and Apellaniz, 2014). ment of the (large) N-qubit cluster state into a (smaller) A central quantity in in this context is the quan- M-qubit target system. tum Fisher information, which was already introduced Therefore, one might be tempted to argue that such in Sec. II.A.9. The quantum Fisher information bounds universal resource states should also be macroscopically the achievable accuracy in the local metrology scenario quantum. However, according to most of the defini- 42 tions for macroscopic quantumness put forward in Sec. II, and noise operators act on different systems and hence 2D cluster states (and other universal resources) are not commute, the effect of noise before the LOCC protocol macroscopically quantum (see example 4). The reason is is the same as after. the lack of two-body correlations necessary for a large This discussion shows that usefulness for certain appli- variance of some local operator. There are, however, cations (such as quantum computation) or entanglement higher-order correlations in the system (more precisely, are different concepts than macroscopic quantumness. five-body correlations) which can be converted into two- Even states that are rendered microscopic according to body correlations via LOCC. many measures can be valuable resources, and allow one As mentioned in example 4, a 2D cluster state of size to perform highly interesting task - such as fault-tolerant N can be transformed into a GHZ state of O(N) particles quantum computation. Note that the choice of free op- using only local measurements (Briegel and Raussendorf, erations is crucial in this respect, as measurements play 2001). Hence, a variance-based measure can increase a central role in MBQC but are usually not considered under LOCC. Entanglement and macroscopicity, though to be free in the context of measures for quantum macro- seemingly related concepts, are therefore intrinsically dif- scopicity. ferent, as noted, e.g., in Fröwis and Dür (2012b); Yadin and Vedral (2015). Hence a resource theory for macro- scopic quantumness should take this into account, and 2. States occurring in quantum computation and metrology cannot allow LOCC as free operations if the variance should be a proper measure (see Sec. II.G for further dis- In spin systems, large variance of local operators im- cussion). On the other side, quantum entanglement and ply strong two-body correlations (i.e., entanglement for quantum macroscopicity are not completely independent, pure states). Shimizu et al. (2013); Ukena and Shimizu see Sec. II.F. (2005) conjectured that this kind of entanglement should Cluster states are also highly robust against noise be present in circuits for quantum algorithms that out- (Hein et al., 2006), which is again in contrast to macro- perform classical computers. The first nontrivial step scopic quantum states. In particular, if one assumes that is to precisely formulate the statement. Every query of each of the qubits of the 2D cluster state interacts with an a quantum algorithm leads to different quantum states independent environment, one can show that the entan- during the computation and not every instance is (expo- glement and other key features of the 2D cluster states nentially) more difficult for a classical device. The au- are maintained up to a certain noise level, independent thors work with a generalized version of the index p (see of the system size. Local depolarizing noise (see foot- Sec. II.A.3) to deal with entire sets of different instances note 38) with error probabilities of up to 10% or more of an algorithm. per particle can be tolerated such that the state remains Shimizu et al. (2013); Ukena and Shimizu (2005) found distillable entangled, i.e., maximally entangled states be- that p = 2 states generically appear in Shor’s factor- tween any pair of qubits can be generated from many ization algorithm (Shor, 1999) and in Grover’s search copies. To generate entangled pairs between neighboring algorithm (Grover, 1997). Since p = 2 states are par- qubits, this can be achieved by measuring the surround- ticularly sensitive to noise generated by local operators ing qubits in the Z-basis, thereby decoupling them from (see Sec. III), the findings emphasize the necessity for the rest of the system and making it obvious that there is a well-designed error correction scheme. Furthermore, no dependence on the total size of the cluster. We remark these results complement other findings about entan- that 3D cluster states have been show to be universal for glement and nonclassicality in quantum enhanced algo- fault-tolerant (encoded) quantum computation using a rithms. For example, Kendon and Munro (2006); Orús 2D surface code, with an error threshold for single qubit and Latorre (2004) found a connection between computa- depolarizing noise of about 0.75% (Briegel et al., 2009; tional speedup and large entropy of entanglement. Note Raussendorf et al., 2007). that there exist states with p = 2 and small entropy of en- This is in contrast to macroscopic superposition states tanglement (e.g., the GHZ state) and states with p = 1 such as the GHZ state, which are much more susceptible and large entropy of entanglement (e.g., eigenstates of to noise (see Sec. III.A.1). It is interesting to note that chaotic systems (Sugita and Shimizu, 2005)). a GHZ state obtained from a noisy 2D cluster state is as As discussed before, a large variance is an impor- decohered as a GHZ state to which the same amount of tant property for pure states to be useful in parameter noise has been directly applied. This can be easily seen estimation. Hence, there is connection between quan- as follows: A GHZ state of N qubits can be generated by tum enhanced computation and sensing. However, this measurements on a subset of qubits of a noisy 2D cluster does not imply that particular instances of large-variance state of size O(N). Even if we assume that all measured states are useful for computation and sensing at the qubits have not been affected by noise, and measurements same time. This connection was further investigated by are perfect, single-qubit noise still acts on the remaining Demkowicz-Dobrzański and Markiewicz (2015). The au- qubits that now form a GHZ state. Since measurements thors rephrased Grover’s algorithm in a time-continuous 43 fashion and expressed it as kind of estimation prob- glement. On the one side, the underlying free, allowed lem. Then, they used results from quantum metrology operations differ in these approaches. On the other side, to bound the performance of the algorithm under some macroscopic quantum states seem to appear in crucial generic decoherence and loss channels. Like in quan- steps of quantum algorithms. tum metrology, Demkowicz-Dobrzański and Markiewicz found a loss of the quadratic improvement in such situa- tions. However, unlike in quantum metrology, where the V. IMPLEMENTATIONS Hamiltonian for parameter estimation is typically given by the problem and cannot be changed, a quantum algo- We now review experiments reporting on the creation rithm is theoretically under full experimental control and, and detection of macroscopic quantum states. In partic- therefore, techniques like quantum error correction can ular, we focus in Sec. V.A on photonic experiments, sep- be applied (Arrad et al., 2014; Dür et al., 2014; Kessler arating optical and microwave setups. We then review in et al., 2014). Sec. V.B experiments with spin systems, distinguishing setups where the spins are addressed individually and col- lectively. Sec. V.C is devoted to massive systems includ- 3. Quantum phase transitions ing atom interferometry and recent experiments with op- tomechanical systems. Superconducting experiments are Recent studies further highlight the importance of mentioned in Sec. V.D. Finally, in Sec. V.E, we provide macroscopic quantum states in other quantum algo- comparisons of the size of states based on experimental rithms and paradigms (Yuge, 2017). But macroscopic data. We summarize in Sec. V.F. quantum states are also expected to play a role in ground states of strongly coupled systems, systems with topo- logical order or topologically protected phases of matter. A. Photonic experiments There already exist some works that relate macroscopic quantum states and quantum phase transitions. One ex- Photonic setups can naturally be divided into two ample is Hauke et al. (2016), where it is shown that the groups: The optical setups which mostly rely on sources quantum Fisher information can be obtained by means based on parametric conversions and the microwave se- of the dynamic susceptibility, and can be used to detect tups where strong light-matter interactions are used to entanglement during phase transitions. In Shitara and shape SCS. We provide a quick presentation of experi- Ueda (2016) it was shown how to obtain the quantum ments nicely illustrating the research activities of several Fisher information from linear response functions. groups and invite the reader to look at for more exhaus- tive review papers on optical45 and microwave46 setups, respectively. D. Summary

1. Optical photons Macroscopic quantum states are no longer only an interesting virtual possibility that illustrate puzzling The central tool of optical experiments is spontaneous features of quantum mechanics, as in the times of parametric down conversion, that is, a bulk crystal or a Schrödinger. Nowadays such states are thought of as waveguide for example, with a second order non-linearity valuable resources. The usefulness of states for certain that is used to convert photons of a pump laser into tasks, their entanglement features and their classification photon pairs. These pairs have combined and of being macroscopically quantum with respect to certain momenta equal to the ones of the laser photons and measures are however different concepts. Though there are correlated in polarization. We can distinguish Type are certain relations, the goal and merit of these concept I and Type II down converters depending on the pair vary. polarization. We have highlighted in this section possible applica- tions of macroscopic quantum states for testing the lim- Type I down converters produce pairs with identical its and validity of quantum mechanics at large scale - polarization and result in single-mode squeezed vacuum a question that is primary of fundamental interest. We (see Eq. (11)) containing only even photon numbers. have also discussed more practical applications, in par- ticular in the context of quantum metrology, where the usefulness for metrology coincides with some concepts of quantum macroscopicity, in particular measures based 45 Chekhova et al. (2015); De Martini and Sciarrino (2012); Jeong on the variance or quantum Fisher information. The link et al. (2015); and Pan et al. (2012) is less obvious for applications for (measurement-based) 46 Devoret and Schoelkopf (2013); Haroche (2013); Makhlin et al. quantum computation and to different aspects of entan- (2001); Raimond et al. (2001); and Schoelkopf and Girvin (2008) 44

Figure 11 Schematic representation of the experiment re- ported in Eberle et al. (2013) in which two single-mode Figure 12 Schematic representation of the experiment pre- squeezed vacuum states are combined on a beam-splitter be- sented in Yao et al. (2012) using four type II down conver- fore characterizing their correlations in phase space. From sion processes to create 8 photons GHZ states. Laser pulses Eberle et al. (2013). with a short duration and a high repetition rate successively pass through four non-linear (BBO) crystals to produce four photon pairs, which are further combined using a half- and The experiment reported in Eberle et al. (2013) has quarter-wave plate (HWP, QWP) and a polarization beam- splitter (PBS). The photons in mode 1 and 4 are then com- combined two of these squeezed vacuum modes on a bined on PBS1, photons 5 and 8 on PBS2, and finally photons beamsplitter before performing phase P and amplitude 4’ and 8’ on PBS3. The photons are detected by 16 single- X quadrature measurements at each output (A and B) photon detectors and a complete set of 256 eight-fold coin- of the beamsplitter, see Fig. 11. The results exhibit a cidence events are post-selected and registered to perform a 2 ≈ 10dB reduction of noise variances [∆(XA + XB)] and tomography of the post-selected state. From Yao et al. (2012). 2 [∆(PA − PB)] with respect to the sum and difference of the corresponding quadratures for a vacuum state. These correlations in amplitude and anti-correlations in phase has been used to certify entanglement and can II converters pumped by short pulses together with nar- be used to quantify the size of the produced state, see row filters of the output photons to erase their frequency Sec. V.E. Note that record squeezing, down to ≈ 15dB, correlations. They combined the outputs of four of these has recently been reported using type I down converters down converters using linear optical elements and post- in Vahlbruch et al. (2016). select events using photon counting devices, see Fig. 12. Using intense pumps, 8 photons GHZ states have been Iskhakov et al. (2011) used two co-linear type I down postselected with 70% fidelity (Yao et al., 2012) and re- converters but in a Mach-Zehnder interferometer such cently, similar techniques led to creation of up to 10 pho- that they can be excited coherently with orthogonally po- tons GHZ states with 57% fidelity (Wang et al., 2016b). larized pumps (see also (Eisenberg et al., 2004) for a sim- Aside from these experiments aiming to create macro- ilar work). This led to polarization entanglement, that is, P (i)† (i)† (i)† (i)† scopic photonic states directly from spontaneous para- an Hamiltonian of the form i(aH bH +aV bV +h.c.) metric down conversion, various techniques have been (i) (i) where the bosonic operators aH , bH correspond to two proposed to amplify the photon number in few-photon spatial modes with horizontal polarization and simi- quantum states while preserving their quantum nature. (i) (i) larly for aV and bV . The sum means that the emis- The first experiment along this line has been reported in sion is multimode, that is, photons are created in dif- De Martini et al. (2008). Polarization entangled photon ferent angular modes. By collecting tens of thousands pairs were first created from a type II converter and a of modes, hundreds of thousands of entangled photons photon from each pair was subsequently injected into a have been successfully detected (Iskhakov et al., 2012, second type II converter, the latter playing the role of a 2011). Note that, as a result of the multi-mode emis- phase covariant cloner (Sekatski et al., 2009). This ide- sion, the state of these photons corresponds essentially ally creates a photonic state of the form to independent two-qubit maximally entangled states (i)† (i)† (i)† (i)† (aH bH + aV bV )|0i where |0i is the vacuum for all modes. For this reason, all measures discussed in 1  ⊥ φ φ⊥  √ |1φ iA|Φ iB + |1φiA|Φ iB (71) Sec. II.A consider this state not as macroscopically quan- 2 tum even under otherwise ideal circumstances. Intensive effort has been dedicated to the generation of φ φ⊥ photon pairs in single modes. Yao et al. (2012) used type after amplification. |Φ iB and |Φ iB are two orthogo- 45 PHYSICAL REVIEW LETTERS week ending PRL 100, 253601 (2008) 27 JUNE 2008

* DA P 0.74 is accepted (higher P could be accepted ALICE respondingg macroqubit.  B  ? B , i.e., ag mac- roscopicby quantum taking superposition the j secondi [‡10].j modeThei quantum† into states account and/or by of Eqs. (2) and (3) deserve some comments. The multi- k SM changing the branching into |Ai and |Di). In this case, A PBS DA   k /2 particle states  B,  ? B are orthonormal and exhibit P PS /4 an effective size of ≈ 1000 can be found for P = 2/3 observables bearingj i macroscopicallyj i distinct average val- g C1 kB and g = 4.4. All measures based on the variance of DM SM ues. Precisely, for the polarization mode ~  the average 2  numberquadrature of photons is m operatorssinh g for find ? aB large, and 3 effectivem size in the k’P DM kB  ˆ j i ‡ /2 PBS C2 1 for orderB. For of the the-mode photon~  these number. values are Note, inter- however, that † j i ? changedan among experimental the two macrostates. test Onshowing the other this hand, large as macroscopic * shown by [10], by changing the representation basis from BOB * LB PB quantumness was not done so far. In particular, the Coincidence ~ ;~ to ~ H;~V , the same macrostates,  B or OF f coherence?g f ofg the state (71) is difficultj i to prove. In this L Box  are found to be quantum superpositions of two /4 /2 PBS P B ? B B jorthogonali context, states RaeisiH , etV al.which(2011) differ showed by a single that coarse-grained j iB j iB FIG. 2 (color online). Optical configuration of the QI-OPA quantum.measurements This unexpected and cannot quite peculiar reveal combination, the quantum nature of Figure 13 Schematic representation of the setup used in apparatus. The excitation source was a Ti:sapphire Coherent i.e., a largethese difference states, of a a property measured observable that is shared when the by many macro- De Martini et al. (2008) aiming to produce micro-macro en- MIRA mode-locked laser amplified by a Ti:sapphire regenera- states arescopic expressed quantum in one states, basis and see a small discussion Hilbert- in Sec. III.B.1. tanglementtive REGA via device the operating amplification with repetition of micro-micro rate 250 kHz. entangle- The Schmidt distance of the same states when expressed in ment.output A first beam, type frequency-doubled II process converts by second-harmonic photons fromgenera- pumpanotherDe basis Martini turned out and to be Sciarrino a useful and (2015) lucky property argued that the quan- laser intotion, provided photon the pairs OPA with excitation polarization field beam at entanglement. the UV wave- since A ittumness rendered the of coherence this state patterns is of experimentally our system very shown in the length (wl)  397:5 nm with power: 750–800 mW. A type II photon from eachP ˆ pair is used to seed a second type II con-robust towardlow-g couplingregime with with environment, a mean photon e.g., losses. number The of up to twelve. BBO crystal (crystal 1: C1) generates pair of photons with decoherence of our system was investigated experimen- verterwavelength pumped by2 the same795 nm laser.. C1 generates The resulting an average pho-entangle- ˆ p ˆ tally and theoretically in the laboratory: cf.: [12,15,17]. mentton is characterized number per mode by equal recording to about the 0.35, coincidences while the overall between States with similar properties can be obtained with As shown in Fig. 2, the single-particle field on mode k photondetection counting efficiency devices of the that trigger are modepreceded was estimated by a set to of be wave- simpler amplification techniques. SekatskiA et al. (2012) plates (5%λ/2with, λ/ detection4) and a rates polarization of about 5 kHz. beamsplitter The NL BBO (PBS). crys- DM:was analyzed in polarization through a Babinet-Soleil tal’ 2: C2, realizing the optical parametric amplification, is cut for phase-shifterfor example (PS), i.e., proposed a variable birefringent to start with optical path-entangled re- state, dichroic mirror, SM: single-mode fiber, PS: phase shifter.   collinear type II phase matching. Both crystals C1 and C2 are tarder, twothat wave is, plates a micro-micro; and polarizing state of beam the splitter form √1 | i | i − From De Martini et al. (2008). f4 2g ( 0 A 1 B 1.5 mm thick. The fields are coupled to single-mode fibers. The (PBS). It was finally detected by two single-photon detec- 2 overall detection efficiency on mode k has been estimated to be |1iA|0iB) where |0iA and |1iA are the vacuum and single B tors DA and DA (ALICE box). The multiphoton QI-OPA 2%. photon Fock state for the spatial mode A and similarly  amplified field associated with the mode kB was sent, nal states that are defined by throughfor a single-mode B. The idea optical then fiber consists (SM), to in a measure- amplifying the photon   single phase  0; 2 in the basis H ; V . The over- ment apparatusnumber consisting in one of mode a set of through wave plates a displacement4 ; 2 ,a in phase ∞ 2p † fj i j ig f g φall outputX state amplified(1 + 2 byi)!(2 thej!) OPA apparatus is ex-⊥ (PBS) andspace, twoD photomultipliers(α), where the (PM) amplitudePB and PB α(BOBis considered to be |Φ iB = γij |(2i + 1)φ; 2jφ iB pressed, in any polarization equatorial basis ~ ;~ , box). The output signals of the PM’s were analyzed by an i!j! f ?g real without loss of generality. This leads to by the micro-macroi,j=0 entangled state [16]: ‘‘orthogonality filter’’ (OF) that will be described shortly in ∞ p this Letter. 1 ⊥ X 1=2 (1 + 2i)!(2j!)  φ  2ÿ  1?  ? 1 ; (1)⊥ We now investigate the√ bipartite(|0iAD entanglement(α)|1iB − between |1iA|αiB) (72) |Φ iBj =iA;B ˆ γij j iBj iA ÿ j |2jφiB;j (2iA+† 1)φ iB i!j! the modes k and k . We define2 the 1 -spin Pauli operators where thei,j=0 mutually orthogonal multiparticle ‘‘macro- A B 2 ^ for a single-photon polarization state, where the label states’’ are f ig where |αiB = D(α)|0iB is a coherent state for mode B.  i+j i 1; 2; 3 refers to the polarization bases: i i −2 tanh(g) ˆ When† the initial micro-micro state is seenˆ in the rotated with γij = (−1) cosh(g) 2 , g being the gain1 ~ H;~V , i 2 ~ R;~L , i 3 ~ ;~ .  1 1 2i ! 2j ! ()f g ˆ ()f−1/2 g ˆ ()f ‡ ÿg −1/2  ‡ † † 2i 1 ; 2j  ; basis {|1=2+i = 2 (|0i + |1i), |−i = 2 (|0i − |1i)}, B ij ⊥ ? B Here ~ R 2ÿ ~ H i~V , ~ L ~ R? are the right- of thej amplification.i ˆ i;j 0 p|nφi!ijA! (|nφj iA‡)† corresponds † i to a ˆ ÿ † ˆ ˆ and left-handedthe entangled circular state polarizations after the amplificationand ~ involves two X √1 iφ  n photon Fock state with polarization H +(2)e V 1=2  ˆ 2 2ÿ ~componentsH ~ V . It isD found(α)|±i^ i B whosei i mean i? i? photon numbers  1 iφ   † ˆ j ih j ÿ j ih j √ H − e V , H and V standing for horizontal andwhere are i ; separated i? are the by two2α. orthogonalThis proposal qubits corre- has triggered two 2 fj i j ig 1 2i ! 2j ! sponding to the ~ i basis, e.g., 1 ; ? H ; V ,  1 φ experiments (Bruno et al.,1 2013; Lvovsky et al., 2013),  ? ‡ † † 2j ; 2i 1 ? ; fj i j ig ˆ fj i j ig vertical polarizationsB ij respectively.i!j! Interestingly,B |Φ ietc.B By the QI-OPA unitary process the single-photon ^ i j i ˆ 2i;j 0 p j † ‡ † i both using spontaneous down conversion based sources ˆ ^ contains 3 sinh Xg + 1 photons on average with polariza-operators evolve into the ‘‘macrospin’’ operators: i 1 iφ  2 (3) ^ ^ to create i photoni i pairs, i the detectionˆ of a photon from tion √ H + e V and sinh g photons with the orthog-U^ iUy    ?  ? . Since the opera- 2 j eachˆ j pairih j servingÿ j ih to heraldj the creation of its twin 2 ÿ i ÿ ⊥ ^ with ij Cÿ 2 2 , C φcoshg, ÿ tanhg, being2 g tors i are built from the unitary evolution of eigenstates onal polarization ÿ whereas† |Φ iB contains sinh g pho-of ^f ,photon. theyg satisfy theThe same heralded commutation photons rules of were the then sent into the NL gain [13]. There,1 p; q?iφB stands for a Fock2 i tons with polarization √ j H + e i V and 3 sinh g + 1 1 ^ ^ ^ state with p photons with2 polarization ~  and q photons single-particlea balanced-spin: beamsplitter ;  " 2i to, where create" path-entanglementis 2 ‰ i jŠˆ ijk k ijk photonswith with~  over the the orthogonal mode kB. Most polarization. important, any injected The statethe Levi-Civitabefore undergoing tensor density. a displacement The generic operation. state The latter ? H V (71) cansingle-particle thus been qubit considered  B as? aB micro-macrois transformed by entan-  wasB implemented B is a macroqubit with in thean Hilbert unbalanced space beamsplitter the information-preserving j i QI-OPA‡ j operationi † into a cor- B spannedj i ‡ byj iH † ; V , as said. To test whether the gled states in which a polarization mode contains about and coherentfj iB j states.iBg To facilitate the detection needed 3 times more photons than the orthogonal polarization to reveal entanglement, the displacement operation mode (De Martini et al., 2008). In the experiment,253601-2 was undone and the coherence and photon number g = 4.4 was used resulting in thousands of created pho- probability distribution in each arm were obtained tons. by photon counting counting techniques. In Bruno The analysis of the ideal state regarding its macro- et al. (2013), entanglement was recorded as the photon scopic quantumness reveals differences between the number is amplified and the authors succeeded to reveal proposed measures (cf. example 2 in Sec. II.B). For entanglement for up to α2 = 500. The size of the target example, Sekatski et al. (2014c) assigns only a large state (72) has been discussed in Sekatski et al. (2014c), effective size when measuring the photon number in one see Sec. II.A.11 for the corresponding measure. For a polarization mode if a relatively low success probability guessing probability Pg = 2/3 for example, it has been 46

Figure 15 Schematic representation of the setup used in Deleglise et al. (2008) to produce various quantum states of light trapped in a high finesse cavity (C), including superpo- sition of coherent states with opposite phases. A stream of atoms is prepared in box B and cross the R1-R2 cavities play- ing the role of a Ramsey interferometer in which the cavity C is inserted. The source S generates a coherent microwave pulse that can be used to inject into C coherent states with controlled amplitude and phase. Another pulsed source, S’, feeds the interferometer cavities R1 and R2. Information is extracted from the field by state-selective atomic counting in Figure 14 Schematic representation of the setup used in Our- D. Adapted from Deleglise et al. (2008). joumtsev et al. (2006) aiming to produce superposition of co- herent states with opposite phases. A squeezed vacuum beam is first produced in a frequency-degenerate optical parametric amplifier (DOPA) by down-conversion of frequency-doubled tained SCS with larger sizes have been implemented to (SHG) laser pulses. A beamsplitter reflects less than 10% of create superpositions of squeezed coherent states (Etesse the squeezed beam toward a photon detector (APD) through et al., 2015). a filtering system, whereas the transmitted beam is analyzed by a homodyne detection. A tomography of the photon sub- tracted squeezed vacuum state is performed and the obtained 2. Microwave photons state is then be compared with a superposition of coherent states with opposite phases. APD : avalanche photo-diode. Fock states and superpositions of coherent states with Adapted from Ourjoumtsev et al. (2006). opposite phases have been created by letting Rydberg atoms interact one by one with the electromagnetic field of a high finesse cavity (Deleglise et al., 2008; Sayrin shown that its effective size for α2 = 500 is the same et al., 2011). The principle relies on a dispersive light- than the one of √1 (|0i |ni − |1i |0i ) with n = 38. 2 A B A B atom interaction that is used to imprint the information about the photon number in the cavity into the phase of Let us finally mention conditional techniques that can a superposition between two internal atomic states, the also be used for example to create SCS (Eq. (10)) as latter being measured through a Ramsey interferometer, proposed in Dakna et al. (1997). Ourjoumtsev et al. see Fig. 15. Fock states have been prepared by first (2006) for example reported on the creation of such a launching a coherent field in the cavity and by then state by subtracting one photon from a squeezed vacuum letting it interact with atoms, achieving a quantum state. The latter was created by means of a frequency- non-demolition measurement of the photon number that degenerate optical parametric amplifier. Photons are progressively projects the field onto a Fock state |ni. then subtracted from the output in a probabilistic way, Reconstruction of Fock state with up to n = 4 photons using a partially reflecting beamsplitter and a photon have been reported in Deleglise et al. (2008) and up to detector, see Fig. 14. A successful photon subtraction n = 7 photons in Zhou et al. (2012) using an additional projects the transmitted part into a state which is close quantum feedback procedure. to a SCS. A tomography revealed SCS-like states charac- terized by |α|2 = 0.79 and a fidelity of 70% (Ourjoumt- To generate SCS, a coherent field is first injected into sev et al., 2006). Larger sizes, up to |α|2 = 3.2 have the cavity before interacting with an atom prepared in a been reported in Gerrits et al. (2010); Neergaard-Nielsen superposition of two internal states. This results in an et al. (2006); Sychev et al. (2017); Yukawa et al. (2013) atom-light entangled state in which the internal atomic using similar protocols. Note that the usefulness of the states are correlated with a coherent state with different photon subtraction has been discussed in Oudot et al. phases. The projection of the atom into the appropriate (2015) where the size of squeezed vacuum states and SCS state ideally leaves the field in the desired superposition. are compared with different measures. More recently, Deleglise et al. (2008) reported on SCS states with α2 = advanced conditional techniques have been used to cre- 3.5 and a fidelity of 72%. iϕ ate entanglement of the form |+iA|αiB + e |−iA| − αiB Similar techniques were used more recently with super- (Jeong et al., 2014b; Morin et al., 2014). Note also that conducting devices. Vlastakis et al. (2013); Wang et al. iterative conditional techniques that could be used to ob- (2016a) reported on a set of multi-photon operations 47 using a superconducting charge qubit called transmon, coupled to waveguide cavity resonators. The authors succeeded to obtain ideal strong-dispersive coupling, where the strengths of the off-resonant qubit-cavity interactions were several orders of magnitude larger than both the cavity and transmon decay rates. This allowed them to create and detect SCS with unprecedented sizes, that is, α2 ≈ 7.8 and a visibility of ≈ 0.57 (Vlastakis et al., 2013). A similar setup has also been used to implement two-mode SCS of the form |α, βi + |−α, −βi with |α|2 = 9.0, |β|2 = 7.0 with a visibility of 0.58 (Wang Figure 16 Spherical projections of the Wigner function on the Bloch sphere for 100 spins which help understanding how et al., 2016a). the techniques used in Riedel et al. (2010); Schmied et al. (2016) lead to spin squeezing. (A) an initial coherent spin state is prepared along the x direction. (B) Controlled elas- tic collisions in state dependent potentials lead to an effective 2 B. Spin experiments interaction of the from Sz . The corresponding unitary is noth- ing else than a rotation around z whose angle depends on the Regarding experiments with spin systems, we can dis- projection on the z axis. This results in a spin squeezed state tinguish between setups in which the spins can be ad- whose squeezing and anti-squeezing directions can be mea- dressed individually or collectively. As before, we quickly sured by projective measurements are along the vertical (+z) spin axis preceded by the appropriate rotation. Adapted from present techniques nicely illustrating experiments along Schmied et al. (2016). these two lines. Blatt and Wineland (2008); Leibfried et al. (2003); Ritsch et al. (2013) are more exhaustive reviews related to trapped ions and cold atoms respec- optical pumping techniques, coherently coupled through tively. controlled elastic collisions and readout using absorption imaging, see Fig. 16. This allowed one to create spin squeezed states with 1250 atoms (Riedel et al., 2010) 1. Spins with individual addressing with a squeezing parameter ξ2 ≈ −2.5dB, see Eq. (8) for One of the most advanced systems where (pseudo- the definition. Note that the detection of spin squeezing )spins can be addressed individually are trapped-ion sys- is connected to quantum correlations between the spins tems. The spin states are often encoded in a ground state (Kitagawa and Ueda, 1993). Recently, the techniques and a metastable electronic state of each ion. Laser cou- reported in Riedel et al. (2010) have been used to prove pling between these internal states and the vibrational that the internal correlations between 480 atoms in mode produces a collective spin flip that can be used as a spin-squeezed states are strong enough to violate a an entangling gate (Sørensen and Mølmer, 1999). From Bell inequality (Schmied et al., 2016). Note also that high fidelity quantum gates (Benhelm et al., 2008), GHZ spin-squeezed states have been created with larger states (see Eq. (4)) with up to 14 ions (Monz et al., 2011) squeezing parameters. Gross et al. (2010) reported on 2 have been created. The diagonal elements of the corre- spin squeezing with 2300 atoms and ξ ≈ −8.2dB in sponding density matrice have been measured directly a Bose-Einstein condensate, through Feshbach control by fluorescence measurements while the off-diagonal el- of interactions in an optical trap. Note finally that ements have been accessed via the amplitude of parity similar states have also been obtained with trapped oscillations. The measurements allowed to infer the fi- ions (Bohnet et al., 2016), with room-temperature delity of GHZ states for different ion numbers. Fidelities (Vasilakis et al., 2015) and cold (Hosten et al., 2016) larger than 95% and 80% have been observed for up to atoms trapped in cavities. The latter succeeded to 5 4 and 8 ions GHZ states respectively while 14-ion GHZ report spin-squeezing with 5 × 10 atoms with squeezing 2 states have been measured with a fidelity above 50%, parameter ξ ≈ −20.1dB. which is in principle sufficient to violate a Bell inequality (Lanyon et al., 2014).

C. Experiments with massive systems 2. Spins with collective addressing Experiments aiming to bring a massive system in a An example of spin systems where the spins cannot quantum superposition of well distinct positions include be addressed individually is given by Bose-Einstein matter interferometry and quantum optomechanics. condensates where the internal states of the atoms While intensive efforts are devoted to the former for constituting these condensates can be initialized with more than 30 years, the latter is nowadays attracting 48

et al. (2016), such a superposition with α ≈ 5.9 has been reported, which effectively corresponds to a spatial separation between the superposed locations dozen of times larger than the extent of local position fluctuations.

To date, the widest delocalization has been obtained by launching a Bose-Einstein condensate made with about 105 Rubidium atoms in a 10m high atomic fountain, see Fig. 17. Once launched, a sequence of pulses is applied to control the atom momenta so that the wave packet of each atom is split and recombined coherently to form the analogue of a Mach-Zehnder interferometer. In the experiment presented in Kovachy et al. (2015) the wave packets get separated during a Figure 17 Schematic representation of the setup used in Ko- drift time ≈ 1s after which they reach their maximum vachy et al. (2015) to coherently split the wave packet of single separation of up to ≈ 54cm. The wave packet of each atoms up to ≈ 54cm. An ultra-cold atom cloud is launched atom is then recombined to spatially overlap after vertically in magnetic shield using an optical lattice. At time another drift interval ≈ 1s. The contrast of the inter- t = 0, a first sequence of laser pulses split the cloud into a su- ference is determined by measuring the variation of the perposition of states with different momenta. A time T later, the wave packet is spatially separated, and a laser sequence normalized number of atoms in one of the two outputs of reverses the momenta of each superposed component. At time the interferometer. Interestingly, the contrast of ≈ 28% 2T, the clouds spatially overlap, and laser pulses make them reported in Kovachy et al. (2015) is incompatible with interfere. Adapted from Kovachy et al. (2015). an explicit collapse model based on quantum gravity (Minář et al., 2016) in the parameter regime that was initially proposed by the fathers of this collapse model a lot of attention and impressive results have been (Ellis et al., 1989). Note however, that Refs. (Kovachy obtained in the last . Matter interferometry et al., 2016; Stamper-Kurn et al., 2016) clarified that has been reviewed in Arndt and Hornberger (2014); the experiment reported in Kovachy et al. (2015) did Cronin et al. (2009) and the most recent progress in not have a stable phase reference for the interferometer, optomechanics can be found in Aspelmeyer et al. (2014); which is required to constrain models that would Meystre (2013); Yin et al. (2013). introduce overall phase noise. In subsequent work (Asenbaum et al., 2017), the same group introduced a second, spatially displaced interferometer as a stable phase reference. This experiment demonstrated phase 1. Matter interferometry stability of interferometers with 16cm arm separation and it remains to be clarified if the observed interference Interferometry with widely delocalized and more is compatible with the collapse model presented in Ref. and more massive objects is an active research domain (Ellis et al., 1989) . since mid 80’s (Gould et al., 1986; Keith et al., 1988). Impressive results have been obtained along this line Despite using a Bose-Einstein condensate, the inter- by the trapped ion community (Wineland, 2013). The ference reported in Kovachy et al. (2015) depends only starting point of these experiments is the use of laser on the wavelength of a single atom. The experiment cooling techniques to bring an ion down to its motional is essentially single atom interferometry with ≈ 105 ground state |0i . Using a displacement whose direc- interferences performed at each experimental run. The tion depends on the spin state of the ion’s outermost relevant mass is thus limited to that of a single atom, electron, an ion prepared in a superposition of spin that is 87 amu for 87Rb. However, larger masses states (|↑i + |↓i) ends up in a superposition of two are required to test a broad class of collapse models. positions (Monroe et al., 1996). More precisely, the This provides motivation to perform interferences with spin-depend force promotes an initial state (|↑i + |↓i) |0i macromolecules and clusters. One of the first results into |↑i |αi + |↓i |−αi where the coherent states |±αi along this line was obtained with a Talbot-Lau type represent the amplitude and phase of the ion in interferometry using fullerenes (Arndt et al., 1999), a its local harmonic trapping potential. The two positions carbon with 720 amu. The basic principle of can then be recombined to form the analog of an such an experiment is shown in Fig. 18. First molecules interferometer which allows one to access the coherence from a thermal source are evaporated into a vacuum of the entangled state |↑i |αi + |↓i |−αi through the chamber and a velocity selection is done using narrow coherence of the spin state superposition. In Kienzler slits. The selected molecules are then sent into an 49

Figure 18 Schematic representation of the setup used in Figure 19 Schematic representation of the setup envisioned Eibenberger et al. (2013) to make interferometry with massive for matter interferometry with a nanosphere. A cooled sphere molecules combining 810 atoms with a molecular weight ex- 4 is first optically trapped before being released to let its wave- ceeding 10 amu. Three narrow slits D1-D3 select a particular packet expand. The sphere then enters a second cavity where particle velocity following a specific parabola in the gravita- a pulsed interaction is performed using a quadratic optome- tional field. G1-G2 are gratings. G2 is a standing wave used chanical coupling. A homodyne measurement of the output to change the phase of the interferometer through an opti- field phase performs a quadratic measurement of the sphere cal dipole force. The transmitted molecules are detected via position and prepares it in a quantum superposition of two po- a ionization technique (QMS : quadrupole mass spectrome- sitions whose spread depends on the measurement outcome. ter) after G3 which can be shifted to sample the interference Then, the sphere is again released before its center-of-mass fringes. From Eibenberger et al. (2013). position is measured and interference fringes are observed. From Arndt and Hornberger (2014).

Talbot-Lau interferometer which is made with three been realized to trap and cool such a nanosphere47, we gratings. The first grating preselects a molecular trans- are not aware of experiments reporting on quantum verse coherence. Diffraction at the second grating then interference with such a system. produces a molecular density pattern at the location of the third grating through the Talbot effect. If the molec- ular pattern and the third grating mask are aligned, the transmission is high. When the third grating is shifted 2. Quantum optomechanics by half a grating period, the total transmitted signal is minimal. The signature of molecular interference is thus While many aspects of quantum cavity optomechan- obtained by counting the molecule number as a function ics started to be explored theoretically in the early of the position of the third grating. An advanced version 90’s (Fabre et al., 1994; Mancini and Tombesi, 1994), of the Talbot-Lau interferometer currently holds the proposals (Bose et al., 1997, 1999) have been done in mass record in matter-wave interference, with molecules the late 90’s to create a superposition of mechanical combining several hundreds of atoms with a molecular states with a distance of the order of the mechanical weight of thousands (Gerlich et al., 2011) and even tens zero-point fluctuation where the effects of unconven- of thousands amu (Eibenberger et al., 2013). tional decoherence might be observed (Kleckner et al., 2008; Marshall et al., 2003). The basic idea is to use a Michelson interferometer with a high-finesse cavity in To conclude this section, let us mention that matter- each arm, see Fig. 20. One of the two cavities is made interferometry experiments are nowadays envisioned with a movable mirror, that is, a high-quality oscillator with nanosphere exceeding 107amu. Romero-Isart et al. with a mechanical frequency larger than the cavity (2011) for example propose to trap a dielectric sphere in decay rate such that it can initially be prepared in its the standing wave of an optical cavity, see also Barker motional ground state by sideband cooling. A single and Shneider (2010); Chang et al. (2010) for related photon is then launched in the Michelson interferometer proposals. The mechanical motion of the sphere’s center and its energy is stored coherently in both cavities. of mass is predicted to be a high-quality mechanical The radiation pressure force shifts the mirror position oscillator due to the absence of thermal contact and and a maximal displacement is achieved after half a dissipation arising from clamping. This is expected to mechanical period. After a full mechanical period, the facilitate laser cooling. The cooled levitating object mirror comes back at its original position and in the can then be released by switching off the trap to let the wave-function expand. A quadratic measurement of the mechanical position finally creates a scenario similar to matter wave interferometry experiments, see 47 Arita et al. (2015); Gieseler et al. (2012); Kiesel et al. (2013); Fig. 19. While significant experimental progress has Millen et al. (2015); Ranjit et al. (2015) 50

D. Superconducting quantum interference devices

The question of whether or not macroscopic quantum- ness can be realized in superconducting devices (Fried- man et al., 2000; Hime et al., 2006; van der Wal et al., 2000) was intensively discussed.49 While superconductiv- ity itself was argued to be a classic example of a micro- scopic quantum effect (Leggett, 1980), there have been attempts to create macroscopic superpositions of clock- wise and anti-clockwise circulating currents in supercon- ducting quantum interference devices (SQUIDs). In its Figure 20 Schematic representation of the interferometer en- simplest version, a SQUID is a superconducting ring with visioned for studying the creation and decoherence of a mirror conductance L interrupted by a single Josephson junc- in a spatial superposition. A cavity is placed on each arm of tion with capacitance C and critical current Ic, which al- a Michelson interferometer. One of the two cavities is made lows electrons to pass through by tunneling (see Leggett with a high-quality mechanical resonator whose motion is af- (1987) and references therein for a detailed description of fected by a single photon through its radiation pressure. A different variant of SQUIDs). In Friedman et al. (2000), photon entering in the interferometer leads, after half a me- the single junction is replaced by two parallel junctions chanical period, to an entangled state in which the photon is stored in cavity B and the mechanical oscillator is in its (see Fig. 21 (c)). Even though many electrons (up to 10 motional ground state and the photon is stored in A and the 10 ) are involved, the only relevant degree of freedom is mechanical motion is excited. After a full mechanical period, the total flux in the ring. This gives rise to a simple phe- the photon mode and the mirror position disentangled, re- nomenological model that is mathematically equivalent sulting in a maximum photon interference in the absence of to a particle with effective mass C in a one-dimensional oscillator decoherence. Recording the photon interference for double-well structure potential. The energy eigenfunc- multiple mechanical periods allows one to infer the mechani- tions that are localized in one well corresponds to cur- cal decoherence. From Marshall et al. (2003). rents circulating either clockwise or anti-clockwise. By controlling the height of the potential barrier as well as the difference between the left and right local minima absence of decoherence, full interference is expected. (see Fig. 21 (a)), experimenters prepared the device in a The mirror decoherence, however, alters the interference localized ground state (state |ii in Fig. 21 (a)) and drove of the photon. In other words, by observing the photon it to localized excited states (either |0i or |1i). By tuning interference, we can infer the mirror decoherence rate. the local minimum (), the authors observed an avoided For an optomechanical coupling rate larger than the crossing in the energy spectrum which is an indication mechanical frequency, the maximum displacement is of the coherence between clockwise and anti-clockwise expected to be larger than its zero-point motion at half circulating currents (see Fig. 21 (b) for the theoretical the mechanical period. If in addition, the device operates prediction). Note that other experiments, for example, in the strong coupling regime so that the photon can be Hime et al. (2006); van der Wal et al. (2000) differ in the stored long enough, this device could be used as a test details, but also work with a double-well potential and bench for unconventional decoherence models. While prove the coherence via an avoided crossing. massive oscillators with eigenfrequencies in the kilohertz regime were initially envisioned, first experiments now succeeded in entering the field of optomechanics in the E. Comparing the size of states describing photonic, spins quantum regime using lighter and more rigid mega and massive systems or gigahertz oscillators. This includes ground state cooling of the mechanical motion48, electromechanical In this section, we summarize estimates of the effec- entanglement (Palomaki et al., 2013) or squeezing of a tive size of states detected experimentally. Ideally, this micromechanical state (Wollman et al., 2015). As far as is done as much as possible directly from the experimen- we can tell however, no experimental realization so far tal data with a minimum of additional modeling. In this entered a regime relevant for ruling out unconventional sense, the first section V.E.1 misses this ideal as all dis- decoherence models. cussions so far mainly focus on the theoretical model and less on the experimental results. Even more importantly,

49 In particular, in Björk and Mana (2004); Dür et al. (2002); Ko- 48 Chan et al. (2011); Meenehan et al. (2015); O’Connell et al. rsbakken et al. (2007, 2009, 2010); Leggett (1980, 2002); Mar- (2010); Teufel et al. (2011) quardt et al. (2008); Nimmrichter et al. (2011) 51

conceptual disagreement50, we note that measures are partially defined for different scales51 This further com- plicates a comparison between the measures. Leggett (2002) claims about the SQUID experiments that “any reasonable” measure for macroscopic quantum states would assign an effective size that is in the order of the number of involved electrons. For the experiments Friedman et al. (2000); van der Wal et al. (2000), this is 1010 (in units of Bohr magneton in the case of Leggett’s extensive difference). Knee et al. (2016) found an ex- tensive difference of roughly 1.3 × 105 in units of Bohr magneton for their experiment. This number is strongly contrasted by Marquardt et al. (2008) using their framework. Analyzing only van der Wal et al. (2000), the authors find an effective size be- tween one and two, that is, it suffices to apply at most two basic steps to map, say, the clockwise circulating state to the anticlockwise one. Similarly low is the num- ber found by Nimmrichter et al. (2011), who assign an effective size of µ ≈ 5 to the experiment of Friedman et al. (2000). Again different results are obtained by Björk and Mana (2004), who apply their proposal to Friedman et al. (2000). They take the width of the coherent superposi- tion of the wavefunction living in both wells and divide this by the width of the ground state localized in only one Figure 21 Illustration of the physical principle behind the well. They find that the spread of the apparent macro- experiment reported in Friedman et al. (2000) aiming to scopic quantum state is only about 33 times larger than create coherent superpositions between clockwise and anti- that of the “classical” ground state. clockwise circulating currents using a SQUID. (a) Illustration of eigenenergies of the device as a function of the current flux. To apply the approach of Korsbakken et al. (2007), By controlling the parameters of the potential, the SQUID can a microscopic model of the experiment is necessary to be prepared in the eigenstate |ii corresponding to a current answer the question of how many electrons are effectively flux with a well defined circulation. Driving the transitions different in the two branches of the superposition. A from |ii to excited states |0i and |1i and controlling the pa- detailed analysis is provided in Korsbakken et al. (2009, rameters , ∆U0 of the double-well, the state ideally ends 2010), which leads to effective sizes of roughly 3800 - 5750 up in a superposition of clockwise and anti-clockwise circu- for Friedman et al. (2000), 42 for van der Wal et al. (2000) lating current. (b) The coherence of this superposition state is revealed thought an avoided crossing between the systems and 124 for Hime et al. (2006). The authors explain eigenenergies when  is tuned. (c) Schematics of the exper- the difference to Leggett’s result by taking into account iment with the SQUID inside the dashed box, the external the fermionic nature of the electrons, which reduces the control and an additional SQUID that operates as a magne- number of effectively different electrons. tometer to probe the energy of the neighboring SQUID. From It is not astonishing that these different results pro- Friedman et al. (2000). voked many discussions. The quantitative results of Leggett (2002) are critically seen by Björk and Mana (2004); Korsbakken et al. (2007, 2009); and Marquardt it shows the difficulties to find an agreement when evalu- et al. (2008). Leggett (2016) remarks that the results of ating the size of a given experiment or state with different Korsbakken et al. (2009) strongly depend on the choice measures. In Sec. V.E.2, we present results for several ex- some characteristic quantities such as the Fermi veloc- periments by applying the frameworks of Fröwis and Dür ity. To illustrate his claims, he discusses the hypothet- (2012b); Nimmrichter et al. (2011). ical example of a dust particle in the superposition of two macroscopically different momenta. By introducing

1. Discussions on the size of flux states in SQUID systems 50 For example, the disagreement between Korsbakken et al. (2009); Leggett (2002); Marquardt et al. (2008). Here, we summarize different contributions that aim 51 In particular, Björk and Mana (2004) (which was the square root for assigning an effective size to various SQUID experi- of, e.g., Korsbakken et al. (2007) in spin examples, Sec. II.B.1); ments. While there are clearly differences that arise from or Nimmrichter et al. (2011). Résultats de la collecte de fonds par vendeur

PARTICIPANT UNITÉS VENDUES

32

11

71

5

10

49 52 20

“reasonable” characteristic scales, he shows that the ap- 80 proach of Korsbakken et al. (2007) lead to trivially low effective size.52 Ne↵ To continue a critical dialogue, we would like to add 60 that the proposals of Korsbakken et al. (2007) and Mar- quardt et al. (2008) depend on the splitting of the total wave function into “dead” and “alive”, which can lead to 40 ambiguous situations (see example 2 in Sec. II.B). Fur- thermore, the choice of the “basic step” in the framework of Marquardt et al. (2008) seems to be intuitive, but 20 needs further justification. The approach of Björk and , 2016 et al. 2013 spin squeezing ion based cat like state Mana (2004) has a clear operational meaning, but the photonic squeezing microwave cat like state microwave cat like state ion based GHZ state et al. , 2016 photonic squeezing , 2011 connection to the idea of an effective size as followed by optical squeezinget al. , 2013 Vahlbruch Eberleoptical squeezinget al. et al. , 2016 the other works is unclear. A similar argument holds for Hostenspin squeezing Monzion - GHZ state et al. , 2016 the collapse model used by Nimmrichter et al. (2011). Vlastakis ion cat stateet al. microwaveKienzler cat state The unresolved answers regarding the interpretation of Wangmicrowave cat state the experimental evidence should be seen as a motivation Figure 22 Summary of effective sizes for several experiments to further improve the theory of macroscopic quantum- with (pseudo-)spins and photons. In various experimental se- 1 ness. tups, the macroscopic quantumness measured with the quan- tum Fisher information (Fröwis and Dür, 2012b) could be significantly increased over recent years. The numbers are 2. Comparing the size of observed states taken from Fröwis (2017).

Some measures presented in Sec. II.A are applicable to real experimental situations. In particular the measure of for the 8 photon GHZ state (Wang et al., 2016b) allows Fröwis and Dür (2012b) has been used in Fröwis (2017) the rough estimate Neff ≈ 2.3. A complete analysis of several experiments was done to compare the effective size Neff of various experimen- tal photonic and spin states. In Nimmrichter and Horn- in Fröwis (2017), see Fig. 22. For example, the spin- berger (2013), the effective size µ of states obtained in squeezing experiment of Hosten et al. (2016) results in various experiments with massive and SQUID systems an effective size of ≈ 71, which is comparable with an has been evaluated using the measure presented in the ideal GHZ state made out of 71 particles. Note the work same manuscript. We quickly summarize the main re- of Kienzler et al. (2016) realizes a spatial superposition sults of these two studies separately. Note that the two of a single atom, which is in apparent contradiction to measures are defined for different scales and hence a di- our initial premise that only large systems can have a large effective size. However, in this case, the system rect comparison between Neff and µ is meaningless. The measure of Fröwis and Dür (2012b) is based on the size is taken as the mean number of phononic excitations quantum Fisher information. This is a promising mixed- in a harmonic trap, which is similar to the treatment of state extension of the variance because of tight and acces- single-mode photonic systems. sible lower bounds (see Sec. II.H.2). For (spin)squeezed The measure of Nimmrichter and Hornberger (2013) states, the left-hand side of the tighter Heisenberg un- is based on the capability of a state to test modifica- certainty relation, Eq. (63), becomes identical to ξ−2, tions of quantum theory. In case of interference experi- Eq. (8). For experiments targeting superpositions of two ments with objects of total mass M whose expansion is “classical” states (e.g., GHZ state or SCS), Eq. (62) can much smaller than the path separation of the interferom- be used to bound the quantum Fisher information by wit- eter, the modification of the master equation shown in nessing large susceptibility to small external influences. Eq. (36) leads to a decay of the coherence which scales ⊗N ⊗N 2 Measuring the coherence terms C = 2| h0| ρ |1i | as (me/M) where me is a reference mass taken as the for the GHZ and C = 2| hα| ρ |−αi | for the mono- mass of an electron. By comparing this decay with the mode SCS, allows to derive (approximate) lower bounds actual period t (in second) during which the coherence 2 2 2 is maintained and taking the contrast f of the observed Neff (GHZ) ≥ C N and Neff (SCS) & 4C |α| + 1, respec- tively. Since these quantities are frequently measured, interference pattern into account, a simple approximate a first estimate can often be directly done with the data expression is obtained for evaluating the effective size given in the publication. For example, the published data h 1  M 2 i µ = log10 t . (73) | ln f| me

52 Similar arguments hold for the analysis of Nimmrichter et al. For example, for the experiment reported in Kovachy (2011). et al. (2015) where the wave packet of 87Rb atoms (86.91 53 a.m.u.) gets separated over 54cm during a drift time of with very high squeezing parameters. Matter-wave inter- t = 2.08s and led to an interference pattern with a con- ferometry and quantum optomechanics with more and trast of 28%, the above expression gives to µ = 10.6. more massive systems are also at the core of intensive Note that higher values have been obtained in the same experimental efforts. experiment with smaller spatial separations. In partic- According to a measure based on the capability of a ular, µ = 12.3 has been obtained for a separation of Résultatsstate de tola collecte test modificationsde fonds par vendeur of quantum theory (Nimm- ∼ 1cm where the observed contrast is of 97.5%. This richter and Hornberger, 2013), very large effective size PARTICIPANT UNITÉS VENDUES value is comparable to interferometry experiments with have been obtained almost 20 years ago, with molecule 12 cluster and molecules (Arndt et al., 1999; Gerlich et al., interferometry. Interestingly, such a measure also witness 7 2011). By describing the superposition of currents with large effective size for single-atom interferometry where 11 displaced Fermi spheres of Cooper pairs, Nimmrichter atomic wave packets gets coherently separated for sec- 5 and Hornberger found µ = 5.2 for the SQUID experiment onds. As noted in Arndt and Hornberger (2014), there is 12 reported in Friedman et al. (2000), see Nimmrichter and plenty of room for improvement and unprecedented sizes Hornberger (2013) for details. These results are shown could be obtained in a near future using e.g. levitating graphically in Fig. 23. nano-spheres. µ 15 VI. DISCUSSION AND OUTLOOK 12 In this review we have summarized different ap- 9 proaches to qualify and quantify the notion of macro- scopic quantum superpositions and macroscopic quan- 6 tum states. While we have seen that there are a mul- 3 titude of proposals that differ strongly in their intuition and its formalization, many of the measures agree on , 2011 et al. , 1988 , 1999 some core features. In particular, it seem that the vari- et al. et al. , 2000 Gerlich ance with respect to linear observables plays a key role in Keith Arndt et al. this respect. It is probably to early to say that we have Cluster (PFNS8)Atom interferometry (Na) interferometry Friedman obtained an agreement on a specific measure, or even all Molecule (C60) interferometryKovachy87 Rb) et interferometry al., 2015 relevant or desirable features of such measures, but there current superposition - SQUIDS Atom ( has been significant progress in the last couple of years. Figure 23 . Effective size of various states associated to mas-1 As we have discussed in detail, there are many facets of sive systems evaluated with the measure in Nimmrichter and the problem, and perhaps there is no single measure that Hornberger (2013). The numbers are taken from Nimmrichter takes all these aspects into account. Establishing a re- and Hornberger (2013) except for the experiment reported in source theory for quantum macroscopicity might seems Kovachy et al. (2015) where the calculation is shown in the to be an attractive avenue, but a good choice of free op- main text. erations remains a challenge. Nevertheless, we are now much closer to be capable of judging experiments, or pro- viding guidelines in which direction to go. The latter is of particular importance given the mul- F. Summary titude of fundamental limitations to prepare, maintain and measure macroscopic quantum states that have been Many attempts have been realized to create and de- identified (see Sec. III). While certain macroscopic quan- tect macroscopic quantum states. Photonic experiments tum states, in particular certain kinds of macroscopic have been implemented both in the optical domain using superposition states, seem to be notoriously difficult to parametric processes and in the microwave domain in the maintain and certify even within the framework of stan- framework of cavity quantum electrodynamics. While dard quantum mechanics, other states were identified the size of optical SCS has been improved by a factor ≈4 where such limitations do not apply. It is still very chal- in the last decade now achieving α2 ≈ 3.2, unprecedented lenging to perform experiments with such macroscopic sizes α2 ≈ 7.8 has been achieved in the microwave do- quantum states, however there seem to be no principle main. However, optical squeezed state holds the largest obstacles to prevent us from observing quantum effects effective size for photonic states, according to a measure on much larger scales than today. In some cases only based on the quantum Fisher information (Fröwis and a relatively small change is required: working with rel- Dür, 2012b). The same measure witnesses unprecedented ative degrees of freedom rather than absolute degrees sizes in spin systems, in particular in Bose-Einstein con- of freedom suffices, or preparing two copies of a state densates where spin squeezed states have been obtained enables one of them to act as a kind of self-reference. 54

Also encoded macroscopic states pose an interesting per- Bassi, A., K. Lochan, S. Satin, T. P. Singh, and H. Ulbricht spective. While the notion of macroscopic quantumness (2013), Rev. Mod. Phys. 85, 471. might slightly change, they possess the desired features Baumgratz, T., M. Cramer, and M. Plenio (2014), Phys. Rev. on a coarse-grained level. If these insights can be har- Lett. 113 (14), 140401. Benhelm, J., G. Kirchmair, C. F. Roos, and R. Blatt (2008), nessed for applications such a quantum metrology or Nat Phys 4 (6), 463. quantum computation remains to be seen. Bhattacharyya, A. (1946), Sankhy¯a: The Indian Journal of At the level of experiments, recent years have seen Statistics (1933-1960) 7 (4), 401. tremendous progress with different physical setups, Björk, G., and P. G. L. Mana (2004), Journal of Optics B: bringing us ever closer to a true macroscopic regime. Quantum and Semiclassical Optics 6 (11), 429. Which system is most suited to demonstrate truly macro- Blatt, R., and D. Wineland (2008), Nature 453 (7198), 1008. Bohnet, J. G., B. C. Sawyer, J. W. Britton, M. L. Wall, scopic quantum effects depends on the goal one has in A. M. Rey, M. Foss-Feig, and J. J. Bollinger (2016), Science mind. However, we are now not only able to test the 352 (6291), 1297. fundamental principles of quantum mechanics or its va- Bose, S., K. Jacobs, and P. L. Knight (1997), Phys. Rev. A lidity at larger and larger scales, but also harness some 56, 4175. of its features for practical applications. Its perhaps the Bose, S., K. Jacobs, and P. L. Knight (1999), Phys. Rev. A mixture of fundamental interest and the possibility of 59, 3204. practical applications that makes the study of macro- Braunstein, S. L., and C. M. Caves (1994), Phys. Rev. Lett. 72 (22), 3439. scopic quantum states so appealing. The road ahead still Briegel, H. J., D. E. Browne, W. Dür, R. Raussendorf, and promises many challenges, but also new insights and sur- M. Van den Nest (2009), Nat. Phys. , 19. prises. While it seems that it is impossible to ever realize Briegel, H. J., and R. Raussendorf (2001), Phys. Rev. Lett. the thought experiment of Schrödinger – performing ex- 86, 910. periments where the spirit of his proposal is maintained Bruno, N., A. Martin, P. Sekatski, N. Sangouard, R. T. Thew, might at least plausible. and N. Gisin (2013), Nat Phys 9 (9), 545. Bruß, D., M. Cinchetti, G. Mauro D’Ariano, and C. Macchi- avello (2000), Phys. Rev. A 62 (1), 012302. Caldeira, A. O. (2014), An Introduction to Macroscopic Quan- ACKNOWLEDGMENTS. tum Phenomena and Quantum Dissipation (Cambridge University Press). We thank all colleagues with whom we have dis- Caldeira, A. O., and A. J. Leggett (1985), Phys. Rev. A 31, cussed about this topic over the last years. This 1059. Carlisle, A., H. Kwon, H. Jeong, A. Ferraro, and M. Pater- work was supported by the European Research Coun- nostro (2015), Phys. Rev. A 92, 022123. cil (ERC-AG MEC), the Swiss National Science Founda- Cavalcanti, D., R. Chaves, L. Aolita, L. Davidovich, and tion through grant No. P300P2_167749, 200021_149109 A. Acín (2009), Phys. Rev. Lett. 103, 030502. and PP00P2_150579 and through Quantum Science Cavalcanti, E. G., and M. D. Reid (2006), Phys. Rev. Lett. and Technology (QSIT) and the Austrian Science Fund 97 (17), 170405. (FWF) through project P28000-N27. Cavalcanti, E. G., and M. D. Reid (2008), Phys. Rev. A 77 (6), 062108. Chan, J., T. P. M. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Groblacher, M. Aspelmeyer, and O. Painter REFERENCES (2011), Nature 478 (7367), 89. Chang, D. E., C. A. Regal, S. B. Papp, D. J. Wilson, J. Ye, Agarwal, G. (1998), Physical Review A 57 (1), 671. O. Painter, H. J. Kimble, and P. Zoller (2010), Proceedings Aolita, L., D. Cavalcanti, R. Chaves, C. Dhara, L. Davidovich, of the National Academy of Sciences 107 (3), 1005. and A. Acín (2010), Phys. Rev. A 82, 032317. Chekhova, M., G. Leuchs, and M. Żukowski (2015), Optics Arecchi, F., E. Courtens, R. Gilmore, and H. Thomas (1972), Communications 337, 27 , macroscopic quantumness: the- Physical Review A 6 (6), 2211. ory and applications in optical sciences. Arita, Y., M. Mazilu, T. Vettenburg, E. M. Wright, and Chou, C. H., B. L. Hu, and Y. Subaşi (2011), J. Phys.: Conf. K. Dholakia (2015), Opt. Lett. 40 (20), 4751. Ser. 306 (1), 012002. Arndt, M., and K. Hornberger (2014), Nat Phys 10 (4), 271. Cramér, H. (1945), Mathematical Methods of Statistics (PMS- Arndt, M., O. Nairz, J. Vos-Andreae, C. Keller, G. van der 9) (Princeton University Press). Zouw, and A. Zeilinger (1999), Nature 401 (6754), 680. Cronin, A. D., J. Schmiedmayer, and D. E. Pritchard (2009), Arrad, G., Y. Vinkler, D. Aharonov, and A. Retzker (2014), Rev. Mod. Phys. 81, 1051. Phys. Rev. Lett. 112 (15), 150801. Dakić, B., and M. Radonjić (2017), Phys. Rev. Lett. 119, Asenbaum, P., C. Overstreet, T. Kovachy, D. D. Brown, J. M. 090401. Hogan, and M. A. Kasevich (2017), Phys. Rev. Lett. 118, Dakna, M., T. Anhut, T. Opatrný, L. Knöll, and D.-G. 183602. Welsch (1997), Phys. Rev. A 55, 3184. Aspelmeyer, M., T. J. Kippenberg, and F. Marquardt (2014), D’Ariano, G. M., and C. Macchiavello (2003), Phys. Rev. A Rev. Mod. Phys. 86, 1391. 67 (4), 042306. Barker, P. F., and M. N. Shneider (2010), Phys. Rev. A 81, Davis, E., G. Bentsen, and M. Schleier-Smith (2016), Phys. 023826. Rev. Lett. 116, 053601. 55

De Martini, F., and F. Sciarrino (2012), Rev. Mod. Phys. 84, Fröwis, F., N. Sangouard, and N. Gisin (2015), Optics Com- 1765. munications 337, 2 . De Martini, F., and F. Sciarrino (2015), Optics Communi- Fröwis, F., R. Schmied, and N. Gisin (2015), Phys. Rev. A cations 337, 44 , macroscopic quantumness: theory and 92 (1), 012102. applications in optical sciences. Fröwis, F., P. Sekatski, and W. Dür (2016), Phys. Rev. Lett. De Martini, F., F. Sciarrino, and C. Vitelli (2008), Phys. 116 (9), 090801. Rev. Lett. 100, 253601. Fuchs, C. A., and J. Van De Graaf (1999), Information The- Deleglise, S., I. Dotsenko, C. Sayrin, J. Bernu, M. Brune, J.- ory, IEEE Transactions on 45 (4), 1216. M. Raimond, and S. Haroche (2008), Nature 455 (7212), Fujiwara, A., and H. Imai (2008), Journal of Physics A: Math- 510. ematical and Theoretical 41 (25), 255304. Demkowicz-Dobrzanski, R., J. Kolodynski, and M. Guta Gerlich, S., S. Eibenberger, M. Tomandl, S. Nimmrichter, (2012), Nature Communications 3, 1063 EP , article. K. Hornberger, P. J. Fagan, J. Tüxen, M. Mayor, and Demkowicz-Dobrzański, R., and M. Markiewicz (2015), Phys. M. Arndt (2011), Nat Commun 2, 263. Rev. A 91 (6), 062322. Gerrits, T., S. Glancy, T. S. Clement, B. Calkins, A. E. Lita, Devi, A. R. U., H. S. Karthik, Sudha, and A. K. Rajagopal A. J. Miller, A. L. Migdall, S. W. Nam, R. P. Mirin, and (2013), Phys. Rev. A 87 (5), 052103. E. Knill (2010), Phys. Rev. A 82, 031802. Devoret, M. H., and R. J. Schoelkopf (2013), Science Ghirardi, G. C., P. Pearle, and A. Rimini (1990), Phys. Rev. 339 (6124), 1169. A 42, 78. Dicke, R. H. (1954), Phys. Rev. 93 (1), 99. Ghirardi, G. C., A. Rimini, and T. Weber (1986), Phys. Rev. Diósi, L. (1989), Phys. Rev. A 40, 1165. D 34, 470. Diósi, L. (2015), Phys. Rev. Lett. 114, 050403. Gieseler, J., B. Deutsch, R. Quidant, and L. Novotny (2012), Duan, L.-M., M. D. Lukin, J. I. Cirac, and P. Zoller (2001), Phys. Rev. Lett. 109, 103603. Nature 414 (6862), 413. Giovannetti, V., S. Lloyd, and L. Maccone (2011), Nat Pho- Dür, W., C. Simon, and J. I. Cirac (2002), Phys. Rev. Lett. ton 5 (4), 222. 89 (21), 210402. Gisin, N. (1989), Helvetica Physica Acta 62, 363. Dür, W., M. Skotiniotis, F. Fröwis, and B. Kraus (2014), Glauber, R. J. (1963), Physical Review 131 (6), 2766. Phys. Rev. Lett. 112 (8), 080801. Gong, J. (2011), arXiv:1106.0062 [quant-ph] . Dür, W., G. Vidal, and J. I. Cirac (2000), Phys. Rev. A Goold, J., M. Huber, A. Riera, L. del Rio, and P. Skrzypczyk 62 (6), 062314. (2016), J. Phys. A: Math. Theor. 49 (14), 143001. Eberle, T., V. Händchen, and R. Schnabel (2013), Opt. Ex- Gould, P. L., G. A. Ruff, and D. E. Pritchard (1986), Phys. press 21 (9), 11546. Rev. Lett. 56, 827. Eibenberger, S., S. Gerlich, M. Arndt, M. Mayor, and Gour, G., M. P. Müller, V. Narasimhachar, R. W. Spekkens, J. Tuxen (2013), Phys. Chem. Chem. Phys. 15, 14696. and N. Yunger Halpern (2015), Physics Reports The re- Eisenberg, H. S., G. Khoury, G. A. Durkin, C. Simon, and source theory of informational nonequilibrium in thermo- D. Bouwmeester (2004), Phys. Rev. Lett. 93, 193901. dynamics, 583, 1. Ellis, J., S. Mohanty, and D. V. Nanopoulos (1989), Physics Gour, G., and R. W. Spekkens (2008), New J. Phys. 10 (3), Letters B 221 (2), 113 . 033023. Emary, C., N. Lambert, and F. Nori (2014), Reports on Greenberger, D. M., M. A. Horne, and A. Zeilinger (1989), in Progress in Physics 77 (1), 016001. Bell’s Theorem, Quantum Theory and Conceptions of the Escher, B. M., R. L. de Matos Filho, and L. Davidovich Universe, Fundamental Theories of Physics No. 37, edited (2011), Nat Phys 7 (5), 406. by M. Kafatos (Springer Netherlands) pp. 69–72. Etesse, J., M. Bouillard, B. Kanseri, and R. Tualle-Brouri Gross, C., T. Zibold, E. Nicklas, J. Estève, and M. K. (2015), Phys. Rev. Lett. 114, 193602. Oberthaler (2010), Nature 464 (7292), 1165. Everett, H. (1957), Rev. Mod. Phys. 29 (3), 454. Grover, L. K. (1997), Phys. Rev. Lett. 79 (2), 325. Fabre, C., M. Pinard, S. Bourzeix, A. Heidmann, E. Gia- Gühne, O., P. Hyllus, O. Gittsovich, and J. Eisert (2007), cobino, and S. Reynaud (1994), Phys. Rev. A 49, 1337. Phys. Rev. Lett. 99 (13), 130504. Farrow, T., and V. Vedral (2015), Optics Communications Halliwell, J. J. (2017), Phys. Rev. A 96 (1), 012121. Macroscopic quantumness: theory and applications in op- Haroche, S. (2013), Rev. Mod. Phys. 85, 1083. tical sciences, 337, 22. Hauke, P., M. Heyl, L. Tagliacozzo, and P. Zoller (2016), Nat Fleming, G. N. (1973), Il Nuovo Cimento A 16 (2), 232. Phys 12 (8), 778, article. Friedman, J. R., V. Patel, W. Chen, S. K. Tolpygo, and J. E. Hein, M., W. Dür, J. Eisert, R. Raussendorf, M. V. den Lukens (2000), Nature 406 (6791), 43. Nest, and H. J. Briegel (2006), in Proceedings of the In- Fröwis, F. (2012), Phys. Rev. A 85 (5), 052127. ternational School of Physics "Enrico Fermi", Vol. 162 Fröwis, F. (2017), J. Phys. A: Math. Theor. 50 (11), 114003. (Varenna, Italia) Chap. 5, pp. 115 – 218. Fröwis, F., and W. Dür (2011), Phys. Rev. Lett. 106, 110402. Helstrom, C. (1976), Quantum Detection and Estimation The- Fröwis, F., and W. Dür (2012a), Phys. Rev. Lett. 109, ory (Academic Press). 170401. Hime, T., P. A. Reichardt, B. L. T. Plourde, T. L. Robertson, Fröwis, F., and W. Dür (2012b), New Journal of Physics C.-E. Wu, A. V. Ustinov, and J. Clarke (2006), Science 14 (9), 093039. 314 (5804), 1427. Fröwis, F., G. Grübl, and M. Penz (2008), Journal of Physics Holevo, A. S. (2011), Probabilistic and Statistical Aspects of A: Mathematical and Theoretical 41, 405201. Quantum Theory (Springer). Fröwis, F., M. van den Nest, and W. Dür (2013), New Journal Holstein, T., and H. Primakoff (1940), Phys. Rev. 58 (12), of Physics 15 (11), 113011. 1098. 56

Horodecki, R., P. Horodecki, M. Horodecki, and K. Horodecki Korsbakken, J. I., F. K. Wilhelm, and K. B. Whaley (2009), (2009), Rev. Mod. Phys. 81, 865. Phys. Scr. 2009 (T137), 014022. Hosten, O., N. J. Engelsen, R. Krishnakumar, and M. A. Korsbakken, J. I., F. K. Wilhelm, and K. B. Whaley (2010), Kasevich (2016), Nature 529 (7587), 505. EPL 89 (3), 30003. Hotta, M., and M. Ozawa (2004), Phys. Rev. A 70 (2), Kovachy, T., P. Asenbaum, C. Overstreet, C. A. Donnelly, 022327. S. M. Dickerson, A. Sugarbaker, J. M. Hogan, and M. A. Husimi, K. (1940), Nippon Sugaku-Buturigakkwai Kizi Dai 3 Kasevich (2015), Nature 528 (7583), 530. Ki 22 (4), 264. Kovachy, T., P. Asenbaum, C. Overstreet, C. A. Donnelly, Hyllus, P., W. Laskowski, R. Krischek, C. Schwemmer, S. M. Dickerson, A. Sugarbaker, J. M. Hogan, and M. A. W. Wieczorek, H. Weinfurter, L. Pezzé, and A. Smerzi Kasevich (2016), Nature 537 (7618), E2. (2012), Phys. Rev. A 85, 022321. Kumari, S., and A. K. Pan (2017), Phys. Rev. A 96 (4), Iskhakov, T. S., I. N. Agafonov, M. V. Chekhova, and 042107. G. Leuchs (2012), Phys. Rev. Lett. 109, 150502. Kwon, H., C.-Y. Park, K. C. Tan, and H. Jeong (2017), New Iskhakov, T. S., M. V. Chekhova, G. O. Rytikov, and J. Phys. 19 (4), 043024. G. Leuchs (2011), Phys. Rev. Lett. 106, 113602. Laghaout, A., J. S. Neergaard-Nielsen, and U. L. Andersen Jaeger, G. (2014), American Journal of Physics 82 (9), 896. (2015), Optics Communications Macroscopic quantumness: Jeong, H., M. Kang, and H. Kwon (2015), Optics Commu- theory and applications in optical sciences, 337, 96. nications 337, 12 , macroscopic quantumness: theory and Lambert, N., K. Debnath, A. F. Kockum, G. C. Knee, W. J. applications in optical sciences. Munro, and F. Nori (2016), Phys. Rev. A 94 (1), 012105. Jeong, H., M. Kang, and C.-W. Lee (2011), arXiv:1108.0212 Landini, M., M. Fattori, L. Pezzè, and A. Smerzi (2014), New [quant-ph] . Journal of Physics 16 (11), 113074. Jeong, H., C. Noh, S. Bae, D. G. Angelakis, and T. C. Ralph Lanyon, B. P., M. Zwerger, P. Jurcevic, C. Hempel, W. Dür, (2014a), J. Opt. Soc. Am. B, JOSAB 31 (12), 3057. H. J. Briegel, R. Blatt, and C. F. Roos (2014), Phys. Rev. Jeong, H., M. Paternostro, and T. C. Ralph (2009), Physical Lett. 112, 100403. review letters 102 (6), 060403. Lee, C.-W., and H. Jeong (2011), Phys. Rev. Lett. 106, Jeong, H., A. Zavatta, M. Kang, S.-W. Lee, L. S. Costanzo, 220401. S. Grandi, T. C. Ralph, and M. Bellini (2014b), Nat Pho- Leggett, A. J. (1980), Prog. Theor. Phys. Suppl. 69, 80. ton 8 (7), 564. Leggett, A. J. (1987), Le hasard et la matiere = chance and Joos, E. (2003), Decoherence and the Appearance of a Classi- matter: les houches session XLVI, 30 juin-1 er aout 1986, cal World in Quantum Theory (Springer Science & Business edited by J. Souletie, R. Stora, and J. Vannimenus (North- Media). Holland Physics Publishing, Amsterdam) Chap. 6. Keith, D. W., M. L. Schattenburg, H. I. Smith, and D. E. Leggett, A. J. (2002), Journal of Physics: Condensed Matter Pritchard (1988), Phys. Rev. Lett. 61, 1580. 14 (15), R415. Kendon, V. M., and W. J. Munro (2006), Quantum Info. Leggett, A. J. (2016), arXiv:1603.03992 [quant-ph] . Comput. 6 (7), 630. Leggett, A. J., and A. Garg (1985), Phys. Rev. Lett. 54 (9), Kessler, E., I. Lovchinsky, A. Sushkov, and M. Lukin (2014), 857. Phys. Rev. Lett. 112 (15), 150802. Leibfried, D., R. Blatt, C. Monroe, and D. Wineland (2003), Kholevo, A. (1974), Theory Probab. Appl. 18 (2), 359. Rev. Mod. Phys. 75, 281. Kienzler, D., C. Flühmann, V. Negnevitsky, H.-Y. Lo, Lvovsky, A. I., R. Ghobadi, A. Chandra, A. S. Prasad, and M. Marinelli, D. Nadlinger, and J. P. Home (2016), Phys. C. Simon (2013), Nat Phys 9 (9), 541. Rev. Lett. 116, 140402. Makhlin, Y., G. Schön, and A. Shnirman (2001), Rev. Mod. Kiesel, N., F. Blaser, U. Delic, D. Grass, R. Kaltenbaek, Phys. 73, 357. and M. Aspelmeyer (2013), Proceedings of the National Mancini, S., and P. Tombesi (1994), Phys. Rev. A 49, 4055. Academy of Sciences 110 (35), 14180. Mandelstam, L., and I. Tamm (1945), J. Phys. (UDSSR) Kitagawa, M., and M. Ueda (1993), Phys. Rev. A 47 (6), 9 (4), 249. 5138. Marquardt, F., B. Abel, and J. von Delft (2008), Phys. Rev. Kleckner, D., I. Pikovski, E. Jeffrey, L. Ament, E. Eliel, A 78, 012109. J. van den Brink, and D. Bouwmeester (2008), New Jour- Marshall, W., C. Simon, R. Penrose, and D. Bouwmeester nal of Physics 10 (9), 095020. (2003), Phys. Rev. Lett. 91, 130401. Klimov, A. B., and S. M. Chumakov (2009), A Group- Marvian, I., and R. W. Spekkens (2016), Phys. Rev. A 94 (5), Theoretical Approach to Quantum Optics (John Wiley & 052324. Sons). Meenehan, S. M., J. D. Cohen, G. S. MacCabe, F. Mar- Knee, G. C., K. Kakuyanagi, M.-C. Yeh, Y. Matsuzaki, sili, M. D. Shaw, and O. Painter (2015), Phys. Rev. X H. Toida, H. Yamaguchi, S. Saito, A. J. Leggett, and W. J. 5, 041002. Munro (2016), Nature Communications 7, 13253. Mermin, N. D. (1980), Phys. Rev. D 22, 356. Kofler, J., and Č. Brukner (2008), Physical review letters Meystre, P. (2013), Annalen der Physik 525 (3), 215. 101 (9), 090403. Milburn, G. J., and C. A. Holmes (1986), Phys. Rev. Lett. Kofler, J., and C. Brukner (2010), arXiv:1009.2654 . 56, 2237. Kofler, J., and Č. Brukner (2013), Phys. Rev. A 87 (5), Millen, J., P. Z. G. Fonseca, T. Mavrogordatos, T. S. Mon- 052115. teiro, and P. F. Barker (2015), Phys. Rev. Lett. 114, Kofler, J., and i. c. v. Brukner (2007), Phys. Rev. Lett. 99, 123602. 180403. Monroe, C., D. M. Meekhof, B. E. King, and D. J. Wineland Korsbakken, J. I., K. B. Whaley, J. Dubois, and J. I. Cirac (1996), Science 272 (5265), 1131. (2007), Phys. Rev. A 75, 042106. 57

Monz, T., P. Schindler, J. T. Barreiro, M. Chwalla, D. Nigg, Riedel, M. F., P. Böhi, Y. Li, T. W. Hänsch, A. Sinatra, and W. A. Coish, M. Harlander, W. Hänsel, M. Hennrich, and P. Treutlein (2010), Nature 464 (7292), 1170. R. Blatt (2011), Phys. Rev. Lett. 106, 130506. Ritsch, H., P. Domokos, F. Brennecke, and T. Esslinger Morimae, T. (2010), Phys. Rev. A 81 (1), 010101. (2013), Rev. Mod. Phys. 85, 553. Morimae, T., A. Sugita, and A. Shimizu (2005), Phys. Rev. Robens, C., W. Alt, D. Meschede, C. Emary, and A. Alberti A 71 (3), 032317. (2015), Phys. Rev. X 5 (1), 011003. Morin, O., K. Huang, J. Liu, H. Le Jeannic, C. Fabre, and Romero-Isart, O. (2011), Phys. Rev. A 84, 052121. J. Laurat (2014), Nat Photon 8 (7), 570. Romero-Isart, O., A. C. Pflanzer, F. Blaser, R. Kaltenbaek, Neergaard-Nielsen, J. S., B. M. Nielsen, C. Hettich, N. Kiesel, M. Aspelmeyer, and J. I. Cirac (2011), Phys. K. Mølmer, and E. S. Polzik (2006), Phys. Rev. Lett. 97, Rev. Lett. 107, 020405. 083604. Saha, D., S. Mal, P. K. Panigrahi, and D. Home (2015), Phys. Van den Nest, M., W. Dür, A. Miyake, and H. J. Briegel Rev. A 91 (3), 032117. (2007), New Journal of Physics 9 (6), 204. Sayrin, C., I. Dotsenko, X. Zhou, B. Peaudecerf, T. Rybar- Van den Nest, M., K. Luttmer, W. Dür, and H. J. Briegel czyk, S. Gleyzes, P. Rouchon, M. Mirrahimi, H. Amini, (2008), Phys. Rev. A 77, 012301. M. Brune, J.-M. Raimond, and S. Haroche (2011), Nature Nielsen, M. A., and I. L. Chuang (2000), Quantum Compu- 477 (7362), 73. tation and Quantum Information (Cambridge University Schlosshauer, M. (2005), Rev. Mod. Phys. 76 (4), 1267. Press, New York, NY, USA) 2010 : 10th anniversary ed. Schmied, R. (2017), Habilitationsschrift, University of Basel . Nimmrichter, S., and K. Hornberger (2013), Phys. Rev. Lett. Schmied, R., J.-D. Bancal, B. Allard, M. Fadel, V. Scarani, 110, 160403. P. Treutlein, and N. Sangouard (2016), Science 352 (6284), Nimmrichter, S., K. Hornberger, and K. Hammerer (2014), 441. Phys. Rev. Lett. 113 (2), 020405. Schoelkopf, R. J., and S. M. Girvin (2008), Nature Nimmrichter, S., K. Hornberger, P. Haslinger, and M. Arndt 451 (7179), 664. (2011), Phys. Rev. A 83, 043621. Schrödinger, E. (1935), Die Naturwissenschaften 23 (48), 1. O’Connell, A. D., M. Hofheinz, M. Ansmann, R. C. Bialczak, Scully, M., and M. Zubairy (1997), Quantum Optics (Cam- M. Lenander, E. Lucero, M. Neeley, D. Sank, H. Wang, bridge University Press). M. Weides, J. Wenner, J. M. Martinis, and A. N. Cleland Scully, M. O., E. S. Fry, C. H. R. Ooi, and K. Wódkiewicz (2010), Nature 464 (7289), 697. (2006), Phys. Rev. Lett. 96 (1), 010501. Orús, R., and J. I. Latorre (2004), Phys. Rev. A 69, 052308. Sekatski, P., M. Aspelmeyer, and N. Sangouard (2014a), Oudot, E., P. Sekatski, F. Fröwis, N. Gisin, and N. Sangouard Phys. Rev. Lett. 112, 080502. (2015), J. Opt. Soc. Am. B 32 (10), 2190. Sekatski, P., N. Brunner, C. Branciard, N. Gisin, and C. Si- Ourjoumtsev, A., R. Tualle-Brouri, J. Laurat, and P. Grang- mon (2009), Phys. Rev. Lett. 103, 113601. ier (2006), Science 312 (5770), 83. Sekatski, P., N. Gisin, and N. Sangouard (2014b), Phys. Rev. Ozawa, M. (2000), Physics Letters A 268 (3), 158. Lett. 113, 090403. Palomaki, T. A., J. D. Teufel, R. W. Simmonds, and K. W. Sekatski, P., N. Sangouard, and N. Gisin (2014c), Phys. Rev. Lehnert (2013), Science 342 (6159), 710. A 89, 012116. Pan, J.-W., Z.-B. Chen, C.-Y. Lu, H. Weinfurter, A. Zeilinger, Sekatski, P., N. Sangouard, M. Stobińska, F. Bussières, and M. Żukowski (2012), Rev. Mod. Phys. 84, 777. M. Afzelius, and N. Gisin (2012), Phys. Rev. A 86, 060301. Park, C.-Y., and H. Jeong (2016), “Disappearance of Sekatski, P., M. Skotiniotis, J. Kołodyński, and W. Dür macroscopic superpositions in perfectly isolated systems,” (2017a), Quantum 1, 27. arXiv:1606.07213. Sekatski, P., B. Yadin, M.-O. Renou, W. Dür, N. Gisin, and Park, C.-Y., M. Kang, C.-W. Lee, J. Bang, S.-W. Lee, and F. Fröwis (2017b), arXiv:1704.02270 [quant-ph] . H. Jeong (2016), Phys. Rev. A 94 (5), 052105. Shimizu, A., Y. Matsuzaki, and A. Ukena (2013), J. Phys. Penrose, R. (1996), Gen. Relativ. Gravit. 28, 581. Soc. Jpn. 82 (5), 054801. Peres, A. (2002), Quantum Theory: Concepts and Methods Shimizu, A., and T. Miyadera (2002), Phys. Rev. Lett. 89, (Springer). 270403. Pezzé, L., and A. Smerzi (2009), Phys. Rev. Lett. 102 (10), Shimizu, A., and T. Morimae (2005), Phys. Rev. Lett. 95, 100401. 090401. Pezzè, L., A. Smerzi, M. K. Oberthaler, R. Schmied, and Shimizu, A., and T. Morimae (2016), Phys. Rev. Lett. P. Treutlein (2016), arXiv:1609.01609 [quant-ph] . 117 (21), 219903. Piani, M. (2012), Phys. Rev. A 86 (3), 034101. Shitara, T., and M. Ueda (2016), Phys. Rev. A 94, 062316. Radhakrishna Rao, C. (1945), Bull. Calcutta Math. Soc. 37, Shor, P. (1999), SIAM Rev. 41 (2), 303. 81. Simon, C., and D. Bouwmeester (2003), Phys. Rev. Lett. 91, Raeisi, S., P. Sekatski, and C. Simon (2011), Phys. Rev. Lett. 053601. 107, 250401. Simon, C., and J. Kempe (2002), Phys. Rev. A 65, 052327. Raimond, J. M., M. Brune, and S. Haroche (2001), Rev. Skotiniotis, M., W. Dür, and P. Sekatski (2017), Quantum Mod. Phys. 73, 565. 1, 34. Ranjit, G., D. P. Atherton, J. H. Stutz, M. Cunningham, and Sørensen, A., L.-M. Duan, J. I. Cirac, and P. Zoller (print A. A. Geraci (2015), Phys. Rev. A 91, 051805. January 04, 2001), Nature 409 (6816), 63. Raussendorf, R., and H. J. Briegel (2001), Phys. Rev. Lett. Sørensen, A., and K. Mølmer (1999), Phys. Rev. Lett. 82, 86, 5188. 1971. Raussendorf, R., J. Harrington, and K. Goyal (2007), New Sørensen, A. S., and K. Mølmer (2001), Phys. Rev. Lett. J. Phys. 9 (6), 199. 86 (20), 4431. Reid, M. D. (2016), arXiv:1612.08883 [quant-ph] . 58

Stamper-Kurn, D. M., G. E. Marti, and H. Müller (2016), L. Frunzio, S. M. Girvin, L. Jiang, M. Mirrahimi, M. H. Nature 537 (7618), E1. Devoret, and R. J. Schoelkopf (2016a), Science 352 (6289), Sudarshan, E. (1963), Physical Review Letters 10 (7), 277. 1087. Sugita, A., and A. Shimizu (2005), J. Phys. Soc. Jpn. 74 (7), Wang, T., R. Ghobadi, S. Raeisi, and C. Simon (2013), Phys. 1883. Rev. A 88, 062114. Sychev, D., A. Ulanov, A. Pushkina, M. Richards, I. Fedorov, Wang, X.-L., L.-K. Chen, W. Li, H.-L. Huang, C. Liu, and A. Lvovsky (2017), Nat Photon 11, 379. C. Chen, Y.-H. Luo, Z.-E. Su, D. Wu, Z.-D. Li, H. Lu, Tatsuta, M., and A. Shimizu (2017), arXiv:1703.05034 [cond- Y. Hu, X. Jiang, C.-Z. Peng, L. Li, N.-L. Liu, Y.-A. Chen, mat, physics:physics, physics:quant-ph] . C.-Y. Lu, and J.-W. Pan (2016b), Phys. Rev. Lett. 117, Teufel, J. D., T. Donner, D. Li, J. W. Harlow, M. S. Allman, 210502. K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, Wheeler, J. A. (1957), Rev. Mod. Phys. 29 (3), 463. and R. W. Simmonds (2011), Nature 475 (7356), 359. Wigner, E. (1932), Phys. Rev. 40 (5), 749. Tóth, G. (2012), Phys. Rev. A 85 (2), 022322. Wigner, E. (1961), in The Scientist Speculates, edited by I. J. Tóth, G., and I. Apellaniz (2014), Journal of Physics A: Good (Heineman) pp. 284–302. Mathematical and Theoretical 47 (42), 424006. Wilde, M. M., and A. Mizel (2012), Foundations of Physics Tóth, G., and D. Petz (2013), Phys. Rev. A 87 (3), 032324. 42 (2), 256. Uhlmann, A. (1991), Lett Math Phys 21 (3), 229. Wineland, D. J. (2013), Rev. Mod. Phys. 85, 1103. Ukena, A., and A. Shimizu (2004), Phys. Rev. A 69 (2), Wollman, E. E., C. U. Lei, A. J. Weinstein, J. Suh, A. Kron- 022301. wald, F. Marquardt, A. A. Clerk, and K. C. Schwab (2015), Ukena, A., and A. Shimizu (2005), quant-ph/0505057 . Science 349 (6251), 952. Vahlbruch, H., M. Mehmet, K. Danzmann, and R. Schnabel Wootters, W. K. (1981), Phys. Rev. D 23 (2), 357. (2016), Phys. Rev. Lett. 117, 110801. Yadin, B., J. Ma, D. Girolami, M. Gu, and V. Vedral (2016), Vasilakis, G., H. Shen, K. Jensen, M. Balabas, D. Salart, Phys. Rev. X 6 (4), 041028. B. Chen, and E. S. Polzik (2015), Nat Phys 11 (5), 389. Yadin, B., and V. Vedral (2015), Phys. Rev. A 92, 022356. Vlastakis, B., G. Kirchmair, Z. Leghtas, S. E. Nigg, L. Frun- Yadin, B., and V. Vedral (2016), Phys. Rev. A 93, 022122. zio, S. M. Girvin, M. Mirrahimi, M. H. Devoret, and R. J. Yao, X.-C., T.-X. Wang, P. Xu, H. Lu, G.-S. Pan, X.-H. Bao, Schoelkopf (2013), Science 342 (6158), 607. C.-Z. Peng, C.-Y. Lu, Y.-A. Chen, and J.-W. Pan (2012), Vogel, W., and D.-G. Welsch (2006), Quantum optics (John Nat Photon 6 (4), 225. Wiley & Sons). Yin, Z.-Q., A. A. Geraci, and T. Li (2013), International Volkoff, T. J. (2015), J. Exp. Theor. Phys. 121 (5), 770. Journal of Modern Physics B 27 (26), 1330018. Volkoff, T. J., and K. B. Whaley (2014a), Phys. Rev. A Yu, S. (2013), arXiv:1302.5311 . 90 (6), 062122. Yuge, T. (2017), arXiv:1705.08117 [cond-mat, physics:quant- Volkoff, T. J., and K. B. Whaley (2014b), Phys. Rev. A ph] . 89 (1), 012122. Yukawa, M., K. Miyata, T. Mizuta, H. Yonezawa, P. Marek, Minář, J. c. v., P. Sekatski, and N. Sangouard (2016), Phys. R. Filip, and A. Furusawa (2013), Opt. Express 21 (5), Rev. A 94, 062111. 5529. van der Wal, C. H., A. C. J. ter Haar, F. K. Wilhelm, R. N. Yurke, B., and D. Stoler (1986), Phys. Rev. Lett. 57, 13. Schouten, C. J. P. M. Harmans, T. P. Orlando, S. Lloyd, Zhou, X., I. Dotsenko, B. Peaudecerf, T. Rybarczyk, and J. E. Mooij (2000), Science 290 (5492), 773. C. Sayrin, S. Gleyzes, J. M. Raimond, M. Brune, and Wang, C., Y. Y. Gao, P. Reinhold, R. W. Heeres, N. Ofek, S. Haroche (2012), Phys. Rev. Lett. 108, 243602. K. Chou, C. Axline, M. Reagor, J. Blumoff, K. M. Sliwa, Zurek, W. H. (2003a), arXiv:quant-ph/0306072 . Zurek, W. H. (2003b), Rev. Mod. Phys. 75, 715.