This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Theory-independent realism

D. M. Fucci [email protected]    R. M. Angelo [email protected] Department of Physics, Federal University of Paraná, P.O. Box 19044, 81531-980 Curitiba, Paraná, Brazil
Abstract

The distinctive features of quantum mechanics, which set it apart from other physical theories, challenge our notions of realism. Recovering realism from purely philosophical grounds, a quantitative and operational criterion was proposed in the past, but solely for the context of quantum mechanics. We use a framework of generalized probabilistic theories to expand the notion of realism for a theory-independent context, providing a criterion uniquely based on the probabilities assigned to measurement outcomes. More so, using robustness and the Kullback-Leibler divergence, we propose quantifiers for the realism of arbitrary physical properties given a particular state of a generic physical theory. These theory-independent quantifiers are then employed in quantum mechanics and we investigate their relation with another well-established irrealism measure.

preprint: APS/123-QED

I Introduction

Scientific knowledge is assumed to go beyond mere appearances and to convey reliable information about that which our senses cannot directly apprehend. In Physics, a large fraction of our scientific knowledge stands upon quantum theory, which remains coherent with the data acquired for more than a century. However, unlike classical mechanics, whose mathematical formulation aligns with our everyday experiences, quantum theory is challenging to make sense of. To abstract a consistent story from its formalism, or to interpret it, we are required to give up at least one of the metaphysical assumptions founding our classical intuition. One of such assumptions is realism.

Broadly speaking, realism tells us about the definiteness of the physical properties of a system at any instant of time independently of observers. Its seminal definition was given by Einstein, Podolsky, and Rosen (EPR) in 1935 [1], without which the first hint of entanglement appearing in the literature would not have been possible. There, it refers to the capability of predicting with certainty the value assigned to a physical property without disturbing the system, such that, whenever this is the case, it is said that there is an element of reality assigned to that physical property. Conceiving an experiment where measurements are performed on two parts of an entangled system, with these measurements being events located outside of each other’s light cones, EPR argue that incompatible observables should be simultaneous elements of reality. The authors then concluded that this would constitute a proof of the incompleteness of quantum mechanics. Put under scrutiny by Bell in 1964, this claim was discredited in Ref. [2], where it was shown that the predictions made by quantum mechanics conflicted with any theory of local hidden variables. Quantum systems that violate Bell’s inequalities, now understood to violate the hypothesis of local causality [3], have been experimentally verified through loophole-free tests [4, 5, 6, 7, 8, 9].

Even though it is possible to envision fully realist interpretations of quantum mechanics at the expense of other metaphysical requirements for the theory, such as in Bohmian mechanics [10], which sacrifices locality, the concept of realism underpins the debate around entanglement or Bell nonlocality. Later developments of the concept of realism, such as that proposed by Fine in Ref. [11], refer to joint probability distributions assigned to the measurements of different observables. The probabilistic profile of the joint measurability of non-commuting observables contrasts with this definition, highlighting a clash between incompatibility and realism. Furthermore, the measurement problem, or what makes a measurement a measurement [12], is an issue that invites us to answer how realism, displayed in classical phenomena, emerges.

Newer proposals for realism criteria, such as the one devised by Bilobran and Angelo (BA) in Ref. [13], make the concept more tangible. BA, in particular, do this by providing a quantifiable and operational criterion that was later employed in the development of a concept of nonlocality [14, 15, 16, 17] and axiomatized [18]. Further developments applied this notion to quantum resources theory [19] and foundational elements in quantum mechanics [20, 21, 22, 23, 24, 25, 26, 27]. However, every proposal in the physics literature for a realism criterion pertains to the context of quantum mechanics. A definition for realism which is theory-independent is lacking, and this work aims at filling this gap.

Besides deepening our understanding of realism itself, a theory-independent approach allows us to better understand not only quantum mechanics per se but also the relation between quantum mechanics and other physical theories. More so, it equips us with a framework suitable to scrutinize physical theories that are still to be developed, once quantum mechanics may not be the fundamental theory of Nature. Work [28] gives us a clear example of a theory-independent proposal shedding light on the concept of macrorealism [29].

A concise and meaningful criterion for realism, fully encoded in a mathematical syllogism, that is independent of any particular theory but still pertains to the realm of physics, is introduced in this work through generalized probabilistic theories (GPT). The GPT framework is constructed as a mathematical structure that describes generic states, transformations, and measurements. In this framework, any physical theory can be viewed as a particular GPT defined by a set of rules that assign probabilities to outcomes for the measurements of physical properties. Our criterion relies on the probabilities themselves, rather than on the rules that determine them. Notwithstanding, we go further and, inspired by the works of BA, we propose two ways for quantifying, in a non-realist theory, the extent to which a state deviates from our criterion of realism. These sought-after measures are obtained using the concepts of robustness and the Kullback-Leibler divergence [30], and they constitute two theory-independent quantifiers of irrealism.

The article is structured as follows. In Section II we review preliminary concepts, summarizing BA’s criterion of realism and skimming through some building blocks of the GPT framework, presenting only the minimal amount necessary for the understanding of the following sections. Our criterion is then laid out in Section III, and in Section IV, the theory-independent quantifiers of irrealism are presented along with some numerical analysis. They are tested in quantum mechanics and contrasted with BA’s notion of irrealism. Section V closes the paper with our conclusions.

II Preliminary Concepts

II.1 BA’s criterion of realism

Put forward in Ref. [13], this is an operational criterion based on the premise that measurements are not able to change the epistemic description of an already installed reality. A protocol is devised in which a source prepares two sets of a very large number of identical copies of a bipartite quantum system, with one of the sets being intercepted by an agent performing an unrevealed measurement of an observable A=aaAaA=\sum_{a}aA_{a}, acting on partition 𝒜\mathcal{A}, over each copy in the set. Here, AaA_{a} represents the projectors of observable AA with assigned outcomes aa. Quantum state tomography performed at the end will result in ΦA(ρ)\Phi_{A}(\rho) for the intercepted ensemble and ρ\rho for the non-intercepted one. If these two descriptions are indistinguishable, then AA is termed an element of reality. Mathematically, BA’s criterion of realism is expressed as

ρ=ΦA(ρ)\rho=\Phi_{A}(\rho) (1)

with

ΦA(ρ)a(Aa𝟙)ρ(Aa𝟙)=apaAaρ|a\Phi_{A}(\rho)\coloneqq\sum_{a}(A_{a}\otimes\mathbbm{1}_{\mathcal{B}})\,\rho\,(A_{a}\otimes\mathbbm{1}_{\mathcal{B}})=\sum_{a}p_{a}A_{a}\otimes\rho_{\mathcal{B}|a} (2)

where \mathcal{B} is the second partition of each system, pap_{a} is the probability associated to the outcome aa, and ρ|a\rho_{\mathcal{B}|a} is the post-measurement state for the partition \mathcal{B}. A deviance measure from the criterion (1) is then defined employing relative entropy:

A(ρ)S(ρ||ΦA(ρ))=S(ΦA(ρ))S(ρ),\mathfrak{I}_{A}(\rho)\coloneqq S(\rho||\Phi_{A}(\rho))=S(\Phi_{A}(\rho))-S(\rho), (3)

where S(ρ)=Tr(ρlogρ)S(\rho)=-\operatorname{Tr}(\rho\log{\rho}) stands for the von Neumann entropy of ρ\rho and S(ρ||ϱ)=Tr[ρ(logρlogϱ)]S(\rho||\varrho)=\operatorname{Tr}\big{[}\rho(\log{\rho}-\log{\varrho})\big{]} for the quantum relative entropy of ρ\rho and ϱ\varrho. The measure A(ρ)\mathfrak{I}_{A}(\rho) reads as the irreality of AA given ρ\rho. By now, irreality is a widely explored concept, supported by numerous investigations, both theoretical [14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25] and experimental [26, 27].

II.2 Generalized probabilistic theories

A general framework that has shown to be useful in dealing with generic physical theories, thus allowing a more solid background for talking about theory-independent properties, is called generalized probabilistic theories. It stands on three main building blocks:

  • a)

    State spaces generalize the concept of the set of all possible density operators in quantum mechanics. They may be understood as a set of equivalence classes of preparation procedures, a list of instructions to be followed wielding a particular physical state. Mathematically, it is expressed as a convex subset of a real finite-dimensional vector space.

  • b)

    Effect algebras map states inhabiting a state space into probabilities, generalizing what in quantum mechanics are the positive operator valued measures (POVMs). They are a set of equivalence classes of “yes” or “no” questions that may be made about a physical state in an experimental setting. So, let 𝒦\mathscr{K} be a state space and EE an effect algebra acting over 𝒦\mathscr{K}. E(𝒦)E(\mathscr{K}) is the set of all affine functions such that f:𝒦[0,1]f\colon\mathscr{K}\to[0,1].

  • c)

    Channels generalize quantum channels, transforming physical states from one system into physical states of another. For state spaces 𝒦A\mathscr{K}_{A} and 𝒦B\mathscr{K}_{B}, a channel Φ\Phi is an affine map Φ:𝒦A𝒦B\Phi\colon\mathscr{K}_{A}\to\mathscr{K}_{B}.

Together with a composition rule, corresponding to the tensor product in quantum mechanics, which describes the formation of joint state spaces, those three elements give a mathematical framework capable of representing the operational features of arbitrary physical theories. A thorough introduction on the subject is given in Ref. [31].

III Theory-Independent Realism

In a theory where a given observable is real (has a definite value regardless of observation), the role of a measurement is merely to reveal the already established element of reality. If the information about the outcome of a performed measurement is lost or somehow inaccessible, then the state of knowledge remains the same as before the measurement. In other words, unrevealed measurements are innocuous in realist theories. This will be the fundamental premise behind the criterion we now propose.

To implement this principle, consider a generic state ϵ\epsilon of a state space 𝒦\mathscr{K} with dimension dd, and physical properties 𝒳={xi}i=1d\mathcal{X}=\{x_{i}\}_{i=1}^{d} and 𝒴={yj}j=1d\mathcal{Y}=\{y_{j}\}_{j=1}^{d} defined by the set of their possible outcomes. The criterion of realism requires that the probabilities assigned to any outcome xix_{i} upon the measurement of any property 𝒳\mathcal{X} on a state ϵ\epsilon remains unchanged if an unrevealed measurement of 𝒴\mathcal{Y} was performed beforehand. Symbolically, if

pϵ(xi)=pΦ𝒴(ϵ)(xi)(𝒳={xi}i=1d),p_{\epsilon}(x_{i})=p_{\Phi_{\mathcal{Y}}(\epsilon)}(x_{i})\qquad\left(\forall\,\mathcal{X}=\{x_{i}\}_{i=1}^{d}\right), (4)

then 𝒴\mathcal{Y} is termed an element of reality given the state ϵ\epsilon. In the right-hand term, sub-index Φ𝒴(ϵ)\Phi_{\mathcal{Y}}(\epsilon) stands for a post unrevealed measurement of 𝒴\mathcal{Y} over state ϵ\epsilon. The term is expressed via conditional probabilities as

pΦ𝒴(ϵ)(xi)jpϵ(xi|yj)pϵ(yj).p_{\Phi_{\mathcal{Y}}(\epsilon)}(x_{i})\coloneqq\sum_{j}p_{\epsilon}(x_{i}|y_{j})\,p_{\epsilon}(y_{j}). (5)

Note that the knowing of yjy_{j} may alter the preparation ϵ\epsilon. The notation for conditional probabilities denotes that ϵ\epsilon was the original state before the knowledge yielded by a precise measurement of 𝒴\mathcal{Y}. Criterion (4) gives the core of this work. It does not demand every physical state ϵ\epsilon to satisfy (4). When that is the case, that physical theory is considered realist. If only a set of states ϵ\epsilon meets (4), those states are said to be 𝒴\mathcal{Y}-realist.

Before moving on, it is opportune to note that our realism criterion can be validated through the assumption that joint probability distributions p(xi,yj)p(x_{i},y_{j}) exist for all 𝒳\mathcal{X}, which makes contact with Fine’s approach to determinism [11]. If p(xi,yj)p(x_{i},y_{j}) exists without any concern a priori with the order with which 𝒳\mathcal{X} and 𝒴\mathcal{Y} are probed, and if the definition of conditional probability holds, then one can write p(xi,yj)=pϵ(xi|yj)pϵ(yj)=pϵ(yj|xi)pϵ(xi)p(x_{i},y_{j})=p_{\epsilon}(x_{i}|y_{j})\,p_{\epsilon}(y_{j})=p_{\epsilon}(y_{j}|x_{i})\,p_{\epsilon}(x_{i}) (Bayes’ rule). Since jpϵ(yj|xi)=1\sum_{j}p_{\epsilon}(y_{j}|x_{i})=1, substituting Bayes’ rule into the definition (5) immediately retrieves the realism criterion (4).

Let 𝒞\mathscr{C} be the state space where generic ϵ\epsilon reside and 𝒞𝒴\mathscr{C}_{\mathcal{Y}} the state space for 𝒴\mathcal{Y}-realist states. While 𝒞\mathscr{C} is convex by construction, the convexity of 𝒞𝒴\mathscr{C}_{\mathcal{Y}} remains to be proved. Considering the one-parameter state ϵλ=(1λ)ϵ1+λϵ2\epsilon^{\prime}_{\lambda}=(1-\lambda)\epsilon^{\prime}_{1}+\lambda\epsilon^{\prime}_{2}, with λ[0,1]\lambda\in[0,1] and ϵ1,ϵ2𝒞𝒴{\epsilon^{\prime}_{1},\epsilon^{\prime}_{2}}\in\mathscr{C}_{\mathcal{Y}}, the proof is complete if ϵλ𝒞𝒴\epsilon^{\prime}_{\lambda}\in\mathscr{C}_{\mathcal{Y}}. Suppose a binary random variable Λ\Lambda with assigned probabilities (1λ)(1-\lambda) and λ\lambda. A rule may be given such that, depending on the outcome of Λ\Lambda, 𝒳\mathcal{X} will be measured in either ϵ1\epsilon^{\prime}_{1} or ϵ2\epsilon^{\prime}_{2}. The probabilistic profile for this scenario is represented by pϵλ(xi)=p(1λ)ϵ1+λϵ2(xi)p_{\epsilon^{\prime}_{\lambda}}(x_{i})=p_{(1-\lambda)\epsilon^{\prime}_{1}+\lambda\epsilon^{\prime}_{2}}(x_{i}). Due to the 𝒴\mathcal{Y}-realism status of ϵ1,2\epsilon^{\prime}_{1,2},

p(1λ)ϵ1+λϵ2(xi)=p(1λ)Φ𝒴(ϵ1)+λΦ𝒴(ϵ2)(xi).p_{(1-\lambda)\epsilon^{\prime}_{1}+\lambda\epsilon^{\prime}_{2}}(x_{i})=p_{(1-\lambda)\Phi_{\mathcal{Y}}(\epsilon^{\prime}_{1})+\lambda\Phi_{\mathcal{Y}}(\epsilon^{\prime}_{2})}(x_{i}). (6)

However, the right-hand side of this equation can also represent a preparation and mixing of two ensembles of ϵ1\epsilon^{\prime}_{1} and ϵ2\epsilon^{\prime}_{2} with relative populations (1λ)(1-\lambda) and λ\lambda, respectively, with the measurement of 𝒴\mathcal{Y} first and 𝒳\mathcal{X} second from a randomly selected state from the mixed ensemble, ϵλ\epsilon^{\prime}_{\lambda}. That is,

p(1λ)Φ𝒴(ϵ1)+λΦ𝒴(ϵ2)(xi)=pΦ𝒴(ϵλ)(xi).p_{(1-\lambda)\Phi_{\mathcal{Y}}(\epsilon^{\prime}_{1})+\lambda\Phi_{\mathcal{Y}}(\epsilon^{\prime}_{2})}(x_{i})=p_{\Phi_{\mathcal{Y}}(\epsilon^{\prime}_{\lambda})}(x_{i}). (7)

This last equality, which mirrors the fact that Φ\Phi is an affine map, can be connected to the previous equations to complete the proof.

Classical mechanics gives the simplest case for realist theories. The state space 𝒞\mathscr{C} becomes a 2n2n-dimensional phase space, where nn represents the dimensions of position and momentum. A state ϵ\epsilon in this space is one point that characterizes a configuration of a physical system. The realism of the framework is implied by the space of 𝒴\mathcal{Y}-realist states being equal to the total phase space, 𝒞=𝒞𝒴\mathscr{C}=\mathscr{C}_{\mathcal{Y}} for every 𝒴\mathcal{Y}. In classical statistical mechanics—that is, classical mechanics supplemented with subjective uncertainties and Liouville’s equation—the diagnosis is no different. For every generalized coordinate qiq_{i} and canonical momentum πj\pi_{j}, a joint probability distribution 𝒫(qi,πj)dqidπj\mathcal{P}(q_{i},\pi_{j})\,dq_{i}d\pi_{j} can be found, which, as discussed above, validates the realism criterion (4).

In the mathematical framework of quantum mechanics (QM), the left and right-hand sides of Eq. (4) become:

pϵ(xi)\displaystyle p_{\epsilon}(x_{i}) QMTr[Xiρ],\displaystyle\xrightarrow{\text{QM}}\operatorname{Tr}[X_{i}\rho], (8)
pΦY(ϵ)(xi)\displaystyle p_{\Phi_{Y}(\epsilon)}(x_{i}) QMTr[XiΦY(ρ)],\displaystyle\xrightarrow{\text{\scriptsize QM}}\operatorname{Tr}[X_{i}\Phi_{Y}(\rho)], (9)

with ΦY\Phi_{Y} defined as in Eq. (2). A state ϵ\epsilon is represented by a density operator ρ\rho and a physical property 𝒳\mathcal{X} by an observable X=ixiXiX=\sum_{i}x_{i}X_{i}, being XiX_{i} projectors corresponding to measurement outcomes xix_{i}. In this context, criterion (4) reads Tr[Xiρ]=Tr[XiΦY(ρ)]\operatorname{Tr}[X_{i}\rho]=\operatorname{Tr}[X_{i}\Phi_{Y}(\rho)] and, because of the linearity and cyclic properties of the trace, this equation can be written as:

Tr[Xi(ΦY(ρ)ρ)]=0.\operatorname{Tr}[X_{i}(\Phi_{Y}(\rho)-\rho)]=0. (10)

A sufficient condition for this equation to hold is for ρ\rho to be a YY-realist state according to criterion (1). The fact that this condition is also necessary is proven in Appendix A.

IV Quantifying irrealism

If two generic states in a physical theory fall outside 𝒞𝒴\mathscr{C}_{\mathcal{Y}}, thus violating criterion (4) for a physical property 𝒴\mathcal{Y}, it should be possible to quantify and compare the degrees of violation of these states. Thus, we propose two different methodologies for quantifying irrealism, the complement of realism.

IV.1 Robustness of Irrealism

Here, we employ the usual approach of robustness under state perturbation. For a state ϵ\epsilon under scrutiny, we introduce another state ϵ𝒞\epsilon^{\prime}\in\mathscr{C} in a convex sum weighted by a factor η\eta, yielding a state ϵη=(1η)ϵ+ηϵ\epsilon_{\eta}=(1-\eta)\epsilon+\eta\epsilon^{\prime}. We ask, what is the minimal η\eta such that ϵη𝒞𝒴\epsilon_{\eta}\in\mathscr{C}_{\mathcal{Y}}? Hereafter referred to as the robustness of irrealism of state ϵ\epsilon regarding the quantity 𝒴\mathcal{Y}, the optimal value of η\eta that answers this question is mathematically formulated as

𝒴(ϵ)minϵ{η[0,1]|(1η)ϵ+ηϵ𝒞𝒴}.\mathcal{R}_{\mathcal{Y}}(\epsilon)\coloneqq\min_{\epsilon^{\prime}}\big{\{}\eta\in[0,1]\;\big{|}\;(1-\eta)\epsilon+\eta\epsilon^{\prime}\in\mathscr{C}_{\mathcal{Y}}\big{\}}. (11)

We resort to the robustness 𝒴(ϵ)\mathcal{R}_{\mathcal{Y}}(\epsilon) because it takes advantage of the convex structure of the spaces 𝒞\mathscr{C} and 𝒞𝒴\mathscr{C}_{\mathcal{Y}} without relying on a particular definition of a metric for such spaces. Moreover, it satisfies desirable conditions for an irrealism quantifier. Specifically, 𝒴(ϵ)0\mathcal{R}_{\mathcal{Y}}(\epsilon)\geq 0, with equality holding if and only if 𝒴\mathcal{Y} is an element of reality, i.e., ϵ𝒞𝒴\epsilon\in\mathscr{C}_{\mathcal{Y}}. The requirement for ϵ𝒞\epsilon^{\prime}\in\mathscr{C} and not 𝒞𝒴\mathscr{C}_{\mathcal{Y}} is because if dim(𝒞𝒴)<dim(𝒞)\dim(\mathscr{C}_{\mathcal{Y}})<\dim(\mathscr{C}), then 𝒴(ϵ)\mathcal{R}_{\mathcal{Y}}(\epsilon) would always be zero, and this measure would not provide any relevant information about the irreality status of an ϵ\epsilon.

Refer to caption
Figure 1: Bloch sphere representation for robustness of irrealism of Sz^S_{\hat{z}} for a qubit. The longer vertical axis represents the set of states 𝒞Sz^\mathscr{C}_{S_{\hat{z}}}. The three colored arrows correspond to the vectors that parametrize the states in Eq. (11) after the minimization of η\eta over ρ\rho^{\prime}. For ρ=12(𝟙+ρ)σ\rho=\frac{1}{2}(\mathbbm{1}+\vec{\rho})\cdot\vec{\sigma}, given by ρ=(x,y,z)3\vec{\rho}=(x,y,z)\in\mathbbm{R}^{3}, the robustness Sz^(ρ)=η¯\mathcal{R}_{S_{\hat{z}}}(\rho)=\bar{\eta} [see Eq. (12)] is found for ρ=12(𝟙+ρσ)\rho^{\prime}=\frac{1}{2}(\mathbbm{1}+\vec{\rho}^{\prime}\cdot\vec{\sigma}) with ρ=(x,y,0)3\vec{\rho}^{\prime}=(-x,-y,0)\in\mathbbm{R}^{3}.

One can gain insight into the physical intuition behind the concept of robustness of irrealism by envisioning a simple physical scenario. Suppose you have a given state ϵ\epsilon and, at your disposal, every other state ϵ\epsilon^{\prime}. Robustness tells you the least amount η\eta you could add of a state ϵ\epsilon^{\prime} to ϵ\epsilon such that the resulting ensemble (1η)ϵ+ηϵ(1-\eta)\epsilon+\eta\epsilon^{\prime} is a realist state for a given physical property. That is, it quantifies the minimum amount of mixing required for the irrealism of a given state to be lost.

In quantum mechanics, for a single qubit, realism as defined in Eq. (1) reduces to coherence [13]. The same is true for the robustness of realism. Indeed, our findings for such a case are consistent with those reported in [32]. To show this, let us compute the robustness of irrealism of the spin observable in the z^\hat{z} direction, Sz^S_{\hat{z}}, for a preparation ρ\rho. For a minimization, as defined in Eq. (11), returning the optimal value η¯\bar{\eta}, we have ρη¯=(1η¯)ρ+η¯ρ\rho_{\bar{\eta}}=(1-\bar{\eta})\rho+\bar{\eta}\rho^{\prime}, where ρη¯\rho_{\bar{\eta}} is a Sz^S_{\hat{z}} realist state. As found in [32], the representation of such states in the Bloch sphere reveals a geometric pattern (see Fig. 1): being ρ=12(𝟙+ρσ)\rho=\frac{1}{2}(\mathbbm{1}+\vec{\rho}\cdot\vec{\sigma}) with ρ=(x,y,z)3\vec{\rho}=(x,y,z)\in\mathbbm{R}^{3}, one finds ρ=12(𝟙+ρσ)\rho^{\prime}=\frac{1}{2}(\mathbbm{1}+\vec{\rho}^{\prime}\cdot\vec{\sigma}) with ρ=(x,y,0)3\vec{\rho}^{\prime}=(-x,-y,0)\in\mathbbm{R}^{3}, where σ\vec{\sigma} is the vector composed of Pauli matrices. Because η¯\bar{\eta} is numerically equivalent to the ratio between |ρη¯ρ||\vec{\rho}_{\bar{\eta}}-\vec{\rho}^{\prime}|, where ρη¯\vec{\rho}_{\bar{\eta}} is the vector representing ρη¯\rho_{\bar{\eta}}, and |ρρ||\vec{\rho}-\vec{\rho}^{\prime}|, it is a simple matter of geometry to derive the expression

Sz^(ρ)=η¯=|z^×ρ|1+|z^×ρ|,\mathcal{R}_{S_{\hat{z}}}(\rho)=\bar{\eta}=\frac{|\hat{z}\times\vec{\rho}|}{1+|\hat{z}\times\vec{\rho}|}, (12)

where |z^×ρ|=rsinθ|\hat{z}\times\vec{\rho}|=r\sin{\theta}, with rr being the radius and θ\theta the polar angle defining the spherical coordinates of the vector ρ\vec{\rho}. For a generic spin observable, Sn^S_{\hat{n}}, Eq. (12) still yields the robustness of irrealism under a basis rotation, aligning z^\hat{z} with n^\hat{n}.

Numerical comparisons between the irreality defined in Eq. (3) and the robustness of irrealism are noteworthy. Figure 2 shows the results of a case study of the irrealism of Sz^S_{\hat{z}} for a pure state of a qubit, where the polar angle θ\theta in the Bloch representation varies from 0 to π\pi. In this computation, we normalized the robustness of irrealism so that its maximum matches the irreality’s. We see that both criteria reach zero at the same points, the maximum value at the same point, and the parametric curve for both functions, displayed in Fig. 3, confirms a monotonic relation.

Refer to caption
Figure 2: Plot of Sz^(ρ)\mathfrak{I}_{S_{\hat{z}}}(\rho), Sz^(ρ)\mathcal{R}_{S_{\hat{z}}}(\rho), and Sz^(ρ)\mathcal{I}_{S_{\hat{z}}}(\rho) for a pure qubit state with its representing polar angle θ\theta in [0,π][0,\pi].

IV.2 Divergence of Realism

Refer to caption
Figure 3: Three parametric plots for the same case of a pure qubit state, with the polar angle θ\theta ranging from 0 to π\pi in the Bloch sphere representation. The top plot shows Sz^(ρ)\mathfrak{I}_{S_{\hat{z}}}(\rho) against Sz^(ρ)\mathcal{R}_{S_{\hat{z}}}(\rho), the middle plot shows Sz^(ρ)\mathfrak{I}_{S_{\hat{z}}}(\rho) against Sz^(ρ)\mathcal{I}_{S_{\hat{z}}}(\rho), and the bottom plot shows Sz^(ρ)\mathcal{R}_{S_{\hat{z}}}(\rho) against Sz^(ρ)\mathcal{I}_{S_{\hat{z}}}(\rho).

Definition (3) hints in another direction by which a theory-independent irrealism quantification may be conceived. We can simply take a divergence between the left and right-hand terms appearing in the criterion (4), and, for this purpose, the Kullback-Leibler divergence [30] provides us with a standard framework. Because the criterion is defined for any physical property 𝒳\mathcal{X}, the violation of realism is given by the property that wields the maximal divergence between the probability distributions. This leads us to the following quantifier for the irreality of 𝒴\mathcal{Y} given a physical state ϵ\epsilon:

𝒴(ϵ)=max𝒳𝒟(Pϵ𝒳PΦ𝒴(ϵ)𝒳).\mathcal{I_{Y}}(\epsilon)=\max_{\mathcal{X}}\,\mathcal{D}\left(P^{\mathcal{X}}_{\epsilon}\Big{\|}P^{\mathcal{X}}_{\Phi_{\mathcal{Y}}(\epsilon)}\right). (13)

Here, 𝒴(ϵ)\mathcal{I_{Y}}(\epsilon) reads as the divergence of realism of 𝒴\mathcal{Y} for the state ϵ\epsilon, with the Kullback-Leibler divergence between probability distributions Pϵ𝒳={pϵ(xi)}i=1dP_{\epsilon}^{\mathcal{X}}=\{p_{\epsilon}(x_{i})\}_{i=1}^{d} and Qϵ𝒳={qϵ(xi)}i=1dQ_{\epsilon}^{\mathcal{X}}=\{q_{\epsilon}(x_{i})\}_{i=1}^{d} computed as 𝒟(Pϵ𝒳||Qϵ𝒳)=ipϵ(xi)ln[pϵ(xi)/qϵ(xi)]\mathcal{D}(P_{\epsilon}^{\mathcal{X}}||Q_{\epsilon}^{\mathcal{X}})=\sum_{i}p_{\epsilon}(x_{i})\ln\big{[}p_{\epsilon}(x_{i})/q_{\epsilon}(x_{i})\big{]}. It follows from the definition of the Kullback-Leibler divergence that 𝒴(ϵ)\mathcal{I_{Y}}(\epsilon) is always non-negative and vanishes iff the two probability distributions are the same, that is, if the criterion (4) is met or, equivalently, ϵ𝒞𝒴\epsilon\in\mathscr{C}_{\mathcal{Y}}.

While we appealed to the convex geometric structure of the state spaces and to physical intuition to derive the robustness of irrealism, the divergence of realism relies solely on classical probability theory. Specifically for quantum mechanics, it is shown in Appendix B that

Y(ρ)=maxXS(ΦX(ρ)||ΦX,Y(ρ)),\mathcal{I}_{Y}(\rho)=\max_{X}\,S\big{(}\Phi_{X}(\rho)||\Phi_{X,Y}(\rho)\big{)}, (14)

where ΦX,Y(ρ)=ΦX(ΦY(ρ))\Phi_{X,Y}(\rho)=\Phi_{X}\big{(}\Phi_{Y}(\rho)\big{)}, for observables XX and YY. Equation (14), for the case of qubits, peaks at the same point state as Eq. (3), just as the robustness of irreality. A case study employing the same states as in the preceding subsection was conducted, and the findings are illustrated in Figs. 2 and 3. It is worthwhile noting that the apparent plateau in Sz^(ρ)\mathcal{I}_{S_{\hat{z}}}(\rho) in Fig. 2 is not a plateau. Instead, it is a region where the absolute value of the slope of the curve continuously approaches zero, with zero occurring only at the peak, in a gradually decreasing fashion. The monotonic relations found in Fig. 3 among the irrealism quantifiers Sz^(ρ)\mathfrak{I}_{S_{\hat{z}}}(\rho), Sz^(ρ)\mathcal{R}_{S_{\hat{z}}}(\rho), and Sz^(ρ)\mathcal{I}_{S_{\hat{z}}}(\rho) suggest that these measures may be qualitatively equivalent.

V Conclusion

Despite a rich body of literature on violations of hypotheses such as locality, causality, or even macrorealism [29] or local-friendliness [33], realism is a notion that sparks debates over its very definition. Additionally, quantifiable notions of realism transcending the scope of quantum theory are still lacking. This work fills this gap by bringing together different criteria and aiming to provide material for the emergence of a consensus on the subject.

By only requiring the probabilities assigned to the outcomes of measurements of physical properties in a physical system, we have crafted a theory-independent realism criterion, Eq. (4). It presents itself as a legitimate generalization of BA’s criterion and also Fine’s criterion, and it retrieves the results expected for both quantum and classical mechanics. Deviance measures for this criterion, the so-called irrealism measures, were developed in two independent ways, making use of the notion of robustness and the divergence of Kullback-Leibler. Although the robustness of irrealism [Eq. (11)] and the divergence of realism [Eq. (13)] are not conceptually equivalent, neither to each other nor to BA’s irreality measure, our case study highlighted a similar behavior for the three quantities. This provides evidence that both the robustness of irrealism and the divergence of realism are candidates for bona fide irreality quantifiers.

The numerical analysis of the robustness of irrealism in a qubit system, as presented in this work, is computationally feasible. However, calculations for higher-dimensional systems require the development of sophisticated computational strategies. We expect future research efforts in this direction to yield insights. Concurrently, our ongoing research delves into the interplay between realism and context incompatibility, particularly from a theory-independent perspective. Finally, this work paves the way for a deeper understanding of the emergence of classical realism, and we expect a revisit of Wigner’s friend scenarios under this light to bear fruit.

Acknowledgements.
This research was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior, Brasil (CAPES), Finance Code 001. D.M.F. thanks Eduardo Hoefel and Ana C. S. Costa for insightful conversations. R.M.A. thanks the financial support from the National Institute for Science and Technology of Quantum Information (CNPq, INCT-IQ 465469/2014-0) and the Brazilian funding agency CNPq under Grant No. 305957/2023-6.

Appendix A

Our intent here is to show that Eq. (3) is a necessary and sufficient condition for Eq. (10). It is an established result, shown in Ref. [31], that there is a duality between the state space and the effect algebra. In other words, given a state space 𝒦\mathscr{K}, it is possible to construct the effect algebra E(𝒦)E(\mathscr{K}) and vice versa. The proof shown below is a particular case of this result, relying exclusively on the formalism of quantum mechanics.

It is evident that if Eq. (1) holds, Eq. (10) will hold as well. It is not immediately clear, however, that this is the only case where Eq. (10) is met, and this is what we prove. To this end, we first introduce some aspects of the formalism of the generalized Bloch representation. A more complete yet concise exposition is given in sources [34] and [35].

A basis for the linear operators acting on the state space is given by the set of matrices {𝟙,Λ1,,Λd21}\left\{\mathbbm{1},\Lambda_{1},\ldots,\Lambda_{d^{2}-1}\right\}, which are the generators of the special unitary group of degree dd, SU(dd). The Λi\Lambda_{i} are complex d×dd\times d self-adjoint orthogonal traceless matrices. Together with the normalization Tr(ΛiΛj)=2δij\operatorname{Tr}(\Lambda_{i}\Lambda_{j})=2\delta_{ij}, a generic quantum state can be expressed as

ρr=1d(𝟙+CdrΛ),\rho_{\vec{r}}=\frac{1}{d}\left(\mathbbm{1}+C_{d}\vec{r}\cdot\vec{\Lambda}\right), (15)

where Cd=d(d1)/2C_{d}=\sqrt{d(d-1)/2} and, with an orthonormal basis in d21\mathbbm{R}^{d^{2}-1} denoted by {e^i}i=1d21\left\{\hat{e}_{i}\right\}_{i=1}^{d^{2}-1}, r=i=1d21rie^i\vec{r}=\sum_{i=1}^{d^{2}-1}r_{i}\hat{e}_{i} is a vector and Λ=i=1d21Λie^i\vec{\Lambda}=\sum_{i=1}^{d^{2}-1}\Lambda_{i}\hat{e}_{i} is a vector with matrices as components. This allows for any quantum state to be represented as a real vector r\vec{r} in a real ball B(d21)B(\mathbbm{R}^{d^{2}-1}) of dimension d21d^{2}-1. Projective operators, characterized by jXj=𝟙\sum_{j}X_{j}=\mathbbm{1} and Tr(XiXj)=δij\operatorname{Tr}\left(X_{i}X_{j}\right)=\delta_{ij} can be written as

Xi=1d(𝟙+CdxiΛ),X_{i}=\frac{1}{d}\left(\mathbbm{1}+C_{d}\vec{x}_{i}\cdot\vec{\Lambda}\right), (16)

where ixi=0\sum_{i}\vec{x}_{i}=\vec{0} and xixj=(δijd1)/(d1)\vec{x}_{i}\cdot\vec{x}_{j}=(\delta_{ij}d-1)/(d-1). Being xix_{i} the eigenvalues of a traceless observable XX, one can write X=ixiXi=xΛX=\sum_{i}x_{i}X_{i}=\vec{x}\cdot\vec{\Lambda} with x=(Cd/d)ixixi\vec{x}=(C_{d}/d)\sum_{i}x_{i}\vec{x}_{i}. With a similar construction for Y=yΛY=\vec{y}\cdot\vec{\Lambda}, where y=(Cd/d)jyjyj\vec{y}=(C_{d}/d)\sum_{j}y_{j}\vec{y}_{j}, it can be shown that [36]

ΦY(ρr)=1d(𝟙+CduΛ),u=d1dj=1d(yjr)yj,\Phi_{Y}(\rho_{\vec{r}})=\frac{1}{d}(\mathbbm{1}+C_{d}\vec{u}\cdot\vec{\Lambda}),\quad\vec{u}=\frac{d-1}{d}\sum_{j=1}^{d}(\vec{y}_{j}\cdot\vec{r})\vec{y}_{j}, (17)

Substituting Eqs.(15), (16), and (17) into Eq.(10), multiplying the resulting equation by xix_{i} and summing over ii yields

Tr[(xΛ)(vΛ)]=0\operatorname{Tr}[(\vec{x}\cdot\vec{\Lambda})(\vec{v}\cdot\vec{\Lambda})]=0 (18)

with vru\vec{v}\coloneqq\vec{r}-\vec{u}. Now, using the identity Tr[(r1Λ)(r2Λ)]=2(r1r2)\operatorname{Tr}[(\vec{r}_{1}\cdot\vec{\Lambda})(\vec{r}_{2}\cdot\vec{\Lambda})]=2(\vec{r}_{1}\cdot\vec{r}_{2}), the last equation becomes simply

xv=0.\vec{x}\cdot\vec{v}=0. (19)

We complete the proof by observing that this equation should hold for every possible observable XX, characterized by x\vec{x}, which is only possible if v=0\vec{v}=\vec{0}, or, equivalently, r=u\vec{r}=\vec{u}, implying Eq. (1).

Appendix B

Here, we demonstrate that, within the context of quantum mechanics, Eq. (13) can be reformulated as shown in Eq. (14). To this end, we employ the relation f(A)|a=f(a)|af(A)\ket{a}=f(a)\ket{a}, where ff is a generic function and AA is a Hermitian operator with orthonormal eigenbasis |a\ket{a} and eigenvalues aa. Now, we note that ΦX(ρ)=ipρ(xi)Xi\Phi_{X}(\rho)=\sum_{i}p_{\rho}(x_{i})X_{i} and ΦX,Y(ρ)=ipΦY(ρ)(xi)Xi\Phi_{X,Y}(\rho)=\sum_{i}p_{\Phi_{Y}(\rho)}(x_{i})X_{i}, implying that these post-measurement states commute and therefore share the same set of eigenvectors. Then,

S(ΦX(ρ)||ΦX,Y(ρ))\displaystyle S\big{(}\Phi_{X}(\rho)||\Phi_{X,Y}(\rho)\big{)}
=Tr[ΦX(ρ)logΦX(ρ)]Tr[ΦX(ρ)logΦX,Y(ρ)]\displaystyle=\operatorname{Tr}\big{[}\Phi_{X}(\rho)\log\Phi_{X}(\rho)\big{]}-\operatorname{Tr}[\Phi_{X}(\rho)\log\Phi_{X,Y}(\rho)] (20)
=ipρ(xi)logpρ(xi)ipρ(xi)logpΦY(ρ)(xi)\displaystyle=\sum_{i}p_{\rho}(x_{i})\log p_{\rho}(x_{i})-\sum_{i}p_{\rho}(x_{i})\log p_{\Phi_{Y}(\rho)}(x_{i}) (21)
=𝒟(PρXiPΦY(ρ)Xi).\displaystyle=\mathcal{D}\left(P^{X_{i}}_{\rho}\Big{\|}P^{X_{i}}_{\Phi_{Y}(\rho)}\right). (22)

Since the argument inside the maximization function is equal for both expressions, so is its result. This proof was originally presented in Ref. [37].

References

  • Einstein et al. [1935] A. Einstein, B. Podolsky, and N. Rosen, Can quantum-mechanical description of physical reality be considered complete?, Phys. Rev. 47, 777 (1935).
  • Bell [1964] J. S. Bell, On the Einstein Podolsky Rosen paradox, Physics Physique Fizika 1, 195 (1964).
  • Bell et al. [2001] M. Bell, K. Gottfried, and M. Veltman, John S Bell on the Foundations of Quantum Mechanics (World Scientific, 2001) pp. 50–60, https://www.worldscientific.com/doi/pdf/10.1142/4757 .
  • Hensen et al. [2015] B. Hensen, H. Bernien, A. E. Dréau, A. Reiserer, N. Kalb, M. S. Blok, J. Ruitenberg, R. F. Vermeulen, R. N. Schouten, C. Abellán, et al., Loophole-free Bell inequality violation using electron spins separated by 1.3 kilometres, Nature 526, 682 (2015).
  • Giustina et al. [2015] M. Giustina, M. A. M. Versteegh, S. Wengerowsky, J. Handsteiner, A. Hochrainer, K. Phelan, F. Steinlechner, J. Kofler, J.-A. Larsson, C. Abellán, W. Amaya, V. Pruneri, M. W. Mitchell, J. Beyer, T. Gerrits, A. E. Lita, L. K. Shalm, S. W. Nam, T. Scheidl, R. Ursin, B. Wittmann, and A. Zeilinger, Significant-loophole-free test of Bell’s theorem with entangled photons, Phys. Rev. Lett. 115, 250401 (2015).
  • Shalm et al. [2015] L. K. Shalm, E. Meyer-Scott, B. G. Christensen, P. Bierhorst, M. A. Wayne, M. J. Stevens, T. Gerrits, S. Glancy, D. R. Hamel, M. S. Allman, K. J. Coakley, S. D. Dyer, C. Hodge, A. E. Lita, V. B. Verma, C. Lambrocco, E. Tortorici, A. L. Migdall, Y. Zhang, D. R. Kumor, W. H. Farr, F. Marsili, M. D. Shaw, J. A. Stern, C. Abellán, W. Amaya, V. Pruneri, T. Jennewein, M. W. Mitchell, P. G. Kwiat, J. C. Bienfang, R. P. Mirin, E. Knill, and S. W. Nam, Strong loophole-free test of local realism, Phys. Rev. Lett. 115, 250402 (2015).
  • Hensen et al. [2016] B. Hensen, N. Kalb, M. Blok, A. Dréau, A. Reiserer, R. Vermeulen, R. Schouten, M. Markham, D. Twitchen, K. Goodenough, et al., Loophole-free Bell test using electron spins in diamond: second experiment and additional analysis, Scientific reports 6, 30289 (2016).
  • Rauch et al. [2018] D. Rauch, J. Handsteiner, A. Hochrainer, J. Gallicchio, A. S. Friedman, C. Leung, B. Liu, L. Bulla, S. Ecker, F. Steinlechner, R. Ursin, B. Hu, D. Leon, C. Benn, A. Ghedina, M. Cecconi, A. H. Guth, D. I. Kaiser, T. Scheidl, and A. Zeilinger, Cosmic Bell test using random measurement settings from high-redshift quasars, Phys. Rev. Lett. 121, 080403 (2018).
  • Li et al. [2018] M.-H. Li, C. Wu, Y. Zhang, W.-Z. Liu, B. Bai, Y. Liu, W. Zhang, Q. Zhao, H. Li, Z. Wang, L. You, W. J. Munro, J. Yin, J. Zhang, C.-Z. Peng, X. Ma, Q. Zhang, J. Fan, and J.-W. Pan, Test of local realism into the past without detection and locality loopholes, Phys. Rev. Lett. 121, 080404 (2018).
  • Bohm [1952] D. Bohm, A suggested interpretation of the quantum theory in terms of “hidden” variables. i, Phys. Rev. 85, 166 (1952).
  • Fine [1982] A. Fine, Hidden variables, joint probability, and the Bell inequalities, Phys. Rev. Lett. 48, 291 (1982).
  • Brukner [2017] Č. Brukner, On the quantum measurement problem, in Quantum [Un]Speakables II: Half a Century of Bell’s Theorem, edited by R. Bertlmann and A. Zeilinger (Springer International Publishing, Cham, 2017) pp. 95–117.
  • Bilobran and Angelo [2015] A. L. O. Bilobran and R. M. Angelo, A measure of physical reality, EPL 112, 40005 (2015).
  • Gomes and Angelo [2018] V. S. Gomes and R. M. Angelo, Nonanomalous measure of realism-based nonlocality, Phys. Rev. A 97, 012123 (2018).
  • Gomes and Angelo [2019] V. S. Gomes and R. M. Angelo, Resilience of realism-based nonlocality to local disturbance, Phys. Rev. A 99, 012109 (2019).
  • Orthey and Angelo [2019] A. C. Orthey and R. M. Angelo, Nonlocality, quantum correlations, and violations of classical realism in the dynamics of two noninteracting quantum walkers, Phys. Rev. A 100, 042110 (2019).
  • Fucci and Angelo [2019] D. M. Fucci and R. M. Angelo, Tripartite realism-based quantum nonlocality, Phys. Rev. A 100, 062101 (2019).
  • Orthey and Angelo [2022] A. C. Orthey and R. M. Angelo, Quantum realism: Axiomatization and quantification, Phys. Rev. A 105, 052218 (2022).
  • Costa and Angelo [2020] A. C. Costa and R. Angelo, Information-based approach towards a unified resource theory, Quantum Information Processing 19, 1 (2020).
  • Dieguez and Angelo [2018] P. R. Dieguez and R. M. Angelo, Information-reality complementarity: The role of measurements and quantum reference frames, Phys. Rev. A 97, 022107 (2018).
  • Engelbert and Angelo [2020] N. G. Engelbert and R. M. Angelo, Hardy’s paradox as a demonstration of quantum irrealism, Foundations of Physics 50, 105 (2020).
  • Łukasz Rudnicki [2018] Łukasz Rudnicki, Uncertainty-reality complementarity and entropic uncertainty relations, Journal of Physics A: Mathematical and Theoretical 51, 504001 (2018).
  • Savi and Angelo [2021] M. F. Savi and R. M. Angelo, Quantum resource covariance, Phys. Rev. A 103, 022220 (2021).
  • Engelbert and Angelo [2023] N. G. Engelbert and R. M. Angelo, Considerations on the relativity of quantum irrealism, Entropy 2510.3390/e25040603 (2023).
  • Paiva et al. [2023] I. L. Paiva, P. R. Dieguez, R. M. Angelo, and E. Cohen, Coherence and realism in the aharonov-bohm effect, Phys. Rev. A 107, 032213 (2023).
  • Mancino et al. [2018] L. Mancino, M. Sbroscia, E. Roccia, I. Gianani, V. Cimini, M. Paternostro, and M. Barbieri, Information-reality complementarity in photonic weak measurements, Phys. Rev. A 97, 062108 (2018).
  • Dieguez et al. [2022] P. R. Dieguez, J. R. Guimarães, J. P. Peterson, R. M. Angelo, and R. M. Serra, Experimental assessment of physical realism in a quantum-controlled device, Communications Physics 5, 82 (2022).
  • Schmid [2024] D. Schmid, A review and reformulation of macroscopic realism: resolving its deficiencies using the framework of generalized probabilistic theories, Quantum 8, 1217 (2024).
  • Leggett and Garg [1985] A. J. Leggett and A. Garg, Quantum mechanics versus macroscopic realism: Is the flux there when nobody looks?, Phys. Rev. Lett. 54, 857 (1985).
  • Kullback and Leibler [1951] S. Kullback and R. A. Leibler, On information and sufficiency, The Annals of Mathematical Statistics 22, 79 (1951).
  • Plávala [2023] M. Plávala, General probabilistic theories: An introduction, Physics Reports 1033, 1 (2023).
  • Napoli et al. [2016] C. Napoli, T. R. Bromley, M. Cianciaruso, M. Piani, N. Johnston, and G. Adesso, Robustness of coherence: An operational and observable measure of quantum coherence, Phys. Rev. Lett. 116, 150502 (2016).
  • Bong et al. [2020] K.-W. Bong, A. Utreras-Alarcón, F. Ghafari, Y.-C. Liang, N. Tischler, E. G. Cavalcanti, G. J. Pryde, and H. M. Wiseman, A strong no-go theorem on the Wigner’s friend paradox, Nature Physics 16, 1199 (2020).
  • Aerts and Sassoli de Bianchi [2014] D. Aerts and M. Sassoli de Bianchi, The extended Bloch representation of quantum mechanics and the hidden-measurement solution to the measurement problem, Annals of Physics 351, 975 (2014).
  • Aerts and Sassoli de Bianchi [2016] D. Aerts and M. Sassoli de Bianchi, The extended Bloch representation of quantum mechanics: Explaining superposition, interference, and entanglement, Journal of Mathematical Physics 57, 122110 (2016).
  • Martins et al. [2020] E. Martins, M. F. Savi, and R. M. Angelo, Quantum incompatibility of a physical context, Phys. Rev. A 102, 050201 (2020).
  • Storrer [2023] M. Storrer, Theory-independent context incompatibility, Master’s thesis, Department of Physics, Federal University of Paraná (2023).