0% found this document useful (0 votes)
8 views25 pages

Taxonomy For Physics Beyond Quantum Mechanics

The document proposes a taxonomy to classify interpretations and modifications of quantum mechanics, addressing terminological confusion in the literature. It introduces a general model framework to distinguish between different types of models, including calculational and mathematical models, and aims to facilitate clearer communication among researchers. The authors emphasize that their classification is not a judgment of the validity of the models but a means to aid understanding and discussion in the field.

Uploaded by

indal1
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
8 views25 pages

Taxonomy For Physics Beyond Quantum Mechanics

The document proposes a taxonomy to classify interpretations and modifications of quantum mechanics, addressing terminological confusion in the literature. It introduces a general model framework to distinguish between different types of models, including calculational and mathematical models, and aims to facilitate clearer communication among researchers. The authors emphasize that their classification is not a judgment of the validity of the models but a means to aid understanding and discussion in the field.

Uploaded by

indal1
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 25

Taxonomy for physics beyond

royalsocietypublishing.org/journal/rspa quantum mechanics


Emily Adlam1, Jonte R. Hance2,3, Sabine
Hossenfelder4 and Tim N. Palmer5
Research 1
The Rotman Institute of Philosophy, Western University, 1151 Richmond Street,
London, Ontario N6A 5B7, Canada
2
School of Computing, Newcastle University, 1 Science Square, Newcastle upon
Cite this article: Adlam E, Hance JR, Tyne NE4 5TG, UK
Hossenfelder S, Palmer TN. 2024 Taxonomy for
3
Quantum Engineering Technology Labs, Department of Electrical and Electronic
Engineering, University of Bristol, Woodland Road, Bristol BS8 1US, UK
physics beyond quantum mechanics. Proc. R. 4
Munich Center for Mathematical Philosophy, Ludwig-Maximilians-Universität
Soc. A 480: 20230779. München, Geschwister-Scholl-Platz 1, Munich 80539, Germany
https://doi.org/10.1098/rspa.2023.0779 5
Department of Physics, University of Oxford, Oxford OX1 3PU, UK

JRH, 0000-0001-8587-7618
Received: 24 October 2023
We propose terminology to classify interpretations
Accepted: 10 June 2024
of quantum mechanics and models that modify
or complete quantum mechanics. Our focus is on
models which have previously been referred to
as superdeterministic (strong or weak), retrocausal
Subject Category: (with or without signalling, dynamical or non-
Physics dynamical), future-input-dependent, atemporal and
all-at-once, not always with the same meaning or
context. Sometimes, these models are assumed to be
Subject Areas:
deterministic, sometimes not, the word deterministic
quantum physics has been given different meanings, and different
notions of causality have been used when classifying
them. This has created much confusion in the
Keywords: literature, and we hope that the terms proposed
quantum, taxonomy, interpretations, models, here will help to clarify the nomenclature. The
theories general model framework that we will propose may
also be useful to classify other interpretations and
modifications of quantum mechanics. This document
Author for correspondence: grew out of the discussions at the 2022 Bonn
Workshop on Superdeterminism and Retrocausality.
Jonte R. Hance
e-mail: [email protected]

1. Introduction
Quantum mechanics, despite its experimental success,
has remained unsatisfactory for a variety of reasons,
notably due to its tension with locality, and due to
the measurement problem [1,2]. Different authors have

© 2024 The Author(s). Published by the Royal Society under the terms of the
Creative Commons Attribution License http://creativecommons.org/licenses/
by/4.0/, which permits unrestricted use, provided the original author and
source are credited.
formulated their unease with quantum mechanics in different ways. Analysing the origin of this
2
unease is not the purpose of this present article. The purpose is instead to sort out the confusion
in the terminology used to describe this unease.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


We will below introduce a framework to distinguish between interpretations of quantum
mechanics and modifications thereof. Our hope is that it may offer researchers a guide to
classify their own approach. Our aim here is not to judge the promise or correctness of any
of these approaches, but to make it easier to communicate with each other about what we are
working on in the first place. An accompanying paper [3] will discuss some applications of this
terminology to showcase its use.
This paper is organized as follows. In the next section, we will outline the general model
framework and most importantly introduce different types of models. We believe that this
classification of what we even mean by a model in and by itself will serve to alleviate the
confusion of nomenclature. We will then, in §3 and §4 introduce some properties that such
models typically have. In §5, we will briefly explain how to use this classification scheme, and
then we conclude in §6. A list of acronyms can be found in the appendix.

2. General model framework


We will start with explaining the general framework that we want to use in the following, so
that we can distinguish between different types of models and their properties.

(a) Calculational models


At its most rudimentary level, the task of physics is to provide an useful method for calculat-
ing predictions (or postdictions) for observables in certain scenarios. (Other potential tasks
associated with physics (e.g. explaining phenomena, telling us what exist, examining what is
built up from what, how it changes over time and why, or producing knowledge of physical
reality) are both more debatable than this, and require such an useful method themselves as a
basis, so we stick with this rudimentary description as a minimal example of what the task of
physics is). A scenario is most often an experiment, and this is the situation we are typically
concerned with in quantum foundations. But in some areas of physics—such as cosmology and
astrophysics—one does not actually conduct experiments, but instead one observes a natural
variety of instances that occurred in the past. For this reason, we will use the more neutral term
‘scenario’. A scenario is not itself a mathematical structure; it is the real-world system that we
want to describe using mathematical structures.
The basic logical flow of such a calculation is outlined in figure 1. We will call the centre
piece of this calculation a calculational model, or c-model for short. We do not call it a ‘computa-
tional model’ because ‘computational’ suggests a numerical simulation, which is an impression
we want to avoid. A calculational model could be numerical, but it could also be purely

We will, in the following, use a notation in which curly brackets ⋅ denote sets of mathe-
analytic. For more on the relation between calculational models and computer models, see §2f.

matical assumptions. The set ⋅ is the set of all assumptions that can be derived from ⋅
(including the elements of ⋅ themselves).
The different components of the modelling framework have the following properties:
Inputs (of a calculational model): The inputs I of a calculational model are a (at most countably
infinite) set of mathematical assumptions—each of which is an input, I i—that describes the

We will denote this set as I : = I i | i ∈ K ⊆ ℕ+ , were K is the (at most countably infinite) index
scenario. To be part of the inputs, an assumption must differ between at least two scenarios.

range of the assumptions.


Mathematics 3
Inputs Model Outputs

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Real world
Scenarios Observables

Figure 1. The logical flow of the model framework.

To aid readability, and because the index set will not matter in the following, we will from here

will just write I i rather than I i | i ∈ K ⊆ ℕ+ .


on assume that all sets are at most countably infinite and suppress the index range. That is, we

Loosely speaking, you can think of the inputs as the mathematical representation of a
scenario. A typical example might be the temperature in the laboratory, or the frequency of
a laser. However, the inputs of a model do not necessarily have to correspond to definite
observable properties of the scenario. They could also be expressing ignorance about certain
properties of the scenario and thus be random variables with probability distributions, or they
might indeed be entirely unobservable. We will come back to this point in §3a.

may work with an unspecified variable x that is an element of some space. The input may then
The inputs of a c-model are often assignments of values to variables. For example, a c-model

assign the value x = 3 to this variable for a specific scenario. For such an input, we will refer to
the value it assigns as the input value. We want to stress that the input value is not the same as
the input. The input is ‘x = 3’, the input value is ‘3’.

oscillator. In this case, the model would be the differential equation mẍ = − kx and a complete
A typical example of a c-model that we are all familiar with would be the Harmonic

set of inputs would be value assignments for k and m, plus two initial values for, say, x t = 0
and ẋ t = 0 .
However, inputs of a calculational model are not necessarily value assignments. They might
also be constraint-equations, or boundary conditions, or something else entirely. For example,
when studying stellar equilibrium, it is quite common to enter an equation of state as the input
to the Tolmann–Oppenheimer–Volkov (TOV) equations. In this case, the TOV equations are the
same for all stellar objects that one may want to consider; they are hence part of the model. The
equation of state, on the other hand, changes from one type of star to another; it is hence part of
the input.
While our main interest is in models that describe the real world, it is also possible to
study a model’s properties with scenarios that do not exist in reality. We will refer to those as
hypothetical scenarios. They include, but are not limited to, counterfactual realities, as well as
universes with different constants of nature. (Note Frigg & Nguyen [4], among others, have also
discussed representation in scientific modelling.)
Calculational model: A set C of mathematical assumptions Ax that are independent of the
scenario. We will denote this set as C: = Ak .
Set-up: The union of a calculational model and its inputs. F : = I ∪ C.

but from neither the model nor its inputs in isolation O = F ∖ (I ∪ C).
Outputs (of a calculational model): All mathematical statements that can be deduced from set-up,

Predictions from observables are obtained from the outputs of the model. However, not all
outputs of a model need to be observable. The prime example is quantum mechanics, where
the outputs contain the time-evolution of the wavefunction, but the wavefunction itself is not
observable. But there are many other examples, such as the production of gravitational waves
by a black hole merger. Given suitable inputs, the model (General Relativity) will output a
4
mathematical description for the creation and propagation of gravitational waves, but we only
measure their arrival on Earth, and only measure that through the waves’ influence on matter,

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


which is a small part of all the model outputs. And like the inputs, the outputs of a model do
not necessarily have to result in definite values for observables; they could instead merely give
rise to probability distributions for values of observables.
For a calculational model to be useful, we further need a prescription to encode a scenario
with specific inputs, and a way to identify the observables from the outputs. This identification,
since it is not restricted to the mathematical world, is not one that we can strictly speaking
define. Science, after all, is not just mathematics. What property this identification with the real
world must have is an interesting question, but it is one of many that we will not address here,
because it is not relevant for what follows.
When using hypothetical scenarios, these have to be distinguished from the set-up as a
matter of definition. If the hypothetical scenarios are not identified as such by definition, it
becomes impossible to tell whether one changes the hypothetical scenario or the model.1
Another property, relevant for our purposes, that we want the set-up of a calculational
model to have is: that they are irreducible, in the sense that we cannot split the combined
set C: = I j, Ak into two sets C1 : = I j1, Ak1 and C2 : = I j2, Ak2 , where I j = I j1 ∪ I j1 and
1 1 2 2 1 2

Ak = Ak1 ∪ Ak2 , so that both C1 and C2 are each set-ups of calculational models and the
1 2

combination of their outputs is the same as the outputs of C. A simpler way to put this is that,
if we split the set-up of an irreducible model into two, some of the output can no longer be
calculated. This approach is reminiscent of identifying particles in quantum field theory from
the irreducible representations of the Poincaré group.
We need this requirement because otherwise, we could just join different set-ups to form a
new one, which would make it impossible to classify any. Note that this does not mean the
composition of two set-ups is not a set-up. On the contrary, a composition of two set-ups is a
set-up, it is just that the combined set-up is no longer irreducible. The issue is, if we allowed
reducible combinations of set-ups, then we could not meaningfully assign properties to them. It
would be like asking which fruit is fruit salad. Once we have, however, succeeded in identifying
properties of irreducible set-ups, we can join those with the same properties together, and
meaningfully assign the same property to the reducible set-up. Using the above fruit example,
if we have identified several fruits as apples, we can join them and be confident we have apple
salad.
A particularly relevant case of a reducible set-up is one in which some inputs or assump-
tions can be removed without changing anything about the outputs. This may be because the
assumptions are not independent (in the sense that some can be derived from the others), or
because an assumption is simply not used to calculate the outputs.
One might be tempted to add to the requirements of a model that its assumptions are
consistent (given problems such as the Principle of Explosion with inconsistent models [5]).
However, as is well known, for recursively enumerable sets, Gödel’s theorem [6] tells us that
we cannot in general prove the consistency of the assumptions. We might then try to settle
on the somewhat weaker requirement that at least the assumptions should not be known to
be inconsistent. However, it sometimes happens in physics that a model works well in certain
parameter ranges despite being inconsistent in general. An example may be the Standard

1
An example for this is the case of the Standard Model with its fundamental constants not fixed but taken as variable
inputs. This cannot be a correct model for scenarios in our universe because in our universe the constants are constant,
hence they cannot be inputs. One can, however, consider hypothetical scenarios with different values of the constants, often
interpreted as a type of multiverse. But, of course, one could alternatively interpret these hypothetical scenarios as different
versions of the Standard Model that, alas, happen to not agree with our observations. The point is that whether the
Standard Model with fundamental constants that do not agree with our observations describes a hypothetical alternative
universe, or is just a wrong model for our universe, is a matter of definition.
Model without the Higgs field and its boson [7]. For this reason, we will here not add any
5
requirement about consistency. One may justify this by taking a purely instrumental approach.
We only care whether a set of assumptions is any good at describing observations.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


The set-up of a calculational model whose inputs are all value assignments that, alas, have
not been assigned is what is usually called a model in the causal models literature [8,9].

(b) Mathematical models


We defined a calculational model and its inputs as sets of mathematical assumptions. Such
sets can be expressed in many different ways that are mathematically equivalent. We will
call an equivalence class of calculational models a mathematical model, or m-model for short,

with [ ⋅ ]m and refer to it as an m-class; it is a set of sets. For example, if C = Ak is a


that is the term m-class means the same as m-model. We will denote this equivalence class

calculational model, then M: = [C]m is the m-model that encompasses all mathematically
equivalent formulations

[C]m : = Bj : Bj ⇔ Ai . (2.1)

Calculational models in the same m-class will be called m-equivalent. By a mathematical


equivalence of the sets ( ), we here mean that either one can be derived from the other. We
thereby adopt the notion of model equivalence advocated by Glymour [10,11].2
It was argued by Weatherall [12,13] that mathematical equivalence might overdistinguish
models in certain circumstances, and that it might be better to use a weaker form of structural
equivalence based on category theory. We do not use this proposal of categorial equivalence
here, not because we object to it, but because it is neither widely accepted nor understood. Most
importantly, it is at present not practical because few physicists would know how to apply it.
Mathematical equivalence, in contrast, is widely understood and comparably straightforward
(though not necessarily easy) to prove.
Among the models we will discuss here, mathematical models are closest to what physi-
cists typically mean by a ‘model’. They do not distinguish between the particular mathemati-

Maxwell’s Equations using differential forms with d ⋆ , d ∧ and ⋆d operations, or we could do


cal formulations that one might use to make a calculation. For example, we could express

it with the more old-fashioned ∇ , ∇ ⋅ and ∇ × operators. These two definitions would be two
different calculational models, but the same mathematical model. A similar equivalence covers
switching from the Schrödinger picture to the Heisenberg picture in quantum mechanics. We
believe that most physicists would not call these different models.
The inputs of a mathematical model are likewise an equivalence class: it is the class of all sets
of assumptions which change with the scenario that, together with the mathematical model,
produce equivalent outputs.
The inputs of a mathematical model are usually not just the set of assumptions that are
mathematically equivalent to the inputs of one of the calculational models. This is because the
assumptions of the model itself may render certain inputs equivalent that are not equivalent

time-reversible evolution operator will produce identical outputs for input states at times t1 and
without the model. An example that will be relevant for what follows is that a model with a

t2 > t1, if the input state at t2 is the forward evolution of the state from t1. In this case, the input is
m-equivalent, though the inputs at different times are not equivalent to each other without the
model.
We will refer to different calculational models as representations of their m-class.

2
Glymour uses the term ‘theory equivalence’, not ‘model equivalence’. This is not a relevant distinction for the purposes of
this present paper: please see §2e for discussion.
Representation (of an m-model): A calculational model within the m-class of that m-model. 6
In Argaman [14], a similar concept, that of two c-models within the same m-class, was referred

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


to as ‘reformulations’ of each other.

(c) Physical models


The previous definitions do the heaviest lifting for the model framework, and the remaining
ones are now straightforward. It can happen that mathematical models are different, but they
nevertheless give rise to the same observables for all scenarios. We will combine all such

physical model, or p-model for short. We will denote this equivalence class with [ ⋅ ]p and refer to
mathematical models into yet another, even bigger class and say they constitute the same

it as the p-class. Note that we could take either the p-equivalence class of a computational model
or that of a mathematical model. Models in the same p-class will be called p-equivalent.
By calling these models physical, we do not mean to imply that we only consider observa-
ble properties as physically real. The reader should not take our nomenclature to imply any
statement about realism or empiricism. We call them physical just because they describe what
physicists in practice can distinguish with physical measurements.
This now gives us a way to define what we mean by an interpretation:
Interpretation (of a p-model): Any one of the mathematical models in the p-class.

Note that according to this nomenclature,3 each interpretation has its own representations. That
is, a c-model is a representation of an m-class, but more specifically it is a representation of a
particular interpretation.
For example, if we take the Copenhagen Interpretation (CI) in one of its common axiomatic
formulations (e.g. as given in Zurek [16]), that set of axioms will constitute a particular representa-
tion, hence, a c-model. The CI per se is the class of all mathematically equivalent models. We can then
further construct the physical equivalence class of the CI, which we will in the following refer to as
Standard Quantum Mechanics (SQM:=[CI]P). Any other mathematical model that is physically but
not mathematically equivalent to the CI is also an interpretation of SQM.
For the purposes of this article, we will not need to specify exactly what we mean by CI.
However, we will take it to be only a first-quantized theory. That is, it is not a quantum field
theory. If one were to specify further details, one would in particular have to decide whether
one considers the relativistic version, or only the non-relativistic case, as some alternative
interpretations work only non-relativistically [17].
That we have interpretations and not just representations is a possibility which appears
in quantum mechanics because it has outputs that are unobservable in principle. A different
computational model which affects only unobservable outputs might not be mathematically
equivalent and yet give rise to the same physics.

(d) Empirical models


Finally, we define an empirical model, or e-model, E , as the class of all physical models that cannot
be distinguished by observations so far. We will denote this class as [ ⋅ ]e, refer to it as an
empirical class or e-class. Models in the same e-class are empirically equivalent or e-equivalent. We
will call anything in the same empirical class as SQM, but not in the same physical class, a
modification of quantum mechanics:

3
Note others have given different definitions of interpretations for quantum mechanics, e.g. [15].
Modification or completion or extension (of a p-model): Another p-model that is not physically 7
equivalent but so far empirically equivalent.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Figure 2 summarizes the relations between these various types of models.
We here use the terms modification, completion and extension loosely, and will not
distinguish them below because it is not necessary for what follows. In the literature, these
terms are used with a somewhat different meaning. A modification or extension of a model
typically employs similar mathematical structures, whereas a completion introduces new
mathematical structures. However, the distinction between the two is not always clear. Again,
just how to distinguish them is an interesting question in its own right, but not one we will
address here.
With a slight abuse of terminology that is, however, unlikely to cause misunderstandings,
we can also use the above definition to refer to an m-model or a c-model as a modification.
That is, an m-model (c-model) is a modification of another m-model (c-model) if the two are not
physically equivalent, but so far empirically equivalent.
One could further refine the class of empirically equivalent models by specifying various sets
of experiments. For example, we could ask what models can be distinguished by experiments
in the near future, and what exactly we mean by near future. Or, we could talk about models
that can be distinguished by experiment in principle, and then discuss whether, say, waiting 100
billion years is possible in principle. Or, whether it is possible in principle to make measure-
ments inside a black hole, and so on. All of these are very interesting questions; however, they
will not concern us in the following, so we will not elaborate on them.

(e) Theories
We will in the following not distinguish between models and theories. Loosely speaking, a
theory is a class of models that can be used for a large number of scenarios, congruent with the
semantic approach of Suppes [18] and van Fraassen [19]. However, physicists do not tend to use
the terms theory and model in a well-defined way.

Model in particular. We refer to General Relativity as a theory, and to ΛCDM as a model. This
For example, we use the terminology of Quantum Field Theory in general, and the Standard

agrees with what we said above. But we also refer to Fermi’s theory of beta decay, and the BCS
theory, though those would better be called models. To make matters worse, we also sometimes
call supersymmetry a theory, when it is really a property of a class of m-models, and so on. For
the purposes of this article, we will not need to distinguish theories from models, so we will not
bother with a precise definition of the term ‘theory’.

(f) Computer models and simulations


Another type of model which is commonly used by scientists of all disciplines are computations
performed on computers. There are two ways to think of these models.
One is that they are really simulations that represent one real-world system with another
real-world system. That is to say, they are not models in the sense that we have considered
here—the models we considered here are mathematical. A suitably programmed computer is
instead a physical stand-in for the system one wants to make a prediction for. This is also how
quantum simulations work [20,21].
However, there is another way to look at computational models, which is to use their
program as a definition for a calculational model. This then falls into the classification scheme
we discussed above. But, in this case, the computational model is typically not mathematically
equivalent to the analytical, calculational model that one used for a scenario. This is because
computers are in almost all cases digital, and only approximate the continuum. The exception
8

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Representations of

Interpretations of

Modifications of
Figure 2. A mathematical model Mi jk is an equivalence class of computational models Ci jkl. A physical model P i j is an
equivalence class of mathematical models Mi jk. An empirical model E i is an equivalence class of physical models P i j.

are certain types of analogue computers, which however can better be understood as simula-
tions again.
That is to say, when one wants to classify a computer model according to our scheme, one
should take its algorithmic definition as a set of assumptions, and use that to define a calcula-
tional model and its inputs. This calculational model, defined by its executable algorithm, will
be different from the calculational model that uses an analytic expression.
(One can then further ask to what accuracy will the algorithm, when executed on a physical
computer, approximate the output of the analytical model. This is a relevant point which
was previously brought up in [22]. While an analytically defined calculational model might
be time-reversible, an algorithmic approximation of it run on a computer will in general no
longer be time-reversible. This is because errors will build up differently depending on which
direction of time one runs the algorithm. This is particularly obvious for time-evolutions which
are chaotic. In such cases, the forward-evolution and the backward-evolution of the algorithm
as executed on the computer will actually be two different calculational models, and both are
different from the analytical calculational model that they approximate).

3. General model properties


In this section, we will discuss some properties that we can assign to models and their inputs.
The point of this section is to specify which properties are m-class properties (do not change
with mathematical redefinitions), and which are c-class properties (can change with mathemati-
cal redefinition).
We want to stress that we will not prove that these assignments are correct. To do that,
we would have to add many more assumptions (e.g. about what we mean by space, time,
measurements and so on). What we will do instead is specify what we believe is the way that
an expression has been previously dominantly used in the literature, and this specification will
then in turn imply properties for the concepts we did not specify.
To make this procedure clear, if we state for example that the property of ‘time-reversibility’
does not change with a mathematical redefinition, then this implicitly means: if it did change,
we would not refer to the property as time-reversibility. That is, there are certain properties of
models which we want to not depend on just exactly how we write down the mathematics.
Quite possibly, some readers will disagree with some of our assessments. This is fine. Our
aim here is not to end the discussion, but to propose a language in which we can talk about
these properties in the first place.
The term ‘model’ without further specification (c/m/p/e) will refer to any type of model.
9

(a) Atemporal properties

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Following the terminology of Cavalcanti & Wiseman [23], we will call an m-model ‘determinis-
tic’ if its observables can be calculated with certainty from the outputs obtained from the set-up:
Deterministic: An m-model is deterministic if4 the probabilities for predictions of observables are
all either 0 or 1.

This property was termed ‘holistic determinism’ in Adlam [24] and differs from a more
common definition of determinism, that connects one moment in time with a later one. We
will elaborate on this in the next subsection. For now, we further follow [23] and distinguish
determinism from predictability:
Predictable: An m-model is predictable if it is deterministic, and the inputs are all derived from
observable properties of the scenario.

Both determinism and predictability are m-model properties, because we expect a redefinition
of the mathematics that one uses to process inputs to not change predictions for observable
output. If that was the case, we would assume something is wrong with our idea of what is
observable.
The distinction between deterministic and predictable is that a model may have inputs
that are unknowable in principle. Typically, these are value assignments for variables that are
usually referred to as ‘hidden variables’. A model with such hidden variables, according to the
above terminology, may be deterministic and yet unpredictable.
Hidden variables: Hidden variables, that we will denote κ, are input values to a c-model that
cannot be inferred from any observation on the scenario.

We want to stress that these hidden variables are in general not localized and sometimes not
even localizable in any meaningful way. We will say more about localizable variables in the
next subsection, but a simple example is that we could use Fourier-transforms of space–time
variables, or just extensive quantities that are properties of volumes. Note that hidden variables
are defined for c-models, not for m-models. This is because hidden variables can be redefined
into (not necessarily local) random variables with an equivalent mathematical outcome. That is
to say, mathematically, it makes no difference whether a variable was unknowable or indeed
random.
While it may sound confusing at first to distinguish determinism from predictability, it will
be useful in what follows. Indeed, the reader may have recognized that Bohmian Mechanics
is deterministic yet not predictable. In Bohmian Mechanics, observables can be calculated with
certainty if the inputs are specified, yet the inputs are also assumed to be partly unobservable in
principle, so predictions can still not be made with certainty. Since determinism is a property of
an m-model that cannot be removed with a redefinition, it follows that Bohmian Mechanics is
not a representation of SQM. We elaborate on this further in the companion paper [3].

(b) Local and temporal properties


We will now add some properties of models that we commonly use in physics, starting with
those that refer to locations in space and time. Clearly, we can only speak of locations in space
and time if a model has some notion of space and time, and a distance measure on them, to
begin with. In the simplest case, that would be the usual space–time manifold and its metric

4
We use the common mathematical abbreviation ‘iff’ for ‘if and only if.
distance. But it could alternatively be some other structure that performs a similar function,
10
such as a lattice, or a discrete network with a suitably defined metric.
To make sense of space and time, we will in the following consider only a restricted m-class

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


associated with a calculational model, that is one in which space and time are consistently
defined throughout the entire class. To see why this is necessary, imagine if we were to redefine
one direction of ‘space’ as ‘time’ and vice versa. This is mathematically possible, but it makes no
physical sense. It would screw up any notion of locality and causality just by nomenclature. We
will hence assume that space and time and a distance measure on them are consistently defined
throughout the m-class.
However, sometimes a model just does not have a description of space–time or a notion of
locality. This might sound odd, but is not all that uncommon in the foundations of quantum
mechanics, where many examples deal with qubit states that do not propagate and are not
assigned space–time locations [25]. We will refer to such models as alocal:
Alocal: An m-model that does not have a well-defined notion of space, space–time, locality or
propagation.

One cannot fix a lack of definition by switching to a different but mathematically equivalent


definition, so if a c-model is alocal, then so will be any other c-model its m-class.
For alocal models, we cannot make any statements about whether they are local or not.
Similarly, a model may just not have a notion of time or a time-evolution. This again is not
all that uncommon in the foundations of quantum mechanics. Many elaborations on correla-
tions and classicality bounds, for example, do not specify a time-evolution for states; and the
process matrix formalism is specifically designed to study possible quantum processes without
a predefined time order [26,27].
Atemporal: An m-model that does not have a well-defined notion of time.

For models which have a proper notion of space–time, and a distance measure on it, we then
want to identify how local they are. For this, we first have to identify the input values that
can be assigned to space–times which Bell coined ‘local beables’ [28]. According to Bell, local
beables ‘can be assigned to some bounded space–time region’. We will take ‘assigned to’ to
mean that the variable is the value of a function whose domain is space–time, or whatever the


stand-in for space–time is in the model at hand. Note ‘region’ might be a point set.
For example, if space–time is parameterized by coordinates x , t , and we have a function
→ →
f: x , t → ℂ, then f x , t is a local beable. The domain of the SQM wavefunction is generically
configuration space and it is therefore not a local beable, though under certain circumstances
local beables can be obtained from it (such as the single-particle wavefunction in position
space). Local beables are not necessarily observable.
Bell’s definition of a local beable makes the assignment to a space–time region optional
(can be assigned). However, if the assignment was optional, then it could be omitted, and a
model with an assumption that can be omitted is reducible. Since we only deal with irreducible
models, we therefore already know that if a variable is assigned to a space–time region, then
this assignment is actually necessary. For this reason, we define
Local beable: An input value of a c-model that is assigned to a compact region of space–time.

In the following, I A will refer to the local beables in space–time region A.


It becomes clear here why we required our models to be irreducible: so that it is actually
necessary, and not just possible, to assign the local beables to space–time regions. It has for
example been rediscovered a few times independently [29,30] that any theory can be made
local (in pretty much any reasonable definition of the term) by copying the local beables from
any space–time location into any other space–time location and postulating them to ‘be there’
11

Concretely, suppose we have localized variables f x, t over a space–time manifold M , then


without any other consequences.

we can just define the state f x, t ⊗ f x′, t′ ∈ M ⊗ M and call that a local beable at x, t , just

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


that now we have the information about any x′, t′ also available at the same point. Depending

ultra-local in that we do not need to know the state at any other location than x, t to calculate
on perspective, one might consider such a model as either ultra-local or ultra-non-local. It is

the time-evolution. It is ultra-non-local because any point in space–time contains a copy of the

A similar problem would occur with parameters, such as ℏ, that could be transformed into
entire universe.

fields ℏ: x, t → δ x − x′ δ t − t′ and then be used as inputs with random locations, leading to


seemingly ‘non-local’ interactions in the Hamiltonian.
Such procedures to localize beables create new c-models that are in different m-classes,
because the copying procedure is an additional assumption that could not have been derived
from the original version of the c-model. Since such a localizing assumption is unnecessary to
calculate outputs, the set-up of such a model is reducible and, as we have previously remarked,
properties of reducible set-ups are ambiguous.
A word of caution is in order here. As mentioned above in the elaboration on alocality,
many discussions of violations of Bell’s inequalities do not explicitly state locality assumptions.
These assumptions might hence appear unnecessary. This is indeed correct if one merely looks
at the inequality. However, if one wants to draw conclusions about locality from a violation of
Bell’s inequality, one needs to at least make a statement about which parts of the experiment
are causally connected or space-like separated. That is, we suspect that many alocal models do
implicitly require locality statements to arrive at the desired conclusions. One should not be
deceived by only looking at the explicitly stated assumptions.
Having introduced local beables, we can now define an alternative:
All-at-once (AAO) input: Input of a c-model that constrains or defines properties of at least one
local beable for at least two temporal regions, and that cannot be decomposed into inputs in
separate temporal regions.

Typical examples of such inputs are event relations, consistency requirements for histories,
evolution laws, temporal boundary conditions or superselection rules. Something as mundane
as an integral over time that depends on the scenario would also be an all-at-once input.
All-at-once (AAO) model: A c-model that uses all-at-once input.

The use of all-at-once input is a priori a property of c-models. That is to say, it might be possible
to reformulate a model with AAO input into a mathematically equivalent model that does not
have this property.
The principle of least action in classical mechanics, for example, uses All-At-Once input (the
action), but given that the Lagrangian fulfils suitable conditions, the principle can equivalently
be expressed by the Euler–Lagrange equations which do not require AAO input. Another
example may be the cosmological model introduced by Carroll & Remmen [31], which uses
a constraint on the space–time average of the Lagrangian density. The authors refer to this as
non-local, and in some sense it is, but it is more importantly also an all-at-once input.
Now that we have localized variables, we can define a notion of locality. We will first leave

Argaman [22]. The idea is that if one wants to calculate the outputs O A for a region A, then
aside causality and introduce a notion of Continuity of Action (CoA), as done in Wharton &

it is sufficient to have all the information on a closed hypersurface, S1, surrounding the region,
and adding local beables from another region outside S1 provides no further information.
condition P O A | I S1 , I B = P O A | I S1 for any region S1 that encompasses A but not B
CoA (locality): An m-model has continuous action or fulfils CoA or is local iff it fulfils the 12

(see figure 3a).

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Non-locality: An m-model is non-local if it violates CoA.
Note that fulfilling CoA does not mean that the outputs in A are determined by the input values
on S1 to begin with. It is just that adding information from B does not make a difference.
A natural question to ask here is just how small the regions should be for the model to have
continuous action. This is a most excellent question but luckily one that we do not need to
answer because it resides on the level of empirical adequacy. A model in which variables cannot
be localized sufficiently much will simply not reproduce certain observations. (For example,
a model with a spatial resolution so rough that it cannot distinguish two detectors from each
other can only give a total probability for both, not for each separately.) The typical size of the
regions is what is often referred to as the coarse-graining scale of a model.
A subtlety of the definition for CoA was pointed out in Palmer [32], which is that some
combinations of inputs may be mathematically possible, but are not in the physical state space
of the model and hence, do not correspond to any scenario which can occur in reality. This
might happen, for example, because certain combinations of variables are just forbidden by

might have a situation where, e.g. P I S1 , I B = 1 and P I S1 , I B ′ = 0 just because the latter
a mathematical constraint (as happens with the Fermi exclusion principle). In this case, one

P A | I S1 , I B ≠ P A | I S1 , I B ′ and CoA is violated. However, as argued in Palmer [32],


combination is incompatible with the assumptions of the model [33]. It would then seem that

it is rather meaningless to talk about violations of locality for configurations which do not
physically exist. Hence, if a model has input constraints, one should only apply the locality
requirement to inputs that lead to physically possible scenarios.
We define CoA as an m-model property because, if it could be removed with a
mathematical redefinition, we believe most physicists would not accept the definition as
meaningful.
The term ‘non-locality’ has been used to refer to many other definitions in the foundations
of physics in general, and quantum mechanics in particular. For example, in field theories,
non-locality often refers to dynamical terms of higher order, a definition that is also used in
General Relativity [34]. In quantum field theories, non-locality usually refers to the failure of
operators to commute outside the light cone. Even in quantum foundations, non-locality may
refer to different properties. For example, entanglement itself is sometimes considered non-local
despite being locally created [35]. The latter in particular has created a lot of confusion, because
while it has been experimentally shown that entanglement is a non-local correlation and does in
fact exist, this does not imply that nature is non-local in the sense that the term has been used in
Bell’s theorem, which has of course not been shown [36].
Surveying all those different notions of non-locality is beyond the scope of this present work.
However, we want to stress that as definitions, none of these notions of non-locality are wrong.
They are just that: definitions. We chose our definition to resemble closely the ‘spooky action at
a distance’ that Einstein worried about.
According to our definition, a calculational model fulfils CoA even if it cannot be directly
evaluated whether it fulfils the requirement on the conditional probabilities, so long as a

for our purposes is that [CI]m does not fulfil CoA. This is because making a measurement in B
mathematically equivalent reformulation of the model fulfils it. The most relevant example

provides information about the measurement outcome in A that was not available in S1. This is
Einstein’s ‘spooky action at a distance’.
That is, with the terminology we have introduced so far, the CI and all mathematically
equivalent formulations are equally non-local. The question is then merely whether this is
(a) (b)
S2 13

B B

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


A A
Time

Time
S1 S2

Space Space
Figure 3. (a) CoA. (b) Strong CoA.

something that we should worry about. If one can understand the wavefunction as an epistemic
state (a state of knowledge), then its non-local update is not a priori worrisome.
CoA loosely speaking means that localizable variables can only influence their nearest
neighbours. However, it does not require that this influence lies within the light cones. To

their insides to a space–time region, A. We will denote with L(A) the union of all space–time
arrive at a stronger condition, we will therefore now as is customary assign the light-cones and

points that are light-like or time-like related to any point in A.


We can then cut out regions from the closed surface S1 using the light-cones: S2 : = S1 ∩ L A
and arrive at a stronger version of CoA:
Strong CoA (locally subluminal): An m-model has Strong CoA if local beables outside the forward

P O A | I S2 , I B = P O A | I S2 ∀ S2 that do not overlap with the light cones of B, i.e.


and backward light-cones of a region play no role for calculating outputs in that region

S2 ∩ L(B) = ∅ (see figure 3b).

And correspondingly,
Weak CoA (locally superluminal): An m-model has Weak CoA if it fulfils CoA but not Strong CoA.

Weak CoA, roughly speaking, means that influences happen locally, but sometimes superluminally.
What we call Strong CoA was called Einstein locality in Maudlin [37]. Local and non-local models
can further be distinguished into those which are super- and subluminal. It is rather uncommon
to consider subluminal non-locality, but it will be helpful in what follows to clearly distinguish

local beables necessary to find out what happens at A are those within or on the light-cones of A, and
non-locality from superluminality. We can define subluminal non-locality by requiring that the only

adding the light-cones of B and their inside brings no further information.


Non-locally subluminal: An m-model is non-locally subluminal iff it is non-local, but local

P O(A) | I(L(A) , I(L(B))) = P O(A) | I(L(A) .


beables outside light-cones of a region play no role for calculating outputs in that region

And, consequently,
Non-locally superluminal: An m-model is non-locally superluminal iff it is non-local, but not
non-locally subluminal.

For a summary of those four terms, see figure 4.

inside the light cones, because of the requirement that one only considers regions S2 which
We should also mention here that strong CoA is a weaker criterion than confining CoA

do not overlap with the light cones of region B. The reason for this requirement—as noted by
Bell already—is that otherwise the outputs in B might well provide additional information that
correlates with the inputs from S2 (and hence, the outputs in A) without influences ever leaving

regions A and B far enough, they will always overlap. This means that one can always try to
the light-cones (see appendix for explanation). But, of course, if one extends the light cones of
14

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Non

al
n
umi
-loc

perl
a
lly s

u
ly s
uper

al
lum

Loc
Loc al
ally min

inal
rlu
supe
rlum supe
ally
inal -loc
Non
Time

n al Loc
umi ally
perl s uper
al ly su lum
-loc inal

Non
Non
inal

-loc
rlum

ally
supe

supe
ally

rlum
Loc

inal

Space

Figure 4. Difference between local, non-local and superluminal. Note that local and non-local are here distinguished by
having solid/dotted lines respectively, and that each arrow is only one representative for the entire quadrant (e.g. one can
imagine a reflected version of the left-to-right ‘locally subluminal’ arrow, going instead of right-to-left, which is also valid, so
long as it is in both the past and future light cones).

explain violations of Strong CoA by locating an origin of correlations between A and B in the
overlap of the light cones.5 We will come back to this later.

(c) Properties of temporal order


Like with the local and temporal properties, to talk about temporal order, a model needs to
have been provided with additional information. We need not only a notion of time, but also
an arrow of time that allows us to distinguish past and future. This arrow of time is often
not explicitly defined but implicitly assumed. Typically, it comes in because we assume that an
experimenter can choose a setting in the future, but not one in the past. That is, the arrow of
time sneaks in with what we consider to be a possible scenario.
Since an arrow of time is a priori a matter of definition, we have to specify that this definition
has to be consistent for all mathematically equivalent expressions of a set-up.
There are many notions of arrows of time that have been discussed in the literature [38,39]
and our aim here is not to unravel this discussion. We will merely note that a model needs to
have such a notion for the following properties to be well-defined.

5
This is often called a common ‘cause‘. However, all that is required here is a correlation, not a causation.
Like before, it is possible to have a model that just does not have a temporal order, or that
15
does not distinguish past and future. Indeed, this is the case for many of the simplest models
that we deal with, such as an undamped harmonic oscillator, or the two-body problem in

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Newtonian gravity.
Acausal: An m-model is acausal iff it does not have a well-defined arrow of time, and hence no
notion of past or future.
An atemporal model is also acausal—one cannot have an arrow of time without having time to
begin with. This feature is exhibited in the process matrix formalism, which can even be used
to model processes for which it is impossible to specify a well-defined arrow of time [26,27].
One might plausibly argue that a model which is not time-reversible necessarily has an arrow of
time, and hence cannot be acausal, but then we did not say who or what defines that arrow of
time, so we cannot draw this conclusion.
As stressed earlier, some properties of models only make sense with an arrow of time that
orders times. We now come to the first of those:

an arrow of time, according to which calculating localizable output values at time t′ does not
Temporally deterministic: A c-model is temporally deterministic if it is both deterministic, and has

require inputs that are local beables at t > t′. An m-model is temporally deterministic if at least
one of its c-models is.
This is a complicated way of saying that, in a temporarily deterministic model, a future state
can be calculated with certainty from a past state, but not necessarily the other way round. This
notion of temporally deterministic is what is often meant by the term ‘deterministic’. Note that
a temporally deterministic model might have other inputs besides value-assignments for local
beables. In particular, a model with all-at-once inputs may still be temporally deterministic.
We have defined temporal determinism of an m-model from the requirement that at least
one of its c-models has that property, because temporal determinism is easy to remove by
redefining all variables so that they mix different times, or using (partially) time-like boundary

In the CI, so long as no measurement occurs, the state of the wavefunction at time ta is
conditions.

temporally determined by the state of the wavefunction at time tb ≠ ta and the Hamiltonian
operator. The state of the wavefunction after measurement, on the other hand, is generically not
determined by the state of the wavefunction before measurement. Hence, the CI (c-model) is
not temporally deterministic.
It does not follow from this that SQM, which we defined as [CI]m, is also not temporarily
deterministic. However, this is so because—as we saw earlier—SQM is not deterministic to
begin with, so it cannot be temporarily deterministic either.
For temporal models, we can further define:

deterministic under the replacement t → − t, where t measures time.


Time-reversible: An m-model is time-reversible iff it is both temporally deterministic, and also

As mentioned earlier, this definition implicitly makes a statement about what properties we
expect of time, and hence cannot stand on its own. That is not its purpose. Its purpose is rather
to capture what properties physical properties like time and time-reversibility should have.
Time-reversibility should not be confused with invariance under time-reversal, which is
a stronger requirement, but one that we will not consider here. Just because a model is
time-reversible, does not mean that its time-reversed version is the same as the original.
Next, we recall a term previously introduced in Wharton and Argaman [22]:
Future input dependence (FID): A c-model has a FID iff, to produce output for time t, it uses local
beables from at least one time t′ > t for at least one scenario.
FID is a property of the set-up of the c-model, and may simply be a matter of choice. For
16
example, in any time-reversible c-model, one can replace a future input with a past input and
get the same outputs. We define it here because it was previously used in Wharton & Argaman

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


[22] and we want to make a connection to this definition below. However, we also want to
define a somewhat-stronger property:

producing outputs for time t requires inputs from t′ > t. An m-model has an FIR if all c-models
Future input requirement (FIR): A c-model has an FIR iff there is at least one scenario for which

in its equivalence class have an FIR.


Another way to phrase this is that a c-model with an FIR has at least one scenario for which
the input cannot be entirely chosen in the past. A c-model with an FIR cannot be temporally
deterministic. However, a model that is not temporally deterministic does not necessarily have
an FIR.
A well-known example for an FIR is a space–time that is not globally hyperbolic, in which
case the initial value problem is ill-posed. The time-evolution can then not be calculated unless
further input is provided about what will happen.
Another example might be a model which evaluates possible policy paths to limit global
temperature increase to certain goals within a certain period of time (say, 2, 3 or 4°C by 2050).
Such a model requires specifying the desired boundary condition in the future. Of course, in
this case, one is dealing with hypothetical scenarios, but the example illustrates that FIRs are
used in practical applications.

(d) Properties of causal order

forward light-cones of a region A, that we will denote Lp A and Lf A , respectively. It is


For models with a temporal order, we can distinguish the past/backwards and the future/

L A = Lp A ∪ Lf A , and we define S3 : = S1 ∩ Lp A , the intersection of the shielding region S1


with the past light-cone. With this, we can further refine CoA to a notion of local causality:

localized in the past light cone: P O A | I S3 , I B = P O A | I S3 ∀ S3 that do not overlap


Local causality: An m-model is locally causal iff it fulfils strong CoA with local beables

with the light cones of B, i.e. S1 ∩ L B = ∅ (see figure 5a).


We want to stress that this notion of local causality is based on the mathematical structure
of space–time, and so mixing it with other notions of causality can result in confusion. In
particular, space–time causality is not a priori the same notion of causality that is used in large
parts of the philosophical literature, which concerns itself with the question of cause-and-effect
(one currently popular realization of which is the one based on causal models, that we will refer
to as interventionist causality [8,9]).
According to space–time causality, of two causally related events, the one in the past is the
cause by definition. One thus has to be careful when interpreting local causality using other
notions of causality. This is especially important to keep in mind when we classify retrocausal-
ity.
If one interprets ‘causal’ as referring to space–time causality, then the term ‘retrocausal’
suggests influences that go inside the light cones, but backwards in time. The term retrocausal,
however, does not necessarily imply locality. For example, the Transactional Interpretation
[40,41], often referred to as retrocausal, is also non-local.
We will hence proceed by endowing the classification of the four local properties, summar-
ized in figure 4 with an additional temporal distinction that is either forward in time or
backward in time, according to the arrow of time that we have assumed exists. We would
like to stress that since we have an arrow of time, we can meaningfully distinguish forward and
backward in time also outside the light-cones. The reader may want to think of the arrow of
(a) (b) S2 17

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


A A B
Time

Time
S3
S3

Space Space
Figure 5. (a) Local causality. (b) A subtlety of local causality.

time as a preferred slicing in space–time. This will give us a total of eight distinctions. It is then
straightforward to define:
Local retrocausality: An m-model is locally retrocausal iff it fulfils Strong CoA and has a future
input requirement.
Non-local causality: An m-model is non-locally causal iff it is non-local, subluminal and has no
future input requirement.

Non-local Retrocausality: An m-model is non-locally retrocausal iff it is non-local, subluminal and


has a future input requirement.
As we noted already, the term ‘causality’ might either refer to the light-cone structure of space–
time or to interventionist causality. However, since the term ‘local causality’ is extremely widely
used and refers to space–time causality, we will use the term ‘retrotemporal’ to refer to models
with a future input requirement that do not respect the light-cone structure. This gives us the
remaining four definitions:
Locally superluminal retrotemporal: An m-model is locally superluminal retrotemporal if it fulfils
Weak CoA and has a future input requirement.
Locally superluminal temporal: An m-model is locally superluminal temporal iff it fulfils Weak
CoA but has no future input requirement.
Non-locally superluminal retrotemporal: An m-model is non-locally superluminal retrotemporal iff
is non-locally superluminal and has a future input requirement.
Non-locally superluminal temporal: An m-model is non-locally superluminal temporal if it is
non-local and superluminal, but has no future input requirement.
Since one can combine forward and backward causes to a zigzag [42], a locally retrocausal
model might appear to be superluminal. However, whether such combinations are possible
depends on the model.
The above types of retrocausality and retrotemporality are properties of the mathematical
structure: a future input requirement cannot be removed by a mathematical redefinition, it is
therefore a property of a representation of a model. An example for a locally retrocausal model
might be General Relativity in space–times that have time-like closed curves.
However, there is a weaker notion of retrocausality that concerns the use of a c-model rather
than its mathematical structure, the associated m-model. We can again distinguish four cases,
which are the same as above, except that instead of a future input requirement, we merely have
a future input dependence:
Local pseudo-retrocausality: A c-model is locally pseudo-retrocausal if it fulfils Strong CoA, and
has a future input dependence.

One can similarly define the terms non-local pseudo-retrocausality, locally superluminal
pseudo-retrotemporarity and non-locally superluminal pseudo-retrotemporality, by taking the
definition without the ‘pseudo’ and replacing the future input requirement with a future input
dependence.
The reason we use the prefix ‘pseudo’ is because according to our earlier definition a future
18
input dependence is a matter of choice. It can be replaced with a past input, at least in principle.
This does not mean that a future input dependence is unimportant, however. This is because

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


it could be that removing the future input much increases the complexity of using the model.
That is, future input dependence is linked to the question of whether a model is fine-tuned
which will be further explored in the companion paper [3].
Causal and retrocausal non-locality can only be distinguished in the presence of an arrow
of time, which for all practical purposes defines a space–time slicing. The same is the case
for superluminal causal and superluminal retrocausal models—absent an arrow of time, they
cannot be told apart, since Lorentz-transformations can convert one into the other.
We want to stress that pseudo-retrocausality is a property of a c-model, but not a property of
an m-model. A retrocausal c-model cannot be temporally deterministic. A pseudo-retrocausal
c-model, in contrast, can be temporally deterministic. It follows that temporal determinism
is a simple way to tell pseudo-retrocausality from retrocausality. Note that pseudo-retrocausal-
ity was referred to just as retrocausality in Wharton and Argaman [22]. The practical use of
pseudo-retrocausality will be discussed in the accompanying paper [3].
The reader who at this point despairs over the many different types of retrocausality will
maybe understand now why the literature on the topic is so confusing, and hopefully also why
a common terminology is necessary.
Some properties about causal structure just come from the definition of the arrow of time.
In particular, we have to distinguish oriented and non-oriented arrows of time. A non-oriented
arrow of time is one that allows forming loops in time (figure 6):
Dynamical retrocausality: An m-model is dynamically retrocausal iff its retrocausality is due to a
non-oriented arrow of time.

Dynamical retrocausality may or may not be due to a space–time structure with closed
time-like curves. It is important to emphasize that dynamical retrocausality is a property of
the model, and not a property of the way the model uses inputs. Depending on how the
arrow of time is defined, it may not be particularly meaningful. What makes an arrow of time
meaningful, however, is not a question we want to unravel here.
Just for completeness, we also define:
Counter-causal: An m-model is counter-causal iff its time-reversed version is locally causal.

A model with such a property would makes one strongly suspect that the arrow of time was
just defined the wrong way round to begin with. However, possibly there were other reasons to
define an arrow of time that way.
A general comment is in order here, which is that the term ‘retrocausal’ is somewhat
linguistically confusing. It does not so much refer to causes generally going backwards, but
rather to them sometimes going against an arrow of time that was defined from something
else. That is, it is really a mix of different directions of time that mark a retrocausal model, the
already mentioned property that has previously been referred to as the possibility of zigzags in
space–time [42]. Note that the zigzag property itself is defined against the presumed-to-exist GR
arrow of time.

(e) Agent-based properties


Bell [28] further introduces local beables that are controllable, and those which are not controllable,
a distinction that we will also use below. This notion is somewhat vague, but for our purposes,
we do not need a precise definition. We will take a controllable local beable to be one whose
value can in practice be set by the action of an experimentalist—typically this is a detector
setting. For a more mathematically minded definition, see [43]. Note that for a local beable to be
19

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


Time

Time
Space Space
Figure 6. Orientable and non-orientable arrows of time. The arrows in the figure indicate the hypothetical flow of internal
time of an observer, which might differ from the coordinate time. That is, the arrows are not in all places time-like, according
to the coordinate time. This is supposed to illustrate the often-used example in which a spaceship that can travel faster than
light in one frame actually seems to go back in time in another frame. If that was possible, one could use it to construct loops
that seem to be ‘timelike’ from the point of view inside the spaceship (i.e. permissible motion for massive objects), but not
according to the coordinate time.

controllable, an experimenter need not have free will, whatever that might mean, and they also
do not have to control the local beable themselves; it could be done by some kind of apparatus.
If controllable input is correlated with observable output, we will speak of signalling.
Signalling is particularly interesting if it is outside the forward light cones.

controllable inputs which are local beables in a region A that are correlated with observable
Superluminal signalling: A c-model allows superluminal signalling iff it is superluminal and has

outputs that are local beables in a region outside L(A). An m-model allows superluminal
signalling if at least one c-model in its class does.

SQM is non-local but does not allow superluminal signalling [44].


Since the time-order of space-like separated events can change under Lorentz-transforma-
tions, superluminal signalling is usually assumed to imply the possibility of signalling back in
time. However, non-local signalling does not necessarily imply the possibility of signalling back
in time when one has a time-order given by an arrow of time. In general relativity, for example,
this may be a time-like vector field. One can then constrain non-local signals to only be forward
in time relative to the vector field. For this reason, the two phenomena—signalling non-locally
and signalling back in time—are distinct.
Let us then move on to signals that actually travel back in time:

at least one scenario for which observables at time t are correlated with required controllable
Retrocausal signalling: A c-model allows retrocausal signalling if it is retrocausal and there is

inputs from t′ > t. An m-model allows retrocausal signalling if at least one c-model in its
equivalence class does.
This retrocausality signalling could either be local or non-local. Like for the non-signalling
case discussed in §3d, causal and retrocausal non-local signalling can only be distinguished
in the presence of an arrow of time, and the same is the case for temporal and retrotemporal

The problem with retrocausal signalling is that the observables at time t are, well, observable.
superluminal signalling, since Lorentz-transformations can mix both cases.

If they can be affected by input at a later time t′ > t, then the result may disagree with what one
had already observed. This is what opens the door to causal paradoxes.
We will not introduce a notion of pseudo-retrocausal signalling, as that would be a techni-
cally possible definition, but rather oxymoronic. If a future input was removable and therefore
not necessary for an earlier observable, then no signal was sent (though in such a case an agent
might still have an illusion of signalling).
4. Specific model properties 20

In this section, we will now introduce properties that are specific to models typically used in the

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


foundations of quantum mechanics.
Using the terminology of the previous sections, we can summarize the key conundrum of
SQM by saying that models mathematically equivalent to the CI, [CI]m, do not fulfil Strong
CoA. However, SQM also has the odd property that observables generically do not have
definite values before a measurement. This opens the possibility that one may still find a
physical or empirical equivalent to the CI that fulfils Strong CoA.
Of course, one may be interested in interpretations or modifications of quantum mechanics
for other reasons. For example, one may want to find a realist interpretation, or to return
to determinism. However, the models we are here mainly concerned with are those which
reinstate Strong CoA. Such models usually introduce new input variables. Using the terminol-
ogy of Maudlin [45], we will call them additional variables:
Additional variables: Input values for an interpretation or modification of [CI]m that have no
equivalent expression in [CI]m.
Additional variables are not necessarily localized, or even localizable, and they are not
necessarily hidden, though all of that might be the case. For example, in Bohmian Mechanics,
particle positions are localizable, localized and hidden. Bohmian Mechanics, however, as we
noted previously, does not fulfil CoA.
One might worry here that since the variables are ‘additional’, they are unnecessary to
produce the same output as SQM, and therefore just make a model’s set-up reducible. However,
this does not have to be the case, because one normally introduces the additional variables to
remove another assumption from [CI]m. This is typically the measurement update postulate,
since it is the one that leads to a violation of Strong CoA [1].
The purpose of additional variables reveals itself when one interprets the violation of Strong
CoA in the CI as being due to the epistemic character of the wavefunction. One assumes that
the real (ontic) state is not the wavefunction, but one that respects CoA, one just does not know
which ontic state one is dealing with until a measurement was made. Quantum mechanics, in
this picture, is just an incomplete description of nature.
We know from Bell’s theorem that all such ensemble models with additional variables which
determine the measurement outcome will violate an assumption to this theorem commonly
known as statistical independence (SI) [46]:
SI: An m-model fulfils SI iff P λ | X, Y = P λ , where λ is (a set of) local beables on S3 and X and
Y quantify the settings of two detectors in regions A and B (compare with figure 5a).

From this, we can tell that all local interpretations or modifications of quantum mechanics can
be classified by the ways in which they violate SI.6 We will hence refer to them all as SI-violating
models.
SI is also sometimes referred to as ‘measurement independence,’ or the ‘free choice
assumption’ or the ‘free will assumption’ in Bell’s theorem. In rare occasions, we have seen
it being referred to as ‘no conspiracy’. In recent years, theories which violate SI have also been
dubbed ‘contextual’ [48], though the class of contextual models is larger than just those which
violate SI (there is more ‘context’ to an experiment than its measurement setting).
Not all models that are being used in quantum foundations reproduce all predictions of
quantum mechanics. Many of them only produce output for certain experimental situations,
typically Bell-type tests, interferometers or Stern–Gerlach devices. We can then ask, in a

6
It is sometimes argued that the Many Worlds Interpretation is a counterexample to this claim [47]. However, as laid out in
the companion paper [3], the Many Worlds Interpretation is either not empirically equivalent to SQM, or violates CoA.
D1 D2
21

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


S

Figure 7. Sketch of Cosmic Bell test with source S and two detectors. The settings of the detectors D1 and D2 are chosen by
using photons from distant astrophysical sources, usually quasars. Any past cause that could have given rise to the observed
correlations must then have been very far back in time. Light cones indicated by grey shading.

hopefully obvious generalization of the above classification, whether these models are either
representations or interpretations of [CI]m as it applies to the same experiments.
Traditionally, a distinction has been made between SI-violating models that are either
retrocausal or superdeterministic. However, this distinction has remained ambiguous for three
reasons.
First because—as we have seen already—retrocausality itself has been used for a bewilder-
ing variety of cases. If one then defines superdeterminism as those SI-violating models which
are not retrocausal, one gets an equally bewildering variety. Second, since Bell (who coined the
term ‘superdeterminism’) did not distinguish retrocausality from superdeterminism, one could
reasonably argue that superdeterminism should be equated with SI-violation in general, and
then consider retrocausality to be a variant of superdeterminism. Third, not everyone agrees
that superdeterministic models have to be deterministic to begin with [49,50].
There is no way to define the term so that it agrees with all the ways it has previously been
used. We will therefore just propose a definition that we believe agrees with the way it has most
widely been used, based on the following reasoning:
(i) We are not aware of any superdeterministic model which is not also deterministic.
Leaving aside that it is terrible nomenclature to speak of ‘non-deterministic superdetermi-
nism’, there is not even an example for it. For this reason, we will assume that superdeter-
ministic models are deterministic.
(ii) We want a model that fulfils Strong CoA, because this is the major reason why violations
of SI are interesting, and it is also the context in which Bell coined the expression.
(iii) Most of the literature seems to consider superdeterminism and retrocausality as two
disjoint cases, so we will do the same, even though Bell seems to not have used this
distinction when he coined the term. This point together, with the previous one, implies
that the model has to be locally causal.
(iv) We want a distinction that refers to the m-model, not to its particular realization as a
c-model, to avoid new ambiguities.
(v) The model should reproduce the predictions of quantum mechanics, at least to a
reasonable extent. This assumption is relevant because without it Newtonian mechanics
would also be superdeterministic which makes no sense.
(vi) It is a one-world model that violates SI.

Some readers might argue that requirement 6 is strictly speaking unnecessary because it follows
from Bell’s theorem given the previous five requirements. However, since we did not explicitly
list the assumptions to Bell’s theorem, and it is somewhat controversial which of those have to
be fulfilled in any case, we add it as an extra requirement.
Taking this together, we arrive at the following definition:
Superdeterminism: An m-model with additional variables which is deterministic, locally causal, 22
violates SI, and is empirically equivalent to SQM (as in [CI]e).

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


With this definition, a superdeterministic model is distinct from a retrocausal model. However,
a superdeterministic c-model can still be pseudo-retrocausal. This distinction has previously
been made in Nikolaev and Vervoort [51]. The authors of this article used the term ‘Soft
Superdeterminism’ for what in our nomenclature would be a pseudo-retrocausal superdeter-
ministic c-model, and they use the term ‘Hard Determinism’ which in our nomenclature would
be a not pseudo-retrocausal, superdeterministic c-model. In both cases though, the m-class
belonging to the model is superdeterministic, and not retrocausal.
The notion of common-cause superdeterminism refers even more specifically to a certain
type of superdeterministic models that are not pseudo-retrocausal. In these models, one
assumes that the additional variables which determine the measurement outcomes in a
Bell-type experiment are correlated with the settings, because they had a common cause in
the past.
We want to emphasize that the additional requirement of common-cause superdeterminism
is that there was a common cause, a more or less localized event that serves as what some would
call an ‘explanation’ for the correlation that gives rise to violations of SI. This is opposed to the
general superdeterministic models in which the correlation does not require explanation, but
rather is the explanation (for our observations). The correlation that gives rise to violations of
SI of course always has a space–time cause—in the most extreme case that would be the initial
condition at the beginning of the universe. But, in general, the hidden variables and detector
settings have no common cause in the interventionist sense, nor do they need have to have one.
That said, this particular type of common-cause superdeterminism can be tested by using
methods to choose the detector settings that are so far apart from each other that any common
cause would have had to be in the very early universe. This is the idea behind the Cosmic Bell
test [52,53], sketched in figure 7.
However, while common-cause superdeterminism is a logical possibility, we are not aware
of any superdeterministic model which is of this type, or of anyone who has advocated such a
model.
Unfortunately, the results of Cosmic Bell tests are often overstated in the literature, quite
frequently claiming to ‘close the superdeterminism loophole’, rather than just testing a
particular type of superdeterministic model, which no one works on to begin with.
The relation between pseudo-retrocausality and finetuning (a.k.a. ‘conspiracies’) will be
explored in a companion paper [3].

5. Classification
We want our classification scheme to be practically useful, so we will here provide a short guide
to how it works.
The question we want to address is this: suppose you have a c-model for quantum mechan-
ics—that is the thing you are doing your calculations with—what should you call it? We will
assume that your model is presently empirically equivalent to SQM in the non-relativistic limit.
(If it is not, you have bigger problems than finding a name for it.) You then proceed as follows:
(i) To classify the model, you first have to make sure that its set-up irreducible. If it is
reducible, the set-up cannot be classified, so please remove all assumptions that are not
necessary to calculate outputs. Assumptions that state the ‘physical existence’ (whatever
that may be) of one thing or another are typically unnecessary for any calculation.
(ii) Figure out whether your model is physically equivalent to SQM. If it is not, it is a
modification. If it is, it is an interpretation.
(iii) Figure out whether your model is mathematically equivalent to any already known
23
interpretation. If it is, your model is a representation (of the interpretation it is mathemati-
cally equivalent to). If it is not, you have a representation of a new interpretation.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


(iv) Go through the list of properties in§3 and §4 and note down whether your model has
them, starting with the m-model properties, then the c-model properties.

6. Summary
We have proposed a classification scheme for models in the foundations of quantum mechanics.
It is most central element is the distinction between different types of models: calculational,
mathematical, physical and empirical. After distinguishing these different classes of models, we
have defined some of their properties that are discussed most commonly in the foundations of
quantum mechanics, with special attention to those concerning locality and causality. We hope
that the here proposed terminology can aid to clarify which problems of quantum mechanics
can and cannot be solved by interpretation.

Data accessibility. This article has no additional data.


Declaration of AI use. We have not used AI-assisted technologies in creating this article.
Authors’ contributions. E.A.: conceptualization, formal analysis, investigation, methodology, writing—original
draft, writing—review and editing; J.H.: conceptualization, formal analysis, investigation, methodology,
project administration, writing—original draft, writing—review and editing; S.H.: conceptualization, formal
analysis, investigation, methodology, project administration, supervision, writing—original draft, writing—
review and editing; T.P.: conceptualization, formal analysis, investigation, methodology, supervision,
writing—review and editing.
All authors gave final approval for publication and agreed to be held accountable for the work
performed therein
Conflict of interest declaration. We declare we have no competing interests.
Funding. J.R.H. acknowledges support from Hiroshima University’s Phoenix Postdoctoral Fellowship for
Research, the University of York’s EPSRC DTP grant EP/R513386/1 and the UK Quantum Communications
Hub, funded by the EPSRC grants EP/M013472/1 and EP/T001011/1.
Acknowledgements. We want to thank Louis Vervoort and Ken Wharton for valuable feedback and all
participants of the 2022 Bonn workshop on Superdeterminism and Retrocausality for helpful discussion.

Appendix
If the region S3 was allowed to intersect with the inside of the past-lightcone of B, then in a

not contained on S3. In this case then, local beables at B could provide extra information for
theory which is not temporally deterministic, correlations could be created later which were

what happens at A though the information got there locally and inside the light-cones. For
illustration, see figure 5b.

References
1. Hance JR, Hossenfelder S. 2022 What does it take to solve the measurement problem? J.
Phys. Commun. 6, 102001. (doi:10.1088/2399-6528/ac96cf)
2. Adlam E. 2023 Do we have any viable solution to the measurement problem? Found. Phys.
53, 44. (doi:10.1007/s10701-023-00686-x)
3. Hossenfelder S. 2023 Quantum confusions, cleared up (or so I hope). arXiv. (doi:10.48550/
arXiv.2309.12299)
4. Frigg R, Nguyen J. 2020 Modelling nature: an opinionated introduction to scientific
representation. Cham: Springer. (doi:10.1007/978-3-030-45153-0)
5. MacFarlane J. 2020 Philosophical logic: a contemporary introduction. New York, NY:
24
Routledge. (doi:10.4324/9781315185248)
6. Gödel K. 1931 Über formal unentscheidbare Sätze der Principia Mathematica und

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


verwandter Systeme I. Monatsh. f. Math. Physik. 38–38, 173–198. (doi:10.1007/BF01700692)
7. Smith CL. 1973 High energy behaviour and gauge symmetry. Phys. Lett. B 46, 233–236. (doi:
10.1016/0370-2693(73)90692-8)
8. Pearl J. 2009 Causality. Cambridge, UK: Cambridge University Press. (doi:10.1017/
CBO9780511803161)
9. Spirtes P, Glymour C, Scheines R. 1993 Causation, prediction, and search. New York, NY:
Springer. (doi:10.1007/978-1-4612-2748-9)
10. Glymour C. 1970 Theoretical realism and theoretical equivalence. In PSA: proceedings of the
biennial meeting of the philosophy of science association, pp. 275–288, vol. 1970.
Cambridge, UK: Cambridge University Press. (doi:10.1086/psaprocbienmeetp.1970.495769)
11. Glymour C. 1980 Theory and evidence. NJ: Princeton University Press.
12. Weatherall JO. 2019 Part 1: Theoretical equivalence in physics. Philos. Compass. 14, e12592.
(doi:10.1111/phc3.12592)
13. Weatherall JO. 2019 Part 2: Theoretical equivalence in physics. Philos. Compass. 14, e12591.
(doi:10.1111/phc3.12591)
14. Argaman N. 2018 A lenient causal arrow of time? Entropy 20, 294. (doi:10.3390/e20040294)
15. Muller F. 2013 Circumveiloped by obscuritads. the nature of interpretation in quantum
mechanics, hermeneutic circles and physical reality, with cameos of James Joyce and Jacques
Derrida. https://philsci-archive.pitt.edu/10166/
16. Zurek WH. 2003 Decoherence, einselection, and the quantum origins of the classical. Rev.
Mod. Phys. 75, 715–775. (doi:10.1103/RevModPhys.75.715)
17. Wallace D. 2022 The sky is blue, and other reasons quantum mechanics is not
underdetermined by evidence. Eur. J. Philos. Sci. 13, 54. (doi:10.48550/arXiv.2205.00568)
18. SuppesP. 1961 In a comparison of the meaning and uses of models in mathematics and the
empirical sciences, pp. 163–177. Dordrecht: Springer Netherlands. (doi:10.1007/978-94-010-
3667-2)
19. van Fraassen B. 1989 Laws and symmetry. Oxford, UK: Clarendon Press. (doi:10.1093/
0198248601.001.0001)
20. Georgescu IM, Ashhab S, Nori F. 2014 Quantum simulation. Rev. Mod. Phys. 86, 153–185.
(doi:10.1103/RevModPhys.86.153)
21. Hangleiter D, Carolan J, Thébault KP. 2022 Analogue quantum simulation. Cham: Springer.
(doi:10.1007/978-3-030-87216-8)
22. Wharton KB, Argaman N. 2020 Colloquium: bell’s theorem and locally mediated
reformulations of quantum mechanics. Rev. Mod. Phys. 92, 021002. (doi:10.1103/
RevModPhys.92.021002)
23. Cavalcanti EG, Wiseman HM. 2012 Bell nonlocality, signal locality and unpredictability (or
what Bohr could have told Einstein at Solvay had he known about Bell experiments). Found.
Phys. 42, 1329–1338. (doi:10.1007/s10701-012-9669-1)
24. Adlam E. 2022 Determinism beyond time evolution. Eur. J. Philos. Sci. 12, 73. (doi:10.1007/
s13194-022-00497-3)
25. Spekkens RW. 2007 Evidence for the epistemic view of quantum states: a toy theory. Phys.
Rev. A 75, 032110. (doi:10.1103/PhysRevA.75.032110)
26. Oreshkov O, Giarmatzi C. 2016 Causal and causally separable processes. New J. Phys. 18,
093020. (doi:10.1088/1367-2630/18/9/093020)
27. Oreshkov O, Costa F, Brukner C. 2012 Quantum correlations with no causal order. Nat.
Commun. 3, 1092. (doi:10.1038/ncomms2076)
28. Bell JS. 1975 The theory of local beables. Technical report CERN-TH-2053 CERN. See https:/
/cds.cern.ch/record/980036?ln=en.
29. Brassard G, Raymond-Robichaud P. 2019 Parallel lives: a local-realistic interpretation of
“nonlocal” boxes. Entropy 21, 87. (doi:10.3390/e21010087)
30. Ciepielewski GS, Okon E, Sudarsky D. 2020 On superdeterministic rejections of settings
independence. Br. J. Philos. Sci. 74, 435–467. (doi:10.1086/714819)
31. Carroll SM, Remmen GN. 2017 A nonlocal approach to the cosmological constant problem.
Phys. Rev. D 95, 123504. (doi:10.1103/PhysRevD.95.123504)
32. Palmer TN. 2019 Bell inequality violation with free choice and local causality on the
25
invariant set. arXiv (doi:10.48550/arXiv.1903.10537)
33. Hance JR, Hossenfelder S, Palmer TN. 2022 Supermeasured: violating bell-statistical

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 480: 20230779


independence without violating physical statistical independence. Found. Phys. 52, 81. (doi:
10.1007/s10701-022-00602-9)
34. Hehl FW, Mashhoon B. 2009 Nonlocal gravity simulates dark matter. Phys. Lett. B 673, 279–
282. (doi:10.1016/j.physletb.2009.02.033)
35. Croke S. 2022 Multipartite subspaces containing no locally inaccessible information. Phys.
Rev. A 106, 032421. (doi:10.1103/PhysRevA.106.032421)
36. Hance JR, Hossenfelder S. 2022 Bell’s theorem allows local theories of quantum mechanics.
Nat. Phys. 18, 1382. (doi:10.1038/s41567-022-01831-5)
37. Maudlin T. 2012 Philosophy of physics: space and time. Princeton, NJ: Princeton University
Press. (doi:10.1515/9781400842339)
38. Price H. 1996 Time’s arrow & Archimedes’ point: new directions for the physics of time.
USA: Oxford University Press.
39. Albert DZ. 2000 Time and chance. MA: Harvard University Press. (doi:10.4159/
9780674020139)
40. Cramer JG. 1986 The transactional interpretation of quantum mechanics. Rev. Mod. Phys. 58,
647–687. (doi:10.1103/RevModPhys.58.647)
41. Kastner RE. 2012 The transactional interpretation of quantum mechanics: the reality of
possibility. Cambridge University Press. (doi:10.1017/CBO9780511675768)
42. Price H, Wharton K. 2015 Disentangling the quantum world. Entropy 17, 7752–7767. (doi:10.
3390/e17117752)
43. Walleczek J, Grössing G. 2016 Nonlocal quantum information transfer without superluminal
signalling and communication. Found. Phys. 46, 1208–1228. (doi:10.1007/s10701-016-9987-9)
44. Ghirardi GC, Rimini A, Weber T. 1980 A general argument against superluminal
transmission through the quantum mechanical measurement process. Lett. Nuovo Cim. 27,
293–298. (doi:10.1007/BF02817189)
45. Maudlin T. 1995 Three measurement problems. Topoi 14, 7–15. (doi:10.1007/BF00763473)
46. Hance JR, Hossenfelder S. 2022 The wave function as a true ensemble. Proc. R. Soc. A 478,
20210705. (doi:10.1098/rspa.2021.0705)
47. Waegell M, McQueen KJ. 2020 Reformulating Bell’s theorem: the search for a truly local
quantum theory. Stud. Hist. Philos. Sci. B Stud. Hist. Philos. Mod. Phys. 70, 39–50. (doi:10.
1016/j.shpsb.2020.02.006)
48. Kupczynski M. 2023 Contextuality or nonlocality: what would John Bell choose today?
Entropy 25, 280. (doi:10.3390/e25020280)
49. Sen I, Valentini A. 2020 Superdeterministic hidden-variables models I: non-equilibrium and
signalling. Proc. R. Soc. A 476, 20200212. (doi:10.1098/rspa.2020.0212)
50. Sen I, Valentini A. 2020 Superdeterministic hidden-variables models II: conspiracy. Proc. R.
Soc. A 476, 20200214. (doi:10.1098/rspa.2020.0214)
51. Nikolaev V, Vervoort L. 2023 Aspects of superdeterminism made intuitive. Found. Phys. 53,
17. (doi:10.1007/s10701-022-00648-9)
52. Gallicchio J, Friedman AS, Kaiser DI. 2014 Testing Bell’s inequality with cosmic photons:
closing the setting-independence loophole. Phys. Rev. Lett. 112, 110405. (doi:10.1103/
PhysRevLett.112.110405)
53. Rauch D et al. 2018 Cosmic bell test using random measurement settings from high-redshift
quasars. Phys. Rev. Lett. 121, 080403. (doi:10.1103/PhysRevLett.121.080403)

You might also like