arXiv:1412.8323v8 [quant-ph] 13 Dec 2017
Toolbox for reconstructing quantum theory from
rules on information acquisition
Philipp Andres Höhn
Perimeter Institute for Theoretical Physics, 31 Caroline Street North, Waterloo, Ontario, Canada N2L 2Y5
December 14, 2017
We develop an operational approach
for reconstructing the quantum theory of
qubit systems from elementary rules on in-
formation acquisition. The f ocus lies on an
observer O interrogating a system S with
binary questions and S’s state is taken as
O’s ‘catalogue of knowledge’ about S. The
mathematical tools of the framework are
simple and we attempt to highlight all un-
derlying assumptions. Four rules are im-
posed, asserting (1) a limit on the amount
of information available to O; (2) the mere
existence of complementary information;
(3) O’s total amount of information to be
preserved in-betwee n interrogations; and,
(4) O’s ‘catalogue of knowledge’ to change
continuously in time in-between interroga-
tions a nd every consistent such evolution
to be possible. T his approach permits a
constructive derivation of quantum theory,
elucidating how the ensuing independence,
complementarity and compatibility struc-
ture of O’s questions matches that of pro-
jective measurements in quantum theory,
how entanglement and monogamy of en-
tanglement, non-locality and, more gener-
ally, how the correlation structure of arbi-
trarily many qubits a nd rebits arises. The
rules yield a reversible time evolution and
a quadratic measure, quantifying O’s infor-
mation about S. Finally, it is shown that
the four rules admit two solutions for the
simplest case of a single elementary sys-
tem: the Bloch ball and disc as state spaces
for a qubit and rebit, respectively, together
with their symmetries as time evolution
Philipp Andres Höhn:
hoephil@gmail.com, Present address:
Institute for Quantum Optics & Quantum Information, Aus-
trian Academy of Sciences; and Vienna Center for Quan-
tum Science and Technology, University of Vienna, Boltz-
manngasse 3, 1090 Vienna, Austria
groups. The reconstruction for arbitrarily
many qubits is completed in a companion
paper [
1] where an additional rule elimi-
nates the rebit case. This approach is in-
spired by (but does not rely on) the re-
lational interpretation and yields a novel
formulation of quantum theory in te rms of
questions.
1 Introduction
Tools and concepts from information theory have
seen an ever growing number of applications in
modern physics, often proving useful for under-
standing and interpreting specific physical phe-
nomena. Among a vast number of examples,
black hole entropy, or more generally space-time
horizon entropies, can be understood in terms
of entanglement entropy [
37], thermodynamics
naturally adheres to entropic and thus informa-
tional perspectives [812], and, above all, the en-
tire field of quantum information and computa-
tion is the natural phys ical arena for applications
of information theoretic tools [
13].
This manuscript, by contrast, is motivated by
the question whether information theoretic con-
cepts, apart from their useful applications to con-
crete physical s ituations, can also tell us s ome-
thing deeper about physics, namely about the
physical content and architecture of theories. The
overriding idea is that elementary rules or restric-
tions of certain informational activities, e.g. infor-
mation acquisition or communication, should be
deeply intertwined with the structure of the ap-
propriate theory. In this article, we shall address
this question by means of the concrete example
of quantum theory. Our ambition is to develop
a novel informational framework for deriving the
formalism and structure of quantum theory for
systems of arbitrarily many qubits from elemen-
Accepted in Quantum 2017-11-27, click title to verify 1
tary operational postulates a task which is com-
pleted in the companion paper [
1]. While neither
this question nor the fact that one can recon-
struct quantum theory from elementary axioms is
new and has been extensively explored before in
various contexts [
1425], we shall approach both
from a novel constructive perspective and with
a stronger emphasis on the conceptual content
of the theory. The ultimate goal of this work is
therefore very rudimentary: to redo a well estab-
lished theory albeit in a novel way which is es-
pecially engineered for exposing its informational
and logical structure, physical content and dis-
tinctive phenomena more clearly. In other words,
we shall attempt to rebuild quantum theory for
qubit systems from scratch.
In such an information based context it is nat-
ural to follow an operational approach, describing
physics from the perspective of an observer. Ac-
cordingly, we shall work under the premise that
we may only speak about the information an ob-
server has access to in an experiment. Our ap-
proach will thus be purely operational and epis-
temic (i.e. knowledge based) by construction and
shall survive without ontic statements (i.e. ref-
erences to ‘reality’). We shall thus say nothing
about whether or not ‘hidden variables’ could give
rise to the experiences of the observer. Under
these circumstances we adopt an ‘inside view of
physics’, holding properties of systems as being
relationally, rather than absolutely defined.
Indeed, more generally the replacement of ab-
solute by relational concepts goes in hand with
the establishment of universal (i.e. observer in-
dependent) limits. For instance, the crucial step
from Galilean to special relativity is the realiza-
tion that the speed of light c constitutes a univer-
sal limit for information communication among
observers. The fact that all observers agree on
this limit is the origin of the relativity of space
and time. Similarly, the crucial step from classi-
cal to quantum mechanics is the recognition that
the Planck constant ~ establishes a universal limit
on how much simultaneous information is acces-
sible to an observer. While less explicit than in
the case of special relativity, this simple observa-
tion suggests a relational character of a system’s
quantum properties. More precisely, the process
of information acquisition through measurement
establishes an informational relation between the
observer and system. Only if there was no limit
on the acquisition of information would it make
sense to speak about an absolute state of a sys-
tem within a purely operational approach (unless
one accepts the existence of an omniscient and
absolute observer as an external standard). But
thanks to the existence of complementarity, im-
plied by ~, an observer may not access all conceiv-
able properties of the system at once. Further-
more, the observer can choose the experimental
setting and thereby which property of the system
she would like to reveal (although, clearly, she
cannot choose the experimental outcome). Under
our purely operational premise, we shall treat the
situation as if the system does not have any other
properties than those accessible to the observer at
any moment of time. In particular, the system’s
state is naturally interpreted as representing the
observer’s state of information about the system.
These ideas are in agreement with earlier propos-
als in the literature [
26,27] and, most specifically,
with the relational interpretation of qu antum me-
chanics [28 , 29].
Of course, in order for different observers who
may communicate (by physical interaction) to
have a basis for agreeing on the description of
a system, some of its attributes must be observer
independent such as its state space, the set of
possible measurements on it and possibly a limit
on its information content. But without adhering
to an external standard against which measure-
ment outcomes and states could be defined, it is
as meaningless to assert a sy stem’s physical state
to be independent of its relations to other sys-
tems as it is to relate a system’s dynamics to an
absolute Newtonian background time.
It is worthwhile to investigate what we can
learn about physics from such an operational
and informational approach. For this endeavour
we shall adopt the general conviction, which has
been voiced in many different (even conflicting)
ways before in the literature [
26, 28, 3041], that
quantum theory is best understood as an opera-
tional framework governing an observer’s acquisi-
tion of information about a system. While most
earlier works take quantum theory as given and
attempt to characterize and interpret its physical
content with an emphasis on information infer-
ence, here and in [
1] we take a step back and show
that one can actually derive quantum theory from
this perspective. This will require a focus on the
informational relation between an observer and a
Accepted in Quantum 2017-11-27, click title to verify 2
system and the rules governing the observer’s ac-
quisition of information. More precisely, our ap-
proach will be formulated in terms of the observer
interrogating a sys tem with elementary questions.
While the present work has been inspired by re-
lational ideas, we emphasize that the sequel does
not actually rely on them s o this should not dis-
courage a reader unsympathetic with relational
interpretations of quantum theory. Our approach
will reconstruct and produce a novel formulation
of quantum theory, but will clearly not single out
the relational interpretation as ‘the right one’.
However, it will support a partial interpretation,
namely that quantum theory is a law book gov-
erning an observer’s acquisition of information
about physical systems.
This is clearly not the only physical situation
to which such a relational approach applies; it
likewise opens up a novel perspective on elemen-
tary s pace-time structure which is encoded in the
informational relations among different observers
and can be exposed by a communication game.
For instance, without presupposing a particular
space-time structure and thus without assuming
an externally given transformation group between
different reference frames one can also derive
the Lorentz group as the minimal group trans-
lating between different observer’s descriptions
of physics from their informational relations, es-
tablished by communication with quantum sys-
tems [
42].
More fundamentally, relational ideas are actu-
ally required and commonly employed in the con-
text of background independent quantum gravity
approaches where the notion of coordinates dis-
appears together with a classical notion of space-
time within which a dynamics could be defined.
Instead, one has to resort to dynamical degrees
of freedom to define physical reference frames
(i.e., dynamical ‘rods’ and ‘clocks’) relative to
which a meaningful dynamics can be formulated
in the first place. This constitutes the relational
paradigm of dynamics [
4351] but goes beyond
a purely informational and operational approach
and thus clearly beyond the scope of this work.
The remainder of this manuscript is organized
as follows.
Contents
1 Introduction
1
2 Why a(nother) reconstruction of quantum theory? 5
3 Landscape of inference theories 5
3.1 The standard landscape of generalized probabilistic theories . . . . . . . . . . . . . . . 6
3.2 A novel landscape of inference theories . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2.1 Questions and answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2.2 From information to probabilities: the state of S relative to O . . . . . . . . . . 9
3.2.3 ‘Belief updating and ‘collapse’ of the state . . . . . . . . . . . . . . . . . . . . 10
3.2.4 Elementary structure on Σ and Q . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2.5 Logical compositions of questions and rules of inference . . . . . . . . . . . . . 14
3.2.6 Parametrization of S’s state and tomography . . . . . . . . . . . . . . . . . . . 16
3.2.7 Compos ite s ystems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.8 Time evolution of S’s state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.9 The theory landscape L . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4 Principles for the quantum theory of qubits as rules on information acquisition 21
4.1 The rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2 Strategy for building the necessary tools and proving the claim . . . . . . . . . . . . . 25
5 Question structure and correlations 26
5.1 A single gbit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.2 Two gbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.2.1 Logical connectives of s ingle gbit questions . . . . . . . . . . . . . . . . . . . . . 26
Accepted in Quantum 2017-11-27, click title to verify 3
5.2.2 Independence, complementarity and entanglement . . . . . . . . . . . . . . . . 28
5.2.3 A logical argument for the dimensionality of the Bloch-sphere . . . . . . . . . . 31
5.2.4 An informationally complete set for N = 2 qubits . . . . . . . . . . . . . . . . . 32
5.2.5 An informationally complete set for N = 2 rebits . . . . . . . . . . . . . . . . . 33
5.2.6 Entanglement and non-local tomography for rebits . . . . . . . . . . . . . . . . 34
5.2.7 A Bell scenario with questions: ruling out local hidden variables . . . . . . . . . 36
5.3 Three gbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.3.1 Three qubits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.3.2 Independence and compatibility for three qubits . . . . . . . . . . . . . . . . . 40
5.3.3 An informationally complete set for three qubits . . . . . . . . . . . . . . . . . 42
5.3.4 Entanglement of three qubits and monogamy . . . . . . . . . . . . . . . . . . . 43
5.3.5 Maximal entanglement for three qubits . . . . . . . . . . . . . . . . . . . . . . . 44
5.3.6 Three rebits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3.7 Independence and compatibility for three rebits . . . . . . . . . . . . . . . . . . 45
5.3.8 An informationally complete set for three rebits . . . . . . . . . . . . . . . . . . 47
5.3.9 Monogamy and maximal entanglement for three rebits . . . . . . . . . . . . . . 48
5.4 Correlation structure for N = 2 gbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.4.1 The logical mirror image of an inference theory . . . . . . . . . . . . . . . . . . 49
5.4.2 Collecting the results: odd and even correlation structure for N = 2 . . . . . . 52
5.5 The general case of N > 3 gbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.5.1 An informationally complete set and entanglement for N > 3 qubits . . . . . . 54
5.5.2 An informationally complete set and entanglement for N > 3 rebits . . . . . . . 58
6 Information measure, time evolution and state space 59
6.1 Why the Shannon entropy does not apply here . . . . . . . . . . . . . . . . . . . . . . 60
6.2 Elementary conditions on the measure . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
6.3 The state of no information is an interior state . . . . . . . . . . . . . . . . . . . . . . 61
6.4 Time evolution of the ‘Bloch vector’ . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.5 Time evolution is injective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.6 Time evolution is also reversible . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.7 Time evolution defines a group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.8 The squared length of the Bloch vector as information meas ure . . . . . . . . . . . . . 65
6.9 The set of all time evolutions is a subgroup of SO(D
N
) . . . . . . . . . . . . . . . . . . 67
6.10 Pure and mixed states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
7 The N = 1 case and the Bloch ball 68
7.1 A single qubit and the Bloch ball . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.2 A single rebit and the Bloch disc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8 Discussion and outlook 71
8.1 An operational alternative to the wave function of the universe . . . . . . . . . . . . . 73
Acknowledgements 74
References 74
The first part of the article up to and including
section
4 includes a substantial amount of con-
ceptual elaborations. The second part, by con-
trast, will become more technical upon putting
the novel postulates to use in sections
5-7. The
reconstruction for arbitrarily many qubits is per-
formed in the companion article [
1] where an ad-
ditional postulate eliminates rebits (two-level sys-
tems over real Hilbert spaces) in favour of qubit
quantum theory. The rebit case is considered sep-
arately in [
2].
Accepted in Quantum 2017-11-27, click title to verify 4
2 Why a(nother) reconstruction o f
quantum theory?
Given that we have a beautifully working theory,
one may wonder why one should bother to re-
construct quantum theory from operational state-
ments. There are various motivations for this en-
deavour:
1. To equip the standard, physically obscure
textbook axioms for quantum theory with
an operational sense. In particular, in ad-
dition to its empirical success, a derivation
from operational statements can conceptu-
ally justify the f ormulation of the theory in
terms of Hilbert spaces, complex numbers,
tensor product rule for composite systems,
etc.
2. To better understand quantum theory within
a larger context. By singling out quan-
tum theory with operational s tatements one
can answer the q uestion “what makes quan-
tum theory special?”, thereby establishing a
bird’s-eye perspective on the formalism and
conceivable alternatives.
3. It may help to understand why or why not
quantum theory in its present form should
be fundamental and thus why it should or
should not be modified in view of attempt-
ing to construct fundamental theories. By
dropping or modifying some of its defining
physical principles, one obtains a handle for
systematic generalizations of quantum the-
ory. This may also be interesting in view of
quantum gravity phenomenology (away from
the deep quantum regime). More fundamen-
tally, the question arises whether an infor-
mational perspective could be beneficial for
quantum gravity in general.
4. The hop e has been voiced that a clear inter-
pretation of the theory may finally emerge
from a successful reconstruction, in analogy
to how the interpretation of special and gen-
eral relativity follows naturally from its un-
derlying principles [
28,31,36].
It is fair to say that the hope alluded to un-
der point 4 has not been realized thus far because
the existing successful reconstructions [1425] are
fairly neutral as far as an interpretation is con-
cerned. While most of them emphasize the op-
erational character of the theory, a particular in-
terpretation of quantum theory is not strongly
suggested.
The language and concepts of the present re-
construction are different. It will emphasize and
concretize the view (or partial interpretation)
that quantum theory is a framework governing
an observer’s acquisition of information ab out the
observed s ystem [26, 28, 3041]. The new postu-
lates are simple and conceptually comprehensible,
concerning only the relation between an observer
and the system. Due to the simplicity, the ensu-
ing derivation is mathematically quite elementary
and, in contrast to previous derivations, yields
the f ormalism, state s paces and time evolution
groups explicitly and in a more constructive man-
ner. The disadvantage, compared to other recon-
structions, is that a large number of detailed steps
is required. The advantage, on the other hand, is
the simplicity of the principles and mathematical
tools, and the fact that the reconstruction affords
natural explanations for many quantum phenom-
ena, including entanglement and monogamy, and
elucidates the origin of the unitary group.
3 Landscape of inference theories
The ambition of a (re-)construction of quantum
theory is to derive its formalism, state spaces,
time evolution groups and permissible operations
from physical principles in some rough analogy
to the construction of relativity theory from the
principle of relativity and the eq uivalence prin-
ciple. But in order to formulate physical postu-
lates, we clearly have to presuppose some mathe-
matical s tructure within which a precise meaning
can be given to them. (For example, also the
construction of special and general relativity cer-
tainly presupposed a substantial amount of me-
chanical structure.)
The procedure is thus to firstly define some
landscape of theories, which hopefully contains
quantum theory and classical information theory,
but within which theories are generally not for-
mulated in terms of the usual complex Hilbert
spaces, tensor product rules, etc. The mathe-
matical formulation of the landscape must there-
fore be more elementary and, in particular, op-
erational. That is, for the time being, we have
Accepted in Quantum 2017-11-27, click title to verify 5
to forget about the usual mathematically crisp
but physically rather obscure textbook axioms
of quantum theory. While different theories will
have the mathematical and physical structure of
the landscape in common, they may have oth-
erwise very different physical and informational
properties; e.g., they may admit much stronger
correlations than quantum theory [
52], or weaker
correlations as classical probability theory, they
may allow exotic communication and information
processing tasks to be accomplished [
53,54], and
so on. Secondly, given the language of this land-
scape, one can attempt to formulate comprehen-
sible physical statements which single out q uan-
tum theory from within it. Going to a larger the-
ory landscape and beyond the language of Hilbert
spaces is precisely what allows us to ask the ques-
tion “what makes quantum theory special?” and,
ultimately, to find an operational and physical
justification for the usual textbook axioms and
the s tandard Hilbert space formulations.
The goal of this section is precisely to build
such an appropriate landscape of inference theo-
ries both conceptually and mathematically from
scratch within which we shall subsequently for-
mulate those elementary rules, governing an ob-
server’s acquisition of information about a sys-
tem, that single out qubit quantum theory. This
novel landscape of inference theories employs a
different language and is conceptually distinct
from the by now standard landscape of general-
ized probabilistic theories (GPTs) which are com-
monly employed f or characterizations (or gener-
alizations) of quantum theory. For contrast and
to put the new tools into a larger perspective, we
begin with a brief synopsis of the GPT language
before we establish the landscape of inference the-
ories underlying this manuscript.
3.1 The standard landscape of generalized
pro babilistic theories
It has become a standard in the literature to
employ the formalism of generalized probabilistic
theories (GPTs) for operational characterizations
or derivations of quantum theory [
1423, 5561].
The setup of GPTs is exclusively operational and
one considers three kinds of operation devices (s ee
figure
1): (1) a preparation device which can spit
out systems in some set of states defined by a
vector of probabilities for the outcome of fiducial
measurements. The state spaces of the systems
are necessarily required to be convex to permit
convex mixtures of states. (2) A transformation
device can perform physical operations on the
prepared systems (e.g., a rotation) which may
change the state of the system (e.g., by some
group action on the state vector), but must al-
low it to continue its journey to (3), a measure-
ment device, which detects certain experimental
outcomes. The measurement devices are math-
ematically described by so-called ‘effects’ which
are ass umed to be dual to, i.e. linear functionals
on, the states. (For the interested reader we note,
however, that Holevo has shown how the math-
ematical incarnation of measurements as linear
functionals on states follows from other simple
operational as sumptions [
62].)
PSfrag rep lacements
preparation
transformation
measurement
systems
convex state sp aces
cbit square bit rebit qubit
‘effects’
dual t o states
Figure 1: The standard op e rational setup of generalized
probabilistic theories with examples of allowed convex
state spaces of elementary two-level systems.
Measurements and states are at the heart of
GPTs; ‘effects’ directly determine the outcome
probabilities of measurements and thereby, when
a complete set is engaged, reveal the probabilistic
state of a system. An observer assumes a support-
ing role, giving intuitive meaning to the notion
of preparation, transformation and measurements
of sy stems. The reconstructions of quantum the-
ory within the GPT formalism depart from rather
global operational axioms, restricting the sets of
possible preparations (i.e. state spaces), transfor-
mations and measurements. The concrete acqui-
sition of information of the observer about the
systems is otherwise not accentuated. In par-
ticular, since the outcome probabilities are the
primary concept of GPTs, these axioms say lit-
tle about what an observer experiences in indi-
vidual experimental runs, but instead f ocus on
the totality of a large number of experimental
runs. This has led to a whole wave of s uccessful
Accepted in Quantum 2017-11-27, click title to verify 6
quantum theory reconstructions, employing GPT
concepts in one way or another [
1423], most of
which, however, are fairly abstract on account of
the rather global axioms.
It is perhaps one of the great strengths of GPTs
that they constitute a functional and purely op-
erational framework which is interpretationally
fairly neutral. It is not the ambition of this frame-
work to elucidate the measurement problem, to
clarify what happens to a state during a mea-
surement, what probabilities are or, ultimately,
how to interpret quantum mechanics (except that
it highlights its operational character). As such,
this framework is compatible with most interpre-
tations of quantum theory.
3.2 A novel landscape of inference theories
The success of G PTs notwithstanding, we shall
now change semantics and perspective to define a
new landscape of theories and a novel framework
for (re)constructing and understanding quantum
theory. Henceforth, we shall fully engage the
observer and give primacy to his acquisition of
fundamentally limited information from observed
systems. This will include being more explicit
about what general sort of information will be
available to the observer even in individual exper-
imental runs. Probabilities, on the other hand,
can be viewed as secondary and as a consequence
of the limited information available to the ob-
server although, clearly, probabilities will as-
sume a pivotal role too (after all we want to re-
construct quantum theory).
3.2.1 Questions and answers
As schematically depicted in figure
2, we shall
consider an observer O who can only interact with
a system S through interrogation via questions
Q
i
from some set of questions Q which we shall
further constrain below. (At this stage we make
no assumption about whether Q is continuous or
discrete.) The only information which we allow
O to acquire about S is by asking questions from
this set Q.
In principle, of course, O could conceive of and
ask S all kinds of questions, but S could not al-
ways give a meaningful answer; S may simply
not have the desired properties or carry the in-
formation O is inquiring about (e.g., S may not
be complex enough, or S does not interact with
other systems carrying the information in ques-
tion), or p ossibly the question is not a senseful
one in the first place. We shall call a question Q
physically implementable on S if O can acquire a
‘meaningful’ answer from S to Q. An elementary
restriction on Q is that any question Q
i
from this
set be implementable on S and that, whenever O
asks Q
i
to S, S will give an answer to O. Clearly,
Q depends on S.
But how does O know whether S will give an
answer and how can he judge whether the latter
is ‘meaningful’? This requires O, like any experi-
menter, to have developed, from previous experi-
ences, a theoretical model by means of which he
describes and interprets his interactions with S.
An answer can only be ‘meaningful’ in the con-
text of this model such that our notion of imple-
mentability is actually dependent on O’s model.
We shall come back to this model frequently.
The central ingredients of this framework will
thus be questions and answers and O’s informa-
tion ab out answer outcomes and their relations.
This will lead to a novel question calculus from
which many crucial q uantum properties will be
derived. That is, rather than focusing mostly
on probabilistic properties as in GPTs, this novel
framework will connect more directly to what can
be measured in experiments through its empha-
sis on questions and their relations. As a conse-
quence, this framework will also produce opera-
tionally more compelling explanations of typical
quantum phenomena.
In the s eq uel, we shall solely speak about the
information O has about S and, correspondingly,
about the state that O assigns to S based on
this information. Such a state of S is then de-
fined relative to O (whether or not some hidden
variables give rise to this state is a question we
shall not address). The act of information ac-
quisition establishes a relation between O and S
and this will be the center of our attention. (A
priori, a different observer O
may establish a
different relation with S.) Although this frame-
work will also not give rise to a unique interpre-
tation, it will support the partial interpretation
that quantum theory is a law book governing an
observer’s acquisition of information about phy s-
ical systems and might therefore be considered
interpretationally less neutral than GPTs. While
this connects with general ideas underlying, e.g.,
the relational [
28,29], Brukner-Zeilinger informa-
Accepted in Quantum 2017-11-27, click title to verify 7
PSfrag rep lacements
Preparation Interrogation
S
O
Q
i
?
Figure 2: Schematic representation of an observer O interrogating a system S.
tional [
3135], or QBist [3639] interpretations of
quantum mechanics, we emphasize that this re-
construction does not rely on them and should
thus not discourage readers uncomfortable with
these specific interpretations.
We shall not explicitly deal with transforma-
tion and measurement (‘effect’) devices as in
GPTs; instead, these will be replaced more gener-
ally by time evolution and questions, respectively.
The s et of all possible (information preserving)
operations that O could perform on S can later
be identified with the set of all poss ible time evo-
lutions of S. H owever, in analogy to GPTs, we
will assume that O has access to some method of
preparing S in different ways such that S’s an-
swers to O’s q uestions depend on the precise way
of preparation. (O could either control himself a
preparation device or have a distinct observer O
prepare systems for him.)
When constructing the new landscape L of the-
ories describing O’s acquisition of information
about S within which we shall later, in section
4,
formulate our postulates, we will make a number
of restrictions and assumptions. In order to fa-
cilitate f uture generalizations and improvements
of the present construction of quantum theory
which constitutes a proof of principle we shall
attempt to be as clear as possible about the as-
sumptions made throughout this work.
As a starter, we would like to keep Q as sim-
ple as possible, while still having non-trivial ques-
tions. In particular, we do not wish to consider
trivial propositions which are always true or al-
ways false. We shall therefore assume the follow-
ing.
Assumption 1. The set of questions Q which
we shall permit O to ask S only contains binary
questions Q
i
. Any Q
i
Q is a non-trivial ques-
tion such that S’s answer (‘yes’ or ‘no’) is not
independent of i ts preparation. Furthermore, any
Q
i
Q is repeatable such that O, b y asking the
same S the same Q
i
m times in succession will
receive m times the same answer.
The restriction to elementary ‘yes-no’-
questions greatly simplifies the discussion and,
ultimately, will give rise to the quantum theory
of qubit systems. For instance, in quantum
theory, a binary question could be ‘is the spin of
the qubit up in x-direction?’ However, it will not
be too difficult to generalize Q to also consist of
ternary, quaternary, quinary, etc. questions, but
we shall not attempt to do so here. Since trivial
questions are not considered, we already see
that not even all implementable questions will
be taken into account. We shall impose further
restrictions on Q such that, ultimately, it will be
a strict subset of all possible binary questions
which O could, in principle, ask S.
Since this is an operational approach, it is fair
to assume that O can record the answers to his
questions asked to any system (e.g., by writing
them on a piece of paper) and that he can do
statistics over the outcomes (e.g., by counting the
frequency of outcomes). We shall require that
every possible way of preparing S will give rise
to a particular s tatistics over the answers to all
Q
i
Q; O could test these statistics by interro-
gating a large number n of identically prepared
systems
1
S
a
, a = 1, . . . , n, s ufficiently often with
1
Given the present structure, two systems S
1
and S
2
could only be considered as distinct in nature if either t he
(maximal) set of questions Q which O can ask the sys-
Accepted in Quantum 2017-11-27, click title to verify 8
(at least ideally) all Q
i
Q. In fact, this is pre-
cisely how O will operationally distinguish differ-
ent preparations of systems.
By having interrogated, in this manner, the
n always identically prepared S
a
for all possible
ways of preparation, we shall assume O to have
gained a ‘stable’ knowledge of the set Σ of all pos-
sible answer statistics for S over Q, i.e., the an-
swer statistics for all Q Q and all possible ways
of preparing a system S.
2
While ideally n is
necessary, ‘practically’ n should be large enough
for O to develop a theoretical model of both Q
and Σ up to some accuracy which agrees with
his observations. It is not our ambition to clarify
further what n is nor how precisely O has de-
veloped his theoretical model, instead, we shall
henceforth just assume that O has puzzled out
the pair (Q, Σ). As any experimenter in an ac-
tual laboratory, O shall interpret the outcomes
of his interrogations by means of his model for
(Q, Σ) and he can decide whether a given ques-
tion is contained in the set Q or not. Henceforth,
we assume that O only asks questions from Q.
Our task will be to establish what this model is,
based on the ensuing assumptions and postulates.
3.2.2 From information to probabilities: the state
of S relati v e to O
The previous subsection did not refer to the no-
tion of probabilities. But it defined two prepa-
rations as being identical if they give identical
statistics. This permits us to regard Σ equiva-
tems or the totality of answer statistics for all p ossible
preparations were distinct for S
1
and S
2
(if always the
same two S
1
, S
2
are interrogated and thereafter freshly
prepared again). If O can not distinguish S
1
and S
2
in this
way for sufficiently many trials (ideally infinitely many),
we shall call them identical. Let S
1
and S
2
be identi-
cal and O prepare both systems with the same procedure
(for instance, the setting of the preparation device is the
same for both systems). If the answer statistics (frequ en-
cies) for S
1
, S
2
for all Q
i
Q become indistinguishable af-
ter sufficiently many trials (of preparing and interrogating
the same systems with the same procedure), we consider
these systems as ‘identically prepared’. (After completion
of this work, the author was made aware that this notion
is similar to the definition of ‘operational equivalence’ put
forward in [
63].)
2
We assume the preparation method to be ideal in the
sense that it accounts for any possible answer statistics
which S can admit in O’s world for questions in Q. That
is, there do not exist other methods which can prepare S
in ways th at O’s method does not en compass.
lently as the set of all possible answer statistics
or as the set of all (distinct) preparations.
Based on this notion, we shall now identify
probabilities as degrees of belief. The ‘knowledge’
of what Σ is for a given S will permit O to assign
probabilities to the outcomes of his questions. It
is very natural for O to assign probabilities to
questions because he deals with statistical fluctu-
ations and furthermore, as we shall see later, with
systems about which he always has incomplete in-
formation in the sense that the corresponding Σ
is such that he can never know the answers to all
Q
i
Q at the same time.
More precisely, for a s pecific S and any Q
i
Q
that he may ask the system next, O can assign a
probability y
i
that the answer will be ‘yes’ (or a
probability n
i
that the answer will be ‘no’), ac-
cording to
(i) O’s knowledge of Σ, and
(ii) any prior information that O may have about
the s pecific S.
Given our setup, the only prior information that
O may have about the particular S (apart from
what the associated Σ and Q are) must result ei-
ther from having interrogated some ensemble of
systems identically prepared to S with some sub-
set of Q beforehand
3
and from the corresponding
accumulated statistics of the asked Q
j
Q (e.g.,
that O may have recorded on a piece of paper),
or from any other prior information about the
method of preparation. In particular, this previ-
ous ensemble could have been empty.
For instance, if every time that O asked the
specific Q
i
to any of the identically prepared sys-
tems gave a ‘yes’ answer before, he will assign
the prior probability y
i
= 1 to Q
i
and to the next
identically prepared S that he will interrogate. If,
on the other hand, the number of ‘yes’ and ‘no’
answers to some other Q
j
was equal for the previ-
ously identically prepared systems, O will assign,
as a best guess, the prior probability y
j
=
1
2
to
this Q
j
and to the next S. Similarly, for any
other answer statistics, O would assign y
i
to the
next S depending on the recorded frequencies of
‘yes’ answers. But, thanks to his knowledge of
Σ and therefore of any possible relations in the
3
If O asks more than one question to any S, the order-
ing of the questions may matter. But then O could ask
the questions for any S always in the same order.
Accepted in Quantum 2017-11-27, click title to verify 9
answer statistics, O can also assign prior proba-
bilities y
k
to questions Q
k
that he did not ask the
previous set of identically prepared systems. For
example, Σ may b e such that whenever S gives
a ‘yes’ answer to Q
i
, it will give a ‘no’ answer to
an immediately following Q
k
. Accordingly, if O
assigns a prior probability y
i
= 1 to Q
i
as above,
he will also assign a prior y
k
= 0 to Q
k
without
previously having asked Q
k
. But other relations
between questions will be permitted too. In par-
ticular, it may be that the information gained
from the questions he previously asked the iden-
tically prepared systems and the structure of Σ
make it equally likely that the answer to Q
k
asked
to the next S will be ‘yes’ or ‘no’. In this case,
O will assign y
k
=
1
2
to Q
k
that he may ask the
next S. This will become more precise along the
way.
We therefore take a broadly
4
Bayesian perspec-
tive on probabilities: O assigns probabilities to
questions according to his ‘degree of belief about
S. These probabilities y
i
are thereby relative to
the observer O. A different observer O
may have
different information about S and thereby assign
different probabilities to the various outcomes of
questions posed to S (for a discussion, within
quantum theory, of the consistency of different
observers having different information about a
system, see [
28,29,64,65]). E.g., O
could be the
one preparing S. She could ‘know’ the statistics
for the specific preparation setting (from previous
tests) and then send O the specifically prepared
S without informing him about her knowledge.
For consistency, we tacitly assume the set Σ of
all possible answer statistics to coincide with the
set of possible ‘beliefs’. That is, to every equiva-
lence class of preparations of S (with ‘identically
prepared’ defining the equivalence relation) there
shall correspond a unique ‘belief {y
i
}
Q
i
∈Q
and
vice versa. Since the only way for O to acquire in-
formation about S is by interrogation with ques-
tions in Q, the prior probabilities y
i
that O as-
signs to every Q
i
Q encode the entire informa-
tion that O has about S. Hence, we shall make
the f ollowing identification.
Definition 3.1. (State of S relative to O)
The collection of all probabilities y
i
Q
i
Q is
the state of S relative to O. Accordingly, the set
4
We add the qualifier ‘broadly’ here since we also allow
for the typical laboratory situation of ensembles of systems
(which may or may not contain more than one element).
Σ of all possible answer statistics on Q which S
admits is the state space of S.
Of course, ultimately not all y
i
will be inde-
pendent such that the full collection of probabili-
ties will yield a redundant parametrization of the
state. However, this is not imp ortant for the mo-
ment and we shall come back to this shortly.
This definition of the state of a system S ex-
plicitly identifies it with the ‘state of informa-
tion’ that O has acquired about S; O assigns
this state to S according to his information about
the Q
i
Q. As such, the state of system S
is epistemic (i.e. a ‘state of knowledge’) and a
priori only meaningful relative to the observer
O. The interpretation of the quantum state as
a ‘state of information’ is certainly not new and
has been proposed in various ways before (see
also, e.g., [
3638, 40]). However, the above def-
inition is closest in spirit to the ideas underly-
ing Relational Quantum Mechanics [28, 29] and
the Zeilinger-Brukner interpretation [
15, 3134]
and thereby generalizes them to the landscape of
theories describing O’s acquisition of information
which we are in the process to establish.
While the state of S is thus a priori only mean-
ingful relative to O, we emphasize that both the
set of questions Q which O may ask and the s tate
space Σ are to be intrinsic to the system S. Oth-
erwise, it would be difficult for two observers to
agree on the description of a given S.
3.2.3 ‘Belief updating and ‘collapse’ of the state
At this stage it is important to distinguish single
from multiple shot interrogations. In a
single shot interrogation O interrogates a
single s ystem S, in some prior s tate, with
a number of questions from Q without in-
termediate re-preparations of S. The defi-
nite answers to these questions give O def-
inite information about this specific S after
the interrogation. Furthermore, his knowl-
edge of Σ and any prior knowledge of S
(acquired through previous interrogations of
identically prepared systems) give him statis-
tical information about any questions he did
not ask S. In conjunction, the new answers
and his prior knowledge thus determine the
state of S after the interrogation. This will
constitute a posterior state update rule, and
thereby a ‘belief update form a prior to a
Accepted in Quantum 2017-11-27, click title to verify 10
posterior s tate for a single system which we
shall turn to shortly (and which clearly de-
pends on the specific way O interrogates S).
This posterior state of S will reflect O’s defi-
nite information about every asked question
Q
i
by featuring either y
i
= 0 or y
i
= 1 due to
repeatability, depending on whether the an-
swer was ‘no’ or ‘yes’, respectively (assuming
for now, of course, that Σ is such that the
answers to the selection of questions that O
asked can be known simultaneously).
If this posterior state does not coincide with
the prior state that O assigned to S before
the interrogation, based on his prior infor-
mation about S, then S’s s tate has ‘col-
lapsed’ relative to O during the interroga-
tion. Hence, a state ‘collapse’ only occurs
if O’s posterior information about S does
not coincide with his prior information about
S, i.e. if O experienced an information gain
about S via the interrogation. We shall
therefore view a state ‘collapse’ as O’s infor-
mation gain about this specific S rather than
a ‘disturbance’
5
of S (we refer the reader also
to [
26,60,66,67] for a related discussion).
mult iple shot interrogation O interrogates
an ensemble of identically prepared systems
S
a
, a = 1, . . . , n, where the interrogation
of every S
a
is a single shot interrogation.
O will carry out such a multiple shot
interrogation to do state tomography, i.e.
to estimate the state of the ensemble {S
a
}
for the specific setting of preparation or, in
other words, the state of any of the systems
prior to being interrogated by O.
This will also be a ‘belief updating, how-
ever, a prior (or ensemble) state updating.
After every interrogation of a system in the
ensemble, O will assign probabilities y
i
to the
Q
i
in the manner described above. This will
then define the prior state of the next sys-
tem in the ensemble to be interrogated. By
interrogating more and more systems, O will
gain more and more information about the
5
A ‘disturbance’ of the system S is only meaningful if
there was an underlying ontic state (i.e. ‘state of reality’)
to which, however, O would have no access. Here we shall
merely speak about the information th at O has access to
and therefore not make any ontic statements, regarding
them as excess baggage for our purposes.
ensemble state such that his ass ignments of
the y
i
will fluctuate less the larger the num-
ber of interrogated systems. This process
gives rise to an ensemble state updating. In-
dependent of this updating, the prior state
that O assigns to any individual system may
experience a ‘collapse’ during the interroga-
tion of that specific system because his infor-
mation about the specific system may have
changed. Accordingly, O will have to distin-
guish the ensemble state from the posterior
state of any system in the ensemble. But the
collection of posterior states determines the
ensemble state.
3.2.4 Elementary s tructure on Σ and Q
The present structure is still too rudimentary for
a quantum (re)construction. We therefore need
more.
Firstly, as in GPTs we will need O to be able to
assign a single prior state to any pair of identical
systems whenever he flips a biased coin in order
to decide which of the two systems he will in-
terrogate. (Equivalently, the preparation method
could involve a biased coin toss , the outcome of
which determines the preparation setting.) That
is, O will be permitted to build convex combina-
tions of states.
Assumption 2. The state space Σ of S is a
closed convex set.
Closure of Σ is assumed because points lying in
the boundary of Σ can be arbitrarily well approx-
imated by states in the interior. Perfect prepa-
ration and arbitrarily good preparation are op-
erationally indistinguishable and so adding the
boundary does not change the physical predic-
tions [
16].
Next, we need to define some elementary struc-
ture on Q in order to meaningfully speak about
relations among questions (and answers). To this
end, we shall establish additional structure on
the pair (Q, Σ). We declared in assumption
1
that there are to be no trivial questions in Q the
answers to which would be independent of S’s
preparation. We also insisted before that O will
distinguish the different ways of preparing a sys -
tem S and thus the different states he can assign
by the particular answer statistics. We shall
now strengthen these requirements, by asserting
Accepted in Quantum 2017-11-27, click title to verify 11
that there exists a distinguished state of ‘no infor-
mation’, corresponding to the situation that O’s
prior knowledge about S makes it equally likely
for him that the answer to any Q Q is ‘yes’ or
‘no’.
6
We note that this ass umption is a restric-
tion on O’s model for the pair
(Q, Σ).
7
Assumption 3. There exists a special state in
Σ, called the state of no information, which is
given by y
i
=
1
2
, Q
i
Q.
This state shall be the prior state that O as-
signs to S (or an ensemble thereof) in a belief
updating whenever he has ‘no prior information’.
For example, a distinct observer O
may prepare
a system S and send it to O in such a way that
the latter knows only the associated Σ but not
the preparation setting. In this case, as a prior,
O will assign the s tate of no information to S,
i.e. y
i
=
1
2
, Q
i
Q. Similarly, there will exist
a special preparation setting which is such that a
multiple shot interrogation on an ensemble pre-
pared in this setting will give totally random an-
swers to O such that he will assign the state of no
information to the ensemble. But note that the
state of any individual system after the interroga-
tion of that system will not be the state of no in-
formation because, through the interrogation, O
will have acquired information about that specific
system (see the discussion of the state ‘collapse’
above).
Given the state of no information, we shall pre-
liminarily quantify the amount of information α
i
that the definite answer to any Q
i
Q defines as
one bit. Similarly, whenever y
i
=
1
2
, as in the
state of no information, we shall say that O has
α
i
= 0 bits of information about Q
i
. In general,
under the premise that information can neither
be negative nor complex, O’s information about
6
We emphasize that GPTs are more general by, in prin-
ciple, permitting state spaces which do not contain such
a distinguished state. However, most operationally inter-
esting GPTs do possess such a state.
7
For instance, we note that on account of the exis-
tence of binary POVMs with an inherent bias, such as
(E
0
= 2/3 · 1, E
1
= 1/3 · 1), the pair given by Q =
{binary POVMs} and Σ = {unit trace density matrices}
cannot satisfy this condition because no unit t race
density matrix exists which yields probability 1/2 for
(E
0
, E
1
). This pair will therefore ultimately not be the
solution of this reconstruction. However, a subset of
{binary POVMs} together with the full quantum state
space will be the solution.
Q
i
should satisfy
0 bit α
i
1 bit. (1)
We shall not propose an explicit information mea-
sure α
i
(as a function on Σ) here because this
must follow from the rules on information acqui-
sition postulated b elow and we shall indeed derive
it therefrom later in section
6.8. Until then it will
be sufficient to work with this implicit notion of
quantify ing O’s information about any Q
i
.
But the questions O can ask, and the informa-
tion that the corresponding answers define, may
not be independent. However, a priori the no-
tion of independence of questions, in the sense of
stochastic independence, is state dependent. For
example, for a pair of qubits in quantum theory
the questions Q
1
, ‘is the spin of qubit 1 up in x-
direction?’, and Q
2
, ‘is the spin of qubit 2 up in
x-direction?’, are stochastically independent rel-
ative to the completely mixed state, but fully de-
pendent relative to an entangled state (with cor-
relation in x-direction). As a result of the state
dependence, the independence of questions may
also be viewed as observer dependent. For in-
stance, in quantum theory an observer O
could
send O an entangled pure state (with correlation
in x-direction) and refuse to tell O which state it
is. Relative to O
, Q
1
and Q
2
will be dependent,
but they will be independent relative to O be-
cause the latter will ass ign the completely mixed
state to the pair prior to measurement.
Since the notion of independence of q uestions
is state dependent, we need a distinguished state
in order to unambiguously define it this is the
second purpose of assumption
3 and brings us in
contact with a state update rule. Indeed, suppose
O acquires S in the state of no information and
poses the q uestion Q
i
Q. By assumption
1,
any Q
i
Q is repeatable such that, upon receiv-
ing either the answer ‘yes’ or ‘no’, O will assign
y
i
= 1 or y
i
= 0, respectively, as the probability
for S giving the answer ‘yes’ if posing Q
i
again.
The posterior state update rule, which enables O
to update his information about a specific S in
compliance with the given answers, must respect
this repeatability. It depends on the details of this
update rule what the probabilities y
j
for all other
Q
j
Q is after having asked only Q
i
. Rather
than fully specifying at this stage what this rule
is, we shall simply assume that O employs one
which is consistent. Whatever this posterior state
update rule, we shall refer to Q
i
, Q
j
Q as
Accepted in Quantum 2017-11-27, click title to verify 12
independent if, after having asked Q
i
to S in
the state of no information, the probability
y
j
=
1
2
. That is, if the answer to Q
i
relative
to the state of no information tells O ‘noth-
ing’ about the answer to Q
j
. We shall require
this relation to be symmetric, i.e., Q
i
is in-
dependent of Q
j
if and only if Q
j
is indepen-
dent of Q
i
.
8
This is equivalent to saying that
Q
i
, Q
j
are stochastically independent with
respect to the state of no information, i.e. the
joint probabilities factorize relative to the
latter, p(Q
i
, Q
j
) = y
i
· y
j
=
1
2
·
1
2
=
1
4
. Here
p(Q
i
, Q
j
) = p(Q
j
, Q
i
) denotes the probabil-
ity that Q
i
and Q
j
give ‘yes’ answers if asked
in sequence to the same S arriving in the
state of no information and y
i
, y
j
are the
individual ‘yes’-probabilities in this distin-
guished state.
9
dependent if, after having asked Q
i
to S in
the state of no information, the probability
y
j
= 0, 1. That is, if the answer to Q
i
rel-
ative to the state of no information implies
also the answer to Q
j
. Again, we require this
relation to be symmetric. This is equivalent
to saying that, relative to the state of no in-
formation, Q
i
, Q
j
are stochastically fully de-
pendent as either p(Q
i
, Q
j
) = y
i
= y
j
=
1
2
= p(¬Q
i
, ¬Q
j
) or p(Q
i
, ¬Q
j
) = y
i
= y
j
=
1
2
= p(¬Q
i
, Q
j
), where ¬Q is the negation of
Q.
10
partially dependent if, after hav ing asked Q
i
to S in the state of no information, the prob-
ability y
j
6= 0,
1
2
, 1. That is, if the answer
to Q
i
relative to the state of no information
8
It should be noted that, while this is true for projec-
tive measurements in quantum theory, it does not hold
for generalized measurements. I thank Tobias Fritz for
pointing this out.
9
We emphasize that this is a definition of maximal ind e-
pendence. For example, for a qubit two linearly indepen-
dent directions ~n
1
, ~n
2
in the Bloch sphere with ~n
1
· ~n
2
6= 0
would d efine two spin observables ~n
1
· ~σ and ~n
2
· ~σ whose
corresponding projectors would not be maximally inde-
pendent according to this definition. The corresp onding
questions would be partially d ependent (see below) be-
cause whenever the observer knows the answer to one
question, the probability for a ‘yes’ answer to the other
would be distinct from
1
2
.
10
The most trivial example of a pair of dep endent ques-
tions are clearly Q and ¬Q. But there exist less trivial
ones. E.g., see theorem
5.7 below.
gives O partial information about the answer
to Q
j
. Again, this relation is required to be
symmetric. This is equivalent to saying that
the joint probabilities p(Q
i
, Q
j
) relative to
the state of no information do not factorize
and the answer to one question does not fully
imply the answer to the other.
We shall henceforth assume that Σ is such that
two questions Q
i
, Q
j
which are fully or partially
dependent relative to the state of no information
are also fully or partially dependent relative to
any other state in Σ. These definitions of (in-
)dependence depend a priori on the update rule
through the j oint probabilities. For our purposes
it will turn out not to be necessary to fully specify
this update rule.
Next, we need a notion of compatibility and
complementarity. Q
i
, Q
j
are
maximally compatible if O may know the an-
swers to both Q
i
, Q
j
simultaneously, i.e. if
there exists a state in Σ such that y
i
, y
j
can
be simultaneously 0 or 1.
maximally complementary if maximal infor-
mation about the answer to Q
i
forbids O to
have any information about the answer to Q
j
at the same time (and vice versa). That is,
every state in Σ which features y
i
= 0, 1 will
necessarily have y
j
=
1
2
(and vice versa).
Consequently, complementary questions are in-
dependent, but independent questions are not
necessarily complementary. Finally, Q
i
, Q
j
are
partially compatible (or complementary) if max-
imal information about one precludes maximal,
but permits non-maximal information about the
other.
This permits us to further constrain the up-
date rule. Firstly, maximal complementarity has
an obvious consequence for an update rule. Sec-
ondly, we shall ass ume the following.
Assumption 4. If Q
i
, Q
j
are maximally com-
patible and independent then asking either shall
not change O’s information about the other, re-
gardless of S’s state. That is, asking Q
i
must
leave y
j
invariant and vice versa for i, j inter-
changed.
This is to prevent O from losing or gaining in-
formation about some question by asking another
Accepted in Quantum 2017-11-27, click title to verify 13
question which is compatible with but indepen-
dent of the first. These constraints on the update
rule turn out to be sufficient for our reconstruc-
tion.
Assumption
4 leads to a first result which we
shall use at times: it implies what sometimes is
referred to as ‘Specker’s principle’ [68] (see also
[
6972]). We shall call Q
1
, . . . , Q
n
Q mutually
maximally compatible if there ex ists a state of S
where the answers to all of Q
1
, . . . , Q
n
are known
simultaneously to O.
Theorem 3.2. (‘Specker’s principle’) If
Q
1
, . . . , Q
n
Q are pairwise maximally compat-
ible and pairwise independent then they are also
mutually maximally compatible.
Proof. Let Q
1
, . . . , Q
n
Q be pairwise indepen-
dent and pairwise maximally compatible. On ac-
count of repeatability, after asking S the ques-
tion Q
1
, O will assign th e probability y
1
to be
either 0 or 1, depending on the answer. O may
subsequently ask S the question Q
2
upon w hich
the probability y
2
will likewise be either 0 or 1.
According to assumption
4, this will not change
y
1
because, by assumption, both are maximally
compatible and independent. By the same ar-
gument, upon next asking and receiving an an -
swer to Q
3
, O will assign either y
3
= 1 or
y
3
= 0, while both y
1
, y
2
are unchanged. Re-
peating this argument recursively, it is clear that
O can thereby generate a state of S in which all
y
i
, i = 1, . . . , n, are either 0 or 1 which represents
a state of S where O knows the answers to all of
Q
1
, . . . , Q
n
.
Notice that this property is satisfied for classi-
cal bit theory and in quantum theory for projec-
tive, however, not for generalized measurements.
3.2.5 Logical compositions of questions and rules
of inference
Given multiple questions, nothing, in principle,
stops O from composing them via logical con-
nectives to all kinds of propositions which them-
selves constitute questions. The issue is, however,
whether he will be able to get an answer from S to
such a question, i.e. whether it is implementable
on S and thus whether it could b e contained in
Q (see section
3.2.1). For example, given any two
questions Q
i
, Q
j
Q, O could consider the cor-
relation question Q
ij
, ‘are the answers to Q
i
, Q
j
the same?’. Do we allow Q
ij
to also be imple-
mentable on S? Clearly, if Q
i
, Q
j
are maximally
compatible, then Q
ij
is implementable because O
can always find the answer to the latter by asking
Q
i
, Q
j
. In this case, since Q
i
, Q
j
are simultane-
ously defined relative to O, we can also write
Q
ij
:= Q
i
Q
j
, (2)
where is the logical biconditional or
(XNOR).
11
Q
ij
will then automatically be maxi-
mally compatible with both Q
i
, Q
j
.
Let us now contrast this with the situation in
which Q
i
, Q
j
are (partially or maximally) comple-
mentary. The structure introduced thus far does
not preclude the ‘correlation question’ Q
ij
to also
be implementable and, ultimately, contained in Q
even if Q
i
, Q
j
are complementary. In fact, it is not
possible to decide without a theoretical model f or
describing and interpreting (Q, Σ) whether Q
ij
is
implementable on S in the case that Q
i
, Q
j
are at
least partially complementary. What, however, is
unambiguously clear is that for all practical pur-
poses Q
ij
will not be implementable in this case.
Namely, Q
ij
would be a statement about the cor-
relation of two complementary questions Q
i
, Q
j
and O could say ‘the answers are the same’, but
he can never directly test them individually and
see that they are actually ‘the same’ at the same
time. Indeed, Q
i
, Q
j
, Q
ij
would need to form
a mutually (partially or maximally) complemen-
tary set. Thus, from a purely operational per-
spective alone, O could never tell, even in princi-
ple, that Q
ij
was implementable; he simply can-
not get an answer to Q
ij
which he could inter-
pret, based on operationally accessible informa-
tion alone, as a correlation of the complementary
Q
i
, Q
j
.
It thus depends on the model for (Q, Σ )
whether Q
ij
is implementable when Q
i
, Q
j
are
complementary. Since it is impossible for O to
settle this question op erationally, it would re-
quire a mo del which employs hidden and opera-
tionally inaccessible ontic information and which
is devoid of complementarity at an ontic level
to conclude that Q
ij
Q in this case. For ex-
ample, Spekkens’ elegant toy model [
40] and the
‘black boxes’ of [
54] employ ontic states, satisfy
the structure established thus far (at least at the
11
That is Q
ij
= ‘yes’ if Q
i
= Q
j
= ‘yes’ or ‘no’ and
Q
ij
= ‘no’ otherwise.
Accepted in Quantum 2017-11-27, click title to verify 14
epistemic level and modulo restrictions on the no-
tion of convexity), and explicitly feature such a
triple of q uestions.
12
However, such a model is
in conflict with our premise of following a purely
operational approach which only speaks about in-
formation that O has access to via direct inter-
rogation. The model by means of which O in-
terprets the answers that he gets from S and,
hence, the information he can acquire about S
shall be based entirely on operational statements
that O can, in principle, check through interroga-
tion, and not on propositions that require hidden
and inaccessible ontic information. Accordingly,
O’s theoretical model for Q should not contain
(a question which is logically equivalent to) Q
ij
whenever Q
i
, Q
j
are at least partially complemen-
tary.
Of course, the ‘correlation (represented by the
XNOR connective ) is only one of many possi-
ble logical connectives. More generally, we shall
require that O’s model for Q does not contain any
logical connectives of complementary questions.
He can only logically connect questions which are
maximally compatible and thus are simultane-
ously defined with respect to him such that he
could meaningfully write down a truth table for
the questions to be connected and the connective
question. If he cannot connect questions, he can
also not ask for the connective.
Assumption 5. Let Q
i
, Q
j
Q and be a logi-
cal connective. According to O’s theoretical model
for Q, Q
i
Q
j
is implementable if and only if
Q
i
, Q
j
are maximally compatible.
12
The reader familiar with Spekkens’ toy model [
40] will
recall the simplest (epistemic) 1-bit system which has four
ontic states ‘1’, ‘2’, ‘3’ and ‘4’. An epistemic restriction
forbids an observer to know the ontic state. Instead, the
epistemic states of maximal knowledge correspond to ei-
ther of the following three questions (and their negations)
Q
1
: “1 2”
, Q
2
: “2 3” ,
Q
3
: “2 4”
,
where is to be read as ‘or’. Q
1
, Q
2
, Q
3
are mutually
complementary and it can be easily checked that Q
3
coin-
cides with the ‘correlation’ Q
12
of Q
1
and Q
2
: it gives ‘yes
when the (ontic) answers to Q
1
, Q
2
are equal and ‘no’ oth-
erwise. This relation is cyclic: Q
1
is also the ‘correlation’
Q
23
of Q
2
, Q
3
and Q
2
is the ‘correlation’ Q
13
of Q
1
, Q
3
.
Accordingly, there are three complementary questions for
this 1-bit system, in principle the correct number for a
qubit. Later in section
5.2.3, we will develop a different
approach, without ontic states, in order to reason for the
three-dimensionality of the Bloch-sphere.
We recall from section
3.2.1 that O’s theoretical
model for Q only contains implementable ques-
tions (but not necessarily all implementable ques-
tions).
Since is only applied to connect maximally
compatible questions which have also opera-
tionally simultaneous truth values relative to O,
we shall allow to be any of the 16 binary connec-
tives (or binary Boolean functions) ¬, , , , . . .
of classical Boolean or propositional logic.
We stress that assumption
5 is a restriction on
O’s model for Q; this assumption clearly does
not rule out that there may exist other (namely,
ontic) models which also y ield a consistent de-
scription of O’s experiences with the s ystems he
is interrogating, yet which ascribe Q
i
Q
j
to be
(logically equivalent to a question) contained in
Q regardless of whether Q
i
, Q
j
are compatible
or not. However, we shall not worry about such
models here.
We note that as sumption
5 does not only ap-
ply to logical connectives of two q uestions, but
arbitrarily many. For example, if Q
i
Q
j
is im-
plementable (and contained in Q), according to
O’s model, then (Q
i
Q
j
) Q
k
is implementable
too iff Q
i
Q
j
and Q
k
Q are compatible, and
so on. It is also important to note that assump-
tion
5 is a statement about which questions can
be directly connected logically. It does not entail
that O, according to this model, can only form
meaningful logical expressions containing exclu-
sively questions which are mutually maximally
compatible. Namely, questions which are com-
patible might be logically further decomposable
into other questions for which different compat-
ibility relations hold. For instance, it might be
that Q
k
in the expression (Q
i
Q
j
) Q
k
is only
compatible with Q
i
Q
j
, yet not with Q
i
or Q
j
in-
dividually. In this case, there is no harm in never-
theless forming the expression (Q
i
Q
j
)Q
k
even
though, say, Q
j
and Q
k
might be complemen-
tary s uch that O cannot write down a truth table
with all three Q
i
, Q
j
, Q
k
separately. However, it
is clear that in this case there is a hierarchy in
which the logical connectives are to be executed.
More specifically, the connective in (Q
i
Q
j
)Q
k
cannot be associative since Q
i
(Q
j
Q
k
) is not
implementable according to O’s model because
Q
j
, Q
k
are, by assumption, complementary and
can therefore not be directly connected. That is
to say, the left in (Q
i
Q
j
)Q
k
must be executed
Accepted in Quantum 2017-11-27, click title to verify 15
first and Q
k
can only be logically connected with
the result. Thus, assumption
5 does admit logical
expressions containing also mutually complemen-
tary questions as long as the latter are not directly
connected through a logical connective. What as-
sumption
5 entails is that the ordering of the exe-
cution of the various logical connectives in a com-
position has to respect a hierarchy which ensures
that at every step two compatible subexpressions
are connected. This will become important later
in the reconstruction where we shall encounter
concrete such examples.
This brings us to the rules of inference by
means of which O may transform or evaluate
logical expressions of questions and derive logi-
cal identities, e.g. to establish possible logical re-
lations among various question outcomes. The
state update rules need, in particular, respect
these rules of inference. Again, the rules of in-
ference which O may employ depend on his the-
oretical model for (Q, Σ). For instance, in ontic
models all q uestions will typically have simultane-
ous values, in contrast to purely operational ones,
which will require different rules of inference.
In line with our operational premise, we shall
require that the rules of inference of O’s model
must not only be consistent with O’s experiences
with the s ystems he is interrogating, but also
testable. In particular, any logical identities de-
rived from these rules must be testable through
interrogations. It is thus appropriate to call them
operational rules of inference.
Classical rules of inference require that the
questions or propositions in a logical expression
have truth values simultaneously. In O’s model
any truth value must be operational such that
only mutually maximally compatible questions
have simultaneous meaning. Nothing stops O
from applying classical rules to such kind of ques-
tions. We shall thus require that it is appropri-
ate for O to employ classical rules of inference,
according to Boolean logic, and only those
for any logical expression or subexpression which
only contains questions that are mutually maxi-
mally compatible. In all other cases (which still
must abide by assumption
5), any possible rule of
inference transforming the composition of ques-
tions will directly involve mutually complemen-
tary questions whose truth values, however, have
no simultaneous operational meaning to O. Ac-
cordingly, classical rules will be operationally in-
appropriate in such cases and we shall demand
more precisely the following.
Assumption 6. O’s model for Q allows him to
apply exclusively classical rules of inference, ac-
cording to the rules of Boolean logic, to (trans-
form or evaluate) any logical expression (or
subexpression of a larger expression) which is
composed purely of mutually maximally compati-
ble questions. This holds regardless of whether
the mutually compatible q uestions can be logi-
cally decomposed into other questions which fea-
ture different compatibility relations. In all other
cases, no classical rules of inference (and thus no
classical logical identities) are valid.
In consequence, O can take any set of mutually
compatible questions and treat them, with clas-
sical logic, as a Boolean algebra. The ass ump-
tion states what is possible for compositions of
compatible questions and what is not possible f or
comp os itions involving also complementary ques-
tions. This will become useful later in section
5.2.7 where we shall also see what is possible for
comp os itions of complementary questions.
We emphasize that as sumption 6 by itself does
not severely constrain the nature of any conceiv-
able ‘hidden variable model’ which could also con-
sistently describe O’s world. However, in con-
junction with two of the subsequent quantum
principles, it will rule out local ‘hidden variables’
in section 5.2.7.
3.2.6 Parametrization of S’s state and t omogra-
phy
Now that we have a notion of independence on Q
we can say more about the parametrization and
thus representation of S’s state relative to O. Not
all Q
i
Q will be necessary to describe the state;
the pairwise independent questions shall be the
fundamental building blocks of the landscape L
of inference theories.
Supp ose there is a maximal set Q
M
=
{Q
1
, . . . , Q
D
Q} of D pairwise independent
(but not necessarily compatible) questions, such
that no further question Q Q\Q
M
exists which
is pairwise independent from all members of Q
M
too. Then, every other Q Q \ Q
M
is either
(i) dependent on exactly one Q
j
Q
M
and in-
dependent of all other Q
M
Q
l
6= Q
j
(if Q
was dependent on Q
j
Q
M
and partially de-
pendent on Q
M
Q
l
6= Q
j
, then Q
j
, Q
l
could
Accepted in Quantum 2017-11-27, click title to verify 16
not be independent), (ii) partially dependent on
some and independent of the other questions in
Q
M
, or (iii) partially dependent on all Q
j
Q
M
.
While O will not be able to infer maximal infor-
mation about the answers to questions of cases
(ii) and (iii) from his information about individ-
ual (or even subsets of) memb ers of Q
M
alone,
the question is whether his information about the
full set Q
M
will be sufficient to do so.
Definition 3.3. (Informational Complete-
ness) A maximal set Q
M
= {Q
1
, . . . , Q
D
Q}
of pairwise independent questions is said to be in-
formationally complete if O’s information about
the questions in Q
M
determines his information
about all other Q Q \ Q
M
in such a way
that the probabilities y
i
which O assigns to eve ry
Q
i
Q
M
are sufficient in order for him to com-
pute the probabilities y
j
Q
j
Q for all prepa-
rations of S. In this case, the set of probabilities
{y
i
}
D
i=1
of the Q
i
Q
M
parametrizes the state
that O assigns to S and thereby yields a com-
plete description of the state space Σ. We shall
call D the dimension of Σ.
We emphasize that the dimension D of Σ will
ultimately not be the Hilbert space dimension in
quantum theory, but the dimension of the set of
density matrices.
If Q
M
was not informationally complete, O
would require further questions that are partially
dependent on at least some of the elements in
Q
M
in order to fully describe the system S and
its state. This situation cannot be precluded,
given the structure we have devis ed so far. How-
ever, we deem it undesirable, given that we would
like to employ pairwise independent questions as
building block s for system descriptions. We shall
therefore require that no more independent infor-
mation about S can be learned from any question
in addition to a maximal set Q
M
.
Assumption 7. Every maximal set Q
M
of pair-
wise independent questions is informationally
complete.
There may exist (even continuously) many such
informationally complete sets of questions on Q
which, at this stage, may still be either discrete or
continuous. However, their dimensions are equal.
Lemma 3.4. The dimensions of all maximal sets
Q
M
on Q are equal (if finite).
Proof. Consider any Q
M
. On account of pairwise
independen ce, for each Q
i
Q
M
there must exist
a state such that O only knows the answer to
this Q
i
with certainty, but nothing about any
other Q
j6=i
Q
M
, i.e. y
i
= 0 or 1 and all other
y
j6=i
= 1/2. (Namely, O could have asked Q
i
to
S prepared in the state of no information.) But
since also in the state of no information y
j6=i
=
1/2, y
i
cannot in general be computed fr om the
y
j6=i
. Hence, given an inform ationally complete
set Q
M
, all associated probabilities {y
i
}
D
i=1
are
necessary to parametrize the full state space Σ.
Now Σ is a convex set (see assum ption
2).
Given the lack of redundancy in {y
i
}, a given
informationally complete Q
M
with D elements
therefore encodes Σ as a D-dimensional convex
subregion of R
D
. Su ppose there is a second Q
M
with D
elements. In terms of Q
M
, Σ would be
described as a D
-dimensional convex s ubregion
of R
D
. But clearly, th ese two descriptions of Σ
must be isomorphic which, for D finite, is only
possible if D D
.
Finiteness of D will later be established in sec-
tion
5 with the help of the principles.
Any such Q
M
establishes a question refe rence
frame on Q (see also [73]) and thereby also a ‘co-
ordinate s ystem’ on Σ. In particular, in order
to do state tomography with a multiple shot in-
terrogation, as outlined in section
3.2.3, it will
be sufficient for O to interrogate an ensemble of
identically prepared S with the questions within
a given Q
M
only. Given a specific Q
M
, there
are now three equivalent ways for O to describe
S’s state: he could represent it by either the D-
dimensional ye s- or no-vector
~y
OS
:=
y
1
y
2
.
.
.
y
D
, ~n
OS
:=
n
1
n
2
.
.
.
n
D
, (3)
of probabilities y
i
and n
i
, i = 1, . . . , D, that the
answers to question Q
i
Q
M
are ‘yes’ and ‘no’,
respectively. That is,
~y
OS
+ ~n
OS
= p
~
1, (4)
where
~
1 is a D-dimensional vector with a 1 in each
of its entries. Here we have introduced a new pa-
rameter: p is the probability that S is present
at all. (For example, the method of preparation
Accepted in Quantum 2017-11-27, click title to verify 17
could be such that O cannot always tell with cer-
tainty when a system is spit out by the prepara-
tion device.) This probability will rescale all y
i
and n
i
simultaneously. However, it only becomes
relevant in subsection
3.2.8 and otherwise is usu-
ally taken to be p = 1. Evidently, the assignment
of which answer to Q
i
is ‘yes’ and which is ‘no’ is
arbitrary, but any consistent such assignment is
fine for us.
But he could also represent the state redun-
dantly as a 2D-dimensional vector
~
P
OS
=
~y
~n
!
(5)
which will turn out to be convenient especially
when p < 1. We shall write a state with a sub-
script O S to emphasize that it is the state of
O’s information about S.
Lastly, this structure also puts us into the po-
sition to specify O’s total amount of informa-
tion about S. Clearly, the total amount of in-
formation must be a function of the state. Let
Q
M
= {Q
1
, . . . , Q
D
} be an informationally com-
plete set of questions in Q. Given that these ques-
tions carry the entire information O may know
about S, we define the total information I
OS
as
the sum of O’s information about the Q
i
Q
M
,
as measured by the α
i
(
1):
I
OS
(~y
OS
) :=
D
X
i=1
α
i
. (6)
(We imagine O as an agent who can write down
results on a piece of paper and add these up.) The
specific relation between α
i
and ~y
OS
will be de-
rived later in section 6.8; it will not be the Shan-
non entropy which, as discussed in section
6.1,
describes average information gains in repeated
experiments rather than the information content
in the state ~y
OS
.
3.2.7 Composite systems
For later purpose we need to clarify the notion
of a composite system. Since we are pursuing a
purely operational approach, the notion of a com-
position of systems must be defined in terms of
the information accessible to O through interro-
gation. O should, in principle, be able to tell
a composite system apart into its constituents.
Accordingly, we require that the set of questions
which can be pos ed to the composite system con-
tains all questions about the individual subsys-
tems and that the remaining questions are lit-
erally composed of these individual questions or
iteratively composed of compositions of them.
13
Definition 3.5. (Composite System) Let
Q
A
, Q
B
denote the full question sets associated
to S
A
, S
B
, respectively. Two systems S
A
, S
B
are
said to form a composite system S
AB
if any
Q
a
Q
A
is maximally compatible with and in-
dependent of any Q
b
Q
B
and if
Q
AB
= Q
A
Q
B
˜
Q
AB
, (7)
where
˜
Q
AB
only contains composite ques-
tions which are i terative compositions, Q
a
1
Q
b
, Q
a
2
(Q
a
3
Q
b
), (Q
a
4
Q
b
)
5
Q
b
, (Q
a
6
Q
b
)
7
(Q
a
8
Q
b
), . . ., via some logical con-
nectives
1
,
2
,
3
, · · · , of individual questions
Q
a
, Q
a
, . . . Q
A
about S
A
and Q
b
, Q
b
, . . . Q
B
about S
B
. For an N -partite system, we use this
definition recursively.
Of course, thanks to assumption
5, O can only
directly compose questions with a logical connec-
tive if they are compatible.
There are further repercussions: let Q
M
A
, Q
M
B
be informationally complete sets for S
A
, S
B
, re-
spectively. Then an informationally complete set
Q
M
AB
for a composite S
AB
can be formed itera-
tively by joining (in a set union sense) Q
M
A
, Q
M
B
and adding to it a maximal pairwise independent
set of questions which are the logical connectives
of members of Q
M
A
with elements of Q
M
B
and/or
iterative compositions of such compositions and
possibly questions from Q
M
A
, Q
M
B
. In section
5.2.1 below, we shall determine which logical con-
nectives we may employ to build up Q
M
AB
from
Q
M
A
, Q
M
B
.
3.2.8 Time evolution of S’s state
We shall assume that O has access to a clock and
begin with a definition:
static states: A state
~
P
OS
which is constant
in time corresponding to the situation that
13
For the GPT specialists, we emphasize that this defi-
nition has noth ing to do with the usual requirement of local
tomography [
1518,20,55,59] in GPTs. To give a concrete
example, we shall see later that two-level systems over real
Hilb ert spaces (rebits) satisfy this definition while violat-
ing local tomography.
Accepted in Quantum 2017-11-27, click title to verify 18
O will always assign the same probability
to each of his question outcomes is called
static.
If all the states O would assign to any system
S were static according to his information
O’s world would be a rather boring place. In or-
der not to let O die of boredom, we allow the
probability vector
~
P
OS
to change in time. (But
we require that Q and Σ are time independent.)
We shall now make use of operational reasoning,
to briefly consider how
~
P
OS
evolves under time
evolution. The following argument bears some
analogy to an argument typically employed in
the GPT framework [
1418, 55] concerning con-
vex mixtures and measurements, however, is dis-
tinct in nature as we instead consider states and
their mixtures under time evolution.
Let O have access to two identical
14
non-
interacting systems S
1
and S
2
. O’s information
about S
1
and S
2
can be different such that they
are allowed to be in distinct states ~y
OS
1
and
~y
OS
2
relative to O. N ow let O perform a (bi-
ased) coin flip which yields ‘heads’ with probabil-
ity λ and ‘tails’ with probability (1 λ).
15
Given
that S
1
and S
2
are identical, O can ask the same
questions to both s ystems. If the coin flip yields
‘heads’, O will interrogate S
1
, if it yields ‘tails’ O
will interrogate S
2
.
PSfrag replacements
O
S
1
S
2
λ 1 λ
In particular, before toss ing the coin, say at
time t
1
, the probability he assigns to receiving a
‘yes’ answer to question Q
i
(asked to either S
1
or S
2
depending on the outcome of the coin flip)
is simply the convex s um y
12
i
(t
1
) = λ y
1
i
(t
1
) +
(1 λ) y
2
i
(t
1
). (O is allowed to build this con-
vex combination thanks to assumption
2.) But
this holds at any time t before O tosses the coin
(which O can also choose never to do).
16
As will
14
That is, both S
1
and S
2
carry the same Q and Σ.
15
Equivalently, O m ay use another system to which he
has assigned a stable state vector by repeated interroga-
tions on identically prepared systems. Given an arbitrary
elementary question Q, he can use the probability that the
answer is ‘yes as λ and the probability that the outcome
is ‘no’ as (1 λ) and thus use Q as the ‘coin’.
16
We assume λ to be constant in time.
become clear shortly, it is convenient to consider
the states
~
P
OS
with 0 p 1 for now. We
shall return to the yes-vector later in section
6.
We then have for all t
~
P
OS
12
(t) = λ
~
P
OS
1
(t) + (1 λ)
~
P
OS
2
(t). (8)
This convex combination is the state of an ‘effec-
tive’ system S
12
(identical to S
1
, S
2
) describing
O’s information about the coin flip scenario. Note
that S
12
is not a composite system according to
definition
3.5 because Q
12
= Q
1
= Q
2
.
Denote by T
k
the time evolution map of the
state of system S
k
, k = 1, 2, 12, from t = t
1
to
t = t
2
. Under the assumption that
~
P
OS
1,2
(t)
evolve independently of each other, given that S
1
and S
2
do not interact, equation (
8) implies
T
12
[
~
P
OS
12
(t
1
)] = T
12
[λ
~
P
OS
1
(t
1
)
+ (1 λ)
~
P
OS
2
(t
1
)]
= λ T
1
[
~
P
OS
1
(t
1
)] (9)
+ (1 λ) T
2
[
~
P
OS
2
(t
1
)].
All three states
~
P
OS
k
, k = 1, 2, 12, are elements
of the same Σ. As such, we assume the following.
Assumption 8. Every time evolution map can
act on any state in Σ.
Next, we restrict to the situation where O ex-
poses both S
1
, S
2
to evolve under the same time
evolution T := T
1
= T
2
. Thanks to assumption
8, (9) becomes in this case
T [
~
P
OS
12
(t
1
)] = λ T [
~
P
OS
1
(t
1
)]
+ (1 λ) T [
~
P
OS
2
(t
1
)]
and T is convex linear. This is clear since O may,
in particular, prepare S
1
, S
2
in identical states
~
P
OS
1
(t
1
) =
~
P
OS
2
(t
1
) =
~
P
OS
12
(t
1
). This spe-
cial case, when inserted into (
9), y ields T
12
= T .
This scenario can be easily generalized to ar-
bitrarily many identical systems. Remarkably, it
can be shown that convex linearity of operations
on probability vectors, in fact, implies full lin-
earity of the operation [
14, 55]. Hence, the time
evolution of the state must be linear:
~
P
OS
(t
2
) = T [
~
P
OS
(t
1
)]
= A(t
1
, t
2
)
~
P
OS
(t
1
) +
~
V (t
1
, t
2
),
where A(t
1
, t
2
) is a 2D × 2D matrix and
~
V some
2D-dimensional vector. Given our assumption
above that O’s world is not static, A(t
1
, t
2
) will
Accepted in Quantum 2017-11-27, click title to verify 19
generally be a non-trivial matrix. Requiring that
the last relation als o holds at t
1
= t
2
for all
initial states, it follows that
~
V (t, t) = 0 and
A(t, t) = 1. Demanding further that the sp e-
cial state
~
P
OS
=
~
0, corresponding to p = 0
and the absence of a system, is invariant un-
der time evolution (such that no system or any
information is created out of ‘nothing’), finally
yields
~
V
~
0, t
1
, t
2
.
17
Given that
~
P
OS
(t) is a
probability vector for all t and (
4) must always
hold, A(t
1
, t
2
) must be a nonnegative (real) ma-
trix which is stochastic in any pair of components
i and i + D of
~
P
OS
.
Consequently, the elementary and natural as-
sumption
8 rules out non-linear and state de-
pendent time evolution, such as in Weinberg’s
non-linear extension of quantum mechanics [74],
and thereby also eliminates the many operational
problems that arise with it (e.g., superluminal
signaling) [
7577].
18
For later purpose, we impose a further condi-
tion on the evolution of S’s state.
Assumption 9. (Temporal Translation In-
variance) There exist no distinguished instants
of time in O’s world such that O is free to set
any instant he desires as the instant t = 0. Time
evolution, as perceived by O, is therefore (tempo-
rally) translation invariant A(t
1
, t
2
) = A(t
2
t
1
).
Since the time evolution matrix A can only de-
pend on the duration elapsed, but not on the par-
ticular instant of time, we can collect the above
results in the simple form:
~
P
OS
(t
2
) = A(∆t)
~
P
OS
(t
1
), t = t
2
t
1
. (10)
We do not yet have sufficient evidence to con-
clude that a given time evolution will be described
by a continuous one-parameter matrix group be-
cause we neither know (i) that A is invertible for
17
A similar argument would not work for the evolution
of ~y
OS
because, in principle, ~n
OS
6=
~
0 is possible even
if ~y
OS
=
~
0 (where
~
0 is a D-dimensional zero vector).
This is the reason why here it is more convenient to work
with
~
P
OS
, which contains the information about both
~y
OS
, ~n
OS
, rather than ~y
OS
alone. In fact, we shall
see later in section 6 that the evolution of the latter does
involve an affine D-dimensional vector
~
V
6=
~
0.
18
It should be emphasized that assumption
8 is also tac-
itly made in the GPT framework [
1418, 55] for any kind
of transformations on states, such that the present setup
is in this regard not less potent.
all t, (ii) that the evolution is actually continu-
ous, nor (iii) that every composition of evolution
matrices is again an evolution matrix. We shall
return to this question in section
6 with the help
of the set of principles for quantum theory and
also defer the discussion of the time evolution of
~y
OS
until then.
For now we note, however, firstly, that any
A(∆t) acting on
~
P
OS
implies a unique map
T
t
(~y
OS
) acting on the yes-vector thanks to (
4,
5). Secondly, a multiplicity of time evolutions
of S is possible, depending on the physical cir-
cumstances (interactions) to which O may sub-
ject S. The set of all p ossible time evolutions
which O shall henceforth be able to implement
will be denoted by T . This set need not contain
all physically possible time evolutions. Indeed,
upon imposing the quantum principles, T will be-
come the set of unitaries (rather than of arbitrary
completely pos itive maps). Clearly, T is part of
O’s model for describing S; his model is thus a
triple (Q, Σ, T ).
3.2.9 The theory landscape L
The landscape L of theories describing O’s acqui-
sition of information about a physical system S
is now the set of all theories which comply with
the structure and assumptions established in this
section.
In the sequel, we shall restrict O’s attention
solely to composite systems of N N general-
ized bits (gbits), where a single gbit is character-
ized by the fact that O can maximally know the
answer to a single question at once such that it
can carry at most one bit of information. Ev-
ery inference theory specifies for every system of
N gbits a triple (Q
N
, Σ
N
, T
N
). We shall be con-
cerned with the landscape of gbit theories L
gbit
which contains all gbit theories, satisfying as-
sumptions
19. To name concrete examples at
this stage, L
gbit
contains, among a continuum of
other theories, classical bit, rebit and qubit the-
ory for all N N. We shall see more of this later,
but for now we summarize their characteristics
for N = 1,
classical bit theory gives Q
1
= {Q, ¬Q},
Σ
1
[0, 1] (for normalized states) with ex-
tremal points corresponding to Q = ‘yes’ and
‘no’, respectively, and T
1
Z
2
.
19
19
There are precisely two states of maximal information
Accepted in Quantum 2017-11-27, click title to verify 20
rebit theory (two-level systems on real Hilbert
spaces) yields Q
1
S
1
and every two maxi-
mally complementary questions are informa-
tionally complete. Σ
1
D
2
(for normalized
states) and T
1
SO(2).
qubit theory has Q
1
S
2
and every triple
of mutually maximally complementary ques-
tions are informationally complete. Σ
1
B
3
(for normalized states) and T
1
SO(3).
With all the assumptions made along the way,
the theory landscape devised here is perhaps not
quite as general as the commonly employed land-
scape of generalized probability theories [
1418,
23, 55, 60, 61]. In particular, GPTs easily handle
arbitrary finite dimensional systems, while this
remains to be done within the present frame-
work. N evertheless, the new landscape L
gbit
is
large enough and prov ides new tools for a non-
trivial and instructive (re)construction of qubit
quantum theory that approaches the latter from a
conceptually and technically new angle. To facil-
itate future generalizations of this work, we have
also attempted to spotlight the assumptions un-
derlying L
gbit
as clearly as possible in this s ection.
4 Principles for the quantum theory of
qubits as rules on information acquisi-
tion
We shall now use the landscape L
gbit
of gbit theo-
ries and formulate the principles for the quantum
theory of qubit systems on it as rules on an ob-
server’s information acquisition. These principles
constitute a set of ‘coordinates’ of qubit quantum
theory on L
gbit
.
that O can assign to a classical bit
~
P
OS
=
1
0
,
~
P
OS
=
0
1
,
corresponding to ‘yes’ and ‘no’ answers to Q. Time evo-
lution is described by the abelian group Z
2
, given by
1 =
1 0
0 1
, P =
0 1
1 0
,
where P swaps the two states. For classical bit theory, the
permitted time evolution is therefore discrete.
4.1 The rules
As we are reconstructing a technically and empir-
ically well-established theory, we are in the fortu-
nate position to avail ourselves of empirical ev-
idence and earlier ideas on characterizing quan-
tum theory in order to motivate a set of basic
postulates. However, the ultimate justification
for these postulates will be their success in sin-
gling out qubit quantum theory within the infer-
ence theory landscape L
gbit
which will be com-
pleted in [
1]. As ‘coordinates’ on a theory space
these principles will not be unique and one could
find other equivalent sets. (As usual, at least
many roads lead to Rome.) But we shall take
the below ones as a first working set which re-
stricts the informational relation between O and
S, where S will be a composite system (c.f. def-
inition
3.5) of N N gbits. Each principle will
be first expressed as an intuitive and colloquial
statement, followed by its mathematical meaning
within L
gbit
.
We take it as an empirical fact that there ex-
ist physical systems about which only a limited
amount of information can be known at any one
moment of time. The standard quantum example
is a spin-
1
2
particle about which an experimenter
may only know its polarization in a given spatial
direction, but nothing independent of that; the
polarization ‘up’ or ‘down’ corresponds to one bit
of information. But there is also a typical classi-
cal example, namely a ball which may be located
in either of two identical boxes, the definite po-
sition ‘left’ or ‘right’ corresponding to one bit of
information. Informationally, these two examples
incarnate the most elementary of systems, a gbit,
supporting maximally just one proposition at a
time. But, clearly, there are more complicated
systems supporting other limited amounts of in-
formation. We shall take this simple observation
and raise the existence of an information limit to
the level of a principle.
More precisely, in analogy to von Weizsäcker’s
‘ur-theory’ [
7880], we shall restrict O’s world
to be a world of elementary alternatives which
thereby consists only of systems which can be de-
comp os ed into elementary gbits.
20
All physical
20
However, in contrast to [
7880], we shall be much
less ambitious here and will not attempt t o deduce the
dimension of space or space-time symmetry groups from
systems of elementary alternatives. Recent developments
[
21, 42, 59, 81], on the other hand, unravel a deep rela-
Accepted in Quantum 2017-11-27, click title to verify 21
quantities in O’s world are to be finite such that
he can record his information about them on a
finite register. We wish to characterize the com-
posite systems in O’s world according to the finite
limit of N bits of information he can maximally
inquire about them.
Rule 1. (Limited Information) “The observer
O can acquire maximally N N independent
bits of information about the system S at any
moment of time.
There exists a maximal set Q
i
, i = 1, . . . , N, of
N mutually independent and maximally compat-
ible questions in Q
N
and no su bset in Q
N
can
contain more than N questions with that prop-
erty.
In other words, O can ask maximally N inde-
pendent questions
21
at a time to S. Accordingly,
this rule immediately implies that O can distin-
guish maximally 2
N
states of S in a single shot
interrogation because there will be 2
N
possible
answers to the Q
1
, . . . , Q
N
. Since Q
N
, Σ
N
are in-
trinsic to S, also the maximal amount of N bits
that each S can carry must be intrinsic to it and
thus be observer independent. In fact, this rule
can be regarded as a defining property of all in-
ference theories in the gbit landscape L
gbit
and
can thus clearly not distinguish between classical
bits, rebits, qubits, etc.
We take another empirical f act and elevate
it to a fundamental principle: despite the lim-
ited information accessible to an experimenter
at any moment of time, there always exists ad-
ditional independent information that she may
learn about the observed system at other times.
This is Bohr’s complementarity principle [
82].
Consider, for instance, the prototypical quantum
physics experiment: Young’s double slit exper-
iment. The experimenter can choose whether
to obtain which-way-information or an interfer-
ence pattern, but not both; the complete knowl-
edge of whether the particle went through the
left or right slit is at the expense of total ig-
norance about the information pattern and vice
tion between (a) simple conditions on operations with sys-
tems carrying fin ite information (e.g., communication with
physical systems) and (b) the dimension and symmetry
group of the ambient space or space-time.
21
Obviously, combinations (e.g., ‘correlations’) of the
compatible Q
i
will define other bits of information which,
however, will be dependent once the Q
1
, . . . , Q
N
are asked
(we shall return to this in detail in section
5).
versa. (For an informational discussion of the
particle-wave duality in Young’s double slit ex-
periment and a Mach-Zehnder interferometer, see
also [
32,35].) Similarly, in a Stern-Gerlach exper-
iment an experimenter may determine the polar-
ization of a spin-
1
2
particle in x-direction, but will
be entirely oblivious about the polarization in y-
and z-direction. A subsequent measurement of
the spin of the same particle in y-direction will
render her previous information about the polar-
ization in x-direction obsolete and keep her ig-
norant about the spin in z-direction and so on.
That is, systems empirically admit many more
independent questions than they are able to an-
swer at a time thanks to the information limit.
We shall now return to the relation between O
and S and accordingly stipulate that complemen-
tarity exists in O’s world, however, we shall say
nothing more about how much complementary in-
formation may exist.
Rule 2. (Complementarity) “The observer O
can always get up to N new independent bits of
information about the system S. But whenever
O asks S a new question, he experiences no n et
loss in his total information about S.
There exists another maximal set Q
i
, i =
1, . . . , N, of N mutually independent and
maximally compatible questions in Q
N
such
that Q
i
, Q
i
are maximally complementary and
Q
i
, Q
j6=i
are maximally compatible and indepen-
dent.
22
That is, after asking S a set of N indepen-
22
Alternatively, one could formulate the technical part
of this rule as follows: In the N = 1 case of a single
gbit there exists, for every Q
1
Q
1
, another Q
1
Q
1
such that Q
1
, Q
1
are maximally complementary. Defini-
tion
3.5 would then immediately imply that a compos-
ite sy stem of N gbits features two sets of questions, Q
i
,
i = 1, . . . , N, and Q
j
, j = 1, . . . , N , with the following
properties: Q
i
, Q
i
are questions about the elementary gbit
labeled by i, all Q
i
are maximally independent and max-
imally compatible, all Q
j
are maximally independent and
maximally compatible and any pair Q
i
, Q
j6=i
is maximally
independent and compatible and every pair Q
i
, Q
i
corre-
sponding to th e same gbit is maximally complementary.
This is a priori not equivalent to the current formulation
of the technical part of rule
2 in the main text as the lat-
ter does not necessarily refer to questions about individual
subsystems of a composite system. However, the alterna-
tive formulation would also be enough to get the correct
structure of the informationally complete sets for N qub its
and N rebits below. While the alternative formulation is
simpler, we keep the other one as it is also published in
this form in the companion article [
1].
Accepted in Quantum 2017-11-27, click title to verify 22
dent elementary propositions Q
i
, i = 1, . . . , N,
in a single shot interrogation, O can pose a new
(N + 1)th elementary question Q
1
to S. Since O
can only know N independent bits of informa-
tion about S at a time and asking a new question
does not lead to a net loss of information, the
single bit of his previous information about Q
1
must have become obsolete upon learning the an-
swer to Q
1
, while O’s total information about S
is still N bits. The rule allows O to perform the
same procedure until he replaces his N old by N
new bits of information about S. As O’s infor-
mation about S has changed, the state of S rel-
ative to O will necessarily experience a ‘collapse’
whenever he asks a new complementary question.
As a result, it is the observer who decides which
information (e.g., spin in x- or y-direction) he will
obtain about the system by asking specific ques-
tions. But clearly O will have no influence on
what the answer to these questions will be and
any answer will come at the price of total igno-
rance about complementary questions.
A few further explanations concerning this
complementarity rule are in place. First of all,
this rule clearly rules out classical bit theory. Sec-
ondly, the postulate asserts the existence of max-
imally complementary questions in O’s world,
however, makes no statement about whether par-
tially independent or partially complementary
questions may exist too. Thirdly, the p eculiar
requirement that every question of rule
2 is com-
plementary to exactly one from rule
1 is chosen
such that each s ubsystem of a composite system
features complementarity. For example, consider
the N = 1 case, corresponding to a single gbit,
for which rules
1 and 2 entail a complementary
pair Q
1
, Q
1
. This complementarity per gbit gen-
eralizes to arbitrary N since every Q
i
in rule
1
and every Q
j
in rule
2 may correspond to one of
N gbits. More complicated complementarity re-
lations arising from rules
1 and 2 will be discuss ed
in section
5.
Notice that the complementarity rule implies
the existence of a notion of ‘su perposition even
in states of maximal information. For example,
take N = 1 and let O know the answer to Q
1
with certainty, y
1
= 1, such that his informa-
tion about S saturates the limit of one bit. This
will leave him oblivious about the complementary
Q
1
such that he would have to assign y
1
=
1
2
.
The state of information he has about S can be
interpreted as being in a ‘superposition’ of the
Q
1
alternatives ‘yes’ and ‘no’, but in this case
with ‘equal weight’ because both alternatives are
equally likely according to O’s knowledge.
23
We
emphasize that the necessary presence of super-
positions of elementary alternatives, as perceived
by O, is a consequence of the information limit
and complementarity.
We would like to stress that rules
1 and 2
are conceptually motivated by related propos-
als which have been put forward first by Rov-
elli within the context of relational quantum me-
chanics [
28] and later, indep endently, by Brukner
and Zeilinger within attempts to understand the
structure of quantum theory via limited informa-
tion [
31,32,34,35,73]. However, in order to com-
plete these ideas to a full reconstruction of qubit
quantum theory, we have to impose further rules.
Specifically, rules
1 and 2 say nothing about
what happens in-between interrogations. We re-
quire that O shall not gain or lose information
about an otherwise non-interacting S without
asking questions.
Rule 3. (Information Preservation) “The to-
tal amount of information O has about (an oth-
erwise non-interacting) S is preserved in -between
interrogations.
I
OS
is constant in time in-between interroga-
tions for (an otherwise non-interacting) S.
Corresp ondingly, I
OS
is a conserved charge
of time evolution; this is a simple observation
which will become extremely us eful later. In fact,
notice that rule
3 could also be viewed as defin-
ing the notion of non-interacting systems (as per-
ceived by O).
Finally, we come to the last rule of this
manuscript which is about time evolution. Em-
pirical evidence suggests, at least to a good ap-
proximation, that the time evolution of an ex-
perimenter’s ‘catalogue of k nowledge’ about an
observed system is continuous in-between mea-
surements. More precisely, the specific prob-
23
We shall later see that, even in a state of maximal
information of N bits, O m ay possibly have only partial
or incomplete knowledge about the outcomes of any ques-
tion in an informationally complete set Q
M
N
of pairwise
independent questions (c.f. assumption
7). In this case,
O’s information about all questions in Q
M
N
will be in a
general state of superposition because their correspond-
ing probabilities cannot all be
1
2
, otherwise it would be
the state of no information.
Accepted in Quantum 2017-11-27, click title to verify 23
abilistic statements an experimenter can make
about the outcomes of measurements , i.e. his ac-
tual information about the s ystems, change con-
tinuously in time. We promote this to a further
postulate for O’s world, however, with a further
requirement.
We note that the state space and the time evo-
lutions are interdependent as every legal time evo-
lution must map a legal state to a legal state, i.e.
T
t
(~y
OS
) Σ
N
, T
t
T
N
and ~y
OS
Σ
N
.
Accordingly, what is the set of legal states de-
pends on what is the set of legal time evolutions
and vice versa. We would like the pair
N
, T
N
)
to be as ‘big’ as compatibility with the other rules
allows in order to equip the other rules with as
general a validity as possible. But there are mul-
tiple ways of ‘maximizing’ the pair. Namely, the
interdependence implies that the larger the num-
ber of states, the tighter the constraints on the set
of time evolutions and vice versa. We required
before that O’s world should not be ‘boring’ and
therefore feature a non-trivial time evolution of
states. We shall now sharpen this requirement: it
is more interesting for O to live in a world which
‘maximizes’ the number of possible ways in which
any given state of S can change in time, rather
than the number of states which it can be in rela-
tive to O. We shall thus require that any consis-
tent, continuous time evolution of O’s ‘catalogue
of knowledge’ about S in-between interrogations
is physically realizable. Since different ways of
evolving the state of a system correspond to dif-
ferent interactions (e.g., among the members of a
comp os ite system) we thereby maximize the set
of possible interactions among systems. O’s world
is a maximally interactive place!
Rule 4. (Time Evolution) O’s ‘catalogue of
knowledge’ about S evolves continuously in time
in-between interrogations and every consistent
such evolution is physically realizable.
T
N
is the maximal set of transformations T
t
on
states which is continuous in t and compatible
with rules
1-3 and the structure of L
gbit
.
Of course, during an interrogation, ~y
OS
may
change discontinuously, i.e. ‘collapse’, on account
of complementarity. As innocent as the require-
ment of continuity of time evolution in rule
4 ap-
pears, it turns out to be absolutely crucial in or-
der to single out a unique information measure
α
i
(~y
OS
). Notice also that classical bit theory
violates this postulate due to its discrete time
evolution group (see section
3.2.9).
It is now our task to verify what the triples
of (Q
N
, Σ
N
, T
N
) for each N N are which obey
rules
14. Remarkably, it turns out that these
four rules cannot distinguish between real and
complex quantum theory, i.e. between real and
complex Hilbert spaces. But the state spaces and
the orthogonal and unitary groups of time evo-
lutions of rebit and qubit quantum theory are,
in fact, the only ‘solutions’ within L
gbit
to these
rules. In the sequel of this manuscript we shall
employ rules
14 in order to develop the necessary
tools, within L
gbit
, for eventually proving the fol-
lowing more precise claim in [1, 2]. In fact, as a
simple example of the newly develop ed tools, we
shall already prove the claim for N = 1 at the
end of this article.
Claim. L
gbit
contains only two solutions for the
pair
N
, T
N
) which are compatible with rules
1
4:
1. rebit quantum theory [
2], where Σ
N
co-
incides with the space of 2
N
× 2
N
density
matrices
24
over (R
2
)
N
and T
N
is PSO(2
N
).
2. qubit quantum theory [1], where Σ
N
co-
incides with the space of 2
N
× 2
N
density
matrices over (C
2
)
N
and T
N
is PSU(2
N
).
Furthermore, states evolve unitarily accord-
ing to the von Neumann evolution equ ation.
We remind the reader that the time evolution
groups T
N
for density matrices in rebit and qubit
quantum theory are projective because they cor-
respond to ρ 7→ U ρ U
, where for (1) rebits ρ is a
2
N
× 2
N
real symmetric matrix, U SO(2
N
) and
denotes matrix transpos e; and (2) for qubits ρ
is a 2
N
× 2
N
hermitian matrix, U SU(2
N
) and
denotes hermitian conjugation.
Furthermore, using one additional rule on Q
N
which allows O to ask S any question that “makes
(probabilistic) sense”, we show in [
1,2] that
1. in the rebit case [
2], Q
N
is (isomorphic
to) the set of projective measurements on to
the +1 eigenspaces of Pauli operators
25
over
R
2
N
. This is the set of rank-(2
N1
)-
projectors.
24
A real density matrix is a symmetric, positive semidef-
inite matrix on (R
2
)
N
.
25
The set of Pauli operators over R
2
N
is the set of trace-
less real symmetric 2
N
×2
N
matrices with eigenvalues ±1.
Accepted in Quantum 2017-11-27, click title to verify 24
2. in the qubit case [1], Q
N
is (isomorphic
to) the set of projective measurements on to
the +1 eigenspaces of Pauli operators
26
over
C
2
N
(i.e., the set of rank-(2
N1
)-projectors)
and the outcome probability for any Q Q
N
to be answered with ‘yes’ by S in some state
is given by the Born rule for proj ective mea-
surements.
Since both real and complex quantum theory
come out of rules
14, we shall impose a further
rule on O’s information acquisition in [
1] in order
to also eliminate rebit theory. However, we note
that rebit theory is mathematically contained in
qubit quantum theory and that rebits can ac-
tually be produced in laboratories. Hence, one
might also hold rules
14 and the rule that O
may ask S any question that “makes sense” as
physically sufficient.
Notice that it is necessary to directly recon-
struct the space of density matrices over Hilbert
spaces from the rules rather than the underly-
ing Hilbert spaces themselves. The reason is that
the latter contain physically redundant informa-
tion (norm and global phase), while the rules refer
only to information which is directly accessible to
O.
4.2 Strategy for building the necessary tools
and proving the claim
Our strategy and procedure f or developing tools
and intermediate results before proving the main
claim is best summarized as a diagram (see figure
3).
PSfrag rep lacements
limited information complementarity information preservation time evolution
time evolution
independen ce, compatibility and cor-
relation structure on Q
N
(in sec.
5)
reversible time evolution (in sec.
6)
and information measure (in sec.
6.8)
Σ
1
is a ball with D = 2, 3, and
T
1
is either SO(2) or SO(3) (in sec.
7)
T
2
is either PSO(4) or PSU(4)
Σ
2
convex hull of RP
3
or CP
3
(in [1,2])
(Q
N
, Σ
N
, T
N
) for N > 2 (in [1, 2])
Figure 3: Strategy an d steps for the reconstruction.
That is, we shall firstly ascertain, in section 5,
the independence, compatibility and correlation
structure on Q
N
which is induced by rules 1 and
2. This section will be particularly instructive
and deliver many elementary technical results
26
The set of Pauli operators over C
2
N
is the set of trace-
less hermitian 2
N
× 2
N
matrices with eigenvalues ±1.
which, besides becoming important later, yield
simple explanations f or typical quantum phenom-
ena such as, e.g., entanglement, non-locality and
monogamy relations. This is also where we shall
determine the dimensionality of Σ
1
by a simple
argument and, using this result, derive the di-
mensionalities of all other state spaces too. Next,
in section
6, we shall use rules 3 and 4 in order
Accepted in Quantum 2017-11-27, click title to verify 25
to conclude that a specific time evolution, is re-
versible and, in fact, described by a continuous
one-parameter matrix group. These results will
then be used in section
6.8, together with rules
3 and 4, to derive an explicit information mea-
sure α
i
(~y
OS
) which is unique under elementary
consistency conditions. In section 7, the conjunc-
tion of these outcomes will then be employed to
show that for N = 1 one either obtains a two-
or three-dimensional Bloch ball as a state space
Σ
1
and that the group of all possible time evo-
lutions T
1
is accordingly either SO(2) or SO(3).
The claim for N > 1 will require s ubstantial addi-
tional work and will be proven in the companion
paper [
1] for qubits and in [2] for rebits.
5 Question structure and correlations
In this section, we shall solely employ rules
1 and
2 in order to deduce the independence, compati-
bility and correlation structure for the questions
in an informationally complete set Q
M
N
(c.f. as-
sumption
7). To this end it is instructive to look
at the individual cases N = 1, 2, 3 in some de-
tail before considering general N N. We shall
denote by D
N
the dimension of the state space
Σ
N
.
5.1 A single gbit
Rules
1 and 2 taken together imply that for N = 1
there exists at least one maximally complemen-
tary pair Q
1
, Q
1
. Since N = 1 is fixed, it is now
more convenient to count the indep endent ques-
tions by an index, rather than a prime.
27
Let us
therefore slightly change the notation and write
Q
1
henceforth as Q
2
, such that the rules imply
the existence of a maximally complementary pair
Q
1
, Q
2
. But, applied to a single gbit only, rules
1 and 2 tell us nothing more about how many
questions maximally complementary to Q
1
exist.
There may arise another Q
3
maximally comple-
mentary to both Q
1
, Q
2
etc. Notice that, for a
single gbit, an informationally complete set of
pairwise independent questions must be given by
a maximal set of mutually maximally comple-
mentary questions (i.e., no further Q Q
1
can
be added to this set without destroying mutual
27
Rules
1 and 2 count compatible questions, here we
need to count maximally complementary questions.
maximal complementarity)
Q
M
1
= {Q
1
, Q
2
, . . . , Q
D
1
} (11)
because rule
1 prohibits further pairwise indepen-
dent (partially or maximally) compatible ques-
tions. (Recall that maximally complementary
questions are automatically independent.) Rules
1 and 2 imply D
1
2. It turns out that we need
to consider two gbits in order to upper bound D
1
.
We shall call such questions for a single gbit
individual questions.
5.2 Two gbits
The N = 2 case requires substantially more work.
First of all, since this is a composite system (c.f.
definition 3.5), the individual questions about the
two single gbits must be contained in an infor-
mationally complete set Q
M
2
. We shall again
slightly change the notation, as compared to sec-
tion
4.1: the maximal complementary set Q
M
1
(
11) for gbit 1 will be denoted by Q
1
, . . . , Q
D
1
,
while the elements of Q
M
1
for gbit 2 will be de-
noted with a prime Q
1
, . . . , Q
D
1
. It will be conve-
nient to depict the q uestion structure graphically.
Representing individual questions henceforth as
vertices, Q
M
2
certainly contains the following
PSfrag replacements
gbit 1 gbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
.
.
.
.
.
.
.
.
.
. (12)
This structure abides by rules
1 and 2 and any Q
i
will be independent of and maximally compatible
with any Q
j
(cf. definition
3.5).
5.2.1 Logical connectives of single gbit questions
But a composite system may also admit compos-
ite questions, built with logical connectives of the
subsystem questions. We must now determine
which composite questions are allowed and which
must be added to the individual questions in or-
der to form an informationally complete Q
M
2
of
pairwise independent questions. To this end, we
must firstly unravel which logical connective O
can employ at all in order to construct compos-
ite questions which are pairwise independent of
Accepted in Quantum 2017-11-27, click title to verify 26
the individual ones. Recall assumption 5 that
O, according to his theoretical model, may only
comp os e compatible questions with logical con-
nectives and consider Q
i
, Q
j
as an exemplary
compatible pair. In truth tables, we shall hence-
forth symbolize ‘yes’ by 1 and ‘no’ by 0’.
Remark. (‘= in logical expressions.) To
avoid confusion, we emphasize that whenever we
write an equality sign = in a logical expression
in the sequel, we do not interpret it as another
logical connective , but as an actual eq uality of
the values of the propositions on the left and right
hand side. For example, the expression Q
i
= Q
j
is not itself a proposition which takes truth val-
ues 0 or 1, but is intended to mean that the truth
value which Q
i
takes is identical to that which Q
j
takes.
While permitted binary connectives, it is clear
that, e.g., the AND and OR operations and ,
respectively, cannot be employed alone to build
a new independent q uestion from Q
i
, Q
j
because
Q
i
Q
j
= 1 implies Q
i
= Q
j
= 1 and Q
i
Q
j
= 0
implies Q
i
= Q
j
= 0 such that certain answers
to these two connectives imply answers to the in-
dividuals. But if Q
i
Q
j
is to be pairwise inde-
pendent from Q
i
, Q
j
, an answer to it must not
imply the answer to either of Q
i
, Q
j
and, con-
versely, the answer to only Q
i
or only Q
j
cannot
imply Q
i
Q
j
. As can be easily checked, it must
satisfy the following truth table:
Q
i
Q
j
Q
i
Q
j
0 1 a
1 0 a
1 1 b
0 0 b
a 6= b a, b {0, 1}. (13)
The only logical connectives satisfying this
truth table are the XNOR for a = 0 and b = 1
and its negation, the XOR for a = 1 and b = 0.
As Q
i
Q
j
= ¬(Q
i
Q
j
), we may only choos e
one of the two binary connectives to define new
pairwise independent questions. We henceforth
choose to use the XNOR , already introduced in
(
2), and operationally interpret it as a ‘correlation
question’; Q
ij
:= Q
i
Q
j
is to be read as ‘are
the answers to Q
i
and Q
j
the same?’. But the
XOR could equivalently be employed.
28
Since, according to assumption 7, O must be
able to build an informationally complete set
for a composite system which, in particular, in-
volves pairwise independent composite questions,
we would like to conclude that Q
i
Q
j
are not
only implementable, but also contained in a Q
M
2
together with the individuals. However, this is
subject to a few consistency checks on pairwise
independence.
The XNOR is a symmetric logical connec-
tive, Q
ij
= Q
i
Q
j
= Q
j
Q
i
(but note that
Q
ij
6= Q
ji
), and, thanks to its associativity,
Q
i
Q
ij
= Q
i
(Q
i
Q
j
)
= (Q
i
Q
i
)
|
{z }
1
Q
j
= Q
j
,
Q
ij
Q
j
= Q
i
(14)
such that defines a closed relation on
Q
i
, Q
j
, Q
ij
. This has the following ramifications:
(1) For any compatible pair Q
i
, Q
j
, the ‘correla-
tion’ operation gives rise to precisely one ad-
ditional question Q
ij
, and (2) the set Q
i
, Q
j
, Q
ij
will indeed be pairwise independent, according to
the definition in section
3.2.4, because neither Q
i
nor Q
j
alone can determine Q
ij
relative to the
state of no information (not even partially), oth-
erwise, together with the determined Q
ij
, they
would (at least partially) determine each other
via (
14) in contradiction with the independence
of Q
i
, Q
j
. That is, Q
ij
is independent of both Q
i
and Q
j
and since we assume independence to be a
symmetric relation, it must also be true the other
way around.
Consequently, the D
2
1
correlation questions
Q
ij
, i, j = 1, . . . , D
1
, are candidates for addi-
tional questions in Q
M
2
besides the 2D
1
indi-
vidual ones of (
12). Graphically, we shall repre-
sent Q
ij
as the edge connecting the vertices cor-
responding to Q
i
and Q
j
. For example, the fol-
lowing question graphs
28
As an aside, the XNOR can be expressed in terms of
the basic Boolean op erations as Q
i
Q
j
= (¬Q
i
Q
j
)
(Q
i
¬Q
j
).
Accepted in Quantum 2017-11-27, click title to verify 27
PSfrag replacements
gbit 1 gbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
31
Q
22
Q
23
Q
D
1
D
1
.
.
.
.
.
.
.
.
.
PSfrag replacements
gbit 1 gbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
31
Q
22
Q
23
Q
D
1
D
1
.
.
.
.
.
.
.
.
.
PSfrag replacements
gbit 1 gbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
31
Q
22
Q
23
Q
D
1
D
1
.
.
.
.
.
.
.
.
.
represent legal sets of questions. Note that due
to assumption
5, there can only be edges between
gbit 1 and gbit 2 but not between the vertices of
a single gbit, e.g., Q
1
and Q
2
, because they are
complementary.
Our next task is to analyse the independence
and compatibility structure of the Q
ij
.
5.2.2 Independence, complementarity and entan-
glement
We begin with a simple, but important observa-
tion which we will frequently make use of.
Lemma 5.1. Let Q
1
, Q
2
, Q
3
Q
N
be such that
Q
1
is maximally compatible with and pairwise in-
dependent of both Q
2
, Q
3
and that Q
3
is maxi-
mally complementary to Q
1
Q
2
. Then Q
2
, Q
3
are maximally complementary.
Proof. Suppose Q
2
, Q
3
were at least partially
compatible. In that case there must exist a state
of S in which O has maximal information α
2
= 1
bit about Q
2
and at least partial information
α
3
> 0 bit about Q
3
(or vice versa). Let O
now ask Q
1
which is maximally compatible with
and pairwise independent of both Q
2
, Q
3
to S
in this state. Hence, according to assumption
4,
by asking Q
1
, O cannot change his in formation
about either Q
2
, Q
3
. Thus, after asking Q
1
, O
will have m aximal information about Q
1
, Q
2
and
at least partial in formation about Q
3
. But max-
imal information about Q
1
, Q
2
implies maximal
information about Q
1
Q
2
which is maximally
complementary to Q
3
. Consequently, Q
2
and Q
3
must be maximally complementary too.
This has immediate useful implications.
Lemma 5.2. Q
i
is maximally compatible with
Q
ij
, j = 1, . . . , D
1
and maximally complemen-
tary to Q
kj
, k 6= i and j = 1, . . . , D
1
. That
is, graphically, an individual question Q
i
is max-
imally compatible with a correlation question Q
ij
if and only if its corresponding vertex is a vertex
of the edge corresponding to Q
ij
. By symmetry,
the analogous result holds for Q
j
.
Proof. Q
i
and Q
ij
are maximally compatible by
construction. Consider therefore Q
i
and Q
kj
for
k 6= i and j = 1, . . . , D
1
. Clearly, Q
i
and Q
j
are
maximally compatible and pairwise independent
and so are Q
j
and Q
kj
. Maximal complemen-
tarity of Q
i
and Q
kj
for k 6= i now follows from
Q
k
= Q
kj
Q
j
(see (
14)), which is maximally
complementary to Q
i
, an d lemma
5.1.
For example, Q
1
and Q
12
are maximally com-
patible, while Q
1
and Q
22
are maximally com-
plementary. The intuitive explanation for the
incompatibility of Q
1
and Q
22
is as follows: if
O knew the answers to both simultaneously, he
would know more than one bit of information
about gbit 1 because Q
1
defines a full bit of in-
formation about it while Q
22
could be regarded
as defining half a bit about each of gbit 1 and
2. But in view of rule
1, O should never know
more than one bit about a single gbit, even in a
comp os ite system. Considering the information
contained in correlation questions to equally cor-
respond to gbit 1 and 2, lemma
5.2 suggests that
O’s information will always be equally distributed
over the two gbits for states of maximal knowl-
edge, i.e. O will know eq ually much about gbit 1
and 2.
29
We shall make this more precise in [1 ].
We saw before that Q
i
, Q
j
, Q
ij
are pairwise in-
dependent. Lemma
5.2 implies that also Q
i
and
Q
kj
for i 6= k are independent. We can make use
of this result to show the following.
Lemma 5.3. The Q
ij
, i, j = 1, . . . , D
1
are pair-
wise independent.
Proof. Consider Q
ij
and Q
kl
and suppose i 6=
k. Then Q
i
and Q
ij
are maximally compatible,
29
Clearly, this is not tru e for non-maximal knowledge.
E.g., let O always ask only Q
1
.
Accepted in Quantum 2017-11-27, click title to verify 28
while Q
i
and Q
kl
are maximally complementary
by lemma
5.2. Hence, when knowing Q
i
and Q
ij
O cannot know Q
kl
such that Q
ij
and Q
kl
must
be independent. (Recall from section
3.2.4 that
dependence requires the answer to Q
ij
to always
imply at least partial knowledge about Q
kl
, i.e.,
y
ij
= 0, 1 implies y
kl
6=
1
2
). The analogous argu-
ment holds for when j 6= l.
Recalling from subsection 5.2.1 that the set of
comp os ite questions in Q
M
2
must be non-empty,
lemmas
5.2 and 5.3 have an important corollary.
Corollary 5.4. Q
i
, Q
j
, Q
ij
are pairwise indepen-
dent for all i, j = 1, . . . , D
1
and, thanks to as-
sumption
7, will thus be part of an information-
ally complete set Q
M
2
.
Next, we consider the compatibility and com-
plementarity structure of the correlation ques-
tions Q
ij
.
Lemma 5.5. Q
ij
and Q
kl
are maximally com-
patible if and only if i 6= k and j 6= l. That is,
graphically, Q
ij
and Q
kl
are maximally compat-
ible if their corresponding edges do not intersect
in a vertex and maximally complementary if they
intersect in one vertex.
Proof. Suppose Q
ij
and Q
kj
with i 6= k were
at least partially compatible. Then there must
exist a state of S in which the answer to Q
ij
is fully known to O and in which also α
kj
> 0
bit (or vice versa). Let O ask Q
j
to S in this
state. Since, by lemma
5.2, this question is max-
imally compatible with both Q
ij
, Q
kj
(and pair-
wise independ ent of both), according to assump-
tion
4, by asking Q
j
, O cannot change his infor-
mation about both Q
ij
, Q
kj
. T hat is, after hav-
ing also asked Q
j
, O must have maximal informa-
tion about Q
ij
, Q
j
and partial inform ation about
Q
kj
. But maximal information about Q
ij
, Q
j
im-
plies also maximal information about Q
i
which,
however, is maximally complementary to Q
kj
by
lemma
5.2. Hence, Q
ij
, Q
kj
with i 6= k must be
maximally complementary.
Consider now Q
ij
and Q
kl
with i 6= k and j 6= l.
Let O ask S both Q
i
and Q
j
the answers to which
imply the answer to Q
ij
according to (
2). Thanks
to rule
1, this defi nes a state of maximal knowl-
edge of 2 independent bits and O may not know
any further independent information. Next, let
O ask the same S the question Q
kl
. From lemma
5.2 it follows that Q
kl
is maximally complemen-
tary to both Q
i
and Q
j
such that the answer to
Q
kl
will give one new independent bit of infor-
mation but renders O’s information about Q
i
, Q
j
obsolete. But by rule
2 O cannot experience a net
loss of information by asking a new question and
after asking Q
kl
he must still know 2 bits about
S. Hence, after acquiring the answer to Q
kl
he
must still know the answer to Q
ij
such that both
are m aximally compatible.
To give graphical examples, Q
11
and Q
21
are
maximally complementary due to the intersection
in Q
1
, while Q
13
and Q
21
are maximally compati-
ble because their edges do not intersect in vertices
PSfrag replacements
gbit 1
gbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
21
Q
22
Q
13
Q
D
1
D
1
.
.
.
PSfrag replacements
gbit 1
gbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
21
Q
22
Q
13
Q
D
1
D
1
.
.
.
This question structure has significant conse-
quences: since the correlation questions Q
ij
and
Q
kl
are both independent and maximally com-
patible for i 6= k and j 6= l, O can ask both of
them simultaneously, thereby spending the maxi-
mal amount of N = 2 independent bits of infor-
mation he may acquire according to rule
1 over
comp os ite questions. For example, he may ask S
Q
11
and Q
22
simultaneously
PSfrag replacements
gbit 1
gbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
21
Q
22
Q
13
Q
D
1
D
1
.
.
.
(15)
upon which he must be entirely oblivious about
the individual gbit properties represented by
Q
i
, Q
j
. (Lemma
5.2 implies that no individual
question is maximally compatible with two non-
intersecting bipartite questions at once.) That is,
O has only composite, but no individual infor-
mation about the two gbits; they are maximally
correlated. But this is precisely entanglement.
Indeed, the question graph (
15) will ultimately
correspond to four Bell states in either of the two
Accepted in Quantum 2017-11-27, click title to verify 29
question bases {Q
1
, Q
1
} and {Q
2
, Q
2
}, represent-
ing the four p os sible answer configurations ‘yes-
yes’, ‘yes-no’, ‘no-yes’ and ‘no-no’ to the corre-
lation questions Q
11
, Q
22
(see [
31] for a related
perspective within quantum theory). In fact, if
O now ‘marginalized’ over gbit 2, corresponding
to discarding any information about questions in-
volving gbit 2, he would find gbit 1 in the state
of no information relative to him because the
discarded inf ormation also contained everything
he knew about gbit 1. Other compatible edges
will correspond to Bell states in different ques-
tion bases.
Inspired by the compelling proposal within
quantum theory in [
33,34], we shall define s tates
~y
OS
of a bipartite gbit system in L
gbit
as
entangled if the composite information satisfies
P
D
1
i,j=1
α
ij
> 1 bit,
where α
ij
(~y
OS
) is O’s information about the
correlation question Q
ij
. While there will be
states with
P
D
1
i,j=1
α
ij
1 bit which can indeed
be considered as classically composed,
30
in con-
trast to [
33,34], we emphasize that the above will
only be a sufficient, but not a necessary condi-
tion for entanglement, as in the quantum case
there will exist entangled states violating it.
31
It
30
E.g., consider a state with y
11
= 1 and y
i
=
1
2
for all
other questions in Q
M
2
, corresponding to O knowing with
certainty that Q
11
= 1 and nothing else. The discussion
surrounding (
1) implies already that
P
D
1
i,j=1
α
ij
= 1 bit,
while the individual information is zero. In quantum the-
ory, this state would correspond to the separable state
ρ =
1
4
(1 + σ
x
σ
x
).
31
For the quantum case, states with
P
D
1
=3
i,j=1
α
ij
1 bit
where defin ed as classically composed in [
33, 34]. This is,
however, in conflict with the usual definition of entangled
states as those which are non-separable. Namely, consider
the family of quantum states (on the r.h.s. expressed in
z-basis) for
1
3
a
1
3
ρ
a
:=
1
4
1 +
1
3
(σ
z
1 + 1 σ
z
σ
z
σ
z
)
+2a(σ
x
σ
x
+ σ
y
σ
y
)) =
1
3
0 0 0
0
1
3
a 0
0 a
1
3
0
0 0 0 0
.
This clearly is a family of non-separable states for a 6= 0.
In particular, for a =
1
3
this is the reduced state which
one obtains from a W-state upon marginalizing over a
qubit. These states y ield y
z
1
= y
z
2
=
2
3
, y
zz
=
1
3
and
y
xx
= y
yy
= a +
1
2
and all other y
i
=
1
2
, where, e.g., y
z
1
and y
zz
are th e probabilities that the spin of qu bit 1 is
up in z-direction and that th e spins of both qubits are
would be an interesting question to investigate
what the precise sufficient and necessary condi-
tions for entanglement are in the quantum case
in this informational formulation. However, we
shall leave this open for now.
Since there are only N = 2 independent bits of
information to be gained, according to rule
1, the
above sufficient condition for entanglement thus
means, in particular, that O has more compos-
ite than individual information. In general, two
gbits will be referred to as maximally entangled
relative to O if he has only composite informa-
tion about them which is incompatible with any
individual information. By contrast, two gbits
would be in a ‘product state’ if O has maximal
individual knowledge of 2 bits about them. For
instance, if he knows that Q
1
= Q
1
= 1, S would
be in such a product state relative to him. But
notice that even in this case, O would have one
dependent bit of information about the correla-
tions because clearly Q
11
= 1 too.
Note also that O can ‘collapse’ two gbits into
an entangled state: as the most extreme exam-
ple, consider the case that O receives a ‘product
ensemble state’ of two gbits in a multiple shot
interrogation. O may then ask the next pair of
gbits the questions Q
11
, Q
22
. Upon receiving the
answers, the two gbits will have ‘collapsed’ into
a maximally entangled posterior state relative to
O, despite having been in a ‘product state’ prior
to interrogation. Entanglement is thus a property
of O’s information about S.
We emphasize that, for systems with limited in-
formation content, entanglement is a direct con-
sequence of complementarity at least for states
of maximal information. To illustrate this obser-
vation, consider two classical bits (cbits). Since
the same in z-direction, resp ectively. Using the quadratic
information measure α
ij
= (2 y
ij
1)
2
which we derive in
subsection
6.8 and which was also proposed for quantum
theory by the authors of [
33, 34], these states give
I
comp
:=
3
X
i,j=1
α
ij
= (2 y
xx
1)
2
+(2 y
yy
1)
2
+(2 y
zz
1)
2
+0 · · · + 0 =
1
9
+ 8a
2
bit 1 bit,
and for the individual information I
indiv
:= (2y
z
1
1)
2
+
(2y
z
2
1)
2
+ 0 · · · + 0 = 2/9 bit. I
comp
can be arbitrar-
ily close to 1/9 bit for a sufficiently small a 6= 0 so that
even I
comp
< I
indiv
, and yet the corresponding state is en-
tangled in violation of the p roposed condition of classical
composition in [
33, 34].
Accepted in Quantum 2017-11-27, click title to verify 30
there is no complementarity in this case, there are
only three pairwise independent questions: the
individuals Q
1
, Q
1
about cbit 1 and cbit 2, re-
spectively, and the correlation Q
11
would form an
informationally complete Q
M
N
, graphically rep-
resented as
PSfrag replacements
cbit 1 cbit 2
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
21
Q
22
Q
13
Q
D
1
D
1
.
.
.
Q
1
, Q
1
, Q
11
are mutually maximally compatible
and this composite system satisfies rule
1 too such
that O may only acquire maximally N = 2 inde-
pendent bits of information about this pair of
cbits. Clearly, also in this classical case it is pos-
sible for O to acquire only composite information
about the pair of cbits, namely by only asking
Q
11
and not bothering about the individuals. O
would be able to find out whether identically pre-
pared pairs of cbits in an ensemble are correlated
by alway s asking Q
11
in a multiple shot interro-
gation. There will indeed exist states such that
y
11
= 1 and y
1
= y
1
=
1
2
, however, the pair of
cbits will not be in a state of maximal informa-
tion relative to O because he has only spent one of
his two available bits. A classical state of max-
imal information corresponds to O knowing the
answers to two questions of the three Q
1
, Q
1
, Q
11
,
but then by (
2, 14) O will also know the answer
to the third. That is, in a state of maximal in-
formation about two cbits, O will always have
maximal individual information about the pair.
For cbits, O cannot spend the second bit also in
comp os ite information. By contrast, it is a conse-
quence of the complementarity rule 2 that in our
case here, O can actually exhaust the entire infor-
mation available to him by composite questions,
thereby giving rise to entanglement.
Thus far, we have only considered individual
and correlation questions. All are part of Q
M
2
.
But can there be more pairwise independent ques-
tions in Q
M
2
? These would have to be obtained
from the individuals and correlations via another
comp os ition with an XNOR. Composing the indi-
viduals Q
i
, Q
j
with the correlation questions via
XNOR, e.g. Q
i
Q
il
, will yield nothing new
because questions need to be maximally compat-
ible in order to be logically connected by O and
for compatible pairs it already follows from (
14)
that the individual and correlation questions are
logically closed under . However, what about
correlations of correlation questions, e.g., what
about Q
11
Q
22
?
5.2.3 A logical argument for the dimensionality of
the Bloch-sphere
The answer to the last question, in fact, is inter-
twined with the dimensionality D
1
of the state
space Σ
1
of a single gbit and thus, ultimately,
with the dimensionality of the Bloch-sphere. De-
riving the dimension of the Bloch sphere has also
been a crucial step in the successful reconstruc-
tions of quantum theory via the G PT framework.
While H ardy’s pioneering reconstruction [
14] did
not fully settle this issue (it required an opera-
tionally somewhat obscure ‘simplicity axiom’ to
obtain D
1
= 3), it was later explicitly solved
by impressive group-theoretic arguments in [59]
which show that, under certain information the-
oretic constraints, qubit quantum theory is the
only theory with non-trivial entangling dynam-
ics. This result was then exploited to relate the
dimension of the Bloch-sphere to the dimension
of space via a communication thought experi-
ment [
21]. However, unfortunately, the ingredi-
ents of these arguments neither fit mathemati-
cally nor conceptually fully into our framework
which relies on distinct structures than GPTs.
On the other hand, although not yielding full
quantum theory reconstructions, approaches as-
serting an epistemic restriction (i.e. a restriction
of knowledge) over ontic states admit very sim-
ple and elegant arguments for a 1-bit system to
have a state space spanned by three indep endent
epistemic states [
40, 41, 54]. These arguments
can be carried out by considering a single sy stem
over a 2-bit ontic state which at the epistemic
level, however, is effectively a 1-bit system due
to epistemic restrictions. The arguments involve
binary connectives (at the ontic level) of comple-
mentary questions which, while non-problematic
when dealing with ‘hidden variables’, are, how-
ever, deemed illegal in O’s model of the world ac-
cording to our assumption
5. We do not wish to
make any ontological commitments in our purely
operational approach. Consequently, we must
reason differently and without ontic states to de-
duce the dimensionality of Σ
1
. In our case, this
requires to consider two gbits rather than just
one and involves entanglement, in analogy to the
group-theoretic GPT derivation in [
21,59] which
likewise requires two gbits and entanglement.
Accepted in Quantum 2017-11-27, click title to verify 31
Theorem 5.6. D
1
= 2 or 3.
Proof. By lemma
5.5, Q
ij
and Q
kl
are indepen-
dent and maximally compatible if i 6= k and
j 6= l. But any maximal set of pairwise maxi-
mally compatible correlation questions then con-
tains precisely D
1
questions. Graphically, this
is easy to see: one can always find D
1
non-
intersecting edges between the D
1
vertices of gbit
1 and the D
1
vertices of gbit 2. For example, the
D
1
‘diagonal correlations’ Q
ii
, i = 1, . . . , D
1
PSfrag rep lacements
.
.
.
Q
11
Q
22
Q
D
1
D
1
Q
33
are pairwise maximally compatible and pairwise
independent (lemma
5.3) su ch that, by theorem
3.2 (Specker’s principle), they must also be mu-
tually maximally compatible. Hence, O m ay ac-
quire the answers to all D
1
Q
ii
at the same time.
However, rule
1 forbids O’s information about
S to exceed the limit of N = 2 independent
bits. Accordingly, the D
1
correlation questions
Q
ii
cannot be mutually independent if D
1
> 2
and the answers to any pair of them must imply
the answers to all others. For instance, suppose
O has asked Q
11
and Q
22
. The pair must de-
termine the answers to Q
33
, . . . , Q
D
1
D
1
such th at
the latter must be Boolean f unctions of Q
11
, Q
22
.
Hence, Q
jj
= Q
11
Q
22
, j 6= 1, 2, for some logical
connective w hich preserves that Q
11
, Q
22
, Q
jj
are pairwise independ ent. Table (
13) implies that
must either be th e XNOR or the XOR
such that either
Q
jj
= Q
11
Q
22
, or
Q
jj
= ¬(Q
11
Q
22
), j = 3, . . . , D
1
.
But th en clearly Q
jj
, j = 3, . . . , D
1
could not
be pairwise independent if D
1
> 3. The same
argument can be carried out for any set of D
1
pairwise independent and maximally compatible
correlations Q
ij
, i.e. for any qu estion graph with
D
1
non-intersecting edges, and any choice of a
pair of maximally compatible correlations which
O may ask rst. We must therefore conclude that
D
1
3, while rule
2 implies that D
1
2.
We shall refer to D
1
= 2 as the ‘rebit case’ and
to D
1
= 3 as the ‘qubit case’; for the moment
this is just suggestive terminology, however, we
shall see later and in [
2] that the D
1
= 2 case
will, indeed, result in rebit quantum theory (two-
level s ystems over real Hilbert spaces), while the
D
1
= 3 case will give rise to s tandard q uantum
theory of qubit systems [1].
5.2.4 An informationally complete set for N = 2
qubits
The two cases have distinct properties as regards
questions which as k for the correlations of bi-
partite questions and it is necessary to consider
them separately. We begin with the simpler qubit
case D
1
= 3 for which correlations of correlation
questions do not define new information f or O
about S such that six individual and nine cor-
relation questions constitute an informationally
complete set Q
M
2
. These will ultimately cor-
respond to the propositions ‘the spin is up in
x, y, zdirection’ for the individual qubits 1
and 2 and ‘the spins of qubit 1 in i and qubit 2
in j direction are the same’ where i, j = x, y, z.
Note that the density matrix for two qubits has
15 parameters.
Theorem 5.7. (Qubits) If D
1
= 3 then the correlation questions Q
ij
are logically closed under
and Q
M
2
= {Q
i
, Q
j
, Q
ij
}
i,j=1,2,3
constitutes an informationally complete set for N = 2 with D
2
= 15.
Furthermore, f or any two permutations σ, σ
of {1, 2, 3} either
Q
σ(3)σ
(3)
= Q
σ(1)σ
(1)
Q
σ(2)σ
(2)
, or Q
σ(3)σ
(3)
= ¬(Q
σ(1)σ
(1)
Q
σ(2)σ
(2)
) (16)
and either
Q
σ(3)σ
(3)
= Q
σ(1)σ
(2)
Q
σ(2)σ
(1)
, or Q
σ(3)σ
(3)
= ¬(Q
σ(1)σ
(2)
Q
σ(2)σ
(1)
) (17)
Accepted in Quantum 2017-11-27, click title to verify 32
such that either
Q
σ(1)σ
(1)
Q
σ(2)σ
(2)
= Q
σ(1)σ
(2)
Q
σ(2)σ
(1)
, or
Q
σ(1)σ
(1)
Q
σ(2)σ
(2)
= ¬(Q
σ(1)σ
(2)
Q
σ(2)σ
(1)
). (18)
Proof. Statements (16, 17) are an immediate
consequence of the argument in the proof of the-
orem 5.6 which can be applied to any correla-
tion question graph with D
1
= 3 non-intersecting
edges and thus to any two permutations σ, σ
of
{1, 2, 3}. From this it directly follows that the
correlation questions Q
ij
are logically closed u n-
der such that in the case D
1
= 3 correlations
of correlations such as Q
ij
Q
kl
do not define
any n ew independent questions. Accordingly,
Q
M
2
= {Q
i
, Q
j
, Q
ij
}
i,j=1,2,3
is an informationally complete set of questions
for N = 2 and D
1
= 3 such that D
2
= 15.
For example, if σ, σ
are both trivial then the
theorem applied to the following two question
graphs
PSfrag replacements
.
.
.
Q
11
Q
22
Q
12
Q
21
Q
33
PSfrag replacements
.
.
.
Q
11
Q
22
Q
12
Q
21
Q
33
implies for D
1
= 3 either
Q
33
= Q
11
Q
22
, or Q
33
= ¬(Q
11
Q
22
) (19)
and either
Q
33
= Q
12
Q
21
, or Q
33
= ¬(Q
12
Q
21
) (20)
such that either
Q
11
Q
22
= Q
12
Q
21
, or Q
11
Q
22
= ¬(Q
12
Q
21
) (21)
In sections 5.2.7 and 5.4 we shall determine
whether negations ¬ occur in the expressions (
16
18). For Q, Q
, Q
′′
maximally compatible and re-
lated by an XNOR, we shall henceforth distin-
guish between
even correlation: if Q = Q
Q
′′
, and
odd correlation: if Q = ¬(Q
Q
′′
).
32
5.2.5 An informationally complete set for N = 2
rebits
Next, let us consider the rebit case D
1
= 2. O can
ask the four individual questions Q
1
, Q
2
, Q
1
, Q
2
and the four correlations Q
11
, Q
12
, Q
21
, Q
22
. But
O can also define the two new correlation of cor-
relations questions
Q
33
:= Q
12
Q
21
,
˜
Q
33
:= Q
11
Q
22
(22)
32
We note that this also implies Q
= ¬(Q Q
′′
) and
Q
′′
= ¬(Q Q
).
corresponding to the two correlation questions
graphs
PSfrag replacements
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
21
Q
22
Q
12
Q
33
˜
Q
33
.
.
.
PSfrag replacements
Q
1
Q
2
Q
3
Q
D
1
Q
1
Q
2
Q
3
Q
D
1
Q
11
Q
21
Q
22
Q
12
Q
33
˜
Q
33
.
.
.
(we add the correlation of correlations as a new
edge without vertices to the graph). The dif-
ference to the qubit case is that Q
33
,
˜
Q
33
can
now not b e written as correlations of individ-
ual questions Q
3
, Q
3
since the latter do not ex-
ist. Clearly, Q
12
, Q
21
, Q
33
and Q
11
, Q
22
,
˜
Q
33
are
two pairwise independent sets. The question is
whether Q
33
,
˜
Q
33
are independent of each other
and of the individuals. The following lemma even
asserts complementarity to the latter. (Note that
this lemma holds trivially in the case D
1
= 3.)
Accepted in Quantum 2017-11-27, click title to verify 33
Lemma 5.8. The questions asking for the cor-
relation of the bipartite correlations, Q
12
Q
21
and Q
11
Q
22
, are maximally complementary
to and independent of any Q
1
, Q
2
, Q
1
, Q
2
.
Proof. We firs tly demonstrate independence of
Q
1
and Q
11
Q
22
. This f ollows from not-
ing that Q
22
is maximally complementary to
Q
1
(lemma
5.2) and maximally compatible with
Q
11
Q
22
and arguments completely analogous
to those in the proof of lemma 5.3. Next, we es-
tablish complementarity of Q
11
Q
22
and Q
1
.
Suppose the contrary, namely, that Q
11
Q
22
and Q
1
were at least partially compatible such
that there must exist a state with α
Q
11
Q
22
= 1
bit and α
1
> 0 bit. Clearly, both are maxi-
mally compatible with and independent of Q
11
.
According to assumption
4, asking Q
11
to this
state does not change O’s information about
Q
11
Q
22
and Q
1
. However, maximal infor-
mation about Q
11
Q
22
and Q
11
also implies
maximal information about Q
22
which is max-
imally complementary to Q
1
. Hence, Q
1
and
Q
11
Q
22
must be maximally complementary.
The arguments works analogously for the other
cases.
Finally, we show that (1921) hold analogously
for the rebit case D
1
= 2. From this it follows
that the four individual questions Q
i
, Q
j
, the four
correlations Q
ij
, i, j = 1, 2, and the correlation
of correlations Q
33
form an informationally com-
plete set Q
M
2
for rebits. That is, in contrast
to the qubit case, there is now one non-trivial
correlation of correlations question which is pair-
wise independent from the individuals and corre-
lations.
Theorem 5.9. (Rebits) If D
1
= 2 then Q
M
2
=
{Q
i
, Q
j
, Q
ij
, Q
33
}
i,j=1,2
constitutes an informa-
tionally complete set for N = 2 with D
2
= 9 and
either
Q
11
Q
22
= Q
12
Q
21
, or
Q
11
Q
22
= ¬(Q
12
Q
21
).
Proof. O can begin by asking S the maximally
compatible Q
11
and Q
22
upon which he also
knows the answer to
˜
Q
33
. O then possesses the
maximal amount of N = 2 independent bits
of information about S. Next, O can ask Q
12
(or Q
21
) which, according to lemma
5.5, is max-
imally complementary to Q
11
, Q
22
. But, by rule
2, O is not allowed to experience a net loss of
information. Hence, Q
12
(or Q
21
) must be max-
imally compatible with
˜
Q
33
(they are also in-
dependent). That is, Q
12
, Q
21
,
˜
Q
33
are mutu-
ally maximally compatible according to theorem
3.2 (Specker’s principle) and O may also ask all
three at the same time. But then the same ar-
gument as in the proof of theorem 5.6 applies
such that either
˜
Q
33
= Q
12
Q
21
or
˜
Q
33
=
¬(Q
12
Q
21
). This implies either Q
33
=
˜
Q
33
or
Q
33
= ¬
˜
Q
33
. Accordingly, there is only one in-
dependent correlation of correlations Q
33
. Since
Q
11
, Q
12
, Q
21
, Q
22
and Q
33
are then, by construc-
tion, logically closed under the XNOR and Q
33
is maximally complementary to all individuals
Q
1
, Q
2
, Q
1
, Q
2
, no fu rther pairwise independent
question can be built from this set such that it is
informationally complete. Hence, D
2
= 9.
5.2.6 Entanglement and non-local tomography for
rebits
The rebit case D
1
= 2 thus has a very spe-
cial question and correlation structure: Q
33
=
Q
12
Q
21
is the only composite question which
is maximally complementary to all individuals,
but maximally compatible with all correlations
Q
ij
. By contrast, e.g., Q
11
is maximally comple-
mentary to Q
2
, Q
2
, Q
12
, Q
21
and maximally com-
patible with Q
1
, Q
1
, Q
22
and maximally compati-
ble with the correlation of correlations Q
33
. Con-
sequently, Q
33
assumes a special role in the en-
tanglement structure: once O knows the answer
to this question he may no longer have any fur-
ther information about the outcomes of individ-
ual questions, but may have information about
correlation questions. That is, two rebits can be
in a state of non-maximal information of 1 bit
relative to O, corresponding to the latter only
having maximal k nowledge about the answer to
Q
33
and no information otherwise, and still be
maximally entangled because any individual in-
formation is forbidden in that s ituation. Notice
that this is not true for pairs of qubits because
even if everything O knew about the pair was the
answer to Q
33
, he could still acquire information
about the individuals Q
3
, Q
3
such that one could
not consider such a state as maximally entangled.
Even stronger, O will always know the answer
to Q
33
if the two rebits are maximally entangled
and he has maximal information about S. This
follows from (
6): for a maximally entangled state
Accepted in Quantum 2017-11-27, click title to verify 34
of maximal information (no information about in-
dividuals), the following must hold
I
OS
=
2
X
i=1
(α
i
+ α
i
)
|
{z }
=0
+
2
X
i,j=1
α
ij
+ α
33
= α
11
+ α
12
+ α
21
+ α
22
+ α
33
= 3 bits.
The last equality follows from the fact that once
O knows the answers to two maximally compat-
ible questions, he will also k now the answer to
their correlation and since this correlation is in
the pairwise independent question set the total
information defined in (
6) will always yield 3 bits
for N = 2 independent bit systems in states of
maximal knowledge.
33
If we now require that O
cannot k now more than a single bit about max-
imally complementary questions then, in a state
of maximal information of N = 2 independent
bits, we must have
α
11
+ α
12
= α
21
+ α
22
= 1 bit
α
33
= 1 bit
such that O must have maximal information
about Q
33
. Therefore, the correlation of corre-
lations Q
33
can be v iewed as the litmus test f or
entanglement of two rebits.
We note that the individuals Q
1
, Q
2
, Q
1
, Q
2
will ultimately correspond to projections onto the
+1 eigenspaces of σ
x
, σ
z
, while the Q
ij
corre-
spond to projections onto the +1 eigenspaces of
σ
i
σ
j
, i, j = x, z, and Q
33
corresponds to the
projection onto the + 1 eigenspace of σ
y
σ
y
,
where σ
x
, σ
y
, σ
z
are the Pauli matrices [
1,2]. Ob-
servables and density matrices on a real Hilbert
space correspond to real symmetric matrices.
This is the reason why σ
y
is not an observable
on R
2
(it corresponds to the ‘missing’ Q
3
), but
σ
y
σ
y
is a real symmetric matrix and thus an
observable on R
2
R
2
.
This gives a novel and simple explanation for
the discovery that σ
y
σ
y
determines the entan-
glement of rebits [
83]: a two-rebit density matrix
ρ is separable if and only if Tr(ρ σ
y
σ
y
) = 0,
i.e. if the state has no σ
y
σ
y
comp onent. This
statement means in our language that a state is
separable if and only if α
33
= 0 and is consistent
with our observation above because any informa-
tion about the individuals is incompatible with
information about Q
33
due to complementarity.
33
We are implicitly using here also rules
3 and 4.
This question structure also has severe reper-
cussions for rebit state tomography: it must ul-
timately be non-local. For rebits, D
1
= 2, the
probability that Q
33
= 1 could be written as
y
33
= p(Q
12
= 1, Q
21
= 1) + p(Q
12
= 0, Q
21
= 0)
where p(Q
i
, Q
j
) denotes here the joint probabil-
ity distribution over Q
i
, Q
j
. That is, in a multiple
shot interrogation, O could ask both Q
12
, Q
21
to
the identically prepared rebit couples and from
the statistics over the answers, he could also de-
termine y
33
. But this probability cannot be de-
comp os ed into joint probabilities over the individ-
ual q uestions according to Q
33
= “(Q
1
Q
2
)
(Q
2
Q
1
)” because Q
1
, Q
2
and Q
1
, Q
2
are max-
imally complementary. Therefore, O would not
be able to determine y
33
by only asking individ-
ual questions to the two rebits and the statis-
tics over these answers.
34
That is, for rebits,
state tomography would always require correla-
tion questions and in this sense be non-local.
Note that this stands in stark contrast to qubit
pairs where Q
33
= Q
3
Q
3
can be written
in terms of individual questions such that also
y
33
= p(Q
3
= 1, Q
3
= 1) + p(Q
3
= 0, Q
3
= 0) can
be determined by the statistics over the answers
to Q
3
, Q
3
only. Qubit systems (and quantum the-
ory in general) are thus tomographically local.
The requirement of tomographic locality, ac-
cording to which the state of a composite system
can be determined by doing statistics over mea-
surements on its subsystems, is a standard con-
dition in the GPT framework [
1518, 20, 55, 59]
and thus directly rules out rebit theory. How-
ever, in contrast to derivations within the GPT
landscape, we shall not implement local tomog-
raphy here because it will be interesting to see
the differences between real and complex quan-
tum theory from the perspective of information
inference and we shall thus carry out the recon-
struction of both here and in [
1,2].
34
For example, in a multiple shot interrogation O could
first ask Q
1
, Q
2
on a set of identically prepared rebit cou-
ples to find out whether the answers are correlated. On a
second identically prepared set, O could then ask Q
2
, Q
1
which are maximally complementary to the first questions
he asked. From the statistics over the answers, O would be
able to determine also the probabilities for Q
12
and Q
21
.
But since he needed two separate interrogation runs to
determine the statistics for Q
12
and Q
21
, he would not be
able to infer from this any information whatsoever about
the statistics of answers to Q
33
.
Accepted in Quantum 2017-11-27, click title to verify 35
Local tomography is usually taken as the ori-
gin of the tensor product structure for composite
systems in quantum theory [16, 17,55,59]. How-
ever, one has to be careful with this statement
because there exist two distinct tensor products:
there is (a) the tensor product of Hilbert spaces,
e.g., (C
2
)
N
for qubits and (R
2
)
N
for rebits, and
(b) the tensor product of unnormalized probabil-
ity vectors or density matrices. A tensor prod-
uct of type (b) defines a sufficient support only
for composite qubit systems, but not for compos-
ite rebit systems; the space of hermitian matrices
over (C
2
)
N
is the N-fold tensor product of her-
mitian matrices over C
2
, but the space of sym-
metric matrices over (R
2
)
N
is not the N-fold
tensor product of symmetric matrices over R
2
(see
also [
83]). What local tomography implies is the
tensor product (b), but as the example of rebits
shows, the tensor product of type (a) also exists
without it.
5.2.7 A Bell scenario with questions: ruling out
local hidden variables
We shall now settle the issue of the relative nega-
tion ¬, i.e. whether
Q
11
Q
22
= Q
12
Q
21
, or
Q
11
Q
22
= ¬(Q
12
Q
21
),
for both rebits and q ubits; one of these equations
must be true (see theorems 5.7 and 5.9). It is
instructive to illustrate these equations by means
of a question configuration analogous to a Bell
scenario.
(A) Suppose Q
11
Q
22
= Q
12
Q
21
was
true. Before we determine whether it is consis-
tent with our background assumptions and pos-
tulates, we note that this is the case of classi-
cal (or realist) logic: O can consistently interpret
any such configuration by means of a local ‘hid-
den variable’ model. For example, consider the
case Q
11
= Q
22
= 1 (which is compatible with
the rules) such that Q
11
Q
22
= 1 and, conse-
quently, also Q
12
Q
21
= 1. We represent the
last two equations by the following two graphs
PSfrag replacements
Q
12
Q
21
Q
11
Q
22
PSfrag replacements
Q
12
Q
21
Q
11
Q
22
where both edges solid and of equal colour means
that the corresponding questions are correlated.
O could consistently read this configuration as
follows: “Since Q
1
= Q
1
and Q
2
= Q
2
, I
would precisely have to conclude that Q
12
=
Q
21
, i.e. Q
12
Q
21
= 1, if the individuals
Q
1
, Q
2
, Q
1
, Q
2
had definite values which, how-
ever, I do not know.” For instance, Q
11
= Q
22
=
Q
12
Q
21
= 1 would be consistent with the four
ontic states:
PSfrag replacements
1
1
1
1
0
PSfrag replacements
11
0 0
PSfrag replacements
1
0
00
0
PSfrag replacements
1 1
00
.
Therefore, O could join the two graphs consis-
tently (without knowing the truth values of the
individuals)
(23)
i.e., draw all four edges simultaneously which cor-
responds to all four edges having definite (al-
beit unknown) values at the same time. Sim-
ilarly, O can interpret any other configuration
with Q
11
Q
22
= Q
12
Q
21
in terms of a
local ‘hidden variable’ model which assigns defi-
nite values to the individuals and, conversely, as
can be easily checked, any assignment of definite
(or ontic) values to the individuals Q
1
, Q
2
, Q
1
, Q
2
leads necessarily to Q
11
Q
22
= Q
12
Q
21
.
We now preclude this possibility. We note that
Q
11
Q
22
= Q
12
Q
21
can be rewritten in
terms of the individuals as
(Q
1
Q
1
) (Q
2
Q
2
) = (Q
1
Q
2
) (Q
2
Q
1
) (24)
Accepted in Quantum 2017-11-27, click title to verify 36
which is a classical logical identity that is al-
ways true for classical Boolean logic. It s tates
that the brackets can be equivalently regroup ed,
i.e. that the order in which the XNOR
is executed in the logical ex pression can be
switched. In classical Boolean logic, this iden-
tity follows from the asso ciativity and symme-
try of the XNOR. Namely, suppose for a moment
that Q
1
, Q
2
, Q
1
, Q
2
had simultaneous truth val-
ues. Then Boolean logic would tell us that
(Q
1
Q
1
) (Q
2
Q
2
) =
(Q
1
Q
1
) Q
2
Q
2
=
Q
1
(Q
2
Q
1
)
Q
2
= Q
2
Q
1
(Q
2
Q
1
)
= (Q
1
Q
2
) (Q
2
Q
1
). (25)
Of course, relative to O, Q
1
, Q
2
, Q
1
, Q
2
do not
have simultaneous truth values. Instead, assump-
tion
6 implies that, for a classical logical iden-
tity to hold in O’s model, either (a) all ques-
tions in the involved logical expressions are mu-
tually maximally compatible or (b) the identity
follows from applying classical rules of inference
to (sub-)expressions which are entirely written in
terms of mutually maximally compatible ques-
tions and subsequently decomposing the result
logically into other questions (with possibly dis-
tinct compatibility relations). However, since
Q
1
, Q
2
and Q
1
, Q
2
form maximally complemen-
tary pairs, neither is the case here. In particu-
lar, the right hand side in (
24) does not follow
from applying any rules of Boolean logic to the
total expression Q
11
Q
22
or those subexpres-
sions Q
1
Q
1
and Q
2
Q
2
on the left hand
side which are fully composed of maximally com-
patible q uestions.
35
Hence, (24) violates assump-
tion
6 such that we must rule out the possibility
Q
11
Q
22
= Q
12
Q
21
.
(B) We are thus already forced to the conclu-
sion that
Q
11
Q
22
= ¬(Q
12
Q
21
). (26)
must hold. Indeed, its equivalent representation
(Q
1
Q
1
) (Q
2
Q
2
)
= ¬(Q
1
Q
2
) (Q
2
Q
1
)
does not constitute a classical logical identity
and is thus consistent with ass umption
6. It
35
The only relevant rules in this case, as in (
25), follow
from the properties of the XNOR, namely its symmetry
and associativity in Boolean logic. However, it is clear
that, th anks to the occurring complementarity, the argu-
ments of (
25) no longer apply. For instance, the first step
in (
25) is illegal, according to assumption 5, because Q
2
and Q
1
Q
1
are complementary such that they may not
be d irectly connected.
is also consistent with as sumption
5 since only
maximally compatible questions are directly con-
nected.
It is impossible for O to interpret this situation
in terms of local ‘hidden variables’ which assign
definite values to the individuals simultaneously.
Namely, consider, again, the case Q
11
= Q
22
=
1 such that Q
11
Q
22
= 1 and, consequently,
now Q
12
Q
21
= 0. This configuration may be
graphically represented as
PSfrag replacements
Q
12
Q
21
Q
11
Q
22
PSfrag replacements
Q
12
Q
21
Q
11
Q
22
where one edge solid the other dashed, but same
colour, means that the corresponding answers are
anti-correlated. One can easily check that, in con-
trast to (
23), it is impossible to consistently join
the two diagrams by drawing all four edges si-
multaneously (which would correspond to all in-
dividuals and thus all edges having definite truth
values), as one would obtain a frustrated graph
PSfrag replacements
<
PSfrag replacements
. (27)
O must conclude that neither the four individu-
als nor the four edges can have definite (but un-
known) values all at the same time. The same
verdict holds for any other configuration with
Q
11
Q
22
= ¬(Q
12
Q
21
).
Since either (A ) or (B) must be true by theo-
rems
5.7 and 5.9, and (A) violates assumption 6,
while (B) is consistent with both assumptions
5
and 6, we conclude that (26) is the correct rela-
tion.
Accepted in Quantum 2017-11-27, click title to verify 37
Note that this argument holds for both rebits
and qubits and, for qubits, also for any other pairs
Q
i
, Q
j6=i
and Q
k
, Q
l6=k
of pairs of maximally com-
plementary individuals (see theorem
5.7). In fact,
even more generally, the same argument can be
made for any f our q uestions (i.e., not necessar-
ily individuals) Q, Q
, Q
′′
, Q
′′′
which are such that
Q, Q
and Q
′′
, Q
′′′
are maximally complementary
pairs, while Q and Q
are each maximally com-
patible with both Q
′′
, Q
′′′
; also in this case one
would have to conclude that
(Q Q
′′
) (Q
Q
′′′
) = ¬
(Q Q
′′′
) (Q
Q
′′
)
. (28)
This will become relevant later on. For now we
observe that exchanging the positions of the com-
plementary Q
′′
, Q
′′′
from the left to the right hand
side introduces a negation ¬.
This relative negation ¬ in (
28) precludes a
classical reasoning for the distribution of truth
values over O’s questions. In fact, as just seen,
it rules out local ‘hidden variable models’, analo-
gously to the Bell arguments. We also recall that
assumption
6 was a statement about which rules
of inference and logical identities are not appli-
cable to logical comp ositions involving mutually
complementary questions. By contrast, (
28) is
now a first non-classical logical identity showing
one possible non-classical way of reasoning in the
presence of complementarity.
The correlation structure for rebits is now clear
from these results; (26) implies Q
33
= ¬
˜
Q
33
for
the correlations of correlations defined in (
22).
This settles the fate of all possible relative nega-
tions ¬ for rebits. However, these results do
not fully determine the odd and even correla-
tion structure for qubits: we still have to clarify
whether there is an overall negation ¬ relative to
Q
33
in (
19, 20) and more generally in theorem 5.7
for other permutations of non-intersecting edges.
This difference between rebits and qubits results
again from the fact that Q
33
is defined as the
correlation of the individuals Q
3
, Q
3
for qubits,
while it is a correlation of correlations for rebits.
This has a remarkable consequence: rebit theory
is its own ‘logical mirror image’, while qubit the-
ory’s ‘logical mirror image’ is distinct from qubit
theory. This topic will be deferred to section
5.4
because we firstly need to understand the ques-
tion structure for the N = 3 case in order to
discuss the odd and even correlation structure of
two qubits further.
5.3 Three gbits
It will be both useful and instructive to explicitly
consider the N = 3 case for rebits and qubits.
As a composite system, we can view three gbits,
labeled by A, B, C, either as three individual sys-
tems, as three combinations of one individual and
a bipartite composite system or as a tripartite
system:
PSfrag replacements
A
B
C
PSfrag replacements
A
B
C
PSfrag replacements
A
B
C
PSfrag replacements
A
B
C
PSfrag replacements
A
B
C
(29)
According to definition 3.5 of a composite system,
Q
3
must then contain the individual questions of
all three gbits, any bipartite correlation questions
and any permissible logical connectives thereof.
This results in different structures for rebits and
qubits.
Accepted in Quantum 2017-11-27, click title to verify 38
5.3.1 Three qubits
We shall begin with the qubit case D
1
= 3.
Clearly, according to definition
3.5, all 3 × 3 =
9 individual questions, henceforth denoted as
Q
i
A
, Q
j
B
, Q
k
C
, and all 3 × 9 = 27 bipartite
correlation questions, from now on written as
Q
i
A
j
B
, Q
i
A
,k
C
, Q
j
B
k
C
, i, j, k = 1, 2, 3, will be part
of an informationally complete set Q
M
3
. As be-
fore, we represent individuals and correlations
graphically as vertices and edges, respectively,
e.g.
PSfrag replacements
A B C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
A
1
C
Q
1
A
3
B
Q
1
B
3
C
Q
2
A
2
B
Q
2
B
2
C
Q
111
Q
333
Q
322
PSfrag replacements
A B C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
A
1
C
Q
1
A
3
B
Q
1
B
3
C
Q
2
A
2
B
Q
2
B
2
C
Q
111
Q
333
Q
322
depict valid question graphs.
But in order to complete the individuals and
bipartite correlations to Q
M
3
we now have to
consider logical connectives of these questions
which are pairwise independent. This will nec-
essarily involve tripartite questions because the
bipartite structure is already exhausted with in-
dividuals and bipartite correlations. Clearly, we
cannot add a question
˜
Q
111
, representing the
proposition “the answers to Q
1
A
, Q
1
B
, Q
1
C
are the
same" to the individuals and bipartite correla-
tions because, e.g.,
˜
Q
111
= 1 would alway s imply
Q
1
A
1
B
= Q
1
A
1
C
= Q
1
B
1
C
= 1. Also, Q
1
A
1
B
= 0
would imply
˜
Q
111
= 0 such that the bipartite
correlations and
˜
Q
111
would not be pairwise in-
dependent.
From (
13) we already know that the logi-
cal connective yielding new pairwise independent
questions must either be the X NOR or the XOR.
For consistency with the bipartite structure, we
continue to employ the XNOR . There is an obvi-
ous candidate for an independent tripartite ques-
tion, namely
Q
ijk
= Q
i
A
Q
j
B
Q
k
C
(30)
which thanks to the associativity and symmetry
of can also equivalently be written as
Q
ijk
= Q
i
A
Q
j
B
k
C
= Q
i
A
j
B
Q
k
C
= Q
i
A
k
C
Q
j
B
. (31)
(Since the notation for this tripartite ques-
tion is unambiguous from the ordering of
i, j, k we drop the subscripts A, B, C in Q
ijk
.)
This structure is also natural from the differ-
ent compositions in (
29). Q
ijk
thus defined
is by construction maximally compatible with
Q
i
A
, Q
j
B
, Q
k
C
, Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
and, for sim-
ilar reasons to the independence of Q
i
A
j
B
from
Q
i
A
, Q
j
B
(see the discussion below (14 )), also
pairwise independence of the latter. Note that
this question does not stand in one-to-one cor-
respondence with the proposition “the answers
to Q
i
A
, Q
j
B
, Q
k
C
are the same"; e.g., Q
i
A
= 1
and Q
j
B
= Q
k
C
= 0 also gives Q
ijk
= 1. It is
easier to interpret this question via (
31) as ei-
ther of the three questions “are the answers to
Q
i
A
, Q
j
B
k
C
/Q
i
A
j
B
, Q
k
C
/Q
i
A
k
C
, Q
j
B
the same?".
There are 3 × 3 × 3 = 27 such tripartite
questions Q
ijk
, i, j, k = 1, 2, 3. We shall repre-
sent them graphically as triangles. For example,
Q
111
, Q
322
, Q
333
are depicted as follows:
PSfrag replacements
A B C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
A
1
C
Q
1
A
3
B
Q
1
B
3
C
Q
2
A
2
B
Q
2
B
2
C
Q
111
Q
333
Q
322
Accepted in Quantum 2017-11-27, click title to verify 39
5.3.2 Independence and compatibility for three
qubits
We have to delve into some technical details on
the independence and compatibility structure to
explain and understand monogamy and the en-
tanglement structure of three qubits. These re-
sults will also be used to prove the analogous re-
sults by induction in section
5.5.1 for N qubits.
Individual questions from qubit A are maxi-
mally compatible with the individual questions
from qubits B and C, etc. But what about the
compatibility of bipartite and tripartite correla-
tions? The compatibility structure of bipartite
correlations of a fixed qubit pair is clear from
lemma
5.5, but we have to investigate compat-
ibility of bipartite correlation questions involving
all three qubits.
Lemma 5.10. Q
i
A
j
B
and Q
l
B
k
C
are maximally
complementary if j 6= l. On the other hand,
Q
i
A
j
B
and Q
j
B
k
C
are maximally compatible and
it holds
Q
i
A
j
B
Q
j
B
k
C
= Q
i
A
k
C
.
The analogous statements hold for any permuta-
tion of A, B, C. That is, graphically, two bipar-
tite correlations involving three qubits are max-
imally compatible if the corresponding edges in-
tersect in a vertex and maximally complementary
otherwise.
Proof. Maximal complementarity of Q
i
A
j
B
and
Q
l
B
k
C
for j 6= l is proven by noting that both
are maximally compatible with and independent
of Q
i
A
, the relation Q
j
B
= Q
i
A
Q
i
A
j
B
and
lemma
5.1.
Q
i
A
j
B
and Q
j
B
k
C
are evidently maximally com-
patible since Q
i
A
, Q
j
B
, Q
k
C
are maximally com-
patible. Moreover,
Q
i
A
j
B
Q
j
B
k
C
= Q
i
A
(Q
j
B
Q
j
B
)
|
{z }
=1
Q
k
C
= Q
i
A
k
C
thanks to the associativity of .
For example, Q
2
A
2
B
and Q
2
B
2
C
intersect in
Q
2
B
and are thus maximally compatible, while
Q
2
A
2
B
and Q
1
B
1
C
do not share a vertex and are
therefore maximally complementary:
PSfrag replacements
A
B
C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
B
1
C
Q
1
A
3
B
Q
1
B
3
C
Q
2
A
2
B
Q
2
B
2
C
Q
111
Q
333
Q
322
PSfrag replacements
A
B
C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
B
1
C
Q
1
A
3
B
Q
1
B
3
C
Q
2
A
2
B
Q
2
B
2
C
Q
111
Q
333
Q
322
We continue with the tripartite questions.
Lemma 5.11. Q
ijk
is maximally compatible with
Q
i
A
, Q
j
B
, Q
k
C
and maximally complementary to
Q
l
A
6=i
A
, Q
m
B
6=j
B
, Q
n
C
6=k
C
. That is, g raphically,
Q
ijk
is maximally compatible with an individ-
ual Q
i
A,B,C
if the corresponding vertex is one of
the ve rtices of the triangle representing Q
ijk
and
maximally complementary otherwise.
Proof. Q
ijk
is by construction maximally com-
patible with Q
i
A
, Q
j
B
, Q
k
C
. On the other hand,
complementarity of Q
ijk
and Q
l
A
6=i
A
is shown by
noting that both are maximally compatible with
and independent of Q
j
B
k
C
and lemma 5.1. One
argues analogously for the individuals of qubits
B and C.
For instance, Q
111
is maximally compatible
with Q
1
C
and maximally complementary to Q
2
C
:
PSfrag replacements
A
B
C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
B
1
C
Q
1
A
3
B
Q
1
B
3
C
Q
2
A
2
B
Q
2
B
2
C
Q
111
Q
333
Q
322
This lemma also directly implies that any in-
dividual and any tripartite correlation question
are pairwise independent because (1) maximally
complementary questions are in particular inde-
pendent, and (2) for maximally compatible ques-
tion pairs such as Q
i
A
, Q
ijk
independence argu-
ments analogous to those surrounding (
14) apply.
Accepted in Quantum 2017-11-27, click title to verify 40
Next we consider bipartite and tripartite cor-
relation questions.
Lemma 5.12. Q
ijk
is maximally compatible
with Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
and, furthermore, with
Q
m
B
n
C
, Q
l
A
n
C
and Q
l
A
m
B
for l 6= i, m 6= j
and k 6= n. On the other hand, Q
ijk
is max-
imally complementary to Q
i
A
m
B
, Q
i
A
n
C
, Q
l
A
j
B
,
Q
j
B
n
C
, Q
l
A
k
C
, Q
m
B
k
C
for l 6= i, m 6= j and
k 6= n. That is, graphically, Q
ijk
is maximally
compatible with a b i partite correlation if the edge
of the latter is either an edge of the triangle cor-
responding to Q
ijk
or if the edge and triangle do
not intersect. Q
ijk
is maximally complementary
to a bipartite correlation question if the edge of
the latter and the triangle corresponding to Q
ijk
share one common vertex.
Proof. Q
ijk
is by construction maximally com-
patible with Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
. Q
ijk
= Q
i
A
Q
j
B
k
C
and Q
m
B
n
C
are also maximally compat-
ible for j 6= m and k 6= n because Q
m
B
n
C
is
maximally compatible with Q
i
A
and thanks to
lemma
5.5 also with Q
j
B
k
C
. Complementarity of
Q
ijk
and Q
i
A
m
B
for j 6= m follows from noting
that both are maximally compatible with and in-
dependent of Q
k
C
and lemma
5.1. The reasoning
for all other cases is analogous.
To give a graphical example, Q
111
is maximally
compatible with Q
1
B
1
C
and Q
3
A
3
C
and maxi-
mally complementary to Q
1
A
2
B
:
PSfrag replacements
A
B
C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
B
1
C
Q
1
A
3
B
Q
1
A
2
B
Q
3
A
3
C
Q
2
B
2
C
Q
111
Q
333
Q
322
PSfrag replacements
A
B
C
Q
1
A
Q
2
A
Q
3
A
Q
1
B
Q
2
B
Q
3
B
Q
1
C
Q
2
C
Q
3
C
Q
1
B
1
C
Q
1
A
3
B
Q
1
A
2
B
Q
3
A
3
C
Q
2
B
2
C
Q
111
Q
333
Q
322
We still have to check pairwise independence of
the bipartite and tripartite correlation questions.
Lemma 5.13. Any bipartite
Q
m
B
n
C
, Q
l
A
n
C
, Q
l
A
m
B
and any tripartite corre-
lation question Q
ijk
are independent from one
another.
Proof. Lemma
5.12 implies that we only
have to check pairwise independence of
Q
m
B
n
C
, Q
l
A
n
C
, Q
l
A
m
B
from Q
ijk
for l 6= i,
m 6= j and k 6= n because maximally comple-
mentary questions are by definition independent
and Q
ijk
and Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
are pairwise
independent. Consider therefore Q
ijk
and
Q
m
B
n
C
for j 6= m and k 6= n. By lemma
5.11,
Q
k
C
is maximally compatible with Q
ijk
and by
lemma 5.2 m aximally complementary to Q
m
B
n
C
.
This implies, using the arguments from the proof
of lemma 5.3, independence of Q
ijk
, Q
m
B
n
C
.
The other cases follow similarly.
Lemma 5.14. The tripartite correlation ques-
tions Q
ijk
, i, j, k = 1, 2, 3 are pairwise indepen-
dent.
Proof. Consider Q
ijk
and Q
lmn
for i 6= l. By
lemma
5.11, Q
i
A
is maximally compatible with
Q
ijk
and maximally complementary to Q
lmn
. Us-
ing the analogous arguments from the proof of
lemma
5.3, this implies that Q
ijk
, Q
lmn
are inde-
pendent. The same reasoning holds when j 6= m
and k 6= n.
This has an immediate consequence:
Corollary 5.15. The individuals Q
i
A
, Q
j
B
, Q
k
C
,
the bipartite Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
and the tripar-
tite Q
ijk
, i, j, k = 1, 2, 3 are pairwise independent
and thus, thanks to assumption
7, contained in
an informationally complete set Q
M
3
.
Lastly, we consider the complementarity and
compatibility structure of the tripartite correla-
tions.
Lemma 5.16. Q
ijk
and Q
lmn
are maximally
compatible if {i, j, k} and {l, m, n} overlap in one
or three indices and maximally complementary if
{i, j, k} and {l, m, n} overlap in zero or two in-
dices. That is, graphically, Q
ijk
and Q
lmn
are
maximally compatible if their corresponding tri-
angles intersect in one vertex (or coincide) and
maximally complementary i f the triangles share
an edge or do not intersect.
Proof. Compatibility for an overlap in all three
indices is trivial. But also Q
ijk
= Q
i
A
Q
j
B
k
C
and Q
imn
= Q
i
A
Q
m
B
n
C
are clearly maxi-
mally compatible for j 6= m and k 6= n because
by lemma
5.5 Q
j
B
k
C
, Q
m
B
n
C
are maximally com-
patible in this case. Compatibility for th e other
cases of an overlap of Q
ijk
and Q
lmn
in one index
follows by permutation.
Accepted in Quantum 2017-11-27, click title to verify 41
The proof of the complementarity of Q
ijk
and
Q
lmn
for i 6= l, j 6= m and k 6= n follows from
lemma 5.1. One m ay use the fact that, by lemma
5.12, both questions are maximally compatible
with and independent of Q
j
B
k
C
and that, by
lemma
5.11, Q
i
A
6=l
A
is maximally complementary
to Q
lmn
.
Similarly, one proves complementarity of
Q
ijk
= Q
i
A
Q
j
B
k
C
and Q
ljk
= Q
l
A
Q
j
B
k
C
for i 6= l by using that both are maximally com-
patible with and independent of Q
j
B
k
C
. Com-
plementarity of tripartite correlations for other
overlaps in precisely two indices follows by per-
mutation.
For example, Q
111
and Q
212
intersect in the
vertex Q
1
B
and are thus maximally compatible.
By contrast, Q
111
shares the edge Q
1
A
1
B
with
Q
113
and does not intersect at all with Q
333
such
that Q
111
is maximally complementary to both.
Q
113
and Q
333
intersect in the vertex Q
3
C
and are
therefore maximally compatible:
PSfrag replacements
Q
111
Q
333
Q
113
Q
212
PSfrag replacements
Q
111
Q
333
Q
113
Q
212
5.3.3 An informationally complete set for three
qubits
We now show that the 9 individuals, the 27 bi-
partite and the 27 tripartite correlation questions
form an informationally complete set Q
M
3
for
N = 3 qubits.
Theorem 5.17. (Qubits) The individuals
Q
i
A
, Q
j
B
, Q
k
C
, the bipartite Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
and the tripartite Q
ijk
, i, j, k = 1, 2, 3 are logi-
cally closed under such that they form an in-
formationally complete set Q
M
3
with D
3
= 63 for
D
1
= 3.
Proof. For individuals and bipartite correlations
of any pair of qubits the logical closure under the
XNOR was already shown in section
5.2. Cor-
relation of individuals with a bipartite correla-
tion question from another pair of qu bits yields
the tr ipartite correlation questions. Lemma
5.10
shows that also the b ipartite correlation ques-
tions involving distinct pairs of qubits are logi-
cally closed und er . We thus only have to check
logical closure of XNOR combinations involving
tripartite correlation questions.
Lemma
5.11 shows that tripartite correlations
and individuals are only maximally compatible if
the individual is a vertex of the tripartite trian-
gle. But then a combination such as
Q
ijk
Q
i
A
= (Q
i
A
Q
i
A
) Q
j
B
k
C
= Q
j
B
k
C
produces another bipartite correlation, thanks to
the associativity of . Similarly, lemma
5.12 as-
serts that tripartite and bipartite correlations are
only maximally compatible if the edge of the b i-
partite correlation is either contained in the tri-
partite triangle or if the edge and triangle do n ot
intersect. However, combinations of such maxi-
mally compatible pairs also do not yield any new
questions because, e.g.,
Q
ijk
Q
i
A
j
B
= Q
k
C
and, us ing again the associativity of XNOR and
theorem
5.7,
Q
σ
A
(1)σ
B
(1)k
Q
σ
A
(2)σ
B
(2)
= (Q
σ
A
(1)σ
B
(1)
Q
σ
A
(2)σ
B
(2)
)
| {z }
= Q
σ
A
(3)σ
B
(3)
or ¬Q
σ
A
(3)σ
B
(3)
Q
k
C
= Q
σ
A
(3)σ
B
(3)k
or ¬Q
σ
A
(3)σ
B
(3)k
,
where σ
A
, σ
B
are permutations of {1, 2, 3}. Fi-
nally, lemma 5.16 entails that tripartite correla-
tions are only maximally compatible if they inter-
sect in one vertex. But, using theorem
5.7 once
more,
Accepted in Quantum 2017-11-27, click title to verify 42
Q
σ
A
(1)σ
B
(1)k
Q
σ
A
(2)σ
B
(2)k
= Q
σ
A
(1)σ
B
(1)
Q
σ
A
(2)σ
B
(2)
= Q
σ
A
(3)σ
B
(3)
or ¬Q
σ
A
(3)σ
B
(3)
.
Permu ting these examples implies that all in-
dividuals, bipartite and tripartite correlations
are logically closed under which by (
13) is
the only independent logical connective possibly
yielding new independent questions. These ques-
tions therefore form an informationally complete
set Q
M
3
with D
3
= 63.
5.3.4 Entanglement of t hree qubits and
monogamy
Let us now put all these detailed results to good
use and reward ourselves with the observation
that they naturally explain monogamy of entan-
glement for the q ubit case. This is b est illus-
trated with an example. Let O have asked the
questions Q
1
A
1
B
and Q
2
A
2
B
to qubits A, B such
that the two are in a state of maximal informa-
tion of N = 2 independent bits and maximal
entanglement relative to O.
PSfrag replacements
Q
1
A
1
B
Q
2
A
2
B
Intuitively, this already suggests monogamy of
entanglement because O has spent the maximally
attainable amount of information over the pair
A, B s uch that even if now a third qubit C enters
the game he should no longer be able to ask any
question to the triple which gives him any fur-
ther simultaneous independent information about
the pair A, B. But any correlation of A or B
with C would constitute additional information
about A or B which would violate the informa-
tion bound for the subsystem of the composite
system of three qubits.
We shall now make this more precise. The
question is, which bipartite or tripartite correla-
tions involving qubit C are maximally compatible
with Q
1
A
1
B
, Q
2
A
2
B
. Firstly, lemma
5.10 states
that bipartite correlations of two distinct qubit
pairs are only maximally compatible if their cor-
responding edges intersect in a vertex. Clearly,
there can then not exist any edge which is max-
imally compatible with both Q
1
A
1
B
, Q
2
A
2
B
. Sec-
ondly, lemma
5.12 asserts that a bipartite correla-
tion is only maximally compatible with a tripar-
tite correlation if the corresponding edge is either
contained in the tripartite triangle or does not
intersect the triangle. This means that the only
tripartite correlations maximally compatible with
Q
1
A
1
B
, Q
2
A
2
B
are (a) the correlations of the lat-
ter with the individuals Q
1
C
, Q
2
C
, Q
3
C
of qubit
C and (b) Q
33k
, k = 1, 2, 3.
For example, in the first of the following two
question graphs Q
2
B
2
C
is maximally compati-
ble with Q
2
A
2
B
but maximally complementary to
Q
1
A
1
B
, while Q
3
B
3
C
is maximally complementary
to both. But O could ask Q
222
together with
Q
1
A
1
B
, Q
2
A
2
B
, as depicted in the second graph:
PSfrag replacements
Q
1
A
1
B
Q
2
A
2
B
Q
222
Q
2
C
Q
3
B
3
C
Q
2
B
2
C
PSfrag replacements
Q
1
A
1
B
Q
2
A
2
B
Q
222
Q
2
C
Q
3
B
3
C
Q
2
B
2
C
However, asking Q
1
A
1
B
, Q
2
A
2
B
and Q
222
simul-
taneously is equivalent to asking Q
1
A
1
B
, Q
2
A
2
B
and Q
2
C
because Q
2
A
2
B
Q
222
= Q
2
C
.
The analogous conclusion holds for any other
tripartite question maximally compatible with
Q
1
A
1
B
, Q
2
A
2
B
. Therefore, once O has asked
the last to bipartite correlations about qubits
A, B, he can only acquire individual informa-
tion corresponding to Q
1
C
, Q
2
C
, Q
3
C
about qubit
C. Clearly, the same state of affairs is true for
any permutation of the three qubits. This is
monogamy of entanglement in its most extreme
form: two qubits which are maximally entangled
can not be correlated whatsoever with any other
system.
Accepted in Quantum 2017-11-27, click title to verify 43
The non-extremal form of monogamy, namely
the case that a qubit pair is not maximally en-
tangled and can thus share a bit of entanglement
with a third q ubit, can also be explained. To this
end, we recall from quantum information theory
that monogamy is generally described with so-
called monogamy inequalities, e.g., the Coffman-
Kundu-Wootters inequality [
84]
τ
A|BC
τ
AB
+ τ
AC
, (32)
where 0 τ
A|BC
= 2(1 Tr ρ
2
A
) 1 measures
the entanglement between qubit A and the qubit
pair B, C and ρ
A
is the marginal s tate of q ubit A
obtained after tracing qubit B and C out of the
(not necessarily pure) tripartite state. τ
AB
, τ
AC
,
on the other hand, measure the bipartite entan-
glement of the pairs A, B and A, C and are s imi-
larly obtained by tracing either C or B out of the
tripartite qubit state. The inequality (32) formal-
izes the intuition that the correlation between A
and the pair B, C is at least as strong as the cor-
relation of A with B and C individually. This
is the general form of monogamy. One usually
also defines the s o-called three-tangle [
84] as the
difference between left and right hand side
τ
ABC
= τ
A|BC
τ
AB
τ
AC
[0, 1]
to measure the genuine tripartite entanglement
shared among all three qubits. This three-tangle
turns out to be permutation invariant τ
ABC
=
τ
BCA
= τ
CAB
.
In our current case, a measure of entanglement
sharing must be informational. At this stage,
in line with our previous definition of entangle-
ment, we simply define the analogous entangle-
ment measures to be the sum of O’s information
about the various independent bipartite or tripar-
tite correlation questions
˜τ
A|BC
:=
3
X
i,j,k=1
α
ijk
+
3
X
i,j=1
α
i
A
j
B
+
3
X
i,k=1
α
i
A
k
C
˜τ
AB
:=
3
X
i,j=1
α
i
A
j
B
, ˜τ
AC
:=
3
X
i,k=1
α
i
A
k
C
. (33)
With this definition, an informational inequality
analogous to the inequality (
32) is trivial
˜τ
A|BC
˜τ
AB
+ ˜τ
AC
as is the fact that the informational three-tangle
˜τ
ABC
:= ˜τ
A|BC
˜τ
AB
˜τ
AC
=
3
X
i,j,k=1
α
ijk
is permutation invariant and measures only the
information contained in the tripartite questions
and thus genuine tripartite entanglement. This
is how one can describe the general form of
monogamy of entanglement in our language. Of
course, at this stage the information measure α
i
about the various questions Q
i
is only implicit,
but we shall derive its explicit form in section
6.8, upon which the above statements become
truly quantitative. These informational versions
of monogamy inequalities and tangles naturally
suggest themselves for simple generalizations to
arbitrarily many qubits, thereby complementing
current efforts in the quantum information liter-
ature (e.g., see [
85]).
Before we move on, we briefly emphasize that
monogamy is a consequence of complementarity
as is entanglement. For example, three classi-
cal bits also s atisfy the limited information rule
1, but due to the absence of complementarity,
O could ask all correlations Q
AB
, Q
AC
, Q
BC
and
Q
ABC
at once
PSfrag replacements
A B
C
which, however, is eq uivalent to asking the three
individuals Q
A
, Q
B
, Q
C
.
5.3.5 Maximal entanglement for three qubits
Let us elucidate maximal tripartite entanglement
for three qubits, corresponding to O asking three
mutually maximally compatible and independent
tripartite correlation questions such that he ex-
hausts the information bound of N = 3 indepen-
dent bits with tripartite information. Lemma
5.16 asserts that tripartite questions are maxi-
mally compatible if and only if they intersect in
exactly one vertex. For example, O could ask
Q
211
, Q
121
, Q
112
simultaneously. These are mu-
tually independent because by (19, 20, 26)
Q
211
Q
121
= Q
3
A
3
B
or ¬Q
3
A
3
B
Q
211
Q
112
= Q
3
A
3
C
or ¬Q
3
A
3
C
Q
121
Q
112
= Q
3
B
3
C
or ¬Q
3
B
3
C
, (34)
i.e., their binary connectives with an XNOR do
not imply each other and the bipartite correla-
tions are pairwise independent from the tripar-
tite ones. Accordingly, asking Q
211
, Q
121
, Q
112
Accepted in Quantum 2017-11-27, click title to verify 44
will provide O with three independent bits of
information about the qubit triple. We note that
lemma 5.11 implies that once O has posed the
questions Q
211
, Q
121
, Q
112
, every individual q ues-
tion Q
i
A
, Q
j
B
, Q
k
C
is maximally complementary
to at least one of the three tripartite correlations.
That is, O cannot acquire any individual inf or-
mation about the three qubits and will only have
comp os ite information. This is a necessary con-
dition for maximal entanglement.
In addition to the three implied binary corre-
lations (
34), O would clearly also know the tri-
partite correlation of all three Q
211
, Q
121
, Q
112
,
which amounts to
Q
211
Q
121
Q
112
= Q
222
or ¬Q
222
.
This exhausts the list of questions about
which O would have information by asking
Q
211
, Q
121
, Q
112
. The lis t can be represented by
the f ollowing question graph:
(35)
This graph will ultimately correspond to
the eight possible Greenberger-Horne-Zeilinger
(GHZ) states in either of the three question
bases {Q
2
A
, Q
1
B
, Q
1
C
}, {Q
1
A
, Q
2
B
, Q
1
C
} and
{Q
1
A
, Q
1
A
, Q
2
C
}, corresponding to the eight
answer configurations ‘yes-yes-yes’, ‘yes-yes-no’,
‘yes-no-no’, ‘yes-no-yes’, ‘no-yes-yes’, ‘no-yes-no’,
‘no-no-yes’ and ‘no-no-no’ to the three tripar-
tite correlations Q
211
, Q
121
, Q
112
. Similarly, any
other three maximally compatible tripartite cor-
relations will correspond to GHZ states in other
question bases.
36
The correlation information in the graph (35) is
clearly democratically distributed over the three
qubits A, B, C. (Intuitively, one could even
36
As an aside, we note that these propositions solve the
little riddle in Zeilinger’s festschrift for D. Greenberger
[
31].
view the bit stemming from, say, the answer to
Q
2
A
1
B
1
C
as being carried to one-third by each
of A, B, C.) Furthermore, if O now wanted to
‘marginalize over qubit C, i.e. discard any infor-
mation involving qubit C, all that would be left of
his knowledge about the qubit triple would be the
answer to Q
3
A
3
B
. But, as argued in section 5.2.6,
this resulting ‘marginal s tate’ of qubits A, B rel-
ative to O cannot be considered entangled. The
analogous results hold for ‘marginalization’ over
either A or B. As a result, the question graph
(
35) corresponds to genuine maximal tripartite
entanglement.
5.3.6 Three rebits
We shall now repeat the same exercise for three
rebits but can benefit from the results of the N =
3 qubit case. We shall thus be briefer here. The
reader exclusively interested in qubits may jump
to section
5.4.
According to definition
3.5, O’s questions to
the three rebit system must contain the 6 indi-
viduals Q
i
A
, Q
j
B
, Q
k
C
, 12 bipartite correlations
Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
, i, j, k = 1, 2, and 3 bipartite
correlations of correlations Q
3
A
3
B
, Q
3
A
3
C
, Q
3
B
3
C
.
To render this set informationally complete, we
have to top it up with tripartite questions. There
are now 8 tripartite correlations Q
ijk
, i, j, k =
1, 2, of the kind (
30, 31) and, moreover, 6 tri-
partite correlations of a rebit individual question
with the question Q
33
asking for the correlation
of bipartite correlations of the other rebit pair
Q
i33
:= Q
i
A
Q
3
B
3
C
,
Q
3j3
:= Q
j
B
Q
3
A
3
C
,
Q
33k
:= Q
3
A
3
B
Q
k
C
, i, j, k = 1, 2.
Given that the individuals Q
3
A
, Q
3
B
, Q
3
C
do not
exist for rebits there is now also no tripartite cor-
relation Q
333
.
5.3.7 Independence and compatibility for three
rebits
The independence, compatibility and comple-
mentarity structure for the questions not involv-
ing an index i, j, k = 3 directly follows from the
qubit discussion as lemmas
5.105.16 also hold in
the present case for i, j, k = 1, 2. But we now
have to clarify the question structure once two
indices are equal to 3 (an odd number of indices
Accepted in Quantum 2017-11-27, click title to verify 45
cannot be 3). The status of any such purely bi-
partite relations was clarified in section
5.2 such
that here we have to consider the case that all
three rebits are involved.
Lemma 5.18. Q
3
B
3
C
is maximally complemen-
tary to Q
i
A
j
B
, i, j = 1, 2, and maximally compat-
ible with Q
3
A
3
B
. Furthermore,
Q
3
A
3
B
Q
3
B
3
C
= Q
3
A
3
C
. (36)
The same holds for any permutations of A, B, C.
Proof. One proves complementarity of Q
i
A
j
B
and
Q
3
B
3
C
= Q
1
B
2
C
Q
2
B
1
C
by employin g that
both are maximally compatible with and inde-
pendent of Q
i
A
and lemma
5.1. T his also requires
invoking that Q
j
B
and Q
3
B
3
C
are maximally com-
plementary by lemma
5.8.
Compatibility of Q
3
A
3
B
and Q
3
B
3
C
fol-
lows ind ir ectly. Namely, Q
1
A
2
B
, Q
2
B
1
C
and
Q
2
A
1
B
, Q
1
B
2
C
are two maximally compatible
pairs by lemma 5.10. We can then apply the
reasoning around (
28) to find
Q
3
A
3
B
Q
3
B
3
C
= (Q
1
A
2
B
Q
2
A
1
B
) (Q
1
B
2
C
Q
2
B
1
C
)
= ¬((Q
1
A
2
B
Q
2
B
1
C
)
|
{z }
= Q
1
A
1
C
(Q
1
B
2
C
Q
2
A
1
B
)
|
{z }
= Q
2
A
2
C
)
= ¬(Q
1
A
1
C
Q
2
A
2
C
) = Q
3
A
3
C
.
The last equality holds thanks to theorem 5.9.
The same reasoning applies to any permuta-
tion of A, B, C.
Next, we discuss the tripartite correlations of
individuals with the questions Q
33
asking for the
correlation of bipartite correlations of rebit pairs.
Lemma 5.19. Q
i33
is maximally compatible with
Q
i
A
and maximally complementary to Q
j
B
, Q
k
C
.
The same holds for any permutation of A, B, C.
Proof. Compatibility of Q
i33
with Q
i
A
is true
by construction. Complementarity of Q
j
B
and
Q
i33
= Q
i
A
(Q
1
B
2
C
Q
2
B
1
C
) follows from
the observation that Q
i33
and Q
j
B
are both max-
imally compatible with and independent of Q
i
A
and a similar reasoning to the previous proof.
Lemma 5.20. Q
i33
is maximally compatible with
Q
3
B
3
C
, Q
j
B
k
C
, j, k = 1, 2, and Q
l
A
k
C
, Q
l
A
j
B
for
i 6= l. On the other hand, Q
i33
is maximally com-
plementary to Q
i
A
j
B
, Q
i
A
k
C
and Q
3
A
3
B
, Q
3
A
3
C
.
Furthermore, Q
ijk
is maximally compatible with
Q
3
A
3
B
. The same holds for any permutation of
A, B, C.
Proof. Compatibility of Q
i33
with Q
j
B
k
C
and
Q
3
B
3
C
, as well as compatibility of Q
ijk
with
Q
3
A
3
B
is obvious. Compatibility of Q
i33
with
Q
l
A
k
C
for i 6= l follows indirectly by noting that
Q
i
A
, Q
l
A
and Q
3
B
3
C
, Q
k
C
are two pairs of max-
imally complementary questions, but that the
questions in one pair are maximally compatible
with both questions of the other. In this case
the reasoning of (
28) applies and entails the cor-
relation of Q
i
A
with Q
3
B
3
C
must be maximally
compatible with the correlation of Q
l
A
with Q
k
C
.
Compatibility of Q
i33
with Q
l
A
j
B
follows simi-
larly.
Complementarity of Q
i33
and Q
i
A
j
B
follows
from the fact that both are maximally compati-
ble with and independent of Q
i
A
and using argu-
ments as in previous proofs. Likewise, Q
i33
and
Q
3
A
3
B
follows similarly by noting that both are
maximally compatible with Q
3
B
3
C
.
Lemma 5.21. Any of Q
i33
, Q
3j3
, Q
33k
is
independent from any bipartite correla-
tion question Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
and
any bipartite correlation of correlations
questionQ
3
A
3
B
, Q
3
A
3
C
, Q
3
B
3
C
. Q
ijk
is also
pairwise independent from the latter. Further-
more, the Q
ijk
and Q
i33
, Q
3j3
, Q
33k
are pairwise
independent.
Proof. The proof is completely analogous to the
proofs of lemmas
5.3, 5.13 and 5.14.
As before this has an important consequence.
Corollary 5.22. The individuals Q
i
A
, Q
j
B
, Q
k
C
,
the bipartite correlations Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
,
the bipartite correlations of correlations
Q
3
A
3
B
, Q
3
A
3
C
, Q
3
B
3
C
, the tripartite correla-
tions Q
ijk
and the tripartite Q
i33
, Q
3j3
, Q
33k
,
Accepted in Quantum 2017-11-27, click title to verify 46
i, j, k = 1, 2, are pairwise independent and
thus, thanks to assumption
7, part of an
informationally complete set Q
M
3
.
The compatibility and complementarity struc-
ture of questions involving the ‘correlation of cor-
relations’ Q
33
is analogous to lemma
5.16.
Lemma 5.23. Q
ijk
is maximally compatible with
Q
i33
, Q
3j3
, Q
33k
and maximally complementary
to Q
l33
, Q
3m3
, Q
33n
for i 6= l, j 6= m and k 6= n.
Furthermore, Q
i33
is maximally compatible with
Q
3j3
, Q
33k
, but Q
133
and Q
233
are maximally
complementary. The analogous result holds for
all permutations of A, B, C.
Proof. Compatibility of Q
ijk
with
Q
i33
, Q
3j3
, Q
33k
follows from the fact that
the constituents of the latter Q
i
A
, Q
3
B
3
C
, . . .
are maximally compatible with Q
ijk
. Comple-
mentarity of, e.g., Q
ijk
and Q
l33
for i 6= l can
be shown by noting that both are maximally
compatible with and independent of Q
3
B
3
C
and
lemma
5.1. Compatibility of, say, Q
i33
and
Q
3j3
can be demonstrated by using (
28) and
noting that Q
i
A
, Q
3
A
3
C
and Q
j
B
, Q
3
B
3
C
are two
pairs of maximally complementary questions
which are such that each question in one pair
is maximally compatible with both questions
of the other pair. Finally, Q
133
and Q
233
are
maximally complementary because both are
maximally compatible with Q
3
B
3
C
and Q
1
A
, Q
2
A
are m aximally complementary.
This finishes our considerations of the inde-
pendence and complementarity structure of three
rebits.
5.3.8 An informationally complete set for three
rebits
The rebit questions considered thus far comprise
an informationally complete set of 35 elements.
Theorem 5.24. (Rebits) The individu-
als Q
i
A
, Q
j
B
, Q
k
C
, the bipartite correlations
Q
i
A
j
B
, Q
i
A
k
C
, Q
j
B
k
C
, the bipartite correlations
of correlations Q
3
A
3
B
, Q
3
A
3
C
, Q
3
B
3
C
, the tri-
partite correlations Q
ijk
and the tripartite
Q
i33
, Q
3j3
, Q
33k
, i, j, k = 1, 2, are logically closed
under and thus form an informationally
complete set Q
M
3
with D
3
= 35 for D
1
= 2.
Proof. Logical closure under the XNOR for any
pair of rebits follows f rom section
5.2. Combin-
ing individuals with bipartite questions of an-
other rebit pair prod uces the tripartite qu estions.
Lemma
5.10 and 5.18 assert logical closure of
bipartite correlation questions an d the bipartite
questions Q
33
asking for the correlation of bipar-
tite correlations among all thr ee rebits. We must
therefore only check logical closure of combina-
tions involving tripartite questions which involve
indices taking the value 3 (the other cases are
covered by theorem
5.17 for i, j, k = 1, 2).
Using theorem
5.9 and lemmas 5.185.23, the
proof is mostly analogous to the proof of theorem
5.17 such that here we will only show two non-
trivial cases which are treated differently. For
example, by lemma 5.20 Q
133
and Q
2
A
2
B
are
maximally compatible. The conjunction with
yields
Q
133
Q
2
A
2
B
=
(
22,26)
(Q
1
A
Q
3
B
3
C
) ¬(Q
1
A
1
B
Q
3
A
3
B
)
=
(
28,36)
Q
1
B
Q
3
A
3
C
= Q
313
.
Similarly, by lemma 5.23, Q
i33
, Q
3j3
are max- imally compatible. Their XNOR conjunction
gives
Q
i33
Q
3j3
= (Q
i
A
Q
3
B
3
C
) (Q
3
A
3
C
Q
j
B
) =
(
28)
¬(Q
3
A
3
B
Q
i
A
j
B
)
which again coincides with some Q
l
A
m
B
with i 6= l and j 6= m (or the negation thereof). In
Accepted in Quantum 2017-11-27, click title to verify 47
complete analogy to these explicit examples and
to the proof of theorem
5.17, one shows the log-
ical closure under the XNOR of all other cases.
This gives the desired result.
5.3.9 Monogamy and maximal entanglement for
three rebits
Rebits are considered as non-monogamous in the
literature [86, 87]. However, this conclusion de-
pends somewhat on one’s notion of monogamy
and can be clarified within our language. Firstly,
rebits, just as qubits, are clearly monogamous in
the following sense: if two rebits A, B are maxi-
mally entangled in a state of maximal information
relative to O, then they cannot share any entan-
glement whatsoever with a third rebit C. This
can be seen by repeating the argument of section
5.3.4 which holds analogously for three rebits.
The situation changes slightly for entangled
states of non-maximal information involving ei-
ther of Q
3
A
3
B
, Q
3
A
3
C
, Q
3
B
3
C
and for tripartite
maximally entangled states. As an example for
the former case, O could ask Q
3
A
3
B
, Q
3
B
3
C
simul-
taneously which gives him two independent bits
of information about the rebit triple and implies
Q
3
A
3
C
by (
36) as well such that his information
could be s ummarized as
PSfrag replacements
Q
3
A
3
B
Q
3
A
3
C
Q
3
B
3
C
(we depict the Q
3
A
3
B
, Q
3
A
3
C
, Q
3
B
3
C
without ver-
tices to emphasize the absence of the individu-
als Q
3
A
, Q
3
B
, Q
3
C
for rebits). This graph cor-
responds to non-monogamously entangled rebit
states: by lemma
5.8 all individuals are maxi-
mally complementary to two of the three known
questions such that O cannot acquire any indi-
vidual information about the three rebits at the
same time. The three rebits are pairwise maxi-
mally entangled, albeit not in a state of maximal
information (see also section
5.2.6).
However, the three rebits can be similarly
non-monogamous in a tripartite maximally en-
tangled state of maximal information. As in
the qubit case (
35) in section 5.3.5, one can
write down a question graph corresponding to the
rebit analogues of GHZ states, representing the
eight possible answers to the tripartite q uestions
Q
211
, Q
121
, Q
112
PSfrag replacements
Q
3
A
3
B
Q
3
A
3
C
Q
3
B
3
C
(to justify this graph one has to employ the-
orem
5.24). Thanks to the information about
the answers to Q
3
A
3
B
, Q
3
A
3
C
, Q
3
B
3
C
, such a state
would contain a non-monogamous distribution of
entanglement over A, B, C. In particular, if O
‘marginalized’ over rebit C, by discarding all in-
formation involving C, he would be left with the
answer to Q
3
A
3
B
which defines a maximally en-
tangled two-rebit state of non-maximal informa-
tion (see section 5.2.6) in contrast to the qubit
case of section
5.3.5.
5.4 Correlation structure for N = 2 gbits
While we were able to settle the relative negation
¬ between the correlations of bipartite correla-
tion questions for both qubits and rebits in (
28)
(e.g., Q
11
Q
22
= ¬(Q
12
Q
21
)), we s till have
to clarify the odd and even correlation structure
for q ubits in (
19, 20) and more generally in theo-
rem
5.7. For example, we have to clarify whether
Q
33
= Q
11
Q
22
or Q
33
= ¬(Q
11
Q
22
), etc.
This will involve the notion of the ‘logical mir-
ror’ of an inference theory and require the tools
from the previous section, covering the N = 3
case. We recall that the odd and even correla-
tion structure for rebits has already been settled
in section
5.2.7 through (28). This, as we shall
see shortly, is a consequence of rebit theory being
its own mirror image.
Accepted in Quantum 2017-11-27, click title to verify 48
5.4.1 The logical mirror image of an inference the-
ory
Consider a single qubit, described by O via an in-
formationally complete set Q
1
, Q
2
, Q
3
which can
be viewed as a question basis on Q
1
. As such, it
defines a ‘logical handedness’ in terms of which
outcomes to Q
1
, Q
2
, Q
3
O calls ‘yes’ and which
‘no’.
37
Clearly, this is a convention made by
O and he can easily change the handedness by
simply swapping the assignment ‘yes’’no’ of
one question, say, Q
1
, which is tantamount to
Q
1
7→ ¬Q
1
. This corresponds to taking the log-
ical mirror image of a single qubit. For a single
qubit this will not have any severe consequences
and O is free to choose whichever handedness he
desires.
However, the handedness of a single qubit ques-
tion basis does become important when consider-
ing composite systems. Consider, e.g., two ob-
servers O, O
each having one qubit and describ-
ing it with a certain handed basis. They can also
consider correlations Q
ij
of their qubit pair. If O
now decided to change the handedness of his local
question basis by Q
1
7→ ¬Q
1
this would result in
Q
11
, Q
12
, Q
13
7→ ¬Q
11
, ¬Q
12
, ¬Q
13
,
Q
ij
7→ Q
ij
, i 6= 1
and therefore
Q
11
Q
22
7→ ¬(Q
11
Q
22
), and
Q
12
Q
21
7→ ¬(Q
12
Q
21
).
Since Q
33
would remain unaffected by the change
of local handedness (Q
3
, Q
3
remain invariant),
O’s change of local handedness would have re-
sulted in a switch between even and odd correla-
tion structure f or (
19, 20). Since the handedness
of the local question basis is just a convention
by O or O
, we shall allow either. Consequently,
whether three correlation questions are related by
even or odd correlation (see theorem
5.7) depends
on the local conventions by O, O
and both are, in
fact, consistent. H owever, (28) will always hold.
Usually, of course, one would favour the situation
that both O, O
make the same conventions such
that their local question bases are equally handed
(or that one observer O describes all qubits with
the same handedness). Nevertheless, physically
all local conventions are fully equivalent, yet will
lead to different representations of composite sys -
tems in the inference theory.
But there are important consistency condi-
tions on the distribution of odd and even cor-
relations. This becomes obvious when examining
three qubits A, B, C. To this end, consider the
trivial conjunction
Q
3
A
3
B
Q
3
A
3
C
Q
3
B
3
C
=
lemma
5.10
1. (37)
From (
19) we know that the correlation is either
even or odd, respectively,
Q
3
A
3
B
= Q
1
A
1
B
Q
2
A
2
B
, or
Q
3
A
3
B
= ¬(Q
1
A
1
B
Q
2
A
2
B
) (38)
and analogously for A, C and B, C. Suppose now
that all three bipartite correlations in the con-
junction (37) were even. Then we immediately
get a contradiction because, using lemma
5.10,
Q
3
A
3
B
Q
3
A
3
C
Q
3
B
3
C
= (Q
1
A
1
B
Q
2
A
2
B
) (Q
1
A
1
C
Q
2
A
2
C
) (Q
1
B
1
C
Q
2
B
2
C
)
=
(
28)
¬(Q
1
A
1
B
Q
1
A
1
C
)
|
{z }
= Q
1
B
1
C
(Q
2
A
2
B
Q
2
A
2
C
)
|
{z }
= Q
2
B
2
C
(Q
1
B
1
C
Q
2
B
2
C
)
= 0.
But this violates the identity (37) and results
from the relative negation between the left and
right hand side in (28). One would get the same
37
Ultimately, these three questions will define an or-
thonormal Bloch vector basis in the Bloch sphere (see sec-
tion
7). The handedness or orientation of this basis will
depend on the lab eling of question outcomes.
contradiction if one of the correlations in (
37)
was even and two were odd because then the two
negations from the odd correlation would cancel
each other and one would still be left with the
negation coming from (
28).
On the other hand, everything is consistent if
either all bipartite correlations in (
38) are odd or
one is odd and two are even because then the odd
Accepted in Quantum 2017-11-27, click title to verify 49
number of negations from the odd correlations cancels the negation coming f rom (28):
Q
3
A
3
B
Q
3
A
3
C
Q
3
B
3
C
= ¬(Q
1
A
1
B
Q
2
A
2
B
) ¬(Q
1
A
1
C
Q
2
A
2
C
) ¬(Q
1
B
1
C
Q
2
B
2
C
)
=
(
28)
¬(Q
1
A
1
B
Q
1
A
1
C
)
|
{z }
= Q
1
B
1
C
(Q
2
A
2
B
Q
2
A
2
C
)
|
{z }
= Q
2
B
2
C
¬(Q
1
B
1
C
Q
2
B
2
C
)
= 1.
We can also quickly check that this is consistent
with (
20)
Q
3
A
3
B
= Q
1
A
2
B
Q
2
A
1
B
, or
Q
3
A
3
B
= ¬(Q
1
A
2
B
Q
2
A
1
B
) (39)
and (
26) (and analogously for A, C and B, C)
Q
1
A
2
B
Q
2
A
1
B
= ¬(Q
1
A
1
B
Q
2
A
2
B
).
That is, if all correlations in (38) are odd, then
all correlations in (
39) must be even. Indeed,
Q
3
A
3
B
Q
3
A
3
C
Q
3
B
3
C
= (Q
1
A
2
B
Q
2
A
1
B
) (Q
1
A
2
C
Q
2
A
1
C
) (Q
1
B
2
C
Q
2
B
1
C
)
=
(
28)
¬(Q
1
A
2
B
Q
1
A
2
C
)
|
{z }
= Q
2
B
2
C
(Q
2
A
1
B
Q
2
A
1
C
)
|
{z }
= Q
1
B
1
C
(Q
1
B
2
C
Q
2
B
1
C
)
= Q
3
B
3
C
Q
3
B
3
C
= 1
is consistent.
In conclusion, if O wants to treat the bipar-
tite relations among all three A, B, C i dentically,
then the following distribution of odd and even
correlations
Q
3
A
3
B
= Q
1
A
2
B
Q
2
A
1
B
= ¬(Q
1
A
1
B
Q
2
A
2
B
), (40)
and analogously for A, C and B, C, is the only
consistent solution. We shall henceforth make the
convention that the bipartite correlation structure
among any pair of q ubits be the same such that
(
40) holds. This turns out to be the case of qubit
quantum theory.
The tacit assumption, underlying the stan-
dard representation of qubit quantum theory, is
that the handedness of each local qubit ques-
tion basis for A, B, C is the same, e.g., all ‘left’
or ‘right’ handed. But we emphasize, that it
would be equally consistent to choose one basis as
‘left’ (‘right’) and the other two as ‘right’ (‘left’)
handed. A qubit pair with equally handed bases
will be described by the odd and even correla-
tion distribution as in (
40), while a qubit pair
with oppositely handed bases will be described
by the opposite distribution of odd and even cor-
relations. In terms of whether Q
33
= Q
11
Q
22
is even or odd Q
33
= ¬(Q
11
Q
22
), the three
qubit relations yield only four consistent graphs
for the distribution of ‘left’ and ‘right’ handedness
PSfrag replacements
odd odd
odd
even
‘left’
‘left’ ‘left’
‘right’
PSfrag replacements
odd
odd odd
even
‘left’
‘right’
‘right’‘right’
PSfrag replacements
odd
even
even
‘left’
‘left’‘right’
PSfrag replacements
odd
even
even
‘left’ ‘right
‘right’
(41)
This gives a simple graphical explanation for
the consistency observations above. The first
two graphs correspond to quantum theory. The
framework with the oppos ite correlation structure
of quantum theory, corresponding to the last two
graphs, is sometimes referred to as mirror quan-
Accepted in Quantum 2017-11-27, click title to verify 50
tum theory [15]. While mirror quantum theory
was considered inconsistent in [
15], we see here
that inconsistencies would only arise if the bi-
partite correlation structure of mirror quantum
theory (and in particular the even correlation
Q
33
= Q
11
Q
22
) was used for all three pairs of
qubits. However, the proper formulation of mir-
ror quantum theory for three qubits corresponds
to the last two graphs in (
41) and gives a per-
fectly consistent framework.
38
It could be easily
generalized in an obvious manner to arbitrarily
many q ubits.
39
Obviously, the s ame argument can be carried
out f or any other triple of maximally compatible
bipartite correlations appearing in theorem 5.7.
In conclusion, the two distinct consistent distri-
butions of odd and even correlations, correspond-
ing to quantum theory and mirror quantum the-
ory,
(a) result from different conventions of local ba-
sis handedness, and
(b) are in one-to-one correspondence through a
local relabeling ‘yes’‘no’ of one individual
question.
As such, the two distinct correlation structures
ultimately give rise to two distinct representa-
tions of the same physics and are thus fully equiv-
alent. The transformation (b) between the two
representations being a translation between two
conventions/descriptions is a passive one and
38
In fact, one could even produce the correlation struc-
ture of mirror quantum theory in the lab by using oppo-
sitely handed bases for two qubits in an entangled p air.
The resulting state would be represented by the partial
transpose of an entangled qubit state. This state would
not be positive if represented in terms of the standard
Pauli matrices and thus not correspond to a legal quan-
tum state [
88]. However, the point is that mirror quantum
theory would not be represented in the standard Pauli
matrix basis but in the partially transposed basis which
corresponds to replacing
σ
y
=
0 i
i 0
by σ
T
y
=
0 i
i 0
,
for one of t he two qubits (this is a switch of the y-axis
orientation). In this basis, the state would be positive.
39
Notice, however, that for more than three qubits one
would get more than two different consistent distributions
of odd and even correlations. For ex ample, for four qubits
there will be three cases: (1) all four have equally handed
bases, (2) three have equally hand ed bases, (3) two have
equally handed bases.
can be carried out on a piece of paper; it is always
allowed. However, clearly, there cannot exist any
actual physical transformation in the laboratory
which maps states from one convention into the
other.
Within the formalism of quantum theory the
transformation (b) Q
1
7→ ¬Q
1
corresponds to
the partial transpose (e.g., see [
15]). It is well
known that the partial transpose defines a sepa-
rability criterion for quantum states which is both
necessary and sufficient for a pair of qubits [
88]:
a two-qubit density matrix ρ is separable if and
only if its partial transpose is positive relative to
a basis of standard Pauli matrix products (i.e.,
represents a legal q uantum state). This criterion
holds analogously in our language here: as seen at
the beginning of this section, the transformation
Q
1
7→ ¬Q
1
changes between odd and even corre-
lations of bipartite correlations Q
ij
. This would,
in fact, be unproblematic if O only had individ-
ual information about the two qubits; it would
map a classically composed state even of maxi-
mal information, say, Q
1
= 1 and Q
1
= 1 and
thus Q
11
= 1, to another legal classically com-
posed state Q
1
= 0, Q
1
= 1 and thus Q
11
= 0.
Both states exist within both conventions. How-
ever, applying this transformation to a maxi-
mally entangled state with odd correlation, say,
Q
11
= Q
22
= 1 and thus, by (
40), Q
33
= 0 yields
an even correlation Q
11
= 0 = Q
33
and Q
22
= 1.
The former state only exists in the quantum the-
ory representation, while the latter exists only in
the mirror image. For other entangled states one
would similarly find that Q
1
7→ ¬Q
1
necessar-
ily maps from one representation into the other.
The same conclusion also holds for a transfor-
mation {Q
1
, Q
2
, Q
3
} 7→ Q
1
, ¬Q
2
, ¬Q
3
}, which
one might call total inversion, because the odd
number of negations involved in the transforma-
tion would likewise lead to a swap of odd and
even correlations.
Lastly, we note that the situation is very differ-
ent for rebit theory because it is its own mirror
image, i.e. rebit theory and mirror rebit theory
are identical representations. O will describ e a
single rebit by a question basis Q
1
, Q
2
. Suppose
O decided to swap the ‘yes’ and ‘no’ assignments
to the outcomes of Q
1
, such that equivalently
Q
1
7→ ¬Q
1
. For a pair of rebits, this would have
Accepted in Quantum 2017-11-27, click title to verify 51
the f ollowing ramification
Q
11
, Q
12
7→ ¬Q
11
, ¬Q
12
,
Q
21
, Q
22
7→ Q
21
, Q
22
,
and therefore
Q
12
Q
21
7→ ¬(Q
12
Q
21
)
Q
33
7→ ¬Q
33
.
In contrast to the qubit case, Q
33
is defined as
a correlation of correlations Q
33
:= Q
12
Q
21
(
22) and can not be written in terms of local
questions Q
3
, Q
3
. Hence, Q
33
also changes un-
der this transformation by construction. Accord-
ingly, Q
1
7→ ¬Q
1
does not lead to a swap of odd
and even correlations for the rebit case. This
‘partial transpose’ therefore always maps states
to other states within the same representation.
For example, even a maximally entangled state
of maximal information and even correlation, say,
Q
12
= Q
21
= 1 and Q
33
= 1, would be mapped to
another evenly correlated state Q
12
= 0, Q
21
= 1
and Q
33
= 0. As a consequence, the Peres sep-
arability criterion [
88] which is valid for qubits
does not hold in analogous fashion for rebits. For
completely equivalent reasons, the total inver-
sion, corresponding to {Q
1
, Q
2
} 7→ Q
1
, ¬Q
2
},
is also a transformation which preserves the rep-
resentation.
5.4.2 Collecting the results: odd and even corre-
lation s tructure for N = 2
After the many technical details it is useful to
collect all the results concerning the compatibil-
ity, complementarity and correlation structure for
two qubits, derived in lemmas
5.2 and 5.5, theo-
rem
5.7, equation (28) and in the previous section
5.4.1 in a graph to facilitate a visualization. We
shall henceforth abide by the convention that all
bipartite relations for arbitrarily many qubits be
treated eq ually such that (
40) must hold. For
the other relations of theorem
5.7 one finds the
analogous results. As can be easily verified, the
ensuing question s tructure has the lattice pattern
of figure
4, where the triangles
PSfrag replacements
+
Q Q
Q
′′
Q = ¬(Q
Q
′′
),
PSfrag replacements
+
Q Q
Q
′′
Q = Q
Q
′′
(42)
denote odd and even correlation, respectively.
We recall that (26, 28) imply alternating odd
and even correlation triangles for the bipartite
correlation questions. However, we emphasize
that the Bell scenario argument of section
5.2.7
does not require the correlation triangles involv-
ing individual questions to also admit such an
alternating odd and even pattern. For instance,
the f ollowing relations of Q
33
Q
33
= Q
3
Q
3
= Q
12
Q
21
= ¬(Q
11
Q
22
)
are consistent with assumptions
5 and 6, despite
the absence of a negation in Q
3
Q
3
= Q
12
Q
21
because the latter is not a classical logical
identity. The graph corresponding to the last re-
lation,
PSfrag replacements
.
.
.
Q
11
Q
22
Q
12
Q
21
Q
33
Q
3
Q
3
is also different from the graphs (
23) as it in-
volves all six individual questions and the argu-
ment leading to (28) does not apply. Clearly,
given the definition Q
ij
:= Q
i
Q
j
, all such
triangles must be even.
The lattice structure in figure
4 contains 15 tri-
angles and 15 × 3 = 45 distinct edges correspond-
ing to compatibility relations. There are 60 ‘miss-
ing edges’ corresponding to complementarity re-
lations. Every question resides in three compat-
ibility triangles and is therefore maximally com-
patible with the six other questions in these three
triangles and maximally complementary to the
eight remaining q uestions not contained in those
three triangles. This means that once one ques-
tion is fully known, all other information available
to O must be distributed over the three adjacent
Accepted in Quantum 2017-11-27, click title to verify 52
PSfrag rep lacements
++
+
Q
1
Q
2
Q
3
Q
1
Q
2
Q
3
Q
11
Q
22
Q
12
Q
33
Q
13
Q
13
Q
21
Q
23
Q
23
Q
31
Q
32
identify identify
PSfrag rep lacements
+
+
+
+
+
+
+
+
+
Q
1
Q
2
Q
3
Q
1
Q
1
Q
2
Q
2
Q
3
Q
3
Q
11
Q
22
Q
12
Q
33
Q
13
Q
21
Q
23
Q
31
Q
32
identify
identify
identify
Figure 4: A lattice representation of the compl e t e compati bility, complementarity and correlation s t ructure of the
informationally comple t e s e t Q
M
2
for two qubits. Every vertex corresponds to one of the 15 pairwise independent
questions. If two questions are conn e c ted by an edge, they are maximally compatible. If two questions are not
connected by an edg e , they are maximally complementary. Thanks to the logical structure of the XNOR , defining
a question as a correlation of two other questions, the compatibility structure results in a lattice o f triangles. As
clarified in (
42), red triangles deno te odd, while green triangl e s denote even c orrelation. Note that the two lattices
represented here are c onnected through the nine correlation questions Q
ij
and form a s ingle c losed l attice (which,
however, is easier to represent in this disconnecte d manner). Every question resides in exactly three triangles and is
thereby maximally compatible with six and maximally complementary to eight other qu e stions.
triangles. In particular, if O knows the answers
to two of the questions in the lattice, he will also
know the answer to the third sharing the same
triangle such that every triangle corresponds to
a specific set of states of maximal information.
(Although, as we shall see shortly in section
6.8
and, in more detail, in [1], the time evolution rule
4 implies that s tates of maximal information will
likewise exist such that two independent bits can
be distributed differently over the lattice.)
Notice that every triangle is connected by an
edge (adjacent to one of its three vertices) to ev-
ery other vertex in the lattice. This embodies
the statement ‘whenever O asks S a new ques-
tion, he experience no net loss of information’ of
the complementarity rule
2. For instance, if O
has maximal knowledge about two questions in
the lattice, corresponding to maximal informa-
tion about one triangle, it is impossible for him
to loose information by asking another question
from the lattice because it will be connected to
one of the questions from the previous triangle.
As a concrete example, suppose O knew the an-
swers to Q
21
, Q
13
, Q
32
. H e could ask Q
22
next.
Since Q
22
is connected by an edge to Q
13
, he
would k now the answers to both upon asking Q
22
and thereby then also the answer to Q
31
no net
loss of information occurs.
Accepted in Quantum 2017-11-27, click title to verify 53
It is straightforward to check, e.g., using the
ansatz
|ψi = α |z
z
i + β |z
+
z
+
i + γ |z
z
+
i + δ|z
+
z
i
for a two qubit pure state and translating it into
the various basis combinations xx, xy, yx, yy, . . .,
that the lattice structure of figure
4 is pre-
cisely the compatibility and correlation struc-
ture of qubit q uantum theory. For instance,
Q
11
, Q
22
, Q
33
correspond to projectors onto the
+1-eigenspaces of σ
x
σ
x
, σ
y
σ
y
and σ
z
σ
z
.
The three questions sharing a red triangle means,
e.g., that Q
11
= Q
22
= 1 and Q
33
= 0 is an al-
lowed state while Q
11
= Q
22
= Q
33
= 1 is illegal.
Indeed, ignoring normalization, in quantum the-
ory one finds
|x
+
x
+
i |x
x
i = i|y
+
y
+
i + i|y
y
i
= |z
+
z
i + |z
z
+
i
for α = β = 0 and γ = δ = 1, corresponding to
the prop os itions Q
11
= 1: “the spins are corre-
lated in x-direction"; Q
22
= 1: “the spins are
correlated in y-direction"; and Q
33
= 0: “the
spins are anti-correlated in z-direction".
40
But
no quantum state exists such that the spins are
also correlated in z-direction if they are correlated
in x- and y-direction (this would be mirror quan-
tum theory). Every other triangle in the lattice
corresponds similarly to four pure quantum states
(representing the answer configurations ‘yes-yes’,
‘yes-no’, ‘no-yes’, ‘no-no’ to the two independent
questions per triangle).
Finally, we also collect the results on the com-
patibility, complementarity and correlation struc-
ture of two rebits, derived in lemmas
5.2, 5.5 and
5.8, theorem 5.9 and equation (26), in a lattice
structure in figure
5. There are six triangles and
6 × 3 = 18 edges representing compatibility rela-
tions. Every question resides in exactly two tri-
angles and is thereby maximally compatible with
four and maximally complementary to four other
questions. As in the qubit case in figure
4, every
triangle is connected by an edge to every other
vertex in the lattice, in conformity with rule 2
40
Similarly, |ψi = α |z
z
i + β |z
+
z
+
i =
α+β
2
(|x
+
x
+
i +
|x
x
i) +
βα
2
(|x
+
x
i + |x
x
+
i) =
α+β
2
(|y
+
y
i +
|y
y
+
i)+
βα
2
(|y
+
y
+
i+ |y
y
i). But |x
+
x
+
i+ |x
x
i =
|y
+
y
i+ |y
y
+
i and |y
+
y
+
i+ |y
y
i = |x
+
x
i+ |x
x
+
i.
Hence, once Q
33
= 1, one indeed gets Q
11
Q
22
= 0
even though α, β are unspecified such that Q
11
, Q
22
may
be u nknown.
PSfrag rep lacements
++
+
+
+
Q
1
Q
2
Q
3
Q
1
Q
2
Q
3
Q
22
Q
11
Q
12
Q
33
Q
1
Q
1
Q
21
Q
2
Q
2
Q
1
Q
2
identify identify
Figure 5: A latti ce representation of the complete com-
patibility, complementarity and correlation structure of
the informationally complete set Q
M
2
for two rebits.
The explana tions of the caption of figure
4 also apply
here.
asserting that O shall not experience a net loss of
information by asking further questions.
5.5 The general case of N > 3 gbits
We are now prepared to investigate the indepen-
dence, compatibility and correlation structure en-
suing from rules
1 and 2 on a general Q
N
, i.e. of
O’s possible questions to a system S composed of
N > 3 gbits. In particular, we shall exhibit an in-
formationally complete set Q
M
N
for both qubits
and rebits. O can view the N gbits as a composite
system in many different ways: as N individual
gbits, as one individual gbit and a composite sys-
tem of N 1 gbits, as a system composed of 2
gbits and a system composed of N 2 gibts, and
so on. All these different compositions yield, of
course, the same question structure. It is simplest
to interpret the N gbits recursively as being com-
posed of a composite system of N 1 gbits and a
new individual gbit. Definition
3.5 of a composite
system then implies that Q
N
must contain (1) the
questions of Q
N1
for N 1 gbits, (2) the set Q
1
of the new gbit, and (3) all logical connectives of
the maximally compatible questions of those two
sets. This entails slightly different repercussions
for qubits and rebits.
5.5.1 An informationally complete set and entan-
glement for N > 3 qubits
A natural candidate f or an informationally com-
plete question set is given by the set of all possible
XNOR conjunctions of the individual questions of
Accepted in Quantum 2017-11-27, click title to verify 54
the N gbits
Q
µ
1
µ
2
···µ
N
:= Q
µ
1
Q
µ
2
· · · Q
µ
N
(43)
(we recall from (
13) that the logical connective
yielding independent q uestions is either or ).
Here we have introduced a new index notation:
µ
a
is the question index for gbit a {1, . . . , N }
and can take the values 0, 1, 2, 3. As before the
index values i = 1, 2, 3 correspond to the indi-
viduals Q
1
a
, Q
2
a
, Q
3
a
. On the other hand, the
index value µ
a
= 0 implies that none of the three
individual questions of gbit a appears in the con-
junction (
43), i.e. always Q
0
a
1. For instance,
Q
100000···000
:= Q
1
1
, Q
0003020···0
:= Q
3
4
Q
2
6
,
etc. Note that Q
000···000
corresponds to no ques-
tion. The set (
43) thus contains all individual,
bipartite, tripartite, and up to N-partite cor-
relation questions for N gbits. We emphasize
that, due to the special multipartite structure
(and associativity) of the XNOR, a question such
as Q
111···111
does not incarnate the question ‘are
the answers to Q
1
1
, Q
1
2
, · · · , Q
1
N
all the same?’.
For ex ample, for N = 4, Q
1
1
= Q
1
2
= 0 and
Q
1
3
= Q
1
4
= 1 also yield Q
1
1
1
2
1
3
1
4
= 1. This is
important for the entanglement structure.
We begin with an important result.
Lemma 5.25. The 4
N
1 questions
41
Q
µ
1
···µ
N
,
µ = 0, 1, 2, 3, are pairwise independent.
Proof. Consider Q
µ
1
···µ
N
and Q
ν
1
···ν
N
. The two
questions must disagree in at least one index oth-
erwise they would coincide. Let the questions
differ on the ind ex of gbit a, i.e. µ
a
6= ν
a
. Sup-
pose µ
a
, ν
a
6= 0. Then Q
µ
a
is maximally compat-
ible with Q
µ
1
···µ
a
···µ
N
= Q
µ
1
· · · Q
µ
a
· · · Q
µ
N
and maximally complementary to
Q
ν
1
···ν
a
···ν
N
= Q
µ
1
· · · Q
ν
a
· · ·
Q
ν
N
since Q
µ
a
, Q
ν
a
are maximally complemen-
tary. Using the same argument as in the proof of
lemma
5.3 this implies independence of Q
µ
1
···µ
N
and Q
ν
1
···ν
N
. Lastly, suppose now µ
a
= 0 and
ν
a
6= 0 (as we have µ
a
6= ν
a
not both can be 0).
Then Q
i
a
6=ν
a
with i 6= 0 is maximally compatible
with Q
µ
1
···µ
N
and maximally complementary to
Q
ν
1
···ν
N
. By the same argument, this again im-
plies independence of Q
µ
1
···µ
N
and Q
ν
1
···ν
N
.
Consequently, the set (43) will be part of an
informationally complete set. We note that a
hermitian matrix of trace equal to 1 on a 2
N
-
dimensional complex Hilbert space (i.e. qubit
density matrix) is described by 4
N
1 param-
eters.
Next, we must elucidate the compatibility and
complementarity structure of this set.
Lemma 5.26. Q
µ
1
···µ
N
and Q
ν
1
···ν
N
are
maximally compatible if the index sets
{µ
1
, . . . , µ
N
} and {ν
1
, . . . , ν
N
} differ by an
even numbe r (incl. 0) of non-zero indices,
maximally complementary if the index sets
{µ
1
, . . . , µ
N
} and {ν
1
, . . . , ν
N
} differ by an
odd number of non-zero indices.
For example, for N = 2, Q
11
and Q
22
differ
by two non-zero indices and are thus maximally
compatible. By contrast, Q
10
= Q
1
and Q
22
dif-
fer by an odd number of non-zero indices and are
thereby maximally complementary.
Proof. Let Q
µ
1
···µ
N
and Q
ν
1
···ν
N
disagree in an
odd number, call it 2n + 1, of non-zero indices.
We can always reshuffle the index labeling of the
N qubits such that now a = 1, . . . , 2n + 1 cor-
responds to the q ubits on which Q
µ
1
···µ
N
and
Q
ν
1
···ν
N
differ by non-zero indices, i.e. µ
a
6= ν
a
and µ
a
, ν
a
6= 0. The remaining qubits labeled
by b = 2n + 2, . . . , N are then such that Q
µ
1
···µ
N
and Q
ν
1
···ν
N
either agree on the non-zero in dex,
µ
b
= ν
b
6= 0 or at least one of µ
b
, ν
b
is 0. That
is, after the reshuffling the index labeling, we can
write the questions as
Q
µ
1
···µ
N
= (Q
µ
1
· · · Q
µ
2n+1
)
| {z }
disagreement
(Q
µ
2n+2
· · · Q
µ
N
)
| {z }
maximally compatible
Q
ν
1
···ν
N
=
z }| {
(Q
ν
1
· · · Q
ν
2n+1
)
z }| {
(Q
ν
2n+2
· · · Q
ν
N
) . (44)
41
We deduct the trivial question Q
000···000
. Obviously,
one arrives at the same number by counting the distri-
bution of individuals, bipartite,...and N -partite correla-
tions over N qubits as a binomial series
P
N
k=1
N
k
3
k
=
(3 + 1)
N
1.
Accepted in Quantum 2017-11-27, click title to verify 55
The parts of the questions where the index
sets either agree or feature zeros coincide with
Q
µ
2n+2
···µ
N
and Q
ν
2n+1
···ν
N
and are clearly maxi-
mally compatible (drop ping here zero indices).
We can now proceed by induction. Lemmas
5.2, 5.5, 5.105.16 imply that the statement of
this lemma is correct for n = 0, 1. Let the state-
ment therefore be true for n and consider n + 1.
Then, (
44) reads
Q
µ
1
···µ
N
=(Q
µ
1
···µ
2n+3
)
| {z }
disagreement
Q
µ
2n+4
···µ
N
| {z }
maximally compatible
= (Q
µ
1
· · · Q
µ
2n+1
)
| {z }
maximally complementary
(Q
µ
2n+2
µ
2n+3
)
| {z }
disagreement
Q
µ
2n+4
···µ
N
| {z }
max. compat.
Q
ν
1
···ν
N
=
z
}| {
(Q
ν
1
···ν
2n+3
)
z }| {
Q
ν
2n+4
···ν
N
=
z }| {
(Q
ν
1
· · · Q
ν
2n+1
)
z }| {
(Q
ν
2n+2
ν
2n+3
)
z }| {
Q
ν
2n+4
···ν
N
.
But, by lemma 5.5, Q
µ
2n+2
µ
2n+3
and Q
ν
2n+2
ν
2n+3
are maximally compatible with each other and
therefore also with Q
µ
1
···µ
N
and Q
ν
1
···ν
N
. This
implies that both Q
µ
1
···µ
N
and Q
ν
1
···ν
N
are max-
imally compatible with and, thanks to lemma
5.25, independent of, e.g., Q
µ
2n+2
···µ
N
. Com-
plementarity of Q
µ
1
···µ
N
and Q
ν
1
···ν
N
now fol-
lows from lemma 5.1 and noting th at Q
µ
1
···µ
2n+1
and Q
ν
1
···ν
N
are maximally complementary be-
cause they disagree in 2n + 1 non-zero ind ices
(for which, by assumption, the s tatement of the
lemma holds).
Finally, let Q
α
1
···α
N
and Q
β
1
···β
N
disagree in
an even number 2n of non-zero indices. Using
an analogous reshuffling of the index labeling as
in the odd case above, one can rewrite the two
questions as
Q
α
1
···α
N
= Q
α
1
α
2
|
{z }
differ
Q
α
3
α
4
|
{z }
differ
· · · Q
α
2n1
α
2n
|
{z }
differ
Q
α
2n+1
···α
N
| {z }
maximally compatible
Q
β
1
···β
N
=
z
}| {
Q
β
1
β
2
z }| {
Q
β
3
β
4
· · ·
z }| {
Q
β
2n1
β
2n
z }| {
Q
β
2n+1
···β
N
. (45)
That is, one can decompose the disagreeing parts
of the questions into bipartite correlations. But
thanks to lemma
5.5 two bipartite correlations of
the same qubit pair are maximally compatible if
and only if they differ in both indices. Conse-
quently, all the pairings of question components
of the up per and lower line, as w ritten in (
45),
are maximally compatible and, hence, so must be
Q
α
1
···α
N
and Q
β
1
···β
N
.
Fortunately, it turns out that the 4
N
1 ques-
tions (
43) are logically closed under the XNOR.
Theorem 5.27. (Qubits) The 4
N
1 questions
Q
µ
1
···µ
N
, µ = 0, 1, 2, 3, are logically closed under
and thus form an informationally complete set
Q
M
N
with D
N
= 4
N
1 for the case D
1
= 3.
Proof. We shall prove the statement by induc-
tion. The statement is trivially true for N = 1
and, by theorems
5.7 and 5.17, holds also for
N = 2, 3. Let the statement th erefore be true
for N 1 and consider a composite system of N
qubits. Since O can treat the N qubit system
in many different ways as a composite system
and th e statement holds for N 1, all XNOR
conjunctions of questions involving at least one
zero ind ex will be contained in (
43). We thus
only need to show th at all conjunctions in -
volving at least one N-partite correlation Q
i
1
···i
N
,
i
a
6= 0 a = 1, . . . , N, produce questions already
included in (43).
Consider Q
i
1
···i
N
and Q
ν
1
···ν
N
. Lemma
5.26 im-
plies that these two questions are maximally com-
patible and can thus be connected by if and
only if {i
1
, . . . , i
N
} and {ν
1
, . . . , ν
N
} disagree in
an even number of non-zero indices. There are
now two cases that we must consider:
(a) Suppose Q
i
1
···i
N
and Q
ν
1
···ν
N
disagree in
an even number of non-zero indices and, further-
more, agree on at least one index i
a
= ν
a
. Thanks
to Q
i
a
Q
i
a
= 1, the conjunction then yields
Accepted in Quantum 2017-11-27, click title to verify 56
Q
i
1
···i
a
···i
N
Q
ν
1
···i
a
···ν
N
= Q
i
1
··· 0
position a
···i
N
Q
ν
1
··· 0
position a
···ν
N
.
Hence, the two questions on the right hand side
both contain less than N non-zero indices such
that the result must lie in the set (
43) because
the statement is true up to N 1 by assumption.
(b) Suppose Q
i
1
···i
N
and Q
ν
1
···ν
N
disagree in an
even number 2n of non-zero in dices and do not
agree on any n on -zero index. Reshuffling the in-
dex labelings as in the proof of lemma
5.26, one
can th en write
Q
i
1
···i
N
= Q
i
1
i
2
|
{z}
differ
Q
i
3
i
4
|
{z}
differ
· · · Q
i
2n1
i
2n
|
{z }
differ
Q
i
2n+1
···i
N
Q
ν
1
···ν
N
=
z
}|{
Q
ν
1
ν
2
z}|{
Q
ν
3
ν
4
· · ·
z }| {
Q
ν
2n1
ν
2n
Q
0
2n+1
···0
N
.
(We obtain here Q
0
2n+1
···0
N
because Q
i
1
···i
N
and
Q
ν
1
···ν
N
do not agree on any common in dex and
Q
i
1
···i
N
does not feature zero-indices.) By lemma
5.5, the pairs of bip artite correlations d iffering in
two in dices, e.g., Q
i
1
i
2
and Q
ν
1
ν
2
, are maximally
compatible and by theorem
5.7 their XNOR con-
junction will yield another bipartite correlation of
the same qubit pair. For example, Q
i
1
i
2
Q
ν
1
ν
2
equals either Q
j
1
j
2
or ¬Q
j
1
j
2
for j
1
6= i
1
, ν
1
and
j
2
6= i
2
, ν
2
. Accordingly, up to negation, one finds
(j
a
6= i
a
, ν
a
, a = 1, . . . , 2n)
Q
i
1
···i
N
Q
ν
1
···ν
N
= Q
j
1
j
2
Q
j
3
j
4
· · · Q
j
2n1
j
2n
Q
i
2n+1
···i
N
= Q
j
1
···j
2n
i
2n+1
···i
N
which is another N-partite correlation contained
in the set (
43).
As in the cases N 3, we could represent
the compatibility and complementarity relations
for N > 3 qubits geometrically by a simplicial
question graph where a D-partite question corre-
sponds to a (D 1)-dimensional simplex within
the graph (D N). The compatibility and com-
plementarity relations of lemma 5.26 then trans-
late into abstract geometric relations according
to whether a D-simplex and a D
-simplex share
or disagree on subsimplices. In particular, the
criterion that two distinct questions in Q
M
N
are
maximally compatible if and only if they disagree
on an even number of non-zero indices means ge-
ometrically that the two questions are maximally
compatible if and only if the two subsimplices in
the two question simplices which correspond to
the qubits they share in common either
(a) do not overlap and are odd-dimensional, or
(b) coincide.
For example, for N = 2, Q
11
, Q
22
correspond to
two non-intersecting edges, i.e. 1-simplices; they
are maximally compatible because they disagree
on their 1-simplices, both involving qubit 1 and 2.
On the other hand, for N = 3, Q
1
A
1
B
, Q
2
B
2
C
also
correspond to two non-intersecting 1-simplices
but are maximally complementary because the
two edges involve different qubits A, B for the
first and B, C for the second edge.
This also helps us to understand entanglement
for arbitrarily many qubits. Specifically, maxi-
mal entanglement will correspond to O spending
the N independent bits he is allowed to acquire
about the system S of N qubits on N-partite cor-
relation questions. Lemma
5.26 guarantees that
for every N there will exist N maximally com-
patible N-partite questions. For instance, there
are
N
2
ways of having N 2 of the indices take
value 1 and 2 indices take the value 2. Any two
of the
N
2
corresponding questions will disagree
in two or four non-zero indices (and agree on the
rest) and will thus be maximally compatible (even
mutually according to theorem
3.2). These will
Accepted in Quantum 2017-11-27, click title to verify 57
correspond to a set of
N
2
maximally compati-
ble (N 1)-simplices in the question graph such
that any two of them either disagree on an edge
or a tetrahedron. (There will exist even more
maximally compatible N -partite questions.) For
N 3 it also holds that
N
2
N.
It is definitely possible to choose N such maxi-
mally compatible N-partite correlation questions
out of the
N
2
many such that these N questions
do not all agree on a s ingle index. For similar rea-
sons to the N = 2, 3 cases, this choice will con-
stitute a mutually independent set such that ev-
ery individual question Q
i
1
, . . . , Q
i
N
will be max-
imally complementary to at least one of these N
N-partite questions. (As a consequence of rule
1,
once the answers to these N N-partite questions
are known, they will also imply the answers to the
N
2
N remaining ones by the same reasoning
as in s ection
5.2.3.) Accordingly, O can exhaust
the information limit with these N-partite corre-
lation questions, while not being able to have any
information whatsoever about the individuals a
necessary condition for maximal entanglement.
There will exist many different ways of having
such multipartite entanglement for arbitrary N.
One could describe such different ways of entan-
glement by generalizing the correlation measures
(
33) and informational monogamy inequalities re-
sulting therefrom. These monogamy inequalities
could also be considered as simplicial relations:
they restrict the way in which the available (inde-
pendent and dependent) information can be dis-
tributed over the various simplices and subsim-
plices in the question graph. However, we abstain
from analyzing such relations here further.
5.5.2 An informationally complete set and entan-
glement for N > 3 rebits
We briefly repeat the same procedure for N
rebits. In analogy to (
43), the natural candidate
set for an informationally complete Q
M
N
will con-
tain
Q
µ
1
µ
2
···µ
N
:= Q
µ
1
Q
µ
2
· · · Q
µ
N
,
µ
a
= 0, 1, 2, a = 1, . . . , N, (46)
where the notation should be clear from sec-
tion
5.5.1. However, being a composite sys-
tem, by definition
3.5 we must permit the cor-
relation of correlations Q
3
a
3
b
(
22) for all a, b
{1, . . . , N } because clearly O is allowed to ask
Q
1
a
, Q
2
a
, Q
1
b
, Q
2
b
. Furthermore, thanks to (
36)
we have, e.g.,
Q
3
1
3
2
3
3
3
4
= Q
3
1
3
2
Q
3
3
3
4
= Q
3
1
3
3
Q
3
2
3
4
= Q
3
1
3
4
Q
3
2
3
3
such that no confusion can arise about the mean-
ing of Q
3
1
3
2
3
3
3
4
although there are no individu-
als Q
3
a
into which the question could be decom-
posed. The same holds similarly for any other
even number of indices taking the value 3. Con-
sequently, the candidate set for Q
M
N
can be writ-
ten as
˜
Q
M
N
:=
Q
µ
1
···µ
N
, µ
a
= 0, 1, 2, 3, a = 1, . . . , N
only even number of indices taking value 3
with an evident meaning of each such question.
Let us count the number of elements within
˜
Q
M
N
.
Lemma 5.28.
˜
Q
M
N
contains 2
N1
(2
N
+ 1) 1
non-trivial questions.
Proof. If arbitrary distributions of the values
µ
a
= 0, 1, 2, 3 over the a = 1, . . . , N were per-
mitted, we would obtain 4
N
1 non-trivial ques-
tions upon subtracting Q
0
1
0
2
···0
N
as in the qub it
case. In order to obtain the number of questions
within
˜
Q
M
N
we thus still have to s ubtract all the
possible ways of d istributing an odd number of
3’s over the N indices. There are precisely
O
N
:=
N
1
!
3
N1
+
N
3
!
3
N3
+ · · · +
N
2n + 1
!
3
N(2n+1)
such ways, where 2n+1 is the largest odd number
smaller or equal to N. Similarly, the number of
ways an even number of 3’s can be distributed
over N indices is given by
E
N
:= 3
N
+
N
2
!
3
N2
+ · · · +
N
2m
!
3
N2m
,
Accepted in Quantum 2017-11-27, click title to verify 58
where 2m is the largest even number smaller or
equal to N. We then have
E
N
+ O
N
= (3 + 1)
N
= 4
N
,
E
N
O
N
= (3 1)
N
= 2
N
and thus
O
N
=
1
2
(4
N
2
N
)
which yields
4
N
1 O
N
= 2
N1
(2
N
+ 1) 1
non-trivial questions in
˜
Q
M
N
.
We note that the number of parameters in a
symmetric matrix with trace equal to 1 on a 2
N
-
dimensional real Hilbert space (i.e. rebit density
matrix) is precisely
1
2
2
N
(2
N
+ 1) 1.
Next, we assert pairwise independence as re-
quired f or an informationally complete set.
Lemma 5.29. The 2
N1
(2
N
+ 1) 1 non-trivial
questions in
˜
Q
M
N
are pairwise independent.
Proof. The proof is entirely analogous to the
proofs of lemmas
5.3, 5.13, 5.14 and 5.25.
Likewise, the complementarity and compatibil-
ity structure of
˜
Q
M
N
is analogous to the qubit
case.
Lemma 5.30. Q
µ
1
···µ
N
, Q
ν
1
···ν
N
˜
Q
M
N
are
maximally compatible if the index sets
{µ
1
, . . . , µ
N
} and {ν
1
, . . . , ν
N
} differ by an
even number (incl. 0) of non-zero indices,
and
maximally complementary if the index sets
{µ
1
, . . . , µ
N
} and {ν
1
, . . . , ν
N
} differ by an
odd number of non-zero indices.
Proof. Thanks to lemmas
5.8, 5.185.20 and
5.23, the proof of lemma 5.26 also applies to
the rebit case with the sole difference that only
an even number of indices in the questions can
take the value 3 and that correlations of correla-
tions Q
3
a
3
b
cannot be decomposed into individu-
als Q
3
a
, Q
3
b
.
Finally,
˜
Q
M
N
is indeed logically closed.
Theorem 5.31. (Rebits)
˜
Q
M
N
is logically
closed under and is thus an informationally
complete set
˜
Q
M
N
= Q
M
N
with D
N
= 2
N1
(2
N
+
1) 1 for the case D
1
= 2.
Proof. Thanks to theorems 5.9, 5.24 and lemma
5.30, the proof of theorem 5.27 also applies here,
except that only an even number of indices can
take the value 3 and Q
3
a
3
b
cannot be decomposed
into individuals Q
3
a
, Q
3
b
.
We close with the observation that simi-
larly to the N qubit case in section
5.5.1, one
could represent the compatibility and comple-
mentarity structure v ia a simplicial question
graph. Maximally entangled states (of maximal
information) will correspond to O spending
all N available bits over a mutually inde-
pendent set of N N -partite questions which
is maximally complementary to every indi-
vidual question. In fact, the prescription for
constructing a maximally N -partite entangled
qubit state provided at the end of section
5.5.1
also applies to N rebits since only indices with
values 1, 2 were employed. Furthermore, for
rebits one can similarly generate maximally
entangled states of non-maximal information;
e.g., O could ask only the N 1 questions
Q
3
1
3
2
000···
, Q
03
2
3
3
00···
, Q
003
3
3
4
0···
, . . . , Q
0···03
N1
3
N
.
If O has maximal information about each such
question, every rebit pair will be maximally
entangled, while O has still not reached the
information limit (see also section
5.2.6).
6 Information measure, time evolution
and state space
Thus far we have only applied rules
1 and 2 to
derive the basic question structure on Q
N
. We
shall now slightly switch topic and return to the
problems of time evolution, begun in subsection
3.2.8, and of explicitly quantifying the informa-
tion which O has acquired about S by means of
interrogations with questions. The information
measure and time evolution of S’s state in be-
tween interrogations are intertwined through rule
3 of information preservation and rule 4 of max-
imality of time evolution such that we have to
discuss these topics together.
In (
6) we have implicitly defined the total
amount of O’s information about S once the
latter is in a s tate ~y
OS
as
I
OS
(~y
OS
) =
D
N
X
i=1
α
i
(~y
OS
),
Accepted in Quantum 2017-11-27, click title to verify 59
where α
i
quantifies O’s information about the
outcome of question Q
i
Q
M
N
and satisfies the
bounds (1). We shall begin in subsection 6.1 by
clarifying that the information measure we seek
to derive here is conceptually distinct from the
standard Shannon entropy, before imposing el-
ementary consistency conditions on the relation
α
i
(~y
OS
) in subsection
6.2. Subsequently, we use
these conditions and implement rules
3 and 4 to
establish that the set of possible time evolutions
defines a group and, finally, to derive the explicit
functional f orm of I
OS
in subsection
6.8. This
discussion will also unravel properties of state
spaces and lead to the notion of pure states.
6.1 Why the Shannon entropy does not apply
here
Consider a random variable experiment with dis-
tinct outcomes each of which is described by a
particular probability. Outcomes with a smaller
likelihood are more informative (relative to the
prior information) than those with a larger like-
lihood. The Shannon information quantifies the
information an observer gains, on average, via a
specific outcome if she repeated the experiment
many times. (Equivalently, it quantifies the un-
certainty of the observer before an outcome of
the experiment occurs.) The probability distri-
bution over the different outcomes is assumed to
be known. In particular, none of the outcomes in
any run of the experiment will lead to an update
of the probability distribution.
Here, by contrast, we are not interested in how
informative one specific question outcome is rel-
ative to another and how much information O
would gain, on average, with a specific answer if
he repeated the interrogation with a fixed q ues-
tion many times on identically prepared systems.
Instead, we wish to quantify O’s prior informa-
tion (e.g., gained from previous interrogations)
about the possible question outcomes on only
the next system to be interrogated. The role of
the probability distribution over the various out-
comes (to all questions) is here assumed by the
state which O ascrib es to S. This state is not as-
sumed to be ‘known’ absolutely, instead, as dis-
cussed in section
3.2, it is defined only relative
to the observer and represents O’s ‘catalogue of
knowledge’ or ‘degree of belief about S. In par-
ticular, this state is updated after any interro-
gation if the outcome has led to an information
gain. This update takes the form of a ‘collapse’ of
the prior into a posterior state of s single system
in a single shot interrogation, while it yields an
update of the ensemble state in a multiple shot
interrogation (see subsection
3.2.3). The proba-
bility distribution represented by the state is thus
a prior distribution for the next interrogation and
only valid until the next answer unless the lat-
ter yields no information gain. We thus seek to
quantify the information content in the state of
S relative to O. This is not eq uivalent to the sum
of the average information gains, relative to that
same state used as a fixed prior probability dis-
tribution, which are provided by specific answers
to the questions in an informationally complete
set over repeated interrogations.
The reader should thus not be surprised to find
that the end result of the below derivation will
not yield the Shannon entropy. We emphasize,
however, that this clearly does not invalidate the
use of the Shannon entropy (in the more general
form of the von Neumann entropy) in quantum
theory as long as one employs it what it is de-
signed f or: to describe average information gains
in repeated experiments on identically prepared
systems relative to a prior state which is as-
signed before the repeated experiments are car-
ried out and which represents the ‘known’ prob-
ability distribution.
6.2 Elementary conditions on the measure
There are a few natural requirements on the re-
lation between α
i
and ~y
OS
:
(i) The Q
i
Q
M
N
are pairwise independent.
Accordingly, α
i
should not depend on the
‘yes’-probabilities y
j6=i
of other questions
Q
j6=i
such that α
i
= α
i
(y
i
).
(ii) All Q
i
Q
M
N
are informationally of equiv-
alent status. The functional relation be-
tween y
i
and α
i
should be the same for all i:
α
i
= α(y
i
), i = 1, . . . , D
N
.
(iii) If O has no information about the outcome
of Q
i
, i.e. y
i
= 1/2, then α
i
= 0 bit.
(iv) If O has maximal information about the out-
come of Q
i
, i.e. y
i
= 1 or y
i
= 0, then α
i
= 1
bit; both possible answers give 1 bit of in-
formation.
Accepted in Quantum 2017-11-27, click title to verify 60
(v) The assignment of which answer to Q
i
is
‘yes’ and which is ‘no’ is arbitrary and the
functional relation between α
i
and y
i
(or n
i
)
should not depend on this choice. Hence,
α(y
i
) = α(n
i
) must be symmetric around
y
i
= 1/2 (see also (iv)); α
i
quantifies the
amount of information about Q
i
, but does
not encode what the answer to Q
i
is.
(vi) On the interval y
i
(1/2, 1], the relation
between α and y
i
should be monotonically
increasing such that O’s information about
the answer to Q
i
is quantified as higher, the
higher the assigned probability for a ‘yes’
outcome. Likewise, on [0, 1/2), α(y
i
) must
be monotonically decreasing. In particular,
α shall be continuous and strictly convex;
i.e., for all y
1
i
, y
2
i
[0, 1] and every λ [0, 1]
α(λ y
1
i
+ (1 λ) y
2
i
) λ α(y
1
i
) + (1 λ) α(y
2
i
)
and the inequality shall be strict whenever
y
1
i
6= y
2
i
and 0 < λ < 1. H ence, y
i
=
1
2
is the
global minimum of α on [0, 1].
Since now
I
OS
(~y
OS
) =
D
N
X
i=1
α(y
i
) (47)
and each summand is strictly convex, we also
have that I
OS
is a strictly convex function on
Σ
N
. Accordingly, in the coin flip scenario of sub-
section
3.2.8, O’s information about the mixed
state ~y
OS
12
= λ ~y
OS
1
+(1λ) ~y
OS
2
is smaller
than his maximal inf ormation about any of its
constituents ~y
OS
1
, ~y
OS
2
unless the outcome
of the coin flip was certain, or S
1
, S
2
are in the
same state. This is consistent with the fact that
O’s information about the outcome of any ques-
tion (asked to either S
1
or S
2
, depending on the
outcome of the coin flip) is clouded by the random
coin flip outcome.
6.3 The state of no information is an interior
state
In order to discuss the action of time evolution
on the state space Σ
N
, we have to derive a few
further properties of the latter, some of which
depend on the properties of I
OS
.
Firstly, since D
N
is finite for finite N, Σ
N
is finite dimensional in this case. In fact, Σ
N
contains a basis of R
D
N
42
so that it is a D
N
-
dimensional closed convex subset of R
D
N
. Fur-
thermore, since y
i
[0, 1] for all y
i
in ~y
OS
, Σ
N
is clearly bounded and thus, by the Heine-Borel-
theorem, compact. As a compact convex set with
non-empty interior, Σ
N
is thus a convex b ody
and, in particular, has volume [
89].
Next, we focus on the state of no information
~y
OS
=
1
2
~
1. Given strict convexity of I
OS
on
Σ
N
, the conditions of the previous subsection en-
tail that the state of no information is the global
minimum of I
OS
with I
OS
(
1
2
~
1) = 0 bits
in agreement with its uniqueness. It is also an
interior state.
Lemma 6.1. The state of no information lies in
the interior of Σ
N
.
Proof. The Q
i
Q
M
N
are pairwise indepen-
dent so for each such Q
i
there exist two states
~y
y
i
OS
= (
1
2
, · · · ,
1
2
, 1,
1
2
, · · · ,
1
2
) and ~y
n
i
OS
=
(
1
2
, · · · ,
1
2
, 0,
1
2
, · · · ,
1
2
) corresponding to O having
only asked Q
i
to S in the state of no information
and having received the answers Q
i
= ‘yes’ and
Q
i
= ‘no’, respectively. Since Σ
N
is convex, it
also contains all convex mixtures of these special
states. The D
N
~y
y
i
OS
, as well as the D
N
~y
n
i
OS
each defin e a basis of R
D
N
such that the set of all
their convex mixtures define a D
N
-dimensional
convex polytope contained in Σ
N
R
D
N
. In
particular, given that
1
2
~y
y
i
OS
+
1
2
~y
n
i
OS
=
1
2
~
1
yields the state of no information for each i, it
is clear that it lies in the interior of the convex
polytope.
Given that time evolution preserves the total
information by rule
3, we shall also be interested
in the level sets of I
OS
.
Lemma 6.2. Let I
OS
fulfill conditions (i)–(vi)
of sec tion
6.2. For sufficiently small ε > 0, the
level set
L
ε
:= {~y
OS
Σ
N
|I
OS
(~y
OS
) = ε} (48)
lies in the interior of Σ
N
and is homeomorphic
to S
D
N
1
. In this case, L
ε
constitutes the full
boundary of the fat level set L
Iε
:= {~y
OS
Σ
N
|I
OS
(~y
OS
) ε} which contains the state
42
Each of the following D
N
-dimensional vectors
(1,
1
2
,
1
2
, · · · ,
1
2
), (
1
2
, 1,
1
2
, · · · ,
1
2
), · · · , (
1
2
, · · · ,
1
2
, 1) rep-
resents a legal state ~y
OS
, corresponding to O only
knowing the answer to precisely one question from Q
M
N
with certainty and ‘nothing else’.
Accepted in Quantum 2017-11-27, click title to verify 61
of no information and is homeomorphic to the
closed unit ball in R
D
N
.
Proof. The state of no information is the global
minimum of I
OS
with I
OS
(
1
2
~
1) = 0 and,
thanks to lemma
6.1, lies in the interior of Σ
N
.
Furthermore, condition (vi) of subsection
6.2 re-
quires that α(y
i
) increases monotonically away
from y
i
=
1
2
for each i so that I
OS
in (
47) must
likewise increase monotonically in each direction
away from ~y
OS
=
1
2
~
1 in R
D
N
. Owing to the
continuity of I
OS
, we can therefore find a suffi-
ciently small ε > 0 su ch that any state attained
by moving away from the state of no inform ation
in any direction until reaching I
OS
= ε still lies
in the interior of Σ
N
. Accordingly, for such ε > 0,
L
ε
must lie in the interior of Σ
N
and so must L
Iε
with the state of no information in its own inte-
rior. Given that I
OS
: Σ
N
R is continuous
and Σ
N
is closed, I
OS
is a closed convex func-
tion. Closed convex functions have closed convex
fat level sets [
89] so that L
Iε
is a closed convex
subset in the interior of Σ
N
. Clearly, L
Iε
is D
N
-
dimensional since we moved away in all directions
from ~y
OS
=
1
2
~
1 to reach L
ε
and the states with
I
OS
= ε define the limit points of L
Iε
in each
direction so that L
ε
constitutes its f ull bound-
ary. Any D
N
-dimensional closed convex subset
of R
D
N
with non-empty interior is homeomorphic
to th e closed unit ball and its boundary is home-
omorphic to S
D
N
1
.
These results will become important for estab-
lishing that the set of time evolutions is a group.
6.4 Time evolution of the ‘Bloch vector’
In subsection 3.2.8, we have only considered the
time evolution of the redundantly parametrized
2D
N
-dimensional state
~
P
OS
. According to (10),
this time evolution is linear, described by a ma-
trix A(∆t), and only depends on the interval
t = t
2
t
1
, where t
1
, t
2
are arbitrary instants of
time in between O’s interrogations on a given sys-
tem. However, because of (
4), it clearly suffices
to consider the yes-vector ~y
OS
(or no-vector
~n
OS
) alone to describe the state of S relative
to O. Let us therefore determine, how ~y
OS
and
~n
OS
evolve under time evolution. To this end,
we decompose A(∆t) and
~
P
OS
in (
10),
~y
OS
(∆t)
~n
OS
(∆t)
!
=
a(∆t) b(∆t)
c(∆t) d(∆t)
!
~y
OS
(0)
~n
OS
(0)
!
,
where a(∆t), b(∆t), c(∆t), d(∆t) are nonnegative
(real) D
N
× D
N
matrices. Using the normal-
ization (
4) (with p = 1), one finds that both
~y
OS
, ~n
OS
evolve affinely,
~y
OS
(∆t) = [a(∆t) b(∆t)] ~y
OS
(0) + b(∆t)
~
1,
~n
OS
(∆t) = [d(∆t) c(∆t)] ~n
OS
(0) + c(∆t)
~
1. (49)
We shall now determine relations among the four
matrices a, b, c, d. Firstly, the normalization (
4)
must hold for all initial states and all times.
Hence, for all t R,
~
1 = ~y
OS
(∆t) + ~n
OS
(∆t)
= [a(∆t) b(∆t)] ~y
OS
(0) + [d(∆t) c(∆t)] ~n
OS
(0) + [b(∆t) + c(∆t)]
~
1
=
(
4)
[a(∆t) b(∆t) + c(∆t) d(∆t)] ~y
OS
(0) + [b(∆t) + d(∆t)]
~
1.
Since the set of all possible initial states ~y
OS
(0)
contains a basis of R
D
N
, we must conclude
a(∆t) b(∆t) = d(∆t) c(∆t),
[b(∆t) + d(∆t)]
~
1 =
~
1. (50)
Next, we note that, by assumption
3, the state of
no information is unique and, by rule
3, clearly
preserved under time evolution. Inserting the
state of no inf ormation, e.g., as ~n
OS
(∆t) =
1
2
~
1
Accepted in Quantum 2017-11-27, click title to verify 62
into (49), this entails
[c(∆t) + d(∆t)]
~
1 =
~
1. (51)
In conjunction, (50, 51) therefore imply
43
~y
OS
(∆t) = [a(∆t) b(∆t)] ~y
OS
(0) + b(∆t)
~
1,
~n
OS
(∆t) = [a(∆t) b(∆t)] ~n
OS
(0) + b(∆t)
~
1.
Consequently, defining the D
N
× D
N
time evolu-
tion matrix
T (∆t) := a(∆t) b(∆t), (52)
yields a linear time evolution of what we shall
henceforth call the generalized Bloch v ector
2 ~y
OS
~
1:
2 ~y
OS
(∆t)
~
1 = ~y
OS
(∆t) ~n
OS
(∆t)
= T (∆t)
2 ~y
OS
(0)
~
1
. (53)
Notice that T (∆t) need neither be nonnegative
nor stochastic in any pair of its components (in
contrast to A(∆t)). Henceforth, we shall only
consider the T (∆t) governing the time evolution
of the Bloch vector
~r
OS
:= 2 ~y
OS
~
1 (54)
and often employ the latter to parametrize states.
6.5 Time evolution is injective
We continue by demonstrating that time evolu-
tion of S’s states must be injective on Σ
N
. In
subsection
6.6, we will use this result to prove
that time evolution is also surjective on Σ
N
and
thus, in f act, reversible.
Lemma 6.3. Let T (∆t) : Σ
N
Σ
N
be a time
evolution as given in (
52) for any t R. If
I
OS
is strictly convex, rule
3 entails that T (∆t)
is injective.
Proof. Assume T (∆t) was not injective. T hen
there would exist states ~y
OS
(t
1
) 6= ~y
′′
OS
(t
1
)
such that
~y
OS
(t
2
) := T (∆t)~y
OS
(t
1
)
= T (∆t)~y
′′
OS
(t
1
). (55)
Now consider the coin flip scenario of section
3.2.8. O can prep are S
1
in th e state ~y
OS
1
(t
1
) :=
~y
OS
(t
1
) an d S
2
in the state ~y
OS
2
(t
1
) :=
~y
′′
OS
(t
1
) at time t
1
before tossing the coin. By
(
8),
~y
OS
12
(t
1
) = λ ~y
OS
1
(t
1
) + (1 λ) ~y
OS
2
(t
1
),
and, on account of (55) and rule 3,
I
OS
1
(~y
OS
1
(t
1
)) = I
OS
2
(~y
OS
2
(t
1
))
= I
OS
1
(~y
OS
1
(t
2
))
= I
OS
2
(~y
OS
2
(t
2
)). (56)
Thus, using (56) and s trict convexity of I
OS
,
this yields for a coin flip with 0 < λ < 1
I
OS
12
(~y
OS
12
(t
1
)) < I
OS
1
(~y
OS
1
(t
2
)) = I
OS
2
(~y
OS
2
(t
2
)). (57)
On the other hand, (55) implies that, at time t
2
,
O would nd
~y
OS
12
(t
2
) = ~y
OS
1
(t
2
) = ~y
OS
2
(t
2
)
43
As an aside, note that A(∆t) can thus b e replaced by
A(∆t) =
a(∆t) b(∆t)
b(∆t) a(∆t)
,
so th at time evolution is symmetric under a swap of the
‘yes/no’-labeling for all Q Q
M
N
at once, i.e. under
~y
OS
~n
OS
.
such that
I
OS
12
(~y
OS
12
(t
2
)) = I
OS
1
(~y
OS
1
(t
2
))
= I
OS
2
(~y
OS
2
(t
2
)). (58)
Hence, (
57, 58) entail that O’s information about
S
12
has increased between t
1
and t
2
, despite not
having tossed the coin and asked any qu estions.
This is in contradiction with rule
3. We conclude
that T (∆t) must be injective.
Since Σ
N
contains a basis of R
D
N
(see subsec-
tion
6.3) it is clear that T (∆t) is also injective
on R
D
N
. For finite dimensional square matrices,
T (∆t) being injective is equivalent to it being
Accepted in Quantum 2017-11-27, click title to verify 63
bijective on R
D
N
. H ence, to every T (∆t) there
must exist an inverse matrix T
1
(∆t) such that
T (∆t)T
1
(∆t) = T
1
(∆t)T (∆t) = 1.
However, is this T
1
(∆t) also a legal time evo-
lution?
44
This would be the case if T (∆t) was
also surjective and thereby reversible on Σ
N
R
D
N
. Indeed, we shall demonstrate this property
next.
6.6 Time evolution is also reversible
To establish reversibility of time evolution we
shall resort to tools from topology and convex
sets.
Lemma 6.4. Restricting any time evolution
T (∆t) : Σ
N
Σ
N
to an interior level set L
ε
(
48) defines a continuous, open and closed map
from L
ε
to itself.
Proof. As an injective square matrix T (∆t) de-
fines a continuous invertible map from R
D
N
to it-
self and so the pre-image of any open s et in R
D
N
is another open set. Rule
3 implies [T (∆t)](L
ε
)
L
ε
and that no state from outside L
ε
can evolve
into L
ε
. The open sets of L
ε
in the induced topol-
ogy are the intersections of L
ε
with open sets of
R
D
N
. Consider any such open set V L
ε
and
any open set U R
D
N
such that V = U L
ε
.
The pre-image [T
1
(∆t)](U) is again open as is
therefore [T
1
(∆t)](U)L
ε
which, thanks to rule
3, must be non-empty if V is non-empty. Hence,
T (∆t) defi nes a continuous map from L
ε
to it-
self. For similar reasons, this map is also open.
Finally, the closed map lemma states th at a con-
tinuous map from a compact to a Hausdorff space
is also closed. Thanks to lemma
6.2, we have
L
ε
S
D
N
1
which is both compact and Haus-
dorff and so the map from L
ε
to itself defined by
T (∆t) is also closed.
Next, we employ this result to show that time
evolution is surjective on interior level sets.
Lemma 6.5. Any allowed time evolution
is surjective on the interior level sets, i.e.
[T (∆t)](L
ε
) = L
ε
and thus also [T (∆t)](L
Iε
) =
L
Iε
.
Proof. Lemma
6.4 entails that T (∆t) defines a
continuous, open and closed map from L
ε
to it-
self. It therefore maps all open sets to open sets
and all closed sets to closed sets in L
ε
. In par-
ticular, it must map all sets that are both open
and closed in L
ε
to other sets which are both
open and closed. Since L
ε
S
D
N
1
is con-
nected, the empty set and the full L
ε
are the
only sets in L
ε
which are both closed and open.
Now, by lemma
6.3, T (∆t) is inj ective so that
[T (∆t)](L
ε
) = L
ε
.
The last result is sufficient to finally establish
that time evolution is surjective also on the state
space.
Theorem 6.6. Any allowed time evolution is
surjective on Σ
N
, i.e. [T (∆t)](Σ
N
) = Σ
N
, and
has | det T (∆t)| = 1.
Proof. Given that an interior fat level set L
Iε
is, by lemma
6.2, homeomorphic to the closed
unit ball in R
D
N
it is a convex body and has
volum e [
89]. Lemma 6.5 yields [T (∆t)](L
Iε
) =
L
Iε
and so its volume is left invariant by
T (∆t) which, according to (53), acts linearly on
states ~r
OS
in the Bloch vector parametriza-
tion. Theorem 6.2.14 in [89] then implies that
| det T (∆t)| = 1 and therefore also that th e vol-
ume of [T (∆t)](Σ
N
) is equal to the volume of Σ
N
.
Consider now the volume difference
δ
[T (∆t)](Σ
N
), Σ
N
:= vol
[T(∆t)](Σ
N
) Σ
N
vol
[T(∆t)](Σ
N
) Σ
N
.
Clearly, [T (∆t)](Σ
N
) Σ
N
, and
so δ
[T (∆t)](Σ
N
), Σ
N
= vol
N
)
vol
[T(∆t)](Σ
N
)
= 0. Exercise 6.2.2 in [89]
then sh ows that δ
[T (∆t)](Σ
N
), Σ
N
= 0 is only
possible if [T (∆t)](Σ
N
) = Σ
N
.
Hence, T
1
(∆t) maps all legal states to legal
states and must, by rule
4, be legal also.
Corollary 6.7. Any time evolution permitted by
the rules is reversible.
44
I thank a referee for pointing out that addressing this
question was overlooked in a previous draft version.
Accepted in Quantum 2017-11-27, click title to verify 64
6.7 Time evolution defines a group
Given that any time interval can be decomposed into two time intervals, t = t
1
+ t
2
, and the
evolution of ~r
OS
= 2 ~y
OS
~
1 is continuous by rule
4, O must find
T (∆t
2
) T (∆t
1
) ~r
OS
(0) = T (∆t
2
) ~r
OS
(∆t
1
) = ~r
OS
(∆t) = T (∆t) ~r
OS
(0),
and thus that multiplication is abelian,
T (∆t
1
+ t
2
) = T (∆t
2
) T (∆t
1
) = T (∆t
1
) T (∆t
2
).
(The last equality follows from time translation
invariance.) From the last equation and T (0) =
1, using t t = 0, we can also infer that
T
1
(∆t) = T (t). If we permit O to consider
the time evolution of S for any duration t R,
it f ollows that the product of any two time evolu-
tion matrices is again a time evolution matrix. In
summary, we therefore gather that a given time
evolution, as perceived by O and under the cir-
cumstances to which he has subjected S, is such
that:
(i) T (0) = 1,
(ii) to every T (∆t) there exists an inverse
T
1
(∆t) = T (t),
(iii) the multiplication of any two time evolution
matrices is again a time evolution matrix,
and
(iv) matrix multiplication is obviously associa-
tive.
In conclusion, under the assumptions of section
3.2 and rules 3 and 4, a given time evolution de-
fines therefore an abelian, one-parameter matrix
group. Hence, a given time evolution is described
by a single evolution generator and single param-
eter t parametrizing the duration. We note,
however, that a multiplicity of time evolutions of
S is possible, depending on the physical circum-
stances (interactions) to which O may subject S.
Different time evolutions will be generated by dif-
ferent generators, but each time evolution will
form a one-parameter group as discussed above.
In fact, the full set of time evolutions T
N
which
O is able to implement must likewise be a group.
Namely, if T (∆t), T
(∆t
) correspond to two dis-
tinct interactions, then rule
4 implies that also
T (∆t) · T
(∆t
) is a legal time evolution for any
state and since both T, T
are invertible their full
set must be a group. This implies, in particular,
that T
N
is a group. We shall return to this further
below and in [
1].
6.8 The squared length of the Bloch vector as
information measure
We now have sufficient structure in our hand
to determine the functional relation between α
i
and y
i
. Given that the generalized Bloch vector
2 ~y
OS
~
1 transforms nicely under time evolution
(
53), it is useful to parametrize α
i
by 2y
i
1, i.e.
α
i
= α(2y
i
1). Rule
3 entails that O’s total in-
formation about (an otherwise non-interacting) S
is a ‘conserved charge’ of time evolution
I
OS
(~y
OS
(∆t)) = I
OS
(~y
OS
(0))
which translates into the condition
I
OS
T (∆t)
2 ~y
OS
(0)
~
1

=
D
N
X
i=1
α
D
N
X
j= 1
T
ij
(∆t) (2y
j
(0) 1)
=
D
N
X
i=1
α (2y
i
(0) 1) = I
OS
(2 ~y
OS
(0)
~
1). (59)
If T (∆t) was a permutation matrix, (59) would
hold for any function α(2y
i
1). For exam-
ple, N class ical bits are governed by the evolu-
tion group Z
2
× · · · × Z
2
. However, permutations
Accepted in Quantum 2017-11-27, click title to verify 65
form a discrete group, while in our present case
{T (∆t), t R} constitutes a continuous one-
parameter group. This is where continuity of time
evolution, as asserted by rule
4, becomes crucial.
Under a reasonable assumption on the informa-
tion measure, we shall now show that continuity
of time evolution, together with conditions (i)–
(vi) of subsection
6.2, enforces the quadratic re-
lation α
i
= (2y
i
1)
2
. To this end, we once more
invoke the coin flip scenario.
45
Given our parametrization in terms of the
Bloch vector 2 ~y
OS
~
1, O’s information about
the outcomes of his questions in the coin flip sce-
nario can be written as follows
I
OS
12
2 (λ ~y
OS
1
+ (1 λ) ~y
OS
2
)
~
1
= I
OS
12
λ (2 ~y
OS
1
~
1) + (1 λ)(2 ~y
OS
2
~
1)
=
D
N
X
i=1
α
λ (2y
1
i
1) + (1 λ)(2y
2
i
1)
.
It is instructive to consider the case in which O
is entirely oblivious about S
2
such that the lat-
ter is in the state of no information ~y
OS
2
=
1
2
~
1
relative to him, but that O has some informa-
tion about S
1
. In this case, 2 ~y
OS
12
~
1 =
λ(2 ~y
OS
1
~
1) and (assuming the outcome of the
coin flip is not certain) strict convexity of I
OS
(see subsection
6.2) implies
I
OS
12
(λ (2 ~y
OS
1
~
1)) < λ I
OS
1
(2 ~y
OS
1
~
1)
or, equivalently,
I
OS
12
(λ (2 ~y
OS
1
~
1)) (60)
= f · I
OS
1
(2 ~y
OS
1
~
1),
where f < 1 is a factor parametrizing O’s in-
formation loss relative to the case in which he
does not toss a coin and, instead, directly asks
S
1
.
46
The reason O experiences such a relative
information loss about the outcome of his inter-
rogation is, of course, entirely due to the random-
ness of the coin flip. But the coin flip is indepen-
dent of the s ystems S
1,2
and, in particular, of the
states in which these are relative to O; the fac-
tor λ by which the probabilities ~y
OS
1
become
45
We suspect that this result may be derivable from
purely group theoretic arguments without an operational
setup by employing the mathematical fact that to every
continuous matrix group acting linearly on some space
there corresponds a conserved inner produ ct which is
quadratic in the components of the vectors.
46
In fact, one can equivalently interpret the situation as
follows: O only considers a system S
1
which would be in
the state ~y
OS
1
if it was present with certainty. However,
λ in the state λ(2 ~y
OS
1
~
1) represents the probability
that S
1
‘is there’ at all. Indeed, in this case, (
4) can be
written as λ(~y
OS
1
+~n
OS
1
) = λ·
~
1 such that λ(~y
OS
1
~n
OS
1
) = λ(2 ~y
OS
1
~
1).
rescaled is state independent. For that reason,
the relative information loss should likewise de-
pend only on the coin flip, quantified by λ, and
not on the state ~y
OS
1
. For instance, if we also
considered the case that the coin flip was certain,
i.e. λ = 0, 1, then clearly for λ = 1 we must have
f = 1 ~y
OS
1
Σ
N
and for λ = 0 it must hold
f = 0 ~y
OS
1
Σ
N
. We shall make this into
a requirement on the information measure f or all
values of λ:
Requirement 1. The relative information loss
factor f in (
61) is a state independent (contin-
uous) function of the coin flip probability λ with
f(λ) < 1 for λ (0, 1).
The binary Shannon entropy H(y) =
y log y (1 y) log(1 y) in the role of α would
fail this requirement. Namely, the measure α thus
factorizes, α(λ (2y
i
1)) = f (λ) α(2y
i
1), while
H(y) would not. Setting λ = λ
1
· λ
2
yields
47
f(λ
1
· λ
2
) α(2 y
i
1) = α(λ
1
· λ
2
(2 y
i
1))
= f (λ
1
) α(λ
2
(2 y
i
1))
= f (λ
1
)f(λ
2
) α(2 y
i
1)
and therefore f(λ
1
) · f(λ
2
) = f (λ
1
· λ
2
), which
implies f(λ) = λ
p
, for some power p R. But
then, α must be a homogeneous function α(2 y
i
47
Such a factorization of coin flip probabilities could be
achieved, e.g., if O decided to use one coin, with ‘heads’
probability λ
1
, to firstly decide which of two possible con-
vex mixtures to prepare where both possible mix tures are
generated with a second coin with ‘heads’ probability λ
2
.
If t hree of the four states within the two mixtures are
chosen as th e state of no information, one would obtain
precisely such an equation.
Accepted in Quantum 2017-11-27, click title to verify 66
1) = k (2 y
i
1)
p
with some constant k R. In
consequence, the information I
OS
(2 ~y
OS
~
1) is
(up to k) the p-norm of the Blo ch vector 2 ~y
OS
~
1.
We can rule out that p (−∞, 0] because in
this case, as one can easily check, it is impossi-
ble to s atisfy all the consistency conditions (i)–
(vi) of subsection
6.2. Hence, p > 0. At this
stage we can make us e of (
59) and a result by
Aaronson [
90] which implies that the only vector
p-norm with p > 0 which is preserved by a contin-
uous matrix group is the 2-norm. Since any given
time evolution of the Bloch vector 2 ~y
OS
~
1 is
governed by a continuous, one-parameter matrix
group, we conclude that α(2 y
i
1) = k (2 y
i
1)
2
.
Imposing condition (iv) yields k = 1 and there-
fore ultimately
I
OS
(~y
OS
) =
D
N
X
i=1
(2 y
i
1)
2
. (61)
It is straightforward to convince oneself that all
of (i)–(vi) of subsection
6.2 are satisfied by this
quadratic information measure. O’s total amount
of information about S is thus the sq uared length
of the generalized Bloch vector, thereby assuming
a geometric flavour. It is important to emphasize
that, had we not imposed continuity of time evo-
lution in rule
4, we would not have been able to
arrive at (
61); if time evolution was not continu-
ous, many solutions to α in terms of y
i
would be
possible.
The quadratic information measure (
61) has
been proposed earlier by Brukner and Zeilinger
in [32,34,91,92] from a different perspective, em-
phasizing that this is the most natural measure
taking into account an observer’s uncertainty
due to statistical fluctuations about the out-
come of the next trial of measurements on a
system in a multiple shot experiment. Further-
more, taking the formalism of quantum theory
as given, Brukner and Zeilinger [
73] later singled
out the quadratic measure from the set of Tsal-
lis entropies by imp osing an ‘information invari-
ance principle’, according to which a continuous
transformation among any two complete sets of
mutually complementary measurements in quan-
tum theory should leave an observer’s informa-
tion about the system invariant. While [
73] is
certainly compatible with the present framework,
here we come from farther away to the same re-
sult: we do not pre-suppose quantum theory and
derive the quadratic measure more generally by
starting from the landscape of inf ormation infer-
ence theories and imposing rule
3 of information
preservation and rule
4 of maximality of time evo-
lution thereon. In this regard, the present deriva-
tion may similarly be taken as a strong justifica-
tion for the original Brukner-Zeilinger proposal.
6.9 The set of all time evolutions is a subgroup
of SO(D
N
)
Thus far, we only gathered that a given time evo-
lution is described by a one-parameter group and,
by theorem
6.6, must have | det T (∆t)| = 1. Now,
given (
61), we are in the position to say q uite a bit
more. The full matrix group leaving (
61) invari-
ant is O(D
N
). However, a given time evolution
is a continuous one-parameter group and must
therefore be connected to the identity. Hence,
T (∆t) SO(D
N
), T (∆t). Consequently, the
group corresponding to any fixed time evolution
is a one-parameter subgroup of SO(D
N
). We
also noted before that the set of all possible time
evolutions T
N
is a group thanks to rule
4. It
must therefore likewise satisfy T
N
SO(D
N
).
This topic is thoroughly discussed in the com-
panion articles, where it is shown that the rules
imply T
N
= PSU(2
N
) for the D
1
= 3 [
1] and
T
N
= PSO(2
N
) for the D
1
= 2 case [
2]. Note
that PS U(2
N
) is a prop er subgroup of SO(D
N
=
4
N
1) for N > 1. The generators of T
N
are the
set of possible time evolution generators of S’s
states.
6.10 Pure and mixed states
The explicit quantification of O’s information
now permits us to render the distinction between
three informational classes of S’s states which
we already loosely referred to as ‘states of max-
imal knowledge’ or ‘states of non-maximal infor-
mation’ in previous sections precise.
Firstly, we determine the maximally attain-
able (independent and dependent) information
content within a state ~y
OS
of a sy stem of N
gbits. This can be easily counted: once O knows
the answers to N mutually independent questions
(these do not need to be individuals), he will also
know the answers to all their bipartite, tripar-
tite,... and N-partite correlation questions all
of which are contained in Q
M
N
too by theorems
Accepted in Quantum 2017-11-27, click title to verify 67
5.27 and 5.31. But these are then
N
1
!
+
N
2
!
+ · · ·
N
N
!
=
N
X
i=1
N
i
!
= 2
N
1
answered questions from Q
M
N
, while all remain-
ing q uestions in Q
M
N
will be maximally comple-
mentary to at least one of the known ones. O’s to-
tal information, as quantified by I
OS
(
47), thus
contains plenty of dependent bits of information
a result of the fact that the questions in Q
M
N
are pairwise but not necessarily mutually inde-
pendent.
Using this observation, we shall characterize
S’s states according to their information content,
i.e. squared length of the Bloch vector. By rules
3
and 4, this distinction applies to all states which
are connected via some time evolution to the
states above, including those for which the infor-
mation may be distributed partially over many
elements of a fixed Q
M
N
. Specifically, we shall
refer to a state ~y
OS
as a
pure sta te: if it is a state of maximal information content, i.e. maximal length
I
OS
(~y
OS
) =
D
N
X
i=1
(2 y
i
1)
2
= (2
N
1) bits,
mixed state: if it is a state of non-extremal information content, i.e. non-extremal length
0 bit < I
OS
(~y
OS
) =
D
N
X
i=1
(2 y
i
1)
2
< (2
N
1) bits,
totally mixed state: if it is the state of no information ~y
OS
=
1
2
~
1 with zero length
I
OS
~y
OS
=
1
2
~
1
= 0 bit.
We note that these characterizations of states in
terms of their length are indeed true in quantum
theory. In particular, N qubit pure states actu-
ally have a Bloch vector sq uared length equal to
2
N
1. Our reconstruction gives this peculiar
fact a clear informational interpretation.
One may wonder whether the above definition
of a pure state is directly equivalent to being ex-
tremal in Σ
N
and thereby to the usual defini-
tion of pure states in GPTs or quantum theory.
This is not obvious from what we have estab-
lished thus far. However, since clearly we as-
sume I
OS
= (2
N
1) bits to be the maxi-
mally attainable information, we can already con-
clude that pure states so defined must lie on the
boundary of Σ
N
because a non-constant convex
function cannot have its maximum in the inte-
rior of its convex domain [
89]. Showing that the
pure states, as defined above, of the theory sur-
viving the impos ition of all rules quantum the-
ory are, indeed, the extremal states, requires
more work. For N = 1 we shall demonstrate this
shortly, while the discussion for N > 1 is deferred
to [
1].
7 The N = 1 case and the Bloch ball
Before closing this toolkit for now, we quickly give
a flavor of the capabilities of the newly developed
concepts and tools by applying them to show that
in the simplest case of a single gbit (N = 1) rules
14 indeed only have two solutions within L
gbit
,
namely the qubit and the rebit state space includ-
ing their respective time evolution groups. This
proves the claim of section
4.1 for N = 1.
To this end, recall theorem
5.6 which asserts
that the dimension of the N = 1 state space Σ
1
is either D
1
= 2 or D
1
= 3 which thus far we
suggestively referred to as the ‘rebit’ and ‘qubit
case’, respectively.
Accepted in Quantum 2017-11-27, click title to verify 68
7.1 A single qubit and the Bloch ball
We begin with the D
1
= 3 case. Σ
1
will be parametrized by a three-dimensional vector
~y
OS
=
y
1
y
2
y
2
,
where y
1
, y
2
, y
3
are the ‘yes’-probabilities of a mutually maximally complementary question set
Q
1
, Q
2
, Q
3
constituting an informationally complete Q
1
(
11).
The informational distinction of states introduced in section
6.10 reads in this case
pure sta tes: ~y
OS
such that
I
OS
= (2 y
1
1)
2
+ (2 y
2
1)
2
+ (2 y
3
1)
2
= 1 bit,
mixed states: ~y
OS
such that
0 bit < I
OS
= (2 y
1
1)
2
+ (2 y
2
1)
2
+ (2 y
3
1)
2
< 1 bit,
totally mixed state: ~y
OS
=
1
2
~
1 such that
I
OS
= (2 y
1
1)
2
+ (2 y
2
1)
2
+ (2 y
3
1)
2
= 0 bit.
Recall from section 6.9 that the set of possible
time evolutions T
1
must be contained in SO(3)
and that the time evolution rule 4 requires the
set of states into which any state ~y
OS
can evolve
to be maximal, while being compatible with the
rules. Thus, in particular, the image of the cer-
tainly legal pure state (1, 0, 0) (rules
1 and 2 im-
ply its existence in Σ
1
) under T
1
must be maxi-
mal. Applying all of SO(3) to this state generates
all states with |2~y
~
1|
2
= 1 bit and these cer-
tainly abide by the rules. Consequently, the set
of all possible time evolutions, compatible with
rules
14, is the rotation group
T
1
SO(3) PSU(2).
This is the component of the isometry group of
the Bloch ball which is connected to the identity
and it is required in full in order to maximize the
number of states into which (1, 0, 0) can evolve.
However, since time evolution is state indep en-
dent (c.f. assumption
8), this is the time evolu-
tion group f or all states. PSU(2) is precisely the
adjoint action of SU(2) on density matrices ρ
2×2
over C
2
, ρ
2×2
7→ U ρ
2×2
U
, U SU(2) and thus
coincides with the set of all possible unitary time
evolutions of a single qubit in standard quantum
theory.
The set of allowed states populates the entire
unit ball in the three-dimensional Bloch ball, i.e.
Σ
1
B
3
. This follows from the fact that apply-
ing the full group T
1
= SO(3) to (1, 0, 0) gener-
ates the entire Bloch sphere as a closed set of ex-
tremal states and the fact that Σ
1
must be closed
convex according to assumption
2. Hence, we
recover the well-known three-dimensional Bloch
ball state space of a single qubit of standard quan-
tum theory with the set of all pure states defining
the boundary sphere S
2
, the totally mixed state
(as the state of no information) constituting the
center and the set of mixed states filling the inte-
rior in between, as illustrated in figure
6a. This
is precisely the geometry of the set of all normal-
ized density matrices on C
2
. Notice that the pure
state space S
2
CP
1
indeed coincides with the
set of all unit vectors in C
2
(modulo phase).
Given the complete symmetry of the Blo ch ball
as the state space for N = 1, there should not ex-
ist a distinguished informationally complete ques-
tion set Q
M
1
, corresponding to a distinguished or-
thonormal Bloch vector basis, by means of which
O can interrogate S. While it is an additional
assumption, it is thus natural to stipulate that
Accepted in Quantum 2017-11-27, click title to verify 69
PSfrag rep lacements
totally mixed state
mixed states
pure states
~r
~r = 2 ~y
OS
~
1
(a) 3D qubit Bloch ball
PSfrag rep lacements
totally mixed state
mixed states
pure states
~r
~r = 2 ~y
OS
~
1
(b) 2D rebit Bloch disc
Figure 6: The three-dimensional Blo ch ball (a) and the
two- dimensional Bloch disc (b) are the correct state
spaces Σ
1
of a single qubit in standard quantum theory
and a single rebit in real quantum theory, respectively.
The vector ~r parametrizing the state s is the Bloch vector
2 ~y
OS
~
1.
to e very Q Q
1
there exists a unique pure state
in Σ
1
which represents the truth value Q = ‘yes’
and, conversely, that every pure state of this sys-
tem corresponds to the definite answer to one
question in Q
1
. But then Q
1
S
2
which also
coincides with the set of all possible projective
measurements onto the +1 (or, equivalently, the
1) eigenspaces of the Pauli operators ~n · ~σ over
C
2
which is parametrized by ~n R
3
, |~n|
2
= 1,
and where ~σ = (σ
x
, σ
y
, σ
z
) are the usual Pauli
matrices. This set of permissible questions Q
N
for all N will be discussed more thoroughly in [
1]
together with a derivation of the Born rule for
projective measurements.
The ‘ballness’ and three-dimensionality of the
state space of a single qubit can also b e derived
from various operational axioms within G PTs
[
15, 16, 18, 21, 23, 59, 93] and constitutes a cru-
cial step in most GPT based reconstructions of
quantum theory [1416, 20, 21]. The principle of
continuous reversibility, according to which every
pure state of the convex set can be mapped into
any other by means of a continuous and reversible
transformation, usually assumes a crucial role in
such derivations. Here we offer a novel perspec-
tive on the origin of the Bloch ball by deriving it
from elementary rules for the informational rela-
tion between O and S; in particular, we recover
continuous reversibility as a by-product.
7.2 A single rebit and the Bloch disc
The analogous result holds for the D
1
= 2 case:
Σ
1
will be parametrized by a two-dimensional
vector
~y
OS
=
y
1
y
2
!
,
where y
1
, y
2
are the ‘yes’-probabilities of two
maximally complementary questions Q
1
, Q
2
con-
stituting an informationally complete Q
1
(
11).
We then have
pure sta tes: ~y
OS
such that
I
OS
= (2 y
1
1)
2
+ (2 y
2
1)
2
= 1 bit,
mixed states: ~y
OS
such that
0 bit < I
OS
= (2y
1
1)
2
+ (2y
2
1)
2
< 1 bit,
and the
totally mixed state: ~y
OS
=
1
2
~
1 such that
I
OS
= (2 y
1
1)
2
+ (2 y
2
1)
2
= 0 bit.
In analogy to the qubit case, the time evolution
rule
4 implies that
(a) the set of all possible time evolutions is the
(projective) rotation group
T
1
SO(2)
#
PSO(2) SO(2)/Z
2
because SO(2) is the full (connected com-
ponent of the orientation preserving) isom-
etry group of the Bloch disc and all time
Accepted in Quantum 2017-11-27, click title to verify 70
evolutions are permitted which preserve the
Bloch vector length, representing O’s total
information about S, and the orientation of
the Bloch vector basis. Orientation preser-
vation means that O’s convention about the
‘yes’-‘no’-labelling of the question outcomes
is preserved: The isomorphism denoted with
a # requires some explaining because SO(2)
is actually a double cover of PSO(2) as indi-
cated on the right. Indeed, the real density
matrix of a single rebit, ρ
2×2
=
1
2
(1 +r
x
σ
x
+
r
z
σ
z
) on R
2
where r
i
= (2y
i
1), evolves
under the adjoint action of SO(2), ρ
2×2
7→
O ρ
2×2
O
T
for O = e
y
t
SO(2). This is
equivalent to an action of PS O(2) because
the non-trivial center element 1 SO(2)
acts trivially on ρ
2×2
and thus factors out.
However, as a subcase of the single qubit
case this adjoint action of SO(2) on ρ
2×2
is
also equivalent to O
~r for O
SO(2) in the
Bloch vector representation.
(b) the s tate space Σ
1
coincides with the two-
dimensional Bloch disc, as depicted in figure
6b, with the totally mixed state in the center,
the pure states on the boundary circle S
1
and
the mixed states in the interior in between.
This is precisely the geometry of the set of
unit trace, positive-semidefinite, symmetric
matrices over R
2
the space of density ma-
trices of a single rebit. Similarly, the pure
state space S
1
RP
1
coincides with the set
of all unit vectors in the Hilbert space R
2
modulo Z
2
.
Again, given the s ymmetry of the Bloch disc,
there is no reason for a distinguished informa-
tionally complete question basis Q
M
1
on Q
1
, in-
carnated as a distinguished Bloch vector basis,
to exist. Accordingly, we assume that any pure
state on S
1
corresponds to the definite answer
of some question in Q
1
which O may ask the
rebit. In this case, Q
1
S
1
, which coincides
with the s et of projective measurements onto the
+1 eigenspaces of the Pauli operators ~n · ~σ over
R
2
which are parametrized by ~n R
2
, |~n|
2
= 1
and where ~σ = (σ
x
, σ
z
) are the two real and sym-
metric Pauli matrices.
8 Discussion and outlook
Based on the premise to only speak about in-
formation which an observer (or more generally
system) has access to by interaction with other
systems, we have developed a novel framework
for characterizing and (re-)constructing the quan-
tum theory of qubit sys tems. We have laid down,
without ontological statements, the mathemati-
cal and conceptual foundations for a landscape
of theories describing how an observer acquires
information about physical systems through in-
terrogation with yes-no-questions.
Within this landscape four elementary rules
have been given which constrain the observer’s
acquisition of information f or qubit (and rebit)
systems. The rule of limited information and the
complementarity rule imply an independence and
compatibility structure of the binary questions
which, in fact, reproduces that of projective mea-
surements onto the +1 eigenspaces of Pauli op-
erators in quantum theory. In particular, these
rules entail in a constructive and simple way
1. a novel argument for the dimensionality of
the Bloch ball,
2. a new method for determining the
correct number of independent ques-
tions/measurements necessary to describe a
system of N qubits (or rebits),
3. a natural explanation for entanglement,
monogamy of entanglement and quantum
non-locality,
4. the explicit correlation s tructure of two
qubits and rebits, and
5. more generally the correlation structure for
arbitrarily many qubits and rebits.
Furthermore, the rules of information preserva-
tion and maximality of time evolution are shown
to result
6. in a reversible time evolution, and
7. under elementary consistency conditions, in
a quadratic information measure, quantify-
ing the observer’s prior information about
the answers to the various q uestions he may
ask the system.
This measure has been earlier proposed by
Brukner and Zeilinger from a different perspec-
tive [
32,34,73,91,92], complementing our present
derivation.
Combining these results, we then show, as the
simplest example, how the rules entail that
Accepted in Quantum 2017-11-27, click title to verify 71
8. the Bloch ball and Bloch disc are recovered
as state spaces for a single qubit and rebit,
respectively, together with the correct time
evolution groups and question sets.
The full reconstruction of qubit quantum the-
ory, following from the present four rules (and two
additional ones), is completed in the companion
paper [
1] (the rebit case which violates one of the
additional rules is considered separately [
2]). In
conjunction, this derivation highlights the partial
interpretation of quantum theory as a law book,
describing and governing an observer’s acquisi-
tion of information about physical systems. In
particular, it highlights the quantum state as the
observer’s ‘catalogue of knowledge’.
Certainly, there are some limitations to the
present approach: First of all, we employ a clear
distinction between observer and system which
cannot be considered as fundamental. Secondly,
the construction is specifically engineered for the
quantum theory of qubit systems. While arbi-
trary finite dimensional quantum systems could,
in principle, be immediately encompassed by im-
posing the so-called subspace axiom of GPTs
[
14, 16], the latter does not naturally fit into our
set of principles, mostly because rules
1 and 2
quantify the information content of a system in
terms of N N bits which is only suitable for
comp os ite gbit systems. More generally, the lim-
itation is that the current approach only encom-
passes finite dimensional quantum theory, but not
quantum mechanics. As it stands, the mechani-
cal phase space language does not naturally fit
into the present framework and more sophisti-
cated tools are required in order to cover mechan-
ical systems, let alone anything beyond that.
While this informational construction recovers
the state spaces, the set of possible time evolu-
tions and projective measurements of qubit quan-
tum theory, in other words, the architecture of
the theory, it clearly does not tell us much about
the concrete physics (other than demanding it to
fit into this general construction). This purely
informational framework is rather universal and
information carrier independent. But qubits as
information carriers can be physically incarnated
in many different ways: as electron or muon spins,
photon polarization, quantum dots, etc. The
framework cannot distinguish among the differ-
ent physical incarnations, underlining the obser-
vation that not everything can be reduced to in-
formation and additional inputs are necessary in
order to do proper physics.
Nevertheless, despite its current limitations,
this informational approach teaches us something
non-trivial about the structure and physical con-
tent of q uantum theory.
While this work is more generally motivated by
the effort to understand physics from an infor-
mational and operational perspective and high-
lights a partial interpretation of quantum the-
ory, it clearly does not single out ‘the right one’
(see also the discussion in [
94]). For example, by
speaking exclusively about the information acces-
sible to an observer, we are by default silent on
the f ate of hidden variables and on whether they
could give rise to the as sumptions and rules which
we impose. The status of hidden variables is sim-
ply not relevant here (other than that local hidden
variables are ruled out).
Let us, nonetheless, make a few possible
(but not inescapable) interpretational state-
ments. Given the absence of ontological commit-
ments, this informational approach is generally
compatible with (but does not rely on) the re-
lational [
28, 29], Brukner-Zeilinger informational
[
3135], or QBist [3639] interpretations of quan-
tum mechanics. They depart f rom the traditional
idea that systems necessarily have, absolute, i.e.
observer independent properties (or, more gen-
erally, properties independent of their relations
with other systems). Instead, many physical
properties are interpreted as relational; the inter-
action between systems establishes a relation be-
tween them, permitting an information ex change
which reveals certain physical properties relative
to one another and in the absence of hidden vari-
ables this would be all there is. Certainly, in or-
der not to render such a view hopelessly solip-
sistic, systems ought to have certain intrinsic at-
tributes, e.g. a corresponding state space or set
of permissible interactions/measurements, such
that different observers have a basis f or agree-
ing or disagreeing on the description of physical
objects. However, a state of a system or measure-
ment/question outcome is taken relative to who-
ever performs the measurement or asks the q ues-
tion. Different observers may agree on states or
measurement outcomes by communication (i.e.,
physical interaction) but if one rejects the idea of
an absolute and omniscient observer it is natural
to also abandon the idea of an absolute and ex-
Accepted in Quantum 2017-11-27, click title to verify 72
ternal standard by means of which properties of
systems could be defined.
From such a perspective, one could say that
the relational character of the qubits’ proper-
ties is a consequence of a universal limit on the
amount of information accessible to the observer
and the mere existence of complementary infor-
mation such that the observer can not know the
answers to all his questions simultaneously. It is
the observer who determines which questions he
will ask and thereby what k ind of system prop-
erty the interrogation will reveal and thus, ulti-
mately, which kind of information he will acquire
(although clearly he does not determine what the
question outcome is). But relative to the observer
the system of qubits does not have properties
other than those accessible to him.
8.1 An operational alternative to the wave
function of the universe
Pushing an informational interpretation of quan-
tum theory to the extreme, one may speculate
whether the quantum state could also represent
a state of information in a gravitational or cosmo-
logical context. For instance, is such an interpre-
tation adequate f or the ‘wave function of the uni-
verse’ which is ubiquitous in standard approaches
to quantum cosmology (e.g., see [9597])? Such
an interpretation would require the existence of
an absolute and omniscient observer, an idea
which we just abandoned.
Alternatively, one could adopt one of the cen-
tral ideas of relational quantum mechanics [
28,
29], according to which all physical systems can
assume the role of an ‘observer’, recording infor-
mation about other systems, thereby relieving the
clear distinction used in this manuscript. Ex-
tending this idea to a space-time context, one
could interpret the universe as an abstract net-
work of subsystems/subregions, viewed as infor-
mation registers, which can communicate and ex-
change information through interaction (see fig-
ure
7). In this background independent context,
any information acquisition by any register is in-
ternal, i.e. occurs within the network; a global
observer outside the network becomes meaning-
less. This may appear as a purely philosophical
observation, but it implies concrete consequences
for the description of the network: there should
be no global state (aka ‘wave function of the uni-
verse’) for the entire network at once. Indeed,
PSfrag rep lacements
absolute observer
S
1
S
2
S
3
S
4
S
5
S
6
S
7
communication
Figure 7: T he universe as an information exchange net-
work of subsytems without a b solute observer.
admitting any register in the network to act as
‘observer’, the self-reference problem [
98,99] im-
pedes a given register to infer the global state of
the entire network including itself from its in-
teractions with the rest. Accordingly, relative to
any subsystem, one could assign a state to the
rest of the network, but a global state and thus
a global Hilbert space would not arise. The ab-
sence of a global state in quantum gravity has
been proposed before [100103] albeit from a
different, less informational perspective.
This offers an operational alternative to the
problematic concept of the ‘wave function of the
universe’ which is ubiquitous in quantum cosmol-
ogy. However, while a ‘wave function of the uni-
verse’ might not make sense as a fundamental
concept, it could still be given meaning as an ef-
fective notion, namely as the state of the large
scale structure of our universe on whose descrip-
tion ‘late time’ observers better agree.
Clearly, if this was to offer a coherent picture
of physics, there would need to exist non-trivial
consistency relations among the different regis-
ters’ descriptions such as, e.g., the requirement
that ‘late time observers’ agree on the state of
the large scale structure. This is not a practically
unrealistic expectation as a concrete playground
for this idea has recently been constructed from
a different motivation [
104]: a scalar field on the
background geometry of elliptic de Sitter space
can only be quantized in an observer dependent
manner. A global Hilbert space for the quantum
field does not exist, but consistency conditions
between different observers’ descriptions can be
derived. (Observer consistency has also been dis-
Accepted in Quantum 2017-11-27, click title to verify 73
cussed in general terms within Relational Quan-
tum Mechanics [
28, 29] and QBism [64,65].) Al-
though not being a quantum gravity model, it
may serve as a platform for concretizing this pro-
posal further.
Acknowledgements
This work has benefited from interactions with
many colleagues. First and foremost, it is a
pleasure to thank Markus Müller for uncount-
ably many insightf ul discussions, patient expla-
nations of his own work and suggestions. I am
grateful to Chris Wever for collaboration on [
1,2]
and important discussions, to Sylvain Carrozza
for drawing my attention to Relational Quantum
Mechanics in the first place and many related
conversations, to Tobias Fritz and Lucien Hardy
for asking numerous clarifying questions, and to
Rob Spekkens for many careful comments and
suggestions. I would also like to thank Caslav
Brukner, Borivoje Dakic, Martin Kliesch, Matt
Leifer, Miguel Navascués, Matteo Smerlak and
Rafael Sorkin for useful discussions. Research at
Perimeter Institute is supp orted by the Govern-
ment of Canada through Industry Canada and by
the Province of Ontario through the Ministry of
Research and Innovation.
References
[1] P. A. Höhn and C. S. P. Wever, “Quantum
theory from
questions,”
Phys.Rev. A95 (2017) 012102,
Preprint:
arXiv:1511.01130.
[2] P. A. Höhn and C. S. P. Wever, “A
reconstruction of real quantum theory from
rules on information acquisition," to appear
[3] R. D. Sorkin, On the Entropy of the
Vacuum outside a Horizon,” 10th Int. Conf.
Gen. Rel. Grav. Contributed Papers II
(1983) 734,
arXiv:1402.3589.
[4] L. Bombelli, R. K. Koul, J. Lee, and R. D.
Sorkin, A Quantum Source of Entropy for
Black Holes,”
Phys.Rev. D34 (1986) 373–383.
[5] T. Jacobson, Thermodynamics of
space-time: The Einstein equation of
state,”
Phys.Rev.Lett. 75 (1995) 1260–1263,
arXiv:gr-qc/9504004.
[6] T. Jacobson, Grav itation and vacuum
entanglement entropy,”
Int.J.Mod.Phys. D21 (2012) 1242006,
arXiv:1204.6349.
[7] E. Bianchi and R. C. Myers, On the
Architecture of Spacetime Geometry,”
Class.Quant.Grav. 31 no. 21, (2014) 214002,
arXiv:1212.5183.
[8] E. T. Jaynes, “Information theory and
statistical mechanics,"
Phys. Rev. 106 (1957) 620
[9] C. H. Bennett, “The thermodynamics of
computation—a review,”
Intern. J. Theor. Phys. 21, (1982) 905–940.
[10] K. Maruyama, F. Nori, and V. Vedral,
“Colloquium: The physics of maxwell’s
demon and information,”
Reviews of Modern Physics 81, (2009) 1.
[11] S. Popescu, A. J. Short, and A. Winter,
“Entanglement and the foundations of
statistical mechanics,”
Nature Physics 2, (2006) 754–758.
[12] M. Horodeck i and J. Oppenheim,
“Fundamental limitations for quantum and
nanoscale thermodynamics,”
Nature communications 4 (2013).
[13] M. A. Nielsen and I. L. Chuang, Quantum
computation an d quantum information.
Cambridge university press, 2010.
[14] L. Hardy, Quantum theory from five
reasonable ax ioms,”
arXiv:quant-ph/0101012 [quant-ph].
[15] B. Dakic and C. Brukner, “Quantum theory
and beyond: Is entanglement special?,”
Deep Beauty: Understanding the Quantum
World through Mathematical Innovation,
Ed. H. Halvorson (Cambridge Un iversity
Press, 2011) 365-392 (11, 2009),
arXiv:0911.0695.
[16] L. Masanes and M. P. Müller, “A derivation
of quantum theory from physical
requirements,”
New Journal of Physics 13, (2011) 063001.
[17] M. P. Müller and L. Masanes,
Information-theoretic postulates for
quantum theory,” in Quantum Theory:
Informational Foundations and Foils.
Chiribella G., Spekkens R. (eds)
Accepted in Quantum 2017-11-27, click title to verify 74
Fund. Theories Phys., vol 181, Springer
arXiv:1203.4516
.
[18] L. Masanes, M. P. Müller, R. Augusiak,
and D. Perez-Garcia, “Existence of an
information unit as a postulate of quantum
theory,” PNAS 110, 16373 (2013),
arXiv:1208.0493.
[19] G. Chiribella, G. M. D’Ariano, and
P. Perinotti, “Informational derivation of
quantum theory,”
Physical Review A 84, (2011) 012311.
[20] G. de la Torre, L. Masanes, A . J. Short,
and M. P. Müller, “Deriving quantum the-
ory from its local structure and reversibility,”
Physical Review Letters 109, (2012) 090403.
[21] M. P. Müller and L. Masanes,
Three-dimensionality of space and the
quantum bit: how to derive both from
information-theoretic postulates,”
New J. Phys. 15, 053040 (2013),
arXiv:1206.0630.
[22] L. Hardy, “Reconstructing quantum
theory,”
arXiv:1303.1538.
[23] H. Barnum, M. P. Müller, and C. Ududec,
“Higher-order interference and single-system
postulates characterizing quantum theory,”
New J. Phys. 16 123029 (2014)
arXiv:1403.4147
.
[24] S. Kochen, “A reconstruction of quantum
mechanics,” in Bertlmann R., Zeilinger A .
(eds) Quantum [Un]Speakables II. (2017)
The Frontiers Collection. Springer, Cham,
arXiv:1306.3951.
[25] R. Oeckl, “A local and operational
framework for the foundations of physics ,”
arXiv:1610.09052.
[26] J. B. Hartle, “Quantum mechanics of
individual systems,”
Am. J. of Phys. 36, (1968) 704–712.
[27] Q. Zheng and T. Kobayashi, “Quantum
optics as a relativistic theory of light,”
Physics Essays 9 (1996) 447–459.
[28] C. Rovelli, Relational quantum
mechanics,”
Int.J.Theor.Phys. 35 (1996) 1637–1678,
arXiv:quant-ph/9609002.
[29] M. Smerlak and C. Rovelli, Relational
EPR,”
Found.Phys. 37 (2007) 427–445,
arXiv:quant-ph/0604064.
[30] A. Peres, Quantum theory: concepts and
methods, vol. 57. Springer, 1995.
[31] A. Zeilinger, “A foundational principle for
quantum mechanics,”
Foundations of Physics 29, (1999) 631–643.
[32] C. Brukner and A. Zeilinger, “Operationally
invariant information in quantum
measurements,”
Phys. Rev. Lett. 83 (1999) 3354–3357,
arXiv:quant-ph/0005084.
[33] C. Brukner, M. Zukowski, and A. Zeilinger,
“The essence of entanglement,”
arXiv:quant-ph/0106119.
[34] C. Brukner and A. Zeilinger, “Information
and fundamental elements of the structure
of quantum theory,”
in “Time, Quantum, Information, eds. by
L. Castell and O. I schebeck (Springer,
2003), arXiv:quant-ph/0212084.
[35] C. Brukner and A. Zeilinger, “Young’s
experiment and the finiteness of
information,”
Phil. Trans. R. Soc. Lond. A 360 (2002) 1061,
arXiv:quant-ph/0201026.
[36] C. A. Fuchs , “Quantum mechanics as
quantum information (and only a little
more),”
arXiv:quant-ph/0205039.
[37] C. M. Caves and C. A. Fuchs, Quantum
information: How much information in a
state vector?,”
arXiv:quant-ph/9601025.
[38] C. M. Caves, C. A. Fuchs, and R. Schack,
“Quantum probabilities as bayesian
probabilities,”
Phys. Rev. A 65 (2002) 022305,
arXiv:quant-ph/0106133.
[39] C. M. Caves, C. A. Fuchs, and R. Schack,
“Unknown quantum states: the quantum de
finetti representation,”
J. Math. Phys. 43, (2002) 4537–4559.
[40] R. W. Spekkens, “Evidence for the
epistemic view of quantum states: A toy
theory,”
Physical Review A 75, (2007) 032110.
[41] R. W. Spekkens, “Quasi-quantization:
classical statistical theories with an
Accepted in Quantum 2017-11-27, click title to verify 75
epistemic restriction,” in Quantum Theory:
Informational Foundations and Foils.
Chiribella G., Spekkens R . (eds)
Fund. Theories Phys., vol 181, Springer
arXiv:1409.5041
.
[42] P. A. Höhn and M. P. Müller, “An
operational approach to spacetime
symmetries: Lorentz transformations from
quantum communication,”
New J. Phys. 18 (2016), 063026,
arXiv:1412.8462.
[43] C. Rovelli, Quantum Gravity. Cambridge
University Press, 2004.
[44] C. Rovelli, Time in quantum gravity:
Physics beyond the Schrödinger regime,”
Phys.Rev. D43 (1991) 442–456.
[45] C. Rovelli, What is observable in classical
and quantum gravity?,”
Class.Quant.Grav. 8 (1991) 297–316.
[46] C. Rovelli, Quantum reference systems,”
Class.Quant.Grav. 8 (1991) 317–332.
[47] B. Dittrich, Partial and complete
observables for canonical General
Relativity,”
Class.Quant.Grav. 23 (2006) 6155–6184,
arXiv:gr-qc/0507106.
[48] J. Tambornino, Relational observables in
gravity: A review,” SIGMA 8 (2012) 017,
arXiv:1109.0740.
[49] M. Bojowald, P. A. Höhn, and
A. Tsobanjan, Effective approach to the
problem of time: general features and
examples,” Phys.Rev. D83 (2011) 125023 ,
arXiv:1011.3040.
[50] M. Bojowald, P. A. Höhn, and
A. Tsobanjan, An Effective approach to
the problem of time,”
Class.Quant.Grav. 28 (2011) 035006,
arXiv:1009.5953.
[51] P. A. Höhn, E. Kubalova, and
A. Tsobanjan, Effective relational
dynamics of a nonintegrable cosmological
model,” Phys.Rev. D86 (2012) 065014,
arXiv:1111.5193.
[52] S. Popescu and D. R öhrlich, “Quantum
nonlocality as an axiom,”
Foundations of Physics 24, (1994) 379–385.
[53] M. Pawłowski, T. Paterek, D. Kaszlikowski,
V. Scarani, A. Winter, and M. Żukowski,
“Information causality as a physical
principle,”
Nature 461, (2009) 1101–1104.
[54] T. Paterek, B. Dakic, and C. Brukner,
“Theories of systems with limited
information content,”
New J. Phys. 12 (2010) 053037,
arXiv:0804.1423.
[55] J. Barrett, “Information processing in
generalized probabilistic theories,”
Physical Review A 75, (2007) 032304.
[56] L. Hardy, “Foliable operational structures
for general probabilistic theories,” Deep
Beauty: Understanding the Quantum World
through Mathematical Innovation;
Halvorson, H., Ed (2011) 409,
arXiv:0912.4740.
[57] L. Hardy, “A formalism-local framework for
general probabilistic theories, including
quantum theory,”
Math. Struct. Comp. Science 23, (2013) 399,
arXiv:1005.5164.
[58] G. Chiribella, G. M. D’Ariano, and
P. Perinotti, “Probabilistic theories with
purification,
Physical Review A 81, (2010) 062348.
[59] L. Masanes, M. P. Müller, D. Perez-Garcia,
and R. Augusiak, “Entanglement and the
three-dimensionality of the bloch ball,”
J. Math. Phys. 55, 122203 (2014),
arXiv:1111.4060.
[60] C. Pfister and S. Wehner, “An
information-theoretic principle implies that
any discrete physical theory is classical,”
Nature communications 4 (2013) 1851.
[61] H. Barnum, J. Barrett, L. O. Clark,
M. Leifer, R. Spekkens, N. Stepanik,
A. Wilce, and R. Wilke, “Entropy and
information causality in general
probabilistic theories,”
New J. Phys. 12 (2010) 033024,
arXiv:0909.5075.
[62] A. S. Holevo, Probabilistic and Statistical
Aspects of Quantum Theory,
North-Holland, Amsterdam, 1982
[63] R. W. Spekkens, “Contextuality for
preparations, transformations, and unsharp
Accepted in Quantum 2017-11-27, click title to verify 76
measurements,”
Physical Review A 71, (2005) 052108 .
[64] C. Caves, C. Fuchs, and R. Schack,
“Conditions for compatibility of
quantum-state assignments,”
Phys. Rev. A 66 (Dec, 2002) 062111.
[65] N. D. Mermin, “Compatibility of state
assignments,”
J. Math. Phys. 43, (2002) 4560.
[66] C. A. Fuchs and A . Peres, Quantum s tate
disturbance versus information gain:
Uncertainty relations for quantum
information,” Phys.Rev. A53 (1996) 2038,
arXiv:quant-ph/9512023.
[67] C. A. Fuchs , Information gain versus state
disturbance in quantum theory,”
arXiv:quant-ph/9611010.
[68] E. Specker, “Die logik nicht gleichzeitig
entscheidbarer aussagen,”
Dialectica 14 (1960) 239.
[69] Y.-C. Liang, R. W. Spekkens, and H. M.
Wiseman, “Specker’s parable of the
overprotective seer: A road to contextuality,
nonlocality and complementarity,”
Physics Reports 506, (2011) 1–39.
[70] A. Cabello, “Simple explanation of the
quantum violation of a fundamental
inequality,”
Physical review letters 110, (2013) 060402.
[71] A. Cabello, “Specker’s fundamental
principle of quantum mechanics,”
arXiv:1212.1756.
[72] G. Chiribella and X. Yuan, “Measurement
sharpness cuts nonlocality and
contextuality in every physical theory,”
arXiv:1404.3348.
[73] Č. Brukner and A. Zeilinger, “Information
invariance and quantum probabilities,”
Foundations of Physics 39, (2009) 677–689.
[74] S. Weinberg, Precision Tests of Quantum
Mechanics,”
Phys.Rev.Lett. 62 (1989) 485.
[75] N. Gisin, “Weinberg’s non-linear quantum
mechanics and supraluminal
communications,”
Physics Letters A 143, (1990) 1–2.
[76] J. Polchinski, Weinberg’s nonlinear
quantum mechanics and the EPR paradox,”
Phys.Rev.Lett. 66 (1991) 397–400.
[77] C. H. Bennett, D. Leung, G. Smith, and
J. A. Smolin, Can closed timelike curves or
nonlinear quantum mechanics improve
quantum state discrimination or help solve
hard problems?,”
Phys.Rev.Lett. 103 (2009) 170502,
arXiv:0908.3023.
[78] C. F. von Weizsäcker, The Structure of
Physics. Springer-Verlag, Dordrecht, 2006.
[79] T. Görnitz and O. Ischebeck, An
Introduction to Carl Friedrich von
Weizs äcker’s Program for a Reconstruction
of Quantum Theory.
in “Time, Quantum, Information, eds. by
L. Castell and O. Ischebeck (Springer, 2003)
[80] H. Lyre, “Quantum theory of ur-objects as
a theory of information,”
Int. J. Theor. Physics 34, (1995) 1541.
[81] B. Dakic and C. Brukner, “The classical
limit of a physical theory and the
dimensionality of space,” in Quantum
Theory: Informational Foundations and
Foils. Chiribella G., Spekkens R . (eds)
Fund. Theories Phys., vol 181, Springer
arXiv:1307.3984
.
[82] N. Bohr, Atomic Theory and the
Description of Nature. Cambridge
University Press, 1961 (Reprint).
[83] C. M. Caves, C. A. Fuchs, and P. Rungta,
“Entanglement of
formation of an arbitrary state of two rebits,”
Foundations of Physics Letters 14, (2001) 199 .
[84] V. Coffman, J. Kundu, and W. K.
Wootters, “Distributed entanglement,”
Phys.Rev.A 61 (2000) 052306,
arXiv:quant-ph/9907047.
[85] B. Regula, S. D. Martino, S. Lee, and
G. Adesso, “Strong monogamy conjecture
for multiqubit entanglement: The
four-qubit case,”
Phys. Rev. Lett. 113 (2014) 110501,
arXiv:1405.3989.
[86] W. K. Wootters, “Entanglement sharing in
real-vector-space quantum theory,”
Foundations of Physics 42, (2012) 19.
[87] W. K. Wootters, “The rebit three-tangle
and its relation to two-qubit entanglement,”
J. Phys. A: Math. Theor. 47 (2014) 424037.
Accepted in Quantum 2017-11-27, click title to verify 77
[88] A. Peres, Separability criterion for density
matrices,”
Phys.Rev.Lett. 77 (1996) 1413–1415,
arXiv:quant-ph/9604005.
[89] R. Webster, Convexity. Oxford University
Press, Oxford, 1994.
[90] S. Aaronson, “Is quantum mechanics an
island in theoryspace?,”
arXiv:quant-ph/0401062.
[91] C. Brukner and A. Zeilinger, “Quantum
measurement and shannon information, a
reply to m. j . w. hall,”
arXiv:quant-ph/0008091.
[92] C. Brukner and A. Zeilinger, “Conceptual
inadequacy of the shannon information in
quantum measurements,”
Phys.Rev.A 63 (2001) 022113,
arXiv:quant-ph/0006087.
[93] A. J. P. Garner, M. P. Müller, and O. C. O.
Dahlsten, “The quantum bit from relativity
of simultaneity on an interferometer,”
Proc. R . Soc. A473 (2017) 20170596,
arXiv:1412.7112.
[94] P. A. Höhn, “Reflections on the information
paradigm in quantum and gravitational
physics,”
J. Phys. Conf. Ser. 880, 012014 (2017),
arXiv:1706.06882.
[95] J. Hartle and S. Hawking, Wave function
of the universe,”
Phys.Rev. D28 (1983) 2960–2975.
[96] M. Bojowald, Quantum cosmology,”
Lect.Notes Phys. 835 (2011) 1–308.
[97] A. Ashtekar and P. Singh, Loop Quantum
Cosmology: A status report,”
Class.Quant.Grav. 28 (2011) 213001,
arXiv:1108.0893.
[98] M. L. Dalla Chiara, “Logical self reference,
set theoretical paradoxes and the
measurement problem in quantum
mechanics,”
J. Philosophical Logic 6, (1977) 331.
[99] T. Breuer, “The impossibility of accurate
state self-measurements,”
Phil. Science (1995) 197.
[100] L. Crane, Clock and category: Is
quantum gravity algebraic?,”
J.Math.Phys. 36 (1995) 6180,
arXiv:gr-qc/9504038.
[101] F. Markopoulou, Quantum causal
histories,”
Class.Quant.Grav. 17 (2000) 2059,
arXiv:hep-th/9904009.
[102] F. Markopoulou, Planck scale models of
the universe,”
arXiv:gr-qc/0210086.
[103] F. Markopoulou, New directions in
background independent quantum gravity,”
in Approaches to Quantum Gravity:
Toward a New Understanding of Space,
Time and Matter, ed. D. Oriti (2009),
Cambridge University Press, 129
arXiv:gr-qc/0703097.
[104] L. F. Hackl and Y. Neiman, Horizon
complementarity in elliptic de Sitter space,”
Phys. Rev. D 91 (2015), 044016,
arXiv:1409.6753.
Accepted in Quantum 2017-11-27, click title to verify 78