You are on page 1of 29

The Problem of Time in Quantum Cosmology

and Non-chronometric Temporality


by
Carlos Pedro Gonçalves
Mathematics researcher at UNIDE-ISCTE, in the areas of quantum computation,
quantum formal systems theory, quantum game theory, quantum cosmology and general
risk science
cpdsg.research@gmail.com (primary); cpdsg2@gmail.com; cpdsg@iscte.pt

Maria Odete Madeira


Interdisciplinary researcher in philosophy of science, systems science, complexity sci-
ences, neurocognition, semiotics, ontology and cosmology
mosmg.research@gmail.com (primary), mariaodete.sm@gmail.com

Abstract

We review two lines of argument regarding the problem of time in quantum cosmology and in
quantum gravity, one that invokes the path integral formalism for quantum gravity to state the
absence of time between two three-geometries, and another that defends the absence of time, as
a fundamental notion in physics, in terms of: (a) the configuration space argument , put forward
by Barbour, Smolin and Kauffman, and (b) the Wheeler-DeWitt equation.
We argue that although being correct with respect to a space-time dependent physical
chronometrizable clock-time frame, both of these lines of argument fail with respect to a general
sense of temporality, expressed in terms of the more elementary notions of a before and an after
of a quantum computation.
With respect to the first line of argument, it is shown that the early works on the subject
address two kinds of temporalities, one that is the space-time geometric dependent temporality,
which coincides with the usual definition of a space-time dependent physical chronometrizable
clock-time frame, the other is a temporality associated to the notions of input and output of a
general quantum gravity computation, that is expressed, in the theoretical discourse of quantum
gravity, through the usage of the concepts of: (1) propagation of a wave functional in super-
space, as addressed by Wheeler; (2) transition amplitudes of three-geometries and (3) the path-
integral formalism, used to calculate such amplitudes, as addressed by Hartle and Hawking.
While the first temporality (space-time dependent temporality) disappears from the theory, the
second plays a fundamental role, not only in the several aspects of the theory’s construction, but
in the clock-time independence as well, as Wheeler showed.
Given this notion of time, different from a chronometrizable, space-time geometry internal
notion, we search for a general mathematical and logical structure that is capable of addressing
it from a formal point of view. This is done through a family of mathematical structures that is
more general than the mathematical category. These structures not only will allow us to address
the nature of the temporality present in the transition amplitudes between two three-geometries,
but they will also allow us to refute the configuration space argument and to show how a static
clock-time-independent quantum state, can be put into a non-clock-time processual expression
in terms of fine-grained computational histories, obtained from the relations between different
observable’s bases.

Keywords: Quantum cosmology, time, relational structures, relational nexus

1
1. Introduction
One of the major open problems, within quantum cosmology, is the so-called problem of
time, which is usually expressed in terms of a general statement of the absense of time
in quantum cosmology (Smolin, 2001).

The arguments underlying this claim can be generally split in two lines of argument
that arise from the formalisms used in addressing quantum cosmology and quantum
gravity. One of these lines comes from a path-integral and propagator approach to
quantum gravity, where the transition amplitudes, evaluated between two three-geome-
tries, are obtained by summing over all four-geometries that aggree with the initial and
final three-geometries at the border (Hartle and Hawking, 1983). The second of these
two lines comes from the analysis of the Wheeler-DeWitt equation, and it can be called,
to a good approximation regarding the underlying argument, the time-independent con-
figuration space argument. An example of such argument can be found, for instance, in
Barbour (1994), in Smolin (2001) and in Kauffman and Smolin (1997).

We address, in this work, each of these two lines of argument, and show that,
although the claim for the non-existence of time, at a fundamental level, is correct with
respect to a space-time internal temporal chronometrizable frame, it is not correct as a
general statement with respect to the presence of a temporality independent from space-
time internal temporal chronometrizable frames.

Regarding the first line of argument, the discourse itself of the works that addressed
the propagator and the path integral formalism in quantum gravity trivially shows, upon
closer inspection, the presence of this other temporality, restricting the validity of the
claim, for the absence of time, to chronometrizable temporal frames, internal to a space-
time geometry. For, otherwise, contradictory elements would be present in the physical
discourse about the theory, leading to problems when addressing the formalism.
These “apparent contradictions” already appear in the early works within quantum cos-
mology and quantum gravity, including Wheeler’s 1968 article “Superspace and the
Nature of Quantum Geometrodynamics” and Hartle and Hawking’s 1983 article “The
Wave Function of the Universe”.

In Wheeler’s article, the conclusion that there is no sense of time in quantum


geometrodynamics is defended and sustained by a set of arguments. However, in con-
cluding about this non-existence, Wheeler uses, as a supporting argument, the behavior
of the propagation of a wave packet in superspace, which immediately becomes problem-
atic with the general statement about the non-existence of time in quantum cosmology,
as a quantum propagation exemplifies a temporality property that comes from a
quantum gravity computational process, expressed in terms of a unitary transition.

This “apparent contradiction”, as we argue in the current work (section 2.), is only
an apparent one, due to an ambiguated phrasing of the statement regarding the absense
of time in the theory of quantum gravity. Indeed, the statement should be taken to refer
only to a notion of a physical clock time, internal to a space-time geometry, this phys-
ical clock time is undetermined because the space-time geometry is undetermined. How-
ever, not all temporality is removed from the description, since there is still a non-
chronometric temporality associated with the basic notions of input and output of a uni-
tary transformation of a quantum state defined over the superspace, a temporality that
is of a different nature than that of the internal chronometrizable temporal frame, that
depends upon the space-time geometry for its chronometrization.

2
This is what allows Wheeler to speak of a propagation of a wave packet in superspace,
and what allows Hartle and Hawking to speak of a transition amplitude between two
three-geometries, and work with the path integral formalism to calculate such ampli-
tude, without incurring into the fundamental problem of using theoretically inconsistent
terminology.
As we argue, in greater detail, in section 2., several of the statements and basic for-
malism, worked out by Hartle and Hawking (1983), can only be consistent with a
quantum cosmology if we accept that chronological or space-time dependent clock time
is undetermined in the theoretical description, and, simultaneously, understand that
quantum gravity introduces a second type of temporality different from the clock-time
that is internal to a fixed space-time. This second temporality, with which we are con-
fronted, through these early works on quantum cosmology and quantum gravity, is not
chronometrizable with respect to a pre-given space-time, and has a nature only depen-
dent upon a notion of a before and an after associated with an input and an output of a
general quantum gravity computation.
Indeed, in order for this temporality to be chronometrizable one would have to have a
clock-time frame, which is a notion internal to a pre-given space-time geometry, which is
not well defined here. Between two three-geometries we cannot measure how temporally
separated they are, because that measurement depends upon the four-geometries that
are being summed over in the path integral. All that remains, therefore, are the more
primitive temporal notions of a before and an after , since one three-geometry appears
before the other in the (reversible) propagator formalism, hence the reason for the terms
initial and final three-geometries used by Hartle and Hawking (1983).
Understanding the presence of this temporality within the theoretical description, and
showing it to be present in some of the foundational works on quantum cosmology and
quantum gravity are the main objectives of section 2., where we provide for a brief
review on the conception of time within physics.
In section 3., we propose the usage of a mathematical structure, called a binary rela-
tional structure. These structures should not be confused with the usual mathematical
set theoretical structures known as binary relations that belong to the category Rel.
Indeed, the general family of structures called mathematical categories is a sub-family of
these binary relational structures. In an intuitive, first approximation, one may consider
the binary relational structure as simply composed of a collection of objects and a collec-
tion of arrows or morphisms, along with the identity morphism. However, this would be
just a first approximation, since even the notion of morphism is a particular case of the
relations considered in these structures.
The collection of relations, in the binary relational structures, are not necessarily set-
theoretical binary relations. Indeed, as we show in the main text and in the appendix to
the present work, the mathematical notion of category is a particular case of these struc-
tures. In the appendix, we address the examples of Rel and Mag, in order to make
clear the distinction between the binary relational structures and Rel.
In section 3. we use the binary relational structures in order to understand and for-
malize the fundamental non-chronometrizable temporality that is found within the
works on quantum cosmology and quantum gravity.
Afterwards, we apply the formalism to address the second line of arguments regarding
the absence of time in quantum cosmology. In particular, we show how, even a time
independent quantum state, can be ascribed a computational description with respect to

3
the unitary transformation between two different bases, which results in a processual
and historical nature for the quantum state, that is independent of clock time, but that
can be formulated in terms of a fine-grained quantum computational history, intro-
ducing a consistent histories’ type of formalization for a clock-time static quantum state.

This allows us to address the second line of arguments for the absence of time in
quantum cosmology, based upon the Wheeler-DeWitt equation and upon the notion of
configuration space. In section 3. we review this line of argument, and address its prob-
lems with respect to its foundational notions, along with the consequences of the appli-
cation of the mathematics of the binary relational structures to the formalism underlying
the argument.

Central to the present work is, thus, the concept of binary relational structure, where
a notion of temporality emerges from a partition of a relational nexus of any two of the
structure’s objects, which introduces the notions of a before and an after that result
from the positions of the objects in the relations that connect them.

In section 4. we conclude with some final reflections upon the main results, and
address what could be considered as open issues that may arise with respect to the
mathematics of binary relational structures, in connection to physics and quantum
gravity.

2. A brief reflection on the conceptions of time in


physics
With Galileo the time was conceptualized as a fundamental physical quantity, measur-
able for a whole series of physical systems and, consequently, susceptible of regulating
experiments and relating them mathematically, thus, the time became the measure of
the motion (Klein, 1995).

In Newton, this arithmetizable, physically measurable time was considered to be abso-


lute and to flow uniformly, independently of the reference frame (Klein, 1995). In spe-
cial relativity, these two notions were understood as relative. Apparently, nature seems
to be such that the flow of time depends upon the relative motion of the inertial refer-
ence frames. Which led Einstein to place the event (that, one should add, is simulta-
neous with itself) as the central element of the physical theory.

Special relativity places us before a different perspective, regarding the nature of time.
Indeed, accepting special relativity entails accepting a distinction in the cosmic time
between its arithmetizable chronometric nature, which is relative, and its fundamental
temporal sense linked to causality, which is not relative, for the causality horizon of each
event is a relativistic invariant (Einstein, 1953, [2004]).

In accordance with special relativity, when we try to find a causal explanation of an


event we must seek it in its anteriority, that is, in that which took place before the
taking place of the event, that which is to the causal past of the event, which means
that we can associate a relativistic invariant sense of what came before, as that which
lies in the past light cone of the event, which forms the past causality horizon of the
event. In the same way, each event can be considered to be at the cause of other events,
which means that each event has a future causality horizon, making it a border between
what comes before and what comes after.

4
We are in the horizon of causality of the dinosaurs, because the information and
records of their existence, the whole line of biological, evolutionary and geological effects
of their existence lie in our causal past.
Also, in the auto-biographical record of an individual, this sense is present, of what
came before and what may come after, of what took place in the past, and what may
take place in the future. According to Einstein (1953, [2004]) the states of consciousness
of an individual appear, to that individual, pictured in a series of events in which each
particular state, accessible to the individual’s memory, appears to be placed in accor-
dance with an irreducible criterion of a “before” and an “after”. There is, thus, to each
individual, a sense of personal time, a subjective time that is not, in itself, and, still
according to Einstein (1954, [2004]), measurable, although one can associate numbers to
the different events, such that what takes place after something else, always gets a
higher number, even if only the ordering of events matters, for the subjective time, and
not the particular distance between the two numbers, which means that this time is not
metrizable, but, solely expressibe upon an ordinal scale.
Einstein (1953, [2004]) added that physical time could be considered as a physical
reality since, even if its passage was relative, that passage of time was measurable by a
clock moving with the reference frame, this measurability making its measurable nature
objectively relative. Any physical system with a periodic motion can be considered to
constitute a physical clock, albeit the pure notion of a mechanical clock is an idealized
one, the clock becoming a conceptual device to express an objectivity of a time that is
part, in special relativity, of a space-time continuum (Einstein, 1953, [2004]). In this
way, special relativity allows one to think about time as a component part of the topos
(in the original sense of the Greek term as place of the body) itself of the events. As
Weyl (1952) stressed, the change, the motion and the transformations exist in time
itself. The world is active, according to Weyl, in the sense that its phenomena are
related by a causal connection.
The relativistic invariance of causality in Minkowski’s space-time, and the conceptual
importance of the event, influenced a line of thinking and interpretations of quantum
theory and quantum cosmology where these two notions play a fundamental role.
Markopoulou’s proposal of quantum causal histories, as a basis for quantum cosmology
(Markopoulou, 2000), finds its support in the argument that the standard approach to
quantum cosmology suffers from the problem of following a tradition of addressing
quantum mechanical problems where a single wave function describes the entire system,
pressupposing a reasoning of an observer who has access to the whole system. The idea
of a local quantum evolution, and that, at the fundamental level, there may be no such
thing as a quantum state of the universe, is a distancing from the usual approaches to
quantum cosmology that work with the Wheeler-DeWitt equation (Markopoulou, 2000).
Smolin (2003; 2006), who defends Markopoulou’s approach, which fits well with loop
quantum gravity, considers that the universe may be thought of in terms of relations
between events. The notion of causal relation, in the histories of physical processes, is
justifiable within the quantum unitary evolution, which is deterministic. The notion of
causality, as Smolin and Markopoulou use it, has, however, a double sense.
In a direct and immediate sense, the causality can be understood in terms of a set of
events that constitute a necessary condition for the occurrence of another event, in that
sense, we have that an event E(n) is in a relation of causality with a set of events E(1),
E(2),  , E(n − 1) , if it is physically necessary that:

E(n) → (E(1) ∧ E(2) ∧  ∧ E(n − 1))

5
where → is he symbol for the material implication. In this case, provided with the infor-
mation that the conjunction (E(1) ∧ E(2) ∧  ∧ E(n − 1)) holds (in the sense that it is
true that each event in the conjunction has occurred), then, E(n) must also have
occurred.
Given a network of causal relations, defined by implications of the kind introduced in
the previous paragraph, and given the truth value of the necessary conditions of the
material implications, we know all of the network’s narrative, in terms of what occurred
and what did not occurr. This fact constitutes a central feature of the relation of
causality expressed by the above implications, which is, thus, a deterministic relation.
Smolin (2003), however, presents an additional interpretation of a causal network,
considering it in terms of information processing, such that, each event can be thought
of as taking the information of the other events in its causal past and performing a com-
putation (or a quantum computation in the quantum theoretical setting), sending the
result of such a computation to its future through the (quantum) causal network.
This computational notion, associated with causal histories, is such that we can con-
sider, in a more general sense, the causal past of an event as the set of events from
which it can receive information, and the causal future of an event can be considered to
be the set of events that can process information from that event.
The special theory of relativity establishes a limit to causality, as it distinguishes
between timelike, spacelike and lightlike separations.
Quantum theory, on the other hand, places us before the problem of nonlocal connec-
tions, which are not causal but spacelike correlational, in the sense that we need to con-
sider the entire spacelike surface, where two or more entangled events occur, to be an
interconnected whole, described by a quantum state.
This configures a local arrow of time in the form of local temporal directedness due to
entanglement-induced local decoherence. Information is “exported” to the whole, and
only the whole has the whole information, so to speak.
In general relativity, and in the standard approach to quantum cosmology, the con-
cept of time becomes more problematic, though less so, as we saw, in the quantum
causal histories approach.
The first thing to notice, as Wheeler (1968) stressed, regarding classical geometrody-
namics, is that the concept of event becomes less primitive and less significant. One may
choose to consider space-time to be comprised of elementary objects or points called
events, or, one may, instead, choose to consider the three-geometry to be the primary
concept, and the event the secondary concept, an event lying at the intersection of such
and such three-geometry.
Under this last perspective, the temporal relation between two three-geometries would
be determined by the structure of the four-geometry, which in turn derives from the
inter-crossings of all the other three-geometries.
In classical geometrodynamics, whether one started with the three-geometry as the
primitive concept, or with the event as the primitive concept makes little difference,
according to Wheeler. However, still according to Wheeler (1968), it makes all the differ-
ence when one turns to quantum geometrodynamics, for there is no such thing as a well-
defined four-geometry, because, as far as we know, and still according to Wheeler, no
probability amplitude function can propagate through superspace as an indefinitely
sharp wave packet, the wave packet spreads.

6
The conclusion, following Wheeler’s argument, is that the space-time, the time, the
notion of before and after do not exist as primitive notions within the theory, which is a
problematic statement, from a theoretical discursive point of view, given the fact that
even if a four-geometry is undetermined, and even if the before and after of a space-time
geometry are internal to that geometry, there is still the problem of the propagation of
the wave packet in superspace, that, from the moment in which we use the term propa-
gation, commits us to a temporal sense present in superspace, different from the internal
times definable and chronometrizable within each four-geometry.

It is important to review Wheeler’s argument, with respect to this matter, in order to


better understand this issue. The first thing to notice is that the spreading of the wave
packet means that it associates a finite probability to a domain of superspace of finite
measure, this domain encompassing a set of three-geometries far too numerous to
accommodate in any one four-geometry.

One can express this situation, according to Wheeler (1968), by stating that the prop-
agation takes place in superspace, not by following any one classical history of space but
by summation of contributions from an infinite variety of such histories. It is noticeable
here the unavoidance of the apparent “discursive trap” that talks about quantum evolu-
tion in superspace, which installs, in the discourse, a quantum temporality that is proper
to the notion of unitary quantum evolution of a wave function, to the notion of a propa-
gator in superspace, and to the notion of sum-over-histories of space-time geometries.

However, this “discursive trap” is only an apparent one, due to an ambiguation of two
temporal senses that must be conceptually disentangled. On the one hand, we have the
chronological time that is an internal definition to each four-geometry, and, on the other
hand, we have a temporal sense, distinct and not ontologically committed to an internal
measurable chronometrics, but, instead, associated with the more elementary notion of a
temporal property simply exemplified by an order of a before and an after of a computa-
tion, an input and an output.

Thus, in this case, Wheeler can talk about a unitary connection, and use temporal
discursive elements with respect to it, without incurring into any problem of discursive
bias due to unexpunged terminology, habituated to a regular usage of quantum
mechanics. Indeed, the unitary connection between two three-surfaces introduces a tem-
poral nature of its own, the temporality associated to a notion of a before and an after ,
an input and an output of a non-chronometric connection, which, itself, contains a multi-
tude of probable chronometrics, each associated with different (to be summed-over) his-
tories of space-time geometries (the internal times being a part of these histories).

The three-geometries that occur with significant probability amplitudes do not fit,
according to Wheeler, and cannot be fitted into any single four-geometry. Without that
building plan, what we call the internal cosmic time, that might be definable within a
fixed four-geometry, at least locally, is undefined and, without a building plan to orga-
nize the three-geometries of significance into a definite relationship, one to another, even
the internal geometric-specific notion of a time ordering of events is devoid of meaning.
One must stress that this time ordering, and this before and after, is distinct from that
of a unitary connection that connects an input and output of a (general1) cosmological
quantum computation.

1. General in the sense that it does not necessarily involve a qubit.

7
From this, Wheeler concludes that the concepts of space-time and time itself are not
primary, but secondary in the structure of a cosmological physical theory. Although
Wheeler may be right about the first (space-time), there are some problems with the
conclusion about the second (time), when it is stated categorically or in a general sense
of the term time. The time, to which Wheeler is referring, is a four-geometry-dependent
chronometrizable temporal frame, this time loses meaning within a quantum cosmolog-
ical setting, since there is no well-defined geometrical structure upon which one may
define notions, internal to that structure, such as that of before, after, present, past or
future, where each event occupies a position in a “grand catalog called space-time”, to
borrow Wheeler’s expression.

Nonetheless, as we stated above, we still have to deal with quantum unitarity, and the
path-integral formulation still brings with it a notion of a temporality associated with a
notion of a quantum superspace history, a temporality that is distinct from the arithme-
tizable, chronometrizable time that may be defined with respect to a four-geometry.

This distinction between the two temporalities, the internal space-time and the
quantum evolution temporality, becomes more explicit in later works on quantum cos-
mology, and, in particular, in Hartle and Hawking’s work (Hartle and Hawking, 1983).

As stressed by Hartle and Hawking, when we define the amplitude to go from a three-

geometry hij , to another three-geometry hij , we must sum over all the internal four-

geometries that match hij on an initial surface and hij on a final surface, the two sur-
faces being connected by this sum. One can see here that, through the sum, we are
obtaining, in a first, and crude, approximation of the quantum cosmological problem,
the transition function (Hartle and Hawking, 1983):

Z

hhij |hij i = δg µν exp(iSE [g µν ]) (1)

where SE is the classical action for gravity, including a cosmological constant Λ.

As Hartle and Hawking (1983) noticed, when addressing the above formula, one
cannot specify the time in these transition amplitudes. Which is indeed true, in the
sense of a chronometric time defined within the four-geometry. This time, or rather, this
notion of time is dependent upon the four-geometry, so that it is undetermined. How-
ever, we also see the usage, by the authors, of the term propagator to address the above
definition (Hartle and Hawking, 1983), which is, by definition, a theoretical notion rid-
dled with temporal connotations, that form part of its fundamental semantics.

Although this might seem to introduce a contradiction, or an ambiguated discursive


problem, it
is in
fact
a correct statement, as long as we understand that the transition

amplitude hij hij is, from a mathematical point of view, a relation between an input
state and an output state, which expresses a primitive notion of temporality, neither
committed nor inseparable from a chronometrics based upon a space-time geometry.
Thus, we have a temporality that is different, in its nature, from any global, or even
local, time direction, definable within the context of a space-time geometry, which is a
notion internal to the quantum theory underlying Eq.(1).

Thus, the unitary connection, expressed mathematically by the integral on the right of
Eq.(1), can be thought of in light of an input and an output of a (general) quantum
computation.

8
In this sense, we are confronted with another notion of a time which is different from
the physical, arithmetizable, chronometrizable time, and which is a more primitive
notion associated with a before and after of a computation, where the how much before
and the how much after are senseless statements since we have no way of assigning num-
bers to measure a chronological distance between input and output, because that
chronological distance is part of what is the object of probabilization in the theory,
being summed-over in the computation istelf.
Such a realization leads to the need to find, within the mathematical and logical
inquiry about the time, a more primitive structure than the temporal arithmetic that
was associated to the notion of time in physics. We provide for a contribution for such a
reflexive inquiry, in the next section, from a mostly mathematical point of view, by
addressing a primitive family of mathematical structures that are close to the notion of
mathematical category, and in which primitive non-chronometrizable notions of time are
obtained.

3. Primitive non-chronometrizable temporalities and


relational structures
In defining physical and mathematical notions, intended for applied science, we have
been largely influenced by a fundamental role of the notions of quantity and number.
From a mathematical point of view, this raises the issue that we may run into mathe-
matically addressable problems where the notion of number and the notion of quantity
are not fundamental. In that sense, even though a mathematically inclined science of
something that is not measurable can be developed, that science cannot be considered to
be x-metrics (where x stands for whatever area of application one is addressing).
It is a known mathematical fact, that the notion of order precedes the notion of order-
able quantity in terms of mathematical generality. Indeed, although an orderable quan-
tifiable set can be numerically labeled with respect to the quantifiable property that
defines it, and, thus, ordered in terms of the numeric values assumed by the members of
the set with respect to that underlying property, not all orderable sets can be numeri-
cally labeled in terms of any significant orderable quantity property, that is, one can
numerically label a non-quantifiable set, but those numbers are, in general, arbitrary
assignments. We already saw an example of this, when we addressed Einstein’s views on
the subjective time. In statistics this is especially pertinent, since a qualitative ordinal
variable can be numerically labeled in an arbitrary way, so long as the order of the
objects in the ordinal scale is reflected in the order of the numeric assignments. Without
a theoretically justified interval distance, any numeric interval between levels in the
ordinal scale is valid.
With the development of mathematics and metamathematics we came to realize that
some mathematical notions and structures are very general, primitive and, thus, funda-
mental. As a consequence of these findings, we find ourselves before a major theoretical
change, coming from the displacement of the focus from the notions of quantity and
number to the notions of relation and of morphism (the last in the context of category
theory).
A most primitive formal structure, within mathematics, is the relational structure,
which is nothing more than a collection of objects along with a collection of relations.
We can define the most simple case of binary relational structure, in a structured form,
as follows:

9
Definition 1. A binary relational structure R = (O, R) is composed of a collection of
objects O, whose members are called the objects of R, and a collection of binary rela-
tions R satisfying the following four conditions:
1. For every related pair of objects X , Y ∈ O there is a non-empty set of relations
R ′ ⊆ R, with cardinality at least one, such that, for every R ∈ R ′ , the well formed
formula R(X , Y ), which reads X is related to Y, under R, holds.
2. For each relation R in R, there is at least one pair of objects X,Y in O, such
that, it holds that R(X , Y ).
3. For any R ∈ R, its truth set 2 is composed only of pairs of objects in O.
4. There is a relation I in R such that, for every object X in O, the well formed for-
mula I(X , X) holds, and I(X , Y ) if, and only if, X and Y are the same object in
O.

The extension to n-ary relations follows immediately.


It is important to stress that we are defining relations in terms of their intension 3
rather than in terms of their extension. Usually, in the mathematical theory of binary
relations one works with the extension, namely, with the set of objects that satisfy the
relation, an approach which abstracts away the meaning and definition of the relation,
i.e., its intension, to focus solely upon the truth-set representation, where this set is
defined in its extension. This process leads to a level of analysis where the relation is
reduced to a set theoretical structure and, in particular, to the composition of the set,
which means that two relations become identical, in accordance with the notion of set
identity, if their truth sets are identical, even if, they are indeed different with respect to
their intension.
In the above structures this is not so. We are working with the notion of logical inten-
sion with respect to the relations. Which means that, even if two relations have the
same truth-set extension, they are still considered to be different, if they have different
intensions. For instance, if a married couple, Fred and Wilma, owns a restaurant, then
the set theoretical expression would be {(Fred, Wilma), (Wilma, Fred)} for both the
business partner relation and for the relation of marriage. However, in terms of the
above relational structures the relations “being married to” and “being a business partner
of ” would constitute two different relations with respect to the universe of discourse
composed of {Fred, Wilma}, even if each of these two relations has the (extensionally
same) truth set, given by: {(Fred, Wilma), (Wilma, Fred)}.
This, and the fact that the objects that we work with are not necessarily sets, means
that one should not confuse the relational structures as mathematical objects with the
category Rel, which is the category of relations that has sets for the class of objects
and, as the morphisms, binary relations, defined extensionally as subsets of ordered
pairs. Indeed, we show in the appendix to the present work, that the mathematical cate-
gory is a particular family of binary relational structures. We also address, in the
appendix, the category Rel and the category Mag to make clearer the difference
between the binary relational structures defined above and the category Rel. Further-
more, we address the role played by Rel in the nature and properties of the truth sets of
the above binary relational structures.
2. We call the model of the relation, that is, the set of pairs of objects for which the relation holds, the truth
set. The truth set establishes a connection between the binary relational structures and the category of the binary
relations over the class of sets, even though the binary relational structures are more general than the notion of
mathematical category. In the appendix we address these two issues.
3. This is a logical notion not to be confused with the notion of intention with a t.

10
Besides this point, that should be taken into account when working with the above
definition, it is important to understand each of the four conditions.
Since each of the four conditions that define the binary relational structure can be
extended to n-ary relational structures, we discuss the role and meaning of each condi-
tion in a general sense. The first condition tells us that all the relations between any
two objects (or n-tuples of objects for the general case) are in R, which means that R is
an exhaustive set of binary (n-ary) relations with respect to the underlying collection of
objects O. The second condition limits each relation to be exemplified by at least a pair
of objects in O, or, in a more general sense (for n-ary relations), the truth set of each
member of R must be different from the empty set.
The third condition restricts the relations in R to be solely relations between objects
of O, which means that R is exclusive with respect to the underlying collection of
objects. Finally, the fourth condition introduces a reflexive relation that is only satisfied
for an object and itself. This relation of an object to itself is called an identity. The
identity is always stated in terms of a binary relation, and it is assumed to be included
in every definition of n-ary relational structures.
Indeed, we have to assume the relation of identity to be able to work with the notion
of object, because otherwise the object would lose its integrity (Madeira, 2008b). Any
structure of individuated objects must be such that each object is maintained and sus-
tained in its identity, while it remains an object of the relational structure (Madeira,
2008b).
Now, as it is defined, a binary relational structure does not include, in its definition,
an explicit temporality of any sense, other than the eternal return of an object to itself
via the permanent coincidence of the object with itself, in the identity relation I.
A primitive sense of temporality associated with a before and an after , primitive in
the sense of not being generally measurable in terms of any kind of clock-time, arises
from an analysis of the notion of relational nexus.
This notion can be built in stages. First, we consider, for any object X ∈ O, the
subset of objects with which X is in relation, and write O(X). We know that this set is
non-empty, since each object is at least in relation with itself.
We can first define an identity nexus of an object N (X&X) as the singleton {I }.
Now, for any Y ∈ O(X), distinct from X, define the relational nexus of X and Y to be
the subset of R of the relations that X and Y exemplify, and denote this nexus by
N (X&Y ). Given this relational nexus one defines the partition between: (a) those rela-
tions, if there are any, in N (X&Y ) whose truth sets contain (X ,Y ) as element but not
(Y , X ); (b) those relations, if there are any, whose truth sets contain (Y , X ) as ele-
ment, but not (X ,Y ); (c) those relations, if there are any, whose truth sets contain both
(X ,Y ) and (Y , X ) as elements.
The first element of the partition, that corresponds to the case (a), as defined above,
is the relational nexus denoted by N (X ֌ Y ), the second case (case (b)), is the rela-
tional nexus denoted by N (X ֋ Y ),4 and the third case (case (c)) is the relational
nexus denoted by N (X ↔ Y ).
The union of these three cases recovers the full relational nexus of X and Y , that is:

N (X ֌ Y ) ∪ N (X ֋ Y ) ∪ N (X ↔ Y ) = N (X&Y ) (2)
4. We stress that N (X ֋ Y ) = N (Y ֌X).

11
Once we partition the relational nexus in terms of the above sequence of disjunctions,
we obtain a directional sense for the relations, this directional sense introduces the fun-
damental notion of a before and an after in a relation. Thus, for instance, in the rela-
tional nexus N (X ֌ Y ), Y comes after X, while in the relational nexus N (X ֋ Y ) X
comes after Y , finally, in the relational nexus N (X ↔ Y ) there is a bidirectionality, in
the sense that if we begin in X , then, we see Y coming after X , while that, if we begin
in Y , then, we see X coming after Y .
It is noticeable that we can break up each relation in N (X ↔ Y ) into two unidirec-
tional sub-relations, one leading from X to Y and another from Y to X .
Considering the general partition of Eq.(2), we see the emergence of a notion of before
and a notion of after , as a consequence of the configuration of the relational nexus of
two objects. These notions appear in these structures as primitive ordinal notions, their
mathematical nature is not fundamentally restricted by the imposition of an underlying
chronometrics.
In the above definition, one does not impose any underlying fundamental numeric
property, or general family of properties that would introduce, foundationally, a tem-
poral metric, allowing the measurement of the separation between the two terms in the
relation.
In order for a temporal metrics to emerge, one would have to assume an axiomatics
that restricted the formalism to a subset of relational structures whose objects might
exemplify, fundamentally and foundationally a numeric property that could form the
basis for a constructible chronometrizable relational nexus. But this would be a subset of
the above relational structures, a subset for which the relational nexus is obtained from
the exemplification of an underlying numeric property.
Another restriction could be obtained through an appropriate axiomatic system. It is
known, for instance, from expected utility theory, that an ordinal system can be trans-
formed into a numerical scale if that ordinal system satisfies certain properties with
respect to combination with probabilities. We shall return to this issue at the beginning
of section 4., when we present a final reflection upon the main results of the present
work and address, in that reflection, some of Barbour’s claims and the nature of a phys-
ical clock time in relation to the nature of the temporality of the relational structures,
introduced in this section.
Now, so far, we have the emergence of a notion of before and and a notion of after ,
not necessarily chronometrizable with respect to some underlying physical or mathemat-
ically significant quantifiable property, since we cannot, in general, assign, in a relational
structure, a chronometric sense of how separated two objects are in a relation. Nonethe-
less, we see a before and an after and the emergence of a general temporal sense,
expressible solely in terms what takes place before and what takes place after , in the
order of the relation, without the ability to state how much before or how much after, or
even how late. This makes the temporality of the above partition more general, but,
even so, present in the formalization.
A further inspection of the above partition, however, shows the presence of a more
structured sense of temporality, other than the simple before and after . Indeed, in two
cases the before and after are not interchangeable, while in the last case they are inter-
changeable. This makes evident that the roles of what comes before and what comes
after are frozen in the first two elements of the partition, and we have an arrow of time
and an irreversibility. Indeed, the full relational nexus partitions in two temporally
asymmetric relational nexus (N (X ֌ Y ) and N (X ֋ Y )), and a temporally symmetric
one (N (X ↔ Y )).

12
These results show how a temporal logic may be present without reference to any
chronometrics. The problem of an undefined metric for time, therefore, does not exclude
the existence of a temporality, or even an ability to address such a temporality within a
formal, mathematical theoretical setting.
Besides providing for an intuition and a formalism that is able to address the path-
integral and propagator extra-clock-time temporalities of quantum gravity, these mathe-
matical structures bring to light a problem in how one conceptually addresses configura-
tion spaces and the relational nature of configuration spaces. The natural emergence of
a temporality, not necessarily defined in terms of a clock temporality, but present in the
partition of a relational nexus, contrasting with what appeared to be a timeless struc-
ture of relations, is sufficient to raise debatable issues with respect to the time-indepen-
dent configuration space argument for the elimination of time as a concept, when it is
taken in its full generality, and not in terms of the chronometrizable clock-time frame,
internally definable within a pre-given space-time geometry.
Although this line of argument can be seen as naturally arising within quantum cos-
mology, it has been argued with respect to classical cosmology as well, and, even, to
physics in general (Kauffman and Smolin, 1997; Smolin, 2001; Barbour, 1994).
Indeed, the argument applies to a general system, placing, as a fundamental concept,
the configuration space, which can be defined as the space of possible configurations of
that system (Barbour, 1994). Three elements are essencial to this definition – the notion
of space; the notion of configuration and the category of possibility.
The notion of configuration, in quantum mechanics, is a familiar one, in a more gen-
eral sense, however, which is the one assumed in the general argument, the term configu-
ration must be taken with respect to its etymological genetics in the Latin term
cum+figurare, which means to shape together , where the action of forming or of making
shape – figurare – has the corresponding noun figura which means shape, form, or pat-
tern.
A configuration is not only a shape, form or pattern, it possesses an active sense of
patternization or making pattern, this active sense is committed, in its conceptual and
etymological genetics, to a temporalization, present in the process by which the system
takes on a shape, or form, or produces a certain pattern. The denial of the existence of
such a semantic connection is an error, proceeding from uprooting the terms from their
underlying etymological structure and semantics. The fact that these terms have a dis-
cursive and conceptual pre-existence, rooted in both a linguistic and philosophical tradi-
tion that has addressed them, is a first problem that introduces one element of illegiti-
macy and error in the line of argument.
This is not the only problem with the configuration space argument for the non-exis-
tence of time, but it is a sufficiently important one, since the proponents of this line of
argument argue against the existence of time in a general sense, that is, any kind of
temporality is considered not to be fundamental. This being so, then, one should be
careful in the choice of the key terminology, if that terminology is somehow committed
to a notion of temporality, and if that terminology plays a constructive role in the
theory, then, the statement of no-temporality becomes inconsistent with respect to the
theory itself.
The second important element to the notion of configuration space is the category of
possibility, that, in this case, introduces a physical openness of the system to different
alternative configurations which it can assume. Along with the statement of possibility,

13
comes the final element which introduces a geometrization of the different possible con-
figurations in the form of a space of possible configurations, where each point is, as Bar-
bour (1994) put it, a distinct structured whole, this means that we can indeed consider
each configuration as an object of a relational system.
Then, we have the collection of objects C, which is the configuration space, as our fun-
damental building block. The proponents of the configuration space argument, and, in
particular, Barbour (1994), assume that physics must be built in terms of such a collec-
tion of objects and relations between the elements of such a collection, without reference
to time.
And, indeed, one can do so. In succint terms if one does not consider the temporal
physical clock chronometrized labels, associated to a curve in a configuration space, one
loses any sense of a clock time and one has a simple spatial curve without any reference
to time. From this point to the next step that states that there is no temporality, may
seem as a legitimate direction. However, it is not. Indeed, one loses any track of any
kind of clock time and one has a curve in a space, without any reference to a physical
time, which makes the temporal labelling a secondary aspect of the world. However, the
fact that the curve is a relational object imposes a temporality of its own, which may be
chronometrizable with respect to physical criteria but not chronometric with respect to
its fundamental temporal nature.
Indeed, we may work with the relational structure (C , R), where the relational nexus
between two objects x, y ∈ C can be, naturally defined as:

N (x ↔ y) (3)

where each relation in N (x ↔ y) is a curve in configuration space connecting the two


points/configurations. It is noticeable that this nexus introduces a reversible connection,
and a reversible computation, that can be broken down in two temporal directions:

N (x ֌ y) ∪ N (x ֋ y) (4)

The first nexus expresses the result of “travelling” along each curve in C from x to y, the
configuration x appearing before (input) the configuration y (output). The second
nexus expresses the result of the inverse computation where y appears before (input)
and x after (output).
The passage of time along each curve, defined by some temporal parametrization, is
irrelevant with respect to the fundamental temporality expressed in the fact that the
non-oriented curves connecting two objects can be considered conceptually akin to a
reversible computation, and broken down into two oriented curves, each one imple-
menting a computation that is the inverse computation of the other.
Thus, (all) time is not removed from a configuration space definition. Even if one
cannot label a curve with respect to a pre-given time frame, one still finds the more fun-
damental temporality expressed by the notions of input and output of a computation.
In general terms, one can understand the consequence of these results for a spatial
geometry. Any spatial relation, or figure contains the temporality of its spatiality which
is the spatiality necessary for the figure to be configured.
If, in general, one cannot speak of a time independence with respect to a curve in con-
figuration space, what can one state or know about a static quantum state?

14
In principle, a time-independent, essencially static quantum state is timeless. That is
an important argument as one moves from the general configuration space argument to
quantum theory. In this case, the argument is linked to the Wheeler-DeWitt equation,
that effectively introduces a time-independent wave functional for the geometry plus
field configurations, with compact spatial sections (D’Eath, 1996).
Smolin (2001) addressed the general wave functional of the universe in terms of a set
of assumptions that support a general theory of quantum cosmology, under the Wheeler-
DeWitt framework. We can express these main assumptions as follows:
• The configuration space for the universe, associated with some relevant cosmolog-
ical variable that is expressible as a physically observable quantity, is knowable
for technologized theorizing observers inside the universe, that is, there is a
CUniverse that is knowable, from a theoretical and empirical point of view.
• The wave functional of the universe exists and is defined to be a normalizable
complex functionalR ΨUniverse that is normalizable upon an expansion in terms of
CUniverse (that is C 2
dµ|ΨUniverse| = 1) and the normalizable states define a
Universe
space with a Hilbert space structure, with inner product given by: hΦ|Ψi =
dµΦ∗Ψ.
R
C
Universe

• The wave functional of the universe satisfies the momentum constraints along
with the Wheeler-DeWitt equation.
Under this framework, we are effectively working with a time-independent wave func-
tional, and time seems to disappear from the theory.
Smolin (2001) reviewed several arguments against this approach to quantum cos-
mology and against the argument for timelessness, on several grounds, including:
• The limits in the ability to build a formal mathematical theory that incorporates
the observables for quantum cosmology using the above approach, linked with the
problem that the Hamiltonian constraint observables are extremely difficult to
construct in real field theories of gravitation, made more severe by the need to
consider the conjecture regarding the presence of chaotic behavior in gravitational
systems;
• The possible unobservability and unkowability of the configuration space of the
universe, including the problem of the computational complexity for constructing
a mathematical representation of complex configuration spaces.
It is about the limits to our ability to theorize that Smolin’s objections, against the
argument for the absence of time, ultimately rest.
Smolin’s arguments emphasize the role played by the requirement that a theory of
cosmology must be falsifiable in the usual way that ordinary classical and quantum theo-
ries are. This, in turn, leads to the requirement that a sufficient number of observables
can be determined by information that reaches an observer inside the universe, allowing
that observer to know the quantum state of the universe. Only if this is the case can we
build a quantum theory of cosmology, based upon the above set of assumptions, such
that this theory obeys the standard methodological and epistemological scientific
requirements that any scientific theory must obey.
It is around our ability to build a theory of quantum cosmology that these arguments
orbit, however, the arguments presented above do not, by themselves, refute the argu-
ment for the elimination of time, as Smolin (2001) admits it.

15
The problem of time is addressed by Smolin as a human theorization problem, the
temporality being the result of the human inability to access a timeless quantum state of
the universe.
However, before excluding any temporality with respect to a static clock-time inde-
pendent quantum state, we should begin by looking at the mathematical structure of a
Hilbert space, as a geometric space, raising the problem of how one should interpret a
clock-time independent quantum state.
If the universe can be thought of as being a quantum system with a clock-time inde-
pendent quantum state (corresponding mathematically to a clock-time independent
wave functional), then, even if we are unable to address this state, from a theoretical
point of view, we must consider the nature of static clock-time independent quantum
states, and their relation with temporality before arguing about the problem of time.
At this point, one should look in more deeply at the relational structures of the bases
of the Hilbert spaces, as vector spaces, and at the consequences of these relational struc-
tures for the mathematical nature of the space of normalized kets.
Indeed, a Hilbert space can be considered in terms of the above relational structures.
To understand how this is so, let us consider a single Hilbert space H and the set of
alternative bases for this space B, which, for simplicity’s sake, we take to be such that
each basis in B is discrete.
Now, we can build an example of a relational structure that, although being indepen-
dent of any physical clock time, it still produces temporalities that come from temporally
symmetric relational nexus.
Let us, then, consider the relational structure (B, U ), where U is the set of unitary
transformations associated with the change of basis.
Then, we can see that, given any two bases:

B1 = {|φn i}
B2 = {|ψn i}

we can expand each element of the second basis in terms of the elements of the first
basis as:
X
|ψn i = Umn |φm i (5)
m,n

The expansion coefficients Umn can, then, be expressed as:



(6)

φm ψn = Umn

which can be interpreted as the amplitudes of the transitions of the initial ket (input)
|ψn i to the final ket (output) |φm i. Alternatively, we can interpret each |ψn i as the
transformed state of an initial state |φn i under an appropriate unitary transformation.
Indeed, the coefficients Umn form the entries of a unitary matrix that is the matrix rep-
resentation of a unitary operator Û (2, 1) ∈ U, such that:
X
Û (2, 1)|φn i = Umn |φm i = |ψn i (7)
m,n

16
thus:

hφm |Û (2, 1)|φn i = hφm |ψn i = Umn (8)

Since the unitary transformation is reversible we can see that, while Û (2, 1) establishes a
connection from the first to the second basis, its conjugate transpose Û (1, 2) = Û (2, 1)z
establishes the inverse connection. We, therefore, have the relational nexus:

N (B1 ↔ B2) = {Û (2, 1), Û (1, 2)} (9)

which is typical of the reversible logic implemented by a unitary transformation.


One may notice that although no effective temporal physical clock frame was defined,
and although we are dealing with chronologically spatial relations, there is a temporality
expressed in the transition from one basis to another (Eq.(6)), and in the unitary trans-
formation of each basis element of one basis in the basis element of another basis
(Eq.(7)). These relations possess a temporality, independently of the definition of an
actual physical clock, internally defined within some space-time metrics.
This temporal sense “leaks out” to the quantum formalism. Indeed, given the expan-
sion of a quantum state in either of the two basis:
X
|Ψi = cn |ψn i (10)
n

X
|Ψi = cm |φm i (11)
m

we have that:

(12)


X X
Ψ(φm) = hφm |Ψi = cm = hφm |ψn icn = hφm |Û (2, 1)|φn icn (13)
Xn n
′ ′
X
Ψ(ψn) = hψn |Ψi = cn = hψn |φm icm = hψn |Û (1, 2)|ψm icm (14)
m m

Therefore, in Eq.(13) we can see that the amplitudes Ψ(φm) can be considered as a
superposition of quantum computational histories with quantum logical gate Û (2, 1) ini-
tial kets ranging over the basis B1 and final ket |φm i. Each computational history is
weighted by the amplitude cn with the index n ranging over the basis B1, expressing an
uncertainty associated with the input state.
Therefore, we can see that the static expansion expressed by Eq.(11) can be put into a
processual form, where the amplitudes Ψ(φm) correspond to the probability amplitudes
of the output of the computation being |φm i in a quantum computation with general
propagator expressed by hφm |Û (2, 1)|φn i, and where the cn correspond to the ampli-
tudes associated with the input states. A similar reasoning can be applied to the expan-
sion of Eq.(10) and to Eq.(14).
We can also introduce the projector chains:

Pmn = |φm ihφm ||ψn ihψn | (15)



Pnm = |ψn ihψn ||φm ihφm | (16)

17
and write:
X
Ψ(φm) = hφm |Pmn |Ψi (17)
n

X
Ψ(ψn) = hψn |P nm |Ψi (18)
m

These projector chains can be interpreted in light of the projector chains of the consis-
tent histories’ formalism (Omnès, 1988, 1992; Griffiths, 1993, 1994; Gell-Mann and
Hartle 1993, 1994, 1996, 1998; Hartle, 2007), the Pmn representing a history of a
quantum system, since we have:

Pmn = Û (1, 2)|ψm ihψm |Û (1, 2)z|ψn ihψn | = Û (2, 1)z|ψm ihψm |Û (2, 1)|ψn ihψn | (19)

Pnm = Û (2, 1)|φn ihφn |Û (2, 1)z|φm ihφm | = Û (1, 2)z|φn ihφn |Û (1, 2)|φm ihφm | (20)

which completes the connection to the projector chains of the consistent histories’s for-
mulation, since Eqs.(19, 20) can be considered to be analogous to projector chains for
fine-grained histories of a quantum computational network, with the corresponding
quantum logical gates Û (2, 1) and Û (1, 2), respectively5 (Hartle, 2007).
This deepens the result about the temporality expressed by the relation between the
two bases, that can be considered as a quantum computation. The unitary transforma-
tion that implements this computation is such that the elements one of the basis take
the role of inputs and the elements of the other basis take the role of outputs.
In this way, a single time independent state of superposition, can be put into a tem-
poral expression in terms of a set of quantum computational histories, fine-grained with
respect to the quantum logical gate, which is the unitary transformation that imple-
ments the change of basis.
This makes the result more compelling, as it shows how a temporal sense, present in
the relational nexus, that is different from a clock time, may turn a representation of a
clock-time independent quantum state, that represents a single static object, into a pro-
cessual representation, in terms of a quantum computational history, with a formalism
analogous to the one used in the consistent histories’ approach to quantum theory.
Thus, a problem is raised in regards to the statement made with a character of gener-
ality with respect to the timelessness of the quantum state of a system, described by a
stationary quantum state. Such a state indeed should satisfy a clock-time independent
Schrödinger-like equation, and, thus, be effectively timeless with respect to a physical
clock time, however, the mathematical processual nature of such a state, uncovered by
the relational structure of the observables’ eigenbases, precludes the generalization of the
statement to any kind of temporality.
Ultimately, results such as those of Eqs.(13, 14, 19 and 20), raise the more general
issue of the terminology used to address our physical theories of the universe, and, in
particular, the legitimacy of using the concept of state, rather than the term process, to
refer to a ket, which resends to a snapshot-like staticness influenced by a classical
physics’ tradition (Baugh et al., 2003).

5. Looking at (13) and (19) and at (14) and (20) we see that the second projector in each chain is analogous
to a projector in the Heisenberg picture.

18
4. Final reflections and open issues
Considering the general question placed by Barbour (1994): is time a basic concept? The
result of the analysis of arguments and of theoretical discourse developed in the previous
two sections, as well as the mathematical results obtained in the previous section, leads
to an answer to the above question, with another question: what time?
Indeed, the result of the work developed above shows that temporality and the notion
of time go beyond the more restrictive chronometrizable notion of time, that is inter-
nally definable with respect to a space-time geometry.
The arguments of Barbour (1994), of Kauffman and Smolin (1997), and Smolin
(2001), along with the Wheeler-DeWitt equation show how a chronometrizable physical
clock time may indeed not be a basic concept.
However, whenever we consider a quantum computation, when we address a time-
independent ket, or the relations between different physical observables’ eigenbasis, we
find a basic temporality that is definable with respect to the configuration of the rela-
tional nexus of a system of relations between objects.
In physics, as we saw in the previous section, if we accept the fundamental role of a
configuration space, we find this temporality present, even in the absence of any kind of
fundamental clock time.
Therefore, we are led to a conceptual need of defining a relation time, as a funda-
mental (ordinal) time, which is the time of the order of the objects’ positions in the rela-
tion, and that ultimately proceeds from the connection of two individuations that are
separated, but linked by the relation, and, thus, are temporally connected in the tempo-
rality that is the order of terms in the relation.
For relational structures such that, given any two objects X and Y , N (X&Y ) is
either N (X ֌ Y ), N (X ֋ Y ) or N (X ↔ Y ), it is possible to obtain a numeric scale for
the objects that reflects the relation time, by introducing the structure of prevalences in
relational nexus (O, %N ), defined as:

X %N Y if, and only if, N (X&Y ) = N (X ֌ Y ) ∨ N (X&Y ) = N (X ↔ Y )


X ∼ Y if, and only if, N (X&Y ) = N (X ↔ Y )

for any X , Y ∈ O.
This order expresses the relation time that results from the peculiar order of terms in
the relational nexus N (X&Y ). Now, if we were to combine the objects with probabili-
ties, and introduce von Neumann and Morgenstern’s (1953, [1990]) axiomatic for
expected utility, a chronometric could, then, be assigned to the above structures,
reflecting the order %N , which would result in an axiomatic for expected time. Of
course, this is a special case, but it serves to show how a chronometric time may emerge
from a purely relational temporal background .
In this sense, one may be inclined to aggree with the position that a chronometric
time may not be fundamental, and may emerge from a more fundamental temporal
structure that is purely ordinal. The above result is, at least, sufficient to show that an
ordinal temporality, that is a time that emerges within a relational structure, can be
more fundamental from a physical point of view, playing a foundational role in an emer-
gence of a space-time chronometrizable physical time.

19
In this work we have solely dealt with reversible unitary transformations, which leads
to relational nexus of the kind of N (X ↔ Y ). A nexus of this kind is reversible, in the
sense that we can take the temporal connection in either direction. A second kind of
temporality underlies quantum mechanics, and it is specific to that theory’s logical and
mathematical expression.

To understand this we may take the example of the qubit:

|ψ i = ψ(0)|0i + ψ(1)|1i (21)

where the two weights ψ(i), i = 0, 1 are time independent complex numbers satisfying
the normalization condition |ψ(0)|2 + |ψ(1)|2 = 1. In this “static” expansion, we find that
the weights assign amplitudes to the transitions:

h0|ψ i = ψ(0) (22)


h1|ψ i = ψ(1) (23)

which, under Heisenberg’s interpretation of quantum mechanics, determine the proba-


bility of actualization of the input |ψ i to an output of |0i or |1i, mathematically
expressible through a stochastic selection of a projection.

The temporality inherent in this is a clear one, even though the category of causality
does not apply to the context of the notions of dynamis (potentia) and energeia (act),
we know, from these notions, and from the above mathematical expressions of these
notions that the actualized quantum state is preceded by a corresponding potential
reality, upon which it is founded and grounded.

Since the nature of the dynamis, or potentia, is to tend towards the act that deter-
mines it, there is a temporal sense, associated with the “static” expansion of Eq.(21), and
identifiable in the physical interpretation of that equation, within Heisenberg’s interpre-
tation of quantum mechanics, as a mathematical expression of the Aristotelic notions of
dynamis or potentia and energeia or act.

Understanding the process of actualization, if one accepts Heisenberg’s interpretation


of quantum mechanics, leads one to a discussion about decoherence, which can and
should be extended not only to quantum cosmology6 but, also, to the mathematical
structures introduced in section 3..

Another open problem, that may be raised, regards the nature of the relational struc-
tures when, they, themselves are subject to a quantum description. Several different
issues arise in this case. For instance, a quantum causal history is such that, if we label
the edges by quantum states and the nodes by unitary operators, we have that the
incoming and outgoing quantum states to a single node are in a relation, this relation is
locally reversible (due to the unitarity), that is, we have the general relational nexus
associated with the system of node + edges:

n o
N (|ψ i ↔ |ψ ′i) = Û , Û z (24)

6. See, for instance, Kiefer (2003) for an example of such a discussion.

20
However, in the directionality imposed upon the network, only one of the directions is
usually chosen, that is, the local reversible nexus is partioned in terms of:
n o n o
N (|ψi ֌ |ψ ′i) ∪ N (|ψi ֋ |ψ ′i) = Û ∪ Û z (25)

and one of the directions N (|ψi ֌ |ψ ′i) or N (|ψ i ֋ |ψ ′i) is assigned to the network by
the labelling of the node and the choice of the directionality given to the edges7.
Now, this assumes a fixed causal structure, already frozen by a pre-selection of one of
the nexus of the partition. We may, however, consider this selection not to be a given
and assume a quantum superposition of the partitionned nexus, which would lead to the
local network state:

|ΓLoci = Ψ1|N (|ψi ֌ |ψ ′i)i + Ψ2|N (|ψ i ֋ |ψ ′i)i (26)

which incorporates a quantum extension of the theory of binary relational structures,


introduced in the previous section.
This leads to a non-fixed causal structures in the quantum causal histories approach
to quantum cosmology. An issue already raised by Hardy (2007), regarding the quantum
gravity computer.
The matter of how one of the directions gets to be selected, faces us with the problem
of decoherence with respect to the above local state, and the problem of decoherence in
quantum binary relational structures.
In some paths of research, uncertainty regarding the choice of nexus may lead to the
research problem of decoherence and recoherence. In other paths, we may address a
computation that takes place with respect to the whole network which transcends the
local temporal connection, determining it.
Considering this last case, a second order kind of temporal uncertainty is produced in
Eq.(19), as what comes before and after in a computation is not fixed, since the compu-
tation itself is not yet selected. An environmental decoherence mechanism would pro-
duce a diagonal local network state, but this mechanism would be such that it would
compute the entire temporal connection, which means that it would introduce a second
order temporality, that takes the relation itself as an object of another relation.
As an example, let us consider the above local network state and define N1 ≡
N (|ψ i ֌ |ψ ′i) and N2 ≡ N (|ψ i ֋ |ψ ′i).
Then, we have a local basis {|N1i, |N2i}. Adding a local environment, partitioned in
at least two parts8, we can introduce the computation:

|ΓLoc, A0, E0i ֌ Ψ1|N1, A1, E1i + |N2, A2, E2i (27)

implemented by the entanglement operator ÛEnt.

7. We need the two pieces of information (labelling and directionality given to the edges) since in some cases
we may have Û = Û z .
8. This is necessary in order for a unambiguous flow of information from the system to the environment to
take place, as shown and addressed by Paz and Zurek (2002).

21
Thus, considering system + environment we obtain the relational nexus:
n o n o
z
|ΓLoc, A0, E0i ↔ Ψ1|N1, A1, E1i + |N2, A2, E2i = ÛEnt, ÛEnt (28)

Properly considered, this is a reversible process, where ÛEnt produces entanglement and
z
ÛEnt transforms the entangled state9 into an unentangled state10. However, for a large
enough environment, as discussed by Zeh (2002), the entanglement can be considered to
be almost irreversible, which means that, with respect to the local network we would
have the transition given by the following nexus, expressed in terms of density opera-
tors:

N |ΓLoc ΓLoc| ֌ |Ψ1|2|N1, A1 N1, A1| + |Ψ2|2|N2, A2 N2, A2| (29)





with the environment described by the kets |Ai i playing the role analogous to a record
keeping physical apparatus.
It is noticeable that we are dealing with a quantum computation that computes an
entire local structure of quantum computations, and that the final relational nexus is the
nexus of the quantum states for the relational nexus of a local quantum computation.
Indeed, the result of Eq.(27)
 corresponds to the case where the computation expressed
′ 2
by the nexus N (|ψ i ֌ |ψ i occurs, with probability |Ψ1| , or the reverse computation,
expressed by the nexus N (|ψ i ֋ |ψ ′i) occurs, with probability |Ψ2|2.
The generalization of this type of research to quantum causal histories, and to their
applications in quantum gravity research, may be important to understand the nature
and role of time in the theory of quantum space-time, and, in particular, in the research
program of loop quantum gravity.

9. Producing local decoherence.


10. Producing a recoherence.

22
Appendix
We use, throughout this appendix, the notation and notions addressed in the main text,
including the notions of logical intension and relational nexus.
Stated in a general sense, a relational structure is composed of a collection of objects
and a collection of n-ary relations defined in terms of their intensions.
A category is a particular structure, within a binary relational structure (O, R), where
the set of relations is restricted to morphisms between objects, satisfying a number of
conditions.
The first condition is that each morphism f in R relates a pair of objects in a direc-
tional way, that is, f ∈ N (X ֌ Y ), where X is known as the domain of f , or source
object and Y as the codomain, or target object, and one writes:

f: X Y
The relational nexus N (X ֌ Y ), thus, corresponds to the class of all morphisms from X
to Y . By definition 1., in the main text, we know that every morphism between any
two objects must be in R. Also, we have that there has to be an identity morphism for
each object, corresponding to the identity relation, evaluated for each object:

id: X X
The second condition is that R be closed under composition of morphisms, defined such
that, for three objects X , Y , Z ∈ O, and two morphisms f and g, f ∈ N (X ֌ Y ) and
g ∈ N (Y ֌ Z), then, we have that the composition g ◦ f is such that:

g ◦ f ∈ N (X ֌ Z)

and:

g ◦ f = g : (f : X  Y ) Z
That is, the morphism g ◦ f from X to Z is first obtained by applying the morphism f
from X to Y , and, then, the morphism g from Y to Z .
The third condition imposes that composition of morphims is associative and the
fourth that the composition is commutative with respect to the identity. A structure
satisfying these conditions is called a category.
It follows from these results that a mathematical category is a binary relational sub-
structure, in the sense that the collection of relations is restricted to the collection of
morphisms.
A binary relational structure is more general than a category, and contains in it struc-
tures that are more general than categories. Indeed, we could simply consider a binary
relational substructure composed of a collection of objects and a collection of mor-
phisms, with the identity morphism included, these being the only definable restrictions.
Any mathematical category is such a structure, but the fact that nothing is stated about
composition means that not every collection of objects and morphisms, with the identity
morphism included, constitutes a category, nonetheless, it constitutes a binary relational
structure.

23
Examples of binary relational substructures that are categories include the category of
binary relations Rel, which has sets by objects and, by morphisms, the binary relations,
that are defined as subsets of the set of ordered pairs A × B, for any two sets.
It is important to notice that the structure composed of the truth sets of a binary
relational structure, along with the collection of objects of that structure, is, trivially, a
member of Rel, since each truth set is a set of ordered pairs, that can be expressed as
the Cartesian product of two subsets of the collection of objects.
The structure Mag is another example of a binary relational substructure that is a
category, having, magmas by objects, that is, algebraic structures composed of sets with
a binary operation, and morphisms given by homomorphisms of operations. It is useful,
for illustration purposes, to address this structure.
First, we notice that a magma is an algebraic structure that consists of a set A
equipped with a single binary operation:

ιA: A × A A
where we write, for any a, b ∈ A:

ιA: (a, b) ∈ A × A  aιAb ∈ A


Thus, we have the magma (A, ιA).
Now, it becomes important to consider the following four definitions, from universal
algebra (Burris and Sankappanavar, 1981):

Definition 2. For a nonempty set A, and a nonnegative integer n, we define A0 = ∅ ,




and, for n > 0, An is the set of n-tuples of elements from A. An n-ary operation (or
function) on A is any function F from An to A; n is the arity (or rank) of F.

Definition 3. A finitary operation is an n-ary operation, for some n. The image of the
n-tuple (a1,  , an), under an n-ary operation F, is denoted by F (a1,  , an).

Definition 4. An operation F on A is called a nullary operation (or constant) if its


arity is zero, being completely determined by the image F (∅) in A of the only element ∅
in A0 , and, as such, one can identify it with the element F (∅). Thus, a nullary opera-
tion is thought of as an element of A.

Taking into account these definitions we can define a language or type of algebras as
(Burris and Sankappanavar, 1981):

Definition 5. A language (or type) of algebras is a set F of function symbols such that
a nonnegative integer n is assigned to each member F of F. This integer is called the
arity (or rank) of F, and F is said to be an n-ary function symbol. The subset of n-ary
function symbols in F is denoted by Fn.

From this last definition, the general definition of an algebra, within universal algebra
is given by (Burris and Sankappanavar, 1981):

24
Definition 6. If F is a language of algebras, then, an algebra a of type F is a structure
(A, F ), where A is a nonempty set and F is a family of finitary operations on A indexed
by the language F, such that, corresponding to each n-ary function symbol F in F, there
is an n-ary operation FA on A. The set A is called the universe (or underlying set) of
the algebra a = (A, F ), and the FA’s are called the fundamental operations of a.

If F is finite, such that F = {F1,  , Fn }, we write (A, F1,  , Fn) for the algebra a =
(A, F ), with the convention that arity F1 ≥ arity F2 ≥  ≥ arity Fn.

It is straightforward to see that a magma (A, ιA), with A nonempty, is an algebra in


the above sense, that is, it is an algebra with a family of binary operations indexed by a
singleton language F = {ι}.

The binary relational structure of magmas MAG = (M, RM) has by collection of
objects the magmas.

Let us, then, consider the subset of the set of relations between magmas in MAG
comprised of the set of homomorphisms, defined as follows:

Definition 7. Given two magmas (A, ιA) and (B , ιB ) the homomorphism from (A, ιA)

to (B , ιB ) is defined as the mapping f : A B satisfying:

f (aιAb) = f (a)ιBf (b)

Thus, a homomorphism between two magmas is a mapping that preserves the binary
operation. It is noticeable that, under this restriction, one is no longer working with the
whole relational structure MAG, since the set of relations between magmas is restricted
to operation preserving mappings. The structure within MAG, with which we are
working, contains the same collection of objects as MAG and works with the subset of
homomorphisms for the set of relations. Such a structure is the category Mag.

This example not only shows how the binary relational structures are more general
than the mathematical categories, it also helps to understand how a category is a sub-
structure within the binary relational structure.

It is important to consider the relation between the truth sets of the relations in
MAG and Rel, in connection with Mag.

The relation between MAG and Rel can be established through the generality of the
concept of set, which can be defined as any collection of objects. Taking this into
account, we can indeed consider the collection of objects of MAG as a set. On the other
hand, any binary relation R defined over pairs of magmas has, by truth set, the set of
ordered pairs of magmas that exemplify it.

Since each relation in MAG is defined with respect to its intension, and not with
respect to its extension, two relations that correspond to different binary properties are
not identical, however, they may have the same truth set.

We can, therefore, define the model of RM, A(RM) as the collection of truth sets for
the relations in RM. From the previous paragraph, it follows that the mapping from
RM to A(RM) is onto, but not one-to-one.

25
Given the general notion of set, it is clear that A(RM) is a class of sets of ordered
pairs of magmas, each such set being a subset of the power set M2 = M × M. Which
means that A(RM) can be put in correspondance with a class of morphisms in Rel
defined by the general condition:

f: M  M′
where M and M ′ are sets of magmas in M.
The truth sets corresponding to the homomorphisms in Mag is, thus, a subclass of
A(RM), which can be put into correspondence with a class of morphisms in Rel,
through the above scheme.

26
Bibliography
1. Barbour, J.B. 1994. The Emergence of Time and Its Arrow from Timelessness.
In: The Physical Origins of Time Asymmetry, J.J. Halliwell, J. Perez-Mercader,
and W.H. Zurek (eds.). Cambridge University Press.

2. Barbour, J.B. 2003. Dynamics of pure shape, relativity and the problem of time.
In: Decoherence and Entropy in Complex Systems (Proceedings of the Conference
DICE, Piombino 2002), H.-T Elze. Springer. arXiv:gr-qc/0309089.

3. Baugh, J.; Finkelstein, D.R.; Galiautdinov, A. 2003. The Qubits of Qunivac. Int.
J. Theor. Phys., 177-187. arXiv:hep-th/0206036v1.

4. Burris, Stanley; Sankappanavar, H.P. 1981. A Course in Universal Algebra. The


Millenium Edition. http://www.math.uwaterloo.ca/~snburris/htdocs/UALG/
univ-algebra.pdf.

5. D’Eath, P.D. 2005. Supersymmetric Quantum Cosmology. Cambridge University


Press.

6. Einstein, Albert. 1953. [2004]. The Meaning of Relativity, Fifth Edition:


Including the Relativistic Theory of the Non-Symmetric Field . Princeton Univer-
sity Press.

7. Gell-Mann, Murray; Hartle, James. 1993. Classical Equations for Quantum Sys-
tems, Phys.Rev. D47, 3345-3382.

8. Gell-Mann, Murray; Hartle, James. 1994. Time Symmetry and Asymmetry in


Quantum Mechanics and Quantum Cosmology. In: Physical Origins of Time
Asymmetry, J. Halliwell, J. Perez-Mercader, and W. Zurek (eds.), Cambridge
University Press.

9. Gell-Mann, Murray; Hartle, James. 1996. Equivalent Sets of Histories and Mul-
tiple Quasiclassical Realms, gr-qc/9404013.

10. Gell-Mann, Murray; Hartle, James. 1998. Strong Decoherence, in “Proceedings of


the 4th Drexel Conference on Quantum Non-Integrability: The Quantum-Clas-
sical Correspondence”, D.-H. Feng and B.-L. Hu (eds.). International Press of
Boston.

11. Griffiths, R. B. 1984. Constistent Histories and the Interpretation of Quantum


Mechanics. J. Stat. Phys. 36, 219–272.

12. Griffiths, R. B. 1993. Consistent interpretation of quantum mechanics using


quantum trajectories. Phys. Rev. Lett. 70, 219–272.

13. Griffiths, R. B. 1994. A consistent history approach to the logic of quantum


mechanics. In: Symposium on the Foundations of Modern Physics, K.V. Lau-
rikainen, C. Montonen and K. Sunnarborg (eds.). Éditions Frontières.

27
14. Hardy, Lucien. 2007. Quantum gravity computers: On the theory of computation
with indefinite causal structure. For Proceedings of Quantum Reality, Relativistic
Causality, and Closing the Epistemic Circle: An International Conference in
Honour of Abner Shimony. arXiv:quant-ph/0701019v1.

15. Hartle, James B.; Hawking, Stephen. 1983. Wave Function of the Universe.
Phys. Rev. D, 28, 12, 2960–2975.

16. Hartle, James B. 2003. Theories of Everything and Hawking’s Wave Function of
the Universe. In: The Future of Theoretical Physics and Cosmology, G. W.
Gibons, E.P.S. Shellard and S.J. Rankin (eds.). Cambridge University Press.
arXiv:gr-qc/0209047.

17. Hartle, James B. 2004. Linear Positivity and Virtual Probability. Phys. Rev. A,
70, 02210. arXiv:quant-ph/0401108.

18. Hartle, James B. 2007. Quantym Physics and Human Language. J.Phys. A40
(2007) 3101-3121. arXiv:quant-ph/0610131v3.

19. Hawking, Stephen. 1984. The Quantum State of the Universe. Nucl. Phys.,
B239, 257–276.

20. Hawkins, Eli; Markopoulou, Fotini; Sahlmann, Hanno. 2003. Evolution in


Quantum Causal Histories. Class. Quant. Grav. 20, 3839. hep-th/0302111.

21. Heisenberg, W. 1961. Planck’s Discovery and the Philosophical Problems of


Atomic Physics. In: On Modern Physics, C. N. Polter (ed.). Orion Press.

22. Kiefer, C. 2003. Decoherence in Quantum Field Theory and Quantum Gravity.
In: Decoherence and the Appearance of a Classical World in Quantum Theory, E.
Joos; H.D. Zeh; C. Kiefer; D. Giulini; J. Kupsch and I.-O. Stamatescu (eds.).
Springer.

23. Klein, Étienne. 1995. Le Temps. Flammarion.

24. Lloyd, Seth. 2007. Programming the Universe. Vintage Books.

25. Kauffman, Stuart; Smolin, Lee. 1997. A possible solution to the problem of time
in quantum cosmology. gr-qc/9703026.

26. Madeira, Maria Odete. 2008a. What is the time? . http://cmath-


phil.blogspot.com/2008/02/what-is-time.html

27. Madeira, Maria Odete. 2008b. Time and Relativity. http://cmath-


phil.blogspot.com/2008/03/time-and-relativity.html.

28. Markopoulou, Fotini. 2000. An insider’s guide to quantum causal histories. Nucl.
Phys. Proc. Suppl. 88, 308–31. hep-th/9912137.

29. Markopoulou, Fotini. 2002. Planck-scale models of the Universe. gr-qc/0210086.

30. Omnès, R. 1988. Logical reformulation of quantum mechanics I − III, J. Stat.


Phys. 53, 893 − 975.

28
31. Omnès, R. 1992. Consistent interpretation of quantum mechanics, Rev. Mod.
Phys. 64, 339 − 382.

32. Paz, J.P.; Zurek, W.H. 2002. Environment-Induced Decoherence and the Transi-
tion from Quantum to Classical . In: Fundamentals of Quantum Information –
Quantum Computation, Communication, Decoherence and All That, Dieter Heiss
(ed.). Springer.

33. Smolin, Lee. 2001. The present moment in quantum cosmology: Challenges to the
argument for the elimination of time. gr-qc/0104097.

34. Smolin, Lee. 2003. Three Roads to Quantum Gravity. Phoenix.

35. Smolin, Lee. 2006. The Trouble With Physics. Penguin Books.

36. Von Neumann, John; Morgenstern, Oskar, 1953, [1990], The Theory of Games
and Economic Behavior, Princeton University Press, US.

37. Weyl, Hermann. 1952. Space, time and matter . Translated by Henry L. Brose.
Dover Publications.

38. Wheeler, J. A. 1968. Superspace and the Nature of Quantum Geometrodynamics.


W.A. Benjamin, Inc. In: Quantum Cosmology, Li Zhi Fang and Remo Ruffini
(eds.). World Scientific.

39. Zeh, H. D. 2002. The Wave Function: It or Bit? . In: Science and Ultimate
Reality, J. D. Barrow, P. C. W. Davies, and C. L. Harper Jr. (eds.), Cambridge
University Press.

29

You might also like