Topos Methods in The Foundations of Physics: Chris J. Isham April, 2010
Topos Methods in The Foundations of Physics: Chris J. Isham April, 2010
Chris J. Isham†
April, 2010
1 Introduction
Over forty years have passed since I first became interested in the problem
of quantum gravity. During that time there have been many diversions and,
perhaps, some advances. Certainly, the naively-optimistic approaches have
long been laid to rest, and the schemes that remain have achieved some degree
of stability. The original ‘canonical’ programme evolved into loop quantum
gravity, which has become one of the two major approaches. The other, of
course, is string theory—a scheme whose roots lie in the old Veneziano model
of hadronic interactions, but whose true value became apparent only after it
had been re-conceived as a theory of quantum gravity.
However, notwithstanding these hard-won developments, there are certain
issues in quantum gravity that transcend any of the current schemes. These
involve deep problems of both a mathematical and a philosophical kind, and
stem from a fundamental paradigm clash between general relativity—the
apotheosis of classical physics—and quantum physics.
In general relativity, space-time ‘itself’ is modelled by a differentiable
manifold M: a set whose elements are interpreted as ‘space-time points’.
The curvature tensor of the pseudo-Riemannian metric on M is then deemed
to represent the gravitational field. As a classical theory, the underlying
philosophical interpretation is realist: both the space-time and its points
truly ‘exist’1 , as does the gravitational field.
∗
Based on joint work with Andreas Döring.
†
c.isham@imperial.ac.uk
1
At least, that would be the view of unreconstructed, space-time substantivalists. How-
ever, even purely within the realm of classical physics this position has often been chal-
lenged, particularly by those who place emphasis on the relational features that are inher-
ent in general relativity.
1
On the other hand, standard quantum theory employs a background
space-time that is fixed ab initio in regard to both its differential structure
and its metric/curvature. Furthermore, the conventional interpretation is
thoroughly instrumentalist in nature, dealing as it does with counter-factual
statements about what would happen (or, to be more precise, the probability
of what would happen) if a measurement is made of some physical quantity.
In regard to quantum gravity, the immediate question is:
2
• Would we recognise the/a ‘correct’ theory of quantum gravity even if
it was handed to us on a plate? Certainly, the Popperian notion of
refutation is hard to apply with such a sparsity of empirical data.
• What makes any particular idea a ‘good’ one as far as the community is
concerned? Relatedly, how is it that one particular research programme
becomes well established whereas another falls by the wayside or, at
best, gains only a relatively small following?
The second question is not just whimsical, especially for our younger col-
leagues, for it plays a key role in decisions about the award of research grants,
post-doc positions, promotion, and the like.
In practice, many of the past and present research programmes in quan-
tum gravity have been developed by analogy with other theories and the-
oretical structures, particularly standard quantum field theory with gauge
groups. And as for why it is that certain ideas survive and other do not, the
answer is partly that individual scientists indulge in their own philosophical
prejudices, and partly that they like using theoretical tools with which they
are already familiar. This, after all, is the fastest route to writing new papers.
At this point it seems appropriate to mention the ubiquitous, oft-maligned
‘Planck length’. This fundamental unit of length comes from combining
Planck’s constant ~ (the ‘quantum’ in quantum gravity) with Newton’s con-
stant G (the ‘gravity’ in quantum gravity) and the speed of light c (which
always lurks around) in the form
1/2
G~
LP := (1.1)
c3
which has a value of around 10−35 meters; the corresponding Planck time
(defined as TP := LP /c) has a value or around 10−42 seconds.
The general assumption is that something ‘dramatic’ happens to the na-
ture of space and time at these fundamental scales. Precisely what that
dramatic change might be has been the source of endless speculation and
conjecture. However, there is a fairly wide-spread anticipation that in so far
as spatio-temporal concepts have any meaning at all in the ‘deep’ quantum-
gravity regime, the appropriate mathematical model will not be based on
standard, continuum differential geometry. Indeed, it is not hard to con-
vince oneself that, from a physical perspective, the important ingredient in a
space-time model is not the ‘points’ in that space, but rather the ‘regions’ in
which physical entities can reside. In the context of a topological space, such
3
regions are best modelled by open sets: the closed sets may be too ‘thin’2
to contain a physical entity, and the only physically-meaningful closed sets
are those with a non-empty interior. These reflections lead naturally to the
subject of ‘pointless topology’ and the theory of locales—a natural step along
the road to topos theory.
However, another frequent conjecture is that what we normally call space
and time (or space-time) will only ‘emerge’ from the correct quantum gravity
formalism in some (classical?) limit. Thus a fundamental theory of quantum
gravity may (i) have no intrinsic reference at all to spatio-temporal concepts;
and (ii) be such that some of the spatio-temporal concepts that emerge in
various limits are non-standard and are modelled mathematically with some-
thing other than topology and differential geometry. All this leads naturally
to the main question that lies behind the work reported in this chapter.
Namely:
4
2 The Problem of Using the Real Numbers
in Quantum Gravity
The real numbers arise in theories of physics in three different (but related)
ways: (i) as the values of physical quantities; (ii) as the values of probabilities;
and (iii) as a fundamental ingredient in models of space and time (especially
in those based on differential geometry). All three are of direct concern vis-
a-vis our worries about making unjustified, a priori assumptions in quantum
theory. Let us consider them in turn.
One reason for assuming physical quantities to be real-valued is undoubt-
edly grounded in the remark that, traditionally (i.e., in the pre-digital age),
they are measured with rulers and pointers, or they are defined operationally
in terms of such measurements. However, rulers and pointers are taken to be
classical objects that exist in the physical space of classical physics, and this
space is modelled using the reals. In this sense there is a direct link between
the space in which physical quantities take their values (what we call the
‘quantity-value space’) and the nature of physical space or space-time [7].
Of course, real numbers also arise as the value space for probabilities via
the relative-frequency interpretation of probability. Thus, in principle, an
experiment is to be repeated a large number, N, times, and the probability
associated with a particular result is defined to be the ratio Ni /N, where Ni
is the number of experiments in which that particular result was obtained.
The rational numbers Ni /N necessarily lie between 0 and 1, and if the limit
N → ∞ is taken—as is appropriate for a hypothetical ‘infinite ensemble’—
real numbers in the closed interval [0, 1] are obtained.
The relative-frequency interpretation of probability is natural in instru-
mentalist theories of physics, but it is inapplicable in the absence of any
classical spatio-temporal background in which the necessary sequence of mea-
surements can be made (as, for example, is the situation in quantum cosmol-
ogy).
In the absence of a relativity-frequency interpretation, the concept of
‘probability’ must be understood in a different way. One possibility involves
the concept of ‘potentiality’, or ‘latency’. In this case there is no compelling
5
reason why the probability-value space should be a subset of the real num-
bers. The minimal requirement on this value-space is that it is an ordered
set, so that one proposition can be said to be more or less probable than
another. However, there is no prima facie reason why this set should be
totally ordered.
It follows from the above that, for us, a key problem is the way in which
the formalism of standard quantum theory is firmly grounded in the con-
cepts of Newtonian space and time (or the relativistic extensions of these
ideas) which are essentially assumed a priori. The big question is how this
formalism can be modified/generalised/replaced so as to give a framework
for physical theories that is
6
3 Theories of a Physical System
3.1 The Realism of Classical Physics
“From the range of the basic questions of metaphysics we shall
here ask this one question: What is a thing? The question is
quite old. What remains ever new about it is merely that it must
be asked again and again [8].”
Martin Heidegger
Asking such questions is not a good way for a modern young philosopher
to gain tenure. Plato was not ashamed to do so, and neither was Kant, but
at some point in the last century the question became ‘Wittgensteined’ and
since then it is asked at one’s peril. Fortunately, theoretical physicists are
not confined in this way, and I will address the issue front on. Indeed, why
should a physicist be ashamed of this margaritiferous question? For is it not
what we all strive to answer when we probe the physical world?
Heidegger’s own response makes interesting reading:
Let us see how modern physics approaches this fundamental question about
the beingness of Being.
In constructing a theory of any ‘normal’4 branch of physics, the key in-
gredients are the mathematical representations of the following:
7
In the case of standard quantum theory, the mathematical states are
interpreted in a non-realist sense as specifying only what would happen if
certain measurements are made: so the phrase ‘the way things are’ has to be
broadened to include this view. This ontological concept must also apply to
the ‘truth values’ that are intrinsically probabilistic in nature.
In addition, ‘the way things are’ is normally construed as referring to a
specific ‘moment’ of time, but this could be generalised to be compatible
with whatever model of time/space-time is being employed. This could even
be extended to a ‘history theory’, in which case ‘the way things are’ means
the way things are for all moments of time, or whatever is appropriate for
the space-time model being employed.
As theoretical physicists with a philosophical bent, a key issue is how
such a mathematical framework implies, or encompasses, various possible
philosophical positions. In particular, how does it interface with the position
of ‘realism’ ?
This issue is made crystal clear in the case of classical physics, which is
represented in the following way.
4. The basic propositions are of the form “A ε ∆” which asserts that the
value of the physical quantity A lies in the (Borel) subset ∆ of the
real numbers. This proposition is represented mathematically by the
subset Ă−1 (∆) ⊆ S—i.e., the collection of all states, s, in S such that
Ă(s) ∈ ∆.
8
The mathematical structure of set theory implies that, of
necessity, the propositions in classical physics have the logical
structure of a Boolean algebra.
• A is represented by a function Ă : S → R.
9
• A is represented by a self-adjoint operator, Â, on H.
10
category, known as a topos, in which the sub-objects of any object do have a
logical structure—a remark that underpins our entire research programme.
Broadly speaking, a topos, τ , is a category that behaves in certain critical
respects just like the category, Sets, of sets. In particular, these include the
following:
2. One can form ‘products’ and ‘co-products’ of objects8 These are the
analogue of the cartesian product and disjoint union in set theory.
for all objects C. In set theory, the relevant statement is that a pa-
rameterised family of functions c 7→ fc : A → B, c ∈ C, is equivalent
to a single function F : C × A → B defined by F (c, a) := fc (a) for all
c ∈ C, a ∈ A.
11
‘true’ respectively, so that χA (x) = 1 corresponds to the mathematical
proposition “x ∈ A” being true.
This operation is mirrored in any topos. More precisely, there is a, so-
called, ‘sub-object classifier’, Ωτ , that is the analogue of the set {0, 1}.
Specifically, to each sub-object A of an object X there is associated a
‘characteristic arrow’ χA : X → Ωτ ; conversely, each arrow χ : X → Ωτ
determines a unique sub-object of X. Thus
Sub(X) ≃ Homτ (X, Ωτ ) (4.2)
where Sub(X) denotes the collection of all sub-objects of X.
5. The power set, P X of any set X is defined to be the set of all subsets
of X. Each subset A ⊆ X determines, and is uniquely determined by,
its characteristic function χA : X 7→ {0, 1}. Thus the set P X is in
bijective correspondence with the function space {0, 1}X . Analogously,
in a general topos, τ , we define the power object of an object X, to be
P X := ΩX τ . It follows from (4.1) that
12
4.2 The Mathematics of ‘Neo-realism’
Let us now consider the assignment of truth values to propositions in math-
ematics. In set theory, let K ⊆ X be a subset of some set X, and let x be an
element of X. Then the basic mathematical proposition “x ∈ K” is true if,
and only if, x is an element of the subset K. At the risk of seeming pedantic,
the truth value, [[ x ∈ K ]], of this proposition can be written as
1 if x belongs to K;
[[ x ∈ K ]] := (4.6)
0 otherwise.
[[ x ∈ K ]] = χK ◦ x (4.7)
[[ x ∈ K ]] := χK ◦ x : 1τ → Ωτ (4.8)
13
2. Propositions about S are represented by elements of the Heyting alge-
bra, Sub(Σ), of sub-objects of Σ. If Q is such a proposition, we denote
by δ(Q) the associated sub-object of Σ; the ‘name’ of δ(Q) is the arrow
pδ(Q)q : 1τ → P Σ.
3. The topos analogue of a state is a ‘pseudo-state’ which is a particular
type of sub-object of Σ. Given this pseudo-state, each proposition can
be assigned a truth value in the Heyting algebra ΓΩτ . Equivalently, we
can use ‘truth objects’: see below.
Conceptually, such a theory is neo-realist in the sense that the propositions
and their truth values belong to structures that are ‘almost’ Boolean: in fact
they differ from Boolean algebras only in so far as the principle of excluded
middle may not apply.
Thus a theory expressed in this way ‘looks like’ classical physics ex-
cept that classical physics always employs the topos Sets, whereas other
theories—including, we claim, quantum theory—use a different topos. If the
theory requires a background space-time (or functional equivalent thereof)
this could be represented by another special object, M, in the topos. It
would even be possible to mimic the actions of differential calculus if the
topos is such as to support synthetic differential geometry10 .
The presence of intrinsic logical structures in a topos has another striking
implication. Namely, a topos can be used as a foundation for mathematics
itself, just as set theory is used in the foundations of ‘normal’ (or ‘classical’
mathematics). Thus classical physics is modelled in the topos of sets, and
thereby by standard mathematics. But a theory of physics modelled in a
topos, τ , other than Sets is being represented in an alternative mathematical
universe! The absence of excluded middle means that proofs by contradiction
cannot be used, but apart from that this, so-called, ‘intuitionistic’ logic can
be handled in the same way as classical logic.
A closely-related feature is that each topos has an ‘internal language’
that is functionally similar to the formal language on which set theory is
based. It is this internal language that is used in formulating axioms for the
mathematical universe associated with the topos. The same language is also
used in constructing the neo-realist interpretation of the physical theory.
14
of physical propositions. A key step in constructing a physical theory is to
translate the latter into the former. For example, in classical physics, the
physical proposition “A ε ∆” is represented by the subset Ă−1 (∆) of the state
space, S; i.e.,
δ(A ε ∆) := Ă−1 (∆). (4.9)
Then, for any state s ∈ S, the truth value of the physical proposition
“A ε ∆” is defined to be the truth value of the mathematical proposition
“s ∈ δ(A ε ∆)” (or, equivalently, the truth value of “Ă(s) ∈ ∆”).
Given the ideas discussed above, in a topos theory one might expect to
represent a physical state by a global element s : 1τ → Σ of the state object
Σ. The truth value of a proposition, Q, represented by a sub-object δ(Q) of
Σ, would then be defined as the global element
1. K ∈ Ts (4.11)
T s := {J ⊆ S | s ∈ J}; (4.13)
and
2. {s} ⊆ K. (4.14)
15
Option 1 (the truth-object option): Note that T s is a collection of
subsets of S: i.e., T s ∈ Sub(P S). The interesting thing about (4.11) is that
it is of the form “x ∈ X” and therefore has an immediate generalisation to
any topos. More precisely, in a general topos τ a truth object, T , would
be a sub-object of P Σ (equivalently, a global element of P (P Σ)) with a
characteristic arrow χT : P Σ → Ωτ . Then the physical proposition Q has
the topos truth value
The key remark is that although (4.10) is inapplicable if there are no global
elements of Σ, (4.15) can be used since global elements of a power object
(like P (P Σ)) always exist. Thus, in option 1, the analogue of a classical
state s ∈ S is played by the truth object T ∈ Sub(P Σ).
16
2. To each physical proposition, Q, there is associated a sub-object, δ(Q),
of Σ.
Note that no mention has been made here of the quantity-value object
R. However, in practice, this also is expected to play a key role, not least
in constructing the physical propositions. More precisely, we anticipate that
each physical quantity A will be represented by an arrow Ă : Σ → R, and
then a typical proposition will be of the form “A ε Ξ” where Ξ is some sub-
object of R.
If any ‘normal’ physics is addressed in this way, physical quantities are
expected to be real -valued, and the physical propositions are of the form
“A ε ∆” for some ∆ ⊆ R. In this case, it is necessary to decide what sub-
object of R in the topos corresponds to the external quantity ∆ ⊆ R.
There is subtlety here, however, since although (in any topos we are likely
to consider) there is a precise topos analogue of the real numbers12 , it would
be a mistake to assume that this is necessarily the quantity-value object:
indeed, in our topos version of quantum theory this is definitely not the
case.
The question of relating ∆ ⊆ R to some sub-object of the quantity-value
object R in τ is just one aspect of the more general issue of distinguishing
quantities that are external to the topos, and those that are internal. Thus,
for example, a sub-object Ξ of R in τ is an internal concept, whereas ∆ ⊆ R
is external, referring as it does to something, R, that lies outside τ . Any
reference to a background space, time or space-time would also be exter-
nal. Ultimately perhaps—or, at least, certainly in the context of quantum
cosmology—one would want to have no external quantities at all. However,
in any more ‘normal’ branch of physics it is natural that the propositions
about the system refer to the external (to the theory) world to which the-
oretical physics is meant to apply. And, even in quantum cosmology, the
actual collection of what counts as physical quantities is external to the for-
malism itself.
12
Albeit defined using the analogue of Dedekind cuts, not Cauchy sequences.
17
5 The Topos of Quantum Theory
5.1 The Kochen-Specker theorem and contextuality
To motivate our choice of topos for quantum theory let us return again to the
Kochen-Specker theorem which asserts the impossibility of assigning values
to all physical quantities whilst, at the same time, preserving the functional
relations between them [14].
In a quantum theory, a physical quantity A is represented by a self-adjoint
operator  on the Hilbert space, H, of the system. A ‘valuation’ is defined to
be a real-valued function λ on the set of all bounded, self-adjoint operators,
with the properties that: (i) the value λ(Â) belongs to the spectrum of Â;
and (ii) the functional composition principle (or FUNC for short) holds:
for any pair of self-adjoint operators Â, B̂ such that B̂ = h(Â) for some
real-valued function h.
Several important results follow from this definition. For example, if Â1
and Â2 commute, it follows from the spectral theorem that there exists an
operator Ĉ and functions h1 and h2 such that Â1 = h1 (Ĉ) and Â2 = h2 (Ĉ).
It then follows from FUNC that
and
λ(Â1 Â2 ) = λ(Â1 )λ(Â2 ). (5.3)
The Kochen-Specker theorem says that no valuations exist if dim(H) > 2.
On the other hand, (5.2–5.3) show that, if it existed, a valuation restricted
to a commutative sub-algebra of operators would be just an element of the
spectrum of the algebra, and of course such elements do exist. Thus valu-
ations exist on any commutative sub-algebra13 of operators, but not on the
(non-commutative) algebra, B(H), of all bounded operators. We shall call
such valuations ‘local’.
Within the instrumentalist interpretation of quantum theory, the exis-
tence of local valuations is closely related to the possibility of making ‘simul-
taneous’ measurements on commutating observables. However, the existence
of local valuations also plays a key role in the, so-called, ‘modal’ interpre-
tations in which values are given to the physical quantities that belong to
13
More precisely, since we want to include projection operators, we assume that the
commutative algebras are von Neumann algebras. These algebras are defined over the
complex numbers, so that non self-adjoint operators are included too.
18
some specific commuting set. The most famous such interpretation is that
of David Bohm where it is the configuration14 variables in the system that
are regarded as always ‘existing’ (in the sense of possessing values).
The topos-implication of these remarks stems from the following obser-
vations. First, let V, W be a pair of commutative sub-algebras with V ⊆ W .
Then any (local) valuation, λ, on W restricts to give a valuation on the sub-
algebra V . More formally, if ΣW denotes the set of all local valuations on
W , there is a ‘restriction map’ rW V : ΣW → ΣV in which rW V (λ) := λ|V for
all λ ∈ ΣW . It is clear that if U ⊆ V ⊆ W then
rV U (rW V (λ)) = rW U (λ) (5.4)
for all λ ∈ ΣW —i.e., restricting from W to U is the same as going from W
to V and then from V to U.
Note that, if one existed, a valuation, λ, on the non-commutative algebra
of all operators on H would provide an association of a ‘local’ valuation
λV := λ|V to each commutative algebra V such that, for all pairs V, W with
V ⊆ W we have
λW | V = λV (5.5)
The Kochen-Specker theorem asserts there are no such associations V 7→
λV ∈ ΣV if dim H > 2.
To explore this further consider the situation where we have three com-
muting algebras V, W1 , W2 with V ⊆ W1 and V ⊆ W2 and suppose λ1 ∈ ΣW1
and λ2 ∈ ΣW2 are local valuations. If there is some commuting algebra W
so that (i) W1 ⊆ W and W2 ⊆ W ; and (ii) there exists λ ∈ ΣW such that
λ1 = λ|W1 and λ2 = λ|W2 , then (5.4) implies that
λ1 | V = λ2 | V (5.6)
However, suppose now the elements of W1 and W2 do not all commute
with each other: i.e., there is no W such that W1 ⊆ W and W2 ⊆ W . Then
although valuations λ1 ∈ ΣW1 and λ2 ∈ ΣW2 certainly exist, there is no longer
any guarantee that they can be chosen to satisfy the matching condition in
(5.6). Indeed, the Kochen-Specker theorem says precisely that it is impos-
sible to construct a collection of local valuations, λW , for all commutative
sub-algebras W such that all the matching conditions of the form (5.6) are
satisfied.
We note that triples V, W1 , W2 of this type arise when there are non-
commuting self-adjoint operators Â1 , Â2 with a third operator B̂ and func-
tions f1 , f2 : R → R such that B̂ = f1 (Â1 ) = f2 (Â2 ). Of course, [B̂, Â1 ] = 0
14
If taken literally, the word ‘configuration’ implies that the state space of the underlying
classical system must be a cotangent bundle T ∗ Q.
19
and [B̂, Â2 ] = 0 so that, in physical parlance, A1 and B can be given ‘si-
multaneous values’ (or can be measured simultaneously) as can A2 and B.
However, the implication of the discussion above is that the value ascribed to
B (resp. the result of measuring B) depends on whether it is considered to-
gether with A1 , or together with A2 . In other words the value of the physical
quantity B is contextual. This is often considered one of the most important
implications of the Kochen-Specker theorem.
op
5.2 The topos of presheaves SetsV(H)
To see how all this relates to topos theory let us rewrite the above slightly.
Thus, let V(H) denote the collection of all commutative sub-algebras of op-
erators on the Hilbert space H. This is a partially-ordered set with respect to
sub-algebra inclusion. Hence it is also a category whose objects are just the
commutative sub-algebras of B(H); we shall call it the ‘category of contexts’.
We view each commutative algebra as a context with which to view the
quantum system in an essentially classical way in the sense that the physical
quantities in any such algebra can be given consistent values, as in classical
physics. Thus each context is a ‘classical snapshot’, or ‘world-view’, or ‘win-
dow on reality’. In any modal interpretation of quantum theory, only one
context at a time is used15 but our intention is to use the collection of all
contexts in one mega-structure that will capture the entire quantum theory.
To do this, let us consider the association of the spectrum ΣV (the set
of local valuations on V ) to each commutative sub-algebra V . As explained
above, there are restriction maps rW V : ΣW → ΣV for all pairs V, W with
V ⊆ W , and these maps satisfy the conditions in (5.4). In the language of
category theory this means that the operation V 7→ ΣV defines the elements
of a contravariant functor, Σ, from the category V(H) to the category of sets;
equivalently, it is a covariant functor from the opposite category, V(H)op to
Sets.
Now, one of the basic results in topos theory is that for any category C,
the collection of covariant functors F : C op → Sets is a topos, known as the
op
‘topos of presheaves’ over C, and is denoted SetsC . In regard to quantum
theory, our fundamental claim is that the theory can be reformulated so as
op
to look like classical physics, but in the topos SetsV(H) . The object Σ is
known as the ‘spectral presheaf’ and plays a fundamental role in our theory.
In purely mathematical terms this has considerable interest as the foundation
for a type of non-commutative spectral theory; from a physical perspective,
15
In the standard instrumentalist interpretation of quantum theory, a context is selected
by choosing to measure a particular set of commuting observables.
20
we identify it as the state object in the topos. op
The terminal object 1SetsV(H)op in the topos SetsV(H) is the presheaf
that associates to each commutative algebra V a singleton set {∗}V , and the
restriction maps are the obvious ones. It is then easy to see that a global
element λ : 1SetsV(H)op → Σ of the spectral presheaf is an association to each
V of a spectral element λV ∈ ΣV such that, for all pairs V, W with V ⊆ W
we have (5.5). Thus we have the following basic result:
Of course, identifying the topos and the state object are only the very
first steps in constructing a topos formulation of quantum theory. The next
key step is to associate a sub-object of Σ with each physical proposition. In
quantum theory, propositions are represented by projection operators, and
so what we seek is a map
21
In other words, the pseudo-state corresponding to the unit vector |ψi is just
the daseinisation of the projection operator onto |ψi.
6 Conclusions
We have revisited the oft-repeated statement about fundamental incompat-
ibilities between quantum theory and general relativity. In particular, we
have argued that the conventional quantum formalism is inadequate to the
task of quantum gravity both in regard to (i) the non-realist, instrumentalist
interpretation; and (ii) the a priori use of the real and complex numbers.
Our suggestion is to employ a mathematical formalism that ‘looks like’
classical physics (so as to gain some degree of ‘realism’) but in a topos, τ ,
other than the topos of sets. The ingredients in such a theory are:
22
References
[1] A. Döring. The physical interpretation of daseinisation. To appear in
Deep Beauty, ed. Hans Halvorson (2010).
[2] A. Döring, and C.J. Isham. A topos foundation for theories of physics:
I. Formal languages for physics, J. Math. Phys 49, 053515 (2008).
[3] A. Döring, and C.J. Isham. A topos foundation for theories of physics:
II. Daseinisation and the liberation of quantum theory. J. Math. Phys
49, 053516 (2008).
[4] A. Döring, and C.J. Isham. A topos foundation for theories of physics:
III. Quantum theory and the representation of physical quantities with
arrows Ă : Σ → R . J. Math. Phys 49, 053517 (2008).
[5] A. Döring, and C.J. Isham. A topos foundation for theories of physics:
IV. Categories of systems. J. Math. Phys 49, 053518 (2008).
[6] A. Döring, and C.J. Isham. ‘What is a Thing?’: Topos Theory in the
Foundations of Physics. To appear in New Structures in Physics, ed.
B. Coecke, Springer (2010).
[7] C.J. Isham. Some reflections on the status of conventional quantum
theory when applied to quantum gravity. In Proceedings of the Confer-
ence in Honour of Stephen Hawking’s 60’th birthday, ed. G. Gibbons,
Cambridge University Press, Cambridge (2003).
[8] M. Heidegger. What is a Thing? Regenery/Gateway, Indiana (1967).
[9] C.J. Isham and J. Butterfield. A topos perspective on the Kochen-
Specker theorem: I. Quantum states as generalised valuations. Int. J.
Theor. Phys. 37, 2669–2733 (1998).
[10] C.J. Isham and J. Butterfield. A topos perspective on the Kochen-
Specker theorem: II. Conceptual aspects, and classical analogues. Int.
J. Theor. Phys. 38, 827–859 (1999).
[11] C.J. Isham, J. Hamilton and J. Butterfield. A topos perspective on
the Kochen-Specker theorem: III. Von Neumann algebras as the base
category. Int. J. Theor. Phys. 39, 1413-1436 (2000).
[12] C.J. Isham and J. Butterfield. A topos perspective on the Kochen-
Specker theorem: IV. Interval valuations. Int. J. Theor. Phys 41, 613–
639 (2002).
23
[13] C.J. Isham and J. Butterfield. Space-time and the philosophical chal-
lenge of quantum gravity. In Physics Meets Philosophy at the Planck
Scale, eds. C. Callender and N. Huggett, Cambridge University Press,
33–89 (2001).
[14] S. Kochen and E.P. Specker. The problem of hidden variables in quan-
tum mechanics. Journal of Mathematics and Mechanics 17, 59–87
(1967).
24