You are on page 1of 20

Quasi spaces revisited

Federico Holik1,4 , Juan Pablo Jorge1 and Cesar Massri2


February 24, 2020

1- Instituto de Fı́sica La Plata, CCT-CONICET, and Departamento de Fı́sica, Facultad de Ciencias


Exactas, Universidad Nacional de La Plata - 115 y 49, C.C. 67, 1900 La Plata, Argentina
2- Departamento de Matemática - Facultad de Ciencias Exactas y Naturales
Universidad de Buenos Aires - Pabellón I, Ciudad Universitaria

Abstract
It is usually stated that quantum mechanics presents problems with the identity of particles,
the most radical position -supported by E. Schrödinger- asserting that elementary particles
are not individuals. But the subject goes deeper, and it is even possible to obtain states
with an undefined particle number. In this work we present a set theoretical framework
for the description of undefined particle number states in quantum mechanics which pro-
vides a precise logical meaning for this notion. This construction goes in the line of solving a
problem posed by Y. Manin, namely, to incorporate quantum mechanical notions at the foun-
dations of mathematics. We also show that our system is capable of representing quantum
superpositions.

Key words: set theory-undefined particle number

1 Introduction
The status of identity of quantum systems has been an issue almost since its conception. Many
aspects of the quantum formalism suggest that quantum systems lack identity, in the the sense
that they seem to be non-individuals. The meaning of this last assertion varies among the
different authors. Perhaps, the most radical position is that of utter indistinguishability, which
was supported by E. Schrödinger. But other voices appeared, claiming that elementary particles
can be weakly discerned. The literature on the subject is huge, and Needless to say, Bohmian
mechanics speaks about an ontology in which there is no issue with identity: Bohmian particles
can be considered individuals, which are identified by their hidden trajectories in space-time.
But even in Bohmian mechanics, identity is hidden, and there is no empirical procedure that
allows to identify quantum systems in special states. This is a remarkable feature of quantum
phenomena, independently of any interpretation: under certain circumstances, there is no way
to identify quantum systems of the same kind. Moreover, the symmetrization postulate, closely
related to the indistinguishability principle, can be used to explain remarkable physical processes.
Among them, one can find the Pauli exclusion principle, the Bose-Einstein condensation, and
all phenomena related to quantum statistics in general. Currently, quantum indistinguishability

1
is considered as a resource, and expolited in quantum informational tasks (citar Adesso y Hong
ou mandel).
In any case -independently of any interpretation- the symmetrization postulate and the inca-
pability of distinguishing quantum systems must be reflected in the effective part of the theory
(i.e., that part of the formalism that connects with experience). At this level, the symmetriza-
tion postulate plays a key role, in the sense that it is the mathematical procedure that physicists
found in order to give place to predictions that describe quantum statistics correctly and, at the
same time, reflect the fact that quantum systems cannot be discerned.
But the symmetrization postulate is implemented with a trick. First, quantum systems are
labeled in order to create a tensor product Hilbert space. After that initial labeling, quantum
states are symmetrized (or anti-symmetrized) in order to obtain the correct sates. While correct,
this procedure seems to rather artificial, because the initial labeling plays no real role in the
predictions of the theory. Thus, many authors posed the problem of finding a formulation of
quantum theory in which indistinguishability is taken right from the start. ACA CITAR A
POST, A REDHED Y TELLER, KRAUSE, KRAUSE Y HOLIK. Among the attempts to solve
this problem, one can find KRAUSE, KRAUSE Y DA COSTA, and HOLIK. In this work,
we elaborate on the proposal of Holik et al. We first review the fundamental features of the
Fock-space formulation of quantum mechanics, and the rudiments of quasi-set theory. Next, we
present the quasi-spaces in a new math fashion.

1.1 Fock-space formalism


It can be shown that the F SF may be used as an alternative approach to non relativistic quan-
tum mechanics [53]. Let us review how this is done in order to understand second quantization.
For the sake of simplicity, we will do this using the heuristic approach presented in elementary
expositions like [52, 53] (but see for example [51], [54] and [55] for a mathematically rigorous
presentation). Remember first that the standard wave mechanics approach presupposes a kinetic
energy
~2 ∇2
T1 (r) = −( ) (1)
2m
which for the n particles case takes the form
n
X
Tn = T1 (ri ) (2)
i=1
with a similar equation for the external potential. Suppose that we have a pairwise interaction
potential
n
X
Vn = V2 (ri , rj ) (3)
i>j=1

The total hamiltonian operator is thus given by


n n
X ~2 ∇2i X
Hn = [(− ) + V1 (ri ) + V2 (ri , rj )] (4)
2m
i=1 i>j=1

and the n-particles wave function


Ψn (r1 , . . . , rn , t) (5)

2
is a solution of the Schrödinger’s equation

Hn Ψn = i~ Ψn (6)
∂t
The second quantization approach to QM has its roots in considering equation (6) as a classical
field equation, and its solution Ψn (r1 , . . . , rn ) as a classical field to be quantized. This alternative
view was originally adopted by P. Jordan [56, 57], one of the foundation fathers of quantum
mechanics, and spread worldwide after the Dirac’s paper [58]. And it is a standard way of
dealing with relativistic quantum mechanics (canonical quantization). The space in which these
quantized fields operate is the Fock-space.
The standard Fock-space is built up from the one particle Hilbert spaces. Let H be a Hilbert
space and define:

H0 =C
H1 =H
H2 =H⊗H
..
.
Hn = H ⊗ · · · ⊗ H (7)
The Fock-space is thus constructed as the direct sum of n particle Hilbert spaces:

M
F= Hn (8)
n=0

When dealing with bosons or fermions, the symmetrization postulate (SP ) must be imposed.
In order to do so, given a vector v = v1 ⊗ · · · ⊗ vn ∈ Hn , define:
1 X
σ n (v) = ( ) P (v1 ⊗ · · · ⊗ vn ) (9)
n!
P

and:
1 X p
τ n (v) = ( ) s P (v1 ⊗ · · · ⊗ vn ) (10)
n!
P

where:
1 if p is even,
n
sp = −1 if p is odd.
Calling
Hσn = {σ n (v) : v ∈ Hn } (11)
and:
Hτn = {τ n (v) : v ∈ Hn } (12)
we have the Fock-space

3

M
Fσ = Hσn (13)
n=0
for bosons and

M
Fτ = Hτn (14)
n=0
for fermions. Now the second quantization procedure considers the one particle wave function
ψ(r) and its hermitian conjugate ψ(r)† as operators acting on the Fock-space and satisfying:

[ψ(r), ψ(r0 )]∓ = 0


[ψ(r)† , ψ(r0 )† ]∓ = 0
[ψ(r), ψ(r0 )† ]∓ = δr−r0 (15)

where δ(r − r0 ) is the Dirac delta function. If A and B are operators, the brackets are defined
by:
[A, B]∓ = AB ∓ BA (16)
The n particle wave function Ψn (r1 , . . . , rn ) of the standard formulation is now written as
Z Z
− 12
|ψn i = (n!) 3
d r1 · · · d3 rn ψ(r1 )† · · · ψ(rn )† |0iΨn (r1 , . . . , rn ) (17)

and it is an eigenvector (with eigenvalue n) of the particle number operator:


Z
N := d3 rψ(r)† ψ(r) (18)

and we find the relation


1
Ψn (r1 , · · · , rn ) = (n!)− 2 h0|ψ(r1 ) · · · ψ(rn )|Ψn i (19)
An arbitrary vector of the Fock-space will be a superposition of states with different particle
number of the form

X
|Ψi = |Ψn i (20)
n=0
and will not be in general an eigenstate of the particle number operator. Thus, according to the
standard interpretation of QM, its particle number will be undetermined. This is very important,
because in the presence of particle interactions, the states may evolve into an undefined particle
state like (20).
The kinetic energy operator (1) can now be written as
Z
T = d3 rψ † (r)T1 (r)ψ(r) (21)

4
It can also be shown that:
Z Z n
− 12
X
3 3 † †
T |Ψn i = (n!) d r1 · · · d rn Ψ (r1 ) · · · Ψ (rn )|0i T1 (ri )Ψn (r1 , · · · , rn ) (22)
i=1

The pairwise interaction potential V2 (r, r0 ) can be written as


Z Z
1
V = d r d3 r0 ψ † (r)ψ † (r0 )V2 (r, r0 )ψ(r0 )ψ(r)
3
(23)
2
Its action on |Ψn i is given by:
Z Z
− 21
V |Ψn i = (n!) d r1 · · · d3 r1 [V, ψ † (rn ) · · · ψ † (r1 )]|0iΨn (r1 . . . rn )
3
(24)

and it follows that:

Z Z
1
V |Ψn i = (n!)− 2 d3 r1 · · · d3 r1 ψ † (rn ) · · · ψ † (r1 )|0i
n X
X i−1
× V2 (ri , rj )Ψn (r1 , . . . , rn ) (25)
i=1 j=1

It can be shown that that the following equations holds:


1
Tn Ψn (r1 , · · · , rn ) = (n!)− 2 h0|Ψ(r1 ) · · · Ψ(rn )T |Ψn i (26)

1
Vn Ψn (r1 , · · · , rn ) = (n!)− 2 h0|Ψ(r1 ) · · · Ψ(rn )V |Ψn i (27)
where:
n
X
Tn = T1 (ri ) (28)
i=1
n
X
Vn = V2 (ri , rj ) (29)
i>j=1

The equivalence with wave mechanics can now be established as follows. If Ψn (r1 , · · · , rn )
satisfies the n particle Schrödinger wave equation with Hamiltonian (4), it follows that in the
Fock-space formulation |Ψn i must satisfy the Fock-space Schrödinger equation:

[i~( ) − H]|Ψn i = 0 (30)
∂t
with H = T + V : given by:

5
~2 ∇2
Z
H = d3 rψ † (r)[(− ) + V1 (r)]ψ(r)
2m
Z Z
1
+ d3 r d3 r0 ψ † (r)ψ † (r0 )V2 (r, r0 )ψ(r0 )ψ(r) (31)
2
It is important to remark that the n particle Schrödinger wave equation is not completely
equivalent to its analogue in the Fock-space formalism. Only solutions of the Fock-space equation
which are eigenvectors of the particle number operator with particle number n can be solutions
of the corresponding n particle Schrödinger wave equation. And the other way around, not all
the solutions of the n particle Schrödinger wave equation can be solutions of the Fock equation,
only those which are symmetrized do. Then, both conditions, definite particle number and
symmetrization, must hold in order that both formalisms yield equivalent predictions.
Observables are usually expresed in terms of creation and annihilation operators. The quantized
field may be expanded as
X
ψ(r) = ak uk (r) (32)
k

and the coefficients of the expansion will be the annihilation operators:


Z
ak = d3 ru∗k (r)ψ(r) (33)

A similar expansion stands for the creation operator: a†k . The usual interpretation is that a†k de-
scribes the “creation of a particle” with wave function uk (r), while ak describes the “annihilation
of a particle”. These operators satisfy the commutation relations

[ak , al ]∓ = 0
[a†k , a†l ]∓ = 0
[ak , a†l ]∓ = δkl (34)

Nk = a†k ak (35)
It is possible to put things in a more familiar way by using the “[, ]” symbol for bosonic com-
mutation relations

[aα ; a†β ] = aα a†β − a†β aα = δαβ I (36)

[a†α ; a†β ] = 0 (37)

[aα ; aβ ] = 0 (38)
and the symbol {; } for fermions (with Cα† and Cα playing the role of fermionic creation and
annihilation operators respectivey),

6
{Cα ; Cβ† } = Cα Cβ† + Cβ† Cα = δαβ I (39)

{Cᆠ; C↠} = 0 (40)

{Cα ; Cβ } = 0 (41)
Substitution of (32) in (31) yields:
X † 1X † †
H= ak Tkl al + ak al Vklpq ap aq (42)
2
kl klpq

where the matrix elements Tkl and Vklpq are given by:

2 ∇2
Tkl = d3 ru∗k (r)[(− ~2m
R
) + V1 (r)]ul (r)
Vklpq = d r d r uk (r)ul (r )V2 (r, r0 )up (r0 )uq (r)
R 3 R 3 0 ∗ ∗ 0 (43)

and similar expressions can be found for more general obserbables.

1.2 Coherent states of the electromagnetic field


Once that the second quantization procedure is outlined, we are ready to show a quantum state
which can be realized in the laboratory. Let us consider electromagnetic Maxwell’s equations
(without density charges)
1 ∂B
∇×E=− (44a)
c ∂t
1 ∂E
∇×B= (44b)
c ∂t
∇E = 0 (44c)

∇B = 0 (44d)
and then apply second quantization to them. This means that we must quantize its solutions,
and this means to elevate fields E(x) and B(x)) to the category of operators, call them Ê(x) and
B̂(x). Now, after some algebra, it is possible to decompose Ê(x) and B̂(x) in terms of fourier
expansion
1
(2π~ωk ) 2 {a†k (t)e∗k (x) + ak (t)ek (x)}
X
Ê(x) = (45)
k

The hamiltonian of the mth mode of the field can be written in terms of the creation and
annihilation operators a†k and ak
1
Hn = ~ω(a†k ak + ) (46)
2

7
and so, each a†m (am ) creates (annihilates) a photon in mode m. Then, a fock space state can
be expressed as
|n1 , n2 , . . . , nm , . . .i = |n1 i ⊗ |n2 i ⊗ . . . ⊗ |nm i ⊗ . . . (47)
with ni the number of photons present in each mode of the field. If for simplicity we concentrate
in only one frequency mode of the field, we can create any normalized superposition of states,
and in particular, the famous coherent state
X zn ∞
1
|zi = exp(− |z|2 ) 1 |ni (48)
2 (n!) 2
n=0

which can be realized in laboratory [52]. State (48) is clearly a superposition of different photon
number states and thus is not an eigenstate of the particle number operator. It follows that,
according to the standard interpretation, it represents a physical system formed by an undefined
number of photons. It is important to remark that there are -at least- two interpretations of
(48)

• 1-Equation (48) represents an statistical mixture of states with definite particle number.
• 2-Equation (48) represents an state which has no definite particle number.

The orthodox interpretation of QM points in the direction of the second option and the first one is
very difficult to sustain unless involved hypotheses are made [59]. Regardless the interpretational
debate, it will suffice for us that there exists at least one interpretation compatible with quantum
mechanics in which particle number is undefined. Thus, given that systems in states like (48)
are predicted by QM and can indeed be reproduced in the laboratory, we are going to propose
below a formalism in order to incorporate physical systems in such states in a set theoretical
framework.

2 von Neumann Algebras


C ∗ -algebras are a suitable tool for studying the formal structure of quantum mechanics [?] and
provide a formal manner of dealing with QFT [?]. It has also applications in the study of
statistical mechanics [?]. We will use the algebraic approach for dealing with the Fock-space
formalism of identical particles. Let us review the main characteristics of these algebras, starting
with a linear space A for which there exists a product operation between its elements and an
involution “∗”. A C ∗ -algebra will be an involutive algebra endowed with a norm which satisfies,
for its elements a,

• 1-kabk ≤ kakkbk for all a ∈ A


• 2-ka∗ ak = kak2
• 3-A is complete with the given norm

An important example of a C ∗ -algebra is that of the closed ∗ -algebras of B(H) (the set of
bounded operators “acting” on Hilbert space). In fact, C ∗ -algebras are abstract generalizations
of von Neumann algebras.
If A is a C*-algebra and if a linear form ϕ : A → C is such that ϕ(aa∗ ) ≥ 0 for all a ∈ A, we
say that it is positive. A state on a C*-algebra is a linear form such that it is positive and has

8
norm 1. Let us denote by C to the set of states in A. This is always a convex set. In an infinite
dimensional space H, a trace-class operator O is a compact operator such that
X
hφi |Oφi i < ∞, (49)
i∈I

for all orthonormal basis {|φi i}i∈I . Trace-class operators always have a trace defined as
X
tr(O) = hφi |Oφi i, (50)
i∈I

which is independent of the basis. Given a positive trace-class operator of trace 1, we obtain a
state s on B(H) if we introduce the linear form
s(a) = tr(Oa). (51)
Any normal state in B(H) has this form (a linear form is normal if it is continuous with respect
to the ultra-weak topology, or equivalently, to the ultra-strong topology). There are states on
B(H) which are not normal.
If A is a C*-algebra, denote by L+ A the set of linear forms on A of norm ≤ 1, endowed with the
topology of pointwise convergence. L+ A is a convex and compact space. If A has a unit (which is
the case of B(H)), then C is the intersection of the closed affine hyperplane of equation ϕ(1) = 1
with L+ A , and is thus a convex and compact space.

3 Fock-space formalism in von Neumann algebras


Let us review the Fock-space formalism from the vantage point of view of the algebraic approach,
in the wake of [55]. If H is a Hilbert space, define first
HN = H ⊗ . . . ⊗ H (N times), (52)
and then the direct sum
M
F(H) = Hn . (53)
n≥0

For each integer n, it is possible to introduce a bounded operator

l(ξ) : Hn −→ H(n+1)
η1 ⊗ . . . ⊗ ηn 7→ ξ ⊗ η1 ⊗ . . . ⊗ ηn (54)

of norm kξk. As the norm is independent of n, it is possible to construct a bounded operator


formed by the direct sum of these operators
l(ξ) : F(H) −→ F(H), (55)
also of norm kξk. Its adjoint l(ξ)∗ satisfies
l(ξ)∗ (η0 ⊗ . . . ⊗ ηn ) = hξ|η0 iη1 ⊗ . . . ⊗ ηn , (56)
and we also have

9
l(η)∗ l(ξ) = hη|ξi1F (H) . (57)

These operators l(ξ) generate a C ∗ -algebra to be called C ∗ (l(H)). Now, for Hn , consider the
symmetric group of transformations
sσ (η1 ⊗ . . . ⊗ ηn ) = ησ(1) ⊗ . . . ⊗ ησ(n) , (58)
Vn
and introduce H, the subspace of HV⊗n on which the symmetry-group operator Sn acts by
signature and orthogonal projection to n H. This yields the “permutation”
1 X
Pn = (−1)σ sσ . (59)
n!
σ∈Sn

Now, for any ξ1 , . . . ξn ∈ H one can introduce


√ n
^
ξ1 ∧ . . . ∧ ξn = n!Pn (ξ1 ⊗ . . . ⊗ ξn ) ∈ H, (60)
and
ξσ(1) ∧ . . . ∧ ξσ(n) = (−1)n (ξ1 ∧ . . . ∧ ξn ). (61)
After these preliminaries, the antisymmetric Fock-space of H gets defined as
n
M^
F− (H) = H, (62)
n≥0

and we can attempt to introduce creation-annihilation operators. Accordingly, consider the


operators an (ξ). Define first

An (ξ) : H⊗n −→ n+1 H


V
q
η1 ⊗ . . . ⊗ ηn 7→ n! ξ ∧ η1 ∧ . . . ∧ ηn = (n+1)!
√1
n! Pn+1 (ξ ⊗ η1 ⊗ . . . ⊗ ηn ) . (63)
Vn
Calling the restriction of An (ξ) to H

an (ξ) : n H −→ n+1 H
V V

η1 ∧ . . . ∧ ηn 7→ ξ ∧ η1 ∧ . . . ∧ ηn , (64)

it is possible to show that the norm of an (ξ) is independent of n. Then one can show that for
any ξ, η ∈ H
a∗n (ξ)an (η) + an−1 (η)a∗n−1 (ξ) = hξ|ηi1Vn H (65a)
an+1 (ξ)an (η) + an+1 (η)an (ξ) = 0. (65b)
We need at this point recourse to the well known result

10
Vn
Proposition 3.1. If ξ ∈ H, ξ 6= 0 and for all n ≥ 0, 1
kξk2
a∗ (ξ)an (ξ) is a projection on H,
and kan (ξ)k = kξk,
that allows us to formulate:
Definition 3.2. For all ξ ∈ H the creation operator
a(ξ) : F− (H) −→ F− (H) (66)
Vn
is the direct sum of the an (ξ)’s on the corresponding spaces H. This is a bounded operator
of norm kξk that depends linearly on ξ. Its adjoint operator a∗ (ξ) is the annihilation operator,
anti-linear in ξ. If kξk = 1, then a∗ (ξ)a(ξ) and a(ξ)a∗ (ξ) are projections on B(F − (H))
The canonical anticommutation algebra is the C ∗ -algebra generated by all the a(ξ)’s [54]. The
associated CAR relations (see, for instance, [54, ?]) are
a∗ (ξ)a(η) + a(η)a∗ (ξ) = hξ|ηi1 (67a)

a(ξ)a(η) + a(η)a(ξ) = 0 (67b)


A similar construction can be made for F+ (H) and the symmetric group of transformations.

4 Quasi-Set Theory
In this section, we basically follow the exposition of [?]; for details, see [?], [?, Chap.7]. Quasi-set
theory is a mathematical theory that enables us to deal with collections (quasi-sets) of objects
that may be indiscernible without turning to be identical (being the same object). The only way
of dealing with objects of this kind in a standard theory such as ZFC is to allow the introduction
of some ad hoc devices such as the restriction of the lexicon of properties, that is, by taking a
language with a finite number of them. This is Quine’s famous way of defining identity, namely,
by the exhaustion of the chosen (finitely many) predicates [?, Chap.12]. But this just defines
indiscernibility relative to the language’s predicates, and not identity strictly speaking, for there
may exist other predicates not in the language which distinguish among the entities.
In ZFC (the same can be said of most theories with due qualification) given an object a,
we can always form the unitary set A = {a} and define a unary property (a formula with
just one free variable) of a by posing Ia (x) iff x ∈ A. This formula of course distinguishes
(or discriminates) a from any other object for only a satisfies it. Saying in other words, in
the standard mathematics (in the sense mentioned in the Introduction), whenever we have a
set with cardinal α > 1, its elements are distinct. In short, there are no indiscernible objects,
except with respect to a few chosen predicates.
Quine says that objects may be (strongly) discriminable when there is a formula with only
one free variable that is satisfied by one of the objects but not by the other. Two objects are
moderately discriminable when there exists a formula with two free variables that is satisfied by
the two objects in one order but not in the other order [?, Chap.15]. As he recalls, any two
real numbers are strongly discriminable, although they may be ‘specified’, that is, definable by a
formula. But in the interpretation which supports the received view, there is no way to attribute
an identity to quanta (let us call this a ‘which is which criterion’), something that isolate one of
them from the others in such a way that this chosen object remains with its identity forever.
Thus, how can we deal with, say, the two electrons of an Helium atom in the fundamental
state? We know that they have all the same properties but distinct spins in a given direction,

11
and so they obey an irreflexive and symmetric relation ‘to have different spin of’ and thus they
are moderately discriminable in the sense of Quine. Muller and Saunders think that with this
example, they have found objects (the mentioned electrons) that are moderately but not strongly
discriminable [?]. Thus apparently we have found a way of speaking about two electrons with
different spins and cannot say which is which. But this is a false supposition (see [?]). Within
classical logic (and the mentioned authors assume that they are using ZF C), being objects (in
some sense of the word) in a finite number, they can always be named, say a and b and the
above property ‘to belong to A = {a}’, which is true just for a, distinguish them absolutely! In
Quine’s words, they can be always specified, that is, the language (of ZF C) there is always a
formula in one free variable that is uniquely satisfied by a given object [?, p.134]. You may say
that this property Ia (x) (‘being identical with a’) is not a ‘legitimate property’ of x, but in our
opinion this would be a quite arbitrary answer: which would be the legitimate ones? And why
such an election (if there is one)?
Quasi-set theory offers a different alternative to treat these questions. Although electrons do
present a difference due to Pauli’s exclusion principle (for instance, differences in their spins),
any permutation of them does not change the relevant probabilities, as is well known; in other
words, there is no a which is which criterion. So, it seems that it is better (and apparently
most correct) to say that the electrons may be indiscernible but without assuming that this
makes them the very same object. A paradigmatic example may be that of the atoms in a
Bose-Einstein condensate. Thus, we need to avoid that indiscernibility implies identity (in the
philosophical sense of being the same entity). To cope with this idea, we ‘separate’ the concepts:
indiscernibility, or indistinguishability, is a relation that holds for all objects of our domain, but
identity is not. Certain objects may be indiscernible without turning to be the same object, as
implied by the standard theory of identity, and they may form collections with cardinals greater
than one (this is achieved by the postulates of the theory) but in a way that they cannot be
identified, named, labeled, counted in the standard way.1 There is no space here to provide
the details of the theory, and we suggest [?] and [?] for detailed references and for the axioms.
Anyhow, in the next subsection we outline without the details the main ideas of the theory.

4.1 Basic Ideas Of The Theory Q


Intuitively speaking, a quasi-set is a collection of objects such that some of them may be indis-
tinguishable without turning to be identical.
As mentioned above, quasi-set theory Q has its main motivations in some insights advanced
by Schrödinger in that the concept of identity would make no sense when applied to elemen-
tary particles [?, pp. 17-18]. Another motivation is (in our opinion) the need, stemming from
philosophical worries, of dealing with collections of absolutely indistinguishable items that need
not be the same ones. Spatio-temporal differences could be used in this case, you may say, and
this serves do distinguish them. Without discussing the role of spatio-temporal properties here
(but see [?]), we can argue that these objects are invariant by permutations; in other words, the
world does not present differences in substituting them one from the other. Thus, it would be
difficult to say that they have some form of identity, for they (in principle) lack any identifying
characteristic. Objects of this kind act like those that obey Bose-Einstein statistics, that is,
bosons (we should remember that the indiscernibility hypothesis was essential in the derivation
of Planck’s formula –see [?] once more).
Thus, the first point is to guarantee that identity and indistinguishability (or indiscernibility)
1
We mean: a series of, say, five objects can be counted by proving that the set having them as elements (and no
other element) is equinumerous to a finite ordinal, in the case, the ordinal 5 = {0, . . . , 4}. We remark that in order
to define the bijection, we need to distinguish among the elements being counted –they need to be individuals,
yet sometimes not specified.

12
will not collapse into one another when the theory is formally developed. We assume that
identity (symbolized by ‘=’) is not a primitive relation, but we use a weaker primitive concept
of indistinguishability (symbolized by ‘≡’) instead. This is just an equivalence relation and
holds among all objects of the considered domain. If the objects of the theory are divided up
into groups, namely, the m-objects (standing for ‘micro-objects’) and M -objects (for ‘macro-
objects’) —these are ur-elements— and quasi-sets of them (an probably having other quasi-sets
as elements as well), then identity (having all the properties of standard identity of ZF) can
be defined for M -objects and quasi-sets having no m-objects in their transitive closure (this
concept is like the standard one). Thus, if we take just the part of the theory obtained by ruling
out the m-objects and collections (quasi-sets) having them in their transitive closure, we get a
copy of ZFU (ZF with Urelemente); if we further eliminate the M -objects, we get a copy of the
‘pure’ ZF.
Technically, expressions such as x = y are not always well formed, for they are not formulas
when either x or y denote m-objects (entities satisfying the unary primitive predicate m). We
express that by informally saying that the concept of identity does not always make sense for
all objects (it should be emphasized that this is just a way of speech). The objects (the m-
objects) to which the defined concept of identity does not apply are termed non-individuals for
historical reasons (see [?]). As a result (from the axioms of the theory), we can form collections
of m-objects which have no identity; these collections may have a cardinal (termed its ‘quasi-
cardinal’) but not an associated ordinal. Thus, the concept of ordinal and of cardinal are taken
as independent, as in some formulations of ZF proper. So, informally speaking, a quasi-set of
m-objects is such that its elements cannot be identified by names, counted, ordered, although
there is a sense in saying that these collections have a cardinal (that cannot be defined by means
of ordinals, as usual, which presupposes identity).
When Q is used in connection with quantum physics, the m-objects are thought of as rep-
resenting quanta (henceforth, q-objects), but they are not necessarily ‘particles’ in the standard
sense—associated with classical physics of even with orthodox quantum mechanics; waves, field
excitations (the ‘particles’ in quantum field theory—QFT), perhaps even strings or whatever
entities supposed indiscernible can be taken as possible interpretations of the m-objects. Gen-
erally speaking, whatever ‘objects’ sharing the property of being indistinguishable can also be
values of the variables of Q (see [?, Chap. 6] for a survey on the various different meanings that
the word ‘particle’ has acquired in connection to quantum physics).
Another important feature of Q is that standard mathematics can be developed using its
resources, for the theory is conceived in such a way that ZFU (and hence also ZF, perhaps with
the axiom of choice, ZFC) is a subtheory of Q. In other words, the theory is constructed so that
it extends standard Zermelo-Fraenkel with Urelemente (ZFU); thus standard sets (of ZFU) can
be viewed as particular qsets (that is, there are qsets that have all the properties of the sets
of ZFU, while the objects in Q corresponding to the Urelemente of ZFU are termed M -atoms;
these satisfy another primitive unary predicate M ). The ‘sets’ in Q will be called Q-sets, or
just sets for short. An object is a qset when it is neither an m-object nor an M -object and,
to make the distinction, the language of Q encompasses a third unary predicate Z such that
Z(x) says that x is a qset which is also a set, and they correspond to those objects erected in
the ‘classical’ part of the theory (without m-objects). It is also possible to show that there is
a translation from the language of ZFU into the language of Q so that the translations of the
postulates of ZFU turn to be theorems of Q; thus, there is a ‘copy’ of ZFU in Q, and we refer to
it as the ‘classical’ part of Q. In this copy, all the usual mathematical concepts can be stated,
as for instance, the concept of ordinal (for sets).
Furthermore, it should be recalled that the postulates for the relation of indiscernibility,
when applied to M -atoms or to Q-sets, collapses into standard identity (of ZFU). The Q-sets
are qsets whose transitive closure (defined as usual) does not contain m-atoms (in other words,

13
they are ‘constructed in the classical part of the theory—see Fig. 1).

BB
B
BB
B
BB
B On 

Q
B B B 
B pure qsets B copies of ZFU-sets B copies of ZF-sets 
B B B 
B (sets with M -atoms) B (‘pure’ sets)
'$
B 
B (here, no identity) B B 
B x≡y B a B 
c
'$
B B B 
b
B &%
B B B 
B
∗ (here we can speak ofB x = y) 
'$B
∗ ∗ '$
B B 
B
&%
B d B 
B
∗∗
B f B 
IndiscernibleB qsets ∗
B e B 
B B &% B 
B &% B B 

B m-atoms M -atoms
B B B 
B B ∅

Figure 1: The Quasi-Set Universe Q: On is the class of ordinals, defined in the classical part of
the theory. See [?]
In order to distinguish between Q-sets and qsets that may have m-atoms in their transitive
closure, we write (in the metalanguage) {x : ϕ(x)} for the former and [x : ϕ(x)] for the latter.
In Q, we term ‘pure’ those qsets that have only m-objects as elements (although these elements
may be not always indistinguishable from one another, that is, the theory is consistent with the
assumption of the existence of different kinds of m-atoms), and to them it is assumed that the
usual notion of identity cannot be applied (that is, let us recall, x = y, as well as its negation,
x 6= y, are not well formed formulas if either x or y stand for m-objects). Notwithstanding, the
primitive relation ≡ applies to them, and it has the properties of an equivalence relation.
We have also a defined concept of extensional identity and it has the properties of standard
identity of ZFU. More precisely, we write x =E y (read ’x and y are extensionally identical’) iff
they are both qsets having the same elements (that is, ∀z(z ∈ x ↔ z ∈ y)) or they are both
M -atoms and belong to the same qsets (that is, ∀z(x ∈ z ↔ y ∈ z)). From now on, we shall not
bother to always write =E , using simply the symbol “=” for the extensional equality.
Since m-atoms cannot be identified in the formalism, it is not possible in general to attribute
an ordinal to qsets of such elements. Thus, for certain qsets, it is not possible to define a notion
of cardinal number by means of ordinals. The theory uses a primitive concept of quasi-cardinal
instead, which intuitively stands for the ‘quantity’ of objects in a collection.2 The theory has
still an ‘axiom of weak extensionality’, which states (informally speaking) that those quasi-sets
that have the same quantity of elements of the same sort (in the sense that they belong to the
same equivalence class of indistinguishable objects) are indistinguishable by their own. One of
2
A notion of finite quasi-cardinal can be defined as a derived concept (see [?]).

14
the interesting consequences of this axiom is related to the non observability of permutations
in quantum physics, which is one of the most basic facts regarding indistinguishable quanta
(for a discussion on this point, see [?]). In standard set theories, if w ∈ x, then of course
(x − {w}) ∪ {z} = x iff z = w. That is, we can ’exchange’ (without modifying the original
arrangement) two elements iff they are the same elements, by force of the axiom of extensionality.
In Q we can prove the following theorem, where [[z]] (and similarly [[w]]) stand for a quasi-set
with quasi-cardinal 1 whose only element is indistinguishable from z (respectively, from w –the
reader shouldn’t think that this element is identical to either z or w:
Theorem 1 (Unobservability of Permutations). Let x be a finite quasi-set such that x does not
contain all indistinguishable from z, where z is an m-atom such that z ∈ x. If w ≡ z and w ∈
/ x,
then there exists [[w]] such that
(x − [[z]]) ∪ [[w]] ≡ x
Informally speaking, supposing that x has n elements, then if we ‘exchange’ their elements
z by corresponding indistinguishable elements w (set theoretically, this means performing the
operation (x − [[z]]) ∪ [[w]]), then the resulting quasi-set remains indistinguishable from the one
we started with. In a certain sense, it does not matter whether we are dealing with x or with
(x − [[z]]) ∪ w0 . So, within Q, we can express that ‘permutations are not observable’, without
necessarily introducing symmetry postulates, and in particular to derive ‘in a natural way’ the
quantum statistics (see [?, Chap.7]).

5 The Q-space
As we have seen in Section ??, if particles are indistinguishable they can only access symmetrized
states. But particles are labeled in this construction, and this procedure was criticized (Section
??). It has been claimed that the Fock-space formalism poses a solution to the questions raised
by this criticism [43, 44]. But the Fock-space formalism also makes use of particle labeling uses
particle labeling in order to obtain the correct states [?].
How can we avoid this problem of the Fock-space formulation of QM ? If we could avoid the
individuation of the particles at every step of the construction of a Fock-like formulation of QM ,
we would give a positive answer to the problem posed in [43, 44] (recalled in the Introduction)
which is not affected by the criticisms linked to it. Quasi-set theory can be used for this purpose,
and in fact, this construction has been done [?, ?]. There, an alternative proposal is presented
which resembles that of the Fock-space formalism but based on Q. And thus, genuinely avoiding
artificial labeling. We give a sketch of the construction here, mainly following [?, ?]. And in the
following Section, we will also see that this kind of constructions not only allows us to solve the
problem posed above, but they serve also to plant interesting foundational issues.
Let us consider a set  = {i }i∈I (that is, a “set” in Q), where I is an arbitrary collection of
indexes (this makes sense in the ‘classical part’ of Q). Suppose that the elements i to represent
the eigenvalues of a physical observable. Next, quasi-functions f are constructed, such that
f :  −→ Fp , where Fp is the quasi-set formed of finite and pure quasi-sets. f is a quasi-set
formed by ordered pairs hi ; xi with i ∈  and x ∈ Fp .
These quasi-functions are chosen in such a way that whenever hik ; xi and hik0 ; yi belong
to f and k 6= k 0 , then x ∩ y = ∅. It is further assumed that the sum of the quasi-cardinals of
the quasi-sets which appear in the image of each of these quasi-functions is finite. This means
that qc(x) = 0 for every x in the image of f , except for a finite number of elements of . These
quasi-functions form a quasi-set called F.
A pair hi ; xi is interpreted as the statement “the energy level i has occupation number
qc(x)”. Quasi-functions of this kind are represented by expressions such as fi1 i2 ...im . If the

15
symbol ik appears j-times the level ik has occupation number j. The levels that do not appear
have occupation number zero.
At this point of the construction, the indexes appearing in fi1 i2 ...in has no meaning at all.
But an order can be defined as follows. Given a quasi-function f ∈ F, let {i1 i2 . . . im } be
the quasi-set formed by the elements of  such that hik , xi ∈ f and qc(x) 6= 0 (k = 1 . . . m).
This quasi-set is denoted supp(f ). Consider now the pair ho, f i, where o is a bijective quasi-
function o : {i1 i2 . . . im } −→ {1, 2, . . . , m}. Each one of the quasi-functions o defines an order
on supp(f ). Let OF denote the quasi-set formed by all the pairs ho, f i. OF is the quasi-set
formed by all the quasi-functions of F with ordered support. Using a similar notation as before
(and also repeating indexes according to the occupation number), fi1 i2 ...in ∈ OF refers to a
quasi-function f ∈ F and a special ordering of {i1 i2 . . . in }. But now, the order of the indexes
must not be understood as a labeling of particles, because it can be shown that the permutation
of particles does not give place to a new element of OF [?].
Consider next the collection of quasi-functions C which assign to every f ∈ F (or f ∈ OF)
a complex number. A quasi-function c ∈ C is a collection of ordered pairs hf ; λi, where f ∈ F
(or f ∈ OF) and λ a complex number. Let C0 be the subset of C such that, if c ∈ C0 , then
c(f ) = 0 for almost every f ∈ OF (i.e., c(f ) = 0 for every f ∈ OF except for a finite number of
quasi-functions). A sum and a product can be defined in C0 as follows
Definition 5.1. Given α, β, γ ∈ C, and c, c1 , c2 ∈ C0 , then
(γ ∗ c)(f ) := γ(c(f )) and (c1 + c2 )(f ) := c1 (f ) + c2 (f )
Using the above definitions, (C0 , +, ∗) is endowed with a complex vector space structure. Given
a quasi-function c ∈ C0 such that c(fi ) = λi (i = 1, . . . , n) for some finite set of quasi-functions
{fi } belonging to F or OF the following association is done
c ≈ (λ1 f1 + λ2 f2 + · · · + λn fn ) (68)
Thus, a quasi-function c ∈ C0 is interpreted as a linear combination of the quasi-functions fi
(representing a quantum superposition).
Scalar products must be introduced in order to reproduce the quantum mechanical machinery
of computation of probabilities. It is possible to define two of them, one for bosons (“◦”) and
one for fermions (“•”). In this way (and using norm completion), two Hilbert spaces (VQ , ◦)
and (VQ , •) are obtained. The scalar product for bosons is defined as follows
Definition 5.2. Let δij be the Kronecker symbol and fi1 i2 ...in and fi0 i0 ...i0 two basis vectors,
1 2 m
then X
fi1 i2 ...in ◦ fi0 i0 ...i0 := δnm δi1 pi01 δi2 pi02 . . . δin pi0n
1 2 m
p

The sum is extended over all the permutations of the set i0 = (i01 , i02 , . . . , i0n ) and for each permu-
tation p, pi0 = (pi01 , pi02 , . . . , pi0n ).
and for fermions
Definition 5.3. Let δij be the Kronecker symbol, fi1 i2 ...in and fi0 i0 ...i0 two basis vectors,
1 2 m
then X
fi1 i2 ...in • fi0 i0 ...i0 := δnm sp δi1 pi01 δi2 pi02 . . . δin pi0n
1 2 m
p

where: sp = +1 if p is even and sp = −1 if p is odd.

16
These products can be easily extended to all linear combinations. The second product • is an
antisymmetric sum of the indexes which appear in the quasi-functions and the quasi-functions
must belong to OF. If the occupation number of a product is greater or equal than two, then,
it can be shown that the vector has null norm. Thus reproducing Pauli’s exclusion principle for
fermions [?].
With these constructions within Q, the formalism of QM can be rewritten giving a positive
answer to the problem of giving a formulation of QM in which intrinsical indistinguishability is
taken into account from the beginning, without artificially introducing artificial labels [?, ?].

6 Conclusions
In this work we have presented a renewed version of the quasi-spaces which are formulated using
quasi-set theory. Our construction can be used to build Fock-space quantum mechanics in a
suitable way.

Acknowledgements

References
[1] Manin, Yu. I., “Mathematical Problems I: Foundations”, in Browder, F. E. (ed.) (1976) 36,
cited in [12].
[2] Manin, Yu. I., “A course in mathematical logic”, Springer-Verlag, p. 84 (1977).
[3] Y. Manin, “A course in mathematical logic for mathematicians”, Springer-Verlag, New
York-Heidelberg-Berlin, (2010).
[4] Halmos P., Naive Set Theory, D. Van Nostrand Company (1963).
[5] K. Kunen, “Set theory, an introduction to indpendence proofs”, North-Holland,
Amsterdam-London-New York-Tokio, (1980).
[6] D. Brignole and N. C. A. da Costa, “On supernormal Ehresmann-Dedecker universes”
Mathematische Zeitschrift 122, 4, (1971) 342-350.
[7] N.C.A. da Costa, “Ensaio sobre os Fundamentos da Lógica”, HUCITEC, São Paulo, (1980).
[8] N.C.A. da Costa y O. Bueno, “Non reflexive logics”, Revista Brasileira de Filosofia, 58
(2009) 181-208.
[9] Dalla Chiara, M. L. and Toraldo di Francia, G.: “Identity Questions from Quantum The-
ory”, in Gavroglu, K. et. al., (eds.), Physics, Philosophy and the Scientific Community,
Dordrecht, Kluwer Academic Publishers, (1995) 39-46.
[10] Dalla Chiara, M. L., Giuntini, R. and Krause, D., “Quasiset Theories for Microobjects: A
Comparision”, in Castellani, E. (ed.), Interpreting bodies: Classical and quantum objects in
modern physics, Princeton University Press, Princeton, (1998) 142-152.

17
[11] Santorelli, A., Krause, D. and Sant’Anna, A., “A critical study on the concept of identity
in Zermelo-Fraenkel like axioms and its relationship with quantum statistics”, Logique &
Analyse 189-192, (2005) 231-260.
[12] French, S. and Krause, D., Identity in Physics: A historical, Philosophical, and Formal
Analysis, Oxford University Press (2006).
[13] Krause, D., “Why quasi-sets?”, Boletim da Sociedade Paranaense de Matematica 20, (2003)
73-92.
[14] Post, H., “Individuality and physics”, The listener 70, 534-537; reprinted in Vedanta for
East and West 32, (1963), 14-22, cited in [12].
[15] F. A. Muller and S. Saunders, “Discerning Fermions”, Br. J. Philos. Sci., 59 (2008) 499-548.
[16] F. A. Muller and M. P. Seevinck, “Discerning elementary particles”, Philosophy of science,
76, (2009) 179-200.
[17] Schrödinger, E., “What is an elementary particle?”, reprinted in Castellani, E. (ed.), In-
terpreting bodies: classical and quantum objects in modern physics, Princeton Un. Press,
(1998) 197-210.
[18] G. Birkhoff and J. von Neumann, Annals Math.37 (1936) 823-843.
[19] H. Putnam, Is Logic Empirical? Boston Studies in the Philosophy of Science, vol. 5, eds.
Robert S. Cohen and Marx W. Wartofsky (Dordrecht: D. Reidel, 1968), pp. 216-241.
[20] M. Pavičić and D. Megill, “Is Quantum Logic A Logic?”, published in “Handbook of Quan-
tum Logic and Quantum Structures”, Vol. “Quantum Logic”, Eds. K. Engesser, D. Gabbay,
and D. Lehmann, Elsevier, Amsterdam, (2008) 23-47.
[21] G. W. Mackey, Amer. Math. Monthly, Supplement 64 (1957) 45-57
[22] J. M. Jauch, Foundations of Quantum Mechanics, Addison-Wesley, Cambridge, (1968).
[23] C. Piron, Foundations of Quantum Physics, Addison-Wesley, Cambridge, (1976).
[24] G. Kalmbach, Orthomodular Lattices, Academic Press, San Diego, (1983).
[25] G. Kalmbach, Measures and Hilbert Lattices World Scientific, Singapore, (1986).
[26] V. Varadarajan, Geometry of Quantum Theory I, van Nostrand, Princeton, (1968).
[27] V. Varadarajan, Geometry of Quantum Theory II, van Nostrand, Princeton, (1970).
[28] J. R. Greechie, in Current Issues in Quantum Logic, E. Beltrameti and B. van Fraassen,
eds., Plenum, New York, (1981) 375-380.
[29] S. P. Gudder, in Mathematical Foundations of Quantum Theory, A. R. Marlow, ed., Aca-
demic, New York, (1978).
[30] R. Giuntini, Quantum Logic and Hidden Variables, BI Wissenschaftsverlag, Mannheim,
(1991)

18
[31] P. Pták and S. Pulmannova, Orthomodular Structures as Quantum Logics, Kluwer Academic
Publishers, Dordrecht, (1991).
[32] E. G. Beltrametti and G. Cassinelli, The Logic of Quantum Mechanics, Addison-Wesley,
Reading, (1981).
[33] M. L. Dalla Chiara, R. Giuntini, and R. Greechie, Reasoning in Quantum Theory, Kluwer
Acad. Pub., Dordrecht, (2004).
[34] A Dvurečenskij and S. Pulmannová, New Trends in Quantum Structures, Kluwer Acad.
Pub., Dordrecht, (2000).
[35] Handbook Of Quantum Logic And Quantum Structures (Quantum Logic), Edited by K.
Engesser, D. M. Gabbay and D. Lehmann, North-Holland (2009).
[36] D. Aerts and I. Daubechies, “A characterization of subsystems in physics”, Lett. Math.
Phys., 3, (1979) 11-17.
[37] D. Aerts and I. Daubechies, “A mathematical condition for a sublattice of a propositional
system to represent a physical subsystem, with a physical interpretation”, Lett. Math. Phys.,
3, (1979) 19-27.
[38] C. H. Randall and D. J. Foulis, in Interpretation and Foundations of Quantum Theory, H.
Neumann, ed., Bibliographisches Institut, Mannheim, (1981) 21-28.
[39] G. Domenech, F. Holik and C. Massri, “A quantum logical and geometrical approach to
the study of improper mixtures”, J. Math. Phys., 51, (2010) 052108.
[40] F. Holik, C. Massri, and N. Ciancaglini, “Convex quantum logic”, Int. J. Theor. Phys., 51,
(2012) 1600-1620.
[41] G. Takeuti, “Quantum Set Theory”, Current Issues in Quantum Logic, E. Beltrametti and
B. C. van Frassen, eds., Plenum, New York, (1981) 302-322.
[42] S. Titani and H. Kozawa “Quantum Set Theory”, Int. Jour. Theor. Phys., 42, (2003)
2575-2602
[43] Redhead, M. and Teller, P., “Particles, particle labels, and quanta: the toll of unacknowl-
edged metaphysics”, Foundations of Physics, 21, (1991) 43-62.
[44] Redhead, M. and Teller, P., “Particle labels and the theory of indistinguishable particles in
quantum mechanics”, British Journal for the Philosophy of Science, 43, (1992) 201-218.
[45] Holik, F. “Aportes hacia una incorporación de la teorı́a de cuasiconjuntos en el formalismo
de la mecánica cuántica”, Master Thesis at the University of Buenos Aires, (2006).
[46] Holik, F. “Compound quantum systems: An Algebraic Approach”, PhD. Thesis at the
University of Buenos Aires (2010).
[47] Domenech, G. and Holik, F., “A discussion on particle number and quantum indistinguisha-
bility”, Foundations of Physics, 37, (2007) 855-878.

19
[48] Holik, F. “Neither name, nor number”, to appear in Probing the Meaning of Quan-
tum Mechanics: Physical, Philosophical, and Logical Perspectives, World Scientific;
arXiv:1112.4622v1 (2011).
[49] Domenech, G., Holik, F. and de Ronde, C., “Entities, Identity and the Formal Structure of
Quantum Mechanics”, arXiv:1203.3007v1, (2008).
[50] Domenech, G., Holik, F. and Krause, D., “Q-spaces and the Foundations of Quantum
Mechanics”, Foundations of Physics, 38, (2008) 969 - 994.
[51] R. Clifton and H. Halvorson, “Are Rindler Quanta Real? Inequivalent Particle Concepts
in Quantum Field Theory”, Br. J. Philos. Sci., 52, (2001) 417-470.
[52] L. Ballentine, Quantum mechanics: a modern development, World Scientific Publishing Co.
Pte. Ltd. (1998)
[53] Robertson, B., “Introduction to Field Operators in Quantum Mechanics”, Amer. J. Phys.,
41, (1973) 678.
[54] O. Bratteli, D. W. Robinson, Operator Algebras and Quantum Statistical Mechanics, v.2,
Springer, Berlin, (1997).
[55] P. de la Harpe and V. Jones, An introduction to C∗ -algebras (1995).
[56] B. Schroer, “Pascual Jordan, his contributions to quantum mechanics and his legacy in
contemporary local quantum physics”, arXiv:hep-th/0303241, (2003).
[57] A. Duncan and M. Janssen, “Pascual Jordan’s resolution of the conundrum of the wave-
particle duality of light”, Studies In History and Philosophy of Science Part B: Studies In
History and Philosophy of Modern Physics, Vol. 39, 3, (2008).
[58] P.A.M. Dirac, “The quantum theory of the emission and absorption of radiation”, Proceed-
ings of the Royal Society of London, Series A, Mathematical and Physical Sciences, 114,
(1927) 243-265.
[59] P. Mittelstaedt, The interpretation of quantum mechanics and the measurement process,
Cambridge University Press (1998).
[60] W. V. O. Quine, From a Logical Point of View, Harvard University Press, (1953), Chapter
V.
[61] J. B. Rosser, Logic for Mathematicians, McGraw-Hill, (1953).

20

You might also like