0% found this document useful (0 votes)
192 views94 pages

Quantum Time

This document proposes a theory of quantizing the classical infrastructures of gravity and the standard model, including space-time. It suggests representing the standard model vacuum and space-time continuum as a large thin dome with quantum transverse dimensions and classical longitudinal ones. The theory quantizes these infrastructures while preserving common Lie groups, conservation laws, and experimental accuracy. It represents the quantum state as a typed exterior algebra that merges space-time and quantum Hilbert spaces. Dynamics are specified by a history state vector that is local in a simplicial sense rather than differentially. In the singular organized limit, the history undergoes "superconducting" condensation, and the classical space-time manifold and imaginary i number emerge.

Uploaded by

Nguyen Hoang Vu
Copyright
© Attribution Non-Commercial (BY-NC)
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
192 views94 pages

Quantum Time

This document proposes a theory of quantizing the classical infrastructures of gravity and the standard model, including space-time. It suggests representing the standard model vacuum and space-time continuum as a large thin dome with quantum transverse dimensions and classical longitudinal ones. The theory quantizes these infrastructures while preserving common Lie groups, conservation laws, and experimental accuracy. It represents the quantum state as a typed exterior algebra that merges space-time and quantum Hilbert spaces. Dynamics are specified by a history state vector that is local in a simplicial sense rather than differentially. In the singular organized limit, the history undergoes "superconducting" condensation, and the classical space-time manifold and imaginary i number emerge.

Uploaded by

Nguyen Hoang Vu
Copyright
© Attribution Non-Commercial (BY-NC)
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 94

Quantum time

IN PROCESS
LAST UPDATE 2010.04.28

David Finkelstein
School of Physics
Georgia Institute of Technology

Abstract
Classical infrastructures of gravity and the standard model are quantized,
including space-time and the imaginary i. If one regards the standard-model
vacuum and its space-time continuum as a large thin dome with classical con-
tinuous longitudinal dimensions and quantum transverse ones, it is resolved
here into a truss dome of generalized spins. This quantization preserves the
common Lie groups and conservation laws as accurately as experiment requires
but bounds all variables. Its stat(e)vectors come from a typed exterior algebra
Q of real spinors that merges the space-time and the Hilbert space of quan-
tum theories. Its dynamics is specified by a history statvector in Q, local in
a simplicial sense, not differentially. The classical space-time manifold and
the imaginary i arise in a singular organized limit as the history undergoes a
“superconducting” condensation.
Keywords: Quantum gravity, quantum logic, quantum pragmatics, quantum
set theory, quantum topology, standard model, superconducting vacuum

1
Contents
1 Infraquantization 5
1.1 Quantizing the infrastructure Dated 2010.04.25 . . . . . . . . . . . 5
1.2 Outline Dated 2010.04.25 . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Infinity in, infinity out revised 2010.04.25 . . . . . . . . . . . . . . . 8
1.4 Eliminating the center Dated 2010.04.20 . . . . . . . . . . . . . . . 10
1.5 Merging products Dated 2010.04.20 . . . . . . . . . . . . . . . . . . 11
1.6 The sea of events Dated 2010.04.22 . . . . . . . . . . . . . . . . . . 11
1.7 Atoms of time and energy Dated 2010.04.14 . . . . . . . . . . . . . 12
1.8 Asymptotic freedom? Dated 2010.04.16 . . . . . . . . . . . . . . . . 14

2 Prior studies 15
2.1 Space-time quantization Dated 2010.04.14 . . . . . . . . . . . . . . 15
2.2 Feynman, Yang, and Penrose quantum spaces Dated 2010.04.20 . . 17
2.3 The reform of the Heisenberg-Poincaré algebra . . . . . . . . . 17
2.4 The reform of Bose statistics revised 2010.04.17 . . . . . . . . . . . 19
2.5 The Umklapp problem Dated 2010.04.14 . . . . . . . . . . . . . . . 19

3 The cosmic dome Dated 2010.04.08 20


3.1 Infraquantum relativity Dated 2010.04.09 . . . . . . . . . . . . . . 22

4 The imaginary quantum i Dated 2010.04.20 24


4.1 Time-slicing Dated 2010.04.24 . . . . . . . . . . . . . . . . . . . . . 27

5 Quantum sets Dated 2010.04.25 28


5.1 Spin-statistics equality . . . . . . . . . . . . . . . . . . . . . . . 31
5.2 Numbering the sets . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.3 Reversible infraquantum theory . . . . . . . . . . . . . . . . . . 34
5.4 Digression on Minkowskian forms . . . . . . . . . . . . . . . . . 35

6 Quantization as quantification 35
6.1 Cellular locality Dated 2010.04.25 . . . . . . . . . . . . . . . . . . . 39
6.2 The spin tree Dated 2010.04.25 . . . . . . . . . . . . . . . . . . . . 40
6.3 Infraquantum Pauli adjoints . . . . . . . . . . . . . . . . . . . . 41
6.4 Cumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
6.5 Negative probability . . . . . . . . . . . . . . . . . . . . . . . . 43

7 Dynamical law Dated 2010.04.29 45


7.1 History statvectors Dated 2010.04.29 . . . . . . . . . . . . . . . . . 46
7.2 Infraquantum scattering Dated 2010.04.21 . . . . . . . . . . . . . . 47
7.3 Infraquantum propagators Dated 2010.04.21 . . . . . . . . . . . . . 49

2
7.4 Dynamics as ambience Dated 2010.04.21 . . . . . . . . . . . . . . . 51
7.5 Organizational entropy Dated 2010.04.23 . . . . . . . . . . . . . . . 52
7.6 Infraquantum ground statvector . . . . . . . . . . . . . . . . . . 54
7.7 Infraquantum organization . . . . . . . . . . . . . . . . . . . . . 55
7.8 Unbreaking symmetry . . . . . . . . . . . . . . . . . . . . . . . 55

8 Gauging as quantification 55
8.1 Gauge in infraquantum space-time . . . . . . . . . . . . . . . . 57

9 Infraquantum space-times 59
9.1 The minimum black hole . . . . . . . . . . . . . . . . . . . . . . 59
9.2 Infraquantum spin-orbit analysis . . . . . . . . . . . . . . . . . 60
9.3 Infraquantum chirality . . . . . . . . . . . . . . . . . . . . . . . 61
9.4 The infraquantum dome . . . . . . . . . . . . . . . . . . . . . . 62

10 Generations 63

11 Particle valence 65
11.1 Standard model valences . . . . . . . . . . . . . . . . . . . . . . 65
11.2 Q valence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

12 Infraquantum fields 68
12.1 Q gaugeons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
12.2 The emergence of space-time . . . . . . . . . . . . . . . . . . . 71

13 Infraquantum standard model 71


13.1 Infraquantum fermions . . . . . . . . . . . . . . . . . . . . . . . 72
13.2 Infraquantum fermion dynamics . . . . . . . . . . . . . . . . . . 76
13.3 Constructing the Higgs field . . . . . . . . . . . . . . . . . . . . 77
13.4 Constructing the electroweak gaugeon field . . . . . . . . . . . 77
13.5 Constructing the strong gaugeon field . . . . . . . . . . . . . . 78
13.6 Constructing the gravitational field . . . . . . . . . . . . . . . . 78

14 Infraquantum fermion dynamics 78


14.1 Composite gaugeons . . . . . . . . . . . . . . . . . . . . . . . . 79
14.2 Dirac spin operators . . . . . . . . . . . . . . . . . . . . . . . . 79
14.3 Fermion annihilation operators . . . . . . . . . . . . . . . . . . 79
14.4 Pauli form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
14.5 Integration over space-time . . . . . . . . . . . . . . . . . . . . 80
14.6 Sources ξ, ξ. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
14.7 An imaginary factor . . . . . . . . . . . . . . . . . . . . . . . . 80
14.8 The orbital momentum . . . . . . . . . . . . . . . . . . . . . . . 80

3
14.9 Dirac operator . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

15 Discussion 81
15.1 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . 82

16 Appendices 82

17 Quantum pragmatics Dated 2010.04.28 82


17.1 The quantum i . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
17.2 The quantum cosmos Dated 2010.04.24 . . . . . . . . . . . . . . . . 86
17.3 Clifford algebras of Q . . . . . . . . . . . . . . . . . . . . . . . 87
17.4 Glossary Dated 2010.04.25 . . . . . . . . . . . . . . . . . . . . . . . 90

4
1 Infraquantization
1.1 Quantizing the infrastructure 2010.04.25

Nature seems to exhibit more dimensions in the small than in the large. In
this respect it is a large thin dome of events, with some cosmic ”longitudinal”
dimensions, and some submicroscopic ”transverse” ones. Its long dimensions
support the orbital coordinates xµ , pµ of space-time position and momentum,
and have cosmic extent. Its short dimensions, for example spins, are short in
the absolute sense that their coordinates take on few values. Of these, the
spins are longitudinal in the sense that they take part in rotations and boosts
of the large dimensions. Isospins are transverse in the sense that they do not,
but have their own transformations.
In working theories today, the dome is a continuum with quantum variables
attached to each point. Its infinitude of degrees of freedom causes some sig-
nificant physical computations to diverge, but a finite discrete framework like
a lattice would generally violate important conservation and invariance laws.
Quantization is a way to repair divergent theories by simplifying their Lie
algebras, originally the commutative ones. It has already eliminated infini-
ties like the heat capacity of ovens and the ground energy of hydrogen atoms,
without violating conservation of energy or momentum. Quantizing the in-
frastructure by simplifying its Lie algebras—“infraquantization”— similarly
eliminates the remaining infinities.
It represents the dome as a finite network or complex composed of quan-
tum simplices. Its elementary events, the atomistic elements of history, are
now finite in number and quantum in nature, and connect into simplices with
familiar continuous symmetries.
This unification of the long and short is of a Feynman-Penrose kind rather
than a Kaluza kind, atomistic rather than continuous. It remodels the large
dimensions after the fashion of the small rather than conversely. This process
does not introduce manifold dimensions but removes them, replacing them by
many spinlike variables.
Experiment today, however, seems rather consistent with manifold theo-
ries like those of the canonical commutation relations and general relativity.
The main goal of this work is to find a crucial experiment for the quantum
infrastructure.
The present infraquantization [hopefully] accounts for the huge anisotropy
of the dome as one accounts for that of a soap bubble or a graphene: by
the way atomistic elements organize themselves into a coherent whole. Unlike
bubbles and graphene, however, the cosmic network does sit in a space-time
but replaces the space-time, as well as the observed short dimensions. As a

5
quantum system, the network dome is described statistically by a vector in a
many-dimensional vector space of other possible statistical descriptions, not
by a membrane in a larger classical space.
Such statistical vectors are often called “state vectors”. This term came
from a non-operational interpretation that doesn’t work. We interpret them
statistically and therefore call them “statvectors” or simply “statvectors” (§17).
The familiar axiomatization of quantum theory in Hilbert space omits an
indispensable construct, quantum statistics. Since quantum theory is statis-
tical, a one-quantum theory has little meaning. A quantum theory should
be able to construct descriptions of beams of quanta out of descriptions of its
quanta. Infraquantum statvectors are posited here to have a product represent-
ing the composition of quanta. They reside in a graded Lie algebra Q whose
grade counts quanta, and whose monadics—first-grade elements—represent in-
put or output processes for individual quanta. In the present infraquantization
the multiplication is exterior, so a product can be interpreted as a simplex, and
the addition is quantum superposition, so the simplex is a quantum simplex.
To form a complex of quantum simplices, one does not add them but braces
them and multiplies (§5). A quantum complex is a simplex of simplices, not a
sum of simplices.
The thread that leads us through the maze of possible simplicial complex
theories is a correspondence principle in the sense of Bohr. It is based like
Bohr’s on a Lie algebra homotopy that simplifies its Lie algebra, now of scale
parameters other than ~. This makes it a special case of the theory of group
deformation quantization. The general hypothesis that we adopt here is that
basically simple symmetries are deformed to compound ones, including gauge
symmetries, by singular organizing limits [16]. Examples:
• The group homotopies of Yang [48] and Segal [33] (§2.2).
• The group contraction of Inönü and Wigner [25].
• The symmetry rearrangement of Heisenberg, Umezawa [41], and Nambu-
Goldstone boson theory.
An infraquantization must undeform these deformations, or reform them.
Standard quantum field theories assume that there is a fixed dynamical
law, and that it is Poincaré invariant, unitary, and local. These seem to be
features of a singular limit. Infraquantization changes them all, slightly in the
usual experimental domain but eventually greatly.
Recall that the process of quantification converts a theory of an individual
to a theory of an unspecified quantity of similar individuals, and qualitative
yes-or-no predicates to quantitative how-many predicates. Here the individuals
are quantum. Quantum quantification is discussed in §6.

6
In most developments of physical theory, quantization and gauging are
separate and independent processes. Infraquantum theories carry out both by
iterated quantification.
In a canonical theory, field operators of quantum annihilation and creation
are functions on the domelike infrastructure, and so have a higher set-theoretic
(Quine) type than the dome events. To deal with such type-distinctions be-
tween quantum systems, infraquantization replaces Hilbert space H and its
dual Dual H by a finite-dimensional subalgebra FQ ⊂ Q of a typed exterior
algebra Q, defined in §5. FQ replaces H and Minkowski space-time too, which
return only as singular organized limits of Q. FQ also replaces Newton’s dif-
ferential calculus by a typed extension of Grassmann’s algebraic differential
calculus. Dynamical variables are finite matrices, hierarchically composed of
smaller finite matrices. The vertices of its simplicial complexes are lower-type
simplices, braced to serve as points.
The infrastructure of canonical quantum theories includes the complex
plane C of probability amplitudes as well as a classical space-time manifold.
While C is useful, it fails in structural stability and locality, two working prin-
ciples that seem indispensable. This is reason enough to consider seriously
whether a real quantum theory will not work better than a complex one.
It is proposed that the imaginary i of quantum theory is no more fun-
damental than that of Steinmetz’s electric network theory. The Steinmetz i
represents a quarter-wave phase shift of an external sine-wave generator, en-
abling us to replace d/dt by iω in the network differential equations. It loses
meaning if the external generator is significantly non-sinusoidal. Then one
returns to the more generally valid second-order differential equations. It is
proposed that the Heisenberg i likewise represents a symmetry of the ambience,
and that in a less symmetric ambience the first-order differential equations of
Heisenberg and Schrödinger must be replaced by one involving higher powers
of ∂t . The variable i [hopefully] becomes a Higgs field, imparting mass to
otherwise massless gaugeons that do not annul it.

1.2 Outline 2010.04.25

§1 motivates and outlines the work.


§2 gives some background.
§3describes nature phenomenologically, as a dome.
§4 explains the quantization of the imaginary i of quantum theory.
§5 sets up a quantum set theory Q useful for infraquantization.
§6 reduces quantization to quantification, constructs the spin tree Q, and in-
terprets its negative probabilities.
§7 formulates infraquantum dynamical laws.

7
§8 relates gauging to quantification.
§9 discusses infraquantum space-times, the Umklapp problem and the chiral
spinors of the standard model.
§10 formulates hypotheses for a microscopic quantum theory of the three
fermion generations.
§11 injects the chemical concept of valence and the covalent bond for the
binding of space-time on one type-level and of gaugeons on the next.
§12 composes the standard model gaugeons and gravitons out of four fermions
bound by deep covalent exchanges.
§13 relates the main condensates of the standard model to those of the in-
fraquantum dome.
§14 [hopefully] sets up a fermionic infraquantum dynamics that may suffice to
control the gaugeons too.
§15 sums up consequences of this theory so far, with grateful acknowledgments.
§16 consists of appendices on general quantum theory and on Clifford and
Fermi algebras, and a glossary.

1.3 Infinity in, infinity out revised 2010.04.25

In hindsight the earthquake warnings in classical mechanics are clear, precur-


sors of the quantum theory to come. They can alert us to similar precursors
in the canonical quantum theories of today.
The optical phenomenon of quantum superposition studied experimentally
by Newton and at least theoretically by Malus was not taken seriously, since
the evidence for the existence of photons was not. The earliest warnings to be
heeded were the noxious infinities in oven heat capacity and in the hydrogen
atom ground energy. Planck’s quantum of action entered physics as a cut-off
for the infinite heat capacity of an oven.
To be sure, the standard model of today is renormalizable, but each infinite
renormalization constant renounces one finite prediction of the theory, and
the renormalization process is non-local, involving Fourier transformations.
Current renormalizations frustrate the prediction of the cosmological constant
stress, for example, and canonical quantum gravity seems to require so many
infinite renormalizations that no physical predictions remain. Since an infinity
is never observed, let us refrain from injecting infinities into our theories in
the first place. In the term of Pauli and Villars, let us perform a physical
regularization, not a formal one.
Umezawa [41], for example, speaks of two languages in quantum field the-
ories, “basic” and “phenomenological” (his quotation marks). The “basic”
expresses Lagrangians and fields, and the “phenomenological” describes quan-
tum annihilation and creation processes. Let us constrain ourselves to one

8
language of quantum annihilation and creation processes, taking a phenomeno-
logical language as basic. The Lagrangian language is then merely a singular
limit, badly defined, no more “basic” for field theory than Newtonian mechan-
ics is for quantum mechanics.
The distinction between creation and annihilation, however, is defined rel-
ative to the ambience, including the arrow of the observer’s time. To overide
this distinction relativistically, let us call elementary creation and annihilation
operations, or input and output operations, and their superpositions, collec-
tively, io operations. Stats represent io operations. The most basic statvectors
are monadics, which generate a graded algebra in which the monadics have
grade 1. The monadics of a canonical quantum theory form a Heisenberg Lie
algebra for bosons and a Fermi Clifford algebra for fermions. Collectively these
algebras and their reforms, interpreted statistically, will be calledstatalgebras.
The statvectors of the usual Hilbert-space formulation of a generic quantum
theory are, in this view, elements of a statalgebra whose product is forgotten.
The infraquantum statvectors belong to a finite-dimensional algebra of io
operations, an operational language. In what follows, it is a Clifford algebra FQ
with additional type-structure, and its statvectors concern quantum particles,
space-time events, and sub-events of a lower type.
Quantum theories with finite-dimensional statalgebra (or Hilbert space)
are termed regular after [7]. For example, a theory of a googol spins is regular,
100
since its Hilbert space dimension is only 2(10 ) , but not the theory of one linear
harmonic oscillator, which has ∞ dimensions. To be more phenomenological
than canonical quantum theories, let us assume regularity:

Physical quantum systems have regular theories. (1)

Theories with a statvector space of ∞ dimensions are then irregular, or sin-


gular. They are supposed here to arise in a singular limit, useful for replacing
some hard sums by easier integrals.
Infinite-dimensional spaces are especially useful in the study of physical
organization processes, for example the condensation from iron vapor of a
magnetized crystal of one domain. It is imagined that the magnetic field
of the crystal breaks the so(3) symmetry of the dynamics, the process often
referred to as “symmetry breaking”. More precisely, the condensation does
not break this symmetry but rearranges it [41]. In the Nambu-Goldstone
theory of symmetry rearrangement, the crystal is imagined to grow without
bound, and the group is contracted. This incurs divergences. A regular theory
of symmetry rearrangement does not go all the way to the limit of singular
Nambu-Goldstone bosons, but stops at regular Nambu-Goldstone Palevons (
§7.7 [hopefully]).

9
The regularity postulate encourages us to quantize until we reach a regu-
lar physical theory. Singular limits like the Poincaré group are less physical
than their regular origins. Each quantization changes the working algebra,
with consequences that may be small in the current experimental domain but
ultimately become large.
A working physical theory must revise the structure of field theory at length
scales much larger than the atom. If the Maxwell action is used, the zero-
point energy of the electromagnetic field just for wavelengths larger than the
atom suffices to curve the universe to fit into a sublunary sphere. The normal
ordering that blocks this disaster is non-local from the viewpoint of field theory,
yet natural within the io language. Infraquantization replaces the differential
locality of field theory by a simplicial locality natural to Q. In this sense the
normal ordering is local and the field-theoretic action non-local.

1.4 Eliminating the center 2010.04.20

One portent of impending quantization is excessive commutativity, which re-


sults in structural instability: an arbitrarily small change in the structural
tensor resulting in a non-isomorphic algebra [33]. Excessive commutativity
is marked by a singular Killing form, with determinant 0. This is unstable
because an arbitrarily small change in the structural tensor suffices to make
the determinant non-zero, implying a non-isomorphic algebra.
Quantization reduces the center of the Lie algebra. Canonical quantization
drastically reduced it to the one-dimensional algebra C of complex numbers,
generated by the right-hand side of the canonical commutation relation

[p, q] = i, [i, p] = 0, [q, i] = 0, (2)

among infinitesimal isometries of the quantum theory. Representing this i by


a scalar matrix blocks regularity—to see this, take the trace of the equation—
and is therefore responsible for all the infinities of present-day physics.
Call a system theory simple or semi-simple if all its basic Lie groups are
simple or semi-simple, respectively, and otherwise compound . Among the basic
groups is the automorphism group of the algebra of dunamical variables of the
theory. Therefore only a regular quantum theory can be simple in this sense.
Then its variables are all bounded, and its predictions finite. Therefore let us
provisionally assume a simplification strategy [48, 33]:

Simplify all physical algebras by “small” homotopies. (3)

One of these algebras is that of the infinitesimal automorphisms of the alge-


bra of operators, which must therefore become finite-dimensional. Therefore

10
simplification implies regularity. It also provides a correspondence principle:
The homotopies (3) correspond infraquantum constructs to singular. (4)
The familiar correspondence principle with ~ → 0 is a special case.

1.5 Merging products 2010.04.20

Another portent of quantization in classical mechanics is a plurality of prod-


ucts. This was insistently called to my attention by A. Petersen and E. Grgin,
and was pointed out by L. Flato and others. The variables of classical me-
chanics have two useful products, the ordinary numerical product pq and the
Poisson Bracket [p, q]P of dynamical variables. Canonical quantization fuses
them into one non-commutative operator product involving a constant ~. They
are re-separated by a power-series expansion in ~.
Space-time tangent vector fields have several useful products, such as the
commutative scalar product u·v, the anti-commutative exterior u×v, the tensor
product uv, and the Lie Bracket [u, v]L . The scalar and exterior products
indeed merge cosily into one, a Clifford product, already used in the Dirac
equation. The present infraquantization blends all these products into one
exterior algebra product.

1.6 The sea of events 2010.04.22

Any fermionic statvector algebra (see Appendix 17.4) is also a Clifford algebra.
The canonical Lie algebra of classical physics becomes a door to quantum me-
chanics when we identify it with the Bose statvector algebra of the constituent
quanta. Indeed, it is often called the Heisenberg Lie algebra, as though Hamil-
ton had nothing to do with it. Likewise one may use the Dirac Clifford algebra
of spin as a door to the lower quantum levels, by identifying it with the Fermi
algebra of sub-events, the quantum elements of quantum space-time events.
We assume that the Dirac Clifford elements too include not only spin opera-
tors but also fermionic statvectors, now for sub-events. Sub-events too have a
statvector algebra, for which the following notation is useful:
TQ is the subspace of Q of type T . The statvector algebras EQ1 for events,
DQ1 for event differences, and CQ1 for cell events, are constructed in §5.

Let us make the tentative assumption that the space-time tangent bundle is
a classical organized limit of a Fermi-Dirac system of infraquantum sub-events:
tan M ≺− Fermi EQ1 ; E
Q1 = ext(DQ1 ) = (ext)2 (CQ1 ). (5)
The term “vacuum” recalls the void of Newton, while classical field theory
recalls the plenum of Descartes. A quantum infrastructure, however, is neither

11
vacuum nor plenum, but may be called a “plexus”: a structure described by
a finitude of topological connections, here membership (ι) relations between
quantum simplices of several types.
A classical crystal has a discrete symmetry group. The quanta have a con-
tinuous one, however, currently including the Poincaré group and the standard-
model unitary groups. If the ambient medium is a crystal, as Newton and
Fresnel declared the ether to be, it must be a quantum and relativistic crystal,
a synthesis of the continuous and discrete. We model it as a quantum plexus.
Since this plexus does not collapse at once, even when it is cold, its elements
presumably obey Fermi statistics. They then may form a Fermi event sea .
This is our ambience. It [hopefully] becomes the standard-model vacuum in a
singular organized limit.
The hypothetical Dirac sea is made of particles with charge and mass,
resulting in infinite charge and mass density. We do not assign charge and mass
to events of the plexus. Charge and mass are, however, sources of curvatures
of different kinds. They are not properties of individual events but describe
defects in the organization of many events, like the Burgers(-Volterra) vector
of crystal physics. Events carry charge and mass no more than a gas atom
carries a Burgers vector.

1.7 Atoms of time and energy 2010.04.14

While Aristotle assumed infinitely many infinitely small instants of time, with-
out beginning or end, and infinitely many points of space, this leads to noxious
divergences already mentioned. To avoid infinities, let us seek meanings for
the constructs of space and time that are more operational.
Einstein associated a space-time point, or event, with a smallest possible
occurrence, for example a collision of two small bodies. The standard model
and general relativity still use this classical event construct, but it is obsolete.
Today we analyze the collision of the smallest bodies into supposedly quan-
tum input-output operations represented in standard theories by arrows in a
Feynman diagram. Unlike the classical ideal event of Einstein, which has only
space-time position coordinates, physical operons also manifest space-time mo-
mentum, spin, hypercharge, isospin, color, and a generation number. These
are all quantum variables in the standard model, coordinates in a quantum
space.
Fermion theories can be finite-dimensional and bosonic ones cannot. Sim-
plicity therefore seems to favor fermionic operons over bosonic. To be sure,
Palev has reformed the bosonic statistics [29] and Palev statistics can be finite-
dimensional; but a pair of fermions obey Palev statistics exactly. Let us there-
fore attempt to express all quanta as aggregates of fermions.

12
A fermionic statvector is a monadic—first-grade element—of an exterior
algebra. Let us therefore call the operon it represents a monad, with apology
to Leibniz and Aristotle. What a monad creates or annihilates can then be
called a monon for the nonce. In brief: Monadics are mathematical vectors,
that represent physical monads, that input or output monons; and analogously
for g-adics, tensors of grade g. In the present application to quantum history,
a monad is a single event of history. What it inputs or outputs depends on its
type.
Leptons and quarks are presumably global defects, with strength measured
by the mass and charge they carry. We nevertheless explore the possibility
that they can be associated with individual monons, as global defects like
disclinations and dislocations can sometimes be attached to single atoms of a
crystal.
We draw our statvectors from a typed exterior algebra Q (§6.2). Q replaces
both differential manifold and Hilbert space as a mathematical framework for
the theory of the quantum plexus. The space-time and momentum-energy
spaces of special relativity, and the Hilbert space of quantum theory, all merge
into subspaces of Q and return as singular organized limits.
The typical canonical commutation relation (2), suitable for a particle on a
line or a boson mode, is a singular organized limit of the regular commutation
relations
q , pb] = bı, [bı, qb] = Q2 pb, [b
[b p,bı] = P2 qb (6)
in which Q, P > 0 are small structural constants like ~. Rescaled, these rela-
tions define skew-hermitian so(3) rotation generators
q p bı ~ ×L
~ = L.
~
Lx = , Ly = , Lz = , L (7)
Q P QP
If we introduce an auxiliary external imaginary constant i for the moment,
we may use it to construct hermitian operators iLx , iLy , iLz , dimensionless
angular momenta that can also be regarded as occupation numbers, positive
or negative, counting angular-momentum quanta of size 1, “rotons”, around
the three axes. Then Q is the quantum of q, P is the quantum of p, and QP is
the quantum of bı. The rotator algebra (6) approaches the oscillator algebra (2)
as the physical coefficients Q, P → 0, provided that there are enough rotons of
all three kinds to imitate a continuum, and the vast majority are z-rotons:

so(3; R) −
 h(1, R). (8)

This Lie algebra can be applied to time and energy too, if we take q = t
and p = E. Their quantum units are designated by X and E. A sub-event
possessing one X of duration may be called a chronon To recover the complex

13
quantum theory in a singular limit, one goes to an aggregate of such oscillators
with a cumulative bı ≈ i.
The simple commutation relations (6) have been applied to particle coor-
dinates ([48], [43],[27], and others) and to Bose statistics [29].
The symplectic group Sp(2, R) = SL(2, R), mixing q and p and fixing i, is
a symmetry of (2), but not of (6). It is stable but infraquantization does not
preserve it. Infraquantization regularizes the commutation relations, not their
symmetry.
To recover the canonical quantum theory from such a simple theory, we
must perform a homotopy (3) of the group or Lie algebra, here letting Q, P → 0.
This centralizes the variable bı −
 i. To justify (6) we must find bı in experiment.

1.8 Asymptotic freedom? 2010.04.16

A symmetry between xµ and pµ can remain in the limit X, E → ∞. While


fields depend on xµ alone, not both xµ and pµ , after a Fourier transformation
they depend on pµ alone; this does not break the x ↔ p symmetry. But
canonical field interactions are local in x. They are not usually assumed to
be local in p. Thus canonical kinematics has x ↔ p symmetry but not the
canonical dynamics. We have to consider whether the dome organization can
rearrange the x ↔ p symmetry in the way that graphene formation rearranges
the x ↔ z symmetry, if the z axis is normal to the graphene plane and the x
axis is parallel.
One way to express such a symmetry rerarrangement would be to have
X/E → ∞ as X, E → 0. This would mean that in ordinary experiments we
deal with many fewer chronons than ergons.
Interactions weaken when interactants differ greatly in position. Therefore,
according to x ↔ p symmetry, they should weaken when the interactants
differ greatly in momentum. This suggests asymptotic freedom, though the
connection seems far-fetched: Asymptotic freedom has been understood as a
consequence of shielding arising from the specific nature of the non-abelian
color gauge group, an expression of locality in x alone. The argument given
here is based on locality in both x and p. It does not seem to approach the
non-abelian argument in the singular limit.
Since we assume that the regular variables have finite matrix representa-
tions, the matrices for pb and qb must grow very large on the way to the singular
organized limit, so that they can pass for continuous variables. We accomplish
this by quantifying over sub-events. Then the cumulative bı can polarize, anal-
ogous to spontaneous magnetization, allowing us to approximate bı by a local
constant i. This organization restricts the quasi-continuous variables pb and qb

14
to small values but leaves them variable. The singular organized limit shrinks
the three-dimensional p, q,bı-space to a two-dimensional p, q-phase space.
Let us refer to such a combined process of singular homotopy and organi-
zation as a singular organized limit or Limo . For example, we write schemat-
ically
qb −
 q = Limo qb, pb −  p = Limo pb, bı − hi = Limo r. (9)
In particular, the compound canonical commutation relations and Bose statis-
tics are singular organized limits of regular Lie algebras.

2 Prior studies
2.1 Space-time quantization 2010.04.14

The oldest surviving mathematical theory of physical space is Euclid’s. Me-


chanics enlarged Euclidean space to phase space, and special relativity to
space-time and its tangent bundle. The Euclidean group ISO(3) of space,
the Galilean group and the Poincaré group ISO(3, 1) of space-time, and the
canonical commutation relations are compound and must be simplified, ac-
cording to (3). We recapitulate some relevant history of this project, with
apologies to authors undoubtedly overlooked.
Historic classical simplifications of Euclid include the classical spherical,
hyperbolic, de Sitter, and conformal geometries. Differential geometry and
general relativity, however, introduced a new kind of classical non-simplicity,
with the diffeomorphism group Diff.
Quantum spaces form a newer line of simplified spaces that branched off
the line of classical phase-space geometries in 1924, when the quantum theory
was invented, and further In 1930, when a quantum space-time was proposed
in order to avoid infinities [47, 2].
R. P. Feynman, as graduate student, quantized space-time by replacing
commuting coordinate differentials dxm with Dirac spin operators γ µ as in
(11) [17]. He interrupted this work to study the Lamb shift at Bethe’s behest
but never completely dropped it.
Oppenheimer carried the idea of quantized space-time from Heisenberg to
Snyder, who then quantized space as well as phase space [35, 47]. The Snyder
algebra is still compound, however. C. N. Yang replaced it by the algebra
so(5, 1) precisely to make it simple [48]. Where Feynman had quantized four-
dimensional space-time, Yang quantized eight-dimensional phase space, the
cotangent bundle of space-time. Conformal so(5, 1) acts on a classical space-
time of four dimensions, but Yang so(5, 1) acts on a quantum phase space of

15
15 dimensions. It uses quantum constants of speed c, action h, time X, and
energy E, to make the orbital variables dimensionless.
On another track, Segal [33], attempting to understand and forecast the
evolution of quantum physics, argued on Darwinian grounds that physics is
evolving toward simple Lie algebras: A compound Lie algebra has a singular
Killing form, and an arbitrarily small change in its structure tensor makes it
non-singular. As measurements of the structure constants improve, therefore,
a compound Lie algebra has survival probability 0 relative to its semi-simple
neighbors, which outnumber it ∞ to 1. Natural selection favors simple Lie
algebras.
Since the isometry group of the statvector space of a quantum theory is
one of these Lie algebras, simplicity in the sense of Segal implies regularity in
the sense of Bopp and Haag [7]:

A simple quantum theory is regular. (10)

Gerstenhaber, influenced by Segal, described homologically a rich terrain


of Lie algebras connected by homotopies, such as contractions [25], that carry
groups out of stable valleys of simplicity, to ridges between the valleys, and
up to singular peaks [23]. According to the simplicity principle, physics is
currently a glacier flowing down the simplicity-gradient to valleys in Gersten-
haber land. The Galileo Lie algebra is on a ridge between the valleys of Lorentz
so(3, 1) and orthogonal group so(4).
Group contraction [25] and deformation quantization [4] are also homotopy-
based theories of the evolution of physics. They do not heed the simplicity
principle but may rest on classical space-times.
R. Penrose, as a graduate student, quantized the Euclidean 2-sphere by
representing its points as directions of sums of many Pauli spins [30]. His
space quantization was extended to a space-time quantization [18].
Vilela Mendez, inspired by Gerstenhaber, rediscovered the Yang group [43]
and proposed physical consequences. It has been rediscovered several times
since.
The simple quantum spaces (11) of Penrose [30], Feynman [17], Segal [33],
and Yang [48], encountered in that order, greatly influenced the present study.
Feynman and Penrose both analyze coordinates into a finitude of spins, but not
the conjugate momenta, it seems. Yang quantizes momentum too, replacing
canonical commutation relation like (2) with spin-like commutation relations
like (6).
Segal posed and explained the simplicity principle and rediscovered the
Yang space. The phase space of special relativity was later quantized into
half-spins with Fermi statistics [20].

16
Galiautdinov [22], Shiri-Garakani [34], Bayer ([5], and others have stud-
ied Yang-invariant physics in Q. Baugh [3] represented the Yang so(5, 1) Lie
algebra in sl(6R) much as we do here.

2.2 Feynman, Yang, and Penrose quantum spaces


2010.04.20

Let us compare the quantized orbital variables of the simple spaces mentioned,
using chEX units. Here x b is a quantized x; k ∈ 3; m ∈ 4; and δ indicates a
finite difference to be summed later:
Feynman [17] xm
δb ∼ γm, δpm
= ?
Yang [48] bm
x ∼ i(g mn η 5 ∂n − η m ∂5 ), ∼ i(η6 ∂m − ηm ∂6 ).
pbm
Penrose [30] xk
δb ∼ σk , δbpm
= ?
Present xm
δb ∼ γ m5 , ∼ γm6 .
δbpm
(11)
The Penrose and Feynman quantum spaces still assume an absolute space or
space-time, with coordinates corresponding to xµ but not pµ . Simplicity led
Yang to relativize space-time within a larger quantum phase space, Yang space,
including momenta, boosts and angular momenta, as well as a complex plane.
These spaces are inadequate today. Standard model operons carry a hyper-
charge y, a generation variable Γ, three isospin variables τ k , four space-time
position variables xm , four momentum-energy variables pm , four Dirac spin
variables γ m , and eight color charges χc . In addition there are bosonic fields:
12 gauge 4-vectors Γm A , a Higgs field φ, and gravity. Infraquantizations must fit
all these variables and their commutation relations into the operator algebra
Lin Q (§14).
Yang, like Snyder, represented his algebra by differential operators on an
infinite-dimensional function space, as shown in (11). Simplicity in the present
sense requires a finite-dimensional representation like Feynman’s and Pen-
rose’s. This is the bottom line of (11). These finite-dimensional representations
must have indefinite metrics, and so do not fit into Hilbert space. A physical
interpretation of indefinite metrics and negative probability was indicated by
Dirac [13].

2.3 The reform of the Heisenberg-Poincaré algebra


The commutation relations of the Yang generators Lm0 m = −Lmm0 ∈ so(6 −
n, n), acting on a statvector space 6R with metric tensor gm0 m , are
[Lm000 m00 , Lm0 m ] = gm000 m0 Lm00 m − gm00 m0 Lm000 m + gm00 m Lm000 m0 − gm000 m Lm00 m0 ,
(12)

17
with m, m0 , m”, m000 ∈ 6.
DO: Check signs.
The commutation relations of the Heisenberg-Poincaré Lie algebra of space-
time position xµ , momntum-energy pµ , and Lorentz generator Lµ0 µ , are
[Lµ000 µ00 , Lµ0 µ ] = gµ000 µ0 Lµ00 µ − gµ00 µ0 Lµ000 µ + gµ00 µ Lµ000 µ0 − gµ000 µ Lµ00 µ0 ,
[Lµ00 µ0 , xµ ] = gµ0 µ xµ00 − gµ00 µ xµ0 ,
[Lµ00 µ0 , pµ ] = gµ0 µ pµ00 − gµ00 µ pµ0 ,
[Lµ00 µ0 , i] = 0,
[xµ0 , pµ ] = igµ0 µ ,
[xµ , i] = 0,
[pµ , i] = 0. (13)
with µ, µ0 , µ”, µ000 ∈ 4 and ~ = 1.
To deform (12) −  (13), one considers a representation of (12) on a statvec-
tor space V = 2(2N + 1)R with large dimension 2(2N + 1), so that the spectra
of the Lm0 m are quasicontinuous, and focuses attention on a polarized sector
Vpol ⊂ V , supposed to represent ordinary experience, in which |L65 | is close to
its maximum eigenvalue N and the other components of Lµ0 µ are much smaller,
though still quasicontinuous. Let us set
xµ = XLµ5 ,
pµ = ELµ6 ,
bı = EXL65 . (14)
Then X is the quantum of xµ , a natural unit of time; call it the chron. E is the
quantum of pµ , a natural unit of energy; N is the maximum magnitude of any
component of Lm0 m in this representation, and in that sense is the maximum
number of rotons; and bı2 ≈ −1 in Vpol . In the limit
E, X → 0, N → ∞, with NEX = 1, (15)
the Heisenberg-Poincaré relations (13) follow.
The Casimir operator for Yang so(6 − n, n) is then
0 bµ x
x bµ pbµ pbµ
0
Lm m Lm0 m = Lµ µ Lµ0 µ ++ 2 + g 66 g 55 N2 (bı)2 . (16)
X2 E
This approximates the square of the mass in E units if X  E subject to (15).
DO: Classify representations of so(6 − n, n) as Wigner classified representa-
tions of the Poincaré group.
DO: Find a correspondence between the two families of representations. Note
that the dome breaks so(6−n, n), while the vacuum is invariant under iso(3, 1)
and that Wigner uses Heisenberg statvectors while infraquantization uses his-
tory statvectors.

18
2.4 The reform of Bose statistics revised 2010.04.17

The commutation relations for Bose statistics define a canonical Lie algebra,
and so require reform. Palev [29], also explicitly seeking simplicity, reformed
the canonical algebra of Bose statistics to a simple Lie algebra, such as so(N ).
Pairs of Q fermions do not exactly obey Bose statistics. It is easy to show
that they obey a Palev statistics [29] defining an so(N ) Lie algebra of which
the canonical Lie algebra of Bose statistics is a singular organized limit. In
infraquantum theories, therefore, it is natural to represent all empirical bosons
as even polyads, with fermionic hard cores. It is understood that this hard
core must show up in high-energy photon-photon collisions.

2.5 The Umklapp problem 2010.04.14

When Heisenberg proposed to quantize space-time, Pauli pointed out the


bounce problem (Umklapp Problem) [47], and Feynman pointed it out to me:
The canonical commutation relations imply that the canonical conjugate
to a circular variable is a lattice variable, one with a discrete uniformly spaced
spectrum. For example, the conjugate to an angle θ ∼ = θ + 2π is an angular
.
momentum L = nh. Conversely, if the space coordinates are lattice variables,
x = nX, then the conjugate momenta are circular, p ∼ = p + 2πnE. Then the
high momentum p = πnE in one direction would be indistinguishable from
an equally high momentum −p in the reverse direction. The particle can
bounce from p → −p with no applied force. This bounce is well known for
phonons in crystals. There the phonon bounces off the crystal, which accepts
the lost momentum. But particles are not observed to bounce off the vacuum.
Therefore the ambient event space is probably not a lattice of that kind.
Infraquantization replaces the canonical commutation relations with spin
commutation relations, not a lattice, though each component of total spin has
a spectrum like the coordinates of a lattice. This replaces the bounce by a
bound. While the Umklapp momentum for spacing X is ~/X, the bound in
plexus momentum is another quantum constant NE/c~  ~/X. But spin “up”,
Lm6 = N, is not identified with spin “down”, Lm6 = −N, so there is no bounce.
An infraquantum space-time has no Umklapp problem.
Moreover, to attain the bound Lm6 = N, all N spins in the sum for pm
must align in the +m6 direction. Since [L56 , Lm6 ] ∼ Lm5 , this alignment is
complementary to the dome alignment of L56 that organizes bı − i, unless
.
h Lm5 i = 0.

19
3 The cosmic dome 2010.04.08

The plenum of canonical field theory emerges from a plexus of fermionic sub-
events as a singular organized limit. We do not use this limit here since it
diverges. Instead we study a symmetry rearrangement [41] that preserves the
full Yang symmetry for the dome condensate. This keeps some dimensions of
the statvector space that are usually thrown away, but the total number is still
finite, while for the singular limit it is infinite.
The standard model and the semiclassical theory of gravity describe a dy-
namically variable cosmic dome. The continuous space-time coordinates of
field theory represent its longitudinal dimensions. The unitary charges rep-
resent its transverse dimensions. A quantum spin of the Lorentz group is
attached to each operon, and can be considered longitudinal but short, quan-
tum, and regular; in contrast to the orbital variables, which are longitudinal
and long, and presently singular, still undergoing the singular diffeomorphism
group of general relativity. Chamseddine, Connes, and Marcolli [11] work with
such a singular dome.
The Q language of §5 enables one to resolve this dome into a space truss
of variable topological structure, a typed simplicial complex, with similar spin
variables for both longitudinal and transverse struts, and no continuous vari-
ables.
The Yang SO(6) group (11) implies that in a singular organized limit bı −i
there is a symplectic symmetry between position xµ and momentum pµ . Our
ambient dome reduces the Yang symmetry. Its variables xµ vary over ranges
of cosmological magnitude, while momenta pµ , ordinarily undefined for the
empty manifold, may be taken to be 0 in the ambient “vacuum”.
Something similar occurs for a bubble or a graphene flake. Organized
into a graphene, carbon atoms collectively define a 2 + 1-dimensional quasi-
manifold. The manifold (x,y,t) coordinates range over macrocosmic ranges
defined by the graphene. The carbon momenta obey the constraint px , py ≈ 0
in the manifold coordinate system due to the same binding forces that define
the quasi-manifold (x, y, t) coordinate systems. One cannot carry such results
from a graphene to a quantum plexus because graphene theory assumes a
background space-time manifold; but they are suggestive.
It is natural to enlarge the Yang so(3, 3) Lie algebra with the unitary
charges. In the standard model each operon also has a Lorentz spin, here
termed longitudinal, and there is no symmetry between the longitudinal and
transverse spins. We suppose that such an extended symmetry exists for the
kinematics of the individual operon, and is badly broken as a symmetry by the
organization of the dome. Unitary charges remain after the dome organization,
as spins transverse to the dome, along its smallest dimensions. Lorentz spins,

20
similarly, survive as short longitudinal variables, of magnitude ∼ X. Orbital
dimensions (x, p) are longitudinal and long.
Standard theories, to be sure, assume that the proper times between phys-
ical operons depend solely on their orbital variables and not on their spins
or charges, longitudinal or transverse. Spins and charges are usually given
space-time length 0.
This assumption makes the chronometric form singular. Therefore the
theory is structurally unstable. There is a natural invariant metrical form
for a spin plexus, and it defines non-zero proper times for longitudinal and
transverse spins, though they seem to be too small to be detected yet. Then a
change in a longitudinal spin generally changes the proper time between two
operons, while a change in a transverse (unitary) spin does not.
The plexus then resembles a truss dome or graphene that is as wide and
old as the cosmos but too thin to measure as yet. In an infraquantum theory
the dome is a typed simplical complex, whose vertices are generated by the
membership operation ι and multiplicative composition ∨.
The imaginary spin dyad bı − i should be considered transverse but long,
since it is macrocosmic, involving many cells, but polarized and frozen. If the
longitudinal dimensions are regarded as angular or azimuthal, the transverse
dimension of bı is radial. We are to narrate the the structure and excitations
of the dome within Q as parsimoniously as possible.
In such theories, particles are propagating excitations of the dome. They
may define representations, not of the unstable Poincaré group, as Wigner
proposed, but of the symmetry Lie group G[E] of the dome. The possible Yang
groups SO(5, 1), SO(4, 2), SO(3, 3), are simple but mix space-time, momentum,
and complex-plane dimensions, and so are broken by the dome. One of them
may still serve as group G[C] of a generic cell of the dome, on a lower type-level
C < E that includes the standard model unitary charges. Then the casimir
(operator) of the Yang Lie algebra includes and unifies casimirs of the Lorentz
group, the complex-phase group U(1), and the infraquantized proper time and
proper mass. The casimir of the cell Lie algebra similarly unifies the casimirs
of the Yang group and the unitary charges.
The physical constants entering into the Feynman or Yang groups are the
speed and action units c, h of earlier groups, elementary time and energy units
X and E with XE = h, and a cosmological integer N. Under either group, time
is just energy measured in huge units. The vacuum is supposed to reduce the
Yang casimir to the Minkowski form in an singular organized limit X, E →
0, N → ∞.
Infraquantization iterates Fermi quantification, which is represented by the
functor ext, forming the exterior algebra of a statvector space. Quantization,
general relativization, and gauging then become special cases of quantification.

21
Multiple Bose quantification has also been proposed [40, 46].
Nowadays there seem to be three variables with important non-zero vacuum
values:
• Gravity’s gµν breaks space-time sl(4R).
• Higgs hi breaks weak su(2).
• The quantized imaginary bı breaks complex-plane sl(2R) and time reversal
T ∈ SL(2R).
Let us designate the regional mean values by g, h, ı. There may be close func-
tional relations among these three varables. Both i and gνµ distinguish time
from space in their way; i responds to time reversal but not to space-reflection,
while gνµ responds to neither. Particle masses may derive from couplings to
h, but are defined as couplings to g.
It has been supposed that g and h are statistical parameters of the vacuum
statvector of the system ([38], for h). The global concept of a vacuum statvec-
tor, however, depends on the assumption that the dynamics is homogeneous
under time translation. This assumption may not work for time-shifts that
are too small or too large. In an infraquantum theory it may be impossible in
principle to carry out time translations by 10100 s or 10−100 s, as assumed in
canonical theories.
More conservatively, let us replace “vacuum” by a regional description of
the dome, which is mostly metasystem, and has only approximate symmetries,
to be determined experimentally.

The ambient g, h, ı values are statistical parameters of the dome. (17)

The hypothetical underlying quantum variables g, h, i may depend on spin


variables of the plexus that we ignore when we study its small excitations.
They may be slightly modulated by the system and the experimenter, and
may vanish completely in a local dome melt-down. The ambient means are
closer to experiment than the underlying quantum variables, which are still
highly theoretical.

3.1 Infraquantum relativity 2010.04.09

The special relativity group ISO(3, 1) reduces the observer to but 10 degrees
of freedom, like a speck of crystal adrift in Minkowski space-time M. This
group is simplistic but not simple.
General relativity over-compensates. While it works empirically in the
classical macrocosmic realm, its gauge group Diff represents observers with
a continuous infinity of degrees of freedom, as if they were infinitely elastic

22
organisms, cosmic super-squids, unaffected by gravity or distortion, crossing
light-cones freely in both directions. General relativity is even more singu-
lar than special relativity. But it is also points toward simplicity, in that it
eliminates the toxic commutativity of the covariant or kinetic momentum com-
ponents. And it inspired the Yang-Mills theory of gauge and thus the standard
model.
In standard quantum theory there is a unitary relativity group U(∞) re-
lating different frames in Hilbert space. Usually the quantization of gravity is
taken to require representing Diff within U(∞).
Let us seek a still better form of relativity, with observer somewhere be-
tween infinitesimal speck and infinite squid, that supports an infraquantum
regularization of the standard model. The strategy is to morph our theo-
ries from the language of H and M to that of some finite-dimensional “field
type” FQ ⊂ Q, without changing them drastically. This is the project of an
infraquantum relativity, with memorandum

Infraquantization : H, M 7→F Q. (18)

Simplicity requires an infraquantum relativity to unify all the variables of


the standard model, and to account for their separation in the standard model
by a dynamical organization.
A regular quantum theory in the sense of (1) has to renounce either uni-
tarity or Lorentz invariance, since the non-compact Lorentz group will not fit
into a compact unitary group.
For example, the regular space D ∼ 4R of real Majorana Dirac spinors is
Lorentz invariant but has no invariant definite metric form. Instead it has an
invariant indefinite bilinear form, the Pauli form often designated by β, and
a cone of definite non-invariant Hilbert forms h, associated with possible time
axes.
Since both definiteness and Lorentz invariance accord well with experiment,
this is a significant conceptual problem. It is taken up in §6.5
Infraquantization incurs a heavy start-up obligation: to reconstruct the ex-
perimental g, h, and i, as singular organized limits of their infraquantizations,
g−
b  g, b
h−  h, and bi −
 i.
Yang space at last eliminates the problem of large zero-point energy. The
commutation relations of the Yang space variables are those of a Lie algebra.
They have the generic form [L, L] = cL instead of [L, L] = i. Unlike the
canonical commutation relations, the Yang relations are satisfied by L ≡ 0.
The zero-point energy is zero in Yang space, or in any simple Lie-algebra
space. The singular organized limit of the condensate evidently does not attain
the zero-point of energy, since its commutation relations are canonical. The

23
condensate is indeed important for gentle measurements on large blocks of the
plexus, but it is disorganized at the zero point of energy.

4 The imaginary quantum i 2010.04.20

Electric circuit theory is a real theory. Steinmetz introduced i into it replace


trigonometric functions like cos ωt by exponentials like eiωt , which are easier to
combine. Steinmetz’s i represents a π/2 rotation in the plane of any variable q
and its time-derivative q̇. It enables us to factor the second-order differential
equations of circuits into two first-order equations. Its utility does not cause
engineers to doubt that circuit theory is basically real.
The imaginary i was introduced into quantum mechanics to permit first-
order differential equations, those of Heisenberg and Schrödinger. Hamiltonian
equations are indeed first order but Lagrangian equations need not be, and
history quantum theory is related more directly to the Lagrangian than to the
Hamiltonian.
We must consider the possibility that the quantum physicist’s i is no more
physical than the electrical engineer’s. The meter readings of both are real
numbers. Is it possible that i is really imaginary?
If so, then the now standard postulations of a Hilbert space over C and a
Heisenberg dynamical equation
dq d i
:= [ , q] = [H, q], (19)
dt dt ~
are not in the best possible agreement with practice. They proceed as if a
unique central i had physical meaning.
Most of the rules of quantum kinematics work equally well with statvector
spaces over the real field R, the complex field C, or the quaternion field H. The
absence of a bilinear tensor product seems to exclude H for practical purposes.
Stückelberg formed a quantum theory over R. A one-dimensional projector
of H(C) is a two-dimensional projector in H(R). A statvector ψ ∈ H(C)
becomes a random statvector eiθ ψ ∈ H(R) with a random phase angle θ and
a constant operator i, i2 = −1. There are strong experimental arguments for
H(C) over H(R): the phenomenon of elliptical polarization, the connection
between symmetry and conservation, the uncertainty relation [37] and the
correspondence between classical Poisson Brackets and quantum commutators,
all depend on i.
A quantum superposition of two beams is a third beam that is experimen-
tally recognized by the fact that it passes—with probability 1, this means—
through any filter that passes both beams. Experimentally, the superpositions

24
of two statvectors seem to form a two-real-parameter family of statvectors.
This indicates C; a one-parameter family implies R, a four-parameter family
H. If there are violations of i symmetry in nature, they hide even better than
violations of parity and time symmetry did.
As long as time is continuous, moreover, any discrete variable that is con-
tinuous must be constant. The choice between i and −i could propagate
throughout the connected universe in this way. But here we consider space-
times that are quantum, not continuous, and cannot use that argument. And
as an operator with i2 = −1, i has a continuous infinity of possibilities, not
just two.
Another difficulty of principle with an absolute i is that the centrality
assumption [i, o] = 0 is structurally unstable. A stabler condition is that the
commutator Lie algebra of the operator i and present observables o is simple,
but the commutators with i are small. This instability, however, seems to lead
to no infinities, and could be tolerated. We can muster no hard evidence that
i is not an absolute.
Nevertheless it moves: Assuming a fixed i throughout the cosmos vio-
lates Einstein locality, one of the most fruitful principles of the last century of
physics.
Therefore we continue the construction of a real quantum theory with a
quantized i. This includes replacing the Heisenberg Lie algebras by orthogonal
Lie algebras, which replaces Bose statistics by Palev statistics (§2.4). As (6)
illustrates, simplification may require a quantized imaginary bı −  i.
One especially simple way to quantize i is to imbed it in the quaternions,
where it is no longer central. This idea was put forward by Yang at a Rochester
Conference in 1959. The proper framework for it seems to be a real quantum
theory with an SO(3) gauge group and a dynamical i-field replacing the con-
stant i, defining the electric axis in electroweak isospin space. (One first effort
explored a quaternion quantum theory, which seems to be less relevant to par-
ticle physics.) The quantized i-field of the so(3) gauge theory then turns out to
be a Higgs field [39, 1]. This is the closest we come so far to a phenomenological
justification for quantizing i.
More generally, to correspond with the standard complex theory, a real
quantum theory will have an infraquantized imaginary bı in its operator algebra,
with a correspondent i ≺− bı, possibly depending on the reference frame, as
with the relativization of time. Yang’s bı = L65 [48], suitably normalized, is
such an operator.
bı, the quantized i, has a universal direct coupling to all the other basic
fields, appearing as a multiplier for their action operators (§9.3); unlike the
Higgs field, which is not assumed to couple directly to the strong interactions.
The quantization bı ∼ Σ2 γ 65 also recalls the proposal i = γ 4321 of Hestenes

25
[24].
bı enters into the quantum skew-action operator with any action operator
δA:
δΦ = bı δA/h. (20)
Like angular momentum and energy, bı is a sum of contributions from every
system, including the metasystem. Let us hypothesize an organization akin to
polarization that centralizes (“superselects”) the resultant bı and results in the
imaginary unit of complex quantum theory: bı −  i.
The Heisenberg relation [xm , pm ] = ~i returns in a singular organized limit
but cannot hold for one cell, where the Yang relation is more accurate. There-
fore canonical quantum theory grossly overestimates zero-point energies in the
small, and the vacuum energy density. The coordinates, momenta, and bı are
now cumulated spins, of finite spectra. Measuring a cumulated spin does not
take indefinitely great momentum-energy in the way that measuring a position
coordinate is supposed to. It may merely incur a space-time meltdown.
Each subsystem contributes to bı as it does to the total angular momentum
tensor. When the system is gauged for the sake of locality, bı acts as a Stückel-
berg or Higgs field in giving mass to some components of the gauge field;
namely, those in the plane normal to bı. bı thus serves to define the electric
direction in isospin space. First noted for the quaternions, this feature of the
quantized imaginary generalizes to any Clifford algebra, including Q, and to
any gauge field.
An absolute i seems almost indispensable today because it was assumed
quite early that basic quantum equations like the Heisenberg Equation of Mo-
tion have to be of first order in ∂t , a formally skew-hermitian operator on
functions of time, and the Hamiltonian provided by classical mechanics is her-
mitian. In Dirac’s development, the Equation of Motion has to be first-order
because by definition a state gives complete information, and so should deter-
mine its own future uniquely. The development equation for the state must
then contain no derivatives higher than the first.
But a statvector is not a state. It does not give complete information, only
maximal information, almost all statistical. There is no compelling reason why
a statvector should determine its own development. The concept of history is
more general than the concept of a first-order differential equation. It puts the
construct of Lagrangian before that of Hamiltonian, and Lagrange’s theory is
not partial to first order equations. In a quantum history theory the basic
dynamical equations of quantum physics could be of higher degree than the
first. Then one could reduce them to the first degree by introducing new
variables and an i, but this could be done in many ways, and the differences
among them need have no physical significance.

26
The imaginary is then needed mainly for correspondence: between the gen-
erator of ∂t and the Hamiltonian; between commutators and Poisson Brackets;
and so forth. But then it may be a product of the classical limit, like corre-
spondence itself.
In the existing complex quantum theory, i is central and therefore strictly
conserved. This has non-trivial experimental consequences. Since this central-
ity assumption is not local, let us assume that i is only approximately central,
like the coordinates of a baseball or the components of the metric tensor, as
a consequence of a singular organizing classical limit. Then all singlets of the
complex quantum theory must in principle be resolved into doublets when the
organization of i is broken and the underlying real theory peeps through.
Absent a fixed i, we must give up the Heisenberg first-order form of dy-
namics (19). In a quantum history theory, Lagrange’s equations of motion are
more natural than Hamilton’s and are typically real second-order equations.
For example, Lagrange’s equation for a quantum oscillator takes the valid form
d dq
+ ω 2 q : (4∂bt )2 q + ω 2 q = 0 (21)
dt dt
(using the commutator symbol 4 of §17.4).

4.1 Time-slicing 2010.04.24

In canonical quantum theories assume time coordinates from the start. The
history statvector space V of a canonical quantum theory is then a tensor
product of time slices Vt O
V= Vt (22)
t
The canonical form form for a history dynamics statvector is a tensor product
E = UT, T −1 ⊗ UT −1, T −2 ⊗ . . . ⊗ U−T +1,−T (23)
in which each time t appears twice, for input and output.
If an infraquantum theory an event may still have a time coordinate opera-
tor b
t. Then its statvector space is a direct sum of b
t-eigenspaces, and its exterior
algebra is an exterior product of time slices. Several such slices may be needed
to determine the rest of the history statvector, subject to the dynamics. For
example a second-order theory can relate two final slices t, t − 1 to two earlier
ones t − 2, t − 3. A history quantum theory of order o is one whose dynamical
statvector can be factored into sub-complexes of
An example of a covariant Killing form of Lagrange’s equations of degree
2 in ∂t is
0
4Lm0 m 4Lm m q = M 2 q (24)

27
where Lm0 m is the tensor of generators of a supposed symmetry group, like
the Yang Lie algebra (12), and M 2 is a squared-rest-mass, or an operator
generalization.
In circuit theory one might define i as the operator

i := |∂t |−1 ∂t . (25)

This could work for the quantum theory too, but it does not look relativisti-
cally invariant. Theories like the Yang theory provide a covariant quantized
tensor that includes i as one component. Such theories have a tensor Lm0 m
of Lie-algebra symmetry generators, with a Lorentz subalgebra. The famil-
iar translation generators ∂µ are approximations to some of its components.
Lm0 m is supposed to have a non-zero vacuum expectation Lm0 m dominated in
an adapted frame by one enormous component, so that under ambient condi-
tions
Lm0 m = Lm0 m + δLm0 m , δLm0 m  Lm0 m (26)
0
For a covariant formulation, let the dominant component be θm m Lm0 m . Then
0
θm m Lm0 m
bı := m0 m . (27)
|θ Lm0 m |
—compare (25).
To account for the fact that the i-trick works so spectacularly well, we
should return to Einstein once again:

The dynamical law is of second order in the Lm0 m . (28)

Possibly this is the first approximation in an unending progression of better


ones with higher orders.

5 Quantum sets 2010.04.25

In the standard quantum theories, quantum variables are functions of time, or


space-time. They therefore belong to a higher membership level (= type, rank,
. . .) than the time variable. Quantizing time therefore quantizes a deeper level
than canonical quantization, and by correspondence calls for a typed quantum
theory, the simplest being a quantum set theory.
Classical sets are set as in stone. Quantum sets, however, are more fluid,
and are changed markedly by perception, in a way delimited by commutation
rules.
Let us begin with founded quantum sets. The foundation is some quantum
system with statvector space F, a finite-dimensional vector space, here taken

28
over R, the modification for complex vector spaces being trivial. The functor
ext, by definition, converts the statvector space F to the exterior algebra
Q1 (F) := ext F, said to be the statvector space for the quantum set of type 1
over F. Thus a quantum set is simply an aggregate with Fermi statistics.
We apply ext T times to form the statvector space

QT (F) := (ext)T F (29)

for quantum sets of type T over F. Since (ext)0 := Id, F itself is the statvector
space for type 0. R, the statvector space for the empty set only, can be
regarded as the statvector space of type −1.
The types form a nested sequence. The limit T → ∞, which is merely their
union, is the statvector space for the generic quantum set founded on F:

Q(F) := Limo TQ(F) = dT TQ(F) (30)


T →∞

We write simply Q for the quantum sets without foundation, where F = ∅.


We build on Quine’s concept of type here. Quantum type is represented
by a linear operator Type whose spectrum is N (including 0). To say that a
set has type T is to say that the statvectors x for that set obey Type x = T x.

If x has type T then ιx = {x} has type T + 1.


The null set 1 has type 0.
If Type x = T x and Type x0 = T 0 x0 then Type(x0 x) = sup(T 0 , T )(x0 x)
(31)
The statvectors of type T form a subspace TQ ⊂ Q composed of eigenvectors
of the operator Type. Those of type ≤ T form an exterior subalgebra of TQ
written M 0
≤T T
Q := Q ⊂ Q. (32)
T 0 ≤T

The set y whose element is x (and x alone) is usually written as {x}. Peano
wrote it as
y = ιx, (33)
both in his theory of the natural numbers, where he called ιx the successor
of x, and in his set theory. His functional notation is more convenient for
iteration than braces, and we adopt it here, but continue to call the operation
of ι bracing.
Leptons and quarks have significant structure, described by their spins
and charges. Their statvector spaces are tensor products of statvector spaces
associated with each of these properties. Tensor products are how parts are
assembled into wholes in quantum theory; one may therefore call the spin of an

29
electron a part of the electron. This does not imply that one can disassemble
the electron into such parts. We may see a particle without a spin, but we
never see a spin without a particle.
Such a structured but indivisible unit can be represented by a unit set =
singleton = monad whose fine structure is in its second members, the members
of its sole member.
We construct a typed quantum set algebra Q suitable for framing a finite
quantum theory by enriching a classical set algebra C with quantum super-
position so that it becomes an exterior algebra Q. The exterior product cor-
responds to the disjoint union. The grade-g subspace of Q is written Qg . ι
becomes a linear operator
ι : Q 7→ Q1 (34)
Q is a real exterior algebra over itself, and minimal in that respect:

(1) Q = ext ι’ Q;
(2) If P = ext P, (35)
then P ⊃ Q.

The Q grade corresponds to classical cardinality.


Q is inductively constructed by iterating the functor ext ι’ . The bracing ι
is inserted to separate the exterior product on ext ι’ V from any product that
might already be defined on V . This converts a vector space into its exterior
algebra. The iteration can start with the almost trivial exterior algebra R; or
for that matter with the null set.
Several metrics on Q come to hand:
As a spinor space each level of Q has a natural Pauli metric (63).
Each level Q also has a naturalHilbert metric, defined inductively by assum-
ing that ι preserves the metric, and treating each type level as the many-body
statvector space over the previous.
As a Clifford algebra, Q has a natural Clifford metric, defined as preserved
by ι and obeying the Clifford rule (43).
Q is then used to regularize standard quantum field theory and gravity by
infraquantizing them. This heuristic process is called Q quantization.
Much as canonical quantization represents the canonical Lie algebra of a
classical theory isomorphically in the quasi-Lie algebra su(∞) of Hilbert space
H (Appendix 17), Q quantization represents the Lie algebras of the system in
a (finite dimensional) Lie subalgebra of sl(Q), isomorphically where possible,
approximately where not.

30
5.1 Spin-statistics equality
A useful clue to this quantization is the observed spin-statistics equality

W = X; (36)

W being the spin parity, a homotopy that rotates the system continuously
.
through 2π. W = +1 for even spin, −1 for odd (in units of h/2). And X being
the statistics parity, interchanging two statvectors of the system: Xψφ = φψX.
.
X = +1 for even (Bose) statistics, −1 for odd (Fermi).
Since monadics x ∈ Q1 , the first grade of Q, obey x2 = 0, the monads they
describe statistically have fermionic statistics, X ≡ −1. We may infer from
(36) that monadics are to serve as spinors, with W = −1.

Each level ≤TQ is a spinor space. (37)

The quadratic space W[T ] of its orthogonal group SO(W[T ]) will be defined
(in (53)). All bosons must be pseudo-bosons composed of fermions in Q theory,
obeying a Palev statistics [29], which is tautologically simple, with a Lie algebra
such as so(n + 2) instead of the canonical Lie algebra H(n). The Q operators
for orbital variables (§8), isospin, color, hypercharge, generation, gaugeons,
and the higgs are in the Lie algebra sl(EQ) ⊂ sl(Q) (not in Q) and sums of
atomic terms of a cellular level sl(CQ) ⊂ sl(E)
The typed exterior algebra has a long beard [19, 21]. The present work
uses a simplicity principle (3) and a correspondence principle (4) lacking in
those first efforts.
One disturbing feature of the ι operator is that it forgets the statistics of
its operand. Whether x has odd or even statistics, ιx has odd statistics.
Fortunately we encounter a similar anomaly daily in particle physics. The
statistics of a particle forgets all its parts except its spin—for example, its
isospin, color, and flavor.
Again, in the standard constriction of a spinor space as the exterior algebra
over a semivector space, both odd-grade and even-grade exterior products
constitute spinors and are subject to Fermi statistics.
We may represent this feature of nature by a difference in type. The prop-
erties of a particle that do not matter for its statistics are internal in the sense
that they act on the contents of the outermost ι defining a particle statvector.
They operate on a lower type than spin.
In the topological theory of quantum statistics, one expresses an exchange
of a pair in terms of a macroscopic continuous rotation of the pair through π. In
this infraquantum theory, the explanatory direction is reversed. A macroscopic
rotation of a spin is composed of microscopic input-output operations on the

31
semivectorial parts of the spin. For spin 1/2, the the statvectors are the factors
0 0
γ µ in γ µ µ = [γ µ , γ µ ]/2.
To incorporate the spin-statistics equation, we assume:

All quantum histories are assembled from monads described by spinors. (38)

5.2 Numbering the sets


The sequences of N 2-valued objects and the combinations of any number of
N -valued objects are both 2N in number. This suggests that the combinations
in C might form a natural sequence and have natural identification numbers
or indices.
Indeed, each set s ∈ C can be regarded as a positional notation for a
natural number Ns, the index of s, in a base that grows hyper-exponentially
with position:
X
Ns = an 2n , 0 ≤ an < bn , 2n := 2(2n−1 ) , 20 := 1. (39)
n

(Compare binary numbers N = n an 2n .)


P
Multiplying disjoint sets adds their indices, which are thus logarithms. The
successor of the set indexed with N is the set indexed with 2N .
(40) lists some basic polyadics of C and Q with their numbers. Type 4
would overflow the page; its monadics are listed in (41). Type 5 in Planck-size
type would fill the known universe many times over.

32
6 ...
26 ...
5 ...
25 ...
4 ...
16 17 18 19 20 21 22 23 24 25 26 27 ...
3
4 5 6 7 8 9 10 11 12 13 14 15
2
2 3
1
1

0
T 0
(40)

Table. Some polyadics of type T ≤ 6. All occurences of “ ” in one polyadic


represent the same empty set 1, reduplicated to simplify the graphics.

33
L 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
L
1(2 )
(41)
Table. The 16 basic monadics of type 4.

5.3 Reversible infraquantum theory


In ordinary quantum usage, statvectors in some statvector space V represent
input processes or system sources, dual statvectors represent output processes
or system targets, and system transformations are represented by linear oper-
ators in V ⊗ Dual V = Lin V .
The practice of set theorists, however, is different. They have arbitrarily
agreed to represent a map or function y = f (x) of sets as a set itself, one of
ordered pairs (y, x), and also to represent an ordered pair (y, x) as a set, one
of the form {{y, x}x}; or something of that sort. They represent both sets and
their transformations in the same space C of sets.
This conflicts with the standard quantum division of labor between V and
Lin V , which is the distinction between an algebra of operators and its defining
representation space. The conflict is less than it seems. The operators V → V
belong to Lin V ∼ = V ⊗ DualV , which is the statvector space of a single pair,
namely of a system and a dual system. The classical theory requires a set of
pairs, rather than just one, to represent transformations of the system because
it drops the quantum requirement of coherence.
This standard quantum practice also conceals a major continuous symme-
try of fermionic kinematics, one that coherently mixes annihilation and cre-
ation operators, as in Bogoliuboff transformations. If fermion input statvec-
tors form a real vector space V , its Fermi-Dirac anti-commutation relations
are invariant under so(W ), where W = Dup V ⊃ V is defined in (53) and
has the duplex quadratic form (54). This symmetry of the Clifford algebra
Cliff(W ) = Fermi V is not induced by a linear symmetry of V but mixes V
and Dual V .
Thermodynamic irreversibility is built into ordinary quantum theory as this
absolute distinction between input and output statvectors. It is inescapable
for experimenters like ourselves, but robotic nano-experimenters can be more
nearly reversible, and might warrant a reversible infraquantum theory.
One way to set up a reversible infraquantum theory, with no absolute sepa-
ration between input and output statvectors, is to imbed Q, Dual Q, and Lin Q
all within Q. One can then distinguish between input and output statvectors
by the sign of the energy change: positive for input, negative for output. In a

34
frame where iE = ∂t ≺− L46 and i ≺− L56 /N, a natural approximate energy is

{L46 , L45 }
E≈ . (42)
N

5.4 Digression on Minkowskian forms


This section is not used in this work. It refers to an interpretation of monadics
in Q as space-time vectors rather than spinors.

Proposition 1 The Clifford rule

kxk = nxx = Grade0 x t x (43)

and the isometry rule


kι xk = kxk (44)
define a quadratic form on Q, associated with a bilinear form
1
nyx = Grade0 {y t x + x t y}, (45)
2
the Clifford metric form. Each entry e in (40) has norm kek = ±1. For all
N , the quartet {14N , 14N +1 , 14N +2 , 14N +3 } has a Minkowskian signature with
respect to n: three elements of one sign and one of the other. 2Q ∼ = M, the
Minkowski 4-space.
√ n is asymptotically neutral in the sense that the signature
of n on TQ is 2T ∼ o(2T ).

The bilinear extension of bn to Q is also designated by n; it is not multiplicative,


of course.
If norms were assigned values ±1 independently at random, the signature
of
√ the norm would have expectation 0 and its standard deviation would be
2T , in asymptotic agreement with the actual values for n. The Clifford form
n is considered because its invariance group is much larger than the invariance
group SO(TQ1 ) of the Hilbert form on TQ. But is it of any use?

6 Quantization as quantification
It is well known that “second quantization” is not a quantization proper but a
quantification, passing from one-quantum to many-quantum properties. Here
we reformulate canonical quantization too in terms of quantification, and use
it as a pattern for infraquantization. Canonical quantization in a Hilbert space
H will be referred to as H quantization. It has two steps:

35
1. Atomize. Select a canonical Lie algebra V of basic system variables.
2. Quantify. Form a linear algebra H generated by V .
Step 1 amounts to guessing the atom of the quantized system. It passes from
the classical system to a quantum atom. Step 2 clones that atom many times
using Bose statistics.
For a linear harmonic oscillator, for example, V is spanned by three vectors
q, p, i, and is also the statvector space of one phonon. The quantized oscillator
is then a bosonic assembly of such phonons.
The corresponding steps of infraquantization are:

1 Atomize. Select a Lie algebra V of system variables. (46)


2 Quantify Form the algebra generated by V . (47)

One extends canonical quantization to allow graded Lie algebras, such as ex-
terior algebras, as well as Lie algebras proper, in order to construct variables
without classical correspondents, like spins and fermion annihilators, as well
as those that have classical correspondents. We will assume that V and its
algebra are finite-dimensional subalgebras of Q.

The functor ext quantifies, quantizes, and gauges. (48)


It quantifies when it converts a one-quantum statvector space to a many-
quantum one. It quantizes when it converts the quantum atom inferred from
the classical Poisson Bracket to the quantized system. It gauges when it con-
verts one global Lie algebra to many local ones.
A coordinate system (x1 , x2 , x3 , x4 ) orders space-time events into a row
of rows of rows of rows, like the continuum limit of a cubical lattice in four
dimensions. Rows, however, have no continuous symmetries, besides those
already present in their elements, while space-time exhibits such symmetries
experimentally, as through conservation of momentum-energy. By contrast, a
quantum g-simplex over a statvector space V has a simple group, a represen-
tation of SL(V ) by its action on Gradeg ext V .

Every simple Lie algebra can be so represented. (49)


In the classical theory of space-time, events are ultimately small, fields
are macrocosmic in extent, and there is no organization of space-time of an
intermediate size scale. Therefore let us suppose that an infraquantum theory
has an event type Q[E] ⊂ Q, and a field type FQ, with no type between them.
An event statvector is then a monadic of level E [see (41)] in 1-1 correspondence
with the polyadics of E−1Q

36
and
Dim E−1Q is microcosmic but Dim EQ is cosmological. (50)
Therefore Dim E−1Q must be so large that we can leap from E−1Q to the
macrocosmic in a single quantification. This is the cosmological leap, from
microcosm to macrocosm.
If N is the number of possible quantum monads in a system history, the
dimension of the level E monadics must obey

Dim EQ1 = 2E−1 ≥ N (51)

E = 5 would imply N ≤ 24 = (216 ), too small for a quasi-continuum of


events; while E = 6 implies N ≤ 25 = 264K , much more than needed (K :=
1024). Then the preceding level E − 1 has monadic dimension 64K, and its
characteristic length 64K X may well be microcosmic. Level E = 6 has 26 =
264K distinct monads. The Poincaré group can be approximated within present
experimental error, though non-uniformly, by a subgroup of SL(6Q) but not
5
SL(Q). Let us therefore assume that a system history is a 6Q simplex (of
simplices . . . of simplices), or equivalently a monad of 7Q1 .
The standard model groups can all be defined by their actions on about
16 = 24 monads, in 4Q1 . If there are N monads in history, a random simplex
has N/2 vertices on the average. Since N/2  16, operons of current history
are not random but are highly organized locally, into something like a thin
truss dome or graphene, perhaps only one cell thick. The aspect ratio between
the long and short dimensions of this dome is far greater than for a graphene.
The dome also supports the particle spectrum, sharp bands of highly coherent
transmission, and so is presumably crystalline, as Newton already inferred,
though in a quantum sense that must be clarified.
Classical set-theoretic operations do not produce quantum sets; the classi-
cal tree of sets bears no quantum fruit. Canonical quantum theory therefore
grafts a rootless quantum-dynamical stock onto the classical tree of types, just
below the level of the quantum dynamical system. This graft must occur at
an infinite type for the classical infrastructure to have continuous symmetry
groups, and divergences ensue.
To describe simple systems, we replace the classical tree of sets by a quan-
tum tree of simplices Q that is equally autogenous, but quantum to its roots
and simple on every level. Unlike C, the tree Q can bear quantum theories
with continuous symmetries even on its lowest branches, which are all finite-
dimensional.
A simplex described by a monadic is a monad. Its one vertex is a polyad
of lower level. A collection of simplices is usually described in homological

37
algebra by a formal sum of the simplices, but in Q addition + is mere quantum
superposition, not aggregation. A multiplicity of g Q simplices is represented
by a g-adic, a higher-level simplex. A general simplicial complex is then merely
a general simplex or polyad. To emphasize that we deal with a simplex of
simplices, we sometimes call it a meta-simplex .
We may express the two main hypotheses as one:

The physical system history is a metasimplex. (52)

We write the type level of Q within which the history statvector is represented
as EQ. Classical homology works with simplices of one level, that of its vertices.
Infraquantum simplices of levels 1 – 6 are convenient and apparently sufficient
for field theory.
For any v ∈ Q, L v : Q → Q, u 7→ v ∨ u is the linear operator of left exterior
multiplication by v; R v : Q → Q, u 7→ u ∨ v, that of right multiplication by v
(Appendix 17.4). The duplex space W[T ]

W[T ] := DupL TQ1 ⊂ Lin TQ ∼


= Fermi TQ1 ∼
= Cliff W[T ], (53)

has the neutral quadratic form

Q : W → R, Q(w) = w2 . (54)

This defines a Clifford algebra Cliff W[T ]. Because TQ is an exterior algebra


over TQ1 ∼ = T −1Q, its operator algebra Lin TQ, a Fermi algebra, is also the
Clifford algebra Cliff W[T ]. This Clifford structure includes the Dirac Clifford
algebra of the γ µ as level B. We therefore re-designate its first-grade generators
as
γ w ∈ Cliff 1 (W). (55)
(40, 41) tabulate some Q polyadics of level ≤ 5 and all the monadics of
level 4.
Dirac interpreted γ νµ as a Lorentz group generator, a spin component, and
a dynamical variable of the single electron. Cartan interpreted γ µ as a Lorentz
reflection in the µ axis. Both assumed a background Minkowski space-time
that Q lacks. We seek a root for spin that goes deeper than classical space-
time.
Schur [32] and Frobenius use spinors to represent swaps, not spins. Quan-
tum swaps are dyadics of a statvector algebra. Let us take statvectors as our
primitives, exemplified by the Fermi Clifford algebra operators γ q .
0
Then c, c0 index cell vertex monadics 1c , 1c ∈ CQ1 of level C, which are
exchanged by γc0 c . If we confine the vertices to (say) positive annihilators 1c ,
the swaps generate SL(2C ). If we allow negative annihilators 1c as well, that is,

38
creators, and allow both symmetric or skew-symmetric swaps, corresponding
0
to (γc0 c)2 = ±1, then the swaps generate SO(2C , 2C ). Then Lc c represents
the swap γc0 c on Dual EQ1 , of the event level E.
The statvector algebra of level T is TQ = (ext)T R. Any infraquantum
operon is composed of lower level ones, in the way that sets and simplices
are composed of lower level sets and simplices. Q has commuting operators
Grade and Level, corresponding to classical cardinality and level. A g-adic is
an element of Q of grade g, corresponding to a set of cardinality g. The operon
it describes is called a g-ad or g-simplex.

6.1 Cellular locality 2010.04.25

Well before Einstein, Newton doubted his law of gravity on the basis of a lo-
cality principle that has become an indispensable guide for theoretical physics:
All action is by contact.
In the standard model, the theories of the space-time structure of special rel-
ativity and the complex-phase structure of quantum theory, remain global
rather than local, and the theories of the hypercharge, electroweak, and strong
interactions are local gauge theories built on this global structure. Classi-
cal locality requires the Lagrangian to couple only field variables at the same
space-time point; not, say, at the same momentum-energy point in the Fourier-
transform space; and with time derivatives of bounded order. Diffeomorphisms
respect such locality, but the canonical group violates it. The conflicting alge-
bras of general relativity and canonical quantum theory must blend into one
in a simple theory, for they both work too well to be merely suppressed. Let
us suppose that a reformed version of locality will survive infraquantization.
In a topological quantum theory like that developed here, two cells shall be
defined to be in contact not by coordinate relations but by shared elements,
simplicial vertices. This permits one to sharpen the locality principle to cellular
locality:
The action swaps elements between cells in contact. (56)
Then the fundamental operations are not true spins, which assume an external
quadratic classical space, but swaps, which do not. When Cartan represented
rotations by their action on spinors [9], Schur and others had already repre-
sented swaps in that way [32].
Infraquantum theories are naturally local in this sense. One sees this as
follows.
The action of gauge theories is built up from gauge differentiators Dµ (x).
Its locality depends on the action having a low polynomial degree in the D0 s,
since an infinite power series eiaD is a no-local translation.

39
Infraquantization replaces these differentiators by orthogonal-group gener-
ators Lm0 m , represented by dyadics γm0 γm ∈ Q. Their monadic factors γm are
io operators, creating and annihilating vertices of the complex on which they
act. While the D’s are represented in an orthogonal group of huge dimension,
they are formed by cumulation from the vertices of a simplex of low type with
fewer than 16 vertices. Then the vertices m, m0 being swapped belong to an
image of this basic cell.
A product like Dm000 m00 Dm0 m in the dynamical phase statvector would be
non-local, if it coupled vertices m and m00 that are not in contact. But such
0
products actually appear in the context of connecting tensors γ m”m that locate
m0 and m” in the standard cell of CQ. They are then local in the cellular sense.

6.2 The spin tree 2010.04.25

Q is a geyser of spinor spaces. One sees them as follows.


The classic spinor constructions begin with a duplex vector space

Dup V := V ⊕ Dual V =: W , (57)

with its natural, neutral quadratic form and orthogonal group. In this context,
elements of V and W are called semivectors and vectors, respectively. The
spinors of SO(W ) are then multi-semivectors, the multivectors over V . Their
exterior product violates SO(W ) and is ultimately forgotten in the standard
quantum theory.
The spin tree, the central column of (58), is constructed inductively, by
iterating ext, beginning with the trivial exterior algebra R, of grade and level
≡ 0. A spinor space of any level is the exterior algebra over the spinor space of
the previous level. We may readily reconstruct the orthogonal group of which
this is a spinor space. The spinor space of level T is defined to be the vector
space TQ ∼ = T +1Q1 . Its semivector space is the space V [T − 1] := T −1Q1 , which
is an isotropic (= null) subspace of the neutral vector space W[T − 1] of (53),
and SO(W[T − 1]) is its orthogonal group [10, 8, 12].
Orthogonal groups and Clifford-Fermi algebras sit on each rung of the spin

40
tree:
.. .. .. .. ..
. . . . .
↑ext
3 Fermi 4R 16R 32R SO(16, 16)
↑ext
2 Fermi 2R 4R 8R SO(4, 4)
↑ext (58)
1 Fermi R 2R 4R SO(2, 2)
↑ext
0 Fermi 0 R 0 1
T Lin TQ TQ W[T ] SO(W[T ])
Level Algebra Spinors Vectors Group

The symmetry Lie algebra of TQ as a vector space is written sl(TQ). The


symmetry Lie algebra of TQ as an exterior algebra is only sl(Q[T − 1], expo-
nentially smaller. The monad coordinate operators of level T are elements
of
Lin TQ = Fermi T −1Q = Cliff W[T − 1], W[T ] := Dup TQ1 . (59)
In classical set theory the power set of X is written as 2X . Therefore we
write the exterior algebra of a semivector space V as

Ψ = 2V . (60)

The orthogonal group associated with this spinor space is SO(W) where W =
Dup V. If V ⊂ Q then Ψ ⊂ Q.
The distinction between spinors of SL(n) and vectors of SO(m) is rela-
tivized in Q. Namely, one and the same space TQ1 of any level T is a spinor
space relative to the orthogonal group SO(W[T − 1]) = SO(Dup T −1Q1 ) of the
previous level T − 1, and is a semivector space relative to the orthogonal group
SO(W[T ]) of its own level. Therefore the operator ι can be spinor-valued or
vector-valued depending on context.
Because of their spinorial statvectors, let us represent leptons and quarks as
monons. Their statvectors are monadics of the event type EQ1 . Then gaugeons
and gravitons are tetrons, with statvectors in EQ4 .

6.3 Infraquantum Pauli adjoints


Like the Dirac spinors, each level T of the spin tree Q has an invariant bilinear
form, its Pauli form (or Pauli adjoint)
0
p[T ] ≡ p := pq0 q eq eq : TQ → Dual TQ. (61)

41
The case p = p[2] for four-dimensional spinors, usually designated by p =
β, is special in being skew-symmetric; the higher Pauli forms can be made
symmetric. This requires
[T ] 0 [T ] 0
∀ω ∈ so(TQ) : pq00 q0 ω q q + pq0 q00 ω q q = 0. (62)

In an orthonormal basis where the metric on Dup TQ is diagonal with diagonal


elements ±1, the matrix for p[T ] may be chosen to be a product of the matrices
for the γ q whose squares are negative:
Y
p= γ m in one orthonormal frame. (63)
(γ m )2 =−1

p2 is skew-symmetric, and p[T ] is symmetric for T > 2. The Pauli adjoints do


not have a well-defined limit form for the entire space Q. Since the levels and
their groups nest, however, the γ’s and p of one level serve as well for all lower
levels.
p : Q → Dual Q generates an anti-automorphism of Lin Q that effects a
total reversal, interchanging creators eq ∈ Q and annihilators ∂q∗ ∈ Dual Q:

T(peq p−1 ) = ∂q∗ , T(peq∗ p−1 ) = eq∗ . (64)

It defines an invariant form of pseudo-expectation value Av A := ψpAψ for


any operator A ∈ Lin Q that preserves transformation properties under SO[T ] .
For example, if Am is a vector of operators under SO[T ] then Av Am is a vector
of scalars.
The Pauli form p is neutral, however, not positive definite. Square opera-
tors like A = ψ ⊗ pψ would have positive averages in a pre-relativistic theory
but can have negative averages for ψ ∈ Q.
The orthogonal group SO(4, 4) may be a symmetry of a cell or of the dy-
namics statvector, but is not a symmetry of the ambient dome, which reduces
SO(4, 4) locally to the group .

6.4 Cumulation
A monadic of any level embraces parts of the level below. Lower-level operators
induce higher-level ones; in particular, quantum spin operators induce classical
orbital operators as follows.
Q theories represent the statvectors of simplices of simplices by multivectors
of multivectors, polyadics of polyadics. Each polyadic in level Q[T + 1] is a
unique exterior polynomial in insertions of the basic polyadics of the previous
level TQ, as in (40). Any transformation of one level thus extends naturally

42
to monadics of the next level, and thence to polyadics. This defines a unique
exponentiation

Π : SO(TQ1 ) → SO(T +1Q1 ), SO(Q1 ) → SO(Q1 ) (65)

representing each orthogonal Lie algebra within the next. Iteration results in
an nth exponential representation

Πn : SO(TQ1 ) → SO(T +nQ1 ) (66)

acting on multivectors, over multivectors, over . . . , over TQ.


Cumulation Σ is the infinitesimal version of the exponentiation Π:

Π(1 + λx) = 1 + λΣx + O(λ2 ). (67)

It occurs as θ in [26] for example. If x is an operator on Q, and y is any Q


simplex, Σx · y sums x over the elements of y. It is often written in terms of
ac operators ψ, ψ h as ψ h xψ. Here we may write, schematically,

Σx := ιx (68)
∂ιo
This replaces an outermost (“o”) iota ιo by ιx, and sums all such terms. Σ
is a Lie homomorphism of the Lie algebra of each level in that of the next.
(Σn x) represents the operator x ∈ Lin(TQ) in Lin(Q[T + n]). If x is further
subdivided, its parts no longer support the same Lie algebra.
Let us call x the cumulandum of Σx, and Σx the cumulant of x, and write
x = δy to mean that y = Σx, and x = δ n y to mean that y = Σn x. Thus when
δ is defined, it is a left inverse of Σ.

6.5 Negative probability


Usually one uses unitary representations of (2). These are necessarily ∞-
dimensional. For example, the statistical form for the history statvector of a
Dirac spin-1/2 quantum is an integral
Z
kψk = (dxlmn ψ ∗ (x)β(x)γlmn (x)ψ(x) (69)
S

over an infinite spacelike subspace S ⊂ M.


Infraquantization must trim this infinite structure to fit into the finite bed
FQ without contradicting with the many experiments that endorse it. Regular-

ity demands finite-dimensional representations. In conjunction with Lorentz


invariance, this leads to probabilities outside the usual interval [0, 1].

43
These have nothing to do with the so-called negative probabilities some-
times used to describe ordinary quantum interference of probability ampli-
tudes.
Negative probabilities have a useful physical interpretation [13]. They are
no more problematical than a negative bank deposit or a negative energy.
The number is the change in population of the system, positive for input and
negative for output. This permits us to use a single vector space V for both
input and output operations, instead of V and Dual V , and distinguish input
from output statvectors in V by the signs of their norms.
The Dirac form kψk of (69) as written is not definite. Its density is the
product of a current vector jµ (x) of the electron with a normal vector nµ (x) to
the surface S. If both point everywhere into the future, kψk is positive. But
by the fermionic symmetry between input and output, such a statvector has a
partner whose current vector points everywhere into the past and whose norm
is negative. This is not considered a problem. We interpret such statvectors ψ
of negative norm as describing an output rather than an input. Under further
inspection of this example, especially of the electric charges, we see that if the
input is an electron, the partner output is a positon. We adopt the Dirac rule
in general [13]:

kψk > 0 for input, (70)


kψk < 0 for output. (71)

Let ψ+ ∈ V and φ− ∈ Dual V be input and output statvectors, respectively,


with associated counters whose readings are n+ and n− The probability of a
transition ψ1 → φ is the mean value of the increase in n− per unit increase in
n+ :
δn−
P = Av . (72)
δn+
When a positive input count is followed on the average by a negative output
count, the transition probability is negative. This indicates the conversion
of some input quanta to anti-quanta. This is a meaningful description of a
process, feasible or not. As a result probabilities can theoretically be negative
or greater than unity.
This concept of number-sign may conflict with one determined by the arrow
of time, which counts thermodynamically irreversible inputs and outputs as
positive. But thermodynamics is a singular organized limit of a reversible
theory as a particle number approaches ∞. Experiments carried out within a
minute metasystem may not exhibit irreversibility, which is then manifest in
the meta2 system.

44
The probability of a transition ψ → ψ 0 defined by two monadics depends
on the Fermi sea in which the experiment is carried out. An experimenter can
borrow from the sea; the resulting vacancy in the sea counts as −1 event. A
transition that has negative probability is also one that would be impossible
if the sea were empty.
From the one-quantum viewpoint the Fermi sea is part of the metasystem,
the ambient environment, not the system. It has too many variables to be
the system of any experimenter that the universe can accommodate. In the
many-quantum or field-theoretic viewpoint, however, the system can include
at least a droplet of the Dirac sea, described by a fiduciary polyadic. It then
has entropy 0, and so is highly ordered, frozen stiff, like the ether of Newton
and Fresnel, who may have wondered about how we and the planets move so
freely through such a crystal.
In the standard theories the polyadic statvector EDirac of the Dirac sea is
too divergent to be of much use. For example, its grade, the number of events
in the history of the sea, is infinite. It is usually taken into account roughly
by subtracting an infinite constant from the Hamiltonian that is said to make
the vacuum energy expectation zero, although experimentally it is not exactly
zero, and mathematically neither is ∞ − ∞. Since the subtraction requires a
Fourier transform, it is non-local as well as infinite. In infraquantum theories
the grade of the event sea E is no longer infinite but merely too large to be
counted as yet, and the vacuum energy is finite.

7 Dynamical law 2010.04.29

Canonical quantum theory uses a synchronous description, refering all trans-


formations to one time, say t = 0. All its fields share the variables of that
time-section. The theory represents a process P 0 at another time t0 6= 0 by the
operator for the process P at t = 0 that has the effect of P 0 at t0 , assuming
that the system undergoes a given development in the intervening time inter-
val, with no external intervention. The development of the isolated system
then causes its variables to depend explicitly on t while acting on statvectors
defined at t = 0. This is the Heisenberg picture. One is then free to shift the
time-dependence from variables to statvectors by a time-dependent unitary
transformation, to a Schrödinger picture. This formulation is not relativisti-
cally invariant, but it reduces the number of independent variables infinitely,
from all the x(t) to one x(0), making a finite atomic theory possible despite
the infinitude of times.

45
7.1 History statvectors 2010.04.29

Dirac early on imagined a more general quantum theory based on paths or


histories, that retained the degrees of freedom and the wide covariance of
the Lagrangian theory. Its descriptions are em diachronic, while those of the
canonical quantum theory are synchronic. Diachronic variables at different
events vary independently, so diachronic theories describe a broader class of
experimental situations. Dirac deduced the classical principle of stationary
action metaphorically by corresponding classical paths to quantum paths of
stationary phase. He sacrificed some mathematical meaning for this, speaking
of sums over unsummably many classical paths.
Let us recall how the dynamics of a quantum system is summed up in a
history statvector D. Let U (t0 − t) = exp[−iH(t0 − t)] represent the dynamical
development from time t to t0 in a canonical quantum theory with Hamiltonian
H. A Hilbert metric form h is assumed, and may be prefixed, superfixed, or
infixed: hψ = ψ h = ψh. U (t0 − t) is unitary with respect to h. Then a
discretized path from an initial time T1 to a final time T2 is represented by the
dynamical statvector
N
O T2 − T1
D := U (δt) ⊗ . . . ⊗ U (δt) = U (δt), δt := . (73)
N
1

The Feynman path integral theory would describe this dynamics by the diag-
onalized product

D = U (qn , qn−1 )U (qn−1 , qn−2 ) . . . U (q2 , q1 ), (74)

a function of only n coordinates q. D is not a unitary invariant. It is designed


to give only a space-time view of the development. The matrix element of D
is a function of 2n coordinates q, however. This permits a history statvector
D to be a unitary invariant and give a view of the development in any frame.
An experiment statvector of the dual form

E = ψ(T2 )h ⊗ E(n − 1) ⊗ . . . ⊗ E(1) ⊗ ψ(T1 ) (75)

connects the atoms of D and provides input and output at the ends of the
process. The two together define a path amplitude

D ◦ E = ψ(T2 )h U (δt)E(T2 − δt) ⊗ . . . ⊗ E(T1 + δT )U (δt)ψ(T1 ). (76)

In this example, a well-defined experimental operation E(t) is carried out at


each time tn = nδt, but in principle superpositions of such products are also
allowed, entangling them.

46
The limit as N → ∞ may not exist. Infraquantum theory is finite-
dimensional, and so has no such divergence problem, but it may still give
absurdly large theoretical values for quantities that are experimentally small.
The relativistic covariance of the canonical theory cannot be taken for granted,
since the time slice is not a covariant construct. But we can assure that the
infraquantum theory is covariant under a given simple group by building it
with covariant processes from covariant atoms. To guarantee Lorentz invari-
ance exactly, we choose the group of the infraquantum theory to include the
Lorentz group. To approximate Poincar’e invariance, we choose the group of
the infraquantum theory to have the Poincaré group as a singular organized
limit, as the Yang group SO(3, 3) does.
Let us suppose for now that an infraquantum system history is maximally
described by a history statvector E in a finite-dimensional subspace V ⊂ Q.
This means that V , and hence Q, contains correspondents of both input and
output statvectors of the canonical theory. A classical history is a singular
organized limit.
The correspondence between inputs and outputs that make up an assured
transition is now represented by a linear operator h : V → V , interchanging
creators and annihilators, and reversing order of products. When φ = hψ, this
means that the transition ψ → φ goes on every trial.
Canonical quantum dynamics permits us to measure the energy by measur-
ing a Hamiltonian operator H constructed from canonical variables; perhaps
by weighing the system. But the quantum energy can also be measured by
measuring a frequency, as symbolized by the definition E := i~∂t . This does
not work in the classical limit ~ → 0, where for finite energy the frequency
→ ∞. Nevertheless this conception of energy, based on frequency, seems more
fundamental than the classical notion, based on work; but it is the frequency
of a statvector, a statistical construct. The dynamical equation equates energy
to a Hamiltonian, which can be measured on one system, but is not always
the energy. Weighing the system seems to be a way to determine its energy
non-statistically.

7.2 Infraquantum scattering 2010.04.21

To compute scattering amplitudes let us find history statvectors in Q repre-


senting the input and output phases of the scattering experiment.
Commonly a scattering statvector for a fermion of momentum km has the
singular form
ψ(k) = e−ik·x ψ0 χ (77)
where the statvector χ provides values for the quantum numbers of simple

47
groups like isospin and Lorentz spin, and ψ0 is a singular “seed” statvector of
momentum eigenvalue pm = 0. ψ(k) is made from ψ0 by a translation in mo-
mentum space generated by ix. Metaphorically, ψ0 is the principal statvector
of a position basis.
Infraquantization need not modify χ but simply takes it to be a statvector
in CQ .
Infraquantization plausibly replaces the exponential by a normalized or-
thogonal transformation:
mL
ψ)(k) = e−ik·x ≺− N −1 eX k m5
= ψ(k). (78)

Although the norm of the usual exponential is infinite, that of its reform is
finite, and designated by N 2 .
To deal with the “seed” statvector ψ0 , let us first replace it by a projector
on momentum p = 0, designated by δ(p). The probability operator for the
incoming quantum is then

ρ(p) = e−ik4x · δ(p) (79)

Specifying both the time and the energy of one event is impossible due
to complementarity. One evades this problem in the canonical theory by al-
lowing an infinite time to prepare the system, requiring only that the task be
completed by a specified time t. The time t may then be continuously varied
to define a sharp frequency. In this way time and the energy can both be
known with arbitrarily small indeterminacy product ∆t∆E. Canonical quan-
tization corresponds a classical infinitesimal transformation to a quantum one
by a Lie-algebra homomorphism, and then uses i to correspond observables to
observables.
In an infraquantum theory, we do not have forever to prepare an energy, we
cannot vary the time continuously, and we do not have an absolute i. Never-
theless a correspondence survives between such skew-hermitian operators, not
hermitian operators. A skew-symmetric correspondent ∂bt = E is assumed to
time-translation ∂t , and a skew-symmetric correspondent t to energy transla-
tion it.
Since the correspondent bı of i is not central, we cannot form correspondents
of t and E from those of t and E by simple multiplication. The factor-ordering
problem is even greater for infraquantization than for canonical quantization.
If we have made a time operator b t, we may define an instant by the pro-
jection operator Pt0 on a specific eigenvalue t0 of the operator b t. Since none
of the spatial coordinates x k
b commute with t = x b 4
b , Pt0 has little resemblance
to the classical or canonical-quantum instants.

48
It is axiomatic in classical physics that all instants include the same infinite
number of events. This is incompatible with regularity, which permits only a
finite number of independent events in each time slice. Infraquantum theories
typically describe histories with time-slices that grow and then shrink with
time. This impairs the correspondence between infraquantum dynamics and
the Hamiltonian dynamics of canonical theories. We assume that Hamiltonian
dynamics applies to middle age, t/NX → 0 in the singular limit NX → ∞, X →
0. To extend dynamics to periods comparable to the total time NX, we must
allow for information increase and decrease, for example by replacing some
statvectors by probability operators.
Infraquantization admits a more pragmatic construct of dynamical law than
classical physics. Classical physics traditionally took the “perspective of eter-
nity”, excluding Observer and Law from the physical universe, and imagining
them to be fixed apriori instead. The Law was given much symmetry, the
Observer none. Special relativity and quantum theory pluralized, naturalized,
and activated the Observer, by accepting observers into the physical universe,
relating us by a relativity group, and predicting that our determinations of
a system variable changes some unobserved variable by a finite unpredictable
amount.
In a typical quantum experiment, the experimenter inputs a beam of sys-
tems undergoing the experimental process. Therefore we may represent the
metasystem for this purpose by a virtual population or reservoir of replicas of
the system, and represent an input operation by a selection from this reser-
voir. Then the ideal quantum experimenter, while no longer infinite, is still
exponentially larger than the system, which is therefore a logarithmically small
part of the universe. A quantum theory does not require a universal ontol-
ogy and a deterministic law; a manual of feasible experiments to be done in
a rather coarsely specified ambience, and a statistical prediction of their out-
comes will suffice. The dynamics of the system includes the influence of the
mostly unobserved metasystem on the system undergoing experiment. This
influence is sometimes expressed by external fields or sources. The ambience
may have a finite temperature, but here we approximate it as cold, coherent
in the quantum sense.

7.3 Infraquantum propagators 2010.04.21

Dynamical theories in classical space-time express vacuum autocorrelation am-


plitudes among any number of generating variables x1 , . . . , xn by

h xn . . . x1 i0 = Tr xn . . . x1 D = E ◦ D, E := xn . . . x1 . (80)

49
These become propagators when the variables are monadic statvectors, com-
monly associated with specified space-time points. Let us retain this form in
the infraquantum theory and merely reform D. E is now a statvector speci-
fying an experiment and E ◦ D is the amplitude for E under the dynamical
statvector D.
To reform the usual dynamics, let us absorb the usual imaginary factor
i/h into the action A, which then becomes a dimensionless skew-symmetric
real skew-action operator Φ − iA/h on history statvectors, representing an
imaginary phase:
R
D = eΦ −
 eiA/h = ei (dx)L/h
, Φ ∈ Lin Q[E]. (81)
The history amplitude D is used to compute propagators as in (80). They in
0
turn define the time-translation operator W t t connecting an input statvector
for one time-slice t with that for a later time-slice t0 . Choose independent
variables st making up a complete set with time t, and basic eigenstatvectors est
with the indicated eigenvalues. Then the transition amplitude can be written
as h 00 i
0 0
W s t st := Tr es t est D . (82)
In the continuum theory there is supposed to be a limiting case t0 = t + dt, of
the form
0
ws (t+dt) st = [δ 0 (t) + H(s
e 0 , s, t)]dt. (83)
This defines the skew-Hamiltonian operator H(t). e
Today the term “vacuum” has been thoroughly relativized [41]. It stands
for the part of the universe that is relegated to the metasystem, which is
observed only coarsely. To escape the preconception of a uniquely defined,
absolutely empty, physical vacuum, let us speak of an ambient medium or
ambience rather than “the vacuum”. The ambience enters the theory in the
determination of the physical input/output statvectors and in the dynamical
law of the system.
Sometimes the ambience is supposed to have some great symmetry, and
sometimes a temperature. Such assumptions must be used with some dis-
cretion, since the metasystem includes the experimenter, who breaks all ob-
servable symmetries and is not in thermal equilibrium. The cosmos exhibits
little symmetry, and the symmetry near the Big Bang is less than that of our
ambience today. A vacuum symmetry is an approximation to the universe in
the way that a tangent is an approximation to a curve. It extrapolates local
conditions to the rest of the universe.
To study the quantum structure of space-time we we must include some
of it in the system. We also must have a still greater place to stand, how-
ever; to satisfy the postulates of quantum kinematics, the metasystem must

50
be exponentially larger than the system. So we must leave most of space-time
structure in the metasystem.
If a quantum description of our metasystem would require a statvector
space of (say) 10100 ∼ 2300 dimensions, the largest system we can fully observe
has only about 300 dimensions in its statvector space. Since 3001/4 ∼ 4, this
would restrict the system to a microcosmic space-time cell only 4X on each
edge, imbedded in the ambient space-time. Like an ice-cube in a glacier, its
quantum structure will then be mainly determined by its ambience.
Let us assume that the standard model spins and charges originate as spin-
like operators on a monadic statvector space CQ1 of some “cell type” C within
the typed statvector algebra Q of §5. We require that C = 4 accommodates
not only the Yang Lie algebra so(6 − n, n), which reforms the Poincaré and
canonical Lie algebras, but also the unitary charges of the standard model,
which are already semisimple:

C = 4, Dim CQ1 = 16; E = 6, Dim EQ1 = 25 . (84)

7.4 Dynamics as ambience 2010.04.21

Recognizing that the quantum system is minute compared to its ambience


permits another reading of dynamical relations:

Dynamics is an autocorrelation within the system induced by the ambience.


(85)
The dynamical law for any small part S of the universe is assumed to be a
history statvector D for S that serves as statistical surrogate for the rest of
the universe and assigns amplitudes to experimental statvectors E describing
a coherent experiment. More generally, a probability operator for the history
could appear instead of either statvector D or E, and gives less information
than a statvector.
In a classical framework such a theory of dynamics begins an infinite regres-
sion. It seems that to describe how the ambience acts on the system requires
an extension of the dynamics of the system to the ambience. The principle
(85) seems to call for a dynamics rather than define one.
In a quantum theory, however, the system is described by statvectors that
represent operations of the metasystem, and therefore of the ambience, on the
system. Such statvectors also describe the metasystem, by its operation on
the system. The theory predicts the probability amplitude for a certain action
of the metaystem on the system to be the contraction of the experimental
statvector E with the dynamical statvector D. No additional dynamical theory
is required.

51
7.5 Organizational entropy 2010.04.23

The existing concept of organization is based on the concept of Hamiltonian


operator. It concerns ground statvectors, eigenstatvectors of minimum Hamil-
tonian eigenvalue. The Hamiltonian generates the development from one time
slice to another determined by the history amplitude. In a canonical theory
this is assumed to be a unitary transformation. This builds in a singular the-
ory of time, and may badly break Poincaré invariance. The usual condensation
theory is also based on symmetry rearrangement, the contraction of symmetry
groups to Nambu-Goldstone groups, which are again usually singular. Both of
these theoretical strategies incur infinities.
In the present project, it would seem absurd to contract a regular sym-
metry to a singular one so that we then have to of de-contract it. Here we
[hopefully] base a theory of organization on a regular diachronic dynamical
statvector instead of a singular Hamiltonian, and on rearranged rather than
broken symmetry.
The concept of organization useful for superfluids is off-diagonal long-range
order in a ground statvector of the Hamiltonian. This presents conceptual
problems for an infraquantum theory. The Hamiltonian construct requires
a time variable that the real infraquantum theory lacks, except in just the
singular organized limit we are trying to formulate.
We can measure organization within the framework of history statvectors,
first canonical and then infraquantum, without singling out a time variable.
For a given dynamical statvector D differences in the system are repre-
sented by differences in the experimental statvector E. b Systems are repre-
sented by experimental statvectors confined to one instant of time. The same
system may be represented by a statvector at a different instant of time. Then
the two E statvectors are connected by a time-development operator defined
by D.
Let us study how to analyze an infraquantum system into time slices related
by a time-development. If the history statvector space is EQ, with projection
operator P [E], the first problem is to analyze P [E] into a product of commut-
ing time-projections
Y
P [E] = Pt [E], [Pt [E], Pt0 [E]] = 0 (86)

in a way compatible with the dynamical statvector D. Let us assume that this
defines a factorization of EQ into factor spaces Qt [E]:
_
E
Q= Qt [E] (87)
t

52
Each space has its own operator algebra and trace operation Trt . Given pro-
jectors P and Pt for the history and one instant t, the marginal distribution
ρnot t for the rest of history is defined by

ρnot t = Trt P Pt (88)

One defining property of an instant t is that when Pt is one-dimensional, so is


ρnot t .
Typically the entropy of a one-quantum marginal probability operator
grows without bound with n, the number of quanta, and D, the dimension-
ality of the one-boson statvector space. One well-known invariant measure of
condensation is a marginal entropy that remains finite and bounded in this
limit.
The prototype is a Bose condensation into a single statvector one-boson
statvector φ. Some condensates are approximately represented by a tensor
power of the form
n factors n
z }| { _
Φ = φ ∨ ... ∨ φ = φ. (89)
1

Φ describes a union of n bosons all associated with the same statvector φ. The
associated one-boson marginal probability operator is

ρ := Trn−1 ΦhΦ = φhφ. (90)

Its entropy is
S = − Tr ρ ln ρ (91)
in units of k. This is 0, while the marginal entropy of a random aggregate is
ln D, infinitely greater in the limit of large D.
The entropy of the marginal one-quantum probability operator is defined
for the diachronic quantum theory as simply as for the synchronic, thanks
to the finiteness of the theory. In an infraquantum theory the assembly is
carried out by ext, the quantum correspondent of power-set formation or set-
exponentiation. The trace is over all histories, that is, over all simplicial com-
plexes, which are finite in number.
In some cases of physical interest, the condensation is not of individual
quanta but of Cooper pairs, or tetrads, or of polyads of some even grade g
that is held fixed as n, D → ∞. Then the condensation shows up as a polyad
entropy that is bounded as n, D → ∞.

53
7.6 Infraquantum ground statvector
The ground statvector ψ is usually defined by Hψ = Emin ψ where Emin is the
minimum eigenvalue of the Hamiltonian H. While there is no Hamiltonian
in an infraquantum theory, there is the propagator W c between two adjacent
time-slices. This is given by (82) with the infraquantized dynamical statvec-
tor Db for D. (82) includes a kinematic transformation t → t0 besides the
induced dynamical transformation s → s0 . There seems to be no reason in
an infraquantum theory, which has a first time and a last, for W c to define a
one-parameter group of time translations. The existence of such a group is
likely a consequence of organization and limited to a middle range of temporal
resolution, much longer than X and much shorter than the age of the universe.
The infinitesimal generator of this limiting group is the Hamiltonian operator
of the time slice.
The Hamiltonian mode of description singles out a rest frame and a time
variable. This is harmless when the condensate also defines a rest frame, as
does a superconducting crystal. But our goal is to describe the ambient “vac-
uum” condensate, which is Lorentz covariant. A description that is transpar-
ently Lorentz covariant will sometimes be more useful than one which masks
this covariance.
We must describe how the diachronic dynamics D enters into a condensate
statvector E. We review the canonical theory to guide the infraquantum one.
In the canonical theory, an input process ψ is carried out over a long interval
in the remote past, −T ≥ t > −∞, and its totally time-reversed output process
ψ T in the remote future, ∞ > t ≥ T . If ψ is an energy eigenstatvector, a shift
δt in the time between ψ and ψ T changes the history amplitude by a phase-
shifting factor e±iωδt , and ~ω is the energy eigenvalue of the system.
The infraquantum theory has an operator corresponding to the time shift-
ing operator ∂t , namely the Yang skew-energy L64 =: ∂bt . And one correspond-
ing to the i of the phase-shifting factor, namely bı = L65 /N. However bı is not
central until an appropriate condensation occurs and a singular organized limit
is taken.
It seems that the usual energy construct requires a limit process T → ∞
for its definition. If we copy this in the infraquantum theory, we invalidate the
guarantee of finiteness; the limit may diverge. Instead we choose an appropri-
ate finite time interval for the experiment such as
√ √
NX > t > − NX. (92)

This is short enough to be far from the beginning and end of time, and so
approximate translational invariance. It is long enough to include a great
many X’s and so approximate continuity.

54
7.7 Infraquantum organization
7.8 Unbreaking symmetry
The problem is to describe Nambu-Goldstone symmetry transformations on a
condensate without breaking the simple symmetry of the pre-condensate. Let
us tackle the example of an organization of many spins 1/2.
Symbols: The individual spin vector is s. The statvectors for an individual
spin are spinors ψ, ψ h . The collective spin is S = ψ h s ψ. For a set of N spins
with statvector Ψ, the mean value of the collective spin is S := Ψh S Ψ. This
is normalized to unit length to define the mean spin direction
Ψh S Ψ
e := . (93)
|Ψh S Ψ|
Therefore the collective spin about the mean collective spin is
S · S := S · Ψh S Ψ. (94)
It is plausible to identify s with the generator of the unbroken symmetry.
This leaves two Nambu-Goldstone symmetries, usually approximated by two
translations of the NG boson, a singular limit of the regular symmetry we
seek. More accurately they are two rotations about axes normal to S. They
can be approximated well by translations because they are usually restricted
to small rotations of a long vector. Rotating one spin s through a full 2π turns
S only through 2π/N. It is only when we unthinkingly use the translational
approximation for large rotations that infinities creep in.
The NG rotations are generated by two components of S orthogonal to S;
let us designate them by S1 and S2 .
The three generators S1 , S2 , S have well-defined physical meaning for the
ensemble of ensembles of spins statistically described by the statvector Ψ. They
do not have the usual form of observables. They are certainly not observables
of one spin, since they involve many. And they are not observables of the
collective, since they involve the statvectors Ψ, Ψh of the collective. They may
be interpreted as joint observables of one spin and one collective of spins.

8 Gauging as quantification
The Yang simplification adjoins dimensions to the usual (x, y, z, t). A Yang-
space point y, relative to any frame, carries a space-time point x, a cotangent
vector p, an element a of an so(n), and an infinitesimal complex-plane rotation
 
0 −φ
∈ so(2), (95)
φ 0

55
in the array
0 −φ −x1 . . . −xn
 

 φ 0 −p1 . . . −pn 

 x 1 p1 0 . . . an1 
 
y= .. .. . (96)

 . . a12 ... a n2 

 .. .. .. .. 
 . . . . 
xn pn a1n ... 0
Thus y includes the elements of a gauge connection, which associates a Lie
algebra element a with a direction p (or its dual) at a point x. This suggests
that a singular quantum gauge field might actually be a limo of a regular sea
of events, taken from a quantum space like a Yang space with an appropriate
orthogonal group.
The spin group represented within the Clifford algebra Cliff W[4] of level
4 is the double covering Cov SO(4, 4). This contains no SO(5, 1) subgroup.
Instead let us adopt SO(3, 3) ⊂ SO(4, 4) for the Yang group for now. SO(4, 2)
would also fit into SO(4, 4). The spinors of Yang SO(3, 3) with 8 real dimen-
sions fit into the 16 dimensions of 4Q1 , but not into the monadics of a lower
level.
Q grade counts vertices of the cell, Q level counts nested iotas, and the
basic Q dynamical operators LC ∈ so(W[C]) of the cell level C count com-
ponents of generalized angular momentum in units of the roots of this Lie
algebra. The eight spin-like atoms of orbital angular momentum x bm and pbm
(m ∈ 6), and the quantized imaginary bı, are among the 15 generators γ C of a
Yang so(3, 3) Lie algebra, forming a Lie subalgebra of the 120-dimensional Lie
algebra so(W[C] ⊂ Lin 4Q1 ∼ = Cliff W[C] of level C
The unitary charges of the standard model can be represented using another
6 real dimensions and the quantized imaginary bı already constructed. This still
fits into the 16-dimensional spinor space CQ1 with C = 4.
The matrices of so(W[C]) are too small to usefully approximate our usual
orbital variables. They best represent spins, the “atoms” of the orbital vari-
ables. The orbital variables are cumulants of such spins. A Yang so(3, 3) of
level C is faithfully represented on level E by second cumulants of its gener-
ators LC (§6.4). The LC in one Q frame are, up to constant multipliers, the
operators

xm
δb = X γ m5 , m, n = 1, 2, 3, 4,
δb
pm = E γm6 ,
δ bı = N−1 γ 65 , (97)
δLb nm = hγnm ,
as h(4) ≺− so(3, 3) ≺− sl(6)

56
The infraquantized imaginary bı is normalized to unit magnitude with a factor
N−1 . To form macrocosmic monad coordinates, we must cumulate these atomic
16
cell variables at least twice, to reach at least level 6, with 26 = 2(2 ) points,
ample for a quasi-continuum.
The statvectors for the atoms of momentum do not commute. Neither do
their cumulants, the quantified momenta, which correspond to infinitesimal
translations. This quantum non-commutativity survives into general relativity
as part of the curvature, perhaps including a cosmologically constant part.
Let us posit spontaneous polarizations of bı = Σ2 γ 65 − i and of a Pauli
form βc0 c = −βcc0 that makes γ 4321 β-symmetric:

β : γ 4321 7→ βγ 4321 β −1 = γ 4321 T . (98)

The form β is a special case of the Pauli metric β T of level T , for which see
(63). β singles out the first-grade γ c : they are the elements of Lin 4Q1 that
anticommute with γ 4321 and are β-skewsymmetric. These γ c in turn define the
Minkowski space-time, as a singular organized limit of cumulants of second-
0
grade products γ c c . Thus the Pauli form determines the Minkowski form and
space-time in this context.
While de Sitter relativity can be a useful approximation for patches of
physical space-time the Yang SO(3, 3) relativity is badly broken by the ambient
dome, which is anisotropic, but it can still apply to a quantum cell of the dome.

8.1 Gauge in infraquantum space-time


it has been proposed that the Higgs field, like the BCS pair statvector, is an
order parameter of a bosonic condensate. This is a further development of
the metaphor of the superconducting vacuum. Usually it is supposed that
the condensate fermions move in classical space-time and interact through
gauge forces; for example, in the technicolor theories of Weinberg and Susskind
[45, 38]. In an infraquantum theory the need for a vacuum condensate is even
greater, to account for the vacuum means
• g, the gravitational field
• φ, the Higgs field
• i ≺− bı, the quantum imaginary
• xµ , the quasi-continuum event coordinates,
as order parameters. Here we assume that the ambient dome forms the su-
perconducting vacuum quasi-continuum and the gaugeon fields in a singular
organized limit.

57
We must then transfer the main ideas of Bardeen-Cooper-Schrieffer super-
conductivity theory from classical Minkowski space-time to an infraquantized
space-time like Yang space. Many of the necessary ingredients for a gauge the-
ory of the Yang-Mills kind are already present in the Yang space (96. A gauge
theory associates a Lie algebra element a(x, v) with each infinitesimal tangent
vector v at each point x of space-time, depending linearly on v. δu = a(x, v)u
represents the change in a vector u due to transport of u from x to x + v.
Points of Yang space, however, have even statistics, wrong for a Fermi sea.
Therefore we take for the statvector space of a fermionic sub-event the spinor
space R8 underlying the Yang so(3, 3) space, with odd statistics. The sub-
event variables fulfilling the Yang commutation relations are then the dyadic
0
spin components γ c c of this event. To form an event of Yang space, many
such spinorial sub-events have to pair off and condense, like Cooper pairs.
This raises the question of how the fermion interactions are to be described,
before and after the organization of the ambient dome. Since the coupling of
events is to have as its singular organized limit the dynamics of the standard
model and gravity, our first approach is to infraquantize that dynamics.
It is parsimonious to surmise that all arise in a singular organized limit
from one bosonic condensation of the plexus, with one critical temperature
Tc , and can all be expressed in terms of one order parameter; rather than a
sequence of condensations with distinct critical temperatures Tc n , one for each
of these vacuum mean values.
In support of the single-condensate notion, we first note that

Proposition 2 In a Yang space (11) the singular organized limit of bı uniquely


determines the Minkowski metric.

Proof The Clifford algebra associated with Yang so(6) has six monadic gener-
ators γ y (y = 1, . . . , 6). The Clifford element defining the quantized imaginary
bı is γ 65 in the Yang frame. To be specific let us assume that

(γ 4 )2 = (γ 5 )2 = (γ 6 )2 = −1, (99)

so that the Yang group is SO(3, 3). Then the Clifford complement of γ 65 is
γ > γ 65 where γ > := −γ 654321 . Clearly

γ > γ 65 = γ 4321 (100)

is the top element of the Minkowski Clifford algebra (not to be designated here
by γ 5 for obvious reasons). Thus bı determines γ 4321 .
But γ 4321 in turn determines the subspace of Minkowskian generators γ µ
(µ = 1, 2, 3, 4). They are the monadic generators among the γ y that anticom-
mute with γ 4321 .

58
As usual, the γ µ in turn determine the Minkowski metric by their anticom-
mutators:
0 0
{γ µ , γ µ } = 2g µ µ .  (101)
To develop this proposal into a microscopic theory, we must propose a trial
statvector analogous to the Cooper pair of superconductivity. The classical
Dirac vector field γ µ (x) is not one operator but four. We may without loss of
generality regard it as a surrogate for a more accurate, though more singular,
two-point dyadic γ(x0 , x) ≈ γ µ (x)δxµ , where δx := x0 − x. This in turn is
supposed to be the ambient mean of a dyadic

b ∈ Cliff 2 W[E]
γ (102)

b is second-grade in Cliff W,
of higher type, containing events as well as spins. γ
and therefore fourth-grade in Dup Q, and may be diagramed as an X. The
relevance of two-point fields was pointed out by Einstein and Mayer [14, 15].
0
The angular momentum operator γ c c is then a concomitant of γ b. It pro-
vides the Q representation (97) of the Yang SO(3, 3) group that reforms the
canonical Lie algebra h(4) ≺− sl(R). It thus provides each frame with an in-
fraquantized imaginary bı, a local variant of the global constant i of complex
quantum theory, in the way that general relativity localized the global constant
gnm of special relativity to the local form gnm (x).

9 Infraquantum space-times
9.1 The minimum black hole
Infraquantization modifies the usual estimate of the minimum black hole size.
To review the standard heuristic argument, consider a body of mass M local-
ized within a ball of radius R for a time much greater than R/c. Let us assume
for the moment that Newton’s law of gravity is still approximately valid for
lengths as small as R. If the body is not to disappear into a black hole, it should
be able to reflect light. The kinetic energy mc2 of a photon at its boundary R
must exceed the photon’s gravitational binding energy −V = GmM/R:
> GM m
mc2 ∼ −V = , (103)
R
where G is the (unrationalized) gravitational constant.
If the Heisenberg determinacy limit holds for lengths as small as r, the
body has a root-mean-square momentum
> h
Mc ∼ . (104)
R

59
(103) and (104) imply the Planck-length bound,
r
> 2hG
r∼ =: RP . (105)
c3
Newton’s law of gravity presumably does not hold on the scale of X. Its
error could be enormous if, as infraquantization suggests, there is a short-range
fermionic hard core of repulsion within the long-range attraction of gravity.
The following crude argument uses the Newton law anyway, so it does not
give an estimate but merely invalidates one of the usual arguments that the
observed elementary particles cannot be quantum black holes, and so reopens
the question.
The Yang indeterminacy relation implies not (104) but

> h
r∼ |hbı i|. (106)
2mc
Then, still assuming Newton’s law of gravity,
r
> 2hG
r∼ |hbı i| = |hbı i|RP . (107)
c3
Since bı is the sum of an even number of spins, its spectrum includes 0. If
hbı i → 0, the black hole radius and mass can approach 0. Quantum black holes
much lighter than the Planck mass become possible below the time scale of
the ambient dome organization.
Suppose now that there is a potential energy minimum V instead of the
unbounded Newtonian potential −GmM/R, but equivalence is still approxi-
mately valid. Then V is proportional to both masses:
GmM
V ≈− (108)
R0
with a new physical constant R0 . Then instead of (103) one has

> GM m < R0 c2
mc2 ∼ −V = , M∼ . (109)
R0 G

9.2 Infraquantum spin-orbit analysis


Let us agree provisionally that fermionic statvectors have the data structure

Ψ := spin time charge field (110)

60
For example, the monadic statvector

∈ 6Q1 (111)

i part [ ] ∈ Q[3], the unitary charge part [ ] ∈ Q[4], and


has the Lorentz hspin
the orbital part ∈ Q[5].
This decomposition of Q[6] raises a significant question of covariance. The
standard decomposition into spin, unitary and orbital variables is invariant
under the standard model symmetry group SO(2 × 3) × ISO(3, 1) and under its
gauged form. The corresponding Q decomposition, however, is not invariant
under SO(Q[4]) or SO(6Q1 ), natural candidates for Q gauge groups. It is
understood that the ambient condensate breaks the symmetry and defines the
decomposition in question.
Since the orbital variables are sums of many spin variables, they commute
with one of the spin variables only approximately. In extreme conditions this
may have experimental consequences. The total angular momentum too is
the sum of orbital and spin parts, with a scale factor h, and this results in a
conspicuous spin splitting of angular momentum levels. In a simple theory the
total momentum coordinate Lm6 too is the sum of orbital and spin parts and
should exhibit a spin splitting. This spin splitting will be harder to measure,
however, because under ordinary conditions its relative scale is much smaller
and the spectral lines of momentum are not as sharp as those of angular
momentum.

9.3 Infraquantum chirality


The condensation that produces the central i also reduces general spinors of
the Yang group to chiral spinors of the Lorentz group. We may see this as
follows.
The chirality of a fermion in the standard quantum theory is an operator
at the top of its 16-dimensional Dirac Clifford algebra,
. .
iγ 4321 := iγ 4 γ 3 γ 2 γ 1 = ±1 =: iγ > = ±1. (112)
.
Let us adopt the convention iγ 4321 = 1 for a left-handed electron, the kind
with isospin 1/2, and −1 for a right-handed electron, the kind with isospin 0.
(Or is it the other way around? No matter.)
The Yang SO(3, 3) group has a spinor space ∼ = 8R. Its pseudoscalar volume
> 654321 65 4321 0
element γ := γ =γ γ commutes with so(3, 3) transformations γ y y
.
(y, y 0 = 1, . . . , 6) and has eigenvalues γ > = ±1. It reduces the spinor space to
two eigenspaces ∼ = 4R with γ 4321 = ∓γ 65 = ∓γbı. These are therefore chiral

61
spinors. That is, when Yang SO(3, 3) is reduced to the Lorenz group, the Yang
spinors are reduced to chiral Lorentz spinors.
xm ∼
The atoms of the quantized orbital operators may be represented by δb
m
γ , δbpm = γ 4321 γm , as was suggested also by Marks [28].
Present experience, where
p the canonical relations work, is with a part of
the spectrum of |bı| = + −Qi2 so near to the maximum value N as to be
indistinguishable from it. Yet this narrow band must have a multiplicity that
passes today for infinite. For example the band

1 − N−1/2 < |bı| ≤ 1 (113)



is narrow and crowded, with width N−1/2 → 0 and multiplicity O( N ) → ∞.
In the singular organized limit bı → i, E → 0, and (supposedly) (bı)2 →
−1, the classical time polyad and the canonical commutation relations are to
emerge. We must suppose that
h
XEN = ~, N  1, h/X  1 TeV, ≈ 0, (114)
NX
in the sense that h/NX is presently not resolvable from 0.

9.4 The infraquantum dome


In the standard model physical lepton and quark annihilators corresponding
to Q monads carry
• the classical four space-time coordinates, which have a large spectrum
and vary greatly over the dome,
• four momentum-energy coordinates, which have a large spectrum but are
very small over the dome, and
• several spin and charge coordinates, which have a small discrete spectrum
and no detectible space-time extent.
• a generation index Γ = 1, 2, 3.
Kaluza-Klein theory created a compactification problem: What energies curve
the gauge dimensions into small loops? Q theories have no such problem,
having no such loops. Instead they have a difficult organizational problem:
the self-organization of its spin-struts into the bubble dome on which we live.
Canonical quantum fields and then classical fields are to arise as successive
singular organized limits of the excitations of this dome. The particle spectrum
is to inform us about the fine structure of the dome as the Brillouin zones of
a crystal inform us about the fine structure of the crystal.

62
Infraquantization quantizes the orbital variables left classical in the stan-
dard model and SCCM , using similar atomic quantum elements for orbital, spin,
and charge variables. Flipping the spin of an electron annihilator is supposed
to change the proper time between it and another electron annihilator, possibly
by too little to resolve with present experimental space-time resolution.
Let us assume that the dome has a simple Lie algebra

adome ⊂ so(3, 3)Yang ⊂ EQ2 (115)

to be found.
The Poincaré relativity group does not fit into the groups of EQ. It is a
singular approximation to the Yang relativity group SO(3, 3) at the cellular
level.
DO: Show this.
Since the Poincaré Lie algebra is a reduction of diff, the Yang Lie algebra in
turn is assumed to be a reduction of a higher-level relativity gauge Lie algebra
suitable for quantum gravity and the other gauge forces too:
d = sl(EQ) ⊂ sl(Q).
diff ≺− diff (116)

Level 5 supports sl(216 ) on its first grade. The second-cumulant Σ2 sl(216 )


0
then represents the reformed coordinate and momentum operators Π2 LE E ∈
5Q2 with quasi-continuous but finite spectra.

10 Generations
Empirically, the fermion spectrum is unexpectedly divided into three genera-
tions differing only in their masses, their coupling to gravity. Properly under-
stood, this must tell us something equally peculiar about the fine structure of
the quantum plexus.
The division into generations is not quite unique. One basis diagonalizes
the fermion mass operator M , another the weak charge operator QW , another
the coupling coefficient gH of the Yukawa coupling to the Higgs field. The
unitary transformation from the charge basis to the mass basis is fermion
mixing. It reduces to a lepton-mixing part UW M , the PMNS (Pontecorvo-Maki-

Nakagawa-Sakata) transformation, and a quark-mixing part USM , the CKM


(Cabibbo-Kobayashi-Maskawa) transformation.
The generation number Γ is often described as a radial or principal quan-
tum number of the particles, in analogy to the radial variable r or principal
quantum number n of atomic physics, because Γ commutes with the groups of
the standard model, whose parameters are likened to angles, as n commutes

63
with angular momentum. On the other hand, there are remarkable violations
of this analogy in the relation of Γ to multiplicities and to couplings. The
multiplicities of the standard model group representations are independent of
Γ, while those of the atomic rotation group depend markedly on n, namely
quadratically. For example, the multiplicity Mn of hydrogen shells changes
from M1 = 2 for the K shell to M2 = 6 for the L shell. There seems to be a
unitary operator connecting the generations, but not the shells of hydrogen.
In the limit of very large radius n → ∞, however, the fractional change
from one hydrogenic radial shell to the next becomes small:
Mn+1 − Mn
→ 0. (117)
Mn
As n → ∞, neighboring shells of hydrogen then become approximately isomor-
phic, in that their differences are small fractions of the total. Γ still resembles
a radial variable, but a very large one, compared to the radial distance between
shells. The generations are shells that are flat and isomorphic on the scale of
particle physics.
This fits the dome model reasonably well, for the dome has a cosmic radius
and a subparticle thickness, and so its strata can be practically flat. On the
other hand, instead of one transverse dimension the dome has one for every
standard-model charge.
What singles out the generation coordinate Γ? And why do fermion masses,
to a crude approximation, seem to vary geometrically with Γ? Roughly speak-
ing,
M (Γ) ∼ M (0)eΛΓ , (118)
with a large inter-generational mass ratio eΛ ∼ 102±1 . No radial variable
of elementary quantum mechanics effects the particle energy so powerfully.
Perhaps this is a clue.
Some speculations are obvious in an infraquantum perspective. Though
still poorly founded, they might be worth future study:
The Higgs field singles out one direction in the dome at each event, and
determines some gaugeon masses. Parsimony suggests that Γ indexes eigenval-
ues of the quantized Higgs field. In the infraquantum theory, correspondingly,
it is conceivable that Γ indexes the spectrum of (bı)2 .
One familiar exponential process in superconductors is shielding or pene-
tration. In a boundary layer the electric vector potential falls off as A ∼ e−µξ
with penetration ξ into the superconductor. Here µ is a measure of photon
rest mass hµ/c within the superconductor. In a bulk superconductor the skin
depth λ is a small part of the bulk, and the mass shift of the Cooper pair is a
small part of its mass, so superconducting generations, with various values for

64
the Cooper-pair mass, have not been observed, as far as I know. If this analogy
has some validity, generations of Cooper pairs would be seen more clearly in
a superconducting film or graphene several atoms thick. What is the mass of
Cooper pairs in the boundary layer of a superconductor? Let us return to the
more exact considerations necessary to solidify such conjectures.]

11 Particle valence
11.1 Standard model valences
In this section we use a terminology for the standard model taken from chem-
istry. Each quantum of the standard model has a single-particle Hilbert space
that is a product of several factor spaces on which act orbital, spin, and charge
operators. The factor spaces provide the indices on the statvectors of the par-
ticle, describe formal parts of the particle, and are transported by the gauge
connections, so they have some semblance of physical existence. Yet they are
not quanta in space-time, which is only one of these parts. We may call these
phenomenological parts of the quanta valences. In an infraquantum theory
valences arise naturally as entities of lower type than particles and space-time.
What is known of the valences of the quantum particles today is summed
up in the Feynman vertices of the standard model, such as the electron-photon
vertex
e
γ ∼∼∼• or γ → e + e. (119)
e
In an infraquantum theory this must be a singular organized limit of the actual
process. Photons are at least dyadic, and should be represented by at least
two lines More generally, Q bosons have statvectors of even grade, and are
composed of an even number of fermions. Furthermore, the components of the
momentum do not commute, and cannot all be assigned eigenvalues at once.
De Broglie and Feynman tried assembling photons and gravitons (respectively)
from fermions, without positive results. They worked in a classical space-time
with quanta, subject to the Heisenberg indeterminacy principle, which requires
strong forces for close binding.
But Q weakens the Heisenberg indeterminacy principle at small range, and
simultaneously provides another way to bind monads into polyads: covalent
bonding below the event level. Polyadics cannot be bound with an ι since suc-
cession results in a monadic, but can be bound by sharing second, third, or n-th

65
members of lower type, which may be valences. Let us call this phenomenon
deep covalent bonding.
It results in pseudo-bosons with hard fermionic cores. The dyadics

ψ = abc ± ab c =: ◦ ±
◦ ,
(120)
φ = ab cd ± ac bd =:
• ◦ ±
◦ •,

are simple examples of deep covalent bonds within Q. In the example ψ, the
leftmost circle represents a, the inner circle b, and the rightmost c.
One candidate for the photon of (119) under Q resolution might be


γ ==== γ → e + e− (121)

e.

in which the horizontal lines represent two electron processes held together by
deep covalent bonds like those of (120). This model cannot be right. It would
not account for direct photon emission by particles other than the electron.
The standard model gauge group itself suggests a more complex structure
for the leptons, and a simpler structure for the gaugeons. In the standard
model kinematics, the chiral lepton and quark and the gaugeon have tensor-
product statvector spaces, which can also be written as exterior products:

Lepton = Orbit ⊗ Spin ⊗ Isospin = L2 (R4 ) ⊗ 2C ⊗ 2C,


Quark = Orbit ⊗ Spin ⊗ Isospin ⊗ Color = L2 (R4 ) ⊗ 2C ⊗ 2C ⊗ 3C,
Gaugeon = Orbit ⊗ Spin ⊗ Valence ⊗ DualValence. (122)

This expresses the chiral (left-handed) lepton as a composite of a scalar par-


ticle, a spin, and an isospin; the quark has a color as well. The anti-chiral
right-handed fermions lack the isospin valence. The gaugeon carries a Lie
algebra element that is also a dual pair of valences.
(122) is not a constitutive hypothesis added to the standard model; it is
parr of the standard model. These constituents are not yet physical particles,
but valences for the three forces, generalizing those of chemistry. Let us develop
this terminology further.
Label the three current kinds of force with G (gravitational), W (weak,
electromagnetic, and hypercharge), and S (strong) for brevity. Fermions may
carry approximately conserved charges of several kinds. The G charges are
momentum-energy pµ and spin γ νµ (µ = 1, 2, 3, 4), The W charges are weak
isospin and hypercharge wi ∈ u(2) (i = 1, 2). The S charges are colors sC ∈
SU(3) (C = 1, . . . , 8).

66
Each gaugeon of the standard model relates an orbital direction and a pair
of valences (G, W, or S). The orbital direction can itself be identified with
a pair of G valences, spins. Thus the gaugeon has a tetrad of valences of
appropriate kinds. The diagram (121) is replaced by the finer structure


γ ==== γ → e + e− (123)

The electron line has been broken into a thinner line standing for an inner
valence and a thicker line that carries everything else including the spin. The
gaugeon line is now composed of two valence lines rather than two electron
lines. It can interact with any fermion carrying an appropriate valence, not
only with an electron.

11.2 Q valence
To form a Q correspondent of the valence concept, let us assume tentatively
that each fermion is a product of its valences, held together by deep covalent
bonding.
This raises the question of whether the dynamics can effect such binding.
The relevant pre-reformation skew-action is the fermionic term ΦF of (139).
According to the standard model, W and S gaugeons map external (orbital)
Lie algebra elements into internal ones. Allowing two valences to define a Lie
algebra element, they have the structure L2 T 2 . According to canonically quan-
tized general relativity, G gaugeons have the valence L4 . In a pre-organization
theory, whose symmetry is not yet broken by the split into L and T statvectors,
the three gaugeons merge into one.
At present, however, superselection rules separate the G, W, and S valences.
Superpositions of different valences do not seem to occur. A superselection
rule, like a selection rule, implies a symmetry, but also implies decoherence.
Decoherence can be due to random disturbances that destroy phase relations
between two terms in a direct sum. They occur, for example, when each of
the terms in question is itself a sum of too many sub-terms for us to resolve
their phases accurately. In Q dynamics, the large numbers involved seem to
be the number of paths joining two events, and the number of steps along the
transport path. Too many steps make the space-time coordinates central. It
is possible that too many paths result in the superselection rules between G,
S, and W.

67
The gauge Lie algebra of the standard model is a direct sum of the gauge
Lie algebras of hypercharge Y , isospin I, and color C, of the kind
 
u(1)Y
gS =  su(2)I  = u(1) ⊕ su(2) ⊕ su(3). (124)
su(3)S

This is not the invariance Lie algebra of any quantum entity supposed to exist
in the standard model. Its defining representation acts on a 6-dimensional
direct-sum complex vector space

V⊕ := C(Y) ⊕ 2C(I) ⊕ 3C(C) = 6C. (125)

This is the statvector space of a hypothetical quantum that can be a hyper-


charge Y, an isospin I, or a color C. No such quantum entity is believed to
exist. For example, a quark has a Y, I, and C all at once, as indices on its
statvectors. Its charges act on a 6-dimensional tensor-product space

V⊗ := 2CW ⊗ 3CC = 6C, (126)

not the direct sum (125). The quark theory is invariant under the charge Lie
group
GQ := U(1) ⊗ SU(2) ⊗ SU(3). (127)
To infraquantize the three interactions involves representing these groups in
SO(EQ1 ), along with the groups of general relativity.

12 Infraquantum fields
We express the usual construct of a spinorial quantum field over classical
Minkowski space-time M operating on a Hilbert space H as a singular or-
ganized limit of a polyad with statvectors in EQ. The constructs of field space,
space-time, and Hilbert space thus merge into the typed exterior algebra Q.
The correspondence between the canonical and infraquantum field con-
structs is crucial for a Q theory. It arises from the dome structure of the
ambient plexus as follows.
In a completely classical model with field values in a linear space F and
space-time events in a finite set X , the field state space is the power-set F X .
Its multiplicity is
|F X | = |F||X | . (128)
A Lagrangian dynamical theory then uses the dual linear space F 0 for the
canonical conjugate of the F field. In the canonical quantization, (F ⊕ F 0 )X is

68
used as statvector space for one quantum. Its vectors represent input-output
processes for the quantum.
When Xb is a quantum space, however, as in an infraquantum theory, there
is no satisfactory power-set construct (F ⊕ F 0 )X . There is the binary expo-
b

2X , but this is not a proper power space of b


nential b 2.
b

(The space of linear maps Xb → Yb is a mathematically natural candidate


for the needed power space, but this is merely the statvector space of one pair
of one X quantum and one dual Y quantum, and is not the statvector space
required. Its multiplicity is merely |Y|
b |Xb|.)
The binary exponential can be used, however, to make a quantum theory
that becomes a canonical quantum field theory in a singular organized limit.
Suppose that the dome organization reduces the space of E monadics EQ1 as
vector space into a tensor product of two vector spaces, a transverse part Y
that is of small extent in the dome and has a single point 0 as classical limit,
and a longitudinal part Xb that is large for the dome and has a classical limit
X ≈ M:
E 1
Q ≈ Yb ∨ Xb − Yb ∨ X . (129)
Then the field statvector space factors according to
E
2 Y∨X .
Q = ext EQ1 = b (130)
b b

In the singular organized limit

 2Y∨X ∼
2 Y∨X − = (2Y )X := ZbX , (131)
b b b b
b

a classical field emerges with quantum field variable Zb := b 2Y and classical


b

space-time X .
This limits us to field spaces that have logarithm spaces. It is convenient
therefore that a spinor space Ψ = ext V has a logarithm, namely its semivector
space V (§6.2).
Thus even though there is no natural Q construct of a quantum function
on a quantum space-time, the dome organization naturally provides such a
construct as a singular organized limit of a quantum set of monads; and the
field so defined can be a spinor field, or any polyadic product of monadic spinor
fields. It is so far merely plausible that this path can be retraced to recover
the spinor field on classical space-time from the infraquantum theory in Q.

12.1 Q gaugeons
We turn now to infraquantum gauge theory. The central object of the standard
gauge theory, following Einstein and Weyl, is a gauge differentiator, a reformed

69
version of the Lie partial differentiator ∂µ . Three known gauges, gravitational
(G), electroweak (W), and strong (S), combine without unification into a grand
gauge differentiator

Dµ = ∂µ − Gµ (x) = ∂µ − GµG (x) − GµW (x) − GµS (x), (132)

with grand vector potentials Gµ that act on each tensor according to its nature
and degree. We have absorbed the relevant coupling constants and current
operators into the grand vector potentials Gµ . Then all the coupling constants
appear as appropriate “fine-structure constants” αG , αS , αW standing before
the respective gaugeon skew-action operators, yet to be written.
The infraquantum correspondent of Dµ is the tensor Lc0 c [E] of cumulated
generators of the reformed simple gauge Lie algebra so(CQ1 ). Lc0 c [E] also
includes correspondents of the space-time coordinates, angular momenta, and
the infraquantized imaginary bı.
Since the kinetic energy of a test particle is part of Dm and the potential
energy is part of G m , let us speak of the three terms in any gauge differentiator

Dµ = ∂µ − G µ (133)

as kinetic (D) , total (∂), and potential (G). The gauge invariant one is
the kinetic, Dµ . Einstein’s principle of local equivalence implies three local
equivalence conditions on Dµ :
1. Dµ respects the tetrad vectors γ µ and so the metric tensor.
2. Dµ is atorsional, respects the coordinate tangent vector fields en (x).
3. Dµ agrees with the Lie derivative ∂µ on the scalars xν :

[Dµ (x), en (x)] = 0 = [Dµ , γ ν (x)], [Dν (x), xµ ] = δνµ . (134)

These conditions are all structurally unstable, and must be reformed in an


infraquantum theory. It is sufficient for immediate physical purposes if they
hold in a singular organized limit.
The full skew-action operator Φ for the standard-model with gravity in-
volves a fermion annihilator ψ(x), the Dirac spin operator γ µ (x), the grand
differentiator Dµ (x), a Higgs isospinor scalar φ(x), and other spin-like vari-
ables. As each gaugeon was adduced, an action operator was invented to
control it. Schematically,

ΦF Φ G Φ WS ΦH
z }| { z }| { z }| { z }| {
Φ = iψγDψ + αG iγγ[D, D] + αW,S i[D, D]γγ[D, D] + i(φD2 φ − W(φ)) . (135)

70
The fermion term is linear in Dµ . The gravitational skew-action operator
ΦG is the term quadratic in D, and involves only DµG , the G part of Dµ .
The electroweak term ΦWS is quartic in the grand differentiator Dµ . The
standard-model Higgs term ΦH involves a colorless isospinor spinless scalar φ
in a W-shaped potential, a source of G and W gaugeons but not S. Yang space
provides an isovector scalar bi −
 i instead.
It has often been suggested, since Kaluza, that some of these gaugeons are
expressible in terms of the others. In an infraquantum theory based on Q, the
gaugeons are associated with polyadics of even grade, composed of an even
number of monadics, representing fermions. We should therefore explore the
possibility that all the skew-action operators might be effective surrogates for
the fermion one. Specifically, we have ordered the terms on the right-hand
side of (135) so that each might be a manifestation of the ones preceding it, all
ultimately flowing from the fermion skew-action operator, the only one that
contains all the fields ordinarily considered to be independent.

12.2 The emergence of space-time


The standard model skew-actions are designed to couple gauge fields with
their sources. None was designed to couple the events of nature into a space-
time quasi-continuum. The dynamics that condenses the plexus into a plenum
operates on a deeper level than can even be described in the standard model
language. If we do not need a completely new dynamical theory to describe
this, we need at least a reformation of the existing dynamics.
It seems that the organization of the space-time plenum requires the limits

bı −
 i, pµ , Lµ0 µ −
 0, (136)

the first to recover the canonical commutation relations for orbital variables
and field variables, the rest is to reduce the x, p-phase space to x-space-time.
It is implicit that the variables xµ , which commute in this limit, are not con-
strained; that the plenum is four-dimensional, not a lower-dimensional quasi-
manifold like a wire or a bubble. In §13 we review how a condensate is described
in a history-based quantum theory.

13 Infraquantum standard model


DO: Produce an ambient dome statvector.
DO: Make MISM (Minimal Infraquantum Standard Model), repressing the
desire to unify or innovate.

71
13.1 Infraquantum fermions
Let us begin by infraquantizing the fermion skew-action operator ΦF of (139).
Prior to infraquantization, but after a canonical quantization that includes
gravity, the variables in ΦF represent tensors of the valences shown first in
(122), rendered contravariant for convenience. We write x for the coordinates
of a point in space-time, s for the spins-and-unitary-charges of the fermion
annihilated by ψ, µ for a Minkowski vector index, f for the fermion quantum
occupation numbers and g for gaugeon occupation numbers connected by ψ.
Then simplification results in the valences shown on the right-hand sides of
these equations. It replaces each Minkowski index and event µx of Dµ by a
pair of event indices e0 e, suitably extended to represent the standard model
internal unitary charges as well, and undergoing a split into spin and orbit
in the limo. Let us write the resulting “cell” orthogonal group of the sub-
event as SO[C]. its vector indices as c, c0 , and its Q spinor indices still as
s, s0 . In a canonical theory, ∂µ and xµ are canonically conjugate as single-
0
event operators (Le e ) on the event type level E. Probably E = C + 2 will
suffice. While c need assume but 16 values or fewer, after infraquantization
the spinor index s ranges over cosmologically values, perhaps 26 = 264K . Then
0
covariance requires a corresponding conversion γ µ (x) ≺− γ e e . We designate
fermion occupation quantum numbers by f, f 0 , and infraquantum numbers by
fb, fb0 ; and analogously for gaugeons g and gb. In sum,
0 b0 b
ψ = (ψ xs|f f ) ≺− ψ e|f f ,
xµ|s 0 s|g 0 g 0 0
γ = (γ ) ≺− γ e e|bg gb, (137)
0 0 0 0
D = (Dxµ|s s|g g ) ≺− Le e|bg gb.

Let us also continue to suppose that all forces are ultimately exchange
forces, differing in what is exchanged:

Gauge forces result from valence exchanges. (138)

In Q theory, “empty” space-time will also be supposed to be held together by


exchange of spins at adjacent events. Let us suppose that this binding force is
gravitational.
Since the monadics of Q have spin 1/2, we are back to a question of Feyn-
man: Could a graviton be a composite of four fermions?
The graviton appears most nakedly inside the skew-action operator of a
fermion, thus:
Write ψ(x) for the fermion spinor operator field, and Dµ (x) for the grand
covariant differentiator. Write the Pauli adjoint of ψ as ψ, defined so that
0
ψψ is an invariant symmetric operator and ψγ µ µ ψ is a skew-symmetric tensor

72
of skew-symmetric operators. The Dirac spin operator field γ µ (x) and the
differentiator Dµ can be regarded as field operators for gravitons. Then the
usual skew-action operator coupling them is
Z
ΦF = i (dx) ψγ µ Dµ ψ. (139)

A Q reformation replaces:
• ψ(x) by an operator Lψ where now ψ ∈ EQ1 .
R
• i (dx) ψ . . . ψ by a cumulation ΣL65 (§6.4).
• Dµ (x) by Lc0 c .
0
• γ µ (therefore) by γ c c .
Due to the hanging indices of L65 the result is not yet a tensor.
In classical gauge theory the gauge derivative coincides with the Lie deriva-
tive for coordinate functions:

[Dµ (xλ ), xν ] = [∂µ , xν ] = δµ ν . (140)

This is a singular condition on the theory. The canonical commutation re-


lations between D and its conjugate field variables are even more singular,
involving Dirac delta functions. As a result, the current gauge theories of
gravity and the standard model are compound and singular, and so must be
reformed and simplified. Infraquantum theory replaces both singular commu-
tation relations by regular Lie algebraic commutation relations, of the event
type and field type respectively.
Diff, the gauge group of gravity, can serve as a prototype for them all.
(140) is a singular organized limit of the Lie algebraic relation
00 0 00 A0
[LA , LA ] = cA AL
A
(141)

of some simple Lie algebra A, here of operators on Q, with constant structure


00 0
tensor cA A A . (140) is diff-invariant but (141) is A-invariant.
For definiteness, let A = so(CQ1 ) be the orthogonal group of a cell level,
0 0
represented by generators of the Clifford form Lc c = γ c c . In an infraquantum
0
theory, the gauge differentiator is a cumulation of Lc c from the cell level CQ1 )
to the event level EQ1 . This can be represented by a single element of Q,
a contraction of the cell operator with the event operator, having the Dirac
spin-orbit operator as a singular organized limit:
0
L := γ c c Σ2 γc0 c ≺− γ m Dm . (142)

73
For gravity all four monads in the tetrad are longitudinal to the dome. For
the other gauge interactions, two are transverse to the dome.
Spin and orbit variables undergo the same group ISO(3, 1) in special rela-
tivity, but not in general relativity. Unlike the vector representation, the spinor
representation of so(3, 1) does not extend to one of sl(4R), whose irreducible
unitary double-valued representations are infinite-dimensional. Therefore, to
general-relativize the construct of spinor, Cartan introduced a mobile frame,
a field of tangent-vector quadruples γ 0 (x), . . . , γ 3 (x) with a fixed Minkowski
inner product γm (x)·γn (x) = gnm . This defines a Lorentz group SO(3, 1; x) on
each tangent space, but events transform under Diff. A Cartan vector-sextuple
would appear in a theory of Yang so(3, 3).
The Dirac spin operators γ µ have had several metamorphoses since their
conception:
1. In gravity theory the Cartan vector-quadruple γ µ (x) becomes the dynam-
ical variable for gravity. The classical gravitational metric at x, described
in the mobile frame, is then the tensor
0 1 0
g m m (x) := {γ m (x), γ m (x)}, (143)
2
which is arbitrarily fixed to be a constant Minkowskian form in the mobile
frame.
2. In Dirac’s theory of the single electron, the γ µ become monadics of a
Clifford algebra and represent physical quantum variables of the electron.
3. In the Yang space-time the coefficient γ µ in the Dirac equation become
a sector γ µ6 of the generator γ nm of Yang so(3, 3), contragredient to the
momentum:
pµ ≺− Lµ6 , γ µ ≺− γ µ6 . (144)
4. In Q the single-electron γ µ become annihilators of type B, the predecessor
of spin.
5. In standard quantum field theory the γ µ are constant coefficients of the
electron wave equation, not dynamical variables.
6. In canonical quantum gravity, the matrix elements of the γ µ (x) become
graviton annihilators.
7. Q gravity uses the cumulant spin operators
0 0
γ c c (E) := ΣE−C γ c c (C). (145)
The Dirac operator becomes
0
γ c c (E)Lc0 c (E) −
 γ µ ∂µ (146)

74
0
The Q operators γ c c (E) act on the spinors ψ ∈ Q(E), which carry both
spin and orbital information (§9.2). Therefore each γ c carries information
about two spins and two orbital variables, input and output. The Q gravita-
tional metric corresponding to (143) is evidently the numerical tensor
0 1 0
g c c := {γ c , γ c }, (147)
2
involving four spin and orbital variables. It will be symmetric in its spin
variables if it is skew-symmetric in its orbital variables. It is simply the natural
neutral metric on W(E), referred to a mobile frame. The variability of this
metrical structure derives from the variability of the Q simplicial complex to
which it is applied, in the way that a variable curved surface in a flat space of
constant metric has a variable metrical structure.
The Q correspondent of the gauge covariant differentiator Dµ is the genera-
tor on the field level F induced by the cell Lie algebra generator LC of level C.
It requires F − C iterations of cumulation Σ. The resulting correspondences
are:
tangent space ≺− Q simplex of level C
Cartan n−ad γ n (x) ≺− Q n-adic in EQn
diff ≺− sl(EQ1 )
Dµ (x) ≺− ΣE−C LC =: LC [E]
s0
Ksm 0m ≺− [LC 0 [E], LC [E]] (148)
Because the cell map is an orthogonal transformation, the local equivalence
principle is automatically satisfied.
A Q quantum theory has no supersymmetry between bosons and fermions.
Even operons are composed of odd, and not conversely. The generating vari-
ables obey a Fermi statistics, not a Bose.
The atomistic analysis of the other gauge fields can be modeled on that of
gravity. One adjoins to the six cellular CQ dimensions required for Yang so(5, 1)
as many more dimensions as needed for the unitary groups of the standard
model, raising the cell level C if necessary. This enlarges the generator LC of
sl(CQ1 ) accordingly. The induced generator LE F −C L then includes all
C := Σ C
the coordinate and momentum variables of the full gauge field.
Near the singular organized limit, where one can speak of space-time points,
we have supposed that the gravitational field variable γb is not attached to one
event but two (§8.1). This makes it possible to form a pseudo-boson from
the basic fermions. The field tensor can be symmetric in its space-time indices
because it is skew-symmetric in its space-time points, although these points are
so closely bound that they presently pass for one point. All apparent bosons
are hard-core pseudo-bosons in this model.

75
The four flat-space Dirac γ m ’s of standard spin 1/2 theory have been as-
sumed to be what remains of the cell operators γ w ∈ W[C] after the dome
organizes itself. The γ w transform as so(3, 1) vectors when 4Q1 transforms as
a spinor space:
0
Λγ c Λ−1 = Lc c0 γ c , Λ ∈ sl(4Q1 ) = sl(16R), L ∈ so(3, 1). (149)

Then Q gauging undoes the reduction

so(4Q1 ) ⊗ so(2) ≺− so(6Q1 ). (150)

The function γ m (x) is covariant under diff, so we assume that Qγ m (x) is


covariant under Q diff = sl(6Q1 ). Suspending the polarization that produces i
and p, the γ m merge into left multiplications by any basis element ec of CQ1 ,
with C = 4 provisionally. Upon Q gauging, this becomes

γ m (x) ≺− L ee , (151)
5
where the ee are the basis elements of Q.

13.2 Infraquantum fermion dynamics


In the canonical theory, the gravitational metric gµ0 µ is constructed from γ
and its canonical conjugate from D, and they are connected by symplectic
transformations. While D and x are canonically conjugate on the event level
E, D and g are canonically conjugate as gravitational field operators on the
higher field type level F = E + 1.
In the simple theory canonical conjugates unify into sectors of one tensor,
let us suppose representing an orthogonal group, as x, p,bı do in (2). They re-
separate in the singular organized limit. Therefore the infraquantum tensors
L and γ are the same on each type level. For example on the event level,
0 0 0 0
L = (Le e|g g ) = (γ e e|g g ) = γ. (152)

We will use γ rather than L to remind us that we are using its spinor repre-
sentation. Then the core of the fermion skew-action operator simplifies to
0 00 g 0 0 0
γ · D ≺− γ e e|g γ ee |g g . (153)

When we supply the outermost fermion factors ψ . . . ψ in the skew-action, it


becomes
Z
ΦF = (dx)i ψ γD ψ

76
≺− Φ
bF
h 00 00 00
∼ Tr ψbf ”e” f ”00 e”00 c”00 Lf ” e” c” f ”0 e”0 c”0
0 0 0 00 0 0 0
i
× γ f ” e” c” f ”e”c” Lf ”e”c f 0 e0 c0 ψbf e c f e
∈ Lin FQ. (154)

DO: : Complete this calculation.

13.3 Constructing the Higgs field


It has already been proposed that a quantized imaginary play the role of a
Higgs field (§9.3). Here the quantized imaginary is one element of the Yang
generator L.. , and the differentiator D is another sector of L. Both appear in
the quantization of ΦW . This identification would absorb the H skew-action
operator into the W. The factor i in ΦF then becomes a bı − η in Yukawa
interaction with the fermions.

13.4 Constructing the electroweak gaugeon field


Kaluza and deWitt showed in a classical context how an action linear in the
curvature like ΦG could absorb one quadratic in the curvature like ΦW :
Suppose that some manifold dimensions of the gravitational G theory be-
long to a Lie algebra a. This was electric so(2R) for Kaluza, and an arbitrary
Lie algebra a for de Witt. If A indexes an a basis, the metric component
gmA is also a connection, coupling the transport direction m to a Lie algebra
element A. If a is the appropriate sum of the W and S gauge algebras, the G
skew-action operator ΦG includes a quadratic one of the form ΦWS .
The Kaluza-deWitt strategy works even better in the infraquantized theory:
Where Kaluza had but one Lie-algebra dimension among four other manifold
dimensions, the Yang-Segal simplification converts all the manifold dimensions
to operators in a Yang Lie algebra. For example, ∂m becomes one sector of a
generalized angular momentum tensor Lc0 c . Lc0 c is represented on a cellular
0
level by a dyadic spin operator γ c c =: γ C in the Lie algebra of Cliff 2 W[C];
0
and Lc0 c is the iterated cumulant ΣE−C γ c c ∈ Cliff W[E], the E level.
0
Then every component of the metric form g C C has a Lie-algebra index C
and admits a second interpretation as a connection amplitude. Let us suppose
provisionally that the W skew-action operator is extracted from an extended
G skew-action operator by such a Q version of the Kaluza-deWitt strategy.

77
13.5 Constructing the strong gaugeon field
DO: Do this.

13.6 Constructing the gravitational field


This brings us to another often-proposed unification, the last of this sequence:
The representation of the graviton g as a fermionic tetrad γγ suggests reducing
the gravitational G dynamics to the fermionic F by eliminating the fermion
F
annihilators ψ from the term ∼ (ΦF )2 in eΦ . One eliminate external lines by
taking an ambient expectation value, internal ones by summation.
DO: Do it!

14 Infraquantum fermion dynamics


For practical reasons, Q dynamics divides into two realms:

1 Pre-organization dynamics. Dynamics of a logarithmically small, disor-


ganized patch, relative to a metasystem supported by an organized ambient
dome.

2 Post-organization dynamics. Small excitations in the organized dome, in-


volving only a few defects, [hopefully] including the standard model quanta.
Let us assume that the grand skew-action operator Φ is formed by Q-
quantizing the fermionic skew-action operator ΦF coupling the extended G
gaugeon to the fermion. For the small-excitation theory we accept that the
dome reduces the simple G gauge algebra to semisimple ones, including diff
in a singular organized limit. Yang so(3, 3) is not a symmetry algebra of the
organized dome, but can still be a symmetry algebra of a cell and of the pre-
dome plexus, and can still be a dynamics algebra of the dome.
The emergence of the dome from a less ordered collection of operons is pre-
sumably a phase transition of the plexus, perhaps beyond our theoretical reach,
but we might be able to at least formulate the problem. The sensible strategy,
however, seems to be to infraquantize something known to work, like the Dirac
equation in Minkowski space-time. Let us approach this problem starting from
the standard, unregularized, gauged, fermionic dynamics statvector
Z R 4 m
DF = exp [dψdψ]ei (d x) [ψ(x)γ Dm (x)ψ(x)+ξ(x)ψ(x)+ψ(x)ξ(x)] (155)

78
The Higgs field has been left out for now because [hopefully] it is a cluster of Q
fermions with deep covalent binding, arising from the infraquantized fermion
dynamical statvector D b F.
In the standard theory DF describes a field of spin-1/2 free fermions propa-
gating with the usual spin-orbit coupling γD between the spin of each fermion
and its kinetic momentum. It includes no direct interaction between fermions.
The fermion field operators act as a quantifier, summing over the individual
fermions in the many-fermion field theory. All interactions between fermions
arise from their coupling to the collective gauge fields of D. To describe the
propagation of the gaugeons requires gaugeon skew-action operators ΦG,W,S .
In the Q theory the fermion field operators still form a quantifier, summing
over individual fermions. The gaugeons. however, have to be replaced by
polyadics in the fermions, say of grade g, if only to fit their observed spin and
statistics. The fermion skew-action operator then couples not two fermions,
but g+2 fermions. It describes interaction as well as propagation. The gaugeon
propagator is then no longer separate from the fermion propagator, but derives
from it algebraically. It is then conceivable that ΦF is the entire skew-action
operator.

14.1 Composite gaugeons


For composite gaugeons (§14) we must choose polyadic expressions for the
gaugeons that indeed bind and propagate as gaugeons. We proceed by trial
and error.
DO: ...
We quantize (155) bit by bit, working outward from the center:

14.2 Dirac spin operators


γ m ≺− cell operators γ C ∈ Lin CQ1 , generating a Yang so(3, 3).
DO: Express these gamma’s in Q terms.

1 Differentiator
0
Dm ≺− ΣE−C γ c c ∈ EQ1 , (156)
a cumulant kinetic energy-momentum operator.

14.3 Fermion annihilation operators


ψ, ψ ≺− annihilation operators γ E . Part of item (14.6).

79
14.4 Pauli form
β implicit in ψ ≺− the Pauli form p[E] of EQ.

14.5 Integration over space-time


. This integral becomes a trace (operation) on the algebra generated by the co-
ordinates. One possible Q correspondent is a trace on the subalgebra generated
by the space-time atoms γ 5µ and represented on the event level EQ, an so(3, 2).
(Not the isomorphic anti-deSitter so(3, 2), whose coordinates commute.) This
breaks Yang so(3, 3), as does the dome and its excitation dynamics. A corre-
spondent suitable for the structure problem is a trace on the event algebra of
level EQ.

14.6 Sources ξ, ξ.
Part of item (14.6). R
An exterior integration [dψdψ] over fermionic histories. This is part of a
Fourier transform, a mere change of frame that can be omitted.

14.7 An imaginary factor


i ≺− bı = Σ2 γ 65 /N.

14.8 The orbital momentum


pm (like xm ) ≺− Σ2 γc0 c ∈ 4Q1 .

14.9 Dirac operator


A Dirac spin-orbit operator D := γ m Dm .
The standard action couples singular organized limits of imaginary, spin,
and orbital operons. We face choices for this operon coupling similar to the
early choice among SV T AP beta-decay couplings, modified today to exhibit
maximal parity violation. In β decay theory, the possible grade of the coupled
factors ranges from 0 for S to 4 for P . Here, assuming a Yang SO(6 − n, n)
group, it ranges from 0 to 6. The Q action can be
h 2 c 00 0
A = Σ γ c00 (imaginary) × γ c c0 (spin) × Σ2 γ c c (orbit)
N

80
spin
 
 ∂
z}|{
00 0
= ι  Σ2 γ c c0 × γ c c × Σ2 γ c c00  (157)

| {z } | {z } ∂ιo
imaginary orbit

with the indicated correspondences. A peels one outer ι from its operand, ap-
plies the bracketed coupling, and then restores the ι. The “imaginary” factor
bı is needed for skew-symmetry, just as the standard Dirac action requires a
factor i. Infraquantization incorporates gauging, and thus raises the grade of
the Dirac propagator from the usual two to six, describing physical interaction
as well as propagation. The Q Dirac equation is a specific topological relation
between quantum operons, relating the edges of connected cells. The reduc-
tion γ 6...1 = ±1 (§9.3) results in chiral operon spinors analogous to the chiral
fermion spinors of the standard model.
The standard model algebra s[u(2)×u(3)] of charges becomes s[u(2)×u(4)].
All the physical monads support one representation of this algebra. Black
monads actually happen while colorful monads are individually only virtual.
This is suggestively similar to the way timelike displacements actually happen
while spacelike ones are virtual [31].
While every basic standard fermion has the same spin 1/2, they do not all
have the same color representation, which acts on 1C for leptons and 3C for
quarks; nor the same isospin representation, which acts on 1C for right-handed
leptons and 2C for quarks and left-handed leptons.
This part of the exploration has analyzed the standard singular description
into atoms. This has to be worked out further for gauge variables. Then will
come a largely synthetic task of a rather different character, reassembling the
parts into a coherent whole.

15 Discussion
DO: ... Finish this section.
One possibility among many is that the generations represent three strata
within the transverse structure of the dome. It is well known that particle
masses represent shielding lengths, analogous to the Debye length of an elec-
trolyte. In a continuum, shielding proceeds continuously, Is it conceivable that
in a quantum plexus shielding proceeds in quantum jumps, the generations?
Or is it possible that the difference between the generations can be expressed
as a difference in type? I cannot formulate these questions algebraically yet.

81
15.1 Acknowledgements
I am indebted to James Baugh, Shlomit Ritz Finkelstein, Andry Galiautdinov,
Dennis Marks, Zbigniew Oziewicz, Heinrich Saller, and Sarang Shah for recent
helpful discussions and corrections.

16 Appendices
17 Quantum pragmatics 2010.04.28

In classical thought there is a clearcut distinction between basic theory and


phenomenology, so deeply woven into our traditions that some try to find it
in the quantum theory. It has ancient roots. Parmenides already divided
knowledge, episteme, into logos and doxa, and logos, arrived at by reason, was
deemed exact, while doxa, arrived at by observation and report, was approxi-
mate. In classical mechanics, the concept of the dynamical law that cannot be
learned from one observation descends from the logos, and the system state,
learned by observation only, harks back to the doxa. There is a categorical dif-
ference between the law and the system it governs. This theory of knowledge
is also the core of the perennial philosophy [42].
Several contradictory formulations of the same quantum theory are in wide
circulation. They may be broadly classed as either classical or operational
(pragmatic, if one prefers a Greek root to a Latin one). It is termed classical
here to theorize about reality, objects as they are in their essence, indepen-
dently of observation, and operational to dispense with essences and speak
instead of our operations, like perceptions and measurements, possible or ac-
tual. Operations can fail to commute; objects cannot.
A classical theory can take the universe as system. A quantum theory must
leave most of the universe outside the system so that it can carry out obser-
vations on the system. The quantum law can then be outside the system but
still inside the universe. The quantum theory thus eliminates the categorical
difference between law and system. Both are represented as operators: opera-
tors like the Hamiltonian H for the law, and operators like io operations ψ for
the system. §7 takes up the immanent law of quantum theory that regularizes
the emmanent law of classical mechanics.
The present formulation is pragmatic and processual.
The quantum mode of reasoning could be called quantum logic, but many
still assume that logic is a priori, and here quantum theory is understood as
afortiori, so the term “quantum logic” could seem oxymoronic; and in any case
it was preempted for a certain lattice algebra of little computational utility.

82
The quantum mode of reasoning is based on operations or actions and so is
aptly termed pragmatic by Stapp [36]. Let us speak of quantum pragmat-
ics rather than quantum logics, to avoid the misconceptions mentioned and
emphasize the empirical nature of the study in question.
The system under study is taken here to be specified by how experimenters
can produce it, act on it, and register it; by their input, throughflow, and
output actions. The rest of nature, including the experimenter, instruments,
and life support, constitute the metasystem.
Historically the metasystem was known only coarsely and described in nat-
ural language. Now it is possible to operate at least on a logarithmically small
part of it with quantum precision. If this part can still serve as a metasystem
for a still smaller system, then it may be considered to be a quantum meta-
system. The possibility that such extensions of quantum theory might become
necessary was anticipated by Bohr [6]. But then the rest of nature, including
ourselves as experimenters on this metasystem, remains coarsely resolved, and
can be regarded as meta-metasystem, or meta2 system.
In the present study, space-time coordinate variables are quantized. Space-
time variables of canonical quantum field theory are not measured on the
field system but on individual quanta, one type lower. Their quantization is
therefore an infraquantization.
Let us reserve the term state as in classical mechanics for an observable
property of the isolated system that determines all others. For an isolated
system, knowing the state is complete knowledge of the system, by definition.
When we sharply determine any property of a physical system, however, it
seems that we change other properties of the same system, which are not be-
ing determined, greatly and uncontrollably. Therefore physical systems do not
seem to have states. In quantum pragmatics a ray in a Hilbert space H pro-
vides a maximally informative—sharp—statistical description of a population
of systems, certainly not a state. A ray is conveniently specified by one of its
vectors, which therefore is also not a state.
Formulations of quantum theory using the term “state vector” often say
that every quantum system has a state vector. This assumption seems mean-
ingless or false, but is so widespread that the only practical way to evade it
to drop the term. Here a non-zero vector in the ray describing a sharp input-
output process is called a statistical vector or statvector . Statvectors and their
duals represent sharply defined input and output quantum operations. The
individual system does not “have” a statvector in the sense that it has a mo-
mentum, or in the sense that a classical system has a state. Statvectors also
have a product, by which aggregates are formed from individuals.
The Hilbert space arises because physical filtration operations that are
crisp—meaning P 2 = P —seem to form a projective geometry. The sharp

83
projectors are its points or rays, the next sharpest are its lines, and so on. The
incidence relation defining this projective geometry is easily expressed in terms
of the inclusion relation A ⊂ B, and this is defined in terms of a statistical
null relation AøB meaning that “every” beam that passes the A test entirely
fails the B test, and conversely. Then

A⊂B := ∀C (CøB ⊂ CøA) (158)

Two points fix a line, and two lines in a plane fix a point, just as in classical
logic. But many points lie on every line, as in projective geometry, not classical
logic. They are all called superpositions of two points that fix the line, and
this is the superposition principle, the feature peculiar to quantum pragmatics
and alien to classical logic. What is often called a quantum logic is a Galois
lattice of the null relation.
To be quite explicit:

An individual quantum system has neither a state nor a statvector. (159)

The system has no state because quantum physics is radically indeterministic;


no statvector, because an individual is not a beam or a statistical population,
quantum or no.
What can have a unique statvector are the input, throughflow, and output
phases I, II, III of an ideal quantum beam experiment. The transition metric
form h of the Hilbert space maps each input statvector ψ to the unique output
dual statvector hψ of an experiment that results in a transition on every trial.
If the statvector III and dual statvector I have unit h- norms, the transition
probability amplitude and probability are

A = III · II · I, P = |A|2 . (160)

An operator on H, and now a statvector in Q, is intended to represent an


operon, a quantum physical process carried out on the syatem that respects
quantum superposition relations among statvectors.
Canonical quantization represents the canonical Lie or quasi-Lie algebras
of a classical theory in the Lie algebra su(∞) of bounded hermitian operators
on Hilbert space H. When von Neumann proposed Hilbert space and quan-
tum logics (his plural) as the platform for elementary quantum mechanics and
quantum field theory he recognized that it was non-relativistic, describing all
space at one time, ignoring light lag 1/c [44]. Relativistic theories are field
theories, and therefore use sets of sets, constructed for example by ι and ∨.
For Boole, a class was a process, namely of selection, transforming one popula-
tion to another. Von Neumann, in his doctoral thesis, also formulated classical

84
set theory as a theory of transformations. His quantum logic was also a the-
ory of transformations, namely projection operators, and he later mentioned
“quantum set theory”, but apparently never constructed one. The quantum
set theory Q is a specially simple and limited quantization of classical set
theory, replacing classical logic by a quantum pragmatics, and retaining the
transformational interpretation of Boole and von Neumann.
In quantum theory it is especially important to distinguish mathematical
objects from physical, since they are handled with different logics. In classical
mechanics one may relate mathematical to physical objects by a generalized
commutative diagram:

computation
M0 → M1

input ↓ ↓ output (161)


.............................................................................
P0 → P1
dynamics

Above the dotted line are mathematical operations on symbols, taking


place in the metasystem; below, symbols for physical actions on the system
under study. This interface exists for classical physics as well as for quantum
physics, but is generally omitted from classical theories, which neglect the
influence of measurement on the system and the limitations of measurement.
If the computational transformation is expressed in a set algebra like C, the
roof of (161) now resembles a large truss assembled from ι struts by ∨ and
resting on 1. C softens the unphysical conceptual wall between space-time and
the other variables of physics by making both out of iotas.
For a quantum system, however, (161) fails on the level of the individual,
which is generally unpredictable. For a valid dagram, the theory must confine
itself to averages over many runs of each experiment. One does not merely
accept or reject such a statistical theory but commonly sets up a confidence
level for its acceptance, taking into account the pragmatic consequences of ac-
cepting or rejecting the theory, and one may revise the decision. Kolmogorov
emphasized that probabilistic theories are not objective; this applies to quan-
tum theories in particular, which have been called non-objective.
A metasystem can coherently produce and register a beam of systems only
if the metasystem is exponentially larger than the system. The interface is
therefore not as freely movable as von Neumann said [44]. We must put the
interface where there are few active connections, so that we isolate the system
to be studied rather than destroy it. Often the interface is located in vacuum.

85
The vacuum too is full of connections, but they may be inactive on the scale
of current measurements. And we must leave enough metasystem outside the
system to carry out the study.

17.1 The quantum i


Since the central imaginary i is the center of canonical quantum theories, it
might help to understand its physical role.
The i of quantum theory is mainly important for the classical limit and the
correspondence principle. The Dirac correspondence between classical Poisson
brackets [a, b]P and quantum commutators [b a, bb] is

\ i b
[a, b]P = [b
a, b]. (162)
~
It converts the Hamiltonian dynamical equation into the Heisenberg dy-
namical equation
dq i
= [H, q]. (163)
dt ~
For fermions this i can be cancelled by one in the Hamiltonian, though not for
bosons.
In the present infraquantization, the imaginary i emerges in a singular
organizing limit. This is consistent with its use in the correspondence relation
(162), which also refers to such a limo.

17.2 The quantum cosmos 2010.04.24

If science is understood as the drawing of statistical inferences from a pop-


ulation of experiments, the question arises of how a science of cosmology is
possible, since the cosmos is unique. When W. H. Pitts asked this question
of Norbert Wiener in a lecture at M.I.T in the 1950’s, the response was that
the cosmos itself can be regarded as a population, of regions and eras. The
question applies to a quantum cosmology as well; so does the answer.
On one hand, a statvector for the cosmos seems to make no sense on its face.
An individual system has no statvector—(159)—and the cosmos, understood
to embrace all that exists, is individual. We cannot know that we have sharply
prepared ourselves, since such a preparation wipes out all memory. Therefore
we can have no sharp input or output processes for the cosmos, which includes
us. It is not clear that the cosmos can be regarded as a quantum system in
the strict sense.
On the other hand, a hypothesized cosmic statvector generates reduced
statvectors, or at least probability operators, for carefully selected subsystems

86
of the cosmos, not including ourselves, that can have experimental meaning.
The use of cosmic statvectors does not seem to be intrinsically more absurd
than the use of cosmic states. One must not take it literally, and must exercise
due caution for features peculiar to the quantum case, like non-commutativity.

17.3 Clifford algebras of Q


A Clifford algebra C is a (linear associative) ring Poly(V ) of polynomials
p(v, v 0 , . . .) in the elements of a vector space V whose addition and multi-
plication extend those of V . The coefficients are called the scalars, 0-adics, or
cenadics of C and form a commutative field and a central subring of C. The
vectors of C 1 = V ⊂ C are called the vectors or monadics of C. The sole
additional axiom is the Clifford rule:
The square of every vector is a scalar.
This scalar is also called the norm of the vector:

∀v ∈ C 1 : v 2 = kvk =: v · v ∈ C 0 . (164)

In the present study C 0 = R. The polarization of the quadratic form kvk is a


bilinear form written u · v, so that by definition kvk = v · v.
An exterior algebra is a Clifford algebra with kvk ≡ 0. The exterior algebra
of a Clifford algebra C with product t is the exterior algebra C̆ with the scalars
and vectors of C, with product written as ∨. Then C and C̆ have the same
elements and
∀u, v ∈ C 1 : u t v = u ∨ v + u · v (165)
C̆ has a grade, which is also assigned to C. The g-grade subspace of C is
written C g . ite elements are called g-adics.
A classical algebra of finite sets defines a Clifford algebra over the binary
field. The Clifford sum x + y of two sets is a statistical mix with equal weights.
The Clifford product xty is the symmetric union x∪y \ x∩y, so that Clifford’s
rule holds in the form x t x = 1.
In the primary interpretation of Q, the monadics of Q1 ⊂ Q represent
elementary acts of basic fermion annihilation or output, called monads. A
fermion annihilated by a monad could be called a monon.
In the present application of Q, leptons and quarks are monons. Polyadics
represent composite operons, here including gaugeons and Higgs. Operons
close to monads are generally modified by vacuum polarization with polyadic
corrections, sometimes large but always finite.
C is a small part of set theory, the semigroup of all ancestrally finite sets
without foundation, and is generated by a dyadic associative commutative

87
multiplication operation x ∨ y (∨yx in prefix notation), a free monadic oper-
ation ιx, read successor of x, and the empty set 1, a constant. The classical
product ∨ is the disjoint union, obeying ιx ∨ ιx = 0. Peano’s ι operation is
better suited to an algebraic theory than Cantor’s ∈ relation. ιx = {x} = x is
the set whose only element is x. The familiar set notation {x, y, . . . , z} is here
an abbreviation for ιx ∨ ιy ∨ . . . ∨ ιz.
A quantum variant of ι serves as the microcosmic strut of which a quantum
system history is constructed, an iota indeed. 0 ∈ C serves as a space-holder,
marking the absence of a meaningful expression; a back-formation from the
quantum theory, where 0 ∈ Q is the sole vector that defines no ray and no
physical process, synonymous with the ∞ of C. S. Pierce and the om (omega)
of SETL.
C[T ] ⊂ C is the level-T part of C, the subset of C whose sets can be written
with 1, ∨ and ≤ T nested ι’s. Levels nest: T < T 0 :⊂: C[T ] ⊂ C[T 0 ]. Let
C g denote the subset of C of grade (:=cardinality) g, the g-ads, products of g
monads. Let the (binary) exponential exp2 X be the set of finite subsets of
any set X. Then C is its own exponential: C = exp2 C. Define the random
(finite) set as the mathematical object with state space C.
Now for the quantum set or polyad. This is associated in quantum fashion
with the vector space Q. Vectors with the statistical interpretation of quantum
theory are called statistical vectors or statvectors to distinguish them from
states in the classical sense (Appendix 17). Any non-zero statvectors in Q
statistically described a polyad.
Q is a real exterior algebra generated, by analogy with C, by the associative
exterior product ∨, an injective linear operator ι : Q → Q, the constant
identity 1 for the empty set, and the constant 0 for the undefined. In an
exterior algebra, monadics anti-commute.
Operations that in classical logic were considered mental or mathematical
correspond in quantum pragmatics to physical operations. The Q monadics
represent operations of individual quantum annihilation, monads. The Q prod-
uct b ∨ a means doing a and then b; ιx means annihilating an annihilator. In
the present application let us suppose that annihilations of first-generation
leptons and quarks are near-monads. The higher generations are taken up in
§10.
Q is the exterior algebra over itself:
Q = ext Q. (166)
A certain sequence of rays in Q illustrated in (40) is isomorphic to C. The levels
TQ and Q[≤ T ] are defined analogously to C[T ] and C[≤ T ]. The inclusive levels
Q[≤ T ] are nested exterior subalgebras of Q.
Q has a natural Hilbert space metric, defined inductively:

88
• R = Q[1] is a one-dimensional Hilbert space with norm h0 xx = x2 for
x ∈ R ⊂ Q.
• If H is a Hilbert space then ext H has a unique natural Hilbert space
form extending H.
• Therefore Q =∼ Limo n→∞ (ext)n R has a unique natural Hilbert space
metrical form h extending h0 .

89
17.4 Glossary 2010.04.25

’ F ’ S = {F (s)|s ∈ S}; “the F ’s of S”.


a ≺− A a is a singular organized limit of A, or Limo A. Informal.
A− a a ≺− A.
β The Pauli adjoint for Dirac spinors, a bilinear spinor form.
Qx, x b Infraquantization of x, with x b− x.
1q , 1Q A basic monadic in Q, as in (40).
γw A basic Clifford generator in Cliff W.
2n The second exponential of n. 20 := 1, 2n := 2(2n−1 ) .
Limo Singular limit with organization.
Lin V The algebra of linear operators V → V .
4 L−R . The adjoint representation.

Dup V duplex space1 of V ; V ⊕ Dual V with quadratic form Q.


exp2 binary exponential2 : exp2 S := 2S = {polyS|(ιs)2 = 0}.
ext V exterior algebra of V ; ext = Q exp2 .
Fermi V the Fermi algebra of V , generated by {L v, ∂v | v ∈ V }.
Γ a gravitational constant ∝ h/G.
~ rationalized Planck action constant h/2π.
h(n) canonical Lie algebra of n coordinates, n momenta, and i.
ι successor,brace. ι q = {q} = q.
io Creation or annihilation or superposition thereof.
L ,R left, right exterior multiplication.
Lv 0 v generators of the defining representation of so(V ).
Limo o Limit with organization. See (9).
metasimplex Simplex of simplices of .... See (52).
monadic first-grade statvector in Q.
monad physical process represented by a monadic.
◦ φ ◦ ψ = value of dual vector φ on vector ψ.
operon quantum physical process represented by a linear operator.
p Pauli adjoint for the highest level of Q in use.
polyadic Q statvector of unspecified grade.
Q ext Q = b 2Q , the exterior algebra of simplex statvectors.
Qg Grade g of Q.
TQ Level T of Q; Q[0] = R.
Q Quadratic form Q(v ⊕ u) := u ◦ v, v ⊕ u ∈ Dup V .
SO(V ) The special orthogonal group of the quadratic space V .
SU(V ) The special unitary group of the Hermitian space V .
W[T ] Dup TQ1 . Lin TQ = Cliff W[T ].
1
The antidouble space of [31]
2
Also called power set, and indeed a power of 2, but not of S.

90
References
[1] Stephen L. Adler. Quaternionic quantum mechanics and quantum fields.
Oxford University Press, 1995.
[2] V. Ambarzumian and D. Iwanenko. Zur frage nach vermeidung des un-
endliche selbstrückwirkung des elektrons. Zeitschrift für Physik, 64:563–
567, 1930.
[3] J. Baugh. Regular Quantum Dynamics. PhD thesis, School of Physics,
Georgia Institute of Technology, 2004.
[4] F. Bayen, M. Flato, C. Fronsdal, A. Lichnerowicz, and D. Sternheimer.
Quantum mechanics as a deformation of classical mechanics. Letters in
Mathematical Physics, 1:521–530, 1977.
[5] Gregor W. Bayer. Theory of the cut-off. Symmetry, to appear, 2009.
[6] N. Bohr. Causality and complementarity. Philosophy of Science, 4:293–4,
1936.
[7] F. Bopp and R. Haag. Über die möglichkeit von spinmodellen. Zeitschrift
fr Naturforschung, 5a:644, 1950.
[8] R. Brauer and H. Weyl. Spinors in n dimensions. American Journal of
Mathematics, 57:425, 1935.
[9] E Cartan. Les groupes projectifs qui ne laissent invariante aucune multi-
plicit plane. Bull. Soc. Math. France, 41:53–96, 1913.
[10] E. Cartan. Leçons sur la theorie des spineurs. Hermann, Paris, 1938.
[11] A. H. Chamseddine, A. Connes, and M. Marcolli. Gravity and the stan-
dard model with neutrino mixing. Adv. Theor. Math. Phys., 11:991–1089,
2007.
[12] C. Chevalley. The construction and study of certain important algebras.
The Mathematical Society of Japan, Tokyo, 1955.
[13] P. A. M. Dirac. Spinors in Hilbert Space. Plenum, New York, 1974.
[14] A. Einstein and V. Bargmann. Bivector fields. Annals of Mathematics,
43:1–14, 1943.
[15] A. Einstein and V. Bargmann. Bivector fields II. Annals of Mathematics,
43:15–23, 1943.
[16] L. D. Faddeev. How we understand “quantization” a hundred years after
Max Planck. Physikalische Blätter, 52:689, 1996.
[17] R. P. Feynman, 1941. Personal communication ca 1961. Feynman began
this line of thought in about 1941, before his work on the Lamb shift, and
may have published a similar formula.

91
[18] D. Finkelstein. Space-time code. Physical Review, 184:1261–1271, 1969.
[19] D. Finkelstein, J.M. Jauch, and D. Speiser. Notes on quaternion quantum
mechanics I, II, III. Report CERN 59–7, 59–11, 59–17, CERN, Geneva,
1959.
[20] D. R. Finkelstein. Space–time structure in high energy interactions. In
T. Gudehus, G. Kaiser, and A. Perlmutter, editors, Coral Gables Confer-
ence on Fundamental Interactions at High Energy. Center of Theoretical
Studies January 22—24, 1969. University of Miami, pages 324–343, New
York, 1969. Gordon and Breach.
[21] D. R. Finkelstein. Quantum Relativity. Springer, Heidelberg, 1996.
[22] A. Galiautdinov and D. R. Finkelstein. Chronon corrections to the Dirac
equation. JMP, 43:4741, 2002.
[23] M. Gerstenhaber. On the deformation of rings and algebras. Annals of
Mathematics, 32:472, 1964.
[24] D. Hestenes. Space-time algebra. Gordon and Breach, New York, 1966.
[25] E. Inönü and E.P. Wigner. On the contraction of groups and their repre-
sentations. Proceedings of the National Academy of Science, 39:510–524,
1953.
[26] B. Kostant. Lie algebra cohomology and the generalized Borel–Weil the-
orem. Annals of Mathematics, 74:329–387, 1961.
[27] A. Kuzmich, N. P. Bigelow, and L. Mandel. Europhysics Letters, A 42:481,
1998.
[28] Dennis Marks, 2008. Private communication.
[29] T. D. Palev. Lie algebraical aspects of the quantum statistics. Unitary
quantization (A-quantization). Preprint JINR E17-10550, Joint Institute
for Nuclear Research, Dubna, 1977. hep-th/9705032.
[30] R. Penrose. Angular momentum: an approach to combinatorial space-
time. In T. Bastin, editor, Quantum Theory and Beyond, pages 151–
180. Cambridge University Press, Cambridge, 1971. Also available at
http://math.ucr.edu/home/baez/penrose/. Penrose kindly shared some
of this seminal work with me as his theory of ‘mops’, in 1960.
[31] H. Saller. Operational Quantum Theory I. Nonrelativistic Structures.
Springer, New York, 2006.
[32] I. Schur. Über die Darstellung der symmetrischen und der alternierenden
Gruppen durch gebrochene lineare substitutionen. Journal für die reine
und angewandte Mathematik, 139:155–250, 1911. Translation: On the

92
representation of the symmetric and alternating groups by fractional lin-
ear substitutions. Transl. M-F Otto. International Journal of Theoretical
Physics 40:413–458 (2001).
[33] I. E. Segal. A class of operator algebras which are determined by groups.
Duke Mathematical Journal, 18:221–265, 1951. Especially §6A.
[34] M. Shiri-Garakani and D. R. Finkelstein. Finite quantum kinematics of
the harmonic oscillator. Journal of Mathematical Physics, 2006.
[35] H.P. Snyder. Quantized space-time. Physical Review, 71:38, 1947.
[36] H. P. Stapp. The copenhagen interpretation. American Journal of
Physics, 40:1098–1116, 1972.
[37] E. C. G. Stückelberg. Quantum theory in real Hilbert space. Helvetica
Physica Acta, 33:727–752, 1960.
[38] L. Susskind. Dynamics of spontaneous symmetry breaking in the
weinberg–salam theory. Physical Review D, 20:2619, 1979.
[39] M. Tavel, D. Finkelstein, and S. Schiminovich. Weak and electromagnetic
interactions in quaternion quantum mechanics. Bulletin of the American
Physical Society, 9:435, 1965.
[40] Armin Uhlmann. State vector and quantization in an over-all space-time
view. Wissenschaftliche Zeitschrift der Karl-Marx-Universität Leipzig,
Mathematische-Naturwissenschaftliche Reihe, 2/3:115–123, 1957/58.
[41] H. Umezawa. Advanced field theory: micro, macro, and thermal physics.
American Institute of Physics, 1993.
[42] R. M. UNGER. THE SELF AWAKENED: PRAGMATISM UNBOUND.
Harvard, 2007.
[43] R. Vilela-Mendes. Deformations, stable theories and fundamental con-
stants. Journal of Physics A, 27:8091–8104, 1994.
[44] John von Neumann. Mathematische Begrundung der Quantenmechanik.
Nachr. Ges. Wiss. Göttingen, pages 1–57, 1927. Also in J. von Neumann,
Collected Works, volume 1.
[45] S Weinberg. Implications of dynamical symmetry breaking: An adden-
dum. Physical Review D, 19:1277, 1979.
[46] C. F. v. Weizsäcker, E. Scheibe, and G. Sussmann. Komplementarität und
Logik III. Mehrfache Quantelung. Naturwissenschaften, 13a:705, 1955.
[47] J. Wess. Gauge theories on noncommutative spacetime treated by the
Seiberg-Witten method. In U. Carow-Watamura, Y. Maeda, and S. Wata-
mura, editors, Quantum Field Theory and Noncommutative Geometry,
page 177. Springer, Berlin, 2005.

93
[48] C. N. Yang. On quantized spacetime. Physical Review, 72:874, 1947.

94

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy