You are here

Appendix F

Octan Physics

 
[Fleegello originally wrote his critique of octan physics as section 2.3 of the Principia. It is listed separately here, as it frankly stands apart, and interferes with the general flow of that work. Most scholars agree that Fleegello did not introduce any novel physics in this section. Rather he took ideas already advanced by contemporary physicists, and adapted them to his own philosophical framework. Editorial comments are enclosed in brackets throughout the text, to distinguish them from primary content.]

The physical (pan)universe has been identified as the Physical Consistency Subfield (PCS) of the Consistency Ideo Field (CIF) – the complete set of abstract mathematical objects that both define physical (temporospatial) relationships and are compatible with consistency logic. The PCS may itself be naturally divided into multiple, logically self-contained physical universes, characterized by distinct physical laws and conditions. Our own universe would be one of these worlds.

Mathematical objects in the PCS may manifest (in part) as states of a physical system, or as operators representing observables or other entities that act on those states. If modern physics is a guide, they incorporate a broad class of multidimensional objects known as dimensors [including common scalars, vectors, more general tensors, and spinors]. Many useful dimensor operators (in particular, those representing observables) define linear relations between other dimensors.

In standard Shrodiik [quantum] theory [named in honor of the pre-Dracian physicist Shrodo], a physical state (of any sufficiently isolated system) is denoted by a ket symbol |ψ>, where ψ is an arbitrary label. This entity is supposed to encompass all physical aspects of a system. |ψ> was originally interpreted in terms of the positions and motions of material particles at a time t in a pre-existing three-dimensional (3D) space x [here bold type indicates a physical 3D vector]. Observables then include the positions, energies and momenta of these particles.

Contemporary Shrodiik physics has a most peculiar feature. For any physical state |ψ>, only the probabilities for measuring different values of a given observable can be computed. Even granted complete knowledge of a physical system at a particular moment, the future course as seen by any octan observer cannot in general be predicted with certainty. Detailed prescriptions for computing probabilities may be found in Shrodiik physics texts.

For any physical system, there is a range of possible states. Related to its probabilistic character, |ψ> can consist of a linear combination, or superposition, of these available states. The selection of a set of fundamental basis states is then arbitrary, to some extent; any given set of basis states can be mixed into new combinations, to form distinct sets.

In general, |ψ> can be viewed as a vector in an abstract space that spans all the possible states. If the components of a state vector are defined with respect to a specified set of basis vectors, then the state may be represented by a single-column dimensor array. Linear operators may in turn be represented by square dimensor arrays that transform any given state (by the rules of matrix multiplication) into another state.

Let  represent a (linear) operator corresponding to an observable A. Here the symbol "hat" explicitly denotes operator, versus numeric parameter, status. When  is applied to an arbitrary state vector, the result is typically a linear combination of other state vectors. Suppose, however, that  is applied to a state vector |ψa> characterized by a well-defined value a of A – i.e., a state in which a measurement of A will definitely yield the value a. Then  acts on |ψa> by extracting this value:

      Â|ψa> = a|ψa> .

This is what it means for  to represent an observable. Mathematically, |ψa> is an eigenstate of Â, with a well-defined value a of that observable. From the perspective of a given observer, performing a measurement (observation) reduces a system to an eigenstate of the observed quantity, corresponding to the measured value. The observer must then interact with the system, in a way that selects one of the available eigenstates

In bra-ket notation [physicists used this archaic script during Fleegello's era], the numeric overlap between two states |φ> and |ψ> is represented by <φ|ψ>, where the states are normalized such that <ψ|ψ>=1 for all |ψ>. The overlap value is a probability amplitude for starting with a system in state |ψ>, but observing it in state |φ>. The actual probability is the absolute square of this amplitude, or |<φ|ψ>|2. For example, consider a one-particle system. If |x> is the state with the particle at 3D position x, then the overlap <x|ψ> is the single-particle wavefunction ψ(t,x) of Shrodiik mechanics, and |ψ(t,x)|2 is the probability per unit volume at time t for finding the particle at x.

The expectation value of an observable A, defined as the average value Aavg over repeated measurements on identical states |ψ>, is given by

      Aavg = <ψ|Â|ψ> .

The observable A has a definite value a only if |ψ> is already an eigenstate |ψa> of Â.

In classical physics, material particles were treated as localized entities, distinct from waves (such as light) that propogate through underlying fields or media. Only waves could undergo self-interference, or diffract around obstacles. At the dawn of the Shrodiik revolution, ostensible particles were found to have wave properties, and nominal waves were found to sometimes act like classical particles. Observables that classically had a continuous range of values – e.g., the energy of an electron in an atom – might now be quantized, or restricted to discrete values.

Energy is one of the central observables in quantum physics. It is associated with the Hoobitean operator Ĥ [named for the classical physicist Hoobitu]. For material particles, Ĥ is often written as the sum of a kinetic energy term Ĥo plus a potential energy (interaction) term Ĥint. After wave-particle duality was discovered, the energy E of a (set of) particle(s) in an eigenstate of Ĥ became associated with a temporal frequency f, or angular frequency ω:

      E = f = ω / 2𝜋 = ω ,

where is the minuscule but nonzero Planko constant [named for the pioneering physicist Planko], and is the reduced Planko constant. Conversely, experiments showed that the energy in a traditional wave of frequency f was not continuously distributed over the wave, but carried by discrete quanta with individual energies given by the same equation. Ĥ embodies the way a state changes in time. When Ĥ acts on a state vector |ψ>, the result is the constant (i) multiplied by the time rate of change of |ψ>, where i is the imaginary unit (square root of -1). It is remarkable how imaginary (or complex) quantities arise naturally in the equations of Shrodiik physics!

A 3D vector quantity closely related to energy is linear momentum, represented by the symbol p. Much as energy is associated with temporal frequency, momentum px along a spatial axis x is associated with a spatial frequency equivalent to an inverse wavelength λx or angular wavenumber kx:

      px = / λx = kx .

Shorter wavelength (and so either larger momentum, or a smaller value of ) generally begets more particle-like behavior. When the operator p̂x acts on a state vector |ψ>, the result is the constant (-i) multiplied by the spatial rate of change of |ψ> along x.

For a single material particle, the relationship between time and energy is thus analogous to that between spatial position and linear momentum. For a multiparticle system, however, the situation is more nuanced. Whereas every particle may be assigned its own dynamical position and momentum operators, all particles traditionally share a common time. Time is then treated as a numeric system parameter, and not associated with a true operator.

Using calculus, it can be shown that the (unnormalized) time- and space-dependent wavefunction ψ(t,x) of a particle (or any quantum) with pure angular frequency ω and wavenumber kx (energy E and momentum px) has the exponential, wave-like form

      ψ(t,x) = e-iωt eikxx = cos( kxx-ωt ) + i sin( kxx-ωt ) .

where e is the Eulero number of mathematics, and the "cos" and "sin" terms refer to standard trigonometric functions. More generally, for a 3D wavevector k (momentm p) and position x,

      ψ(t,x) = e-iωt eik·x

where the scalar product  k·x is the scalar length of x multiplied by the projection of k onto x. Note that the relative probability at any moment for finding the particle at a given 3D position (equal to the absolute-square of the wavefunction) is the same for all x values.

With quantum physics, it is found that the order in which observables are measured may be significant. Consider then two observables A and B, represented by operators  and B̂. The observables/operators commute if the order of measurement is irrelevant, or equivalently if the order in which  and B̂ act on an arbitrary physical state is irrelevant – i.e., if ÂB̂ = B̂Â. They do not commute if ÂB̂B̂Â. In this case the very act of measuring  or B̂ introduces uncertainty into the value of the other, complementary observable. A system cannot simultaneously be an eigenstate of two operators that do not commute – neither operator would alter such a state, so their order could not matter. It is then impossible to simultaneously measure the values of two non-commuting observables, since such a measurement must create an eigenstate of both.

In Shrodiik mechanics, the archetypal pair of observables with non-commuting operators are the position x and linear momentum px of a material particle along a given spatial direction. Classically, these quantities commute, and every particle simultaneously has well-defined position and linear momentum. Based on their operator interpretations, the quantum commutation relation between x̂ and p̂x is

      x̂ p̂x - p̂x x̂ = i .

Because the operators do not commute, an eigenstate of p̂x must span a range of x values. As described earlier, an eigenstate of p̂x is indeed totally unlocalized in space. Conversely, it can be shown that an eigenstate of x̂ must include all possible linear momenta px. More generally, if σx is the [standard deviation] uncertainty in x, and σpx the uncertainty in px, it can be shown that

      σx σpx/2 .

Consider then a system |ψ> = |xo>, in which a particle initially has a definite position xo. Suppose an observer measures first the position x of the particle, and then its momentum px. Because the system starts in a state of well-defined position, the particle will be found at xo with 100% probability, and the wavefunction is unchanged. Because this wavefunction contains all possible momentum values, any value of momentum may be observed in the subsequent measurement, with equal probability. Now suppose the order of measurement is reversed – the observer measures momentum first, followed by position. The likelihood of initially observing any value of momentum px is the same as before. But the momentum measurement forces the particle into a state of well-defined momentum. The particle's position is thereby scrambled, and the observer may subsequently find the particle at any location!

Consider now a more general system in which one member of any pair of non-commuting observables is well defined. Mathematically, the system can be considered a superposition of pure eigenstates with different but well-defined values of the other non-commuting quantity. The existence of non-commuting observables is contrary to classical (pre-Shrodiik) physics. The natural law that describes physical evolution applies to superpositions of pure states, rather than to states in which all classical variables have precise values.

The state |ψ> of a physical system can in general be written as a coherent sum

      |ψ> = C1|φ1> + C2|φ2> + C3|φ3> + . . . = Σj Cj|φj>

over a complete set of orthonormal states |φj>, where the Cj are (complex) constants.

The |φj> are orthonormal if <φj|φk>=0 for all jk, and <φj|φj>=1 for all j.

The choice of the |φj> is arbitrary to some extent, but they must be eigenstates of a complete set of commuting observables that cover all physical aspects of the system.

The probability of starting with the system in state |ψ> but finding it in a state |φj> is then Cj*Cj, where Cj* is the complex conjugate of Cj. The expectation value of an observable A is

      Aavg = <ψ|Â|ψ> = ΣjΣkCj*CkAjk   where Ajk = <φj|Â|φk> .

The off-diagonal terms jk in the sum represent nonclassical interference between the different states in the coherent superposition comprising |ψ>. These terms in general vanish only if the |φj> have well-defined values of A (i.e., are eigenstates of Â).

The physical interpretation of the state vector |ψ> has a long and tortuous history. Originally it was viewed merely as a device for computing the probability of observing a given outcome in an experiment. Reality was seen to reside in the observed positions and momenta of individual particles. The physical universe was assumed to evolve in a linear manner, with a single unfolding history, which was deterministic in only a limited, probabilistic sense. The act of observation was divorced from the natural evolution of a physical system, and treated as something special, even magical.

Yet consistency logic requires that the universe be totally deterministic. Recently, the contradictions inherent in the original interpretation of Shrodiik mechanics have led the Evette group to develop an alternative multi-world view, in which reality resides in |ψ> itself. Observers, measuring devices and related processes are now included as integral parts of |ψ>. The physical panuniversal |ψ> is viewed as a superposition of conventional quantum worlds, each represented by a restricted state vector, which have become decohered and mutually orthonormal [or minimally overlapping]. These worlds evolve [almost] independently of each other, and continually split [rarely merge] into new separate, decohered worlds through time. A given observer occupies one conventional world at a given instant. As this world subsequently splits, the observer likewise branches into multiple selves, each with a distinct future experience. An observer does not see physical evolution as completely deterministic simply because no individual mind encompasses all worlds of the unfolding panuniversal state.

[Unfortunately, only the barest references to the original Evette School survive in the historical record. The writings may have been systematically destroyed by conservative, fundamentalist religious sects that flourished at the time, and found the work heretical. These traditionalist factions believed the universe progressed in a linear fashion along a single preordained path, in accordance with a divine plan for the octan race. The random, branching character of the multi-world view demanded an even greater, and to many more threatening, decentering to the octan psyche than the recognition five octujopes earlier that Jopitar was not at the physical center of the universe, but was a nonsingular ball of ordinary matter orbiting a commonplace star adrift on a vast ocean of space and time.]

Let |ϕ0> represent a conventional world in the Evette sense. Then the matrix elements A0j between this world and any other coexisting conventional world |ϕj> must be [essentially] zero for all observables A, including noncommuting quantities. Such conventional worlds evolve independently, with no [minimal] mutual interference.

Suppose that |ϕ0> incorporates a subsystem consisting of a simple superposition ( |B1> + |B2>) of orthonormal eigenstates of an observable B. Then

      |ϕ0> = ( |B1> + |B2> ) |Ɛ>

where |Ɛ> represents the environment of the subsystem. The environment may interact with the subsystem, so as to become correlated with its eigenstates. This happens in particular when |Ɛ> includes an observer who measures the value of B. If B̂ commutes with the interaction Hoobitean Ĥint, the eigenstates of B̂ are not changed by the interaction, and

      |ϕ0> ➜ |B1> |Ɛ1> + |B2> |Ɛ2> = |ϕ1> + |ϕ2>

where |Ɛ1> and |Ɛ2> are themselves eigenstates of observables that commute with Ĥint.

Consider now the matrix elements A12 and A21 with |ϕ1> and |ϕ2> for an arbitrary observable A. If  commutes with B̂, then A12 = A21 = 0, since the eigenstates of B̂ are orthonormal. If  does not commute with B̂, then it acts only on the B subsystem (if  were a product of operators that separately act on the subsystem and its environment, then A would not be a valid observable). In this case  does not affect |Ɛ>, and A12=A21=0 if |Ɛ1> and |Ɛ2> are orthonormal. The states |ϕ1> and |ϕ2> can thus be identified as two new conventional worlds, split off from the original |ϕ0>, if only |Ɛ1> and |Ɛ2> are orthonormal.

[This line of reasoning, which did not originate with Fleegello, helped resolve a problem with the many-world interpretation, involving an apparent ambiguity in the identification of the individual worlds. Some researchers argued that the choice of states |B1> and |B2> in the given example was quite arbitrary. By choosing a rotated basis set, e.g.

      |b1> = (|B1> + |B2>)/√2   and   |b2> = (|B1> − |B2>)/√2 ,

the state |ϕ0> appeared to split into a different set of conventional worlds. Eventually it was realized that the interaction between a system and its environment naturally selects a particular (compatible) basis set. If the operator B̂ does not commute with Ĥint, then |B1>|E> does not evolve into |B1>|E1>, since |B1> is itself transformed by the interaction.]

Conventional worlds can thus be distinguished by non-interfering "memories" of prior branchings. The storage sites of these data may include, but are by no means limited to, animal brains (and recently, scientific apparatus acting as extensions of those brains). The physical structure of a brain determines its interactions with the environment, and thus the types of conventional worlds (i.e., which observables are relevant and well-defined) generated by the observation process. If a brain is so constructed that only one value of a particular observable can communicate with (affect) other elements in a conscious field, then a state including a coherent superposition of different values of that observable at the same moment must correspond to distinct unified ideo fields, or selves, in separate (conventional) worlds. The information stored in a brain does not define the external reality of the associated world – a person may make faulty observations – but it may still be a point of reference by which that world is distinguished from others. Two distinct conventional worlds can even merge, if their distinguishing memories are lost or corrupted so as to become identical. Observers inhabiting the worlds would experience no sense of merger, as all valid memories of a former distinct past would be absent.

What determines useful observables, other than position? The mathematician Noethra has linked many such quantities to symmetries in the equations of motion that describe the temporal evolution of |ψ>. Noethra's first theorem states that for every continuous, differentiable coordinate transformation that does not alter these equations, there is a corresponding observable whose expectation value is conserved, or constant over time. For sufficiently isolated (closed) systems, the equations are in fact generally unaffected by several such transformations, including time displacement, spatial displacement, and spatial rotation. Each of these symmetries is associated with an observable and conserved quantity.

Why are the dynamical equations unaffected by the given transformations? Although physical conditions clearly vary at different locations in time-space, there is nothing else to distinguish points or directions. From an ideobasic perspective, physical law for a sufficiently closed system (which incorporates all relevant causal agents) should then depend only on extant physical conditions. Although distinct physical laws may apply in different physical universes, the same law and dependencies should apply at all times, positions and orientations within a given universe. This leads to the observed symmetries.

When the equations of motion are not affected by displacements in time (i.e., they remain the same over time), then what is commonly called energy is conserved. This is primarily what makes energy a useful observable. Note that only the laws of motion are unchanging; physical conditions and entire systems may change dramatically over time. When the equations of motion are not affected by displacements in spatial position (i.e., physical law is the same at different spatial points), then linear momentum is conserved, and is a useful observable. When the equations of motion are not affected by spatial rotations (orientation in space), then angular momentum is conserved, and useful. It can be shown more generally that the expectation value of any operator that both commutes with the Hoobitean operator Ĥ, and is not explicitly a function of time, is also a constant of motion. Every symmetry in Ĥ is thus associated with a conserved quantity, and a corresponding observable.

Classical observables may have nonclassical analogs that result from a reinterpretation (typically involving commutation relations) of associated operators. In particular, the commutation relations among the three orthogonal (mutually perpendicular) angular momentum operators imply the existence of a nonclassical type of angular momentum, known as spin. Elementary particles are found to inherently possess this type of angular momentum. Particle spin is naturally quantized to discrete values, characterized by a spin number s, which must be an integral multiple of 1/2. Overall spin angular momentum is s(s+1), while the maximum possible component in any 3D direction is s.

Spin angular momentum operators can be represented by irreducible (2s+1) x (2s +1) arrays. The spin aspect of a spin-s particle can then be represented by a (2s +1)-dimensional single-column dimensor known as a pointor, designated by Š. An overall single-particle state may in turn be represented by a pointor wavefunction Š(t,x) of time t and 3D position x.

Spinless (s=0) particles are represented by simple scalar (zero-rank dimensor) functions, with no inherent directionality. [No elementary spin-0 particles were known in Fleegello's era, although composite spin-0 particles (e.g., pions) were certainly recognized.] Spin-½ particles are represented by special two-dimensional pointors known as spinors. Spinors do not transform like geometric vectors under coordinate transformations. Spin-1 particles with mass are represented by three-dimensional pointors, which do transform like geometric vectors. [Because massless spin-1 particles (in particular photons) have no rest frame but are constrained to move at light speed, they must be represented by two-component pointors.] Particles with even larger spin values are represented by distinct pointor classes.

Yet particles do not normally exist in isolation. How then can the state of a multiparticle system be represented? Suppose first that the particles are distinguishable, and motions are much slower than light speed. Such systems have traditionally been represented by a direct product of the pointor functions for the individual particles, in which time t is a common system parameter, but the coordinates xj of the various particles j are distinguished. For example, the state of a two-particle system might be represented by

      Ša(t,x1) Šb(t,x2) ,

where subscripts a and b label two different single-particle states.

Suppose now that two particles in a system are identical. The probability of finding either cannot be affected when their labels are exchanged – they would otherwise be distinguishable. Because the probability is equal to the absolute square of the wavefunction, and the associated exchange operator must (as an observable) be linear, the state can at most acquire a complex phase factor (absolute value one) under particle exchange. Since two successive exchanges must leave the state unchanged, the phase factor is limited to the values ±1. A state must then be either symmetric (unchanged) or antisymmetric (phase factor -1) under identical particle exchange.

The wavefunctions of identical bosons (particles with integral spin) are found to be symmetric, while those of identical fermions (particles with half-integral spin) are antisymmetric. The appropriate symmetry can be achieved if a system is represented by a sum over the direct pointor products, in which the functional dependencies of the particles are suitably interchanged. For example, the state of two identical fermions might be represented by

      Ša(t,x1) Šb(t,x2)-Ša(t,x2) Šb(t,x1) .

Symmetries in the equations of motion are not limited to continuous time-space transformations, but may also include discrete operations, such as time reversal and parity inversion (mirror reversal). [Fleegello stubbornly maintained that various discrete spacetime symmetries should generally hold, despite contrary evidence. For example, experiments seemed to demonstrate that parity is not conserved during certain types of radioactive decay. Parity is conserved if the equations of motion are unchanged when a system is replaced by its mirror image. Fleegello believed that physics could not be affected by such a simple transformation, and felt that crucial elements had been omitted from experimental analyses. Yet physicists soon realized that, since time and space are intimately linked, and antiparticles are equivalent to ordinary particles moving backward in time, the true symmetry involves the CPT transformation – a combination of particle-antiparticle charge exchange, parity inversion, and time reversal – and not any one of these operations in isolation.] Internal symmetries, that do not transform time-space points, can give rise to additional conserved quantities and observables (e.g., electric charge).

Indeed, the fundamental interactions between elementary particles are thought to derive from a variety of internal local gauge symmetries. For example, consider the electromagnetic interaction. Under a local phase transformation, the single-particle wavefunction ψ(t,x) is multiplied by a phase factor e(t,x), where λ(t,x) is an arbitrary function of time-space. The absolute square (probability density) of ψ is unchanged by this transformation. If local gauge symmetry holds, then the new wavefunction must also satisfy the standard equation of motion. The kinetic energy part of that equation generally contains terms involving both the time- and space-rate of change of ψ, so the phase factor in the transformed wavefunction generates new quantities. The equation is invariant under the transformation only if it also contains terms that transform so as to cancel the effect of the (t,x) dependence in λ, while maintaining the original form of the equation. These terms can be identified with the electromagnetic scalar and vector potentials.

The physicist Vigno has argued that symmetries do not merely restrict the laws of physics, but further define much of physical reality. While fundamental forces have been related to symmetries in the equations of motion, elementary particles have themselves been associated with (irreducible) mathematical representations of abstract symmetry groups. Every consistent object and process must coexist with every other consistent object and process somewhere within the PCS. This may involve a natural segregation into distinct, self-contained physical universes.

Coordinate systems do not exist a priori in nature. The choice of a coordinate framework to describe a physical system should thus be arbitrary, from a strictly mathematical viewpoint (although one frame may be more convenient than another for a given purpose). It should then be possible to describe the laws of physics in a coordinate-free manner, in which observables appear only as abstract quantities, with no explicit reference to coordinate components. Expressing physical laws in such a covariant manner simplifies identification of symmetries and conserved quantities.

If the PCS is to respect the inherent arbitrariness in the choice of coordinate system, then fundamental physical constants that appear in the laws of physics should also be the same for all observers within a given physical universe, independent of the choice of reference frame. This applies in particular to dimensionless constants (e.g., the fine structure constant of atomic physics), which carry no physical units, but can be expressed as the ratios or products of dimensional constants that do possess units. Changes in the values of dimensional constants are generally meaningful only with respect to changes in their dimensionless combinations. So long as the values of physical constants are individually changed in a way that maintains the values of all fundamental dimensionless constants, the physical world is unaffected. Dimensionless constants stand independent of any arbitrary choice of measurement units. Indeed, no variations over time or space have thus far been detected.

[Some quantities thought to be fundamental constants in Fleegello's era have since been found to be variable. These have been reinterpreted as functions of truly fundamental constants and local physical conditions.]

Dimensionless fundamental constants need only be the same at all points within a particular physical universe. The values in distinct, non-interacting universes may be different. If there is no fundamental reason a constant should have a particular value, then the PCS must encompass a host of universes covering the range of acceptable values. Yet these values must be countable (either discrete/quantized, or at least represented by rational numbers). All the worlds otherwise could not have meaningful existence within the PCS.

Even fundamental dimensional constants (whose numeric values depend on the choice of physical units) should be the same for all observers in a given universe, when measured with respect to reproducible units characteristic of fundamental physical processes. In particular, the speed of light in a vacuum, commonly denoted by the symbol c, appears to constitute a universal limit to the rate at which information can propagate through space. As first proposed by the physicist Niestu in his theory of inertial invariance, the speed c has the same value for all observers, irrespective of their state of motion. This is contrary to classical expectations, whereby an observer moving toward (away from) a light source detects a higher (lower) relative light speed than an observer at rest with respect to the source. That c is finite may be expected from an ideobasic viewpoint. An infinite speed is a special, limiting case of a general value, and the PCS should opt for the most general conception.

Niestu introduced a major paradigm shift in physics when he showed that a common value for c implies that time (space) intervals measured by one observer may be partially seen as space (time) intervals by an observer in a relative state of motion; time and space do not exist separately, but must be combined into a unified timespace [scientists of Fleegello's era apparently preferred this expression to today's more common term spacetime]. The effect is tiny at low velocities, but becomes significant as speed approaches c (so-called Niestiik speeds). The associated coordinate transformation between reference frames in a relative state of motion is distinct from that of classical physics. If the equations of motion are to remain invariant under a velocity transformation, then those equations must be modified as well. A remarkable consequence of inertial invariance is that any mass m is associated with an energy mc2. For a free particle, the relationship between total energy E, momentum p, and rest mass m becomes

      E2 = p2c2 + m2c4 .

Niestu ultimately expanded his ideas into the theory of general invariance, which describes gravity in terms of distortions in the geometry of timespace.

[Fleegello overlooked a related serious inconsistency in his view of the CIF. The CIF must encompass all possible reference frames. If It experiences the same time as observers in those frames, as Fleegello envisioned, It must integrate the various time lines to maintain a single unified state of being. Yet if speed c is the same for all observers, events that are simultaneous in one frame may be nonsimultaneous in another. Events could then be seen by the CIF as both simultaneous and not simultaneous, a contradiction. This inconsistency is resolved only if the CIF transcends physical time, and experiences it the way corporeal creatures experience space – as block time. All events in the physical panuniverse then span a single, eternal moment in the mind of the CIF. Yet the CIF must still distinguish the time-like and space-like separations among physical events that define causal chains. Primacy resides in these causal chains, and not in the reference frames that observers use to describe them.]

While inertial invariance was readily incorporated into Shrodiik mechanics for single particles, problems arose for multi-particle systems. In particular, time and space coordinates were not treated coequally in the traditional equations of motion. Inertial invariance requires that time and position both be treated either as system parameters, or as formal operators. Currently the most widely adopted solution, based on the first approach, is to reformulate Shrodiik mechanics into a Niestiik quantum field theory (QFT), in which elementary particles of a given type are treated as quantum excitations of an underlying field. QFT covers both traditional particles with mass, like the electron, and zero-mass particles once considered pure waves, like the photon. Different particle types are represented by distinct fields, defined by a variety of attributes including rest mass, spin, and electric charge. For each field type q, a position field operator Φ̂q(t,x) and conjugate momentum field operator 𝜋̂q(t,x), now functions of common timespace parameters (t,x), replace the single-particle 3D position and momentum operators x̂j and p̂j for the discrete particles j of Shrodiik mechanics. Because all fields share one time parameter, QFT (like Shrodiik mechanics) is a single-time theory (STT).

A simple field state in QFT is characterized by the number of (identical) quanta occupying each of a set of allowed levels. The number of quanta is just the number of "particles" of the given type. Field quanta contain no explicit particle labels; QFT respects the exchange symmetry of identical particles in a remarkably natural way. Indeed, in QFT it can be shown that fields with half-integral spin must be antisymmetric, and those with integral spin symmetric [dictated by the distinct Niestiik equations of motion for fermions versus bosons]. Many physicists prefer not to speak of particles at all in QFT, but only quanta. A general field state can be represented by a superposition of simple states. Unlike in Shrodiik mechanics, this is not limited to states with a fixed number of particles; interactions between fields result in the routine creation/destruction of quanta. The overall state of a system is represented by the direct product of its constituent fields, or more generally by a superposition of such products.

Shortly after QFT was introduced, the mathematician Draci proposed a multi-time theory (MTT) alternative, in which both the (observer-based) times and positions (tj, xj) of various particles j are now distinguished, and associated with coequal system operators (t̂j, x̂j). The earliest versions, identified here as MTTa, were simple extensions of Shrodiik STT. For a system of fixed N particles, and an observer in an inertial reference frame (t, x), the associated multi-time wavefunction has the form Ψmta(t1, x1; t2, x2; ...; tN, xN ). This reduces to the single-time (ST) wavefunction of Shrodiik mechanics if all the tj are set to a common time t. Just as the ST wavefunction was defined over a constant-time surface, Ψmta is only defined over appropriate space-like hypersurfaces. As in Shrodiik STT, |Ψmta|2 is a probability per unit 3N-dimensional (3ND) spatial volume. Normalization now requires that the sum [integral] of |Ψmta|2 over all 3N spatial coordinates on a specified hypersurface equals unity.

A later version of MTT was a more radical departure from single-time theory, but more attuned to inertial invariance and the vagaries of measurement, and is adopted here. Every particle j in MTT has an innate proper time dimension, measured along the particle's world line. Different time lines are not inherently synchronized, but correlated only through interactions. To locate at time to the position xj of particle j, an observer must interact with the particle at some particle proper time τj. Due to a synchronization ambiguity [Fleegello elucidates this point later], τj is not in general well defined for any given to and xj. There must then be a probability distribution over a range of τj. Since there should be at least a probabilistic Niestiik-like transformation relating timespace coordinates (τj,0) in the particle rest frame to (tj, xj) in the observer frame, there is a corresponding distribution for tj, centered on but not limited to to.

The wavefunction must then separately incorporate the reference observer time to at which the (tj, xj) are measured, and feature a probability distribution over each tj for a given (to, xj). Inertial invariance requires an additional spatial coordinate xo, paired with the temporal coordinate to. A physical interpretation of xo will be presented later. However, if the observer is a massive, complex entity compared with the observed particles, then (to,xo) should represent some kind of composite timespace self-coordinates, and it should be possible to localize xo. The functional dependence of the wavefunction is now Ψmt(to, xo; t1, x1; t2, x2; ...; tN, xN ).

The given Ψmt is a joint probability amplitude that, when the observer finds itself at time to, it somehow sees itself at xo, and each particle j at tj (corresponding to τj) and xj. |Ψmt|2 is now the probability per (invariant) 4ND timespace (tj,xj) volume and 3D spatial (xo) volume. The sum [integral] of |Ψmt|2 over all these volumes must equal unity. As in STT, Ψmt is defined over spacelike surfaces of constant to. However, Ψmt now involves invariant timespace volumes, and incorporates an observer state. Ψmt does not reduce to the ST wavefunction when all tj are set to a common to. Times to and τj are not assumed to be well-correlated, though for non-Niestiik speeds it may be possible to set the average value of each tj equal to to. The more to and τj are correlated by interactions, the narrower the probability distribution over tj.

Note that MTT does not posit that any material particle evolves along multiple time dimensions. Each particle j evolves along a single proper time line τj. The corresponding tj are observer-based, and measured with respect to one observer time line. Yet from the adopted multi-time (MT) perspective, a multiparticle system can be accurately described using a single time coordinate only if to and the tj are sufficiently correlated.

If there are no particle interactions, Ψmt should separately satisfy the free-particle equations of motion for each particle j, as well as for the observer. If there are also no observer-particle interactions, then there is no way the observer can detect the particles, define observer-particle distances, or establish the relation between to and the tj. Then Ψmt can only be a function of to and proper times τj. In this case, Ψmt can be constructed from products of independent, free energy eigenstates e-iωoto and e-iωjτj, where ωo and ωj are the angular frequencies associated with rest masses mo and mj of the observer and particle j, respectively.

If there are minimal observer-particle interactions but still no interparticle forces, Ψmt can be constructed from products of energy-momentum eigenstates e-iωoto+iko·xo and e-iωjtj+ikj·xj, where ko and kj are wavevectors corresponding to the (observer-based) momenta of the observer and particle j. Now ωo and ωj are frequencies associated with the overall energies of the observer and particle j. The observer has the same chance of seeing any tj at a given to.

With interparticle forces, interaction terms must be introduced into the equations of motion, resulting in N+1 coupled equations of motion for an N-particle plus observer system. Draci has shown that, for any MTT, the respective interaction terms must satisfy a consistency condition, or the evolution of Ψmt along the various tj is not well-defined. Finding consistent MT equations with interactions is a challenge. Most notably, conventional interaction potentials always lead to inconsistency. Interactions must then be implemented by other means, in particular through the use of creation and destruction operators (defined to maintain appropriate symmetry under identical particle exchange), as in QFT. While the simplest forms of MTT are restricted to fixed N, use of these operators allows N to change.

It is often useful to adopt composite MT timespace coordinates to replace the (tj,xj). For an observer plus two particles (j=1 or 2), define

      X = (a1x1 + a2x2)  and  r12 = (x2 - x1) , where a1 and a2 are constants.

Center-of-mass (COM) coordinates with aj = mj/(m1 + m2) are routinely used in non-Niestiik analyses of two-body systems. Remarkably, analogous composite time coordinates may be defined in the MT approach, by

      T = a1t1 + a2t2  and  ρ12 = (t2 - t1) .

The (to,xo; t1,x1; t2,x2) are now replaced by (to,xo; T,X; ρ12,r12), where T serves as an overall composite system time for the particles. At Niestiik speeds, center of momentum (also COM) coordinates are appropriate, with aj = ωj/(ω1 + ω2) for free particles. If the particles interact, individual energies Ej = ωj ,are no longer constant, but a constant-velocity COM can still be defined.

Recall that (to,xo) is already inherently composite, if the observer (including measurement devices) is a macroscopic entity composed of myriad material particles. If the observer reference frame is at rest with respect to the observer spatial COM, and internal time lines are sufficiently synchronized, then (to,xo) may be identified as composite 4D timespace COM observer self-coordinates (where some uncertainty in xo may be expected around xo=0). While this (to,xo) could in principle be included in the definition of (T,X), the observer would then become part of the system being observed, confusing its status. The observed complex would also become a three-body system, requiring a new set of 4D displacement coordinates.

The observer is nonetheless integrally connected to the observed two-particle system. Interactions between the observer and the two particles define the COM motion of that system. If the two observed particles are otherwise isolated, then ideally interactions between these particles directly determine the primary dependence of Ψmt on both ρ12 and r12, but interactions with the observer must inevitably have some residual effect.

The (T,X) and (ρ12,r12) each transform in the same way as any conventional (t,x). If only (a1 + a2)=1, the rate of change of Ψmt with respect to T (with X, ρ12 and r12 fixed) equals the sum of the rates of change of Ψmt with respect to t1 and t2 (with x1, x2, and t2 or t1 fixed). The rate of change with respect to X is similarly the sum of the rates of change with respect to x1 and x2. These results are readily extended to systems of N > 2 particles. Total particle energy E and momentum P, equal to the sums of individual particle energies Ej and momenta pj, are then proportional to the rates of change of Ψmt with respect to T and X, respectively. Just as every (xj, pj) and (in MTT only) every (tj, Ej) form a pair of complementary, non-commuting operators, so too do (X, P) and (T, E).

Energy / linear momentum eigenfunctions for two free particles can now be written

      Ψmt(free) ∼ e-iωoto+iko·xo e-iΩT+iK·X e-iωρρ12+ikr·r12

where (Ω,K)=(ω1+ω2, k1+k2) and (ωρ,kr)=(a1ω2-a2ω1, a1k2-a2k1). Note that ωρ (and so all dependence on ρ12) vanishes if aj=ωj (Niestiik COM)! The free eigenstate shows no preferred value of T, X, ρ12 or r12 at any to. More generally, interactions between the observer and the two-particle system determine the time correlation between to and T. Interactions between particles 1 and 2 determine not only the primary dependence of Ψmt on r12, but also the correlation between t1 and t2, and so the dependence of Ψmt on ρ12. For strong repulsive interactions, Ψmt should be significant only for large values of the Niestiik-invariant timespace separation S12≡ √r122 - c2ρ122 . For strong attractive interactions, there should be solutions of Ψmt localized to small values of S12. If t1 and t2 become tightly correlated, ρ12 is confined to values near zero, and the particle-related time dependence of Ψmt is mainly through T.

Both QFT and standard MTT assume a pre-existing, observer-based, four-dimensional (4D) timespace framework (t, x). Even if this timespace is affected (warped) by matter and energy, it is not created by them. Yet what is the origin of timespace itself, and how are its coordinates meaningfully defined at all for a multiparticle system? Neither time nor space can be measured in absolute terms. Temporal and spatial intervals are gauged only with respect to physical processes and structures, which are traditionally interpreted in terms of elementary particles and their interactions. Stripped of these vestments, timespace loses all meaning. Physical objects and dimensions of relation are inextricably linked.

As mentioned earlier, every elementary particle with mass does have an innate proper time dimension. For every external observer time tj in standard MTT, there is a corresponding internal proper time τj. Indeed, MTT can be reformulated using the more fundamental τj. Yet because distinct proper time lines are not inherently synchronized, and are correlated only by particle interactions, further temporospatial relationships can be defined only if particles interact.

Interactions corresponding to the fundamental forces are associated with gauge symmetries. In QFT, their description can be interpreted in terms of the exchange of phantom elementary gauge bosons by elementary fermions. The electromagnetic, weak, and strong interactions involve the exchange of phantom photons, W and Z bosons, and gluons, respectively (all spin-1). Phantom particles have all the attributes of their real counterparts, except mass; the usual relationship between energy, momentum and rest mass is not followed, making these particles ephemeral. Many physicists consider phantom particles not real in any sense, but merely a bookkeeping device. Elementary fermions include electrons, neutrinos, and quarks (all spin-1/2). [Fleegello's archaic list omits various types of invisible matter, that interact with ordinary matter solely through the gravity.] Only the gravitational force, which is ostensibly associated with the exchange of phantom gravitons (normally massless, spin-2), has eluded incorporation into the QFT framework.

Although phantom particles were introduced in QFT, gauge symmetries in MTT should lead to expressions for interactions with an analogous interpretation. Recall that individual particles can be identified in MTT, but not in QFT. Because phantom bosons in either approach are superpositions spanning energies and momenta that do not respect standard mass relationships, their exchange should not be literally interpreted in terms of particle trajectories. Yet they do link interacting particles, and transfer information at light speed. Fermions may also be linked and transfer information via real gauge bosons. The exchange of either phantom or real particles can thus establish causal links (CLs) – phantom causal links (PCLs) or real causal links (RCLs), respectively. Because individual particles can be identified in MTT, but not in QFT, fully exploiting CLs requires an MT approach.

Consider then a multi-time CL of any type from fermion j at proper time τj to fermion k at τk. Such CLs are universal; all observers recognize the same connections, at the same proper times of the respective particles. CLs are directional; energy and other information flow forward or backward from one particle to another. This interparticle direction may be indicated by a binary parameter θjk=±1, where value +1 signifies forward flow from j to k, and -1 signifies reverse flow. Every [electromagnetic] CL can also be associated with a three-component (equivalent to a 3D) unit vector ujk that defines a 3D spatial direction of flow.

Define an elementary event as any point on the world line of a real elementary particle at which a CL is established with another particle. The physicists Machi and [later] Niestu have promoted the radical idea that the network of CLs among particles does not merely occur within timespace, but even defines timespace. The number of spatial dimensions is set by the number of components in ujk. Particle Interaction For example, consider the pair of electromagnetic links between two charged particles #1 and #2 in the diagram at right. Suppose a CL (dotted line) connects #1 (solid line) at proper time τ1A, defining event A, to #2 at τ2B, marking event B; and a second CL exists from #2 at the same τ2B, back to #1 at τ1C, or event C.

The spatial location of particle 2 from the perspective of particle 1 (the "observer" in this case) at τ1o = (τ1A+τ1C)/2 is

      r12B = r12B u1AB

where u1AB is a 3D unit vector pointing in the spatial direction of flow from A to B from the perspective of particle 1, and r12B is the scalar interparticle distance

      r12B = (τ1C-τ1A) c/2 .

From the same perspective, u1BC = -u1AB. Note that neither τ2B nor the relative interparticle speed v appear in these equations. If τ1 and τ2 are not perfectly correlated (by the probability distributions for all possible interactions, which coincidentally define v), a range of τ2B values could give the same result. Timespace coordinates for event B are (τ2B,0) from the perspective of #2. Time τ1B should be related to τ2B by at least a probabilistic, Niestiik-like (due to v-dependence) transformation. For non-Niestiik motion, one can set τ1B~τ2B based on the single interaction. However, τ1B must equal τ1o only if τ1 and τ2 are perfectly correlated. More generally, τ1B has a range of possible values, corresponding to the range of τ2B, but (with even minimal synchronization) centered on τ1o.

From the perspective of #2, the situation is more nuanced. It is not generally true that the two link paths are equal, or that u2BC = -u2AB. Interparticle distance r21B at τ2B now depends on v, with Niestiik corrections expected. While both r21B and v can be related to the event times (τ2A, τ2B, τ2C) from the perspective of #2 (an interesting exercise!), deriving τ2A from τ1A, and τ2C from τ1C, now depends on the correlation between τ1 and τ2. The interaction defines a definite correspondence τ1o=τ1B=τ2B, with a common interparticle separation r21B = r12B, only if time is absolute (as in classical physics), and the time lines are synchronized.

Multiple serial links are required to correlate time lines and define interparticle distance over time. The properties of successive links must be compatible, as per the equations of motion, to accord with physical principles. In particular, adjacent CLs should reflect a consistent sense in the flow of time, or distance is ill-defined. Relative speeds must not exceed c. Changes in velocity (acceleration) should reflect the transfer of linear momentum by CLs.

Let 12 represent the probability amplitude, from the prespective of particle 1 (the first index), for a CL connecting particles 1 and 2, with [partial] functional dependence 12 (τ1, τ2, θ12, u12). Assume that 12 is a complex quantity, and that the (real) connection probability per τ1, τ2, and solid angle in u12 is |12|2. There may be applications where it is useful to renormalize 12 such that, for any (τ1,θ12), the sum [integral] of Δ|12|2 over (τ2, u12) is unity, where Δ is a (small) interval along τ1 [if proper time is discrete, Δ may be set to the minimum time interval]. In the weak interaction limit, 12 is then the same, minuscule value for all (τ1, τ2).

The amplitude for link AB is 12 (τ1A, τ2B, +1, u1AB), and for BC is 12 (τ1C, τ2B, -1, -u1AB). Measured by 1 at τ1o, distance r12 is defined by link pairs of like AB and BC, with τ1A=τ1o-r12/c and τ1C=τ1o+r12/c. For any τ2B and θAB=+1, the average of |AB|2 over u1AB should peak at some τp1A. For the same τ2B and θCB=-1, the average of |CB|2 over u1CB peaks at some τp1C. This defines a most probable connection between τp1o=(τp1A+τp1C)/2 and τ2B. By evaluating τp1o over a range of τ2B, the function τp1o(τ2) is obtained. The width of the distribution |ABCB|2 along τ2B is a measure of the correlation between τ1 and τ2 at τp1o, which is not in general one-to-one. Single-time theory then cannot most accurately describe the system. The distribution width with respect to r12 at fixed (τp1o,τ2B) is a measure of the uncertainty in r12. As in standard Shrodiik physics, this may be nonzero even when τ1 and τ2 are well correlated.

Remarkably, ||ABCB|2 has units equivalent to a probability per 4D (τ2, r12B) timespace volume, just like the probability of an MT observer (here, particle 1) seeing particle 2 at τ2B and r12B. Apart from x1o dependence, one can identify a CL-related partial wavefunction ɸmt, with

      ɸmt(τ1o; τ2, r12) ~ 12 (τ1o-r12/c, τ2, +1, r12/r12) 12 (τ1o+r12/c, τ2, -1, -r12/r12).

For a given τ1o, the quantity |ɸmt|2 defines a joint probability distribution over (τ2B, r12, u1AB). Amplitudes should be normalized such that the sum [integral] of |ɸmt|2 over all (τ2, r12) is unity.

A student of Draci [likely inspired by QF] has investigated the idea that a physical object may even have PCLs to itself. A link connecting τ1a to τ1b on the world line of particle 1 effectively extends outward a distance r11~(τ1b-τ1a)c/2 at a time τ1o=(τ1a+τ1b)/2. For a lone particle 1, the corresponding amplitude may be represented by 11 (τ1a, τ1b, θ11, u11). The CL-related partial wavefunction associated with self links is then

      ɸmt(τ1o, r1o) ~ 11 (τ1o-r1o/c, τ1o+r1o/c, +1, r1o/r1o) 11 (τ1o+r1o/c, τ1o-r1o/c, -1, r1o/r1o)
                          ~ 112 (τ1o-r1o/c, τ1o+r1o/c, +1, r1o/r1o) .

Self-links suggest a way to physically interpret r1o. Indeed, it may be that r1o is meaningfully defined only by self-links. Note that r1o is defined by a single link rather than a pair of links, and the two link amplitudes associated with ɸmt(τ1o, r1o) are identical. Self links mark elementary particles as extended objects. They may even embody traditional force fields (e.g., the electric field of an isolated charged particle), and contribute to rest-mass energy.

In a two-particle system, a self-link may become correlated with links between the particles, such that the corresponding joint link amplitude is not a product of two simple amplitudes. Let jk represent the full joint link amplitude, including all self-links, for CLs involving two particles j and k, where j and k are either 1 or 2 in the current example. Let τjk now be the proper time on the world line of particle j, for a link connecting j to k. The functional dependence of 12 is 12 (τ11a, τ11b, θ11, u11; τ12, τ21, θ12, u12; τ22a, τ22b, θ22, u22).

Based on its relationship with the standard wavefunction, it is convenient to now redefine jk as a truncated amplitude only including CLs with at least one end anchored to the world line of particle j (the first amplitude index). While the full amplitude jk contains more information and is a more complete representation of a system, jk can be derived from jk. The functional dependence for N=2 particles is 12 (τ11a, τ11b, θ11, u11; τ12, τ21, θ12, u12).

This scheme can be extended to systems with N>2 particles. Links between different particle pairs may themselves become correlated, such that joint link amplitudes cannot be written as products of one- or two-particle amplitudes. The N=3 joint amplitude for particles 1, 2 and 3 may be written 123, and includes links 1-1, 1-2, 1-3, 2-2, 2-3, and 3-3. The truncated amplitude 123 includes links 1-1, 1-2, and 1-3.

Because they depend on proper times with respect to distinct particles, the jk and jk utilize a mixed perspective. Although they may constitute a most fundamental representation of CLs, and proper times are independent of perspective, 3D directions and the values of quantities transferred by CLs do depend on perspective. It is thus customary, and often useful, to express functional dependence solely in terms of quantities defined from a common perspective.

In the traditional external-perspective MT (MTe) approach, a wavefunction is defined from the perspective of an external inertial observer, typically an organization of myriad physical objects that establishes a timespace frame (to, xo), and provides a consistent perspective. To measure the position of a solitary particle 1, this observer must rely on CLs connecting its own world line at to1 with the particle world line at t1o (measured in the observer frame), where t1o may be interpreted as the most likely to corresponding to proper time τ1o. The full observer-based link amplitude may be represented by Γo1 (tooa, toob, θoo, uoo; to1, t1o, θo1, uo1; t11a, t11b, θ11, u11), and the truncated amplitude by γo1 (tooa, toob, θoo, uoo; to1, t1o, θo1, uo1). Function γo1 embodies all information accessible to the observer by measurement, including self-position (whether a displacement or an extension) and particle position as functions of observer time, and defines a partial wavefunction ɸo1, with

      ɸo1(to, xo; t1, x1) ~ γo1 (to-xo/c, to+xo/c, +1, +xo/xo; to-x1/c, t1, +1, +x1/x1)
                                    γo1 (to+xo/c, to-xo/c, -1, -xo/xo; to+x1/c, t1, -1, -x1/x1) .

For two particles 1 and 2, link amplitudes are Γo12 and γo12, defining ɸo12(to, xo; t1, x1; t2, x2). Links involving only particles 1 and 2 are defined in Γo12, but not γo12, though the two amplitudes must be compatible. ɸo12 defines a joint probability that the observer at to sees itself at (to, xo), particle 1 at (t1, x1), and particle 2 at (t2, x2). Units of |ɸmt|2 are probability per 3D xo spatial volume for the observer, and per 4D (tj, xj) timespace volume for every other j. There is no probability per to, as there is no distribution over to at time to.

In the spirit of exploring how CLs might define timespace, members of the Draci group have introduced an alternative internal-perspective MT (MTi) approach, which does not rely on a pre-existing 4D timespace. Instead, the theory assumes only that every particle has its own innate proper time, and that CL (gauge) bosons are associated with mathematical objects that define common multi-D properties. Quantities are defined from the consistent but non-inertial perspective of a given particle, indicated by the first index in the MTi link amplitude.

For two particles 1 and 2, the full MTi link amplitude with respect to particle 1 is represented by Γ12 (t11a, t11b, θ11, u11; t12, t21, θ12, u12; t22a, t22b, θ22, u22), and the truncated link amplitude by γ12 (t11a, t11b, θ11, u11; t12, t21, θ12, u12). Here all times are measured with respect to the proper time of particle 1, so that t1j=τ1j for any j, and t2j is the most probable proper time along the world line of particle 1 corresponding to the linked τ2j. Note that for any (τ12,τ21) there is a probability distribution over t21, and for any (τ12,t21) a distribution over τ21, defined by Γ12.

The quantity γ12 defines a CL-related partial wavefunction ɸ12(t1, r1; t2, r2) in the usual way, except that particle 1 now replaces an external observer. ɸ12 is a joint probability amplitude that particle 1 at t1 sees itself at r1, and particle 2 at (t2, r2). The system may equally be represented by ɸ21(t2, r2; t1, r1) from the perspective of particle 2, with link amplitudes Γ21 and γ21.

The external- and internal-perspective approaches can both be extended to more complex systems with N>2 particles. If the various links become correlated, the CL amplitudes cannot be encoded using two-particle Γojk and γojk, or Γjk and γjk. For N=3, it is necessary to introduce correlated link amplitudes Γojkl and γojkl, or Γjkl and γjkl. Information regarding links not involving the observer is found in Γo123, but not in γo123. Similarly, information regarding links only involving particles 2 and 3 is in Γ123, but not in γ123. Now γ123 (t11a, t11b, θ11, u11; t12, t21, θ12, u12; t13, t31, θ13, u13) is the joint probability amplitude from the perspective of particle 1, that a self-link joins 1 at t11a to itself at t11b, while a second CL joins 1 at t12 to 2 at t21, and a third CL joins 1 at t13 to 3 at t31, in the specified directions. The corresponding partial MTi wavefunction can be represented by ɸ123(t1,r1; t2,r2; t3,r3). This object embodies the joint probability amplitude that particle 1 at proper time t1 sees itself at r1, particle 2 at (t2,r2 ), and particle 3 at (t3,r3 ). The system may equivalently be represented by ɸ213 or ɸ312, from the perspectives of particles 2 or 3. More generally, for N particles, there are N+1 indices and a single distinct perspective and link amplitude in MTe, while there are N subscripts and N distinct perspectives and link amplitudes in MTi.

The internal-perspective approach does not rely on any pre-existing external reference frame; coordinates are defined internally by interconnections among a given set of particles. Yet there are difficulties with this approach. For any system of interacting particles, a reference frame anchored in one particular particle is inherently non-inertial. If particles are identical, exchanging indices further involves a change in coordinate prespective.

For non-Niestiik motion, these problems may be overcome by transforming coordinates to a center-of-mass reference frame through a spatial displacement vector. For two particles, the partial MTi wavefunction becomes ɸ12C(tC1, xC1; tC2, xC2), where index "C" now indicates an inertial COM reference. If the particles are identical, ɸ12C is easily symmetrized. Alternatively, composite COM coordinates (TC,XC; ρC,rC) may be used, as decribed earlier. Both TC and XC are unchanged by particle exchange, while the signs of ρC and rC are reversed.

If CLs transfer information between particles, their amplitudes should be functions of the associated quantities. Consider in this regard a simplified, two-particle MTi link amplitude γ12 (t12, t21, θ12, u12), where correlations with self-links and with other particles are ignored.

In addition to implicitly defining interparticle distance r12 from the perspective of particle 1, the given link may be associated with an energy and linear momentum transfer. Let ω12 be the angular frequency associated with energy transferred along u12, and K12 the 3D wavevector associated with linear momentum transferred in the same direction, again from the perspective of 1. The quantity K12 should contribute to changes in relative particle velocity. The probability per t12, t21, solid angle in u12, frequency, and 3D wavevector volume that a link between 1 and 2 exists is now the absolute square of γ12 (t12, t21, θ12, u12, ω12, K12). [Link amplitudes may also specify angular momentum transfer, and the type (phantom or real) of linking particle.]

Although information carried by a PCL effectively moves at light speed, such links do not represent real particles, so there is no fixed relationship between ω12 and K12. Because a negative frequency ω12 moving in the positive time direction is equivalent to a positive frequency moving in the opposite θ12 direction [as well as a positive-frequency antiparticle moving in the positive time direction], then physically meaningful frequencies ω12 may be restricted to positive values.

The amplitude γ12 may in general be nonzero for K12 pointing in directions other than u12. Whereas the average value of K12 is collinear with u12 for RCLs, this restriction does not apply to PCLs. Define K+12 to be the vector component of K12 pointing along u12, with K+12= K+12u12. Forward wavenumbers K+12 may be either positive, if an interaction is repulsive; or negative, if it is attractive. K+12 respectively points in the same direction as u12, or in the opposite direction.

For a PCL, it is straightforward to conserve linear momentum at both time-ordered link vertices regardless of the value K12. However, energy is then not generally conserved. This is allowed, since PCLs are transient, and do not represent real particles. If frequencies closest to energy conservation are favored, then ω12<<K12/c, and the effective link mass is imaginary. The total energy of two particles may change (mainly due to momentum exchange) during a link interaction, reflecting a conversion of potential energy to or from kinetic energy.

If there is no interparticle motion, then u21=u12. If particle 2 moves relative to 1, then ω21 and the components of u21, r21 and K21 fwith respect to 2 will differ from the values seen by 1. The functions relating the two perspectives should be compatible with Niestiik transformations for the given link (speed-c if a PCL or zero-mass RCL). Those equations maintain any collinearity between K and r. If relative speed is non-Niestiik, then r21 ≅ -r12.

[Fleegello unnecessarily restricted each CL to the transfer of a single quantum of information. A CL may actually transfer any number of information packets. A link amplitude must specify this number, as well as the angular frequency and linear momentum associated with each. Amplitudes may alternatively be reformulated in a way resembling QFT, to specify the number of quanta occupying each of a set of states between each particle pair.]

What are the functional forms of the link amplitudes and wavefunctions in MTT? The explicit functional dependence ultimately derives from the MT equations of motion, which have not yet been identified, but must be distinct from the equations of QFT. It should nonetheless be possible to write any physical link amplitude as a Foorier sum over a complete set of eigenfunctions of ω12 and K12.

The time dependence of a useful frequency eigenfunction should encode the direction of information flow, in a manner consistent with θ12. In quantum physics, a complex phase factor e-iωt indicates flow in the positive t (time) direction, where ω is angular frequency. A practical eigenfunction of ω12 (again ignoring self-links) should then include a temporal phase factor

      ∼ e-i ω12 θ12 (t21-t12) = e-i ω12 δ12

where δ12θ12 (t21-t12).

An eigenfunction of both frequency and wavenumber must also include a spatial phase factor. Distance r12=(t21-t12) u12c, defined at t21=t12+θ12 r12/c from the perspective of 1, is the space-like vector corresponding to δ12. Note that r12=r2+ε12 and t21=t2 for (t2,r2) found in ɸ12(t1,r1; t2,r2), where ε12=c(t2-t1)u12. Both coordinate pairs (δ21,r12) and (t2,r2) have operator counterparts in MTT. The overall eigenfunction becomes

      ∼ e-iω12δ12  e+iθ12K12·r12 .

Functions of this form are akin to energy-momentum eigenstates of free particles. [Note that other complete sets – e.g., angular momentum eigenstates – are also possible.]

The given eigenfunction does not favor any linkage 1 to 2; its absolute value is the same for all (t12,t21). While the corresponding link amplitude may not be a solution to the actual equations of motion, it should be possible to express a physical state of γ12 as a superposition of the eigenfunctions. Mathematically, a summation

      ω  e-i ω (t-to)

over a range of ω peaks at t to. If t1 and t2 are correlated, then |γ12| should peak at some δ12 d12/c, where d12 is the length of an interparticle distance d12. If r12 is bounded, |γ12| should also peak at r12 d12. In a restrictive sum over ω12 and K12, the link amplitude becomes

      γ12e-iω12(δ12-d12/c)  e+iθ12K12·(r12-d12) .

Here d12 is in general a function of t21, as determined by the equations of motion.

Note that under particle exchange, the two-particle amplitude γ12 (t12, t21, θ12, u12, ω12, K12) becomes γ21 (t21, t12, θ21, u21, ω21, K21), where the same quantities are defined from different perspectives in the respective amplitudes. Because γ12 and γ21 represent the same link, the exchange of labels must not alter any physical link characteristic. This condition is in fact satisfied for the given amplitudes, if perspectives are connected by a Niestiik transformation.

How closely can superpositions of simple frequency/wavenumber eigenfunctions correlate the times t1 and t2? In the classical limit, the time lines can be perfectly synchronized, and a common, well-defined 3D distance d12(t21) exists at time t21. The quantities

      η ≡ [θ12(t21-t12) - d12/c]   and

      ξ ≡ (r12-d12)

are then both zero with 100% probability. The corresponding γ12 can be approximated by a maximal summation over (ω12, K12) eigenfunctions. In the classical extreme, the sum is over discrete values of ω12 from 0 to +∞, and each component of K12 from -∞ to +∞ (including zero), at increments Δω12 and ΔK12, in the limit Δω12→0 and ΔK12→0. Each phase factor in the sum is multiplied by the four-dimensional product of the increments.

[This sum is closely related to the product δ(ηδ3(ξ) of four Draci delta functions, where δ(x) is defined by δ(x) = 0 for x≠0, and δ(0)=∞ such that the area under δ(x) is unity. δ3 is the product of three ordinary delta functions, one for each spatial component of ξ. By ultimately limiting the range of ω12 and K12, Fleegello avoids a thorny normalization issue associated with delta-function probability amplitudes.]

Can such a link amplitude be realized in our physical world? RCLs and PCLs actually have distinct minimum and maximum allowed absolute values (ωmin, ωmax) of angular frequency and (Kmin, Kmax) of wavenumber in any superposition of simple states.

For real massless bosons, an RCL travel distance d12 defines a limiting maximum wavelength λmax=2d12, entailing a minimum frequency ωmin=𝜋c/d12 and wavenumber Kmin=𝜋/d12. Limiting maximum values of frequency and wavenumber are defined only by the (inverse) smallest possible size of a timespace interval, and so are huge but presumably finite.

The situation is reversed for PCLs. Because a PCL represents collective phantom processes and does not comprise an independent time line, phase cannot gradually change along its length; only the net shift across a PCL is meaningful. This shift must then be limited to the range -𝜋 to +𝜋. Any outside value would be mathematically indistinguishable from a number inside the range. A PCL travel distance d12 thus defines a limiting maximum absolute frequency ωmax=𝜋c/d12 and wavenumber Kmax=𝜋/d12. The limiting minimum allowed frequency and wavenumber are both zero.

For RCLs, the ω12 and K12 must further satisfy the standard Niestiik energy–momentum relationship with rest mass (zero for a typical massless boson link), and both ω12 and K+12 are restricted to positive values. For PCLs, the ω12 must also be positive, but K+12 is restricted to positive values if an interaction is repulsive, and negative values if it is attractive.

A PCL sum may not favor all frequencies and momenta equally even within the allowed ranges. Analogous sums in QFT electromagnetic calculations include factors like 1/(ω2-c2K2), favoring photon-like states with effective mass near zero. As suggested earlier, sums for PCLs may be restricted to ω12<<K12/c. Applying appropriate limits, but neglecting possible weighting factors, a link summation can be converted to an integral, and evaluated using calculus. Probability distributions may be derived from the absolute square of the result.

For PCLs, the (normalized) probability distribution PPCL(η) associated with the sum over ω12 is

                          2 sin2(Δω η/2)
      PPCL(η  _________________      where      Δω ≡ (ωmax-ωmin).
                              𝜋 Δω η2

This symmetric distribution still peaks at η=0, now with finite value Δω/2𝜋. The uncertainty ση in η ( ~ distribution half-width at first null) is not zero, but ~2𝜋/Δω. For PCL limiting values,

      PPCL(η=0) = 0.5c/d12      and      ση ~ 2d12/c .

The distribution PPCL(ξ) associated with the sum over K12 is more complex, but peaks at ξ=0, with value ~5Kmax/6𝜋 if Kmin=0. The uncertainty σξ in ξ (again defined by the first null) is ~8/Kmax. For PCL limiting values,

      PPCL(ξ=0) ~ 5/6d12      and      σξ ~ 8d12/𝜋.

For RCLs, there is no explicit integration over ω12 , as it is folded into the integration over K12. However, for massless boson links, inertial invariance requires that |η|=ξ/c with ξ≥0, so that PRCL(|η|)=PRCL(ξ). The distribution peaks at at η=0 and ξ=0. If KminKmax, then Kminξ≪1 inside the main peak of the distribution, and the peak value is again ~5Kmax/6𝜋. The uncertainty σξ remains ~8/Kmax, though the Kmax limit is now much larger than for PCLs.

The correlation between t1 and t2 is thus in general not one-to-one, especially for PCLs. As long as elementary physical objects have separate world lines (this conclusion then applies to MTa, but not QFT), STT is an approximation; a multi-time scheme must be more accurate. The classical limit is approached only through a sum of lower-energy interactive phantom processes and higher-energy real boson links. The minimum uncertainty in the correlation between t1 and t2 from phantom processes alone is ~d12/c (the time light travels d12). An uncertainty ~d12 is likewise inherent in any PCL specification of interparticle distance. Values for a macroscopic observer viewing a single fermion could be much smaller, if real particles are used as probes.

Recall that defining distance r2 at time t1 in ɸ12(t1,r1; t2,r2) requires pairs of links AB (1 to 2) and BA (2 to 1), with respective amplitudes (still ignoring self-links)

      γAB = γ12 (t1-r2/c, t2, +1, r2/r2)     and     γBC = γ12 (t1+r2/c, t2, -1, -r2/r2) .

Each amplitude can be separately represented by a weighted sum of pure-frequency eigenfunctions over a range of temporal and spatial frequencies. Define ε12=c(t2-t1) r2/r2. Even if restricted to values ω12<<K12c, a broad sum over K12 in both γAB and γBC causes the absolute value of each amplitude to peak at r2 + ε12 ~ d12 and r2 - ε12 ~ d12, respectively. The distribution peak of |ɸ12(t1; t2, r2)|2 is then subject to the joint condition r2 ~ d12 and t2 ~ t1, as expected.

Consider now the opposite extreme, in which γAB and γBC are each represented by a simple, pure-frequency eigenfunction. Setting ωBC = ωABω, KBC = -KAB ≡ -K/2, ignoring the irrelevant dependence on d12, and imposing the realistic restriction ω12<<K12/c, then

      ɸ12(t1; t2, r2) ∼ e+K·r2 ,

with an r2-dependence similar to that of a traditional wavefunction with a pure wavevector K2=K.

More generally, in both MTe and MTi, partial wavefunctions ɸmt constructed in this manner can embody all information concerning relative positions, but lack the time dependence of the complete wavefunction Ψmt. In the weak-interaction limit for N particles (plus observer if MTe), we expect

      Ψmte-iωoto   e-iω1t1   . . .   e-iωNtN

where ω2 = k2c2 + m2c4/2 for each object, and m is object rest mass. It can be shown that the missing time factors can be recovered in ɸmt by including self-links in the link amplitudes. ɸmt can thus in principle be made equivalent to Ψmt. To the extent they define this wavefunction, links essentially define time-space itself.

The precise form and time evolution of MT causal link amplitudes and wavefunctions are ultimately defined by dynamic equations of motion. What are these equations, and . . .
.
 
.
 
.
 
.
 

[At this point there is a gap in the historic record, in which Fleegello purportedly tried to formulate a regorous mathematical framework for MTT, and demonstrate that CLs provide a complete, novel MT framework for representing multi-particle systems. His failure to complete this project apparently led Fleegello to doubt his own competence and worth as a natural philosopher / physicist, and to destroy all related material in a fit of despair shortly before his death.

Physicists eventually did develop a complete multi-time quantum theory that supplanted standard single-time QFT, in which the connections among elementary non-gauge particles (mainly fermions, plus non-gauge bosons unknown to Fleegello) inherently define spacetime, and PCLs are dictated by gauge symmetries within an MT context. For every non-gauge particle type, there is a pair of operators that create or destroy, respectively, one such particle (including its own proper time line). For every gauge boson, there is a pair of operators that create or destroy, respectively, a CL between two specified points on particle world lines. Operator commutation relations ensure the appropriate symmetries of quantum states.]

Insofar as CLs specify 3D direction vectors, they also define (probability amplitudes for) the relative 3D positions of all causally connected particles in a system, regardless of their number. Yet CLs should be capable of establishing a 3D spatial framework even without these inherent 3D direction vectors.

Consider in this regard an isolated system of N distinguishable particles. If relative speeds are non-Niestiik, then interparticle distances djkdkj, and CLs define amplitudes for N(N-1)/2 such distances. These are sufficient to construct amplitudes for the (3N-6) coordinates of 3D interparticle relative position vectors (up to a rigid rotation and displacement of the entire system), as long as N>3. Indeed, they should be sufficient to determine all relative coordinates for up to (N-1) dimensions! Relative speeds and accelerations are implicit in changes in the djk along world lines. Comparable conclusions apply even to quantum systems of identical particles, and for Niestiik speeds.

The web of CLs and associated proper times among the world lines of elementary particles could thus determine the geometry of timespace, in particular the large-scale geometry, whether or not CLs specify spatial direction. In either case, space can unfold from the relationships among myriad connected events. To the extent this occurs, space does not have independent existence, but is defined by the connections between the mathematical objects we perceive as particles. A world without causal links would be a world without space; any particles would be independent of each other, with no meaningful positional relationships.

Yet why do inter-particle connections defined by interactions specifically generate an overall three-dimensional, nearly flat (under normal conditions) space? Classically, even without inherent 3D directions, the N(N-1)/2 interparticle distances in an N-particle system are sufficient to determine all relative coordinates for up to (N-1) dimensions!

The mathematical physicist Wittuu has proposed a mechanism that both restricts and defines the number of spatial dimensions. Large-scale timespace is defined mainly by the electromagnetic interaction, since it is associated with the exchange of phantom photons of unlimited range. While real photons are massless, and so restricted to two spin states, phantom photons have three spin states, like any spin-1 particle with mass. These correspond to three inherent, independent "directions." Because all photons are identical, exchanging their identities cannot alter a physical system, so they must share the same three directions. This causes every interparticle distance defined by CLs from photon connections to be limited to vectors in a common macroscopic 3D space.

The strong and weak interactions should establish additional spatial dimensions. Although there are eight gluon types, these are not truly independent, and together generate only three new dimensions. The three bosons of the weak force generate the same, making a total of nine spatial, or ten timespace dimensions. Yet the ranges of the strong and weak interactions are so tiny (~10-13 and 10-16 centurets, respectively), they mainly affect the small-scale geometry of timespace. Wittuu suggests that the associated dimensions are "curled up" or "attenuated," and only obvious at very small scales or high energies.

[While Wittuu's argument was sketchy, later generations of physicists demonstrated that his intuition was sound (although he missed a few small-scale dimensions). A rigorous explanation of the origin of macroscopic spatial dimensions was eventually developed. Spacetime essentially arises as an emergent property from the interconnections among primitive, abstract, timeless mathematical forms.]

What about gravity? The carrier of this interaction is ostensibly the massless graviton. Because the graviton is a spin-2 particle with unlimited range, gravity might be expected to generate its own large-scale 5-dimensional space. Yet gravity has an unusual character, related to its incompatibility with standard QFT. All other fundamental forces are carried by spin-1 bosons, and associated with unique and conserved "charges" (e.g., electric charge). But gravity couples to a system's stress-energy tensor, to which every interaction contributes. Gravity reduces all "bare" mass energies, and even couples to itself, leading to nonlinearities in the gravitational field equations. PCLs established by gravity are thus dependent on and flow from the other interactions, so that gravity does not add any new dimensions. It may nonetheless distort the large-scale structure of timespace, as in Niestu's theory of general invariance.

It is clear that neither standard graviton exchange nor general invariance represents the complete fundamental description of gravity, even if both are good approximations in the low-energy limit. A missing element in these theories may involve the small-scale structure of time. Physicists have traditionally considered time to be continuous. In an attempt to avoid divergent (infinite) quantities in QFT calculations [which had previously been removed for forces other than gravity by a dubious procedure known as renormalization], Planko has proposed that proper time is quantized along the world line of every elementary particle with mass. The fundamental (minimum) proper time interval, or chronon, is represented by the symbol Δ. Distance between particles is naturally quantized in integral multiples of Δc/2. [Spacetime volume is thus more generally quantized, and not space or time separately. By inertial invariance, this quantity is unaffected by a velocity transformation.]

Quantized proper time may be required by ideobasic principles. Consistency logic compels the PCS to recognize the most general conception of time. Yet continuous time is only a limiting case of quantized time. The infinity of numbers on a continuous line segment is furthermore countable only for rational values; there is no one-to-one correspondence between irrational numbers and the set of positive integers. If it is impossible to locate irrational values within the PCS field, they cannot have meaningful existence there (although the concept may exist). Note that time is inherent only along particle world lines; it does not meaningfully reside anywhere else.

If timespace is quantized, then the smooth differential equations of Shrodiik mechanics, QFT and MTT must be replaced by discrete difference equations. Observables defined in terms of derivatives must be similarly redefined. Symmetry principles and conservation laws are all affected.

[Other physicists had previously hypothesized that spacetime was quantized, but only with respect to the overall coordinate framework of a given observer, not with respect to individual particles. These lattice approaches were doomed to failure, as they were divorced from the very processes that define space and time.]

A minimum proper time interval Δ along an elementary proper time line implies an associated maximum absolute angular frequency

      ωmax = 𝜋/Δ,

and a range of meaningful frequencies

      0 ≤ ω ≤ +ωmax .

This limit also applies to RCLs, though a smaller cutoff should apply to PCLs.

In MTT, every elementary particle is characterized by a single, separate proper time line. Based on a symmetric version of the modified single-particle equation of motion, Planko has proposed replacing the linear equation relating the energy E of an elementary particle in its own rest frame to its proper time angular frequency ω by the trigonometric formula

      E = (ωmax/𝜋) sin(𝜋ω/ωmax) .

This reduces to the standard equation when ω/ωmax << 1, and can alternatively be written

      E = Emax sin(Eo/Emax)

where Eo is a particle's uncorrected energy Eo = ω , and

      Emax = ωmax /𝜋 = /Δ at the angular frequency ωmax / 2 .

Note that Eo = moc2 for an elementary particle with an uncorrected (bare) rest mass mo.

[Massless particles have no rest frame, and there is no passage of time along their world lines; frequency and energy must be specified with respect to associated particles with mass.]

Proper time quantization reduces and limits a particle's effective rest mass energy, in a manner curiously similar to gravity. For every known elementary particle, Eo / Emax << 1. The sine function in the modified equation for energy can then be approximated by a truncated power series. Including only the first correction term in this expansion,

      EEo - Emax(Eo/Emax)3/6 .

According to standard Shrodiik theory, the uncertainty in the position of a particle of mass m cannot be smaller than /2mc. The bare rest mass mo for an elementary particle thus cannot be meaningfully confined to a volume with a radius smaller than

      rmin/4moc = c/4Eo .

Using this relation to remove one power of mo from the previous equation,

      Emoc2 - (c5Δ2/24)(mo2/rmin) .

The correction term is equivalent to the classical gravitational binding energy of a mass mo distributed over a surface of radius rmin, if one identifies the gravitational constant G as

      Gc5Δ2/12 .

Conversely, the chronon Δ can now be related to the gravitational constant by

      Δ = √12G/c5 .

Indeed, Planko has identified the minimum distance Δc/2 with the Planko length

      LP = √G/c3 ≈ 10-33 centurets,

and the chronon Δ with the Planko time

      TP = √4G/c5 ≈ 10-43 nocs,

which differs from the value derived above by less than a factor of two.

[These estimates are remarkably close (within a factor of eight) to the chronon value obtained from subsequent experiments. The Planko quantities were inferred theoretically from the time/distance scale at which the quantum effects of gravity should become significant.]

Planko has further proposed a natural system of units, in which the equations of motion are simplified. The unit of time is the chronon, and the unit of distance Δc. Interparticle separations are then half-integral multiples of the fundamental unit, and speed c is numerically equal to one. Units of mass and electric charge are selected so that both the Planko and the gravitational constants are numerically equal to one, and an elementary electric charge is the square root of the fine structure constant.

How does time quantization affect the energy of a multiparticle system? Consider a pair of elementary particles 1 and 2, interacting solely by virtue of their mass, that are at rest with respect to an observer, and separated by radial distance r. When the particles are far apart, interactions are negligible, so proper times τ1 and τ2 should be independent and uncorrelated. The two-body wavefunction Ψmt is the product of free-particle wavefunctions, with bare masses m1o and m2o . Because an increment Δ along one time line does not imply an increment along the other, total system energy Efar is the sum of individual particle energies:

      Efar = ω1+ω2m1c2 + m2c2

where m1 and m2 are the respective effective rest masses of each particle, related to the bare rest masses by

      mjc2 = Emax sin(mjoc2/Emax) .

The energy limit Emax applies only to the rest mass energies of the individual particles, along their respective world lines, and not to the overall system, so the maximum possible system energy is 2Emax. In terms of Niestiik COM time coordinates (T,ρ) defined earlier, the discrete rate of change of Ψmt along T no longer equals the sum of the discrete rates of change of Ψmt along τ1 and τ2. The latter instead equals the sum of the discrete rates of change along T for increments a1Δ and a2Δ, respectively. This is equivalent to the discrete rate of change along T for an increment ΔT, where

      Δ/2 (for ω1=ω2) ≤ ΔTΔ (for ω1ω2 or ω2ω1).

Note here that T is not a proper time measure, so the minimum increment Δ does not apply.

At smaller separations, τ1 and τ2 may be correlated by mass-related interactions (also required to define interparticle distance), and increments along τ1 and τ2 are no longer independent. The maximum correlation, at a minimum meaningful distance rmin, is equivalent to the particles merging into a single world line and proper time – increasing τ1 by Δ also increases τ2 by Δ, and vice versa. The time correlation parameter ρ is restricted to values near zero, and the individual particle wavefunctions merge into a single function characterized by a bare mass (m1o+m2o) and proper time T, now with a standard minimum increment Δ. Considering only bare mass and time quantization effects, the combined energy is then

      EnearEmax sin[(m1o+m2o)c2/Emax] .

Whereas a total energy limit corresponding to a minimum system time interval Δ/2 applies when the particles are far apart, a lower limit corresponding to Δ applies when they are near.

At intermediate separations r, one can write

      Er = (1-ε)Efar + εEnear ,

where ε is a function of r, such that ε → 0 as r → ∞, and ε → 1 as rrmin .

This equation can in turn be rewritten

      Er = Efar + Eint ,

where Eint is an effective interaction energy

      Eint = -ε (Efar - Enear) .

If bare mass energies are much smaller than Emax, then to good approximation, correction terms of order higher than (Eo/Emax)2 may be ignored, and

      Eint ≈ -ε c6 m1m2(m1o+m2o)/2Emax2 .

The minimum separation rmin can be estimated from the smallest volume that can confine the total uncorrected rest mass energy,

      rmin/4(m1o+m2o)c .

If the chronon is approximated by the value √8G/c5, then

      Eint ≈ -m1m2G (ε/rmin) .

This is identical to the classical expression for gravitational binding energy, if only

      ε = rmin/r .

This low-energy radial dependence is appropriate for a long-range interaction carried by massless gravitons. It also suggests that the degree of correlation between times τ1 and τ2 is ~ 1/r, consistent with a cutoff frequency ωmax ~ c/r for PCLs.

[While Fleegello used semiclassical reasoning in the above argument, the conclusions do have some qualitative validity.]

Thus, in the low-energy limit, time quantization appears to effect system energy in a way similar to the low-energy limit of standard graviton exchange, assuming only that τ1 and τ2 can be correlated by purely mass-dependent interactions. Gravity would then be similar to the other forces, in that it can be associated with the exchange of a gauge boson at low energies, but distinct in its more fundamental association with time quantization.

What if the interacting particles have electric charges Q1 and Q1? The bare electrostatic interaction energy is then

      Eoelec = Q1Q1 / r .

Again ignoring corrections of order greater than (Eo/Emax)2, and adopting the original ε(r), the residual interaction energy associated with time quantization is now

      Eint ≈ -[m1m2 + (m1+m2)Eoelec/c2 + EoelecEoelec/c4] G/r .

The three terms in brackets can be interpreted as the gravitational interaction energies between the two masses, between the masses and the electrostatic field, and between the electrostatic field and itself. As in more traditional theories, gravity couples to all relevant sources of energy.

The PCLs generated by the electric (or any non-gravitational) force appear to only indirectly affect the correlation of the proper times of the two particles; the derived mass-to-mass interaction energy with and without an electric interaction would otherwise have a different value for a given separation. The energies associated with non-gravitational PCLs apparently engender coincident graviton PCLs, which embody the actual process by which time correlations are established. This may reflect gravity's pivotal role in defining the geometry of timespace, and the fact that gravity does not appear to add any new spatial dimensions.

The origami-like unfolding of timespace from a milieu of interwoven, correlated events may generally result in a non-Euclidean macroscopic geometry. The effective curvature of conventional timespace would then be a natural consequence of the quantization of proper time intervals and the correlation of time lines by graviton exchange.

An elementary particle's rest mass may derive from a variety of sources. Every particle is effectively surrounded by a cloud of phantom exchange bosons, corresponding to all applicable interactions. But this is unlikely the sole source of mass for common particles. For example, if the electron's effective size is the minimum volume that can contain its mass, then the electron electric field contributes less than 1% to the observed mass value (the magnitude of the negative gravitational self-energy contribution is 43 orders of magnitude smaller).

Rest mass may also originate in a particle's underlying geometric character. Wittuu has suggested that elementary particles may not be pointlike, but associated with vibrations of extended (but tiny) geometric forms. The size of such entities should be comparable to the radial parameter rmin computed earlier for a given rest mass.

Can quantized timespace be incorporated into a quantum field theory? Time and 3D space have traditionally been treated as continuous system parameters in QFT. Yet past attempts to include gravity in QFT have failed; the theory is not renormalizable for point particles if system timespace is continuous. Even excluding gravity, the standard renormalized version of QFT predicts an enormous vacuum energy density. By adopting a multi-time framework, and restricting the proper time intervals between events to integral multiples of a chronon, maximum energies are naturally limited, and the divergent quantities in QFT calculations may be tamed.

As discussed previously, maximum frequencies associated with phantom processes in a multi-time setting should be further restricted, and inversely proportional to the distance between interacting particles. If space is defined by and only exists with respect to real material particles, then phantom processes that are completely disconnected from real particles might also be forbidden. Any new formulation should reflect that timespace is meaningfully defined only with respect to particle world lines and interactions. It may even prove necessary to treat elementary particles as finite-sized objects.

[The approach to quantized timespace outlined by Fleegello was naive, and flawed in many respects. It does not effectively address relative particle motion, or modifications to symmetry principles and conservation laws, or how an observer can fully integrate the proper times and spatial separations of individual particles with a global timespace coordinate system, and define a total system energy and wavefunction. Fleegello did acknowledge in private correspondence that his approach to quantized timespace was simplistic and incomplete, and certainly did not comprise a testable theory. Yet by replacing QFT with a multi-time theory, quantizing proper time intervals, and identifying elementary particles with extended (though minuscule) vibrating geometric forms, physicists were at last able to integrate gravity into quantum theory, and accurately compute the rest masses of elementary particles from first principles, avoiding the infinities that had plagued previous attempts.]

Although the fundamental (microscopic) equations of motion are symmetric in time, physical processes on a macroscopic level superficially do not appear to be time-symmetric. For example, if all the air molecules in a room were clustered in a corner, they would rapidly spread out to fill the entire room; yet the reverse process is not observed to happen.

Niestu has proposed that this so-called arrow of time is a purely statistical phenomenon.The universal state witnessed by an observer at a given moment is connected by a single time increment (chronon) to a host of other states. In Niestu's view, the number of less ordered (higher entropy) states corresponding to a "forward" time process is simply much larger than that for a "backward" process. If conscious experience is a random walk from one state to another, a person is much more likely to experience events along the traditional arrow of time. Reverse time steps occur, but are swamped by the sheer number of forward steps. This distinction acquires significance mainly in macroscopic systems, due to the sensitive dependence of the number of states of a given type on the number of particles in a system.

Yet this statistical feature does not in itself guarantee our experience of time. A conscious being that lacked a memory would live in an eternal present, with no sense of time's arrow. Most animal memories in a given universal state are found to be of events in connected states with lower overall entropy. This implies that the creation of memory generally involves a statistically irreversible process. Memories laid down in this direction are normally adaptive, and facilitate survival into an expanding realm of universal states. Memories laid down in the opposite direction could in principle also be adaptive, but only if they overtly present as precognitions, consistent with a person's walk through time.

[Shortly after Fleegello died, the natural philosopher Loh demonstrated that even the observed expansion of the universe could be linked to such statistical considerations, providing a critical link between cosmology, gravitation theory and Shrodiik physics.]

As discussed in section 1.18, any physical universe must have [at least] one basic (self-caused) initial state. This state can imply no previous history; the system would otherwise logically extend to an earlier time. Every physical universe must therefore evolve from an initial state characterized by infinitesimal spatial volume. Our own universe appears to have originally experienced runaway, exponential growth – the primeval hyperburst of modern cosmology – from a minuscule primitive state. Every newborn universe must further incorporate particles or analogous localized objects relative to which distance can be meaningfully defined. It could otherwise not expand (or contract) in any meaningful way.

[Fleegello failed to recognize that, if the experience of the CIF is timeless, a self-contained physical universe may also be cyclic, along a time-like dimension that loops back into itself. Such a system must in its entirety be the cause of itself. His basic argument has nonetheless since been extended to the multiverse of all possible physical worlds, whereby our own universe and its generative hyperburst may have been spawned by a pre-existing, self-caused system.]

Physics continues to evolve. Our understanding may yet be profoundly superficial. Will the physical objects and patterns identified so far prove to be unified by a single underlying entity? The multitudinous facets of one magnificent (mathematical) jewel? Or are they disparate, random elements, fragments tied loosely together only by the principle of consistency? Our descendants will hopefully discover the answer to this compelling question.

[During Fleegello's era, physics was rocked by conceptual revolutions every several jopes. Prominent scientists would periodically announce that a "theory of everything" was at hand, or that all that remained in physics was to clean up a few loose ends. These claims were invariably contradicted by new discoveries. Only after many octujopes of struggle was a viable unified theory in fact attained. Even then, physics was hardly dead. The new vision was so rich in possibilities, that its many veins continue to be mined even to this yad. Indeed, quantum physics is no longer considered the most fundamental of the physical sciences, but is viewed instead as the study of emergent phenomena arising from a still deeper level of mathematical reality. Other higher-level sciences (chemistry, biology, psychology, etc.) likewise continue to flourish, as an effective understanding of complex emergent reality inevitably transcends knowledge of underlying physical processes.]