Appendix A. Operator Formalism and Spinor Field
Construction
A.1 Overview A central feature of the Space–Time Membrane (STM) model is the emergence of fermion-like spinor fields from a purely classical elastic membrane. In this appendix, we detail how the classical displacement field
– whose dynamics are governed by a high-order wave equation including fourth- and sixth-order spatial derivatives, damping, nonlinear self-interactions, Yukawa-like couplings, and external forces – is promoted to an operator
via canonical quantisation. We also define its conjugate momentum and introduce a complementary out-of-phase field
. A bimodal decomposition of these fields subsequently yields a two-component spinor
, which forms the foundation for the emergence of internal gauge symmetries.
A.2 Canonical Quantisation of the Displacement Field A.2.1 Classical Preliminaries The classical displacement field
describes the elastic deformation of the four-dimensional membrane. Its dynamics are derived from a Lagrangian density that incorporates higher-order spatial derivatives to capture both bending and ultraviolet (UV) regularisation. A representative Lagrangian density is
where:
is the effective mass density,
is the membrane tension, entering as
in the Lagrangian. It penalises large-scale deformations, controls the infrared dispersion
, and, via spatial variations in the in-phase/out-of-phase modes, gives rise to the compensating gauge connection required by local phase invariance,
is the scale-dependent baseline elastic modulus,
represents local stiffness variations, The term
yields, via integration by parts, the sixth-order term
,
is the potential energy (e.g.
or more complex forms incorporating nonlinearities such as
),
includes additional interaction terms such as the Yukawa-like coupling
. Damping (
) and external forcing
are introduced separately or via effective dissipation functionals in the complete equation of motion:
A.2.2 Conjugate Momentum The conjugate momentum is defined as
A.2.3 Promotion to Operators Starting from the classical elastic displacement field
and its conjugate momentum
we construct the corresponding quantum theory via a fully deterministic procedure:
Hilbert-Space Domain and Self-Adjointness For physically meaningful observables, and must be self-adjoint operators. One must therefore specify a dense domain (e.g.\ a suitable Sobolev space) on which these operators act. This requirement ensures real eigenvalues and a lower-bounded Hamiltonian spectrum.
-
Determinism of the Mapping
- –
The promotion itself (steps 1–3) is a purely deterministic, one-to-one mapping from the classical phase space to the quantum operator algebra.
- –
Probabilistic outcomes arise only when applying the Born rule during measurements, not from the quantisation prescription.
Unitary Time Evolution Once operators and commutators are fixed, the Hamiltonian operator
A.2.4 Normal Mode Expansion and Dispersion Relation In a near-homogeneous region, the operator
is expressed in momentum space as
Substituting this into the linearised equation of motion
and seeking plane-wave solutions
gives
Here the tension term governs the infrared dispersion (small
), the quartic term encodes bending rigidity, and the sextic term provides ultraviolet regularisation by strongly suppressing high-wavenumber modes.
A.2.5 Hamiltonian Operator Starting from the Lagrangian density
the canonical momentum is
and the Hamiltonian density becomes
where
contains Yukawa and other interaction terms. The full Hamiltonian operator is
By choosing the domain of
to be the Sobolev space
(or higher) and imposing appropriate boundary conditions (for example, fields vanishing at infinity or Dirichlet/Neumann on a finite domain), each differential operator is rendered symmetric under integration by parts. Consequently
is self-adjoint, its spectrum is real and bounded from below, and no Ostrogradsky ghosts appear in the effective low-energy theory.
A.3 Bimodal Decomposition and Spinor Construction To capture additional internal degrees of freedom, we introduce a complementary field
, interpreted as the out-of-phase (or quadrature) component of the membrane’s displacement. We define two new real fields via the linear combinations
These represent the in-phase and out-of-phase components, respectively. They are then combined into a two-component spinor operator
By imposing appropriate (anti)commutation relations between
and
, one can demonstrate—by analogy with Fermi–Bose mappings in certain lower-dimensional systems—that the spinor
exhibits chiral substructures. These substructures are essential for the emergence of internal gauge symmetries.
A.4 Self-Adjointness and Path Integral Formulation The Hamiltonian operator
is shown to be self-adjoint by verifying that all higher-order derivative terms are well defined on the chosen Sobolev space (here,
or higher) and by imposing suitable boundary conditions (e.g. fields vanishing at infinity). This self-adjointness is essential for ensuring a real energy spectrum and the stability of the quantised theory. A complete path integral formulation can then be constructed. The transition amplitude between field configurations is given by
with the action
Integrating out the momentum degrees of freedom yields the configuration-space path integral, which serves as the basis for further extensions, including the incorporation of gauge fields.
A.5 Extended Path Integral for Gauge Fields To incorporate internal gauge symmetries, we augment the effective action with gauge field contributions. For a gauge field
(where
a indexes the generators), the covariant derivative is defined as
with
representing the generators (for example,
for SU(2) or
for SU(3)) and
g the gauge coupling constant. The corresponding field strength tensor is given by
The gauge symmetry is quantised by imposing a gauge-fixing condition (e.g. the Lorentz gauge
) and by introducing Faddeev–Popov ghost fields
and
. The resulting gauge-fixed path integral is
where
includes the original STM Lagrangian, the gauge field Lagrangian, and the ghost contributions.
A.6 Ontological meaning of the bimodal spinor This appendix clarifies the physical interpretation and underlying ontology of the two-component spinor
employed in the STM model, explaining its emergence directly from the dynamics of a four-dimensional elastic spacetime membrane.
A.6.1 Spinor Definition and Physical Interpretation In the STM framework, the fundamental spinor field is explicitly constructed from two measurable elastic deformation modes of the spacetime membrane. We define the spinor as:
where
u and
represent orthogonal displacements of the membrane. Each component is physically real and measurable:
In-phase mode: Represents a local patch of the membrane moving synchronously ("up and down") with the bulk spacetime background deformation.
Quadrature (out-of-phase) mode: Represents the same local patch moving with a 90° phase lag, achieving its maximum displacement precisely when the in-phase component is at zero displacement.
Together, these two components form a classical standing-wave system analogous to the two orthogonal polarisations of electromagnetic waves in a cavity. Crucially, the indivisibility of these modes—no local perturbation can excite one mode independently without affecting the other—is the fundamental elastic origin of quantum spin-½ behaviour.
A.6.2 Local Gauge Phase and Emergent Electromagnetism The spinor supports a local gauge invariance expressed through a point-wise phase transformation:
This gauge transformation corresponds physically to a local rotation of the oscillation ellipse formed by and . To ensure that physical predictions remain invariant under such local rotations, an additional compensating field (gauge connection) naturally emerges, identifiable with the electromagnetic potential. Hence, gauge symmetry in the STM model has a direct and intuitive geometric-elastic meaning. A.6.3 Hidden Elastic Variables and Deterministic Origin At a microscopic level, the instantaneous configuration of the bimodal spinor is entirely determined by the underlying displacement and velocity fields of the membrane. Consequently, the STM model maintains strict determinism—its quantum-like behaviour emerges only through coarse-graining and ensemble averaging. The macroscopically observable quantum spinor thus encodes only the envelope amplitude and relative phase, masking the deterministic hidden variables of the underlying elastic fields. A.6.4 Spin Encoding and the Bloch Sphere Choosing a particular quantisation axis (e.g., along the -direction), spin-up and spin-down states correspond explicitly to membrane oscillation ellipse orientations:
Intermediate orientations of the ellipse naturally map onto the continuum of quantum states represented by points on the standard quantum Bloch sphere. A.6.5 Measurement as Boundary-Condition Selection In the STM interpretation, quantum measurement is fundamentally a boundary-condition selection process. For instance, a Stern–Gerlach analyser temporarily modifies local boundary conditions—specifically altering local stiffness and membrane boundary dynamics—so that only oscillation ellipses with particular orientations can pass through. Thus, measurement outcomes reveal pre-existing elliptical orientations encoded at emission, consistent with a deterministic hidden-variable interpretation, rather than spontaneously creating measurement outcomes upon observation. A.7 Summary and Outlook In summary, the operator quantisation scheme for the STM model proceeds as follows: Displacement Field Promotion:
The classical displacement field and its conjugate momentum are promoted to operators and on a Hilbert space. The domain is chosen as a suitable Sobolev space (e.g. or higher) to ensure that all derivatives up to third order (which produce the term) are well defined. Complementary Field and Spinor Construction:
A complementary field is introduced. By forming the in-phase and out-of-phase combinations and , a two-component spinor is constructed. This spinor structure is central to the emergence of internal gauge symmetries. Self-Adjoint Hamiltonian:
The Hamiltonian includes kinetic, fourth-order, and sixth-order spatial derivatives, along with potential and interaction terms. It is shown to be self-adjoint under appropriate boundary conditions, ensuring a real and bounded-below energy spectrum. Path Integral Formulation:
A configuration-space path integral is derived from the action , serving as the basis for calculating transition amplitudes and for extending the formulation to include gauge fields and ghost terms. This comprehensive operator formalism provides a robust foundation for the STM model’s quantum framework, opening the door to further theoretical investigations and experimental tests of how deterministic elasticity can give rise to quantum-like behaviour.
Appendix B. Derivation of the STM Elastic-Wave Equation and
External
Force
This appendix provides an explicit, yet compact, route from a covariant elasticity energy functional to the second-, fourth-, and sixth-order terms, the nonlinear self-interaction, the Yukawa-like coupling, and the damping force that together define the Space-Time Membrane (STM) partial differential equation (PDE). Every algebraic step needed for independent reconstruction is shown, but purely repetitious index contractions have been suppressed for brevity. In this appendix we derive the governing PDE of the Space–Time Membrane (STM) model, showing how each term arises from clear physical reasoning and mathematical necessity. B.1 Physical Foundations The STM model treats spacetime as a four-dimensional elastic continuum whose local deformation is described by a scalar displacement field . This picture emerged from attempts to reconcile gravitational anomalies, matter–antimatter symmetry, and quantum interference within a single deterministic framework. B.2 Classical Elastic Wave Equation via Newton’s Law for Continuous Media
We start from Newton’s second law in continuum form, including an external force density
:
where the
stress is proportional to the
strain :
In the long-wavelength (Newtonian) limit we set
, so that
To capture short-scale curvature corrections, we include a bending stress proportional to
. In variational form this adds a term
. Altogether,
Thus the STM model recovers general relativity at leading order while naturally extending it into wave- and nonlinear regimes.
B.3 Cubic Nonlinearity (Self-interaction)
To reflect physically realistic scenarios—particularly the formation of stable solitonic structures (interpreted historically as standing-wave black hole cores)—a nonlinear interaction is necessary. A standard choice, both historically motivated by nonlinear elastic materials and essential mathematically to stabilise and confine energy, is a cubic (Kerr-like) self-interaction:
The Euler–Lagrange variation thus introduces a cubic nonlinear term:
This nonlinearity stabilises finite-amplitude solutions, enabling soliton formation, a historically significant step in resolving singularities and black-hole interiors.
B.4 Higher-Order Regularisation
Historically, persistent instabilities at short wavelengths motivated introducing an additional higher-order stabilisation term—specifically, a sixth-order spatial derivative
:
This term suppresses ultraviolet instabilities and divergences, mathematically regularising high-frequency modes, enabling stable numerical integration, and physically modelling the suppression of curvature singularities at sub-Planck scales.
B.5 Energy-Dependent Elasticity
Motivated historically by quantum double-slit interference phenomena, we introduced a position- and amplitude-dependent elastic modulus. Explicitly, the stiffness becomes energy-dependent:
Physically, represents feedback from local energy-density fluctuations. A minimal form capturing this behaviour at linear order is , introducing a nonlinear modulation of stiffness proportional to wave amplitude squared, ensuring persistent coherent wave structures analogous to quantum wavefunctions.
Here, is a phenomenological nonlinear-stiffness coefficient (units ) that controls the strength of feedback from the local displacement amplitude back into the membrane’s elastic modulus. Since in the undamped STM PDE of B.2–B.4 we replace by the constant dark-energy density (so that the term remains linear in u), does not appear among the six fixed coefficients in the PDE. Its numerical value can only be fixed once a reference amplitude for u is chosen, and so is left unspecified in Appendix K.7.
In the PDE this appears as
Full parameter anchoring is detailed in Appendix K.7.
B.6 Damping and Deterministic Decoherence
Although early simulations suggested damping may be optional, rigorous analysis (
Section 3.4 of main text) shows explicit damping is essential for deterministic collapse of wavefunctions and decoherence. Introducing a viscous damping term into the Lagrangian (via a Rayleigh dissipation functional) yields a term proportional to velocity:
Thus, the PDE includes:
enabling the deterministic emergence of measurement outcomes and quantum probabilities.
B.7 Spinor and Gauge-Field Couplings
(see also Appendix A & Appendix M)
A key feature of STM is that a
bimodal decomposition of the displacement field
u into two complementary oscillatory components naturally yields a two-component spinor,
Appendix A provides the full operator formalism and spinor-field construction, showing how
emerges when one splits fast and slow modes, and how canonical commutation relations follow from the membrane’s symplectic structure.
Enforcing
local phase invariance requires introducing a U(1) gauge connection
via the covariant derivative
which yields the familiar Maxwell field strength
.\ Extending this to non-Abelian rotations on the two- and three-dimensional internal mode spaces produces the SU(2) and SU(3) gauge fields
and
, respectively. In STM these gauge connections correspond physically to elastic “twists” or “shears” in the membrane’s internal oscillator basis.
Beyond gauge fields, spinor–mirror-spinor dynamics play a crucial role in energy exchange with the membrane (see Appendix M for the full curved-spacetime derivation). In brief:
Attractive couplings between a spinor on our “face” of the membrane and its mirror antispinor on the opposite face generate localised curvature outside the membrane, drawing elastic energy out of the membrane bulk and into the surrounding spacetime geometry.
Conversely, when spinors and mirror spinors repel or cancel, they relieve spacetime curvature and push energy back into the membrane, accounting for particle–antiparticle annihilation events as elastic energy deposition into the membrane substrate.
Pair production operates in reverse: local energy deposits in the membrane can spontaneously “pop” into spinor–mirror-spinor pairs, reducing the membrane’s stored energy and curving the external spacetime accordingly.
This dynamic, bidirectional energy exchange is encoded in the full covariant action (Appendix M), where the spinor–antispinor stress–energy tensor appears as a source in the Einstein-like equations, and the membrane’s elastic energy density appears in the spinor field equations.
To couple the membrane displacement
u directly to these emergent spinors, we introduce a
Yukawa-like interaction in the Lagrangian:
where
. This term encodes deterministic interactions between fermionic excitations and spacetime geometry, underpinning mass generation, CP-violating phases, and flavour dynamics within the STM framework.
B.8 Complete Lagrangian and Final PDE
Collecting all terms, the full STM Lagrangian density becomes:
Applying the Euler–Lagrange equation:
yields the PDE:
Then subtract the RHS to form
Substitute
and
, and multiply by
to adopt the conventional sign order and yielding the final PDE;
This PDE encapsulates all conservative elastic terms, damping, nonlinearity, spinor coupling, and external forcing used throughout the main text and Appendices D–H.
B.9 Summary of Mathematical Terms and Physical Roles
| PDE Term |
Physical Role |
|
Inertial
(kinetic) response |
|
Newtonian gravitational limit (Poisson’s
equation) |
|
Short-scale curvature regularisation with energy-dependent stiffness |
|
Ultraviolet divergence suppression |
|
Deterministic decoherence (measurement
collapse) |
|
Nonlinear self-interaction stabilising
finite-amplitude solitons |
|
Yukawa-like
coupling to emergent spinor field (deterministic gauge interactions) |
|
External forcing or boundary effects |
Appendix C. Gauge symmetry emergence and CP
violation
C.1 Overview
The Space–Time Membrane (STM) model naturally gives rise to internal gauge symmetries through underlying high-order elasticity
which carries a second-order tension operator
By performing a bimodal decomposition of the displacement field (as described in Appendix A), a two-component spinor is obtained. The internal structure of allows for local phase invariance, which necessitates the introduction of gauge fields. In this appendix, we derive the gauge structures corresponding to U(1), SU(2), and SU(3), including the construction of covariant derivatives, the formulation of field strength tensors, and the implementation of gauge fixing via the Faddeev–Popov procedure.
C.2 U(1) Gauge Symmetry
Local Phase Transformation and Covariant Derivative:
Consider the two-component spinor
derived from the bimodal decomposition. A local U(1) phase transformation is given by:
where
is an arbitrary smooth function. To maintain invariance of the kinetic term in the Lagrangian, we replace the ordinary derivative with a covariant derivative defined by:
where
is the U(1) gauge field and
e is the gauge coupling constant.
Field Strength Tensor:
The corresponding U(1) field strength tensor is defined as:
Under the gauge transformation,
the field strength tensor
remains invariant.
Gauge Fixing and Ghost Fields:
For quantisation, it is necessary to fix the gauge. A common choice is the Lorentz gauge, . The Faddeev–Popov procedure is then employed to introduce ghost fields and that ensure proper treatment of gauge redundancy in the path integral formulation.
C.3 SU(2) Gauge Symmetry
Local SU(2) Transformation:
Assume that the spinor
exhibits a chiral structure such that its left-handed component,
, transforms as a doublet under SU(2). A local SU(2) transformation is expressed as:
where
with
(
) being the Pauli matrices, and
representing the local transformation parameters.
Covariant Derivative for SU(2):
To maintain invariance under this transformation, the covariant derivative is defined as:
where
are the SU(2) gauge fields and
is the SU(2) coupling constant.
Field Strength Tensor for SU(2):
The field strength tensor associated with the SU(2) gauge fields is given by:
where
are the antisymmetric structure constants of SU(2).
Gauge Fixing:
Imposing the Lorentz gauge,
, and applying the Faddeev–Popov procedure, ghost fields
and
are introduced with a ghost Lagrangian of the form:
C.3.1 Electroweak Mixing, theZBoson, and CP Violation via Zitterbewegung
In the STM framework, electroweak symmetry breaking and the emergence of the neutral Z boson can be naturally explained through interactions between the bimodal spinor field residing on one face of the membrane and the corresponding bimodal antispinor field located on the opposite face (the "mirror universe").
Specifically, the displacement field
couples these spinor fields through an interaction Lagrangian of the form:
where:
represents Yukawa-like coupling constants between generations .
is the membrane displacement field, whose vacuum expectation value (VEV), , generates effective fermion masses.
Complex phase shifts arise naturally due to rapid oscillatory interactions—known as zitterbewegung—between the spinor and the mirror antispinor .
When the displacement field
acquires a vacuum expectation value (VEV), denoted
, this interaction yields an effective fermion mass matrix of the form:
where the phases
become averaged into constant effective phases
upon coarse-graining.
Electroweak Mixing and Emergence of theZBoson:
To clearly illustrate the connection with electroweak theory, consider the gauge fields emerging from the bimodal spinor structure. Initially, the theory features separate U(1) and SU(2) gauge symmetries, represented by gauge fields
(U(1)) and
(SU(2)). Through the process described above—where the membrane’s displacement field acquires a vacuum expectation value
—mass terms arise for specific gauge bosons. Explicitly, electroweak mixing occurs via a linear combination of the neutral gauge fields
(from SU(2)) and
(from U(1)):
where
is the Weinberg angle, dynamically determined by membrane parameters, and
is the original U(1) gauge field. The gauge boson corresponding to the
acquires mass directly from the membrane’s elastic structure, analogous to the conventional Higgs mechanism but derived here entirely from deterministic elastic interactions rather than from an additional scalar field.
Finally, note that the gauge-boson kinetic terms inherited from the elastic action carry this same prefactor, so the effective Weinberg-angle mixing and the relative normalisation of the photon and Z kinetic terms are now functions of and .
Emergence of CP Violation:
Under a combined charge conjugation–parity (CP) transformation, the spinor fields transform approximately as:
with analogous transformations applied to the mirror antispinor
. Due to the presence of nontrivial phases induced by the zitterbewegung interaction between spinor and antispinor fields, the effective fermion mass matrix
is generally complex. Diagonalising this matrix yields physical fermion states with mixing angles and phases analogous to the experimentally observed CKM matrix, thus naturally introducing CP violation into the STM framework.
Summary:
Gauge boson masses and electroweak mixing angles emerge naturally via vacuum expectation values of the membrane displacement field.
Z bosons arise explicitly from the SU(2) × U(1) gauge field mixing.
CP violation is introduced through the deterministic zitterbewegung interaction between spinors and antispinors across the membrane, producing effective Yukawa couplings with nonzero complex phases.
A rigorous derivation of chiral anomalies and electroweak parity violation still demands an explicit triangular-loop calculation within the STM framework.
In sections 3.1.4 and Appendix R, we have derived the effective light-neutrino mass matrix via the minimal see-saw mechanism, built upon the bimodal spinor–antispinor formalism developed in Appendices C and N. Utilising the zitterbewegung-induced mass terms, we recover the form of both the CKM and PMNS mixing matrices and extract all three mixing angles along with the corresponding Jarlskog invariants. While absolute neutrino masses are not determined at this stage, the model reproduces the observed neutrino mass-splitting pattern alongside the quark-sector mass-splitting hierarchy
C.4 SU(3) Gauge Symmetry
Local SU(3) Transformation:
For the strong interaction, the spinor
is assumed to carry a colour index and transform as a triplet under SU(3). A local SU(3) transformation is given by:
with
where
(
) are the Gell–Mann matrices, and
are the transformation parameters.
Covariant Derivative for SU(3):
The covariant derivative is defined as:
where
are the SU(3) gauge fields and
is the SU(3) coupling constant.
Field Strength Tensor for SU(3):
The SU(3) field strength tensor is defined by:
where
are the structure constants of SU(3).
Gauge Fixing:
The Lorentz gauge
is imposed, and ghost fields
and
are introduced via the Faddeev–Popov procedure. The ghost Lagrangian is then:
C.4.1 Physical Interpretation — Linked Oscillators and Confinement:
In the main text (
Section 3.1.2), the strong force is depicted by analogy with a “linked oscillator” network, wherein each local site carries a colour-like degree of freedom. From the perspective of continuum gauge theory, this classical picture emerges naturally once we require that
carry a local SU(3) index and that neighbouring “sites” (or regions) remain elastically coupled under deformations. In essence, each SU(3) gauge connection
plays the role of an “elastic link” constraining colour charges, which becomes increasingly stiff (i.e. confining) with separation.
Mathematically, the field strength
enforces local colour gauge invariance, just as tension in a chain of coupled oscillators enforces synchronous motion. When two colour charges are pulled apart, the membrane’s elastic energy—now interpreted as the non-Abelian gauge field energy—rises linearly with distance (up to corrections from real or virtual gluon-like modes). This provides a deterministic analogue of confinement: it is energetically unfavourable for a single “coloured oscillator” to exist in isolation, so colour remains bound. Thus, the formal gauge-theoretic description of SU(3) in this appendix and the intuitive “linked oscillator” analogy of
Section 3.1.2 are two views of the same phenomenon: a deterministic continuum mechanism underpinning the strong interaction.
C.4.2 Derivation of SU(3) Colour Symmetry
In the STM model, spacetime is described as an elastic four-dimensional membrane whose displacement field,
, obeys a high-order partial differential equation:
where
is the effective mass density,
is a scale-dependent elastic modulus,
accounts for local variations in stiffness, and
controls the higher-order spatial derivative terms that serve to regularise ultraviolet divergences.
Plane-wave dispersion: If we try a solution
, the PDE immediately yields
In particular, the term sets the low-k (infrared) “speed” for all three emergent colour-oscillator modes.
At sub-Planck scales, the membrane exhibits rapid deterministic oscillations. Coarse-graining these fast modes yields a slowly varying envelope. Initially, the displacement field is decomposed bimodally:
which can be combined into a two-component spinor,
This spinor naturally exhibits a U(1) symmetry under local phase rotations. However, the strong interaction is described by an SU(3) symmetry, necessitating an extension to three internal degrees of freedom.
Extending to Three Components
The inclusion of higher-order derivative terms (
and
) implies a richer dynamical structure than a simple two-mode system. For example, in a one-dimensional analogue, an equation such as
yields a dispersion relation
that supports a multiplicity of normal modes. In four dimensions, such higher-order dynamics may naturally allow for three distinct, independent oscillatory modes. Label these as
,
, and
(metaphorically corresponding to “red”, “green”, and “blue”). Then the displacement field may be expressed as:
which is recast as a three-component field,
This field now naturally transforms under SU(3) via unitary matrices with determinant 1, preserving the norm .
Anomaly Cancellation and Topological Constraints
A consistent, anomaly-free gauge theory requires that the contributions from all fields cancel potential gauge anomalies. In the Standard Model, the colour triplet structure of quarks ensures anomaly cancellation within QCD. In the STM model, if the three vibrational modes couple to emergent fermionic degrees of freedom analogously to quark fields, then both energy minimisation and anomaly cancellation considerations naturally favour an SU(3) symmetry. Moreover, topological constraints—for instance, those imposed by suitable boundary conditions or by a compactified membrane geometry—can enforce the existence of exactly three independent, stable oscillatory modes.
Thus, by extending the initial bimodal decomposition to include additional degrees of freedom arising from higher-order elastic dynamics, the STM model naturally leads to a three-component field. This field, transforming under SU(3), provides a first-principles, deterministic explanation for the emergence of three colours. Such a derivation not only aligns with the phenomenology of QCD but also reinforces the unified, classical elastic framework of the STM model.
C.5 Prototype Emergent Gauge Lagrangian
While we have described how local phase invariance of our bimodal spinor
induces gauge fields
, we can also hypothesise a Yang–Mills-like action arising at low energies (See
Figure 7):
where
In the STM context, this term would emerge from an effective elasticity-based action once the short-wavelength excitations are integrated out and the spinor fields become nontrivial.
C.6 Summary
In summary, the internal structure of the two-component spinor (derived from the bimodal decomposition of ) leads naturally to local gauge invariance. Enforcing invariance under local U(1) transformations necessitates the introduction of a U(1) gauge field with covariant derivative and field strength . Extending this to non-Abelian symmetries, local SU(2) and SU(3) transformations require the introduction of gauge fields and , respectively, with covariant derivatives defined accordingly. Gauge fixing, typically via the Lorentz gauge, is implemented using the Faddeev–Popov procedure, ensuring a consistent quantisation of the gauge degrees of freedom.
Appendix D. Derivation of the Effective Schrödinger-Like
Equation, Interference, and Deterministic Quantum
Features
D.1 Introduction
This appendix supplies the complete multiple-scale (WKB-type) derivation by which the deterministic Space–Time Membrane (STM) wave equation yields, after coarse-graining, an effective non-relativistic “Schrödinger-like’ ’ evolution law for the slowly varying envelope of the membrane displacement. All intermediate steps are retained, and the next-order (diffusive) corrections—needed for quantitative tests of damping and fringe deformation—are displayed explicitly in terms of the microscopic STM parameters.
D.2 The STM Membrane PDE (one spatial dimension)
The linearised STM PDE take the form;
where
is the effective mass density,
T is the tension coefficient,
is the baseline elastic modulus,
is a slowly varying stiffness modulation,
regularises ultraviolet modes,
is a small damping,
“…” denotes neglected nonlinear or spinor/gauge couplings.
is the scalar (membrane) damping constant. The milder spinor dephasing rateaffects the Dirac-like spinor equations and therefore does not enter the envelope derivation presented here.
D.3 Carrier + Envelope Ansatz and Coarse-Graining
A Gaussian filter
ensures
U varies only on scales
. Derivatives expand as
D.4 Expansion of Derivatives Acting on :
D.5 Substitution and Order-by-Order Balance Insert these into the PDE, divide by , and collect powers of :
D.6 Next-Order Envelope Equation Solving (D.5.3) for
yields the effective Schrödinger-like law,
with STM-parameter expressions
Here satisfies (D.5.2) and solves (D.5.1). In the conservative limit , reproduces , while a small yields deterministic envelope damping via .
D.7 Summary
The leading-order multiple-scale expansion delivers a free-particle Schrödinger equation for the coarse-grained envelope U.
Equation (D.6.1) incorporates next-order damping () and dispersion () in closed form, directly in terms of .
The tension T enters both the carrier dispersion relation (D.5.1) and the diffusion coefficient , modifying effective mass and fringe spacing.
D.8 Physical Interpretation and Onward Links
Coherent quantum-like envelope. The Gaussian filter ensures captures only slow modes; with it propagates exactly like a wavefunction in non-relativistic quantum mechanics, while induces deterministic decoherence.
Born-rule density. Positivity and normalisation of the filter imply obeys a continuity equation to leading order. Appendix E shows how tracing out environmental modes endows P with the standard probabilistic interpretation.
Interference and deterministic collapse. The real part of sets fringe spacing in double-slit analogues; governs contrast loss. The non-Markovian master-equation in Appendix G detail these phenomena.
Parameter sensitivity. Equations (D.5.2)–(D.6.2) tie fringe shifts and damping times directly to . Appendix K uses these to calibrate finite-element simulations against experiments.
Readers interested in entanglement and Bell-inequality violations should proceed to Appendix E; for the cosmological impact of persistent envelopes see Appendix H.
Appendix E. Deterministic Quantum Entanglement and Bell
Inequality
Analysis
E.1 Overview
In the Space–Time Membrane (STM) model the fully deterministic membrane dynamics produce, after coarse-graining, an effective wavefunction that contains non-factorisable correlations. These reproduce the empirical signatures of quantum entanglement even though the underlying evolution is strictly classical. In this appendix we (i) show how such correlated global modes arise, (ii) demonstrate how a simple projection rule at a Stern–Gerlach detector yields the familiar statistics, and (iii) verify that a standard CHSH test exceeds the classical bound.
E.2 Formation of a non-factorisable global mode
Consider two localised excitations on the membrane,
and
. The full displacement field is
with the interaction term
where
is an elastic coupling constant. After Gaussian coarse-graining (Appendix D) the effective state becomes
Because the argument is a genuinely mixed function of and , the state cannot be factorised into ; consequently the two regions are correlated exactly as in standard entanglement.
E.3 Overlap derivation of thelaw
E.3.1 A singlet-like standing wave
Pair creation leaves the membrane in a single global standing-wave packet
where each single-packet field is
The “spin-up” or “spin-down” label is encoded in the internal phase between the two elastic modes and .
E.3.2 Local basis rotation by a Stern–Gerlach magnet
A Stern–Gerlach magnet set at angle
mixes the two modes via
E.3.3 Projection amplitudes
The incoming phase vector
is projected onto the magnet’s eigen-vectors
and
:
E.3.4 Deterministic routing rule
Energy flows into the branch whose instantaneous amplitude is larger, so
Thus the usual detection statistics arise purely from geometric overlap—no intrinsic randomness is required.
Consequently, the pair
follows the deterministic routing
selecting one of the two Bell branches. This branch-selection mechanism is exactly the small-
damping analysed in
Section 3.4: a strictly positive yet Planck-time-scale
turns the routing into an attractor process that yields Born-rule weights.
The small, flavour-sector damping constantis fixed using(see Section 3.4.1); it acts on the Dirac-like evolution ofbetween beam-splitter events but does not modify the geometric overlap amplitudesor the CHSH correlation
.
E.3.5 Joint expectation value
Because the global standing wave enforces the opposite internal phase on the right-hand packet, the joint correlation for magnet settings
a and
b is
exactly matching quantum-mechanical predictions and reaching the Tsirelson value
in a CHSH test.
E.3.6 Photon entanglement
Exactly the same construction applies to polarisation-entangled photons: here the two-component spinor corresponds to the horizontal/vertical membrane sub-modes, and the operator represents a linear polariser set at angle . The resulting correlation function reproduces the standard photonic Bell-test sinusoid
E.4 Measurement Operators and Correlation Functions
To quantitatively probe the entanglement, we introduce measurement operators analogous to those used in quantum mechanics. Assume that the effective state (obtained after coarse-graining) lives in a Hilbert space that can be partitioned into two subsystems corresponding to regions A and B.
For each subsystem, define a spinor-based measurement operator:
where
and
are the Pauli matrices and
is a measurement angle. For subsystems A and B, we denote the operators as
and
, respectively.
The joint correlation function for measurements performed at angles
and
is then given by:
This expectation value is calculated by integrating over the coarse-grained degrees of freedom, taking into account the non-factorisable structure of .
E.5 Detailed CHSH Parameter Calculation
Recalling that the deterministic routing rule of E.3.4 (which—via the tiny, Planck-time-scale damping analysed in
Section 3.4—yields Born-rule probabilities) is in force, we now compute the detailed CHSH parameter as follows;
The CHSH inequality involves four correlation functions corresponding to two measurement settings per subsystem. Define the CHSH parameter as:
A detailed derivation involves the following steps:
State Decomposition:
Express
in a basis where the measurement operators act naturally (e.g. a Schmidt decomposition). Although the state arises deterministically from the coarse-graining process, its non-factorisable nature allows for a decomposition of the form:
where
are effective coefficients that encode the correlations.
Evaluation of:
With the measurement operators defined as above, compute the joint expectation value:
The explicit dependence on the measurement angles enters through the matrix elements of the Pauli matrices.
Optimisation:
Choose measurement angles
to maximise
S. Standard quantum mechanical analysis shows that the optimal settings are typically:
With these settings, the CHSH parameter can be shown to reach:
Interpretation:
The fact that S exceeds the classical bound of 2 is indicative of entanglement. In our deterministic STM framework, this violation emerges from the inherent non-factorisability of the effective state after coarse-graining, despite the absence of any intrinsic randomness.
E.6 Off-Diagonal Elements as Classical Correlations
Within the STM model, the effective density matrix is constructed from the coarse-grained displacement field emerging from the underlying deterministic PDE. In conventional quantum mechanics, the off-diagonal matrix elements (or “coherences”) are interpreted as evidence that a particle has simultaneous amplitudes for distinct paths. In STM, however, these off-diagonals are reinterpreted as a measure of the classical cross-correlations among the sub-Planck oscillations of the membrane.
Specifically, if one considers the effective state formed by the overlapping wavefronts from, say, two slits, the element in the density matrix quantifies the overlap between the states and , which are not distinct quantum paths but rather the coherent classical waves generated by the membrane. When the environment or a measurement apparatus perturbs the membrane, these classical correlations decay, resulting in the vanishing of the off-diagonal elements. Thus, the “collapse” of the effective density matrix is interpreted not as an ontological disappearance of superposition but as a deterministic loss of coherence among real, classical wave modes.
This reinterpretation not only reproduces the standard interference patterns and entanglement correlations—such as those responsible for the violation of Bell’s inequalities—but also demystifies the process by replacing probabilistic superposition with measurable, deterministic wave interference.
E.7 Summary
The effective wavefunction obtained from the deterministic dynamics is non-factorisable due to the coupling term .
Spinor-based measurement operators are defined to emulate quantum measurements.
The correlation functions computed from these operators lead to a CHSH parameter S that, under optimal settings, reaches , thereby violating the classical bound and reproducing the quantum mechanical prediction.
This deterministic entanglement analysis augments the Schrödinger-like interference picture (Appendix D) and sets the stage for further results on decoherence (Appendix G) and black hole collapse (Appendix F)—all approached through an elasticity-based, sub-Planck wave interpretation in the STM framework.
Appendix F. Singularity Prevention in Black
Holes
F.1 Overview
Modern physics typically predicts that gravitational collapse leads to spacetime singularities under General Relativity. In the Space–Time Membrane (STM) model, higher-order elasticity terms—particularly a second-order “tension” operator alongside the operator—regulate both infrared and ultraviolet modes. This combined stiffness mechanism effectively avoids the formation of infinite curvature. Instead of a singularity, the interior relaxes into a finite-amplitude wave or solitonic core. This appendix first outlines how singularity avoidance occurs, then Section F.7 discusses routes toward black-hole thermodynamics within STM.
F.2 STM PDE and Local Stiffening
The STM model’s master PDE often appears in schematic form:
where:
is an effective mass density for the membrane,
T is the tension coefficient penalising large-scale deformations,
is the scale-dependent bending modulus,
imposes a strong penalty on high-wavenumber modes,
introduces damping or friction,
is a nonlinear self-interaction.
As matter density grows in a collapsing region, the local stiffening surges, and the tension term resists large-scale contraction, making further inward collapse energetically prohibitive.
F.3 Role of theTerm
he STM equation includes a sixth-order spatial derivative term,
, which is crucial for ultraviolet regularisation. In configuration space, this term directly penalises short-wavelength deformations. In momentum space, the propagator for
becomes
so that at high momentum the
contribution dominates, ensuring loop integrals remain finite. At low
k, the added
term softens infrared modes and helps prevent large-scale collapse. Consequently, when simulating gravitational collapse, rather than evolving towards a singularity, the system relaxes into a stable configuration characterised by finite-amplitude standing waves. These standing waves manifest as solitonic configurations—localised, finite-energy solutions that effectively replace the classical singularity with a “soft core” in which energy is redistributed into stable oscillatory modes.
Detailed derivations, discussing the formation and stability of such solitons, are provided in Appendix L. This link underscores how the STM model not only circumvents the singularity problem but also lays the groundwork for exploring the thermodynamic properties of black hole interiors.
Appendix F.4 Mode Counting and Microcanonical Entropy
Large-scale numerical work (Appendix K) shows that the solitonic black-hole interior is an extremely stiff region where the displacement field remains small but experiences very high spatial gradients. In this regime the
linearised, time-independent form of the complete STM equation is appropriate. Retaining every spatial-derivative term—tension, bending and sixth-order ultraviolet stiffness—one obtains
with positive constants
. Damping, nonlinear and Yukawa terms are negligible inside the core. We now calculate the number of independent standing-wave modes in a spherical core of radius
and hence its entropy.
F.4.1 Separation of variables
For spherical symmetry (lowest angular harmonic
= 0) write
Setting in (F.4.1) yields the dispersion relation
(F.4.2)
Because all
(by construction of the elastic energy; see Appendix B) and
, (F.4.2) has three real non-negative roots:
and
each of which is strictly positive. The boundary condition
then quantises
for each independent root, giving two towers of radial modes.
F.4.2 Mode count below a physical cut-off
Let
(
is the core mass-density). Define a maximum frequency
where linear theory ceases to be valid and denote the corresponding wavenumbers
. Counting all modes with
yields
Because for astrophysical cores, N grows ∝ , foreshadowing an area law.
F.4.3 Micro-canonical entropy
Assuming equipartition among the
N harmonic oscillators, the micro-canonical entropy is
where
encodes phase-space factors. Introduce the
effective horizon area (F.3) and the crossover length
. Re-expressing (F.4.6) in these terms gives
Hence the leading term exactly reproduces the Bekenstein–Hawking area law, while the full sixth-order operator introduces only suppressed corrections of relative size . Such corrections become relevant only for Planck-scale remnants.
F.4.4 Implications and onward links
The term—vital for singularity avoidance—does not spoil the entropy–area relationship for macroscopic black holes; it merely adds tiny, testable corrections.
Section F.5 discusses how the standing-wave interior implied by (F.4.1) can store information without a curvature singularity.
A detailed treatment of sub-leading terms is deferred to the task list in F.7. Possible logarithmic and power-law corrections, together with thermal stability tests, are enumerated among the outstanding tasks in F.7.
F.5 Implications for the Black Hole Information
Because the PDE remains well-defined (and in principle deterministic) for all times, the usual scenario of a “lost” interior or singular region is avoided. The interior’s standing wave can store or reflect quantum-like information, subject to additional couplings (e.g., spinors, gauge fields). However, how that information might be released back out remains linked to black hole thermodynamics—an ongoing focus described below.
F.6 Summary of Singularity Avoidance
The tension term halts large-scale collapse.
Higher-order elasticity (especially ) halts runaway collapse.
Local stiffening near high density further resists infinite curvature.
Numerical PDE solutions show stable wave or solitonic cores, not a singularity (because the STM modulus never exceeds , strains are capped and the would-be singularity is replaced by a finite-amplitude solitonic core once regularisation and tension dominate).
F.7 Outstanding Thermodynamic Tasks
Sections F.2 – F.6 establish that combined tension and higher-order elasticity prevent singularities. Appendices G and H supply initial analytic ingredients for STM black-hole thermodynamics. Remaining tasks include (cf. leading-order results in F.4).
F.7.1 Entropy Beyond the Solitonic Core
Context. Section F.4 reproduces the leading Bekenstein–Hawking result
by micro-canonical mode counting inside the stiff core.
Outstanding tasks.
Calculate sub-leading logarithmic and power-law corrections when full / elasticity and gauge couplings are retained.
Define an effective horizon radius (surface where outgoing low-frequency waves red-shift sharply) and verify that the dominant density of states accumulates near .
Test thermal stability: confirm that small perturbations of the solitonic interior leave the area–entropy relation intact for .
F.7.2 Hawking-Like Emission and Evaporation
Context. Appendix G.4 derives a near-thermal spectrum and grey-body factors; Appendix G.5 supplies the transmission coefficient.
Outstanding tasks.
Include non-linear mode coupling to determine whether the spectrum remains Planckian once energy loss feeds back on and on local stiffness .
Integrate the flux in time to see whether persists or halts at a remnant mass when damping is sizeable.
Quantify the influence of slow drifts , (as introduced in Appendix H.9) on late-stage evaporation.
F.7.3 Information Release and Unitarity
Correlation tracking. Evolve collapse + evaporation numerically and monitor two-point functions linking interior solitonic modes to the outgoing flux.
Page-curve test. Partition the (quantised) membrane field into interior/exterior regions and compute entanglement entropy versus time, searching for the characteristic rise-and-fall.
Spectral fingerprints. Look for phase correlations, echoes or other deviations from a perfect thermal spectrum that would evidence unitary evolution.
F.7.4 First-Law Checks and Small-Mass Behaviour
Large-mass regime. Perturb or inject spinor/gauge energy; verify that the resulting changes in total energy E, horizon temperature (from Appendix G.4) and entropy S satisfy .
Planck-scale remnants. If evaporation saturates near the stiffness cut-off, derive modified first-law terms incorporating residual elastic strain or non-Markovian damping contributions.
F.7.5 Numerical and Experimental Road-Map
Develop adaptive-mesh finite-element solvers (see Appendix K) capable of tracking the term through collapse, rebound and long-time evaporation.
Construct acoustic or optical metamaterials with tunable fourth-/sixth-order stiffness to emulate horizons and measure grey-body transmission.
Perform parameter surveys in to locate regions where area law, Hawking-like flux and a unitary Page curve coexist.
Appendix G. Non-Markovian Decoherence and
Measurement
G.1 Overview
In the Space–Time Membrane (STM) model, although the underlying dynamics are fully deterministic, the process of coarse-graining introduces effective environmental degrees of freedom that lead to decoherence. Instead of invoking intrinsic randomness, the decoherence in this model arises from the deterministic coupling between the slowly varying (system) modes and the rapidly fluctuating (environment) modes. In this appendix, we provide a detailed derivation of the non-Markovian master equation for the reduced density matrix by integrating out the environmental degrees of freedom using the Feynman–Vernon influence functional formalism. The resulting evolution includes a memory kernel that captures the finite correlation time of the environment.
G.2 Decomposition of the Displacement Field
We begin by decomposing the full displacement field
into two components:
where:
is the slowly varying, coarse-grained “system” field,
comprises the high-frequency “environment” modes (the sub-Planck fluctuations).
The coarse-graining is achieved by convolving
with a Gaussian kernel
over a spatial scale
L:
with
The environmental part is then defined as:
This separation allows us to treat as the primary degrees of freedom while regarding as the effective environment.
G.3 Derivation of the Influence Functional
In the path integral formalism, the full density matrix for the combined system (S) and environment (E) at time
is given by:
To obtain the reduced density matrix
for the system alone, we integrate out the environmental degrees of freedom:
We define the Feynman–Vernon influence functional
as:
where
denotes the interaction part of the action that couples the system to the environment.
For weak system–environment coupling, we can expand
to second order in the difference
. This yields a quadratic form for the influence action:
where
is a memory kernel that encapsulates the temporal correlations of the environmental modes. The precise form of
depends on the spectral density of the environment and the specific details of the coupling.
G.4 Effective Horizon Temperature via Fluctuation–Dissipation
The frequency-domain Green function with Rayleigh damping
and baseline membrane tension
T obeys
For long wavelengths and low frequencies
the imaginary part reduces to
The fluctuation–dissipation theorem then gives an effective horizon temperature
after identifying
. Thus the Hawking temperature is recovered, while Planck-suppressed corrections scale with the ratio
.
G.5 Grey-body Factors from Mode Overlaps
The probability for an exterior wave at frequency
to transmit through the core-horizon region is given by the squared overlap
With
and normalisation constants
, the integral evaluates to
Substituting this into the emission rate yields the full non-thermal spectrum.
G.6 Derivation of the Non-Markovian Master Equation
Starting from the reduced density matrix expressed with the influence functional:
we differentiate
with respect to time
to obtain its evolution. Standard techniques (akin to those used in the Caldeira–Leggett model) yield a master equation of the form:
where:
is the effective Hamiltonian governing the system ,
is a dissipative superoperator that typically involves commutators and anticommutators with system operators (e.g., or its conjugate momentum),
The kernel introduces memory effects; that is, the rate of change of depends on its values at earlier times.
In the limit where the environmental correlation time is very short (i.e., approximates a delta function ), the master equation reduces to the familiar Markovian (Lindblad) form. However, in the STM model the finite correlation time leads to explicitly non-Markovian dynamics.
G.7 Implications for Measurement
The non-Markovian master equation implies that when the system interacts with a macroscopic measurement device, the off-diagonal elements of the reduced density matrix decay over a finite time determined by . This gradual loss of coherence—induced by deterministic interactions with the environment—leads to an effective wavefunction collapse without any intrinsic randomness. The deterministic decoherence mechanism thus provides a consistent explanation for the measurement process within the STM framework.
G.8 Path from Influence Functional to a Non-Markovian Operator Form
We have described in Eqs. (G.3, G.7) how integrating out the high-frequency environment
produces an influence functional
with a memory kernel
. In principle, if this kernel is short-ranged, one recovers a Markov limit akin to a Lindblad master equation,
However, in our non-Markovian STM scenario, the memory kernel extends over times
. We therefore obtain an integral-differential form,
capturing the environment’s finite correlation time (See
Figure 8). Determining explicit Lindblad-like operators
from this memory kernel would require further approximations (e.g., expansions in powers of
, where
T is a characteristic system timescale).
Consequently, a direct closed-form solution of the STM decoherence rates is not currently derived. Nonetheless, numerical simulations (Appendix K) can approximate these integral kernels and predict how quickly off-diagonal elements vanish, giving testable predictions for deterministic decoherence times in metamaterial analogues.
G.9 Summary
Decomposition: The total field is decomposed into a slowly varying system component and a high-frequency environment .
Influence Functional: Integrating out yields an influence functional characterised by a memory kernel that captures the non-instantaneous response of the environment.
Master Equation: The resulting non-Markovian master equation for the reduced density matrix involves an integral over past times, reflecting the system’s dependence on its history.
Measurement: The deterministic decay of off-diagonal elements in explains the effective collapse of the wavefunction observed in quantum measurements.
Thus, the STM model demonstrates that deterministic dynamics at the sub-Planck level, when coarse-grained, can reproduce quantum-like decoherence and the apparent collapse of the wavefunction—all through non-Markovian, memory-dependent evolution of the reduced density matrix.
Appendix H. Vacuum energy dynamics and the cosmological
constant
H.1 Overview
This appendix presents a detailed multi-scale PDE derivation showing how short-scale wave excitations in the Space–Time Membrane (STM) model—including the tension term —produce a near-constant vacuum offset interpreted as dark energy. We cover:
The full STM PDE with tension and scale-dependent elasticity.
A two-scale expansion separating fast sub-Planck oscillations from slow modulations.
The solvability condition that yields an envelope equation.
The sign and damping constraints required for a non-decaying (persistent) mode.
How the resulting locked amplitude acts as a cosmological constant.
The prospect of mild late-time evolution to relieve the Hubble tension (building on the constant-offset result of Appendix M.7).
H.2 Governing PDE with Scale-Dependent Elasticity
H.2.1 Equation of Motion
In flat space, the STM membrane obeys the sixth-order PDE with tension
T, feedback
, UV-regulator
, damping
and weak nonlinearity
:
where
is the membrane’s mass density,
T penalises large-scale curvature,
is the baseline bending modulus at scale ,
encodes fast-wave feedback,
UV-regularises high-k modes,
is a small positive damping,
is a weak cubic stiffness.
H.2.2 Sub-Planck Oscillations and Scale Dependence
Fast, sub-Planck “particle-like” modes drive via renormalisation-group flows (Appendix J). When and , these modes lock in a non-zero mean . Coarse-graining over many cycles then leaves a uniform offset that appears as a cosmological constant in the emergent field equations.
H.3 Multi-Scale Expansion: Fast vs. Slow Variables
Derivatives split as , .
H.3.1 Leading Order
At
,
drop out, giving
A plane-wave ansatz
yields the dispersion relation
H.3.2 Next Order
At one collects terms involving and . Requiring no secular growth in imposes a solvability condition, which reduces to an envelope equation for the slow amplitude .
H.4 Stiffness-Feedback Locking
Model the effective bending modulus as
The envelope equation takes the schematic form
where
. Writing
and separating real parts gives
with
,
. Setting
yields
so a non-zero, persistent envelope is maintained.
H.5 Euclidean Partition Function and Evaporation Law
Wick-rotate
to obtain the Euclidean action
with
. The Gaussian mode sum yields the free energy
reproducing Hawking’s
timescale up to
corrections.
H.6 Envelope Equation and Parameter Criteria
H.6.1 Full Envelope PDE
For
one finds
with
.
H.6.2 Non-Decaying Steady State
A uniform, time-independent envelope
requires
so that the cubic nonlinearity balances any residual damping and enforces
.
H.7 Vacuum Offset and Dark Energy
H.7.1 Coarse-Graining the Persistent Wave
Once
U locks,
splits into an oscillatory part (zero mean) and a constant
. The latter is identified with the vacuum-energy density
H.7.2 Mapping to the Cosmological Term
In four-dimensional Einstein equations, a constant energy density
enters as
in exact agreement with Appendix M.7.
Inserted into this offset drives cosmic acceleration without invoking an independent dark-energy field.
H.8 Maximum STM Stiffness and Dark-Energy Smallness
The baseline modulus peaks at
Thus a tiny fractional offset automatically yields the observed vacuum density
H.9 Late-Time Evolution and Hubble Tension
Building on the constant-offset result of Appendix M.7, we now allow residual time dependence in
or
at low redshift (
). Then
offering a potential resolution of the Hubble-tension discrepancy, provided
and
remain satisfied under fixed boundary conditions.
H.10 Modifications to Traditional EFE, Time Dilation & Tests
Redshift & Time Dilation In the weak-field limit , STM modifies , inducing small anomalies in clock rates near compact or oscillating sources.
High-Frequency Damping The regulator and memory kernels suppress abrupt metric changes, shifting QNM ringdown frequencies by , potentially observable by next-generation detectors.
Local Tests –
Atomic Clocks: Precision clock comparisons may reveal departures from GR’s redshift.
Metamaterials: Laboratory analogues with tunable T can probe short-range modifications to Poisson’s equation.
H.11 Open Challenges
Ghost-Free Quantisation Proving absence of negative-norm modes for the combined , and operators.
Spinor/Gauge Self-Adjointness Ensuring well-posed boundary conditions and positive-definite norms once spinor and gauge couplings are included.
Planck-Scale Completion Bridging the continuum elasticity description to a discrete or microscopic theory at the Planck scale.
H.12 Summary
Full STM PDE: Incorporates , , , damping and nonlinearity.
Multi-Scale Expansion: Yields a dispersion relation and envelope equation with feedback.
Locking Conditions: and enforce a non-decaying amplitude.
Dark Energy: The coarse-grained plays the rôle of .
Hubble Tension: Tiny late-time drifts in or can reconcile discrepant measurements.
This deterministic elasticity framework thus unifies sub-Planck wave persistence with cosmic acceleration, while admitting minimal late-time evolution to resolve cosmological tensions.
Appendix I. Proposed Experimental
Tests
This appendix summarises feasible near-term experiments explicitly designed to test distinctive predictions of the Space–Time Membrane (STM) model, focusing on setups achievable with existing or soon-to-be-available technologies. Each experimental setup includes precise methodologies, clear STM predictions, falsification criteria, and feasibility assessments.
I.1 Reference parameters and context
The laboratory-scale experiments probe only the low-momentum tail of the STM dispersion, so we quote the dimensionless ratios that control the relevant terms in the calibrated June-2025 model:
Quartic coefficients and is its envelope analogue. A small local departure is encoded in the fluctuation . The same nondimensional stiffness appears in the carrier phase and in the envelope evolution.
Sextic regulator. For laboratory wave-numbers the associated phase shift is below and can be neglected.
Scalar damping. Governs the cross-over from algebraic mode damping (undriven membrane) to exponential decay when deliberate viscoelastic loss is introduced.
Flavour-sector damping. Appears only in spinor/CP-violation tests (Appendix E) and plays no rôle in the mechanical or optical set-ups below.
The experimental protocol is therefore:
determine by fitting the pure-tension dispersion ;
measure residual phase and envelope shifts and compare them with the quartic STM prediction fixed by (with .
I.2 Mechanical membrane interferometer (primary laboratory test)
Objective. Validate the STM quartic term by measuring the phase shift and envelope contraction of a single, low-k flexural mode on a purpose-built metamaterial proxy membrane whose long-wavelength dynamics match the STM PDE term-for-term.
| Item |
Value |
Rationale / design handle |
| Film |
Mylar (polyester) |
Stock film; baseline
tension T set during clamping. |
| Outer skin |
epoxy–silica laminate |
Raises the
bending modulus to give the target
. |
| Clear aperture |
|
X-edges
clamped, Y-edges free; . |
| Fundamental mode |
|
Mode index (1,0). |
| Drive frequency |
|
From the measured
k– fit. |
| Probe point |
(= ) |
Maximises and envelope signal. |
| Sextic handle |
1 mm “mass-on-spring’ ’ pillar array at 30 mm pitch |
Tunes the nondimensional without altering
. |
| Damping handle |
viscoelastic paint stripe |
Sets
; removable to recover the undamped limit. |
The laminate thickness is chosen so that
locking the laboratory plate to the Planck-anchored STM quartic coefficient. Pillar resonators supply the sextic regulator while leaving
unchanged; a thin damping stripe controls
.
A voice-coil shaker applies a sinusoidal moment at one clamp; the opposite free edge minimises reflections. A laser-Doppler vibrometer (10 kHz sample rate) records
with spatial precision <
and phase precision < 0.01 rad.
After subtracting the measured quadratic baseline
, the quartic propagation factor
gives
within a 60 ms integration window (= 30 drive cycles; the flexural packet itself crosses the 0.12 m path in = 2 ms). The envelope equation with
predicts an extra contraction
. Sextic effects (
) shift the phase by <
rad and are negligible at this
k.
Commercial LDVs easily meet rad and , surpassing the required sensitivity.
rules out the benchmark quartic term at high confidence.
If the quartic prediction is falsified
| Adjustment |
Effect on PDE |
Consequence elsewhere |
|
Uniform modulus rescale with ; |
Multiplies the quartic coefficient, ,
leaving every other operator untouched. |
Feeds directly into all
Appendix K conversion tables, so the CKM/PMNS fit and FRG flow must be
rerun, yet the algebraic form of the STM equation itself is
preserved. |
|
Add constant stiffness offset
|
Effective ; eq.
(1) reduces to bi-Laplacian-free form at lab scales |
No impact on
Planck-scale physics; quartic term returns at higher k. |
|
Dispersive viscous loss
|
Attenuates quartic phase without affecting quadratic baseline |
Adds
one parameter; Lindblad sector (Appendix P) must be re-checked for ghost
freedom. |
|
Sextic retune
|
Small
can partially cancel the quartic phase at this k
|
Alters high-k stability; soliton analysis (Appendix M) must be
updated. |
Any chosen fix must be propagated through the renormalisation tables of Appendix K and re-validated against flavour data and cosmology, but none threatens the fundamental STM structure.
Why this geometry is preferred
Off-the-shelf hardware (audio-rate shaker, kHz LDV); no RF drive or MS-s-1 data streaming.
Quartic residual large enough to detect yet small enough that sextic and nonlinear terms remain negligible.
Single-mode spectrum simplifies baseline fitting and error budgeting.
A high-frequency (25 kHz) variant could probe sextic and viscous terms once the quartic sector is confirmed.
I.3 Controlled Decoherence on Mechanical Membrane
Objective: Directly test STM prediction of decoherence transitioning from algebraic to exponential decay with introduced damping.
-
Implementation:
- –
Apply a 5 cm × 2 cm felt patch to raise the local nondimensional damping, .
-
Measurement:
- –
Intensity decay over time monitored at fixed membrane antinode, both with and without damping.
-
STM Signature:
- –
Without felt (undamped): algebraic decay pattern observed.
- –
With felt (damped): exponential decay pattern emerges clearly (time constant ~2–3 ms).
-
Falsification Criterion:
- –
Absence of clear algebraic-to-exponential decay distinction invalidates the STM prediction.
I.4 Twin-Membrane Bell-Type Experiment
Objective: Verify deterministic entanglement analogue predicted by STM via macroscopic CHSH inequality measurement.
-
Setup:
- –
Two identical membranes clamped back-to-back along one edge, opposite edges free.
- –
Paddle-shaped analysers near free edges set adjustable measurement angles ().
-
Measurement:
- –
Displacement at membrane endpoints measured as binary outcomes (±½ “spin” states).
-
STM Prediction:
- –
Correlations reproduce quantum-mechanical CHSH parameter, reaching the Tsirelson bound ().
-
Falsification Criterion:
- –
Repeatable shortfall of 1% or more below falsifies STM deterministic entanglement mechanism.
I.5 Slow-Light Optical Mach–Zehnder Test (Optional)
Objective: Provide optical verification of STM quartic dispersion via slow-light enhancement.
-
Method:
- –
Mach–Zehnder interferometer with a 10 cm silicon-nitride slow-light photonic-crystal segment.
-
STM Prediction:
- –
Tiny extra phase shift (~10-4 rad), at the limit of modern homodyne detection capabilities.
-
Feasibility:
- –
Only pursue if mechanical membrane tests (I.2–I.3) provide positive results. Marginal feasibility due to stringent sensitivity requirements.
I.6 Gravitational Wave Echoes from Black Hole Mergers
Objective: Detect STM-predicted gravitational wave echoes indicative of solitonic black-hole cores.
-
Facilities:
- –
Reanalysis of existing gravitational-wave events captured by LIGO and Virgo detectors (e.g., GW150914, GW190521).
-
Predicted Signature:
- –
Echoes post-ringdown at milliseconds intervals, frequency range approximately 100–1000 Hz.
-
Detection Approach:
- –
Matched filtering or Bayesian methods applied to existing strain data to extract subtle echo signals.
-
Falsification Criterion:
- –
Absence of predicted echo signals within detector sensitivity thresholds ( strain) challenges STM predictions.
-
Feasibility:
- –
Immediately feasible; data already collected, existing analysis pipelines available. Main challenge is distinguishing echoes clearly from instrumental or astrophysical noise.
I.7 High-Energy Collider Tests for STM-Induced Spacetime Ripples
Objective: Observe STM-predicted transient spacetime ripples produced in high-energy particle collisions.
-
Facilities:
- –
Large Hadron Collider (LHC) detectors (ATLAS/CMS, proton-proton collisions at 13 TeV)
- –
Pierre Auger Observatory (cosmic-ray events).
-
STM Prediction:
- –
Minute metric perturbations (), detectable via cumulative statistical anomalies over extensive datasets.
-
Measurement Method:
- –
High-statistics analysis to find subtle particle trajectory deviations, timing anomalies, or unexpected photon emissions correlated with specific STM-predicted frequency scales ( Hz).
-
Analysis Technique:
- –
Machine learning and statistical anomaly detection methods developed specifically for STM signature extraction.
-
Falsification Criterion:
- –
Non-detection after comprehensive analysis effectively rules out measurable STM-induced ripples at accessible energy scales.
-
Feasibility:
- –
Data sets and infrastructure already exist; principal challenge is the very small amplitude signals and substantial backgrounds.
I.8 Recommended Experimental Sequence and Feasibility Summary
High feasibility (immediate): Mechanical membrane interferometer and controlled decoherence tests (I.2–I.3); gravitational wave echo searches (I.6).
Moderate feasibility: Twin-membrane Bell-type test (I.4), collider anomaly search (I.7); feasible with careful setup or advanced statistical analysis.
Low feasibility (conditional): Optical slow-light interferometer (I.5); proceed only if strongly justified by positive mechanical test results.
This structured experimental programme provides a robust, multi-platform approach to empirically validating or falsifying distinctive STM predictions, leveraging both scalable laboratory analogues and state-of-the-art astrophysical/collider infrastructures available today.
Appendix J. Renormalisation Group Analysis and Scale-Dependent
Couplings
J.1 Overview
In the Space–Time Membrane (STM) model, spacetime is described as a four-dimensional elastic membrane whose dynamics incorporate scale-dependent elasticity and higher-order spatial derivatives—specifically the and operators—and in addition a second-order “tension” operator that softens the infrared behaviour. These features serve to control ultraviolet (UV) divergences and ensure a well-behaved theory at high momenta. In this appendix, we derive the renormalisation group (RG) equations for the elastic parameters by evaluating one-loop and two-loop corrections, and we outline the extension to three-loop order. We employ dimensional regularisation in dimensions together with the BPHZ subtraction scheme. The resulting beta functions reveal a fixed point structure that may explain the emergence of discrete mass scales—potentially corresponding to the three fermion generations—and indicate asymptotic freedom at high energies.
J.2 One-Loop Renormalisation
J.2.1 Setting Up the One-Loop Integral
Consider the cubic self-interaction term,
, in the Lagrangian. At one loop, the dominant correction to the propagator arises from the bubble diagram. In momentum space, the one-loop self-energy
is expressed as
where the propagator denominator is given by
At high momentum, the
term dominates, so the integral behaves roughly as
For the simplified case in which the term moderates the divergence, one typically encounters a pole in after dimensional regularisation. At intermediate momenta the added term softens infrared divergences, although it does not affect the ultraviolet scaling at large p.
J.2.2 Evaluating the Integral
Using standard results,
and substituting
, one finds
with
the Euler–Mascheroni constant. Hence, the one-loop self-energy contains a divergence of the form
J.2.3 Extracting the Beta Function
Defining the renormalised effective elastic parameter
through
and requiring that the bare parameter is independent of the renormalisation scale
(i.e.
), one differentiates to obtain the one-loop beta function for the effective coupling
(which parameterises
):
where
a is a constant proportional to
.
J.2.4 Tension-Coupling Beta Function:
Defining the renormalised tension via
, and enforcing
, one finds to one-loop order
This shows that the tension coupling runs multiplicatively with the elastic self-coupling.
J.3 Two-Loop Renormalisation
At two loops, more intricate diagrams contribute. We discuss two key contributions: the setting sun diagram and mixed fermion–scalar diagrams. Both diagrams generate divergences in the coefficient of
. In particular, the setting-sun topology yields a two-loop counterterm
, and the mixed fermion–scalar graphs contribute further
pieces to
. Consequently, the renormalisation constant for the tension becomes
feeding into a two-loop correction of the form
.
J.3.1 The Setting Sun Diagram
For a diagram with two cubic vertices, the setting sun contribution to the self-energy is given by:
with
as defined above. To combine the denominators, one introduces Feynman parameters:
After performing the momentum integrations, overlapping divergences manifest as double poles in and single poles in .
J.3.2 Mixed Fermion–Scalar Diagrams
If the Yukawa coupling y (coupling u to ) is included, diagrams involving fermion loops inserted in scalar bubbles contribute additional terms. Such diagrams yield divergences proportional to after performing the trace over gamma matrices and momentum integrations.
J.3.3 Two-Loop Beta Function
Collecting all two-loop contributions, the renormalisation constant
for the effective coupling is expanded as:
yielding the two-loop beta function:
with the coefficient
b incorporating both single and double pole contributions.
Note: Because the elastic origin of all gauge sectors is shared, the FRG flow admits but does not require a single crossing point. Choosing initial elastic ratios that miss that crossing leaves all low-energy observables unchanged and avoids introducing proton-decay channels.
J.4 Three-Loop Corrections and Fixed Points
At three loops, additional diagrams (such as the “Mercedes-Benz” topology) and further mixed fermion–scalar contributions introduce terms of order
. Schematically, the three-loop self-energy takes the form:
Analogously, three-loop diagrams (e.g.\ the Mercedes-Benz topology) induce further poles in the
channel, generating an
correction to
. Thus the full tension beta function reads
which may influence the position and stability of nontrivial fixed points.
Defining the bare coupling as
and enforcing
-independence leads to the full beta function:
The existence of nontrivial fixed points, where , depends on the interplay of these terms. If multiple real solutions exist, the model may naturally produce discrete mass scales, potentially corresponding to the three fermion generations. Moreover, a negative term could imply asymptotic freedom.
J.5 Illustrative One-Loop Example
As a concrete example, consider a bubble diagram in the scalar sector with a cubic self-interaction term (See Figure 9).
The one-loop self-energy is given by:
where
may arise from the second derivative of
.
Isolating the UV pole now proceeds identically, but the IR-regulated denominator improves convergence for small p.
In dimensional regularisation (with
), one isolates the divergence via
where
is the Euler–Mascheroni constant. This divergence determines the running of
and leads to a one-loop beta function of the form:
Higher-loop contributions then add corrections of order and beyond.
J.6 Summary and Implications
Tension Coupling: The new second-order operator generates a one-loop beta function , with higher-loop corrections analogous to those of the elastic self-coupling
One-Loop Corrections:
Yield a divergence , leading to .
Two-Loop Corrections:
The setting sun and mixed fermion–scalar diagrams contribute additional overlapping divergences, resulting in a beta function .
Three-Loop Corrections:
Further diagrams introduce terms , refining the beta function to .
Fixed Point Structure:
Nontrivial fixed points (satisfying ) can emerge, potentially corresponding to distinct vacuum states. These may naturally explain the discrete mass scales observed in the three fermion generations, while also suggesting asymptotic freedom at high energies.
Overall, the renormalisation group analysis demonstrates that the inclusion of higher-order derivatives in the STM model not only tames UV divergences but also induces a rich fixed point structure, with significant implications for particle phenomenology and the unification of gravity with quantum field theory.
Appendix K. Finite-Element Calibration of STM Coupling
Constants
This appendix details the finite-element methodology and physical anchoring used to determine the STM model’s dimensionless coupling constants.
K.1 Finite-Element Discretisation of the STM PDE
K.1.1 Spatial Mesh and Shape Functions
Domain: Choose a geometry (e.g.\ double-slit analogue, black-hole analogue) large enough to capture both local wave features and global displacement.
Mesh: Tetrahedral or hexahedral elements with adaptive refinement in regions of steep gradients (near slits, curvature peaks, soliton cores).
Shape functions: Require at least continuity to support and operators. Use high-order polynomial or spectral bases, or employ mixed formulations that introduce auxiliary fields to lower the derivative order.
K.1.2 Discrete Operator Assembly
Expand
apply the
and
terms element-by-element using high-order quadrature, and assemble the global mass, stiffness and higher-order matrices. Careful assembly preserves self-adjointness and sparsity for numerical stability.
K.2 Time Integration and Non-Linear Solvers
K.2.1 Implicit Time Stepping
K.2.2 Non-Linear and Damping Terms
Include residual contributions from:
Cubic self-interaction .
Yukawa coupling .
Scale-dependent stiffness .
Damping .
At each timestep, solve via Newton–Raphson:
where
R is the residual vector and
J its Jacobian. Very small or time-dependent
is treated as a weakly stiff term alongside dominant spatial stiffness.
K.3 Parameter Fitting via Cost-Function Minimisation
K.3.1 Simulation Outputs
Finite-element runs yield:
Interference patterns and decoherence times in analogue setups.
Ring-down frequencies and solitonic core shapes in gravitational analogues.
Coarse-grained vacuum offsets in persistent-wave experiments.
K.3.2 Cost Function and Optimisation
Define the cost
where
,
are simulated observables and
the corresponding data. Use:
Gradient-based methods (Levenberg–Marquardt, quasi-Newton) for smooth parameter spaces.
Evolutionary algorithms (genetic, particle-swarm) for high-dimensional or non-convex problems.
Multi-objective optimisation when fitting multiple datasets simultaneously.
K.4 Practical Considerations and Limitations
Computational cost: 3D problems require adaptive mesh refinement and parallel solvers.
Boundary conditions: Use absorbing or perfectly matched layers for wave analogues; radial or no-flux conditions for black-hole analogues.
Chaotic sub-Planck fluctuations: May necessitate ensemble averaging over varied initial conditions.
Scale-dependent: For cosmological tests, model globally; laboratory analogues may implement local instead.
K.5 Cosmological-Constant Fit via Persistent Waves
To match the observed dark-energy density:
Sign constraint: Ensure so persistent oscillations neither diverge nor decay too rapidly.
Minimal damping: Choose sufficiently small that oscillation amplitudes remain effectively constant over the age of the Universe.
After each simulation, compute
and iterate
until
.
K.6 Planck-Unit Non-Dimensionalisation
We adopt the conventional Planck units
Any coefficient that still carries dimensions in SI is rendered
dimension-less by
where the exponent pair
is chosen so that the remaining dimensions cancel. Coefficients that are already dimension-free (
and
g) have
.
| STM symbol |
PDE term |
units (SI) |
|
Planck-ND formula |
|
|
|
|
|
| T |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
| g |
|
— |
|
|
|
|
— |
|
|
† Because contains mass, the natural divisor is rather than .
Numerical note.
After the SI → non-dimensional step (Appendix K.6-1), all “Pa-class” coefficients acquire a tiny value . We adjust as follows;
We next rescale so that the term acquires the tidy coefficient .
Because the same rescaling hits every higher-derivative term with the appropriate power,
Only the damping coefficients receive this pragmatic trim; all elastic parameters are fixed purely by the uniform rescaling.
The resulting solver-friendly set is denoted by a superscript .
K.7 Physical Calibration of STM Elastic Parameters
Below each SI coefficient is matched to a familiar constant and then rendered dimensionless via K.6:
-
Mass density
- –
STM symbol: (coefficient of )
- –
Derivation: For plane waves , dispersion with gives
-
TensionT
- –
STM symbol:T (coefficient of )
- –
Derivation: Low-k dispersion fixes
-
Sixth-order stabiliser
- –
STM symbol: (coefficient of )
- –
Derivation: UV cutoff gives
-
U(1) gauge couplingg
- –
STM symbol:g (in minimal substitution )
- –
Derivation: Electromagnetism .
-
Cubic self-interaction
- –
STM symbol: (coefficient of )
- –
Derivation: Higgs quartic self coupling.
In the coarse-grained effective action we add a dissipative term
with
as argued in
Section 3.4.2; this damps residual
zitterbewegung above the flavour-mixing scale while preserving all conserved charges.
| STM symbol |
PDE location |
Calibrated value(SI) |
Planck-ND value |
Solver value |
Physical anchor |
|
|
|
|
1.00 |
plane-wave dispersion
|
|
|
|
|
0.10 (unit) |
|
|
|
|
|
1.00 |
|
|
vacuum offset |
|
|
|
observed dark-energy density |
|
|
|
|
0.02 |
UV cut-off
|
|
|
|
1.00 |
0.01 |
decoherence time
|
| g |
|
0.3028 |
0.3028 |
0.05 |
|
|
|
0.13 |
0.13 |
0.13 |
Higgs quartic self-coupling |
Notes:
Zero damping simulations For numerical cross-checks we also run a formal limit, but the physical model uses the value above.
Spinor dephasing rate inequationsinherits its value directly from and therefore no additional fit is required.
K.8 Usage Notes
Appendix L. Nonperturbative Analysis in the STM
Model
L.1 Overview
While perturbative approaches (such as loop expansions and renormalisation-group analysis in Appendix J) provide significant insights into the running of coupling constants and ultraviolet (UV) behaviour, many crucial phenomena in the Space–Time Membrane (STM) model arise from nonperturbative effects. These include:
Solitonic excitations: Stable, localised solutions arising from the nonlinearity of the full STM equations, now including a tension term .
Topological defects: Long-lived structures that may contribute to vacuum stability and the emergence of multiple fermion generations.
Nonperturbative vacuum structures: Potential mechanisms for dynamical symmetry breaking.
Gravitational-wave modifications: Additional contributions to black-hole quasi-normal modes (QNMs) due to solitonic excitations.
To study these effects, we employ a combination of Functional Renormalisation Group (FRG) techniques (now tracking the running of the tension coupling ), variational methods, and numerical soliton analysis.
L.2 Functional Renormalisation Group Approach
A powerful tool for analysing the nonperturbative dynamics of the STM model is the Functional Renormalisation Group (FRG). The FRG describes how the effective action
evolves as quantum fluctuations are integrated out down to a momentum scale
k. The evolution equation (the Wetterich equation) reads:
where
is an infrared regulator and
the second functional derivative.
L.2.1 Local Potential Approximation (LPA) with Tension
Under the Local Potential Approximation, we posit the ansatz
where
is the running tension coupling (the coefficient of
in momentum space). Its flow can be obtained by projecting the Wetterich equation onto the
operator, yielding
alongside the usual potential flow
Solving these coupled flows reveals how tension-driven stiffness and vacuum structure co-evolve, potentially yielding multiple nonperturbative minima.
L.3 Solitonic Solutions and Topological Defects
L.3.1 Kink Solutions with Tension
Consider the classical double-well potential
The static one-dimensional field equation, now including tension
T, is
This admits the kink solution
which interpolates between
at
and
at
.
L.3.2 Soliton Stability and Energy Calculation
The total energy of this kink is
which is finite and ensures classical stability.
L.3.3 Link to Fermion Generations
Fermions couple to the displacement field via a Yukawa interaction
If
acquires multiple stable vacuum expectation values (VEVs)
(e.g.\ from different soliton types), fermion masses arise as
providing a natural mechanism for a hierarchy of three generations.
L.4 Influence on Gravitational Wave Ringdown
Solitons near a black-hole horizon modify the quasi-normal mode equation:
where the effective potential
now depends on the soliton profile (which in turn depends on
T). The resulting frequency shift scales as
with
. Such shifts could be probed by LIGO/Virgo observations.
L.5 Illustrative toy model: multiple mass scales and deterministic flavour mixing
The numerical curve in Figure 10 is not meant as a precision fit; it is a proof-of-concept showing that the STM functional-renormalisation flow can generate several well-separated condensates even in the simplest truncation. Below we spell out the minimal calculation that produces the three minima quoted in the caption.
L.5.1 Ansatz and flow equation
At each RG scale
k we keep only the local potential
(Local-Potential Approximation, LPA) and postulate a quartic “Mexican-hat’ ’ form
with running couplings
and
.
Using the Litim regulator
and working in
the Wetterich equation reduces to
Matching coefficients of
and
gives two coupled
-functions
where
and
is the UV cutoff (taken as 1 in nondimensional units).
L.5.2 Toy UV data and infrared couplings
For the illustrative run we choose*
A fourth-order Runge–Kutta integration of (L.24) down to
yields
*These numbers are not thecalibrated STM parameters of Appendix K; they merely demonstrate the mechanism.
L.5.3 Effect of the cubic STM correction
The sextic elastic regulator induces a small cubic correction in the scalar envelope, giving
Solving
in (L.25) produces three inequivalent stationary points
The first two are local minima; the third is a shallow but distinct well created by the interplay of quartic and cubic terms. Figure 10 plots on a logarithmic vertical scale so that all three wells are simultaneously visible.
L.5.4 Interpretation as “generational’ ’ mass scales
If a Yukawa term
couples the scalar to a fermion, the dynamical masses are
. Choosing
for illustration gives the hierarchy
mirroring equation (4.11) in the main text. Although those ratios do
not match the measured quark or lepton spectrum, the exercise demonstrates—without fine-tuning—that
three discrete minima can emerge from a single elastic scalar once the STM flow and sextic regulator are engaged.
A full STM calculation would include:
Those ingredients shift both the depths and positions of the wells and are required for quantitative agreement with CKM/PMNS phenomenology. Nevertheless, the toy model already shows why discrete RG basins are natural in STM and how they can underpin a deterministic origin for the three fermion generations.
L.5.5 Mixing angles and deterministic CP phases
The three wells derived above supply only mass scales. Reproducing the observed CKM and PMNS mixing angles and the Jarlskog-type CP phase requires an additional ingredient: the deterministic interaction between each bimodal spinor on our membrane face and its mirror antispinor on the opposite face. These rapid cross-membrane exchanges—zitterbewegung at the Planck scale—imprint scale-averaged complex phases on the effective Yukawa couplings, exactly as worked out for the weak sector in Appendix C.3.1.
When that deterministic phase mechanism is combined with the
three discrete minima generated in (L.26), a complete flavour picture emerges:
Appendix R shows—via a Monte-Carlo scan constrained only by the non-dimensional STM ratios —that this deterministic recipe already fits all nine CKM magnitudes to sub-per-mille accuracy and the PMNS angles to within a few per cent. A full numerical match of the entire fermion spectrum, incorporating the renormalised potential obtained here plus spinor/gauge back-reaction, is deferred to future work; nonetheless, the mechanism laid out in Appendix C.3.1 and demonstrated heuristically in Appendix R remains a primary motivation for extending the STM phenomenology.
Appendix M. Covariant Generalisation and Derivation of Einstein
Field
Equations
M.1 Action and Lagrangian Decomposition
On a Lorentzian manifold
, the total action is
where the scalar ("membrane") sector is
the spinor sector is
and the interactions read
Throughout we work with .
M.2 Variation: Einstein Equations
Varying the total action
with respect to the metric
formally yields
where, for example,
and the
term in
contributes higher-derivative stresses.
Effective low-energy projection. Although the full stress–energy splits into scalar, spinor and interaction pieces, in the four-dimensional, low-energy description
only the spinor–mirror-spinor tensor
appears as a source of observable curvature (see Appendix M.6). Both the scalar-sector and interaction contributions remain internal to the membrane, and any constant offset generated by persistent oscillations is absorbed into the cosmological term
(Appendix M.7). Consequently, the operative Einstein equations become
M.4 Variation: Spinor Field Equation
To couple spinors to the curved STM membrane one introduces a tetrad
and spin connection
, so that the covariant derivative on spinors is
with curved gamma-matrices
satisfying
. Varying the action with respect to
then yields the Dirac equation in curved spacetime,
Under the usual inner product , and assuming vanishes (or obeys appropriate boundary conditions) at , the operator is formally self-adjoint, guaranteeing unitary evolution.
M.5 Flat-Space and WKB Limits
In the weak-field regime, set , replace covariant derivatives , and identify the scalar field with membrane displacement via . The d’Alembertian becomes , and all curvature-related terms vanish.
Under these substitutions, the covariant field equations reduce exactly to the flat-space STM dynamics:
the sixth-order scalar PDE governing membrane elasticity,
the nonlinear envelope equation describing modulated sub-Planck waves, and
the Yukawa-type spinor coupling from .
The parameters map directly to their dimensionless STM counterparts via the natural scaling described in Appendix H.
In the WKB limit, where u exhibits rapid oscillations modulated by a slowly varying envelope, the STM model further yields the emergent Schrödinger-like dynamics and decoherence behaviour, as discussed in Sections H.2–H.5.
Appendix M.6 Spinor–Mirror-Spinor Stress–Energy Tensor
We begin from the curved-spacetime action for the two-component spinor
and its mirror counterpart
, minimally coupled to both the emergent U(1) gauge field
and the background metric
:
where
and
denotes the Levi–Civita connection.
Varying the total action
with respect to
yields the spinor–mirror-spinor stress–energy tensor as the
sole matter source:
with
where
and
.
Physical interpretation.
Attraction between and produces positive curvature outside the membrane, drawing elastic energy out of the bulk into the surrounding spacetime.
Repulsion or cancellation from the interaction between spinors and mirror spinors, as would arise pre annihilation, relieves curvature, pushing energy back into the membrane—modelling annihilation as elastic-energy deposition.
Projected Einstein equations. All observable spacetime curvature in STM arises from fermionic excitations. The membrane’s intrinsic elastic energy remains invisible to the four-dimensional Einstein equations. Consequently, the effective low-energy field equations read
Clarifying remark. The intrinsic membrane stress–energy, though essential for the internal elastic dynamics, does not couple to our macroscopic Einstein equations—only the spinor–mirror-spinor tensor sources curvature. Any constant vacuum term arises solely from the persistent membrane oscillations analysed in Appendix M.7.
Appendix M.7 Vacuum-Energy Offset from Persistent Waves
STM predicts a tiny but non-zero cosmological constant arising solely from phase-locked membrane oscillations. Below we show how the carrier-wave energy gives rise to , with no contribution from the membrane’s static background.
M.7.1 Multi-Scale Locking of Carrier Modes
Consider a carrier oscillation
on the membrane with weak feedback:
where
. Introducing the slow time
and writing
the two-time expansion (cf. H.4–H.6) gives, to
,
so that
locks into a persistent, non-decaying oscillation of amplitude
.
M.7.2 Emergent Cosmological Constant
The
time-averaged energy density of the locked carrier is
Because this offset is
constant in space and time, in the low-energy 4D description it appears exactly as a cosmological term:
No other membrane energy contributes: the static elastic background remains non-gravitating and does not enter .
See Appendix H for the detailed averaging that yields a vacuum-offset
.
M.8 Extended Elastic Action and PDE
In the flat-space limit (M.5), the full membrane dynamics—including elastic stiffness, higher-order regularisation, damping, nonlinearity and matter coupling—is governed by
where the Yukawa interaction
descends from
, and emergent gauge fields
arise upon enforcing local U(1) phase invariance on the two-component spinor (M.6). Mirror-spinor attractions or cancellations shuttle elastic energy out of or into the membrane substrate—sourcing local curvature via
(M.6)—while only the time-averaged, phase-locked carrier-mode energy contributes a constant vacuum term
(M.7).
M.9 Linear Regime: Emergent Einstein-Like Equations
For small
u, drop
and
to get
Under
, this reproduces the linearised Einstein equations
A uniform stiffness shift appears as a cosmological-constant term (Appendix H.6).
M.10 Cosmological Constant and Vacuum Energy
Persistent feedback generates a
uniform extra stiffness
over and above the baseline modulus
. After coarse-graining over many carrier cycles the rapid oscillations average out, leaving the constant energy density
which enters the emergent Einstein equations precisely as a cosmological constant. No separate dark-energy field is required; the bare tension
T merely fixes the gravitational scale
(Appendix H).
M.11 Nonlinear and Damping Effects
Regulators avert singularities (Appendix F).
Non-Markovian damping or memory kernels model horizon-like dissipation, affecting information flow near compact objects.
Strong-field particle–mirror dynamics can repeatedly remove or deposit local stress–energy; fully quantifying such non-linear exchanges remains an open problem.
M.12 Modifications to Standard EFE & Testable Predictions
Extra Stiffness Terms: High-order derivatives and running add novel curvature corrections.
Scale-Dependent : , varying with local stiffness.
Time Dilation & Redshift: Strain–potential mapping is modified by elasticity, yielding small anomalies near compact or oscillating bodies.
Ringdown QNM Shifts: in black hole mergers—future detectors like the Einstein Telescope may observe these.
Laboratory Tests: Metamaterials with tunable T can probe short-range departures from GR in torsion-balance or atomic-clock experiments.
M.13 Progress on Open Challenges
Ghost-free quantisation: Ensuring no negative-norm modes with and .
Spinor/gauge self-adjointness: Constructing well-posed boundary conditions in the presence of T.
Planck-scale completion: Bridging continuum elasticity to a fundamental discrete structure remains to be developed.
M.14 Modifications to Traditional EFE, Time Dilation, and Testable Predictions
While the linearised STM membrane reproduces the familiar weak-field Einstein equations, the inclusion of a tension term and higher-order elasticity yields definite corrections:
Extra stiffness operators: The tension , together with the fourth- and sixth-order terms and , adds new curvature-dependent contributions to the emergent field equations, so that schematically
Potential observational tests
Black-hole ringdown shifts: Quasi-normal mode frequencies acquire corrections; next-generation detectors (Einstein Telescope, Cosmic Explorer) could detect or constrain these shifts.
Localised time-dilation anomalies: Precision atomic-clock comparisons at different altitudes or in strong laboratory-scale potentials might reveal small departures from the GR redshift prediction.
Vacuum-energy inhomogeneities: Spatial fluctuations in across cosmological scales could leave imprints on the CMB power spectrum or lensing maps, providing a handle on variability .
Spatial fluctuations in T across cosmological scales could leave imprints on the CMB power spectrum or lensing maps, providing a handle on variability.
Mirror-interaction signatures: Interferometric or cavity experiments performed in controlled mirror-antiparticle environments may uncover tiny deviations from standard QED if local stress-energy is periodically removed.
M.15 Conclusion
By identifying spacetime curvature with membrane strain and by showing how energy is exchanged between local dents (A + C) and the uniform reservoir (B), the STM model recasts Einstein’s equations within a single deterministic elasticity framework. Crucially, persistent sub-Planck oscillations build a uniform extra stiffness on top of the baseline tension. It is this offset – not T itself – that manifests as the cosmological constant. The higher-order operator then prevents singularities by stiffening the membrane in extreme-curvature regions. Although formal proofs of operator self-adjointness, full anomaly cancellation, and UV completion at the Planck scale remain outstanding, the model already yields concrete, testable deviations—from black-hole ringdown shifts and clock-rate anomalies to CMB inhomogeneity limits—offering a clear experimental roadmap for validating or refuting this higher-order elasticity approach to unifying gravity and quantum phenomena.
Appendix N. Emergent Scalar Degree of Freedom from
Spinor–Mirror Spinor
Interactions
This appendix provides a conceptual outline of how spinor–mirror spinor interplay in the STM framework can yield a single scalar excitation. Such a mode can couple to gauge bosons and fermions in a manner reminiscent of the Standard Model Higgs, potentially matching observed branching ratios and decay channels.
In the STM framework the membrane displacement
satisfies
All subsequent mode decompositions and coarse-graining (including the bimodal split into and ) implicitly inherit this extra term, which dominates the low-k (infrared) dispersion.
N.1 Spinor–Mirror Spinor Setup
Bimodal Spinor
As introduced in Appendix A, the STM model begins with a bimodal decomposition of the membrane displacement field
. This decomposition yields a two-component spinor
, often written:
On the opposite side (the “mirror” face of the membrane), one defines a mirror antispinor . Zitterbewegung exchanges between and create effective mass terms and CP phases.
Effective Yukawa-like Couplings
The total Lagrangian typically contains terms coupling
to the membrane field. Symbolically:
Coarse-graining these rapid cross-membrane interactions can spontaneously break symmetry and leave behind a massive scalar.
N.2 Radial Fluctuations and the Emergent Scalar
Spinor–Mirror Condensate
Once one includes zitterbewegung loops and possible non-Markovian damping, the low-energy effective theory may exhibit a condensate . This is akin to spontaneous electroweak symmetry breaking in standard field theory, except it arises from deterministic elasticity plus spinor–mirror spinor pairing.
Effective kinetic operator for h.
Writing
in a radial decomposition, the quadratic spatial part of the emergent scalar’s effective action reads
so that its low-momentum (“mass”) term receives a direct contribution from
T.
Polar (Amplitude–Phase) Decomposition
Fluctuations around the condensate can be expressed in polar or radial form:
Phase : Would-be Goldstone modes that can be “absorbed” by gauge bosons, giving them mass.
Amplitude : A real scalar field representing the radial component of the condensate. One may write , with a vacuum expectation value and the physical scalar mode.
Couplings to Gauge Bosons and Fermions
If the gauge fields in the STM become massive via this symmetry breaking, the surviving radial fluctuation couples to them proportionally to . Similarly, fermion masses induced by – interactions imply Yukawa-type couplings of h to fermion bilinears. Hence, can play the role of an effective Higgs-like scalar.
N.3 Potential Matching to Higgs Phenomenology
Branching Ratios
In standard electroweak theory, the Higgs boson’s partial widths are tied to its gauge and Yukawa couplings. In STM:
Gauge couplings arise from the local spinor-phase invariance (Appendix C).
Yukawa couplings come from cross-membrane spinor–mirror spinor pairing.
Matching the observed 125 GeV resonance would require calibrating these couplings so that partial widths fit LHC measurements.
Unitarity and Vacuum Stability
The radial mode must also preserve unitarity in high-energy processes (e.g. scattering of ) and ensure vacuum stability. STM’s elasticity-based PDE constraints could supplement or replace the usual “Higgs potential” arguments, but verifying this in detail remains an open theoretical challenge.
Numerical Implementation
A full PDE-based simulation (cf. Appendices K, J) could in principle track how , -regularisation, and spinor–mirror spinor couplings produce a scalar mass near 125 GeV. Fine-tuning or discrete RG fixed points might be involved in setting this scale. Reproducing branching fractions, cross sections, and loop corrections from the STM perspective would then confirm or falsify this emergent scalar scenario.
N.4 Conclusions and Outlook
The emergent scalar arises as a collective radial excitation in spinor–mirror spinor space once the membrane’s background is considered. While the conceptual mechanism is clear—no fundamental Higgs field is required—realistic numerical fits to collider data remain pending. Nonetheless, this approach demonstrates how the deterministic elasticity framework can replicate a Higgs-like sector, further unifying typical quantum field concepts under the umbrella of classical membrane dynamics.
Appendix O. Rigorous Operator Quantisation and
Spin-Statistics
O.1 Introduction and Motivation
A central goal of the Space–Time Membrane (STM) model is to unify gravitational-scale curvature with quantum-like field phenomena, all within a single deterministic elasticity partial differential equation (PDE). However, ensuring that this PDE admits a fully rigorous operator quantisation—particularly once higher-order derivatives (such as ), emergent spinor fields, mirror spinors, and non-Abelian gauge interactions are included—remains a major open task. In conventional quantum field theory (QFT), one enforces:
Self-adjointness (Hermiticity) of the Hamiltonian, ensuring real energy eigenvalues and unitarity.
Spin–statistics correlation so that half-integer spin fields obey Fermi–Dirac statistics while integer spin fields obey Bose–Einstein statistics.
Gauge invariance (for groups such as SU(3) × SU(2) × U(1)), typically handled via BRST quantisation or Faddeev–Popov ghost fields.
Absence of ghost modes or negative-norm states, especially when higher-order derivative operators are present.
Below, we outline how the STM model might satisfy these requirements by focusing on (a) the use of appropriate boundary conditions and function spaces for high-order operators, (b) an effective field theory (EFT) perspective for the term, (c) the implementation of anticommutation rules for spinor fields (including mirror spinors), and (d) the preservation of gauge invariance and anomaly cancellation.
O.2 The STM PDE and Its Higher-Order Operator
The STM model is described by the PDE
where, in addition, the full theory includes non-Abelian gauge fields for SU(3) × SU(2) × U(1) and mirror spinors that couple across the membrane.
In this PDE:
: effective mass density describing inertial response
T: membrane tension, stiffening long-wavelength modes
: baseline elastic modulus at renormalisation scale
: local stiffness variations; its uniform part acts like vacuum energy once fast oscillations are averaged out
: sixth-order regularisation damping ultraviolet modes
: viscous damping, extensible to non-Markovian kernels
: non-linear self-interaction
: Yukawa-like coupling to an emergent spinor field
: external forcing or boundary effects.
Note: the low-k dispersion reads so the tension T governs the infrared behaviour.
O.3 Function Spaces and Boundary Conditions
O.3.1 Higher-Order Sobolev Spaces
Because the PDE includes derivatives up to
, a natural choice is to consider solutions in a Sobolev space of order three. Specifically, we assume
which ensures that all derivatives of
u up to third order are square-integrable. This means
On an infinite domain, we impose that
For a finite domain , we adopt Dirichlet or Neumann boundary conditions on so that integration by parts eliminates boundary terms. This guarantees that the differential operators and are symmetric and well-defined, enabling the construction of a self-adjoint Hamiltonian in the conservative limit.
O.3.2 Elimination of Spurious Modes
With the chosen boundary conditions, partial integrations bringing out or are symmetric. Thus, even if the PDE includes strong damping or additional scale-dependent terms, the field remains within a function space where the operators are well-behaved, crucial for constructing a self-adjoint Hamiltonian.
O.4 Spin–Statistics Theorem in a Deterministic PDE
O.4.1 Anticommutation Relations
In standard QFT, spin–statistics is ensured by imposing the anticommutation relations
For the classically deterministic STM PDE, we require that upon quantisation, the emergent spinor fields obey these same relations. This is enforced by appropriate boundary conditions (such as antiperiodic conditions in finite domains) and projection onto a subspace where these antisymmetric properties hold.
O.4.2 Mirror Spinors and CP Phases
The STM model includes mirror spinors,
, on the opposite face of the membrane. Their interactions, often captured by terms like
must also respect the same anticommutation rules to avoid doubling the physical degrees of freedom. Imposing identical anticommutation structures on both
and
, with additional boundary condition constraints linking them, ensures that the full system upholds the spin–statistics theorem.
O.5 Ghost Freedom and theTerm
O.5.1 Ostrogradsky’s Theorem and EFT Perspective
Higher-order time or spatial derivatives can, in principle, lead to Ostrogradsky instabilities and the appearance of ghost modes (negative-norm states). In the STM model, the term is treated as an effective operator, valid up to a cutoff scale . Provided that and the field u is restricted to a Sobolev space such as , the spurious high-momentum modes that might otherwise cause negative-energy contributions are excluded. Additionally, the damping term further suppresses these modes, preserving unitarity below the cutoff.
O.5.2 Constructing a Hamiltonian
A convenient starting point is the elastic–spinor Lagrangian density, now including the tension term:
The conjugate momentum is
Performing the Legendre transform under our Dirichlet/Neumann boundary conditions (so that all total derivatives vanish) gives the Hamiltonian density
Remark. The above energy functional is manifestly bounded below provided , , and . In particular the new tension term gives an extra infrared-positive contribution, while the sextic term suppresses any high-momentum ghosts. Hence—within the low-energy, effective-field-theory regime set by our cutoff—no negative-norm (Ostrogradsky) states arise.
O.6 Gauge Fields and BRST Quantisation
O.6.1 Non-Abelian Gauge Couplings
The STM model also incorporates non-Abelian gauge fields corresponding to groups such as SU(3) × SU(2) × U(1). Their contribution to the Lagrangian is typically given by
where
is the field strength tensor. To maintain gauge invariance, standard gauge-fixing procedures (e.g. the Lorentz gauge) are applied. Faddeev–Popov ghost fields are then introduced as necessary.
O.6.2 BRST Invariance
By adopting BRST quantisation, the physical states of the theory are defined to lie in the kernel of the BRST charge . This process ensures that gauge anomalies are cancelled and that the resulting physical Hilbert space contains only positive-norm states, preserving the integrity of the spin–statistics for fermions and the consistency of gauge interactions.
O.7 Summary and Outlook
We have proposed a scheme for rigorous operator quantisation of the STM model that addresses the challenges posed by higher-order derivatives, damping, and the incorporation of spinor and gauge fields. In summary:
We restrict the field to suitable Sobolev spaces (e.g. ) and impose boundary conditions to ensure that operators like and are well-defined and symmetric.
We treat the term within an effective field theory framework, valid below a cutoff scale , thereby avoiding ghost modes.
We enforce the proper anticommutation relations for emergent spinor fields (and mirror spinors) to ensure Fermi–Dirac statistics, with additional boundary conditions that maintain the necessary antisymmetry.
For the gauge sector, BRST quantisation guarantees that the inclusion of non-Abelian interactions does not introduce negative-norm states.
While these measures establish a promising framework for a self-adjoint Hamiltonian and unitarity at low energies, further work is required—especially in multi-loop analyses and numerical validations—to conclusively demonstrate full consistency across all energy scales.
This strategy lays a conceptual foundation for combining classical elasticity with quantum field theoretic requirements in the STM model, and it offers a roadmap for future research into a fully unified and rigorously quantised theory.
Appendix P. Reconciling Damping, Environmental Couplings, and
Quantum Consistency in the STM
Framework
In this appendix, we address in detail the challenge of integrating the STM model’s intrinsic damping and environment interactions into a consistent quantum-theoretical framework. Specifically, the STM model is governed by the deterministic elasticity PDE for the displacement field
:
supplemented by interactions with spinor and gauge fields. A significant difficulty arises from the damping term
, representing energy dissipation into a presumed high-frequency environment, and its implications for quantum self-adjointness, positivity, and ghost freedom.
P.1 Quantum-Theoretical Implications of Damping
Classically the Rayleigh term
breaks time-reversal symmetry, so the full membrane equation is not generated by a self-adjoint Hamiltonian. To keep the quantum theory consistent we adopt the standard
open-system split:
where
contains only
conservative terms. The Yukawa coefficient in Appendix S satisfies
.
A single-cell, single-Planck-time coarse-graining of the sub-Planck bath yields
and, after dividing by
and multiplying by
,
With these constants fixed once and for all, the remaining task is to write the dissipator so that it preserves trace, complete positivity, gauge constraints and spin statistics.
P.2 Lindblad operators and environmental couplings
The dissipator is expressed as a sum over local Lindblad operators:
Gauge compatibility These jump operators commute with the Gauss-law constraints, so BRST symmetry and ghost freedom proven in P.6 remain untouched. When gauge-field damping is needed one may add , but in all baseline runs.
Because is quadratic in the L’s it removes exactly the energy that flows into the coarse-grained bath, while the Hamiltonian part stays Hermitian. Setting = = 0 removes the dissipator ; the master equation then reduces to unitary evolution under , which matches the conservative limit of the STM wave equation.
P.3 Time-Reversal Symmetry Breaking and the Thermodynamic Arrow of Time
Although the conservative STM wave equation is time-symmetric in the limit , once one includes realistic damping and environmental couplings the dynamics acquire a built-in irreversibility:
-
Causal, non-Markovian memory kernel
As derived in Appendix G, integrating out the fast “environment” modes produces a master equation for the reduced density matrix
where the memory kernel has support only for . By construction it depends only on past history, not on future states, and so enforces a causal, forward-pointing flow of information and coherence.
-
Reversible limit
Only in the formal limit and does the STM equation recover full time-symmetry. In any realistic setting, however, the combined effect of damping and causal decoherence defines a clear thermodynamic arrow of time.
Together, these two ingredients show that STM dynamics “travel” strictly forward in time: elastic waves dissipate, coherence decays, and entropy increases in a deterministic yet irreversible manner.
P.4 Avoiding Ghost Modes and Ensuring Positivity
The introduction of a higher-order spatial derivative term,
, must not introduce negative-norm ghost states. To ensure ghost freedom, we impose that
, and define the field
u rigorously within Sobolev spaces
. This ensures all energy contributions remain positive and finite:
Because the highest spatial derivative is even (sixth order) and its coefficient is positive, the principal symbol of the linearised operator remains elliptic; all mode energies are therefore bounded below, eliminating Ostrogradsky ghosts.
Thus, we rigorously ensure the model is devoid of Ostrogradsky instabilities. These results hold for a flat background; curved manifolds with non-trivial spinor/gauge structure will be treated in forthcoming work.
Note: The tension operator is only second order in space and introduces no new high-momentum instabilities, so it respects the same Sobolev-space positivity arguments as the lower-order terms
P.5 Non-Markovian Extensions and Memory Effects
Realistic environments might induce non-Markovian effects. To accommodate this, we generalise the Lindblad formalism via time-convolutionless (TCL) approaches, employing time-dependent memory kernels
:
ensuring these kernels remain positive-definite and decay suitably, maintaining quantum positivity and well-posedness of the master equation.
The TCL kernel is constructed so that and ; in the short-memory limit it reduces smoothly to the Markovian Lindblad dissipator introduced in P.2.
P.6 Gauge Symmetry and BRST Quantisation
Gauge invariance remains critical. Damping of gauge fields is treated carefully to maintain gauge symmetry through BRST quantisation, introducing Faddeev-Popov ghost fields to ensure unitarity and positivity within the gauge sector. Gauge-invariant Lindblad operators, e.g.:
ensure damping respects gauge symmetry explicitly.
Because the jump operator commutes with the BRST charge, the dissipator obeys , so nilpotency and gauge invariance are preserved even in the open-system evolution.
P.7 Summary of Quantum-Consistent STM Formulation
Through this carefully constructed open quantum-system approach, the STM model maintains:
Self-adjoint Hamiltonian (excluding dissipative terms explicitly).
Quantum positivity and ghost freedom via rigorously chosen Sobolev spaces and positive Lindblad (presently proven for flat ; curved-space extension remains outstanding).
Spin-statistics compliance and gauge invariance, via fermionic and gauge-compatible Lindblad operators.
Compatibility with realistic non-Markovian environments, ensuring a physically meaningful evolution of quantum states.
STM dynamics “travel” strictly forward in time: elastic waves dissipate, coherence decays, and entropy increases in a deterministic yet irreversible manner.
The conservative Hamiltonian includes both and the higher-order , contributions, all of which remain self-adjoint and positive-definite on .
This resolves a critical ongoing challenge, integrating classical damping terms and environmental interactions into a quantum-consistent framework, significantly strengthening the theoretical foundation and predictive capability of the STM model.
Appendix Q. Toy model PDE
simulations
Q.1 STM dimensionless couplings (See Appendix K.7)
| Symbol |
Physical definition / PDE term |
Dimension-less value used in demos |
|
|
1 |
|
tension coefficient in
|
0.10 |
|
|
1 |
|
sixth-order stabiliser |
0.02 |
|
scalar damping |
0.01 (physics) / 0
(diagnostics) |
|
spinor dephasing rate |
0.005 |
|
gauge (Yukawa) coupling |
0.05 |
|
cubic self-interaction |
0.13 |
|
external forcing amplitude |
|
Characteristic solver scales: (choice fixes the units).
Q.1.3 Conserved quantities (undamped benchmarks)
With
the Hamiltonian
and the skew-adjoint invariants
should stay within ±
.
Q.2 Common numerical pitfalls & remedies
| Pitfall |
Remedy |
|
Tension-mode drift – lets long-waves
blow up. |
Enforce and treat the
term semi-implicitly (Crank–Nicolson). |
|
Stiffblow-up –
excites Nyquist modes. |
Rule
of thumb (§ 3.5.4). Either reduce or apply a smooth
high-k taper to every /
operator. A Butterworth filter |
|
with on a 128² grid keeps
. |
|
|
Gauge-coupling runaway – large injected
instantaneously. |
Ramp
with and cap . |
|
Undamped benchmark crash () |
Use a
fully implicit BDF(3–5) or CN-leap-frog with the corrected CN
half-step stored before advancing. Track . |
Physical reminder All predictive STM runs require , the undamped mode is diagnostic only.
Q.3 Simulation recipes
Q.3.1 2-D spinor-membrane (leap-frog + CN)
Filter and with the Butterworth mask described above.
First CN half-step on .
Leap-frog RHS including damping .
Second CN half-step, store the corrected field, then update .
Switching between damped and undamped simply toggles ; if you may need a smaller or a fully implicit solver.
Q.3.2 1-D STM far-field diffraction
with as in the main text.
Q.4 Damped vs undamped runs
| Simulation |
|
Ramping g |
Observation |
| 2-D spinor |
|
linear
|
Smooth, slightly
dissipative dynamics. |
| 2-D spinor |
0 |
Linear |
Conservative; implicit solver
essential. |
| 1-D slit |
|
– |
Fringe decay plus phase shift. |
| 1-D slit |
0 |
– |
Pure phase correction, no decay. |
Q.5 Implementation guidelines
Spectral taper – if exceeds at any grid point, apply a Butterworth mask (see Q.2).
Crank–Nicolson hand-off – always copy the second CN half-step field into both u and .
Sampling & padding – and ×4 zero-padding suppress Gibbs artefacts in 1-D diffraction.
Windowing – use a Hanning taper on each slit edge.
Resolution rule – To capture tension-dominated modes, choose step size
Q.6 Code (supplied in the supplementary archive)
STM_spinor_damped.py – 2-D spinor membrane, .
STM_spinor_undamped.py – diagnostic conservative run ().
STM_schrodinger_damped.py – 1-D far-field with damping.
STM_schrodinger_undamped.py – 1-D far-field, phase-only variant.
These scripts already implement the Butterworth filter and CN hand-off rules discussed above.
Appendix R. First principles derivations of CKM and PMNS
matrices
R.1 Basis of the three-mode envelope equation
A single trivalent STM wave-packet sits in a region where the stiffness field
possesses three shallow wells at
. Because the well spacing
greatly exceeds the carrier wavelength
, each well supports a localised normal mode
,
, that is exponentially small outside its own well. Expanding the displacement field in this tight-binding basis,
and projecting the linearised STM PDE onto
yields, after time-averaging over the carrier oscillation, the slow-amplitude equation
Diagonal detunings
arise from on-site depth differences of the three wells. Nearest-neighbour mixings
stem from overlap of the sextic operator between adjacent wells; the phases
encode short-range interference. Uniform spinor damping later adds
when the envelopes are promoted to flavour states.
R.2 Elastic-mode couplings and the CKM matrix
With
the flavour-sector effective Hamiltonian is
Let
V contain the right eigenvectors of
; then the polar decomposition
is exactly unitary. A flat-prior Monte Carlo scan over the five elastic parameters with 50 000 draws gives
comfortably inside the PDG 1o bands. Acceptance for a squared error
is
. A secondary phase scan fixes the Jarlskog-invariant deviation
while maintaining unitarity to
. Turning
off shifts any modulus by less than
.
R.3 Seesaw implementation and the PMNS matrix
Choose a Dirac block
and a heavy Majorana mass
The light mass matrix is
and the damped effective operator
Diagonalising and polar-projecting yields
matching global oscillation fits within a few per cent; acceptance for
is
.
R.4 Algorithmic outline
The script STM_flavour_mixing.py supplied with the paper:
Initialises and .
Performs CKM scan over elastic parameters; logs best fit and acceptance.
Carries out phase refinement to pin down J.
Performs PMNS scan over ; logs best fit and acceptance.
Visualises bar charts and residual heat-maps for both matrices.
Runs reproduce the numbers above and confirm that dropping changes mixing magnitudes only at the level.
R.5 Summary
With the calibrated ratios and the fixed damping hierarchy , the STM model: – matches every CKM modulus to sub-per-mille precision, – reproduces the Jarlskog invariant to , – fits the PMNS matrix to a few per cent, all while preserving exact unitarity. Acceptance fractions of order show these fits are highly non-generic, providing quantitative support for STM’s deterministic flavour mechanism without introducing any extra free parameter.
R.6 Code (supplied in the supplementary archive)
Appendix S. Glossary of
Symbols
Fundamental constants
— Speed of light in vacuum.
ℏ — Reduced Planck constant, .
G — Newtonian gravitational constant.
— Boltzmann constant.
— Planck length, .
— Cosmological constant, linked to vacuum-energy density.
— Geometry-dependent coarse-graining factor () that sets the fraction of Planck-frequency jitter surviving a single-cell average and therefore fixes the macroscopic damping .
{Elastic membrane and field variables
— Mass density of the STM membrane.
— Classical displacement field of the four-dimensional elastic membrane.
— Operator form of the displacement field (canonical quantisation).
— Conjugate momentum.
— Scale-dependent baseline elastic modulus; inverse gravitational coupling.
— Local stiffness fluctuation, time- and space-dependent.
— Fourth-order spatial (bending) operator.
— Coefficient of the term; provides ultraviolet regularisation.
T — Baseline membrane tension (energy / length³); governs long-wavelength wave speed when , equal to the coefficient of the term in the STM action.
— Dimensionless shear-to-bulk stiffness ratio appearing in the covariant elastic moduli.
— Small but strictly positive damping coefficient; non-Markovian memory enters via .
— Potential energy density for the displacement field.
— Self-interaction coupling (e.g.\ ); one of the eight calibrated elastic parameters.
— External force density acting on the membrane.
Gauge fields and internal symmetries
— U(1) gauge field (photon-like).
— SU(2) gauge fields, .
— SU(3) gauge fields (gluons), .
— Gauge-group generators, e.g.\ for SU(2).
— Pauli matrices (); satisfy .
— Coupling constants for U(1), SU(2) and SU(3).
— U(1) field-strength tensor, .
— SU(2) field-strength tensor.
— SU(3) field-strength tensor.
— Structure constants of non-Abelian groups ( for SU(2)).
— Levi-Civita symbol (totally antisymmetric).
Fermion fields and deterministic CP violation
— Two-component spinor from bimodal decomposition of u.
— Mirror antispinor on the opposite membrane face.
— Fermion bilinear (Yukawa-like).
v — Vacuum expectation value of u.
— Yukawa coupling between spinors and u.
— Deterministic CP phase between spinor and mirror fields.
— Fermion mass matrix (complex, CP-violating).
Renormalisation group and couplings
— Renormalisation scale.
k — Functional-RG running scale (infrared cut-off).
— Effective coupling at scale .
— Beta function for RG flow.
— Strong coupling constant in the SU(3) sector.
— QCD-like confinement scale in STM.
— Scale-dependent wavefunction renormalisation (FRG).
Path-integral and operator formalism
— Functional integration measures.
Z — Path integral (partition function).
— Gauge-fixing parameter.
— Faddeev–Popov ghost and antighost fields.
Non-perturbative effects and solitonic structures
— Scale-dependent effective action (FRG).
— Infrared regulator suppressing modes with .
— Second functional derivative (inverse propagator).
— Scale-dependent effective potential.
— Scalar field variable in FRG analyses.
— Quasinormal-mode wavefunction near the solitonic core.
— Soliton energy.
— Solitonic mass scale.
— QNM frequency shift due to the soliton core.
Lindblad and open-quantum-system parameters
— Lindbladian acting on density matrix .
— Lindblad jump operators (dissipators).
— Density matrix of the system.
— Memory kernel for non-Markovian damping.
— Fermionic damping rate.
BRST and ghost-free gauge formalism
— BRST charge operator defining physical states.
— Physical Hilbert space satisfying .
F — Ghost-number operator.
s — Nilpotent BRST differential.
Double-slit and interference interpretations
— Off-diagonal coherence elements of an effective density matrix.
— Phase difference between elastic wavefronts at detectors.
— Observed interference intensity at position x.
Black-hole thermodynamics and solitonic horizon
— Bekenstein–Hawking entropy, .
— Effective horizon area in the STM solitonic geometry.
— Hawking-like temperature.
— Surface gravity at the effective horizon.
— Effective horizon radius.
Multi-scale expansion and vacuum-energy terms
— Slow coordinates: .
— Small multi-scale parameter.
— nth-order term in the displacement expansion.
— Slowly varying envelope amplitude.
— Oscillatory part of the stiffness field.
— Residual (vacuum) stiffness offset.
— Scaled damping coefficient, .
— Scaled nonlinear coupling.
— Feedback coefficient linking envelope amplitude to local stiffness perturbation.
— Group velocity of the slow envelope mode.