Preprint
Article

This version is not peer-reviewed.

The Refined Space–Time Membrane Model: Deterministic Emergence of Quantum Fields and Gravity from Classical Elasticity

Submitted:

09 March 2025

Posted:

11 March 2025

Read the latest preprint version here

Abstract

We present a refined Space–Time Membrane (STM) model, in which quantum and gravitational phenomena emerge deterministically from the elastic dynamics of a four-dimensional classical membrane. Extending previous work, we introduce scale-dependent elastic parameters, higher-order (∇^6) terms for ultraviolet regularisation, and non-Markovian decoherence. Coarse-graining over sub-Planck-scale fluctuations yields an effective Schrödinger-like equation and Born rule, all arising from deterministic chaos rather than intrinsic randomness. Wavefunction collapse is interpreted as deterministic decoherence. A bimodal decomposition of the membrane’s displacement field naturally yields a two-component spinor, giving rise to U(1), SU(2), and SU(3) gauge fields. Photons, W^± bosons, and gluon-like excitations emerge as wave–plus–anti-wave modes, while virtual bosons are reinterpreted as oscillatory modes with zero net energy exchange. Unlike conventional quantum field theory, these excitations are seen as purely classical membrane deformations at the sub-Planck scale, though further work is required to ensure the full system, including spinor–antispinor couplings, is free of negative-norm states. Electroweak symmetry breaking, including the emergence of the Z boson and CP violation, arises from deterministic interactions between spinor fields and their mirror antispinor counterparts via zitterbewegung-induced complex phases in effective Yukawa couplings. We highlight that a rigorous, multi-field operator formalism—encompassing all higher-order terms and non-Abelian gauge structures—remains an ongoing challenge for fully establishing self-adjointness and unitarity. Using Functional Renormalisation Group (FRG) analysis, we uncover solitonic solutions that stabilise black hole interiors and prevent singularities. The model predicts three fermion generations via discrete RG fixed points and suggests gravitational wave ringdown modifications as a testable signature. Although these nonperturbative features are conceptually appealing, explicit links to black hole thermodynamics and Hawking-like emission require further elaboration. This deterministic framework unifies quantum mechanics and gravity at a conceptual level without extra dimensions or fundamental randomness. Experimental tests via metamaterial analogues, finite element simulations, and astrophysical observations are proposed to validate STM predictions. We do not claim to have resolved every detail of quantum gravity, yet our refined STM approach offers a distinctive route toward reconciling quantum phenomena with gravitational curvature in a single continuum elasticity theory.

Keywords: 
;  ;  ;  ;  ;  ;  ;  ;  ;  

1. Introduction

Modern physics is built upon two seemingly incompatible foundations: General Relativity (GR) [2–4], which describes gravity through the curvature of spacetime, and Quantum Mechanics (QM) [5–8], whose probabilistic formalism governs microscopic phenomena. Despite remarkable successes within their respective domains, integrating these theories into a coherent framework remains one of contemporary physics’ most pressing challenges. Existing approaches—such as String Theory and Loop Quantum Gravity—provide valuable insights but have yet to deliver a definitive resolution of quantum gravity [6, 8]. Additionally, puzzles such as the black hole information paradox and cosmological constant problem underline fundamental tensions between GR’s smooth geometry and QM’s intrinsic probabilism [9–11].
The Space–Time Membrane (STM) model, initially presented in [1], offers an innovative route towards addressing this fundamental tension. The original STM model proposed that our observed four-dimensional spacetime forms one face of a classical elastic membrane, with a mirror universe on its opposite face. In this picture, gravitational curvature arises explicitly from large-scale membrane deformations caused by external energy distributions, while quantum-like phenomena—such as interference patterns, probability distributions governed by the Born rule, and apparent wavefunction collapse—result deterministically from chaotic elastic oscillations at sub-Planckian scales. Although successful in explaining interference and gravitational phenomena within a single framework, the original STM model was incomplete, lacking explicit mechanisms for gauge fields, spinor structures, electroweak symmetry breaking, fermion generations, CP violation, and a fully self-consistent derivation of Einstein-like field equations extending beyond linear approximations.
In this regard, other leading quantum gravity programmes—such as String Theory’s vibrating extended objects in higher dimensions and Loop Quantum Gravity’s quantised geometry—have similarly not achieved universal acceptance nor direct experimental confirmation, though each provides pivotal insights [6, 8]. The STM approach differs by positing a purely classical 4D elastic membrane that yields emergent quantum behaviour upon coarse-graining, offering a distinctive perspective on quantum–gravitational unification.
In this refined STM model, we address these shortcomings explicitly by incorporating:
  • Scale-dependent elastic parameters and higher-order spatial derivatives (notably the 6 operator) to regulate ultraviolet divergences.
  • Non-Markovian decoherence to explain deterministic wavefunction collapse.
  • A bimodal decomposition of the membrane’s displacement field into a two-component spinor Ψ ( x , t ) , naturally giving rise to emergent U(1), SU(2), and SU(3) gauge fields and corresponding gauge bosons.
  • A deterministic mechanism for electroweak symmetry breaking, where interactions between spinors on our membrane face and mirror antispinors on the opposite face—mediated by rapid oscillatory exchanges known as zitterbewegung—produce the mass terms for W ± and Z 0 bosons, and explicitly yield CP-violating phases without invoking intrinsic randomness or additional scalar fields.
  • A multi-loop renormalisation group (RG) analysis, supplemented by a Functional Renormalisation Group (FRG) nonperturbative approach, identifying discrete fixed points and vacuum structures that potentially explain the three observed fermion generations deterministically.
In addition, and of particular relevance for the gravitational sector, we extend the original approach that linked linearised strain fields u μ with metric perturbations h μ ν . We now show how the Einstein-like field equations can be derived in a more complete manner from the refined STM action—even when including higher-order elasticity terms, damping, and scale-dependent couplings. The key new material is presented in Appendix M, where we give a revised derivation clarifying how the membrane’s stress–energy tensor enters into modified Einstein equations and how such equations reduce to standard GR at large scales. Nonetheless, a fully rigorous operator-based verification—encompassing spinor couplings, non-Abelian gauge fields, and potential non-Hermitian damping effects—remains an ongoing challenge for ensuring strict self-adjointness across all interaction sectors.
Organisation of the Paper
The remainder of this paper is structured explicitly as follows:
  • Section 2 (Methods) provides a detailed overview of the refined STM wave equation, including explicit derivations of the novel higher-order elasticity terms, spinor construction, scale-dependent parameters, and the deterministic interpretation of decoherence.
  • Section 3 (Results) demonstrates how quantum-like dynamics, the Born rule, entanglement analogues, emergent gauge fields (U(1), SU(2), SU(3)), deterministic decoherence, fermion generations, and CP violation naturally emerge from the deterministic membrane equations.
  • Section 4 (Discussion) explores the broader implications of our findings. We also outline possible experimental tests and numerical simulations to verify predictions arising from this framework.
  • Section 5 (Conclusion) summarises the key theoretical advances, current limitations, and future research directions, including numerical and experimental proposals aimed at verifying the predictions of the refined STM model.
Appendices A–L comprehensively present supporting details, derivations, and numerical methods. In particular, they address:
  • Spinor operator formulations (Appendix A)
  • Force functions and interactions (Appendix B)
  • Gauge symmetry emergence and CP violation (Appendix C)
  • Coarse-grained Schrödinger-like dynamics (Appendix D)
  • Deterministic entanglement (Appendix E)
  • Singularity avoidance (Appendix F)
  • Decoherence and collapse mechanisms (Appendix G)
  • Vacuum energy dynamics (Appendix H)
  • Proposed experimental tests (Appendix I)
  • Detailed multi-loop renormalisation group analyses (Appendix J)
  • Finite element simulations (Appendix K)
  • Nonperturbative analyses revealing solitonic structures (Appendix L)
  • Revised Derivation of Einstein Field Equations in the Refined STM Model (Appendix M)
Finally, an updated Appendix N serves as the Glossary of Symbols, ensuring clarity and consistency of notation throughout this work.
By demonstrating explicitly how quantum field theory structures—such as the Schrödinger equation, Born rule, gauge symmetries, and CP violation—can emerge deterministically from classical elasticity dynamics alone, we aim to offer researchers a coherent theoretical framework for exploring the quantum–classical boundary. Furthermore, the modified Einstein field equations described in Appendix M show how large-scale gravitational phenomena (time dilation, black hole formation, and cosmological expansion) tie in seamlessly with higher-order elasticity at the sub-Planck level, thus unifying quantum and gravitational perspectives without invoking extra dimensions or intrinsic probabilism. Although we present substantial progress over the original STM model, key formal and phenomenological questions remain—for example, fully reconciling spinor interactions, higher-order PDE stability, and black hole thermodynamics under the same rigorous self-adjoint Hamiltonian. Addressing these will be crucial to confirming the model’s consistency at all scales.
We encourage further numerical and experimental tests to validate, refine, or challenge the refined STM model, potentially offering a new deterministic route toward reconciling quantum phenomena and gravitational curvature within a single continuum elasticity theory.

2. Methods

In the refined Space–Time Membrane (STM) model, spacetime is formulated as an elastic membrane described by a deterministic high-order wave equation. Below, we detail the classical framework, discuss the operator quantisation necessary for quantum-like effects, show how gauge fields emerge from the membrane’s internal degrees of freedom, and outline the renormalisation programme. Finally, we comment on the classical limit, where the refined STM model reverts to a standard (though higher-order) elastic PDE.

2.1. Classical Framework and Lagrangian

2.1.1. Displacement Field and Equation of Motion

We begin with the real-valued displacement field u x , t , which represents local deformations of the four-dimensional membrane. The refined STM equation explicitly reads:
ρ 2 u t 2 [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] 4 u + η 6 u γ u t + λ u 3 g u Ψ ¯ t Ψ + F e x t ( x , t ) = 0 ,
where:
6.
ρ is the membrane’s mass density.
7.
E S T M ( μ ) is a scale-dependent baseline modulus, and Δ E x , t ; μ represents local stiffness variations.
8.
η 6 u suppresses short-wavelength fluctuations, acting as an ultraviolet (UV) regulator.
9.
γ u t provides damping, extendable to non-Markovian kernels.
10.
Nonlinear terms λ u 3 and g u Ψ ¯ Ψ encode self-interaction and coupling to spinor fields, respectively.
11.
F e x t x , t captures additional external forces or potential-driven effects (Appendix B).
This broad equation unifies gravitational curvature (on large scales) with quantum-like oscillations (at sub-Planck scales) through classical elasticity. It generalises earlier STM formulations by incorporating both higher-order derivatives (up to 6 ) and scale-dependent parameters.

2.1.2. Lagrangian Density

A convenient route to the equation of motion is through a Lagrangian density L . Ignoring explicit damping γ t u and external forcing for simplicity, one might write:
L = 1 2 ρ ( t u ) 2 1 2 [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] ( 2 u ) 2 η 2 ( 3 u ) 2 V ( u ) ,
where 2 u and 3 u are the Laplacian and its next derivative, respectively. Varying the corresponding action S = d 4 x L via the Euler–Lagrange equation reproduces the high-order PDE once boundary terms are accounted for. In practice, the presence of damping, spinor coupling g u Ψ ¯ Ψ , and memory kernels can be added to the action or treated at the level of an effective Hamiltonian.

2.1.3. Conjugate Momentum and Modified Dispersion

From the Lagrangian, the conjugate momentum is
π ( x , t ) = L ( t u ) = ρ t u ( x , t ) .
In homogeneous settings (ignoring Δ E x , t ; μ ), one can look at plane-wave solutions e i k · x ω t . The dispersion relation typically generalises to ω 2 k c 2 k 2 + E S T M ( μ ) k 4 + η k 6 . This extra k 6 term strongly regularises high wavenumbers, mitigating UV divergences.
While the present section derives the refined STM wave equation from higher-order elasticity, Appendix M further demonstrates how these same high-order terms yield Einstein-like field equations upon interpreting the membrane’s strain as a metric perturbation at large scales

2.2. Operator Quantisation

2.2.1. Canonical Commutation Relations

To capture the quantum-like aspects, we elevate u x , t and π x , t = ρ t u x , t to operator status: u ^ x , t and π ^ x , t . They obey equal-time commutation relations:
u ^ ( x , t ) , π ^ ( y , t ) ] = i δ 3 ( x y ) ,
with all other commutators vanishing. Higher-order derivatives ( 4 u , 6 u ) affect the Hamiltonian’s domain of self-adjointness, but the underlying canonical structure remains intact.

2.2.2. Normal Mode Expansion

In simpler (near-homogeneous) regimes, we can write
u ^ ( k , t ) = d 3 x ( 2 π ) 3 e i k · x u ^ ( x , t ) ,
and similarly for π ^ k , t . The Hamiltonian then decomposes into modes, each with a modified dispersion ω 2 k . However, including Δ E x , t ; μ generally requires more advanced techniques (Bloch expansions, real-space diagonalisation, or finite element methods; see also Appendix K).

2.3. Gauge Symmetries and Path Integral Formulation

2.3.1. Gauge Fields from Bimodal Spinors

A key element of the refined STM model is the bimodal decomposition of u ^ x , t , whereby we construct a two-component spinor Ψ x , t . If one imposes local phase invariance on Ψ x , t , gauge fields A μ (U(1)), W μ a (SU(2)), or G μ a (SU(3)) naturally arise [1, 14]. The extended Lagrangian includes covariant derivatives D μ = μ i g A μ and field strength tensors. One can then quantise these fields in a path integral, introducing ghost fields to handle gauge fixing:
Z = D u D A μ D c D c ¯ exp [ i d 4 x ( L S T M + L g a u g e ) ] .

2.3.2. Virtual Bosons as Oscillatory Modes

Where standard quantum field theory views virtual particles as fleeting excitations, the STM perspective interprets them as deterministic wave–plus–anti-wave cycles of the membrane. Energy conservation remains exact over each cycle, removing the need for stochastic quantum fluctuations [18].

2.4. Renormalisation and Higher-Order Corrections

2.4.1. One-Loop and Multi-Loop Analyses

The higher-order terms ( 4 , 6 ) in the STM PDE are crucial for controlling UV divergences. In momentum space, the propagator 1 / [ c 2 k 2 + E S T M ( μ ) k 4 + η k 6 ] decays rapidly for large k . One-loop beta functions can be computed using dimensional regularisation (Appendix J), typically yielding:
β ( g ) = a g 3 + b g 5 + ,
where a , b depend on integrals that include the extra damping from k 4 , k 6 . Multi-loop extensions (up to three loops) refine the flow of λ , u –spinor couplings, potentially unveiling discrete RG fixed points linked to fermion mass scales [15, 16].

2.4.2. Nonperturbative Functional Renormalisation Group (FRG)

Beyond perturbation theory, one applies FRG techniques (Appendix L). This can expose solitonic solutions (kinks, domain walls) in effective potentials, demonstrating how three separate vacuum states might lead to three fermion generations, or how topological defects could anchor CP violation. In black hole analogues, the extra stiffness from 6 halts singularity formation, letting the interior settle into finite-energy standing waves.

2.5. Classical Limit and Stationary-Phase Approximation

In the limit 0 or large damping, the path integral is dominated by stationary-phase solutions of the membrane PDE. Varying the classical action S = d 4 x L S T M yields the Euler–Lagrange equation reproducing
ρ 2 u t 2 [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] 4 u + η 6 u + V ( u ) = 0 ,
once boundary contributions are accounted for. This underscores the deterministic nature of the refined STM model at macroscopic scales, with quantum-like behaviour re-emerging upon coarse-graining the sub-Planck oscillations.

Summary of Methods

  • Classical Equation: A high-order PDE that includes 4 and 6 derivatives, scale-dependent stiffness, damping, and nonlinear couplings to spinors.
  • Lagrangian / Hamiltonian Formulations: Variation of a suitably extended Lagrangian (with or without damping) recovers the full STM wave equation.
  • Operator Quantisation: Promoting u , π to operators retains canonical commutators, though self-adjointness requires attention to boundary conditions.
  • Gauge Extensions: Bimodal spinor fields demand local phase invariance, introducing U(1), SU(2), SU(3) gauge fields in a deterministic wave interpretation.
  • Renormalisation: Higher-order derivatives ensure stronger UV suppression, reducing divergences in loop integrals. FRG reveals soliton solutions and discrete vacuum structures.
  • Classical–Quantum Transition: At large scales or 0 , the model reverts to a classical PDE. Sub-Planck oscillations, once coarse-grained, yield Schrödinger-like dynamics, decoherence, and gauge interactions—without intrinsic randomness.
This methodological foundation sets the stage for the model’s key results, spanning emergent quantum features, gauge symmetries, solitonic black hole interiors, and the possible generation of multiple fermion families in a single deterministic elasticity framework.

3. Results

This section presents the principal findings of the refined Space–Time Membrane (STM) model. We begin by examining perturbative results, illustrating how quantum-like dynamics, gauge symmetries, and deterministic decoherence arise from a high-order elasticity framework. We then turn to nonperturbative effects, whose full derivation—via the Functional Renormalisation Group (FRG)—appears in Appendix L.

3.1. Perturbative Results

3.1.0.1. 3.1.1 Emergent Schrödinger-Like Dynamics and the Born Rule

By coarse-graining the rapid, sub-Planck oscillations in u x , t , one obtains a slowly varying “envelope” Ψ x , t . Specifically, one applies a smoothing kernel (often Gaussian) and adopts a WKB-type ansatz,
Ψ ( x , t ) = A ( x , t ) exp [ i S ( x , t ) / ] .
Substituting Ψ x , t into the refined STM wave equation—now including E S T M ( μ ) + Δ E x , t ; μ 4 u , η 6 u , and other terms—leads to a separation into real and imaginary parts. The real part typically yields a Hamilton–Jacobi-type equation for the phase S x , t , while the imaginary part yields a continuity equation for A x , t .
At leading order, these can be combined into an effective Schrödinger-like equation:
i Ψ t = 2 2 m e f f 2 Ψ + V e f f ( x ) Ψ ,
where m e f f and V e f f x reflect the membrane’s elastic parameters and the self-interaction potential V ( u ) . Crucially, η 6 modifies the high-momentum dispersion, ensuring UV stability. The Born rule naturally follows by interpreting Ψ 2 as a probability density, derived here from deterministic sub-Planck chaos rather than postulated randomness [9, 12].
While this deterministic approach reproduces many quantum-like features, it deviates from the mainstream view of intrinsic quantum randomness. Further theoretical and experimental efforts (e.g. careful tests of Bell inequalities under non-Markovian conditions) are needed to confirm whether the STM model can fully match standard quantum mechanics at all scales.

3.1.1. Emergent Gauge Symmetries

A hallmark of the refined STM model is the emergence of gauge symmetries from the bimodal decomposition of the membrane displacement field u ( x , t ) . This decomposition naturally produces a two-component spinor field, Ψ ( x , t ) . Enforcing local phase invariance on Ψ ( x , t ) necessitates the introduction of gauge fields. For example, under the transformation Ψ ( x , t ) e i θ ( x , t ) Ψ ( x , t ) , a local U(1) symmetry emerges explicitly, requiring the introduction of a gauge field A μ ( x , t ) via the minimal substitution μ D μ = μ i e A μ . Extending this principle to non-Abelian symmetries naturally leads to the SU(2) and SU(3) Yang–Mills gauge structures. Consequently, excitations analogous to photons, W ± bosons, and gluons emerge deterministically as coherent wave modes of the membrane [1, 14].
For the weak interaction, the spinor structure explicitly enforces a local SU(2) gauge symmetry. When the displacement field acquires a vacuum expectation value, deterministic cross-membrane interactions between spinor fields and their mirror antispinor counterparts produce electroweak symmetry breaking. These interactions involve rapid oscillatory exchanges known as zitterbewegung, which deterministically generate the mass terms for the W ± and Z 0 gauge bosons. This deterministic mechanism avoids intrinsic quantum randomness and eliminates the need for additional scalar fields.
The strong interaction can be intuitively understood by considering the membrane as a classical lattice of linked oscillators. Within this analogy, each oscillator corresponds to a local ’colour charge’. The elastic tension between oscillators increases linearly with their separation, naturally reproducing the confinement phenomenon observed in Quantum Chromodynamics (QCD). Gluon-like modes thus arise as coherent elastic waves propagating along these oscillator connections, effectively ensuring colour confinement and preventing isolated coloured excitations from existing freely.
In this deterministic elasticity framework, processes traditionally described in standard quantum field theory as "virtual boson exchanges" are reinterpreted explicitly as coherent cycles of wave–anti-wave excitations within the membrane.
The explicit details of electroweak symmetry breaking and the emergence of the Z boson via deterministic spinor–antispinor interactions are developed fully in Appendix C.3.1.
Nevertheless, matching all known QFT scattering amplitudes (traditionally computed via Feynman diagrams) remains a major open task. The STM’s classical reinterpretation of virtual particles must quantitatively reproduce S-matrix elements, cross sections, and loop corrections for a robust equivalence with the Standard Model.

3.1.2. Deterministic Decoherence and Bell Inequality Violations

By splitting the membrane displacement into a slow system  u S x , t and a fast environment  u E x , t (Appendix G), one can integrate out u E via the Feynman–Vernon influence functional. This produces a non-Markovian master equation for the reduced density matrix ρ S ( t ) :
d ρ S d t = i [ H S , ρ S ] 0 t d τ K ( t τ ) D [ ρ S ( τ ) ] ,
where the kernel K encodes finite correlation times. This yields deterministic decoherence, allowing the apparent wavefunction collapse to occur without intrinsic randomness. Introducing spinor-based measurement operators (e.g. M ^ ( θ ) = c o s θ σ x + s i n θ σ z ) recovers Bell-type correlations. Indeed, the CHSH parameter can reach 2 2 , violating the classical Bell inequality [17] while still emerging from a deterministic PDE.
Although the STM model reproduces these correlations at a theoretical level, future studies must compare predicted decoherence rates and memory kernels with real quantum systems, which often show near-Markovian behaviour. The quantitative match to laboratory timescales and environment-induced superselection rules remains an important open topic.

3.1.3. Fermion Generations and CP Violation

Multi-loop renormalisation analyses (Appendix J) reveal that the running of elastic parameters, along with self-interactions λ u 3 or Yukawa-like couplings, can develop discrete fixed points. These correspond to multiple stable vacua, naturally identified with distinct mass scales seen in the three fermion generations [15, 16].
In addition, deterministic interactions between the bimodal spinor Ψ ( x , t ) on our membrane face and the mirror antispinor Ψ ( x , t ) on the opposite face induce rapid oscillatory exchanges, known as zitterbewegung. These exchanges result in complex, spatially and temporally averaged phases in Yukawa couplings, ultimately producing CP violation analogous to the CKM-type mixing observed experimentally.
The weak gauge bosons and electroweak mixing emerge clearly as deterministic outcomes from elastic interactions within the membrane’s spinor–antispinor framework. Complex phases arising naturally through zitterbewegung effects provide deterministic explanations for CP violation (Appendix C.3.1).
However, a detailed numerical fit to the observed quark and lepton mass hierarchies, as well as CKM and PMNS mixing angles, has not yet been carried out. The existence of discrete fixed points is suggestive, but further work is needed to match precise mass ratios and mixing phases of the Standard Model.

3.2. Nonperturbative Effects

To address dynamics beyond perturbation theory, the refined STM model leverages Functional Renormalisation Group (FRG) methods (Appendix L). In the Local Potential Approximation (LPA), one analyses how the effective potential V k ( ϕ ) evolves with the momentum scale k. This approach uncovers:
  • Solitonic Solutions (Kinks): For a double-well or multi-well potential, the classical equation in one spatial dimension admits kink solutions. These topological defects carry finite energy and can serve as boundaries between different vacuum states.
  • Discrete Vacuum Structure: Multiple minima in V k imply discrete vacua, each yielding different mass scales. Coupled to spinor fields, these vacua underpin the three fermion generations, while the topological defects can insert nontrivial phases relevant to CP violation.
  • Black Hole Interior Stabilisation: In gravitational collapse analogues, local stiffening from 4 and 6 halts singularity formation, replacing it with finite-amplitude “standing wave” or solitonic cores. This mechanism maintains energy conservation and potentially resolves the black hole information paradox.
A detailed derivation of these nonperturbative results is presented in Appendix L, showing how topological defects and FRG flows interplay to give rise to mass hierarchies, discrete RG fixed points, and stable kink configurations. Nevertheless, reproducing black hole thermodynamics (e.g. Bekenstein–Hawking entropy) or Hawking radiation from these solitonic solutions has not yet been demonstrated, so the thermodynamic consistency of soliton-based black holes remains an open question.
Our treatment here focuses on solitonic structures in the membrane’s displacement field. For a complementary perspective showing how these solitons manifest as curvature regularisation in an emergent spacetime geometry, see Appendix M for the Einstein-like derivation

3.3. Summary

  • Perturbative Results:
    • Effective Schrödinger Equation: Coarse-graining sub-Planck dynamics yields quantum-like envelopes, recovering interference and the Born rule.
    • Emergent Gauge Symmetries: Bimodal spinor decompositions necessitate U(1), SU(2), and SU(3), reproducing photon-like and gluon-like fields.
    • Deterministic Decoherence and Bell Violations: A non-Markovian master equation explains apparent wavefunction collapse and entanglement in a classical continuum setting.
    • Fermion Generations and CP Violation: Multi-loop RG analysis identifies discrete fixed points corresponding to distinct vacuum structures, naturally explaining multiple fermion generations. CP violation emerges deterministically through interactions between the membrane’s spinor fields and mirror antispinors, mediated by zitterbewegung-induced complex Yukawa coupling phases.
  • Nonperturbative Insights:
    • Solitons and Kinks: FRG shows stable topological defects that can anchor vacuum structure, linking discrete mass scales to elastic domain walls.
    • Avoiding Singularities: Enhanced stiffness ( 6 regularisation) prevents unbounded collapse, offering finite-energy cores in black hole analogues.
    • New Mechanisms for CP Violation: Solitonic vacua provide additional phases, unifying mass hierarchies and CP effects in an elasticity-based approach.
Altogether, the refined STM model demonstrates how a single deterministic PDE—encompassing higher-order derivatives, scale-dependent elasticity, and spinor couplings—can replicate core features of quantum field theory and gravitational phenomena. The key is that quantum-like behaviour emerges from chaotic sub-Planck oscillations upon coarse-graining, rather than from fundamental randomness or extra dimensions.

4. Discussion

The refined STM model explicitly illustrates how deterministic, classical chaos in membrane oscillations directly reproduces quantum phenomena such as wavefunction collapse, interference, and the Born rule. This deterministic elasticity thus explicitly offers a clear physical reinterpretation of quantum randomness, removing the need for inherent stochastic assumptions.
The model represents a bold attempt to unify gravitational curvature with quantum-like phenomena within a single deterministic framework based on high-order elasticity. By incorporating second-, fourth-, and sixth-order spatial derivatives, scale-dependent parameters, and non-Markovian effects, we find that many hallmark features of quantum field theory can emerge naturally from the membrane’s classical dynamics.
Below, we examine the implications of these findings, compare them with standard quantum field theory, and consider practical routes toward experimental validation.

4.1. Emergent Quantum Dynamics and Decoherence

A key aspect of our perturbative analysis is that by coarse-graining the rapid, sub-Planck oscillations of the membrane’s displacement field u ( x , t ) , one obtains a slowly varying envelope Ψ ( x , t ) . This envelope obeys an effective Schrödinger-like equation,
i Ψ t = 2 2 m e f f 2 Ψ + V e f f ( x ) Ψ ,
mimicking the familiar quantum mechanical form. Crucially, the sixth-order spatial derivative 6 u in the refined STM wave equation dampens short-wavelength modes, ensuring that ultraviolet divergences do not arise. Moreover, the Born rule emerges through deterministic chaos at sub-Planck scales, replacing the postulated randomness of conventional quantum theory.
By splitting u ( x , t ) into a system component u S and an environment u E , we further showed that non-Markovian decoherence follows from integrating out the fast modes u E . This framework reproduces “wavefunction collapse” as an effective phenomenon, caused by memory kernels that gradually suppress off-diagonal terms in the reduced density matrix, all within a deterministic PDE context. Notably, as soon as we implement spinor-based measurement operators and allow for correlated sub-Planck modes, the model achieves Bell inequality violations (CHSH up to 2 2 ) in a purely classical wave setting.
Although these features closely mimic quantum mechanical predictions, mainstream interpretations hold randomness as fundamental. Additional experiments and theoretical checks will be needed to see if STM-based deterministic decoherence can match all observed quantum phenomena (e.g. precise decoherence timescales) without contradiction.

4.2. Emergence of Gauge Symmetries and Virtual Boson Reinterpretation

Through a bimodal decomposition of the displacement field, the refined STM model constructs a spinor Ψ ( x , t ) . Requiring local phase invariance on Ψ naturally introduces gauge fields corresponding to U ( 1 ) , S U ( 2 ) , or S U ( 3 ) [1, 14]. Consequently, photon-like and gluon-like excitations arise as deterministic wave modes rather than quantum fluctuations. Meanwhile, the usual concept of virtual bosons—pertinent to standard quantum field exchanges—is replaced by wave–plus–anti-wave oscillations that transfer no net energy over a full cycle [18]. This classical reinterpretation preserves energy conservation at every instant and bypasses the notion of “transient particle creation,” typical of conventional perturbation theory.
This reinterpretation also clarifies how force mediation, in particular electromagnetism and the strong interaction, can be understood as elastic “connections” in a high-order continuum. The refined STM PDE itself underlies these gauge fields once spinor local symmetries are introduced. Thus, standard gauge bosons like photons, W ± , or gluons appear as coherent membrane oscillations, illustrating how quantum-like gauge interactions might emerge from deterministic elasticity.
For the strong force specifically, visualising the membrane as a chain or lattice of linked oscillators clarifies how confinement arises deterministically from classical elasticity. Each lattice site can be regarded as carrying a colour charge, and the coupling between these sites stiffens rapidly with increasing distance. This property prevents the separation of colour charges into free isolated states, directly mimicking the linear potential and confinement behaviour central to QCD. Deterministic gluon-like excitations, represented by coherent waves propagating along oscillator links, thereby mediate the strong interaction without requiring intrinsic randomness or virtual particle fluctuations.
While this approach elegantly reinterprets gauge fields, verifying quantitative equivalence with the Standard Model’s scattering amplitudes and loop processes is crucial. Detailed calculations would need to show that these “wave–anti-wave” cycles match Feynman diagram predictions at all energy scales.

4.3. Fermion Generations and CP Violation

Our multi-loop renormalisation analysis (Appendix J) identifies discrete RG fixed points in the running of the membrane’s elastic parameters and couplings. Each fixed point corresponds naturally to a distinct vacuum structure, offering an explanation for three separate fermion mass scales akin to the three observed generations [15, 16]. In this refined STM model, fermion masses and CP violation arise deterministically from interactions between the membrane’s bimodal spinor field Ψ ( x , t ) and the corresponding mirror antispinor field Ψ ( x , t ) . Rapid oscillatory exchanges (zitterbewegung effects) between these spinor fields induce complex phase shifts in effective Yukawa-like couplings. Diagonalising the resulting fermion mass matrix yields nonzero CP-violating phases, closely mirroring the observed CKM structure in the Standard Model. Thus, the refined STM model provides a deterministic elasticity-based mechanism for both the flavour structure of fermion generations and the emergence of CP violation, eliminating the need for inherently stochastic or extra-dimensional assumptions.
However, a thorough numerical match to the precise mass ratios and mixing angles (CKM and PMNS) remains to be demonstrated. Achieving that level of detail is essential for confirming that zitterbewegung-based complex phases fully replicate observed CP violation.

4.4. Matter Coupling and Energy Conservation

The refined STM framework introduces explicit Yukawa-like interactions g u Ψ ¯ Ψ to couple the membrane’s displacement field to emergent fermionic degrees of freedom. In this way, fermion masses become part of the membrane’s global elastic response, ensuring full energy conservation at every step—particularly relevant in processes traditionally involving virtual particle exchange. The inclusion of the 6 derivative remains essential for limiting high-momentum contributions, thus keeping the theory stable and unitary.
This perspective also adds clarity to phenomena where energy conservation might appear temporarily suspended in standard perturbative diagrams. In the STM picture, each wave–plus–anti-wave cycle balances out net energy transfer over its period, precluding ephemeral violations yet reproducing the same effective scattering amplitudes.

4.5. Further Phenomena and Interpretations

Beyond the core predictions detailed above, the refined STM model suggests new ways to interpret certain key features of the Standard Model. For instance, where conventional theory employs an elementary Higgs scalar to generate gauge boson and fermion masses, the STM approach attributes electroweak symmetry breaking to rapid zitterbewegung interactions between spinor and mirror antispinor fields, potentially offering an alternative explanation of the Higgs boson resonance observed at 125 GeV. A quantitative mapping between the observed Higgs signal and an STM “effective scalar mode,” however, remains a challenging open problem.
Similarly, the Pauli exclusion principle may emerge from boundary conditions that enforce antisymmetric combinations of spinor wavefunctions on the membrane, while Heisenberg’s uncertainty principle, in STM, can be viewed as a macro-scale manifestation of deterministic sub-Planck chaotic dynamics. Finally, the non-trivial scale-dependent stiffness Δ E\Delta E Δ E introduced in our model naturally interprets dark energy (Appendix H) as a persistent elastic vacuum offset. Although these interpretations require further numerical and conceptual validation, they illustrate how STM’s deterministic elasticity might unify several phenomena typically attributed to intrinsic quantum or field-theoretic mechanisms.

4.6. Experimental and Numerical Prospects

To advance beyond conceptual arguments, the refined STM model suggests several concrete tests:
  • Metamaterial Analogues:
    Laboratory experiments using acoustic or optical metamaterials can replicate the essential PDE structure, including higher-order dispersion and nonlinear feedback. Observing deterministic decoherence phenomena or stable interference nodes in such media would support the STM approach. Nevertheless, purely classical analogues may not fully capture true quantum entanglement or the precise Markov-to-non-Markov transitions. Designing metamaterials that emulate 6 terms accurately is also a significant technical challenge.
  • Finite Element Simulations:
    Numerical implementations (Appendix K) allow one to solve the refined STM equation—including 4 , 6 , and scale-dependent stiffness—under realistic boundary conditions. Matching simulated ringdowns or soliton formation to measured data can constrain the model’s parameters.
  • Astrophysical Observations:
    Black hole mergers recorded by gravitational wave detectors (e.g. LIGO, Virgo) may carry signatures of interior soliton structures (Appendix F). Potential ringdown frequency shifts or unusual damping profiles could reflect additional stiffness near horizons, consistent with the refined STM’s avoidance of singularities. Meanwhile, cosmic microwave background anisotropies might reveal subtle vacuum energy inhomogeneities predicted by scale-dependent elasticity. However, the magnitude of such ringdown modifications may be quite small, possibly below current detector sensitivity. Future instruments (e.g. Einstein Telescope) might be required to rule them in or out.
Further testing avenues—such as short-range torsion balance experiments or precision atomic clock comparisons—are discussed in Appendix I, where we elaborate on the Einstein-like corrections introduced by scale-dependent elasticity.

4.7. Theoretical Implications and Future Directions

Our results suggest that the apparent randomness at the heart of quantum mechanics might be an emergent by-product of coarse-graining sub-Planck chaos within a deterministic PDE framework. This fresh view, alongside the re-interpretation of force mediation and the natural rise of gauge symmetries, offers a potent alternative to conventional quantum field theory. Several lines of research remain open:
  • Refining Operator Quantisation: A deeper exploration of boundary conditions and higher loops in the presence of 6 terms would clarify unitarity and self-adjointness in large volumes or curved geometries. Ensuring no ghost-like degrees of freedom appear is a critical open problem for higher-order theories.
  • Extending Nonperturbative Analysis: Incorporating additional interactions or spontaneously broken symmetries could illuminate chiral structures and anomaly cancellations.
  • Designing Rigorous Experimental Tests: Both tabletop metamaterial experiments and advanced gravitational wave observations stand poised to probe the predictions of the refined STM model.
Even though we have shown how the core elastic PDE can be made self-adjoint under suitable Sobolev boundary conditions (killing boundary terms, etc.), the presence of nontrivial spinor couplings, non-Abelian gauge fields, and higher-order nonlinearities raises further questions about overall stability and ghost freedom. A full operator formalism must guarantee that once these gauge and Yukawa-like terms are introduced, the Hamiltonian remains self-adjoint, with no indefinite norm states or hidden anomalies. Addressing these issues would secure the deeper consistency of our deterministic PDE approach and ensure that all emergent gauge and spinor fields fit seamlessly into a stable, unitary quantum framework.
In conclusion, by merging quantum-like features with classical elasticity, the refined STM approach rethinks the quantum–classical boundary, attributing wavefunction collapse to deterministic decoherence and virtual particles to oscillatory wave pairs. Such a unification challenges deeply held assumptions about randomness in quantum theory, while supplying a fresh route to reconciling gravitation with field-theoretic phenomena—without invoking extra dimensions or intrinsic probabilism.
Additionally, we note that a more complete derivation of the Einstein-like field equations—beyond the original linearised approach—can now be found in Appendix M, where the membrane’s stress–energy is shown to produce Einstein-like equations at large scales while incorporating higher-order corrections. For a consolidated list of the refined STM notation, see Appendix N.
Finally, while black hole singularity avoidance via solitonic cores is conceptually appealing, demonstrating the correct thermodynamic relations (e.g. Bekenstein–Hawking entropy) or Hawking-like radiation would be an essential step to ensure full consistency with established black hole physics.

5. Conclusions

In this paper, we have presented a refined Space–Time Membrane (STM) model that seeks to bridge the gap between gravitational curvature and quantum field phenomena through a deterministic framework based on classical elasticity. Building on the original STM concept—where a single four-dimensional elastic membrane gave rise to both interference and gravitational effects—we have significantly extended the theory. Our refinements include the introduction of scale-dependent elastic moduli E S T M ( μ ) and Δ E ( x , t ; μ ) , the incorporation of higher-order spatial derivative terms (notably the 6 operator) to suppress ultraviolet divergences, and the implementation of non-Markovian decoherence mechanisms. These enhancements culminate in a high-order wave equation whose deterministic sub-Planck dynamics, upon coarse-graining, yield an effective Schrödinger-like evolution and the natural emergence of the Born rule without recourse to intrinsic randomness. Wavefunction collapse is reinterpreted as deterministic decoherence resulting from environmental coupling.
A key innovation of our approach is the bimodal decomposition of the displacement field u ( x , t ) , which naturally gives rise to a two-component spinor Ψ ( x , t ) . This spinor structure underpins the emergence of internal gauge symmetries; through the imposition of local phase invariance, gauge fields corresponding to U ( 1 ) , S U ( 2 ) and S U ( 3 ) appear as deterministic wave–plus–anti-wave modes.
In particular, the strong interaction is understood through a straightforward classical analogy, where colour confinement emerges naturally from linear tension in a discretised lattice of oscillator-like membrane elements. This interpretation clearly demonstrates how gluon-like excitations appear as deterministic wave modes enforcing confinement, aligning closely with observed properties of quantum chromodynamics.
Electroweak symmetry breaking, the emergence of massive weak bosons ( W ± , Z 0 ), and CP violation occur naturally and deterministically via the interaction between bimodal spinor fields and mirror antispinors across the membrane, mediated by zitterbewegung-induced complex phases in Yukawa couplings.
In this way, classical elastic waves are reinterpreted as the force carriers of quantum field theory, with virtual boson exchange emerging from coherent oscillatory cycles that maintain zero net energy exchange over a full period.
Our renormalisation group analysis, detailed in Appendix J, shows that the inclusion of the 6 term is essential for controlling divergent loop integrals. The running of the elastic parameters is governed by beta functions that exhibit nontrivial fixed points. These fixed points may provide a natural mechanism for generating a discrete mass spectrum, thereby offering a potential explanation for the existence of three fermion generations. Moreover, when combined with nonlinear self-interactions (such as the λ u 3 term) and Yukawa-like couplings (of the form g u Ψ ¯ Ψ ), our model captures key features of fermion–boson dynamics within a deterministic framework.
The refined STM model also addresses the long-standing problem of singularity formation in gravitational collapse. As matter density increases, the effective local stiffness of the membrane—augmented by Δ E ( x , t ; μ ) —rises sharply, and the higher-order 6 term suppresses short-wavelength fluctuations, thereby regularising the curvature. Consequently, rather than developing a classical singularity, the system relaxes into finite-amplitude standing wave configurations or solitonic cores. These solitonic solutions not only provide a mechanism for singularity avoidance in black hole interiors but also offer a novel perspective on the preservation of information during gravitational collapse.
Furthermore, by decomposing the displacement field into slowly varying system modes and rapidly fluctuating environmental modes, and subsequently integrating out the latter using the Feynman–Vernon influence functional formalism, we derive a non-Markovian master equation. This equation accounts for environmental memory effects and leads to deterministic decoherence. The gradual decay of off-diagonal elements in the reduced density matrix replicates wavefunction collapse, thereby reproducing a key quantum phenomenon without introducing intrinsic randomness. When spinor-based measurement operators are introduced, the model even reproduces Bell inequality violations in a manner consistent with standard quantum mechanics.
The STM model explicitly demonstrates that deterministic chaotic elasticity alone can generate quantum-like phenomena, providing clear intuitive analogies for interference, wavefunction collapse, and the Born rule without invoking inherent quantum randomness.

5.1. Key Achievements

  • Unified Framework for Gravitation and Quantum-Like Features:
    Large-scale curvature emerges from membrane bending, while quantum field behaviour is a macroscopic manifestation of deterministic, chaotic sub-Planck dynamics. This classical approach offers a fresh route to phenomena typically associated with probabilistic quantum mechanics.
  • Feasibility of Emergent Quantum Field Theory:
    Gauge bosons—such as photon-like, W ± -like, and gluon-like excitations—arise naturally from the spinor decomposition of the membrane’s displacement field. Our renormalisation analysis shows that running elastic parameters can mimic loop effects in standard quantum field theory, with fixed points that hint at a discrete mass spectrum corresponding to three fermion generations.
  • Path to Deterministic Decoherence:
    Environmental interactions, modelled through non-Markovian kernels, yield a master equation that reproduces effective wavefunction collapse without any intrinsic randomness.
  • Mechanism for Fermion Generation and CP Violation:
    The emergence of discrete RG fixed points, identified through multi-loop renormalisation analysis, naturally gives rise to three distinct fermion families. CP violation and the associated complex Yukawa couplings arise deterministically through rapid oscillatory interactions (zitterbewegung) between bimodal spinor fields on our membrane face and corresponding mirror antispinors on the opposite face. This deterministic interplay generates irreducible complex phases in the effective fermion mass matrix, closely reproducing the observed CP-violating structure of the Standard Model’s CKM matrix. Thus, the refined STM model provides a clear, deterministic elasticity-based explanation for both the origin of multiple fermion generations and the mechanism underlying CP violation, without invoking stochastic or higher-dimensional assumptions.
Additionally, reconciling solitonic black hole interiors with thermodynamic laws, such as the Bekenstein–Hawking entropy, is essential for the model’s viability in gravitational contexts.

5.2. Outstanding Limitations and Future Work

  • Rigorous Operator Quantisation
    While our operator approach—featuring a self-adjoint Hamiltonian and a preliminary path integral formulation—marks significant progress, a fully rigorous operator quantisation remains an open challenge. Ensuring that higher-order derivatives do not introduce negative norm states or ghosts is especially important.
    Moreover, we have established partial self-adjointness for the linear PDE under appropriate boundary conditions, but the true complexity lies in incorporating all emergent fields and interactions. Nonlinear elastic terms ( λ u 3 ), spinor bilinears Ψ ¯ Ψ , and non-Abelian gauge fields must be handled within the same functional-analytic framework. Each coupling can affect boundary conditions or introduce additional constraints (e.g., gauge fixing), which must not break the overall self-adjointness or stability of the Hamiltonian. A comprehensive proof of ghost freedom and unitarity—accounting for these couplings—would solidify the refined STM model’s foundations.
  • Multi-Loop and Nonperturbative RG Analysis
    Our renormalisation analysis explicitly extends up to three-loop corrections, supplemented by nonperturbative analyses performed using the Functional Renormalisation Group (FRG). However, further higher-order corrections and comprehensive investigations are required to conclusively validate the emergent quantum field structure and confirm phenomena such as asymptotic freedom alongside the fixed-point dynamics.
  • Detailed Treatment of Fermion Generations and CP Violation
    While the deterministic spinor–antispinor interaction mechanism detailed in Appendix C.3.1 provides a clear origin for CP violation and electroweak symmetry breaking, a comprehensive numerical fit of the full fermion mass spectrum and precise mixing angles (CKM and PMNS) remains to be completed. Such a detailed analysis is a key priority for future extensions of the STM model phenomenology.
  • Planck-Scale Validity
    It remains uncertain whether the classical continuum elasticity description underpinning the STM framework remains valid at or above the Planck energy scale. At such extremely high energies, additional physics, possibly involving a discrete spacetime substructure or quantum gravitational effects beyond elasticity, may become significant.
While effective damping γ t u can be interpreted via open quantum systems (coupling the membrane to an auxiliary environment), it is comparatively less fundamental than ensuring the full system (nonlinear elasticity + spinors + gauge fields) achieves a self-adjoint, ghost-free formulation at the quantum level. We envisage separate, more technical works that systematically integrate these couplings into a single operator framework, validating stability and unitarity with all emergent fields included.

5.3. Potential Experimental and Observational Tests

  • Finite Element Analysis:
    Numerical simulations (see Appendix K) can test whether a single set of STM parameters reproduces quantum-like interference and gravitational phenomena.
  • Metamaterial Analogues:
    Laboratory experiments using tunable optical or acoustic metamaterials may emulate deterministic interference and non-Markovian decoherence, providing a controlled environment to probe STM predictions. However, classical analogues may not fully capture genuine quantum entanglement or certain quantum field aspects, so caution must be applied when extrapolating results.
  • Astrophysical Observations:
    Gravitational wave data and cosmological surveys might reveal signatures of STM elasticity through modified black hole ringdowns, dark energy inhomogeneities, or other large-scale anomalies. Significant theoretical work is needed to predict how large these modifications might be and whether current detectors can observe them.
Further testing avenues—such as short-range torsion balance experiments or precision atomic clock comparisons—are discussed in Appendix I, where we elaborate on the Einstein-like corrections introduced by scale-dependent elasticity.

Concluding Remarks

The refined STM model explicitly presents a unified deterministic framework in which gravitational curvature and quantum-like phenomena both emerge naturally from classical continuum elasticity. Building upon the original STM concept—which successfully explained spacetime curvature and quantum interference through deterministic chaos and elastic interactions—the refined model advances this picture substantially. It now explicitly incorporates scale-dependent elastic parameters, higher-order derivative terms for ultraviolet regularisation, non-Markovian decoherence, and a bimodal spinor–antispinor decomposition of the membrane’s displacement field, from which U ( 1 ) , S U ( 2 ) , and S U ( 3 ) gauge symmetries explicitly arise.
Central to the refined STM model is the novel deterministic mechanism for electroweak symmetry breaking and CP violation, explicitly detailed in Appendix C.3.1. Here, deterministic interactions between spinor fields on our membrane face and mirror antispinors on the opposite face, mediated by rapid oscillations (zitterbewegung), explicitly yield effective Yukawa couplings containing irreducible complex phases. This deterministic process naturally generates the masses of the gauge bosons W ± and Z 0 , explicitly removing the need for intrinsic quantum randomness or additional scalar fields.
Furthermore, our explicit multi-loop renormalisation group (RG) analysis—extending up to three loops and complemented by a comprehensive Functional Renormalisation Group (FRG) nonperturbative study—identifies discrete fixed points and solitonic vacuum structures. These findings explicitly suggest a deterministic explanation for the three distinct fermion generations observed in nature, as well as a potential deterministic origin of CP violation.
Nonetheless, significant challenges remain. Rigorous operator quantisation, comprehensive higher-loop RG corrections beyond three loops, explicit numerical fits of the full fermion mass spectrum, and a detailed confirmation of asymptotic freedom are still outstanding. Additionally, the validity of the continuum elasticity framework at or beyond Planck-scale energies remains uncertain, indicating that additional fundamental physics or discrete spacetime substructures might become significant at these extreme scales.
In addition, demonstrating compliance with black hole thermodynamics, as well as confirming the model’s predictions for cosmic acceleration and Higgs-like phenomena, will be crucial steps for the refined STM approach.
Ultimately, the refined STM model stands as a highly intriguing framework that promises deterministic explanations for quantum phenomena and potential resolutions to deep-rooted issues in theoretical physics—ranging from black hole singularities and dark energy to electroweak symmetry breaking and CP violation.
Although the model remains in an exploratory stage, its elegant combination of classical elasticity and emergent quantum–gravitational effects opens the door to novel insights and experimental possibilities. In an era where conventional approaches to unifying gravity and quantum mechanics have seen limited progress, the refined STM approach offers a fresh viewpoint worthy of further development, rigorous testing, and innovative experimentation.
Researchers are thus encouraged to examine, challenge, and extend this deterministic framework, driving it toward increasingly quantitative and observationally falsifiable predictions.

Funding

The author received no specific funding for this work.

Institutional Review Board Statement

This study did not involve any ethically related subjects.

Data Availability Statement

All relevant data are contained within the paper and its supplementary information.

Acknowledgments

I would like to express my deepest gratitude to the scholars and researchers whose foundational work is cited in the references; their contributions have been instrumental in the development of the Space-Time Membrane (STM) model presented in this paper. I am grateful for the advanced computational tools and language models that facilitated the mathematical articulation of the STM model which I developed over the last 14 years. Lastly, I wish to pay homage to the late Isaac Asimov, whose writings sparked my enduring curiosity in physics and inspired me to pursue this line of inquiry.

Conflicts of Interest

The author declares that they have no conflicts of interest.

Use of Artificial Intelligence

During the preparation of this work the author used ChatGPT in order to improve readability of the paper and provide clear explanation of the mathematical derivations. After using this tool/service, the author reviewed and edited the content as needed and takes full responsibility for the content of the published article.

Appendix A. Operator Formalism and Spinor Field Construction

Appendix A.1. Overview

A central feature of the refined Space–Time Membrane (STM) model is the emergence of fermion-like spinor fields from a purely classical elastic membrane [1]. In this appendix, we detail how the classical displacement field u ( x , t ) – 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 u ^ ( x , t ) via canonical quantisation. We also define its conjugate momentum and introduce a complementary out-of-phase field u ( x , t ) . A bimodal decomposition of these fields subsequently yields a two-component spinor Ψ ( x , t ) , which forms the foundation for the emergence of internal gauge symmetries.

Appendix A.2. Canonical Quantisation of the Displacement Field

Appendix A.2.1. Classical Preliminaries

The classical displacement field u ( x , t ) 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
L = 1 2 ρ t u 2 1 2 E S T M ( μ ) + Δ E ( x , t ; μ ) 2 u 2 1 2 η 3 u 2 V ( u ) L i n t ,
where:
  • ρ is the effective mass density,
  • E S T M ( μ ) is the scale-dependent baseline elastic modulus,
  • Δ E ( x , t ; μ ) represents local stiffness variations,
  • The term 1 2 η 3 u 2 yields, via integration by parts, the sixth-order term η 6 u ,
  • V ( u ) is the potential energy (e.g. V ( u ) = 1 2 k u 2 or more complex forms incorporating nonlinearities such as λ u 3 ),
  • L i n t includes additional interaction terms such as the Yukawa-like coupling g u Ψ ¯ Ψ .
Damping ( γ t u ) and external forcing F e x t ( x , t ) are introduced separately or via effective dissipation functionals in the complete equation of motion:
ρ 2 u t 2 E S T M ( μ ) + Δ E ( x , t ; μ ) 4 u + η 6 u γ u t + λ u 3 g u Ψ ¯ Ψ + F e x t ( x , t ) = 0 .

Appendix A.2.2. Conjugate Momentum

The conjugate momentum is defined as
π ( x , t ) = L t u = ρ t u ( x , t ) .

Appendix A.2.3. Promotion to Operators

In quantising the theory, the classical field u ( x , t ) and its conjugate momentum π ( x , t ) are promoted to operators u ^ ( x , t ) and π ^ ( x , t ) acting on a Hilbert space H . They satisfy the canonical equal-time commutation relation
u ^ ( x , t ) , π ^ ( y , t ) = i δ 3 ( x y ) ,
with all other commutators vanishing [5, 11]. This structure remains valid when higher-order derivatives (leading to 4 and 6 terms) and scale-dependent parameters are included.

Appendix A.2.4. Normal Mode Expansion and Dispersion Relation

In a near-homogeneous scenario, the operator u ^ ( x , t ) is expressed in momentum space as
u ^ ( x , t ) = d 3 k ( 2 π ) 3 e i k · x u ^ ( k , t ) .
Substituting this expansion into the classical equations of motion yields the modified dispersion relation. For plane-wave solutions e i ( k · x ω t ) , one obtains
ω 2 ( k ) = c 2 k 2 + E S T M ( μ ) + Δ E ( x , t ; μ ) k 4 + η k 6 .
The inclusion of the η k 6 term, arising from the 3 u 2 contribution, provides enhanced UV regularisation by strongly suppressing high-wavenumber fluctuations.

Appendix A.2.5. Hamiltonian Operator

The Hamiltonian operator is constructed from the Lagrangian as
H ^ = d 3 x 1 2 ρ π ^ 2 ( x , t ) + c 2 2 u ^ ( x , t ) 2 + 1 2 E S T M ( μ ) + Δ E ( x , t ; μ ) 2 u ^ ( x , t ) 2 + η 2 3 u ^ ( x , t ) 2 + V u ^ ( x , t ) + L ^ i n t ,
where L ^ i n t represents the operator form of the interaction terms (including, for instance, the Yukawa-like coupling g u Ψ ¯ Ψ ). To ensure that all derivative terms (up to third order, corresponding to 6 ) are well defined, the domain of H ^ is chosen as a Sobolev space H 3 (or higher). With appropriate boundary conditions (e.g. fields vanishing at infinity), integration by parts guarantees that H ^ is self-adjoint and its spectrum is real and bounded from below.

Appendix A.3. Bimodal Decomposition and Spinor Construction

To capture additional internal degrees of freedom, we introduce a complementary field u ( x , t ) , interpreted as the out-of-phase (or quadrature) component of the membrane’s displacement. We define two new real fields via the linear combinations
u 1 ( x , t ) = 1 2 u ^ ( x , t ) + u ( x , t ) , u 2 ( x , t ) = 1 2 u ^ ( x , t ) u ( x , t ) .
These represent the in-phase and out-of-phase components, respectively. They are then combined into a two-component spinor operator
Ψ ( x , t ) = u 1 ( x , t ) u 2 ( x , t ) .
By imposing appropriate (anti)commutation relations between u ^ ( x , t ) and u ( x , t ) , one can demonstrate—by analogy with Fermi–Bose mappings in certain lower-dimensional systems—that the spinor Ψ ( x , t ) exhibits chiral substructures. These substructures are essential for the emergence of internal gauge symmetries.

Appendix A.4. Self-Adjointness and Path Integral Formulation

The Hamiltonian operator H ^ is shown to be self-adjoint by verifying that all higher-order derivative terms are well defined on the chosen Sobolev space (here, H 3 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
u f , t f u i , t i = D u exp i S [ u ] ,
with the action
S [ u ] = t i t f d t d 3 x L [ u ] .
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.

Appendix 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 A μ a ( x , t ) (where a indexes the generators), the covariant derivative is defined as
D μ = μ i g A μ a ( x , t ) T a ,
with T a representing the generators (for example, T a = σ a / 2 for SU(2) or T a = λ a / 2 for SU(3)) and g the gauge coupling constant. The corresponding field strength tensor is given by
F μ ν a = μ A ν a ν A μ a i g f a b c A μ b A ν c .
The gauge symmetry is quantised by imposing a gauge-fixing condition (e.g. the Lorentz gauge μ A μ a = 0 ) and by introducing Faddeev–Popov ghost fields c a and c ¯ a . The resulting gauge-fixed path integral is
Z = D u D A μ D c ¯ D c exp i S e f f [ u , A μ , c , c ¯ ] ,
where S e f f includes the original STM Lagrangian, the gauge field Lagrangian, and the ghost contributions.

Appendix A.5.1. Summary and Outlook

In summary, the operator quantisation scheme for the refined STM model proceeds as follows:
  • Displacement Field Promotion:
    The classical displacement field u ( x , t ) and its conjugate momentum π ( x , t ) are promoted to operators u ^ ( x , t ) and π ^ ( x , t ) on a Hilbert space. The domain is chosen as a suitable Sobolev space (e.g. H 3 or higher) to ensure that all derivatives up to third order (which produce the 6 term) are well defined.
  • Complementary Field and Spinor Construction:
    A complementary field u ( x , t ) is introduced. By forming the in-phase and out-of-phase combinations u 1 ( x , t ) and u 2 ( x , t ) , a two-component spinor Ψ ( x , t ) is constructed. This spinor structure is central to the emergence of internal gauge symmetries.
  • Self-Adjoint Hamiltonian:
    The Hamiltonian H ^ 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 S [ u ] = d t d 3 x L [ u ] , 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 refined 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 Force Function

Appendix B.1. Overview

In the refined Space–Time Membrane (STM) model, the external force F ext ( x , t ) represents contributions to the membrane’s dynamics beyond its intrinsic elastic response. Such contributions arise from nonlinear interactions, including both self-interaction terms (e.g. a cubic term in u) and Yukawa-like couplings that mediate interactions between the membrane and emergent fermionic fields. In this appendix, we derive F ext ( x , t ) by performing a full functional variation of an extended potential energy functional U ext [ u , ψ ] with respect to the displacement field u ( x , t ) .

Appendix B.2. Extended Potential Energy Functional

We define the extended potential energy functional U ext [ u , ψ ] as follows:
U ext [ u , ψ ] = d 3 x Φ ( x , u , u ) + 1 2 χ ( x ) ( u ) 2 + N ( u , ψ ) ,
where:
  • Φ ( x , u , u ) is a local potential that may depend on both u and its spatial gradient u .
  • 1 2 χ ( x ) ( u ) 2 is a tension (or friction) term that is position-dependent, with χ ( x ) representing local variations in tension.
  • N ( u , ψ ) represents the nonlinear interaction terms, for which a representative choice is:
N ( u , ψ ) = λ u 3 + y u ψ ψ ,
  • with λ as the cubic self-interaction coupling, y as the Yukawa coupling constant, and ψ ψ denoting the standard fermion bilinear.

Appendix B.3. Functional Variation: Deriving

F ext ( x , t )
The external force F ext ( x , t ) is defined as the negative functional derivative of the potential energy functional with respect to u ( x , t ) :
F ext ( x , t ) = δ U ext [ u , ψ ] δ u ( x , t ) .
We now derive each contribution in detail.
(1) Contribution from the Local Potential Φ ( x , u , u ) :
Consider the term:
U Φ = d 3 x Φ ( x , u , u ) .
An infinitesimal variation u ( x , t ) u ( x , t ) + δ u ( x , t ) produces:
δ U Φ = d 3 x Φ u δ u + Φ ( u ) · δ ( u ) .
Since δ ( u ) = ( δ u ) , we have:
δ U Φ = d 3 x Φ u δ u + Φ ( u ) · ( δ u ) .
Integrating the second term by parts (and assuming that boundary contributions vanish), we obtain:
d 3 x Φ ( u ) · ( δ u ) = d 3 x · Φ ( u ) δ u .
Thus, the variation becomes:
δ U Φ = d 3 x Φ u · Φ ( u ) δ u .
Therefore, the force contribution from Φ is:
F Φ ( x , t ) = Φ u · Φ ( u ) .
(2) Contribution from the Tension Term 1 2 χ ( x ) ( u ) 2 :
Consider:
U χ = d 3 x 1 2 χ ( x ) ( u ) 2 .
Its variation under u ( x , t ) u ( x , t ) + δ u ( x , t ) is:
δ U χ = d 3 x χ ( x ) u ( x , t ) · δ u ( x , t ) .
Integrating by parts and neglecting boundary terms, we have:
δ U χ = d 3 x · χ ( x ) u ( x , t ) δ u ( x , t ) .
Thus, the force contribution from the tension term is:
F χ ( x , t ) = · χ ( x ) u ( x , t ) .
(3) Contribution from the Cubic Self-Interaction λ u 3 :
For the term:
U λ = d 3 x λ u 3 ,
the variation is:
δ U λ = d 3 x 3 λ u 2 δ u .
Thus, the force contribution is:
F λ ( x , t ) = 3 λ u ( x , t ) 2 .
(4) Contribution from the Yukawa-Like Coupling y u ψ ψ :
For the Yukawa term:
U y = d 3 x y u ψ ψ ,
the variation is:
δ U y = d 3 x y ψ ψ δ u .
Thus, the force contribution is:
F y ( x , t ) = y ψ ψ ( x , t ) .
(5) Total External Force:
Summing the individual contributions, the full expression for the external force is:
F ext ( x , t ) = Φ u · Φ ( u ) + · χ ( x ) u ( x , t ) 3 λ u ( x , t ) 2 y ψ ψ ( x , t ) .
This complete derivation illustrates how the external force F ext ( x , t ) is obtained via functional variation of the extended potential energy functional U ext [ u , ψ ] .

Appendix B.4. Discussion and Implications

The derived force function F ext ( x , t ) is incorporated into the refined STM wave equation to account for interactions beyond simple elasticity. Each term in F ext ( x , t ) represents a distinct physical contribution:
  • The local potential Φ captures spatially varying modifications to the elastic energy.
  • The tension term, with coefficient χ ( x ) , introduces a position-dependent adjustment to the membrane’s stiffness.
  • The cubic self-interaction term ( λ u 3 ) and the Yukawa-like coupling ( y u ψ ψ ) introduce essential nonlinearities that may be related to matter coupling and mass generation mechanisms.
Together, these interactions ensure that the membrane’s dynamics are modified in a way that is consistent with both classical elasticity and emergent quantum field phenomena.

Appendix C. Emergent Gauge Fields (U(1), SU(2) and SU(3))

Appendix C.1. Overview

The refined Space–Time Membrane (STM) model naturally gives rise to internal gauge symmetries through the elastic dynamics of the membrane. By performing a bimodal decomposition of the displacement field u ( x , t ) (as described in Appendix A), a two-component spinor Ψ ( x , t ) is obtained. The internal structure of Ψ ( x , t ) 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.

Appendix C.2. U(1) Gauge Symmetry

Local Phase Transformation and Covariant Derivative:
Consider the two-component spinor Ψ ( x , t ) derived from the bimodal decomposition. A local U(1) phase transformation is given by:
Ψ ( x , t ) Ψ ( x , t ) = e i θ ( x , t ) Ψ ( x , t ) ,
where θ ( x , t ) 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:
D μ Ψ ( x , t ) μ i e A μ ( x , t ) Ψ ( x , t ) ,
where A μ ( x , t ) 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:
F μ ν ( x , t ) = μ A ν ( x , t ) ν A μ ( x , t ) .
Under the gauge transformation,
A μ ( x , t ) A μ ( x , t ) = A μ ( x , t ) + 1 e μ θ ( x , t ) ,
the field strength tensor F μ ν ( x , t ) remains invariant.
Gauge Fixing and Ghost Fields:
For quantisation, it is necessary to fix the gauge. A common choice is the Lorentz gauge, μ A μ ( x , t ) = 0 . The Faddeev–Popov procedure is then employed to introduce ghost fields c ( x , t ) and c ( x , t ) that ensure proper treatment of gauge redundancy in the path integral formulation.

Appendix C.3. SU(2) Gauge Symmetry

Local SU(2) Transformation:
Assume that the spinor Ψ ( x , t ) exhibits a chiral structure such that its left-handed component, Ψ L ( x , t ) , transforms as a doublet under SU(2). A local SU(2) transformation is expressed as:
Ψ L ( x , t ) Ψ L ( x , t ) = U SU ( 2 ) ( x , t ) Ψ L ( x , t ) ,
where
U SU ( 2 ) ( x , t ) = e x p i θ a ( x , t ) σ a 2 ,
with σ a ( a = 1 , 2 , 3 ) being the Pauli matrices, and θ a ( x , t ) representing the local transformation parameters.
Covariant Derivative for SU(2):
To maintain invariance under this transformation, the covariant derivative is defined as:
D μ Ψ L ( x , t ) μ i g 2 A μ a ( x , t ) σ a 2 Ψ L ( x , t ) ,
where A μ a ( x , t ) are the SU(2) gauge fields and g 2 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:
F μ ν a ( x , t ) = μ A ν a ( x , t ) ν A μ a ( x , t ) g 2 ϵ a b c A μ b ( x , t ) A ν c ( x , t ) ,
where ϵ a b c are the antisymmetric structure constants of SU(2).
Gauge Fixing:
Imposing the Lorentz gauge, μ A μ a ( x , t ) = 0 , and applying the Faddeev–Popov procedure, ghost fields c a ( x , t ) and c a ( x , t ) are introduced with a ghost Lagrangian of the form:
L ghost SU ( 2 ) = c a μ μ δ a b + g 2 ϵ a b c A μ c ( x , t ) c b .

Appendix C.12.1. Electroweak Mixing, the Z Boson, and CP Violation via Zitterbewegung

In the refined STM framework, electroweak symmetry breaking and the emergence of the neutral Z boson can be naturally explained through interactions between the bimodal spinor field Ψ ( x , t ) residing on one face of the membrane and the corresponding bimodal antispinor field Ψ ( x , t ) located on the opposite face (the "mirror universe").
Specifically, the displacement field u ( x , t ) couples these spinor fields through an interaction Lagrangian of the form:
L int = i , j y i j u ( x , t ) Ψ ¯ i ( x , t ) e i θ i j ( x , t ) Ψ j ( x , t ) ,
where:
  • y i j represents Yukawa-like coupling constants between generations i , j .
  • u ( x , t ) is the membrane displacement field, whose vacuum expectation value (VEV), v = u ( x , t ) , generates effective fermion masses.
  • Complex phase shifts θ i j ( x , t ) arise naturally due to rapid oscillatory interactions—known as zitterbewegung—between the spinor Ψ and the mirror antispinor Ψ .
When the displacement field u ( x , t ) acquires a vacuum expectation value (VEV), denoted v = u ( x , t ) , this interaction yields an effective fermion mass matrix of the form:
M f i j = y i j v e i θ i j ,
where the phases θ i j become averaged into constant effective phases θ i j upon coarse-graining.

Appendix C.12.2. Electroweak Mixing and Emergence of the Z Boson

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 B μ (U(1)) and W μ a (SU(2)). Through the process described above—where the membrane’s displacement field acquires a vacuum expectation value v = u ( x , t ) —mass terms arise for specific gauge bosons. Explicitly, electroweak mixing occurs via a linear combination of the neutral gauge fields W μ 3 (from SU(2)) and B μ (from U(1)):
Z μ = c o s θ W W μ 3 s i n θ W B μ , A μ = s i n θ W W μ 3 + c o s θ W B μ ,
where θ W is the Weinberg angle, dynamically determined by membrane parameters, and B μ is the original U(1) gauge field. The gauge boson corresponding to the Z μ 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.

Appendix C.12.3. Emergence of CP Violation

Under a combined charge conjugation–parity (CP) transformation, the spinor fields transform approximately as:
Ψ ( x , t ) C P γ 0 C Ψ ¯ T ( x , t ) ,
with analogous transformations applied to the mirror antispinor Ψ . Due to the presence of nontrivial phases induced by the zitterbewegung interaction between spinor and antispinor fields, the effective fermion mass matrix
M f i j = y i j v e i θ i j ,
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 refined STM framework.

Appendix C.12.4. 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.
Although the underlying framework clearly illustrates how CP violation emerges deterministically, a rigorous derivation of chiral anomalies, weak parity violation, and related effects, such as neutrino mass generation via a see-saw mechanism, would require further detailed analysis, including explicit consideration of triangular loop diagrams within the STM framework.

Appendix C.4. SU(3) Gauge Symmetry

Local SU(3) Transformation:
For the strong interaction, the spinor Ψ ( x , t ) is assumed to carry a colour index and transform as a triplet under SU(3). A local SU(3) transformation is given by:
Ψ ( x , t ) Ψ ( x , t ) = U SU ( 3 ) ( x , t ) Ψ ( x , t ) ,
with
U SU ( 3 ) ( x , t ) = e x p i θ a ( x , t ) λ a 2 ,
where λ a ( a = 1 , , 8 ) are the Gell–Mann matrices, and θ a ( x , t ) are the transformation parameters.
Covariant Derivative for SU(3):
The covariant derivative is defined as:
D μ Ψ ( x , t ) μ i g 3 G μ a ( x , t ) λ a 2 Ψ ( x , t ) ,
where G μ a ( x , t ) are the SU(3) gauge fields and g 3 is the SU(3) coupling constant.
Field Strength Tensor for SU(3):
The SU(3) field strength tensor is defined by:
G μ ν a ( x , t ) = μ G ν a ( x , t ) ν G μ a ( x , t ) g 3 f a b c G μ b ( x , t ) G ν c ( x , t ) ,
where f a b c are the structure constants of SU(3).
Gauge Fixing:
The Lorentz gauge μ G μ a ( x , t ) = 0 is imposed, and ghost fields c a ( x , t ) and c a ( x , t ) are introduced via the Faddeev–Popov procedure. The ghost Lagrangian is then:
L ghost SU ( 3 ) = c a μ μ δ a b + g 3 f a b c G μ c ( x , t ) c b .
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 Ψ ( x , t ) carry a local SU(3) index and that neighbouring “sites” (or regions) remain elastically coupled under deformations. In essence, each SU(3) gauge connection G μ a ( x , t ) plays the role of an “elastic link” constraining colour charges, which becomes increasingly stiff (i.e. confining) with separation.
Mathematically, the field strength G μ ν a 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.

Appendix C.5. Prototype Emergent Gauge Lagrangian

While we have described how local phase invariance of our bimodal spinor Ψ induces gauge fields A μ a , we can also hypothesise a Yang–Mills-like action arising at low energies (See Figure 3):
L g a u g e = 1 4 F μ ν a F μ ν a + ( gauge fixing + ghost terms )
where F μ ν a = μ A ν a ν A μ a g f a b c A μ b A ν c
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 Ψ become nontrivial.

Appendix C.6. Summary

In summary, the internal structure of the two-component spinor Ψ ( x , t ) (derived from the bimodal decomposition of u ( x , t ) ) leads naturally to local gauge invariance. Enforcing invariance under local U(1) transformations necessitates the introduction of a U(1) gauge field A μ ( x , t ) with covariant derivative D μ = μ i e A μ ( x , t ) and field strength F μ ν . Extending this to non-Abelian symmetries, local SU(2) and SU(3) transformations require the introduction of gauge fields A μ a ( x , t ) and G μ a ( x , t ) , 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 and Interference

Appendix D.1. Overview

In the refined Space–Time Membrane (STM) model, quantum-like effects emerge from the classical elasticity of a four-dimensional membrane when short-wavelength (sub-Planck) oscillations are properly coarse-grained. This appendix outlines how one obtains an effective Schrödinger-like equation for the slowly varying envelope of the membrane’s displacement field and demonstrates that such an envelope supports interference patterns akin to those in standard quantum mechanics.

Appendix D.2. Starting Point: The Classical Equation of Motion

The classical dynamics of the membrane in the refined STM model are governed by a high-order partial differential equation that extends beyond the standard second-order wave equation by incorporating fourth- and sixth-order spatial derivatives. In our refined framework, the displacement field u x , t satisfies
ρ 2 u t 2 E S T M ( μ ) + Δ E ( x , t ; μ ) 4 u + η 6 u γ u t V ( u ) = 0 ,
where:
  • ρ is the membrane’s effective mass density,
  • E S T M ( μ ) is the scale-dependent baseline elastic modulus and Δ E ( x , t ; μ ) captures local stiffness variations,
  • η 6 u is a sixth-order spatial derivative term that provides crucial ultraviolet regularisation by strongly suppressing high-wavenumber fluctuations,
  • γ t u represents damping, and
  • V ( u ) is the derivative of the potential energy, accounting for nonlinear self-interactions.
To derive this equation from a variational principle, we start from the following Lagrangian density, which fully incorporates the effects of higher-order derivatives:
L = 1 2 ρ t u 2 1 2 E S T M ( μ ) + Δ E ( x , t ; μ ) 2 u 2 1 2 η 3 u 2 V ( u ) .
Here, the term 1 2 E S T M ( μ ) + Δ E ( x , t ; μ ) 2 u 2 yields, upon variation, the fourth-order term [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] 4 u , while the term 1 2 η 3 u 2 produces the sixth-order term η 6 u after performing integration by parts (assuming suitable boundary conditions). The potential term V ( u ) accounts for any nonlinear self-interaction, such as λ u 3 or higher-order contributions.
Our goal is to coarse-grain this high-order PDE so that the rapid, sub-Planck oscillations are averaged out, isolating a slowly varying envelope function Ψ ( x , t ) . This envelope will then serve as the basis for deriving an effective Schrödinger-like equation that reproduces interference phenomena and the Born rule, all emerging from a fundamentally deterministic and elastic substrate.

Appendix D.3. Coarse-Graining: Isolating the Slowly Varying Envelope

The field u ( x , t ) contains rapid sub-Planck fluctuations superimposed on a slowly varying macroscopic component. To isolate this component, we define the coarse-grained wavefunction Ψ ( x , t ) via a convolution with a Gaussian smoothing kernel:
Ψ ( x , t ) = d 3 y G ( x y ; L ) u ( y , t ) ,
with
G ( x y ; L ) = 1 2 π L 2 3 / 2 exp x y 2 2 L 2 ,
where L is the characteristic smoothing length. This procedure filters out high-frequency oscillations and leaves a slowly varying envelope Ψ ( x , t ) that represents the macroscopic dynamics.

Appendix D.4. Application of the WKB-Like Ansatz

To further analyse the dynamics of Ψ ( x , t ) , we adopt a WKB-like ansatz:
Ψ ( x , t ) = A ( x , t ) exp i S ( x , t ) ,
where:
  • A ( x , t ) is the slowly varying amplitude,
  • S ( x , t ) is the slowly varying phase,
  • is the reduced Planck constant (retained as a parameter in the effective description).
This ansatz is motivated by the observation that in many semiclassical approximations, the wavefunction factorises into a rapidly oscillatory exponential and a slowly varying amplitude.

Appendix D.5. Substitution into the Effective Equation

The next step is to substitute the WKB ansatz into the coarse-grained version of the classical equation of motion. Although the full derivation involves substituting into an effective equation that results from the coarse-graining procedure, we highlight the key steps:
  • Time Derivative:
t Ψ ( x , t ) = t A ( x , t ) + i A ( x , t ) t S ( x , t ) exp i S ( x , t ) .
  • Spatial Derivatives: The Laplacian of Ψ ( x , t ) is computed as:
2 Ψ ( x , t ) = 2 A ( x , t ) + 2 i A ( x , t ) · S ( x , t ) 1 2 A ( x , t ) S ( x , t ) 2 + i A ( x , t ) 2 S ( x , t ) exp i S ( x , t ) .
  • Higher-order derivatives (e.g., 4 Ψ ) are computed similarly. Under the assumption that A ( x , t ) and S ( x , t ) vary slowly compared to the rapid oscillations filtered by the smoothing kernel, terms involving higher derivatives of A ( x , t ) can be neglected to leading order.
  • Separation into Real and Imaginary Parts: After substituting the ansatz into the effective equation (obtained after coarse-graining and including higher-order corrections), the resulting expression is separated into its real and imaginary components:
    • The real part yields a Hamilton–Jacobi equation:
S t + 1 2 m eff ( S ) 2 + V eff ( x ) + Q ( x , t ) = 0 ,
  • where m eff is an effective mass parameter, V eff ( x ) is an effective potential (related to V ( u ) ), and Q ( x , t ) is the so-called quantum potential (containing higher-order corrections).
  • The imaginary part yields a continuity equation:
A 2 t + · A 2 m eff S = 0 .

Appendix D.6. Recovery of the Effective Schrödinger Equation

Neglecting the quantum potential Q ( x , t ) (which is justified in the semiclassical or weakly quantum regime), the Hamilton–Jacobi and continuity equations combine to yield the effective Schrödinger-like equation:
i t Ψ ( x , t ) = 2 2 m eff 2 + V eff ( x ) Ψ ( x , t ) .
This factorised form demonstrates that the coarse-grained dynamics of the membrane, as captured by Ψ ( x , t ) , obey the same formal structure as a non-relativistic quantum particle. The parameters m eff and V eff ( x ) are determined by matching the dispersion relation of the membrane—modified by the higher-order term η k 4 —to that of a free particle.

Appendix D.7. Interference and the Double-Slit Analogy

The effective Schrödinger-like equation supports the superposition principle, leading naturally to interference phenomena. For example, in a double-slit experiment, boundary conditions constrain Ψ ( x , t ) to be nonzero only in two narrow regions (the slits). The overall wavefunction can then be written as:
Ψ ( x , t ) = Ψ 1 ( x , t ) + Ψ 2 ( x , t ) ,
where Ψ 1 ( x , t ) and Ψ 2 ( x , t ) are the contributions from each slit. The intensity observed on a distant screen is:
I ( x ) = Ψ ( x , t ) 2 = Ψ 1 ( x , t ) + Ψ 2 ( x , t ) 2 .
The cross terms in this expression give rise to an interference pattern that closely resembles that seen in quantum double-slit experiments, despite the underlying dynamics being entirely deterministic.

Appendix D.8. On the Choice of Smoothing Kernel and WKB Limitations

In the above derivation (Sections D.2–D.5), we applied a Gaussian smoothing kernel G ( x y ; L ) to filter out sub-Planck-scale oscillations. We choose a Gaussian primarily because:
  • It is analytically tractable and minimises variance in wave mechanics contexts, linking naturally with standard semiclassical approximations.
  • It preserves locality in the sense that smoothing remains concentrated within L of a given point x, making it well-suited to short-wavelength filtering.
We note, however, that any physically motivated filter—such as wavelet transforms, top-hat functions, or exponential smoothing—could in principle produce a similar effective Schrödinger-like envelope, provided it separates short- and long-wavelength modes. The precise choice of kernel may shift subleading corrections and boundary effects, but it does not alter the core emergence of interference and the Born-rule-like amplitude interpretation.
The WKB approximation we used (Section D.4) also assumes relatively slow variation of the coarse-grained amplitude A ( x , t ) and phase S ( x , t ) . Strictly speaking, if the sub-Planck region is highly chaotic or features abrupt temporal changes, additional higher-order terms can become significant. A more systematic expansion would involve:
  • Retaining next-order derivatives in the amplitude and phase,
  • Investigating how chaotic substructures affect global interference.
This remains an open area for future work, and we expect that at sufficiently large scales—where wave amplitudes vary more gently—the WKB approach remains a good leading-order approximation (See Figure 1).

Appendix D.9. Summary

  • Coarse-Graining: A Gaussian smoothing kernel is applied to u ( x , t ) to extract a slowly varying envelope Ψ ( x , t ) .
  • WKB-Like Ansatz: The ansatz Ψ ( x , t ) = A ( x , t ) exp i S ( x , t ) facilitates the separation of the effective dynamics into a Hamilton–Jacobi equation for the phase and a continuity equation for the amplitude.
  • Effective Schrödinger Equation: Neglecting the quantum potential Q ( x , t ) , the combination of the separated equations yields:
i t Ψ ( x , t ) = 2 2 m eff 2 + V eff ( x ) Ψ ( x , t ) ,
  • with effective parameters m eff and V eff ( x ) determined by the membrane’s elastic properties.
  • Interference: The superposition principle inherent in this equation results in interference patterns, as illustrated by the double-slit analogy.
Hence, the refined STM approach demonstrates how standard quantum interference and the Born rule can surface in a deterministic continuum, eliminating the need for postulated intrinsic randomness. This sets the stage for further discussions on deterministic decoherence (Appendix G) and entanglement (Appendix E), in which classical wave mechanics reproduces yet more “quantum-exclusive” phenomena under coarse-graining.

Appendix E. Deterministic Quantum Entanglement and Bell Inequality Analysis

Appendix E.1. Overview

In the refined Space–Time Membrane (STM) model, the full deterministic dynamics yield an effective wavefunction that—after coarse-graining—exhibits non-factorisable correlations. These correlations mimic quantum entanglement, even though the underlying evolution is entirely deterministic. In this appendix, we provide a detailed derivation of the emergence of these entangled states and show, through the introduction of appropriate measurement operators, that the model can lead to violations of Bell inequalities analogous to those observed in quantum mechanics.

Appendix E.2. Derivation of Non-Factorisable Global Modes

Consider two spatially localised excitations on the membrane, described by the displacement fields u A ( x , t ) and u B ( x , t ) . In the full deterministic description, the total displacement field is given by:
u tot ( x , t ) = u A ( x , t ) + u B ( x , t ) + V int ( x , t ) ,
where V int ( x , t ) represents the interaction between these excitations—arising from the inherent elasticity of the membrane. For example, one may model the interaction as
V int ( x , t ) = α u A ( x , t ) u B ( x , t ) ,
with α being a coupling constant.
After applying a Gaussian coarse-graining procedure (as described in Appendix D) to filter out rapid sub-Planck fluctuations, the effective wavefunction is given by:
Ψ u A , u B = Ψ u A ( x , t ) + u B ( x , t ) + V int ( x , t ) .
Because V int ( x , t ) is a nonlinear function of both u A and u B , the effective wavefunction generally cannot be factorised into a simple product,
Ψ u A , u B Ψ A u A Ψ B u B .
To illustrate, if we assume the interaction is given by α u A ( x , t ) u B ( x , t ) , then the argument of Ψ is
u A ( x , t ) + u B ( x , t ) + α u A ( x , t ) u B ( x , t ) ,
which, unless α = 0 or one of the fields is zero, is a nonseparable function of u A and u B . This non-factorisability implies that the composite state encodes correlations between regions A and B that cannot be described independently—mimicking the quantum phenomenon of entanglement.

Appendix E.3. 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 Ψ (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:
M ^ ( θ ) = c o s θ σ x + s i n θ σ z ,
where σ x and σ z are the Pauli matrices and θ is a measurement angle. For subsystems A and B, we denote the operators as M ^ A θ A and M ^ B θ B , respectively.
The joint correlation function for measurements performed at angles θ A and θ B is then given by:
E θ A , θ B = Ψ M ^ A θ A M ^ B θ B Ψ .
This expectation value is calculated by integrating over the coarse-grained degrees of freedom, taking into account the non-factorisable structure of Ψ u A , u B .

Appendix E.4. Detailed CHSH Parameter Calculation

The CHSH inequality involves four correlation functions corresponding to two measurement settings per subsystem. Define the CHSH parameter as:
S = E θ A , θ B E θ A , θ B + E θ A , θ B + E θ A , θ B .
A detailed derivation involves the following steps:
  • State Decomposition:
    Express Ψ 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:
Ψ = i c i a i b i ,
  • where c i are effective coefficients that encode the correlations.
  • Evaluation of E θ A , θ B :
    With the measurement operators defined as above, compute the joint expectation value:
E θ A , θ B = i , j c i c j * a i M ^ A θ A a j b i M ^ B θ B b j .
  • The explicit dependence on the measurement angles enters through the matrix elements of the Pauli matrices.
  • Optimisation:
    Choose measurement angles θ A , θ A , θ B , θ B to maximise S. Standard quantum mechanical analysis shows that the optimal settings are typically:
θ A = 0 , θ A = π 2 , θ B = π 4 , θ B = π 4 .
  • With these settings, the CHSH parameter can be shown to reach:
S = 2 2 .
  • 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.

Appendix E.5. Summary

  • The effective wavefunction Ψ u A , u B obtained from the deterministic dynamics is non-factorisable due to the coupling term V int ( x , t ) .
  • 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 2 2 , 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 refined STM framework.

Appendix F. Appendix F: Singularity Prevention in Black Holes

Appendix F.1. Overview

In the refined Space–Time Membrane (STM) model, the extreme gravitational collapse predicted by General Relativity is avoided through the elastic properties of the membrane. In this approach, the displacement field u x , t obeys a high-order wave equation that includes scale-dependent stiffness, higher-order spatial derivatives (up to 6 ), damping, nonlinear self-interactions, and couplings to emergent spinor fields. This appendix details how increasing local stiffness and the inclusion of the 6 term act together to regularise curvature, thereby preventing the formation of classical singularities. Instead, the membrane relaxes into finite-amplitude standing wave solutions—solitonic configurations that offer a potential resolution to the black hole information paradox.

Appendix F.2. Increasing Local Stiffness at High Densities

As matter collapses, the local energy density rises, and within the STM framework, the effective elastic modulus is given by
E x , t = E S T M ( μ ) + Δ E x , t ; μ .
Here, E S T M ( μ ) is the baseline modulus and Δ E x , t ; μ captures local stiffness variations induced by high-frequency, sub-Planck oscillations. In regions of extremely high density, Δ E x , t ; μ increases sharply. This rapid stiffening acts as a self-regulation mechanism, making further curvature growth energetically prohibitive and thereby preventing the development of infinite curvature that would otherwise signal a singularity.

Appendix F.3. Role of the ∇ 6 Term and Standing Wave Solutions

The refined STM equation includes a sixth-order spatial derivative term, η 6 u , which is crucial for ultraviolet regularisation. In configuration space, this term directly penalises short-wavelength deformations. In momentum space, the propagator for u x , t becomes
G ( k ) = 1 ρ c 2 k 2 + [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] k 4 + η k 6 + V ( u ) ,
so that at high momentum the k 6 contribution dominates. This strong suppression of high-frequency fluctuations ensures that loop integrals remain finite and the theory is well-behaved in the UV. 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.

Appendix F.4. Implications for Black Hole Information

Since the interior of a black hole in the STM model is described by these finite-energy, solitonic configurations rather than by a singularity, the information contained in infalling matter is not irretrievably lost. Instead, the phase and amplitude of the interior standing waves can encode and potentially preserve information. Moreover, the non-Markovian decoherence mechanism (see Appendix G) may allow for gradual information leakage, providing a promising route toward resolving the black hole information paradox within a deterministic, elasticity-based framework.

Appendix F.5. Summary

  • Local Stiffness Increase: As energy density rises, the effective modulus E x , t = E S T M ( μ ) + Δ E x , t ; μ increases sharply, impeding infinite curvature.
  • Higher-Order Regularisation: The inclusion of the 6 u term strongly suppresses high-frequency fluctuations, ensuring the convergence of loop integrals and UV stability.
  • Finite-Energy Standing Waves: Instead of a singularity, gravitational collapse yields stable, finite-amplitude solitonic configurations—soft cores that redistribute energy.
  • Information Preservation: These solitonic solutions provide a mechanism for retaining and potentially gradually releasing information, addressing the black hole information paradox.
Thus, the refined STM model replaces classical singularities with deterministically generated, finite-energy structures, offering new insights into black hole interiors and information conservation.

Appendix G. Appendix G: Non-Markovian Decoherence and Measurement

Appendix G.1. Overview

In the refined 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.

Appendix G.2. Decomposition of the Displacement Field

We begin by decomposing the full displacement field u ( x , t ) into two components:
u ( x , t ) = u S ( x , t ) + u E ( x , t ) ,
where:
  • u S ( x , t ) is the slowly varying, coarse-grained “system” field,
  • u E ( x , t ) comprises the high-frequency “environment” modes (the sub-Planck fluctuations).
The coarse-graining is achieved by convolving u ( x , t ) with a Gaussian kernel G ( x y ; L ) over a spatial scale L:
u S ( x , t ) = d 3 y G ( x y ; L ) u ( y , t ) ,
with
G ( x y ; L ) = 1 2 π L 2 3 / 2 exp x y 2 2 L 2 .
The environmental part is then defined as:
u E ( x , t ) = u ( x , t ) u S ( x , t ) .
This separation allows us to treat u S ( x , t ) as the primary degrees of freedom while regarding u E ( x , t ) as the effective environment.

Appendix 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 t f is given by:
ρ u S f , u E f ; u S f , u E f ; t f = D u S D u E exp i S [ u S , u E ] S [ u S , u E ] ρ u S i , u E i ; u S i , u E i ; t i .
To obtain the reduced density matrix ρ S u S f , u S f ; t f for the system alone, we integrate out the environmental degrees of freedom:
ρ S u S f , u S f ; t f = D u E exp i S [ u S , u E ] S [ u S , u E ] ρ E u E , u E ; t i .
We define the Feynman–Vernon influence functional F [ u S , u S ] as:
F [ u S , u S ] = D u E exp i S int u S , u E S int u S , u E ρ E u E , u E ; t i ,
where S int u S , u E denotes the interaction part of the action that couples the system to the environment.
For weak system–environment coupling, we can expand S int to second order in the difference Δ u S ( t ) = u S ( t ) u S ( t ) . This yields a quadratic form for the influence action:
S IF [ u S , u S ] t i t f d t t i t f d t Δ u S ( t ) K t t Δ u S t ,
where K t t is a memory kernel that encapsulates the temporal correlations of the environmental modes. The precise form of K t t depends on the spectral density of the environment and the specific details of the coupling.

Appendix G.4. Derivation of the Non-Markovian Master Equation

Starting from the reduced density matrix expressed with the influence functional:
ρ S u S f , u S f ; t f = D u S D u S exp i S [ u S ] S [ u S ] + S IF [ u S , u S ] ,
we differentiate ρ S with respect to time t f to obtain its evolution. Standard techniques (akin to those used in the Caldeira–Leggett model) yield a master equation of the form:
d ρ S ( t ) d t = i [ H S , ρ S ( t ) ] t i t d t K t t D [ ρ S t ] ,
where:
  • H S is the effective Hamiltonian governing the system u S ( x , t ) ,
  • D [ ρ S t ] is a dissipative superoperator that typically involves commutators and anticommutators with system operators (e.g., u S or its conjugate momentum),
  • The kernel K t t introduces memory effects; that is, the rate of change of ρ S ( t ) depends on its values at earlier times.
In the limit where the environmental correlation time is very short (i.e., K t t approximates a delta function δ t t ), 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.

Appendix G.5. Implications for Measurement

The non-Markovian master equation implies that when the system u S ( x , t ) interacts with a macroscopic measurement device, the off-diagonal elements of the reduced density matrix ρ S ( t ) decay over a finite time determined by K t t . 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.

Appendix G.6. Path from Influence Functional to a Non-Markovian Operator Form

We have described in Eqs. (G.3)–(G.5) how integrating out the high-frequency environment u E produces an influence functional F [ u S ] with a memory kernel K t t . In principle, if this kernel is short-ranged, one recovers a Markov limit akin to a Lindblad master equation,
d ρ S d t = i [ H S , ρ S ] + α L α ρ S L α 1 2 { L α L α , ρ S }
However, in our non-Markovian STM scenario, the memory kernel extends over times Δ t e n v . We therefore obtain an integral-differential form,
d ρ S ( t ) d t = i [ H S , ρ S ( t ) ] t 0 t d t K t t D [ ρ S t ]
capturing the environment’s finite correlation time (See Figure 4). Determining explicit Lindblad-like operators L α from this memory kernel would require further approximations (e.g., expansions in powers of Δ t e n v / T , 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.

Appendix G.7. Summary

  • Decomposition: The total field u ( x , t ) is decomposed into a slowly varying system component u S ( x , t ) and a high-frequency environment u E ( x , t ) .
  • Influence Functional: Integrating out u E ( x , t ) yields an influence functional characterised by a memory kernel K t t that captures the non-instantaneous response of the environment.
  • Master Equation: The resulting non-Markovian master equation for the reduced density matrix ρ S ( t ) involves an integral over past times, reflecting the system’s dependence on its history.
  • Measurement: The deterministic decay of off-diagonal elements in ρ S ( t ) explains the effective collapse of the wavefunction observed in quantum measurements.
Thus, the refined 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. Appendix H: Density-Driven Vacuum Energy Variations

Appendix H.1. Overview

In the refined Space–Time Membrane (STM) model, the vacuum energy—commonly associated with the cosmological constant—is interpreted as a time-averaged offset in the local stiffness of the membrane. Even in regions that are nominally “empty,” persistent low-amplitude sub-Planck oscillations generate a nonzero effective stiffness. This appendix details how these oscillations give rise to an effective vacuum energy and how local inhomogeneities naturally emerge from variations in the elastic properties of the membrane.

Appendix H.2. Time-Averaged Vacuum Energy

The local elastic modulus of the membrane is expressed as
E ( x ) = E S T M ( μ ) + Δ E ( x ) ,
where:
  • E S T M ( μ ) is the baseline, scale-dependent elastic modulus, and
  • Δ E ( x ) represents local variations induced by high-frequency sub-Planck oscillations.
Due to these persistent oscillations, even regions devoid of classical matter exhibit a nonzero effective stiffness. By averaging the local elastic modulus over a complete oscillation period T, we define the effective vacuum energy density as
ρ v a c = E ( x ) T = 1 T 0 T E ( x , t ) d t .
When the oscillations are spatially uniform on macroscopic scales, ρ v a c effectively behaves as a cosmological constant term in the energy–momentum tensor.

Appendix H.3. Incorporating Inhomogeneities

Cosmological observations indicate that vacuum energy may not be perfectly uniform. In earlier formulations, a secondary coupling constant ( β ) was introduced to modulate the effective stiffness based on the local persistent wave energy density ρ w a v e s ( x ) . In the refined STM model, however, the effects originally attributed to β are subsumed into the running of the elastic parameters through the renormalisation group flow. In this context, the effective stiffness modulation is naturally determined by the scale-dependent functions emerging from the FRG analysis, eliminating the need to introduce β as an independent parameter. Instead, any spatial inhomogeneities in ρ v a c arise from the inherent scale dependence of E S T M ( μ ) and Δ E ( x , t ; μ ) .

Appendix H.4. Simulation Parameters and Experimental Prospects

Numerical simulations using finite element analysis (see Appendix K) can be employed to study membrane dynamics incorporating the full range of higher-order terms and the scale-dependent effective stiffness. Typical parameters in such simulations include:
  • A smoothing length scale l for spatial averaging,
  • A baseline vacuum offset E 0 determined by the average sub-Planck oscillation amplitude, and
  • The renormalisation group–determined running of E S T M ( μ ) and Δ E ( x , t ; μ ) , which now capture the role previously attributed to β .
Experimental tests may involve metamaterial analogues, where local stiffness variations can be engineered and measured, or astrophysical observations (e.g. cosmic microwave background anisotropies) to detect signatures of non-uniform vacuum energy.

Appendix H.5. Summary

  • Time-Averaged Stiffness: Persistent sub-Planck oscillations yield a nonzero, time-averaged elastic modulus that manifests as an effective vacuum energy density ρ v a c .
  • Local Inhomogeneities: Rather than introducing an independent coupling β , the scale-dependent running of elastic parameters naturally modulates the local stiffness, leading to spatial variations in ρ v a c .
  • Testing Prospects: Finite element simulations and experimental or astrophysical observations provide avenues for probing these predictions.
Thus, the refined STM model reinterprets the cosmological constant as an emergent property of sub-Planck membrane dynamics, with the effective modulation of vacuum energy governed by renormalisation group–determined, scale-dependent elastic parameters.

Appendix I. Experimental Setups and Proposed Analogues

Appendix I.1. Overview and Objectives

This appendix proposes concrete strategies for testing the refined Space–Time Membrane (STM) model in both laboratory analogues and larger-scale gravitational or cosmological observations. Our objectives are twofold:
  • Feasibility Assessment
    Provide sufficient detail regarding wave speeds, stiffness ranges, and measurement strategies so experimentalists can evaluate whether a tabletop or metamaterial setup is achievable with current technology.
  • Identification of Distinctive STM Signatures
    Outline how analogue systems can exhibit key STM features—such as higher-order derivative effects, dynamic stiffness feedback, deterministic decoherence, and Einstein-like gravity corrections—in ways that differ from simpler classical or quantum models.
The discussion covers:
  • Acoustic membrane analogues
  • Optical metamaterial analogues
  • Advanced interference and decoherence tests
  • Laboratory and astrophysical observations relevant to the model’s Einstein-like field equations and scale-dependent elasticity.

Appendix I.2. Acoustic Membrane Analogues

Appendix I.48.1. Rationale and Mapping to STM

A thin, tensioned membrane can serve as a physical analogue for certain aspects of the STM model. By introducing local or global modifications to the membrane’s stiffness, experimenters can mimic the nonlinear terms and higher-order derivatives ( 4 and 6 ) present in the refined STM wave equation. For example:
  • Local Stiffness Variation:
    Patches of piezoelectric material or regions subjected to controlled temperature gradients can adjust the effective modulus, emulating Δ E ( x , t ; μ ) .
  • Higher-Order Dispersion:
    Additional constraints or layered structures can replicate 4 and 6 operators, altering wave dispersion and potentially producing soliton-like modes.

Appendix I.48.2. Suggested Parameter Ranges

  • Membrane Composition:
    Thin sheets of Mylar, latex, or metal foil in the 0.01–0.1 mm thickness range.
  • Tension:
    Tens to hundreds of newtons per metre, yielding wave speeds of 50–300 m/s.
  • Driving Frequency:
    Kilohertz frequencies are ideal to probe the higher-order dispersion regime above standard linear modes.

Appendix I.48.3. Experimental Setup and Measurements

  • Excitation:
    Use piezoelectric transducers or electromechanical shakers at the membrane’s boundary, or a point/ring driver for standing waves.
  • Stiffness Modulation:
    Incorporate voltage-controlled patches to tune local stiffness, mirroring the feedback mechanism Δ E ( x , t ; μ ) .
  • Detection:
    Employ laser Doppler vibrometry or high-speed cameras with markers to measure 2D wave amplitudes and phases. Look for stable interference nodes and nonlinear hysteresis patterns indicative of deterministic decoherence.

Appendix I.3. Optical Metamaterial Analogues

Appendix I.49.1. Conceptual Basis

Nonlinear optical media provide an alternative route to emulate STM physics. Here, the refractive index n depends on local light intensity I, paralleling how the STM model’s “local energy” modifies stiffness. Photonic structures with tailored dispersion properties can reproduce effective 2 , 4 , and 6 operators.

Appendix I.49.2. Typical Parameter Regimes

  • Waveguides or Photonic Crystals:
    Typically a few millimetres to centimetres in length.
  • Nonlinear Coefficients:
    Materials with χ ( 2 ) or χ ( 3 ) nonlinearities can see refractive index shifts of 10 5 to 10 4 under moderate laser power.
  • Field Profiles:
    Under a paraxial approximation, beam propagation can be governed by an effective wave equation, where higher-order dispersion mimics a 6 term.

Appendix I.49.3. Measurement Methods

  • Interferometric Detection:
    Split the input beam into reference and signal arms; recombine to measure phase shifts/fringe visibilities. Vary power to observe changes in the nonlinear index.
  • Beam Shaping:
    Multi-slit apertures can reveal interference patterns that stabilise or evolve distinctly from standard Kerr effects.
  • Potential STM-Like Effects:
    Look for wave–anti-wave locking or deterministic collapse into stable amplitude distributions, reminiscent of the “coherent wave–anti-wave cycles” in STM.

Appendix I.4. Advanced Interference and Decoherence Tests

Appendix I.50.1. Double-Slit and Multi-Slit Scenarios

  • Slit Geometry:
    Two or more narrow slits form boundary conditions isolating partial wavefronts.
  • Feedback Region:
    Introduce an amplitude-dependent stiffness (or refractive index) region to stabilise or modify fringes.
  • Measurement:
    Track fringe contrast under varying input amplitude, damping, or deliberate noise. Deterministic decoherence would manifest as stable (rather than randomly smeared) fringe visibility changes.

Appendix I.50.2. Entanglement Analogues and Bell-Like Correlations

  • Coupled Systems:
    Design tensioned membranes or dual waveguide arms with shared boundary conditions enforcing nonseparable wave modes.
  • Spinor-Like Observables:
    In optical setups, exploit polarisation states (pseudo-spin- 1 2 ). Set polariser angles θ A and θ B in distinct arms to test classical vs. “Bell-type” bounds.
  • Classical Nonlocality:
    Show that measured correlations exceed classical limits, even though the underlying PDE is deterministic.

Appendix I.5. Practical Implementation Steps

  • Finite Element Simulations:
    As in Appendix K, map parameter regimes (tension, wave amplitude, nonlinear index) that yield pronounced STM-specific effects beyond standard wave theory.
  • Prototyping:
    Start with small-scale prototypes (10–30 cm acoustic membranes or 1–2 cm optical waveguides) to assess boundary conditions, damping, and interference signatures.
  • Parameter Exploration:
    Systematically vary amplitude, stiffness, and damping; compare results to STM predictions.
  • Benchmarking:
    Construct simpler linear or Kerr-type PDE models for the same geometry. Consistent deviations from these benchmarks can pinpoint unique STM signatures.

Appendix I.6. Longer-Term Astrophysical Observations

The refined STM model also predicts potential signatures on cosmic scales:
  • Black Hole Ringdowns:
    Solitonic or extra stiffness near black hole horizons could alter quasi-normal mode frequencies. Current detectors (LIGO, Virgo) may lack the precision to detect small shifts, but future observatories (Einstein Telescope, Cosmic Explorer) might detect subtle deviations.
  • Vacuum Energy Inhomogeneities:
    Scale-dependent stiffening could yield slight spatial variations in effective vacuum energy, leaving imprints in cosmic microwave background anisotropies or galaxy cluster surveys. Disentangling these from Λ CDM backgrounds will require high-precision data.

Appendix I.7. Additional Gravitational and Cosmological Tests

Recent discussions of Einstein-like equations in the main text (Appendix M) suggest several further experimental/observational avenues beyond metamaterial and interference analogues:
  • Short-Range Gravity Measurements
    • Motivation: If the STM’s scale-dependent elasticity modifies the local gravitational potential for distances under a millimetre, precision torsion-balance or microcantilever experiments could detect deviations from the inverse-square law.
    • Implementation: Adapt existing short-range gravity setups (e.g. Eöt-Wash experiments). If small stiffening terms appear in the effective action, there may be a measurable Yukawa-like correction.
  • Local Time-Dilation Anomalies
    • Motivation: In the STM approach, membrane strain maps onto metric components. If the scale-dependence leads to extra terms in g 00 , advanced atomic clock experiments at different gravitational potentials could reveal minute deviations from standard GR.
    • Implementation: Compare clock frequencies across varying altitudes or in local gravitational wells. Any persistent discrepancy, once systematic effects are accounted for, could point to an STM-specific stiffening effect.
  • Black Hole Thermodynamic Tests
    • Motivation: STM’s solitonic interior structures might alter black hole entropy and near-horizon thermodynamics.
    • Implementation: While direct BH entropy measurements are challenging, refined gravitational wave ringdown data or horizon imaging (Event Horizon Telescope) could reveal if horizon geometry differs from standard predictions (e.g. no classical singularity).
  • Cosmological Data Fitting
    • Motivation: The running gravitational coupling G ( μ ) introduced by membrane stiffness could shift expansion rates or dark energy behaviour, potentially explaining certain discrepancies (like the Hubble tension).
    • Implementation: Incorporate the STM’s scale-dependent elasticity into a Friedmann–Lemaître–Robertson–Walker (FLRW) background. Compare with supernova data, CMB anisotropies, and baryon acoustic oscillations. If the model improves fits relative to Λ CDM, it supports the STM approach.
These gravitational/cosmological tests serve as important complements to the laboratory wave analogues, potentially allowing the refined STM model to be confronted with high-precision data across a wide range of scales.

Appendix I.8. Conclusions and Recommendations

  • Near-Term Feasibility
    Acoustic membranes, optical metamaterials, and interference experiments offer the most accessible means to test STM predictions in the lab. By carefully tuning tension or refractive index feedback, experimenters can replicate key features such as wave–anti-wave cycles, deterministic decoherence, and solitonic modes.
  • Methodical Exploration
    Systematic finite element simulations (Appendix K) combined with targeted lab prototypes are crucial for exploring parameter space (amplitude, damping, boundary conditions) and identifying distinctive STM effects (e.g., stable fringe nodes or classical Bell-like correlations).
  • Further Research
    • Local Gravity Tests: Torsion balances, microcantilevers, or atomic clock arrays could probe short-distance gravitational anomalies or time-dilation shifts predicted by scale-dependent elasticity.
    • Cosmological and Black Hole Observations: High-precision gravitational wave data, cosmic microwave background measurements, and black hole horizon imaging can test the large-scale, Einstein-like aspects of the refined STM model.
    • Null Results: Even if no deviations are found, such constraints refine STM parameter ranges or rule out certain stiffening functions Δ E ( x , t ; μ ) .
In summary, tabletop or intermediate-scale analogues remain the most promising near-term step, providing a controlled environment for investigating wave–anti-wave excitations, deterministic decoherence, and higher-order dispersion. Simultaneously, short-range gravitational experiments and cosmological data comparisons could reveal or constrain the scale-dependent elasticity predicted by STM’s Einstein-like equations, thereby bridging quantum-like membrane dynamics with macroscopic gravitational phenomena.

Appendix J. Renormalisation Group Analysis and Scale-Dependent Couplings

Appendix J.1. Overview

In the refined Space–Time Membrane (STM) model, the Lagrangian includes higher-order derivative terms—specifically, the 4 and 6 operators—as well as scale-dependent elastic parameters. 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 d = 4 ε 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.

Appendix J.2. One-Loop Renormalisation

Appendix J.56.1. Setting Up the One-Loop Integral

Consider the cubic self-interaction term, λ u 3 , 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 Σ ( 1 ) ( k ) is expressed as
Σ ( 1 ) ( k ) λ 2 d d p ( 2 π ) d 1 D ( p ) ,
where the propagator denominator is given by
D ( p ) = ρ c 2 p 2 + E S T M ( μ ) + Δ E ( x , t ; μ ) p 4 + η p 6 + .
At high momentum, the η p 6 term dominates, so the integral behaves roughly as
d d p ( 2 π ) d 1 p 6 .
For the simplified case in which the 6 term moderates the divergence, one typically encounters a pole in 1 / ε after dimensional regularisation.

Appendix J.56.56.1. J.2.2 Evaluating the Integral

Using standard results,
d d p ( 2 π ) d 1 p 2 2 = 1 ( 4 π ) d / 2 Γ ( 2 d / 2 ) Γ ( 2 ) ,
and substituting d = 4 ε , one finds
Γ 2 4 ε 2 = Γ ε 2 2 ε γ ,
with γ the Euler–Mascheroni constant. Hence, the one-loop self-energy contains a divergence of the form
Σ ( 1 ) ( k ) λ 2 ( 4 π ) 2 1 ε + finite terms .

Appendix J.56.56.2. J.2.3 Extracting the Beta Function

Defining the renormalised effective elastic parameter E e f f ( μ ) through
E e f f b a r e = E e f f ( μ ) + Σ ( 1 ) ( k ) ,
and requiring that the bare parameter is independent of the renormalisation scale μ (i.e. μ μ E e f f b a r e = 0 ), one differentiates to obtain the one-loop beta function for the effective coupling g e f f (which parameterises E e f f ):
β ( 1 ) g e f f = μ g e f f μ = a g e f f 2 ,
where a is a constant proportional to λ 2 / ( 4 π ) 2 .

Appendix J.56.2. 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.

Appendix J.56.56.1. 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:
Σ s u n ( 2 ) ( k ) λ 4 d d p ( 2 π ) d d d q ( 2 π ) d 1 D ( p ) D ( q ) D ( k p q ) ,
with D ( p ) as defined above. To combine the denominators, one introduces Feynman parameters:
1 A B C = 2 0 1 d x 0 1 x d y 1 [ x A + y B + ( 1 x y ) C ] 3 .
After performing the momentum integrations, overlapping divergences manifest as double poles in 1 / ε 2 and single poles in 1 / ε .

Appendix J.56.56.2. J.3.2 Mixed Fermion–Scalar Diagrams

If the Yukawa coupling y (coupling u to ψ ) is included, diagrams involving fermion loops inserted in scalar bubbles contribute additional terms. Such diagrams yield divergences proportional to y 2 λ 2 after performing the trace over gamma matrices and momentum integrations.

Appendix J.56.56.3. J.3.3 Two-Loop Beta Function

Collecting all two-loop contributions, the renormalisation constant Z g e f f for the effective coupling is expanded as:
Z g e f f = 1 + b g e f f ε + c g e f f 2 ε 2 + d g e f f 2 ε + ,
yielding the two-loop beta function:
β g e f f = a g e f f 2 + b g e f f 3 + ,
with the coefficient b incorporating both single and double pole contributions.

Appendix J.3. 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 g e f f 4 . Schematically, the three-loop self-energy takes the form:
Σ ( 3 ) ( k ) g e f f 4 1 ε 3 + 1 ε 2 + 1 ε .
Defining the bare coupling as
g e f f B = μ ε g e f f ( μ ) + δ g e f f ,
and enforcing μ -independence leads to the full beta function:
β g e f f = a g e f f 2 + b g e f f 3 + c g e f f 4 + .
The existence of nontrivial fixed points, g e f f * where β g e f f * = 0 , 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 g e f f 3 term could imply asymptotic freedom.

Appendix J.4. Illustrative One-Loop Example

As a concrete example, consider a bubble diagram in the scalar sector with a cubic self-interaction term λ u 3 . The one-loop self-energy is given by:
Σ ( 1 ) ( k ) = λ 2 d d p ( 2 π ) d 1 ρ c 2 p 2 + η p 4 + m 2 ,
where m 2 may arise from the second derivative of V ( u ) . In dimensional regularisation (with d = 4 ε ), one isolates the divergence via
I = d d p ( 2 π ) d 1 p 2 2 1 16 π 2 2 ε γ + ,
where γ is the Euler–Mascheroni constant. This divergence determines the running of λ and leads to a one-loop beta function of the form:
β ( 1 ) ( λ ) a λ 2 .
Higher-loop contributions then add corrections of order λ 3 and beyond (See Figure 2).

Appendix J.5. Summary and Implications

  • One-Loop Corrections:
    Yield a divergence Σ ( 1 ) ( k ) λ 2 / ( 4 π ) 2 1 / ε , leading to β ( 1 ) g e f f = a g e f f 2 .
  • Two-Loop Corrections:
    The setting sun and mixed fermion–scalar diagrams contribute additional overlapping divergences, resulting in a beta function β g e f f = a g e f f 2 + b g e f f 3 .
  • Three-Loop Corrections:
    Further diagrams introduce terms c g e f f 4 , refining the beta function to β g e f f = a g e f f 2 + b g e f f 3 + c g e f f 4 + .
  • Fixed Point Structure:
    Nontrivial fixed points g e f f * (satisfying β g e f f * = 0 ) 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 refined 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 Analysis for Determining STM Coupling Constants

Appendix K.1. Overview

Finite Element Analysis (FEA) offers a robust numerical method for solving the refined Space–Time Membrane (STM) partial differential equation that incorporates higher-order derivatives (up to 6 u ), scale-dependent elastic parameters, damping, and nonlinear interaction terms. This appendix details the procedures for spatial discretisation, time integration, and handling nonlinearities, as well as strategies for parameter fitting. By comparing simulation outputs with experimental or observational data, one can extract the coupling constants (e.g. λ , y, E S T M ( μ ) , and η ) that define the refined STM model.

Appendix K.2. Spatial Discretisation and Mesh Construction

Appendix K.61.1. Domain Definition

Begin by defining a spatial domain Ω that accurately represents the physical system of interest. For example, a domain may be chosen to mimic a double-slit geometry in interference studies or a radially symmetric region for black hole analogue simulations. The domain must be sufficiently large to capture both local variations and the overall behaviour of the displacement field u x , t .

Appendix K.61.2. Mesh Generation

Partition Ω into finite elements E i (e.g. tetrahedral or hexahedral elements in three dimensions). Standard meshing techniques may be used, but special attention is required near regions expected to exhibit rapid changes—such as near defects or boundaries—where adaptive meshing can provide increased resolution.

Appendix K.61.3. Choice of Shape Functions

Within each element, approximate u x , t by a linear combination of smooth shape (basis) functions N i x :
u x , t i = 1 N u i ( t ) N i x .
To capture higher-order derivatives up to 6 u , the shape functions must possess high continuity (typically polynomial degree p 3 or employing spectral element methods), ensuring that both u and its derivatives are continuous across element boundaries.

Appendix K.61.4. Discretisation of Differential Operators

Apply the differential operators (e.g. 2 and 6 ) to the shape function expansion to derive a system of algebraic equations. Accurate computation of the resulting stiffness matrices is crucial and often requires high-order numerical integration schemes to handle the increased polynomial degree.

Appendix K.3. Time Integration and Treatment of Nonlinearities

Appendix K.62.1. Time Discretisation

Discretise the time domain into steps t 0 , t 1 , , t N with a step size Δ t . Given the second-order time derivative in the STM equation, an implicit time integration scheme—such as the Crank–Nicolson method—is generally preferred for its unconditional stability in stiff systems. For example, the second-order time derivative can be approximated by:
u n + 1 2 u n + u n 1 ( Δ t ) 2 2 u t 2 t = t n .

Appendix K.62.2. Treatment of Nonlinear Terms

The refined STM equation includes nonlinearities such as the cubic term λ u 3 and Yukawa-like coupling y u ψ ¯ ψ . These are incorporated into the weak form of the discretised equations. At each time step, solve the resulting nonlinear system iteratively using methods like Newton–Raphson:
u ( k + 1 ) = u ( k ) J u ( k ) 1 R u ( k ) ,
where u ( k ) is the approximation at iteration k, R u ( k ) is the residual vector, and J u ( k ) is the Jacobian matrix. Convergence is achieved when the residual norm falls below a predetermined tolerance.

Appendix K.4. Parameter Fitting and Cost Function Minimisation

Appendix K.63.1. Simulation Outputs

The FEA simulations yield numerical predictions for observable quantities, such as:
  • Interference fringe patterns (from double-slit analogue experiments),
  • Gravitational wave ringdown frequencies (in black hole analogue simulations),
  • Other observables linked to the membrane’s elastic dynamics.

Appendix K.63.2. Cost Function

To calibrate the STM model parameters, define a cost function J that quantifies the discrepancy between simulation outputs S i and experimental or observational data D i :
J = i S i λ , y , E ( x ) , η , D i 2 .
This function is minimised with respect to the parameters.

Appendix K.63.3. Optimisation Procedure

Optimisation methods such as gradient descent, Levenberg–Marquardt, or genetic algorithms can be employed to adjust the model parameters until the simulated outputs closely match the measured data, thereby determining the best-fit values.

Appendix K.5. Practical Considerations and Limitations

  • Computational Cost:
    The complexity of solving higher-order, nonlinear PDEs in three dimensions necessitates efficient numerical solvers and often parallel computing resources. Adaptive meshing and high-order integration are crucial for managing computational demands.
  • Boundary Conditions:
    Correct implementation of boundary conditions is critical. For interference simulations, absorbing boundary conditions (e.g. perfectly matched layers) are often necessary to minimise spurious reflections.
  • Ensemble Averaging:
    Due to the chaotic nature of sub-Planck dynamics, it may be necessary to perform simulations over multiple initial conditions and average the results to obtain robust, reproducible predictions.

Appendix K.6. Fitting the Cosmological Constant via Persistent Waves

A longer-term goal is to reconcile the observed dark energy density ρ Λ 10 29 g c m 3 with the time-averaged sub-Planck oscillation energy. In principle, one treats the persistent oscillations as yielding an effective vacuum energy. Numerically, this involves:
  • Identifying the key dimensionless STM couplings (e.g. λ k , η ) that most strongly affect the vacuum energy in the Functional Renormalisation Group (FRG) flow.
  • Integrating the RG flow down to low momentum scales (as k 0 ) to extract the final effective offset Δ E e f f .
  • Tuning the parameters so that this offset matches the measured ρ Λ .
While a complete multiscale run is beyond the scope here, the framework offers a promising avenue for reconciling the cosmological constant puzzle within the STM model.

Appendix K.7. Summary

  • Spatial Discretisation:
    The domain Ω is discretised using high-order shape functions to accurately represent higher-order derivatives.
  • Time Integration:
    Implicit schemes, such as Crank–Nicolson, are employed to ensure stability in the face of stiff, nonlinear dynamics.
  • Nonlinear Solvers:
    Iterative methods (e.g. Newton–Raphson) are used to handle nonlinear terms, with convergence monitored via residual norms.
  • Parameter Fitting:
    A cost function quantifies the deviation between simulation outputs and experimental data, allowing for systematic parameter extraction.
  • Long-Term Objectives:
    Future work aims to match the effective vacuum energy from persistent sub-Planck oscillations to the observed cosmological constant, bridging microscale dynamics with cosmic observations.
This FEA approach is pivotal for translating the refined STM model’s theoretical predictions into tangible, testable phenomena.

Appendix L. Nonperturbative Analysis in the Refined STM Model

Appendix 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 refined Space–Time Membrane (STM) model arise from nonperturbative effects. These include:
  • Solitonic excitations: Stable, localised solutions arising from the nonlinearity of the STM equations.
  • 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, variational methods, and numerical soliton analysis.

Appendix 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 Γ k [ ϕ ] evolves as quantum fluctuations are integrated out down to a momentum scale k. The evolution equation, known as the Wetterich equation, is given by:
k Γ k [ ϕ ] = 1 2 Tr k R k Γ k ( 2 ) [ ϕ ] + R k ,
where:
  • R k ( p ) is an infrared (IR) regulator that suppresses fluctuations with momenta p < k ,
  • Γ k ( 2 ) [ ϕ ] is the second functional derivative of the effective action,
  • The trace Tr represents an integration over momenta.

Appendix L.68.1. Local Potential Approximation (LPA) and Nonperturbative Potentials

Applying the Local Potential Approximation (LPA), the effective action takes the form:
Γ k [ ϕ ] = d 4 x 1 2 μ ϕ 2 + V k ( ϕ ) .
The running of the effective potential V k ( ϕ ) follows:
k V k ( ϕ ) = 1 2 d 4 p ( 2 π ) 4 k R k ( p ) p 2 + R k ( p ) + ϕ 2 V k ( ϕ ) .
Solving this equation reveals the scale dependence of vacuum structure and potential dynamical symmetry breaking. In particular, the appearance of nontrivial minima in V k ( ϕ ) signals spontaneous symmetry breaking and the potential emergence of multiple fermion generations.

Appendix L.3. Solitonic Solutions and Topological Defects

Appendix L.69.1. Kink Solutions in the STM Model

One of the most intriguing features of the STM model is the presence of solitonic excitations—stable, localised field configurations. Consider a double-well potential:
V ( ϕ ) = λ 4 ϕ 2 a 2 2 .
The classical field equation for a static solution in one spatial dimension is:
x 2 ϕ = λ ϕ ϕ 2 a 2 .
A kink solution interpolating between the vacua ϕ = ± a is:
ϕ ( x ) = a tanh λ 2 a x .
This represents a topological defect, as the field transitions between different vacuum states at spatial infinity.

Appendix L.69.2. Soliton Stability and Energy Calculation

The total energy of the kink solution is given by:
E = d x 1 2 x ϕ 2 + V ( ϕ ) .
Substituting ϕ ( x ) and solving the integral, we obtain:
E kink = 2 2 λ 3 a 3 .
Since this energy is finite, the kink is stable and does not decay. This provides a mechanism for the emergence of long-lived structures in the STM model.

Appendix L.69.3. Link to Fermion Generations

In the STM model, fermions couple to the displacement field u ( x , t ) via a Yukawa-like interaction:
L Yukawa = y ψ ψ u .
If u ( x ) develops multiple stable vacuum expectation values (VEVs), fermion masses are generated as:
m f = y u .
A hierarchy of solitonic vacua could lead to three distinct fermion mass scales, potentially explaining the existence of three fermion generations.

Appendix L.4. Influence on Gravitational Wave Ringdown

If solitons exist near black hole horizons, they alter the ringdown phase of gravitational waves. The modified quasi-normal mode (QNM) equation for perturbations in the STM model is:
2 V eff ( r ) ψ QNM = 0 .
The presence of solitonic structures modifies the effective potential V eff ( r ) , leading to a frequency shift:
Δ f QNM = β M M sol ,
where M is the black hole mass and M sol is the soliton mass. This shift could be observable via LIGO/Virgo gravitational wave detectors.

Appendix L.5. Illustrative Toy Model for Multiple Mass Scales

As a partial demonstration of how our renormalisation flow might yield more than one stable mass scale, consider a simplified ϕ 4 -type potential
V k ( ϕ ) = λ k ϕ 2 a k 2 2
where λ k , a k run with scale k. Numerically integrating the FRG equation (L.3) can reveal discrete minima ϕ 1 , ϕ 2 , ϕ 3 at a low-energy scale k 0 . Each minimum could correspond to a distinct fermion mass scale m f y ϕ . For instance, in a toy numeric run:
ϕ 1 = 1.0 , ϕ 2 = 3.2 , ϕ 3 = 9.8
m f , 1 : m f , 2 : m f , 3 = 1 : 3.2 : 9.8 .
While this does not match real quark or lepton mass ratios, it demonstrates how three stable vacua can arise (See Figure 5). In a more elaborate model (including Yukawa couplings and gauge interactions), such discrete RG fixed points might align with the observed generational hierarchy.
Mixing Angles & CP Phases: Achieving realistic CKM or PMNS mixing angles and CP-violating phases requires explicitly incorporating deterministic interactions between bimodal spinor fields and their mirror antispinor counterparts across the membrane, mediated by rapid oscillatory (zitterbewegung) effects as detailed in Appendix C.3.1. A complete numerical fit of the Standard Model fermion mass and mixing spectrum within this deterministic STM framework is left to future analysis, but we emphasise this mechanism as a central motivation for extending the phenomenological scope of the refined STM model.

Appendix L.6. Summary and Implications

This appendix provides a detailed nonperturbative analysis of the STM model, highlighting:
  • Functional Renormalisation Group (FRG): Governs the evolution of the effective potential and reveals dynamical symmetry breaking.
  • Solitonic Excitations: Stable kinks arise from the nonlinear potential, with finite energy and topological stability.
  • Fermion Generation Mechanism: Multiple stable vacua suggest a natural explanation for the three fermion generations.
  • Gravitational Wave Modifications: Solitons near black holes alter quasi-normal mode frequencies, providing an experimental test of the STM model.
In summary, the nonperturbative analysis of the refined STM model via FRG and soliton theory reveals a rich vacuum structure with profound implications for particle physics and gravity. These insights provide a deterministic basis for the emergence of multiple fermion generations, CP violation, and the stabilisation of black hole interiors—all without resorting to extra dimensions or intrinsic randomness.

Appendix M. Revised Derivation of Einstein Field Equations in the Refined STM Model

Appendix M.1. M.1 Original Derivation Recap

In the original Space–Time Membrane (STM) model [1], gravitational curvature was shown to emerge from classical elasticity. The core equation was
ρ 2 u t 2 = T 2 u ( E S T M + Δ E ( x , y , z , t ) ) 4 u + F e x t ,
where ρ is the membrane’s mass density, T its tension, and E S T M its baseline elastic modulus (with Δ E capturing local variations). Small metric perturbations h μ ν were defined via a strain tensor ϵ μ ν :
h μ ν = 2 ϵ μ ν = μ u ν + ν u μ .
Under appropriate assumptions, one recovers linearised Einstein equations,
h μ ν = 0 ,
with 4 u acting primarily as an ultraviolet (UV) regulator. This established the notion that elastic deformations can be interpreted geometrically, linking strain fields u μ to gravitational curvature.

Appendix M.2. Refined STM Wave Equation and Extensions

In the refined STM model, the classical PDE is augmented by additional higher-order terms and scale-dependent parameters:
ρ 2 u t 2 [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] 4 u + η 6 u γ u t + λ u 3 g u ( Ψ ¯ Ψ ) + F e x t ( x , t ) = 0 .
Here, η 6 u further regularises short-wavelength modes; γ t u introduces non-Markovian damping; λ u 3 is a self-interaction term; and g u Ψ ¯ Ψ couples the membrane to spinor fields. Compared to the original STM approach, these refinements modify the emergent geometry in strong-field and quantum-scale regimes.

Appendix M.3. Action Formulation and Metric Variation

In many emergent-gravity frameworks, one attempts to cast the membrane dynamics in a covariant action whose variation with respect to a metric g μ ν yields Einstein-like field equations. Below is an outline of how this can work in the refined STM model, with caveats for damping and higher-order terms that do not trivially follow from a standard minimally coupled scalar field.

Appendix M.75.1. Action Proposal

A representative action incorporating the refined STM parameters can be written schematically as
S = { 1 2 [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] ( 2 u ) 2 + η 2 ( 3 u ) 2 γ 2 ( u t ) 2 + λ 4 u 4 g 2 u 2 ( Ψ ¯ Ψ ) + L m a t t e r } g d 4 x ,
where denotes covariant derivatives in a notional 4D “membrane geometry”. The term L m a t t e r accounts for standard matter fields. The factor g is included to ensure covariance under (small) metric variations.
Caveats:
  • Damping γ as a Non-Conservative Term
    Strictly, a friction-like term γ t u 2 does not usually arise from a purely conservative action. In continuum mechanics, damping is often added via a Rayleigh dissipation function rather than an action. Here, one may treat γ as an effective parameter or use a pseudo-action approach. In practice, the PDE in Section M.2 emerges from a combination of variational and dissipative terms, so the “stress–energy” for damping must be interpreted carefully.
  • Higher-Order Operators
    Terms such as 2 u 2 and 3 u 2 typically introduce boundary terms upon integration by parts. One must assume suitable boundary conditions (e.g. fields vanishing at infinity) so that the PDE in Section M.2 is recovered without extra surface contributions.

Appendix M.75.2. Variation with Respect to g μν

Varying the above action w.r.t. the metric yields
δ S = 1 2 g ( T μ ν e l a s t i c + T μ ν ) δ g μ ν d 4 x ,
where T μ ν is the usual matter stress–energy tensor from L m a t t e r . The “elastic stress–energy” part, T μ ν e l a s t i c , arises from the membrane terms:
T μ ν e l a s t i c [ E S T M ( μ ) + Δ E ( x , t ; μ ) ] ( 2 u ) 2 g μ ν η ( 3 u ) 2 g μ ν + γ ( u t ) 2 g μ ν λ u 4 g μ ν + g u 2 ( Ψ ¯ Ψ ) g μ ν .
Note:
  • A fully rigorous variation for higher derivatives often yields terms with μ u ν u , etc. If one assumes isotropic or slowly varying solutions, such cross-terms can be rearranged into an effectively perfect-fluid–type expression ( s c a l a r f u n c t i o n ) × g μ ν .
  • The presence of γ in T μ ν e l a s t i c is also an approximation, as dissipation typically would not show up in a purely Hamiltonian stress–energy. This indicates that the emergent “Einstein-like” system is not purely conservative.

Appendix M.75.3. Resulting Einstein-like Equations

Imposing δ S / δ g μ ν = 0 yields:
R μ ν 1 2 g μ ν R + Λ g μ ν = 8 π G ( μ ) c 4 ( T μ ν + T μ ν e l a s t i c ) ,
where:
  • Λ can be identified with any vacuum contribution (including possible cosmological terms from Δ E if uniform),
  • G ( μ ) 1 / E S T M ( μ ) captures the idea that membrane stiffness inversely sets the gravitational coupling.
In large-scale or low-energy regimes—where the high-order terms 4 and 6 are negligible—the usual Einstein equations re-emerge. Near strong fields or high frequencies, however, Δ E ( x , t ; μ ) , η 6 u , and other corrections become non-trivial, potentially altering singularities and gravitational wave signals.

Appendix M.4. Physical Implications

  • Singularity Regularisation
    The higher-order derivative terms ( 6 u ) can suppress short-wavelength instabilities, preventing the curvature from diverging, thus avoiding classical singularities. This matches the refined STM discussions of black hole interiors (see main text and Appendix F).
  • Modified Gravitational Waves
    Because the membrane stiffness is scale-dependent, wave equations for metric perturbations gain extra terms that may shift quasi-normal mode frequencies in black hole mergers or alter early-universe gravitational wave spectra.
  • Localised Vacuum Energy Variation
    The spatial/temporal fluctuations Δ E ( x , t ; μ ) introduce position-dependent effective vacuum energy, leading to local gravitational effects. This ties in with Appendix H’s idea of “density-driven vacuum energy” in the refined STM approach.

Appendix M.5. Time Dilation and Redshift

In the original STM model, gravitational time dilation followed from the identification g 00 = 1 + h 00 1 + 2 ϵ 00 , where ϵ 00 is the local strain. Because ϵ 00 depends on u ( x , t ) , any change in E S T M or Δ E modifies the local amplitude of u, thus changing h 00 and directly affecting gravitational redshift. In the refined model:
  • Running Couplings: Δ E ( x , t ; μ ) and η can vary with scale, so the relation between the strain u and the time dilation factor is not purely linear.
  • Strong-Field Enhancements: Near compact objects, the additional 4 and 6 terms induce steeper gradients in u, further modifying the redshift. These effects might be testable if accurate gravitational redshift data become available for extreme regions (e.g. near black hole horizons).

Appendix M.6. Summary and Caveats

In summary, the refined STM model preserves and extends the original emergent-gravity interpretation:
  • Scale-Dependent Gravitational Coupling:
    E S T M ( μ ) acts like an inverse G ( μ ) , with Δ E contributing to local vacuum or matter effects.
  • Singularity Avoidance and Extra Terms:
    The 6 u and other higher-order operators ensure UV regularity, offering a path to prevent classical singularities in gravitational collapse.
  • Damping and Non-Hamiltonian Effects:
    The presence of γ t u 2 and other dissipative terms means the action-based derivation of T μ ν e l a s t i c must be understood with care. In practice, certain terms are introduced phenomenologically, reflecting real membrane friction.
  • Phenomenological Agreement:
    At large scales (low energy), these modified Einstein-like equations reduce to standard General Relativity. At smaller scales, additional corrections—potentially observable in gravitational wave ringdowns, black hole interiors, or local redshift anomalies—provide avenues for testing the refined STM.
Overall, while some higher-order and dissipative contributions do not stem from a standard purely conservative action, the approach here indicates how an Einstein-like system emerges from elasticity. The resulting extended field equations reflect both sub-Planck chaotic oscillations (Appendix D, G) and scale-dependent stiffness (Appendix H, J), bridging the gap between classical GR phenomenology and quantum/short-wavelength corrections in the refined STM framework.

Appendix N. Glossary of Symbols

Appendix N.1. Fundamental Constants

]@ >p() * 0.1113 >p() * 0.8887@ () [b]Symbol [b]Definition
() c Speed of light in vacuum.
Reduced Planck’s constant, = h 2 π .
G Newton’s gravitational constant.
ρ Effective mass density of the membrane (used in the refined STM wave equation).
Λ Cosmological constant, often linked to vacuum energy density.
()

Appendix N.2. Elastic Membrane and Field Variables

]@ >p() * 0.1378 >p() * 0.8622@ () [b]Symbol [b]Definition
() u ( x , t ) Classical displacement field of the four-dimensional elastic membrane.
u ^ ( x , t ) Operator form of the displacement field (canonical quantisation).
π ( x , t ) Conjugate momentum, π = ρ t u .
E S T M ( μ ) Scale-dependent baseline elastic modulus, acting as an inverse gravitational coupling at large scales.
Δ E ( x , t ; μ ) Local stiffness fluctuations, possibly time- and space-dependent, also running with renormalisation scale μ .
η Coefficient for the 6 u term providing ultraviolet (UV) regularisation in the refined STM model.
γ Damping parameter (may be extended to non-Markovian damping).
V ( u ) Potential energy function for the displacement field u.
λ Self-interaction coupling constant (e.g. in a λ u 3 or λ u 4 term).
F e x t ( x , t ) External force contributions acting on the membrane’s displacement field.
()

Appendix N.3. Gauge Fields and Internal Symmetries

]@ >p() * 0.1405 >p() * 0.8595@ () [b]Symbol [b]Definition
() A μ ( x , t ) U ( 1 ) gauge field (photon-like).
W μ a ( x , t ) S U ( 2 ) gauge fields, a = 1 , 2 , 3 (weak interaction bosons).
G μ a ( x , t ) S U ( 3 ) gauge fields (gluons), a = 1 , , 8 .
T a Gauge group generators (e.g. T a = σ a 2 in S U ( 2 ) ).
g 1 , g 2 , g 3 Gauge coupling constants for U ( 1 ) , S U ( 2 ) , S U ( 3 ) .
F μ ν U ( 1 ) field strength, F μ ν = μ A ν ν A μ .
W μ ν a S U ( 2 ) field strength tensor.
G μ ν a S U ( 3 ) field strength tensor.
f a b c Structure constants of non-Abelian gauge groups (e.g. ϵ a b c for S U ( 2 ) ).
()

Appendix N.4. Fermion Fields and Deterministic CP Violation

]@ >p() * 0.1140 >p() * 0.8860@ () [b]Symbol [b]Definition
() Ψ ( x , t ) Two-component spinor field arising from the bimodal decomposition of u ( x , t ) ; underlies the emergence of internal gauge symmetries.
Ψ ( x , t ) Mirror antispinor field on the opposite “face” of the membrane.
Ψ ¯ Ψ Standard fermion bilinear (Yukawa-like term), combining spinor Ψ and mirror antispinor Ψ .
v Vacuum expectation value (VEV) of u ( x , t ) (or of a bilinear, depending on the context).
y f Yukawa coupling between spinor fields and the displacement u.
θ i j ( x , t ) Complex phase arising deterministically (e.g. via zitterbewegung) between spinor and mirror antispinor fields.
M f Effective fermion mass matrix that acquires complex phases from deterministic oscillations, yielding CP violation.
()

Appendix N.5. Renormalisation Group and Couplings

]@ >p() * 0.1113 >p() * 0.8887@ () [b]Symbol [b]Definition
() μ Renormalisation scale at which the elastic parameters (e.g. E S T M ) are measured or evolved.
g e f f Effective coupling constant parameterising running elastic or interaction strengths in the STM model.
β ( g ) Beta function governing the scale dependence (RG flow) of the coupling g.
α s Typically denotes the strong coupling constant in the SU(3) sector.
Λ Q C D QCD confinement scale analogue; in STM, one often interprets “colour confinement” via classical oscillator tension.
Z k ( ϕ ) Scale-dependent wavefunction renormalisation factor (FRG context).
()

Appendix N.6. Path Integral and Operator Formalism

]@ >p() * 0.1113 >p() * 0.8887@ () [b]Symbol [b]Definition
() D u , D Ψ Functional integration measures in path integral quantisation.
Z Path integral or partition function.
ξ Gauge-fixing parameter (e.g. in covariant gauges).
c a , c ¯ a Ghost fields introduced by the Faddeev–Popov procedure in non-Abelian gauge theories.
()

Appendix N.7. Nonperturbative Effects and Solitonic Structures

]@ >p() * 0.1113 >p() * 0.8887@ () [b]Symbol [b]Definition
() Γ k [ ϕ ] Scale-dependent effective action in the Functional Renormalisation Group (FRG) framework.
R k ( p ) Infrared regulator function that suppresses fluctuations below the momentum scale p < k in FRG approaches.
Γ k ( 2 ) [ ϕ ] Second functional derivative (w.r.t. fields) of the effective action Γ k .
V k ( ϕ ) Scale-dependent effective potential evolving with the RG scale k.
ϕ Generic scalar field variable used in nonperturbative FRG analyses, which can represent a coarse-grained version of u or an auxiliary field.
ψ Q N M Perturbation wavefunction in black hole ringdown analyses (quasi-normal modes).
E s o l Energy of a solitonic (kink-like) solution emerging in nonperturbative analyses.
M s o l Effective mass or energy scale associated with a solitonic core (e.g. in black hole interiors or topological defects).
Δ f Q N M Frequency shift in quasi-normal modes due to the presence of solitonic structures near black hole horizons.
()

Appendix N.8. Zitterbewegung and Deterministic CP Violation

]@ >p() * 0.1945 >p() * 0.8055@ () [b]Symbol [b]Definition
() Zitterbewegung Rapid deterministic oscillations (classically reminiscent of “Dirac trembling”) between spinor fields on the membrane and mirror fields; crucial for generating CP phases.
()

Appendix N.9. Finite Element Analysis

]@ >p() * 0.1113 >p() * 0.8887@ () [b]Symbol [b]Definition
() Ω Spatial domain (or mesh) used in finite element simulations of the refined STM PDE.
{ x i } Nodal points within the finite element mesh.
N i ( x ) Shape (or basis) functions used to discretise u ( x , t ) over each finite element.
J Cost function quantifying discrepancy between simulation outputs and experimental data; used for parameter fitting (e.g. λ , y , η ).
()

Appendix N.10. Experimental and Observational Signatures

]@ >p() * 0.1113 >p() * 0.8887@ () [b]Symbol [b]Definition
() Δ f Q N M Shift in quasi-normal mode frequencies of black hole analogues, possibly observable in gravitational wave data.
Ω Λ Fractional density of vacuum energy in cosmological observations, e.g. Ω Λ 0.7 in Λ CDM.
δ T / T Relative temperature anisotropies in the cosmic microwave background, possibly linked to inhomogeneous vacuum stiffness.
()

Appendix N.11. Summary

This glossary compiles the symbols used throughout the refined Space–Time Membrane (STM) model, covering:
  • Elastic Membrane Variables, including high-order derivatives 4 , 6 , and scale-dependent parameters Δ E .
  • Gauge and Spinor Fields, under U(1), SU(2), and SU(3) symmetries, all emerging from the bimodal construction of u.
  • Renormalisation Group Tools, used to understand UV suppression, discrete vacuum structures, and multi-loop effects.
  • Nonperturbative Structures, such as solitons and topological defects, which can explain phenomena like black hole interior stabilisation and multiple fermion generations.
  • Experimental and Observational Probes, ranging from metamaterial analogues to gravitational wave ringdowns.
By consolidating these definitions, Appendix N ensures notational clarity and consistency across the entire refined STM model.

References

  1. Swann, P. The Space–Time Membrane Model: Unifying Quantum Mechanics and General Relativity through Elastic Membrane Dynamics. Preprints 2024. [Google Scholar] [CrossRef]
  2. Einstein, A. Die Grundlage der allgemeinen Relativitätstheorie. Annalen der Physik 1916, 49(7), 769–822. [Google Scholar] [CrossRef]
  3. Misner, C. W.; Thorne, K. S.; Wheeler, J. A. Gravitation; W. H. Freeman, 1973. [Google Scholar]
  4. Wald, R. M. General Relativity; University of Chicago Press, 1984. [Google Scholar]
  5. Peskin, M. E.; Schroeder, D. V. An Introduction to Quantum Field Theory; Addison–Wesley, 1995. [Google Scholar]
  6. Weinberg, S. The Quantum Theory of Fields, Volume I; Cambridge University Press, 1995. [Google Scholar]
  7. Mandel, L.; Wolf, E. Optical Coherence and Quantum Optics; Cambridge University Press, 1995. [Google Scholar]
  8. Greenstein, G.; Zajonc, A. G. The Quantum Challenge: Modern Research on the Foundations of Quantum Mechanics; Jones and Bartlett, 2006. [Google Scholar]
  9. Zeh, H. D.; Joos, E.; Kiefer, C.; Giulini, D. J. W.; Kupsch, J.; Stamatescu, I. O. Decoherence and the Appearance of a Classical World in Quantum Theory; Springer.
  10. Feynman, R. P. The Theory of Positrons. Physical Review 1949, 76, 749–759. [Google Scholar] [CrossRef]
  11. Feynman, R. P.; Hibbs, A. R. Quantum Mechanics and Path Integrals; McGraw–Hill, 1965. [Google Scholar]
  12. Hawking, S. W. Particle Creation by Black Holes. Communications in Mathematical Physics 1975, 43, 199–220. [Google Scholar] [CrossRef]
  13. Penrose, R. Gravitational Collapse and Space–Time Singularities. Physical Review Letters 1965, 14, 57–59. [Google Scholar] [CrossRef]
  14. Riess, A. G.; et al. Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant. Astronomical Journal 1998, 116, 1009–1038. [Google Scholar] [CrossRef]
  15. Perlmutter, S.; et al. Measurements of Ω and Λ from 42 High-Redshift Supernovae. Astrophysical Journal 1999, 517, 565–586. [Google Scholar] [CrossRef]
  16. Weinberg, S. The Cosmological Constant Problem. Reviews of Modern Physics 1989, 61, 1–23. [Google Scholar] [CrossRef]
  17. Cramer, J. G. The Transactional Interpretation of Quantum Mechanics. Reviews of Modern Physics 1986, 58, 647–687. [Google Scholar] [CrossRef]
  18. Wheeler, J. A.; Feynman, R. P. Interaction with the Absorber as the Mechanism of Radiation. Reviews of Modern Physics 1945, 17, 157–181. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

Disclaimer

Terms of Use

Privacy Policy

Privacy Settings

© 2025 MDPI (Basel, Switzerland) unless otherwise stated