Preprint
Article

This version is not peer-reviewed.

The Space-Time Membrane Model: Unifying Quantum Mechanics and General Relativity through Elastic Membrane Dynamics

Submitted:

15 December 2024

Posted:

17 December 2024

Read the latest preprint version here

Abstract

We present the Space-Time Membrane (STM) model, which treats our four-dimensional spacetime as thesurface of an elastic membrane, with a mirror universe on the opposite side. Gravitational curvature correspondsto membrane deformation induced by energy external to the membrane, while homogeneous internal energydoes not produce curvature. Particles emerge as oscillatory excitations on the membrane’s surface, with theirmirror antiparticles on the far side. These oscillations modulate the membrane’s local elastic properties, yieldinggravitational and quantum-like phenomena. A modified elastic wave equation, incorporating tension, bendingstiffness, and space-time-dependent elastic variations, reproduces key features of General Relativity (GR) andaspects of Quantum Field Theory (QFT). Identifying strain fields with metric perturbations recovers equationsstructurally identical to the Einstein Field Equations. Time dilation, gravitational effects, and non-singular blackhole interiors arise naturally from these mechanics. Moreover, stable standing waves and controlled stiffnessvariations produce interference patterns and entanglement analogues, resembling quantum experiments withina deterministic, continuum framework. Interpreting photons as composite particle–antiparticle oscillationspreserves their masslessness, correct polarisations, U(1) gauge symmetry, and Lorentz invariance, consistentwith QFT. High-energy processes converting photons into particle pairs support this view. By adjusting anintrinsic coupling constant, time-averaged stiffness variations match observed vacuum energy, reproducing thecosmological constant. Furthermore, spatial variations in persistent wave energy may explain dark matter-likedistributions and address the Hubble tension. The STM model thus offers a geometric, deterministic approach tolinking particle-scale dynamics with cosmological phenomena, potentially resolving long-standing conceptualissues such as the black hole information loss paradox.

Keywords: 
;  ;  ;  ;  ;  ;  

1. Introduction

Modern physics rests on the twin pillars of General Relativity (GR) and Quantum Mechanics (QM). GR attributes gravity to spacetime curvature, while QM and Quantum Field Theories (QFTs) capture phenomena at subatomic scales [1–3]. Despite their successes, a complete reconciliation of these frameworks remains elusive [4,5]. Efforts such as String Theory and Loop Quantum Gravity attempt to unify gravity with quantum principles, yet no consensus has emerged, and challenges like the black hole information paradox persist [6–8].
The Space-Time Membrane (STM) model offers an alternative approach. Instead of introducing additional spatial dimensions (as in String Theory) or quantising spacetime geometry (as in Loop Quantum Gravity), the STM model conceptualises our four-dimensional spacetime as one side of a 4D elastic membrane with a mirror universe on the opposite side (See Figure 1). Here, energy residing “outside” the membrane generates local curvature, analogous to mass-energy distributions in GR, while energy stored uniformly within the membrane does not induce curvature. Particles appear as oscillations on this membrane, and their mirror counterparts reside on the far side. Particle–mirror antiparticle attraction and particle–antiparticle repulsion modulate the membrane’s elastic properties, enabling gravitational and quantum-like effects to emerge naturally (See Figure 2).
Key aspects of the STM model include:
  • Deriving a modified elastic wave equation that governs membrane dynamics, bridging gravitational and quantum phenomena.
  • Interpreting photons as composite oscillations, retaining masslessness, gauge invariance, Lorentz invariance, and correct polarisation states while providing a geometric explanation for phenomena such as interference and entanglement (see Appendices A–E for detailed derivations).
  • Explaining how large curvatures, such as those in black hole interiors, are tamed by increasing membrane stiffness, avoiding singularities and offering routes to resolve the information paradox (see Appendices F–H).
  • Relating vacuum energy and the cosmological constant to a time-averaged stiffness offset, linking particle-scale oscillations to cosmological parameters (Appendix K).
  • Ensuring internal consistency by deriving the force function from a potential energy functional (Appendix B) and showing the emergence of Einstein Field Equations (Appendix C).
This paper is structured as follows; The Methods section details the theoretical framework and derivations. The Results section highlights key findings, while the Discussion and Conclusion situate these findings relative to established theories and emphasises potential empirical avenues for testing. The Appendices provide full mathematical derivations, modifications to wave equations, and further technical details supporting the main text.
A glossary of symbols is also provided in Appendix M for convenience.

2. Methods

2.1. Conceptual Framework and Analogy

The STM model begins with the analogy that our four-dimensional spacetime corresponds to the surface of an elastic membrane, while a mirror universe resides on the opposite side. Each point in spacetime corresponds to a point on this membrane, and gravitational curvature arises as deformations of the membrane’s geometry. Particle–mirror particle interactions occur across the membrane, modulating its elastic properties and enabling energy transfer between uniform (homogeneous) and localised states of deformation. The details of how this analogy is mathematically formalised, including the setup of coordinates and boundary conditions, are presented in Appendix A.

2.2. Elasticity and Material Parameters

Linear elasticity theory describes how small deformations relate to applied forces. We define the displacement field u ( x , y , z , t ) , representing how each point on the equilibrium membrane shifts under applied stresses. The strain tensor ϵ i j measures deformation, and the stress tensor σ i j relates linearly to ϵ i j via Hooke’s law:
σ i j = λ δ i j ϵ k k + 2 μ ϵ i j ,
where λ and μ are Lamé parameters. These parameters relate to the intrinsic elastic modulus E STM and Poisson’s ratio ν . The choice of E STM is crucial for ensuring that, when strains are identified with metric perturbations, the resulting field equations match the structure of Einstein’s theory (Appendix C).
In the STM model, we introduce Δ E ( x , y , z , t ) to capture changes in local stiffness induced by particle oscillations. Unlike a static material constant, Δ E depends on the energy density associated with oscillations E ( x , y , z , t ) and an intrinsic coupling constant α :
Δ E ( x , y , z , t ) = α E ( x , y , z , t ) .
This fundamental parameter α governs how sensitively the membrane’s stiffness responds to local particle oscillation energies, ensuring that short-range oscillation-to-stiffness conversions are captured. Whilst the time-averaged component of Δ E remains predominantly homogeneous, slight spatial variances due to persistent wave energy can introduce minor deviations in vacuum energy across the membrane. These variances may provide a mechanism for explaining dark matter distributions and the Hubble tension observed in cosmological measurements. The rationale, derivation, and implications of Δ E and α are discussed in detail in Appendices A and J.
To determine the appropriate value of the coupling constant α , and to assess whether a single α can describe both photon- and electron-induced interference patterns, we propose a numerical simulation strategy. By implementing a finite element analysis (FEA) of the STM wave equation in a double-slit configuration for both photons and electrons, we can iteratively adjust α to match the observed fringe spacing and intensity distributions in standard quantum interference experiments. If the optimal α value differs significantly between photon- and electron-based scenarios, this may indicate scale-dependent behaviour or the need for additional coupling constants. Thus, the FEA results will guide the refinement of our parameter space, offering empirical routes to identifying whether α alone suffices or if further constants are required (Appendix L).

2.3. Incorporating Particle–Mirror Particle Dynamics

Particles are modelled as oscillations on one side of the membrane, and their mirror antiparticles as oscillations on the opposite side. Attractive interactions between a particle and its mirror antiparticle pull energy “outside” the membrane, creating localised curvature and mass-energy. Conversely, repulsive interactions between a particle and its own antiparticle push energy back into the membrane’s homogeneous background. This interplay ensures a cycle of energy transfer that underpins both gravitational effects and quantum-like patterns. The mathematical formalism capturing these interactions is introduced in Appendix A, with subsequent applications in Appendices D–E (for quantum-like phenomena) and F–H (for black hole interiors and Hawking radiation).

2.4. Deriving the Modified Elastic Wave Equation

Starting from Newton’s second law for a continuum element:
ρ 2 u i t 2 = σ i j x j + f i ,
and substituting the stress–strain relationships and terms for tension T and bending stiffness, we obtain a baseline elastic wave equation. Introducing Δ E modifies the bending term to:
ρ 2 u t 2 = T 2 u [ E STM + Δ E ( x , y , z , t ) ] 4 u + F ext .
The Laplacian 2 term arises naturally from tension, representing wave-like behaviour akin to a vibrating membrane (e.g., a drumhead), while the biharmonic operator 4 encodes bending stiffness effects, ensuring that sharp curvatures cost significant energy. Incorporating Δ E means that oscillatory energy densities directly alter the local bending rigidity. Detailed derivations, including boundary conditions, scaling arguments, and linearisation steps, are found in Appendix A.

2.5. Force Function and Persistent Waves

The external force F ext encapsulates influences on the STM membrane beyond its inherent tension and bending stiffness, such as interactions between particles and mirror particles or energy redistributions that affect the membrane’s geometry. To integrate persistent waves with aligned wavelengths into the STM model, F ext is derived from a potential energy functional. This approach ensures that the force remains conservative and supports sustained oscillations.
Key features of the force function include:
  • Tension Term: Represents the standard wave-like behaviour resulting from membrane tension, analogous to vibrations in a drumhead.
  • Bending Stiffness Term: Accounts for the membrane’s resistance to bending. This term is dynamically modulated by Δ E ( x , t ) , enabling the stabilisation of specific wave modes.
  • Feedback Mechanism: The modulation Δ E ( x , t ) , through α , introduces a feedback loop that reinforces oscillations at desired wavelengths, ensuring the persistence of wave patterns aligned with composite photon wavelengths.
For a comprehensive derivation of F ext from the potential energy functional, including functional variations and the incorporation of Δ E ( x , t ) , please refer to Appendix B.

2.6. Relating Strain to Curvature and Einstein Field Equations

By identifying the metric perturbations h μ ν with the strain fields ϵ μ ν , we map membrane deformation onto geometric perturbations in spacetime. Varying an action constructed from the elastic energy plus matter contributions yields equations identical in form to the Einstein Field Equations in the linearised regime (Appendix C). This establishes that the STM model does not contradict known gravitational laws and, in fact, reproduces them from basic mechanical principles.

2.7. Composite Photons and Persistent Oscillations

In standard quantum field theory, photons are elementary excitations of the electromagnetic field. Here, we treat the photon as a composite oscillation—a particle–antiparticle bound state oscillating across the membrane. Because the net deformation over one oscillation cycle is zero, the photon remains massless and respects U(1) gauge symmetry and Lorentz invariance. This reinterpretation is not forced; it emerges naturally once we allow local stiffness variations and energy transfer through the membrane.
These persistent oscillations enable stable standing waves that replicate phenomena like double-slit interference deterministically (Appendix D) and yield correlated modes analogous to quantum entanglement (Appendix E), all without introducing intrinsic randomness.

2.8. Extreme Regimes: Black Hole Interiors and Cosmological Parameters

When energy concentrates in a tiny region (as inside a black hole), the membrane’s stiffness increases dramatically due to Δ E . Instead of infinite curvature and a singularity, the membrane supports finite-energy standing waves, storing information and offering a potential resolution to the information paradox (Appendix F). Perturbations to Hawking-like radiation emerge, enabling slow information leakage (Appendix G).
On cosmological scales, time-averaging Δ E yields a uniform offset in stiffness interpreted as vacuum energy. By tuning α , we can match the observed value of Λ (Appendix K), linking microscopic oscillations to the universe’s accelerated expansion.

2.9. Introducing a Density-Driven Coupling Constant β

In addition to the intrinsic coupling constant α that governs immediate, local oscillation-to-stiffness conversions, we introduce a second coupling constant, β . While α handles direct, pointwise responses of the membrane to oscillation energy densities, β is posited to couple the spatial and temporal distributions of persistent wave energy to more extended modifications of vacuum energy density.
Formally, consider a measure of the persistent wave energy density ρ waves ( x , y , z ) , obtained by time-averaging and spatially smoothing the underlying oscillation energies. The effective vacuum energy offset Δ E eff ( x , y , z ) can then be expressed as:
Δ E eff ( x , y , z ) = Δ E ( x , y , z , t ) t + β F [ ρ waves ( x , y , z ) ] ,
where · t denotes time-averaging, and F is an integral or smoothing operator that aggregates persistent wave energy distributions. By adjusting β , we can explore how spatial inhomogeneities in persistent wave distributions lead to local vacuum energy variations. Such variations may influence local expansion rates of the universe, offering a mechanical, continuum-based explanation for the Hubble tension (Appendix I).
Thus, the STM model, with both α and β , provides flexibility in capturing phenomena ranging from local quantum-like effects and gravitational curvature to global cosmological parameters. We outline a finite element analysis approach in an additional appendix (Appendix L) to test whether a single α suffices for different particle types or if further constants are necessary, providing a practical route to refining our parameter space.

3. Results

3.1. Unified Emergence of Gravity and Quantum-Like Behaviour

From the outlined methods, the STM model yields a single elastic wave equation capturing both gravitational and quantum-like phenomena. Identifying strain with metric perturbations and varying the corresponding actions leads to equations structurally identical to EFE, ensuring that classical gravitational effects appear as large-scale, low-frequency modes of membrane deformation [1–3]. This approach avoids introducing additional dimensions (as in String Theory) or discretising spacetime (as in Loop Quantum Gravity), providing a complementary continuum-based viewpoint.

3.2. Composite Photons: Ensuring QFT Compatibility and Masslessness

Treating photons as composite oscillations remains consistent with QFT as they remain massless, maintain U(1) gauge symmetry, Lorentz invariance, and exhibit the correct two transverse polarisation states. At low energies, the effective theory is indistinguishable from standard low-energy quantum electrodynamics (QED) [4–6]. High-energy processes that convert photons into particle pairs support this composite picture [7,8]. Thus, the model’s reinterpretation of photons does not conflict with QFT principles; instead, it offers a geometric narrative for photon properties and interactions.

3.3. Deterministic Interference and Entanglement Analogues

Applying double-slit boundary conditions to the STM wave equation results in stable, steady-state solutions that produce interference fringes akin to quantum double-slit experiments [9–11]. Unlike the probabilistic interpretation in QM, these patterns emerge deterministically from the elastic response of the membrane’s varying stiffness (Appendix D).
Similarly, multiple oscillations lead to non-factorisable mode patterns that serve as deterministic analogues of quantum entanglement (Appendix E). While not claiming to replace standard QM interpretations, this demonstrates that complex quantum-like patterns can arise from classical mechanical principles under suitable conditions.

3.4. Black Hole Interiors Without Singularities

In general relativity, gravitational collapse can produce singularities where known physics breaks down. In the STM model, as deformation becomes extreme, the enhanced stiffness from Δ E prevents infinite curvature, replacing singularities with finite-amplitude standing waves [12,13]. This stabilises the interior structure of black holes, providing a coherent explanation for information encoding in stable wave patterns (Appendix F). Thus, the model suggests that no true singularity need form, preserving the consistency of physics even at extreme densities.

3.5. Modified Hawking Radiation and Information Leakage

Classical Hawking radiation is thermal and does not, by itself, restore information. Here, non-thermal corrections arise from the altered internal structure of black holes. Highly redshifted signals can escape gradually, carrying imprints of the interior standing waves, and over long timescales, this process slowly releases information [14]. By breaking the strict thermal nature of emission, the model provides a path to unitary evolution and addresses the black hole information paradox (Appendix G).

3.6. Connecting Vacuum Energy and the Cosmological Constant

Time-averaging Δ E ( x , y , z , t ) over many oscillation cycles leaves a uniform component Δ E interpreted as vacuum energy (Appendix K). However, slight special variances in Δ E due to persistent waves introduce minor deviations in vacuum energy across the membrane. These variances offer potential explanations for dark matter distributions and Hubble tension, linking microscopic wave dynamics to macroscopic cosmological observations. By adjusting α , the observed cosmological constant Λ is matched, aligning with supernova observations of an accelerating universe [15,16] and the broader cosmological constant problem [17]. Furthermore, slight spatial variances in Δ E due to persistent wave energy may address phenomena such as dark matter distributions and the Hubble tension, providing a continuum-mechanics-based explanation for these cosmological discrepancies [15–17]. This insight provides a fresh perspective on the origin of Λ and the universe’s accelerated expansion.

4. Discussion

4.1. Unifying Quantum and Gravitational Concepts

The Space-Time Membrane (STM) model provides a geometrical framework in which both gravitational and quantum phenomena emerge as manifestations of elastic properties in a four-dimensional membrane. By relating membrane strains to metric perturbations, the model recovers the structure of the Einstein Field Equations and thus aligns with General Relativity in the appropriate limits. Simultaneously, by interpreting particles as membrane oscillations and incorporating local stiffness variations Δ E , the STM model captures quantum-like effects—such as interference patterns and entanglement analogues—without invoking intrinsic randomness or modifying existing Quantum Field Theories (QFTs).
This dual success suggests that the elusive goal of unifying GR and QM may not require entirely new fundamental principles, but rather a different perspective on familiar mathematics and physics. Instead of positing a quantised geometry or introducing additional spatial dimensions, the STM model posits that geometry and quantum phenomena arise from continuum elasticity in a membrane-like substrate. Such a viewpoint offers a deterministic underpinning to phenomena traditionally considered probabilistic.

4.2. Photons, Gauge Invariance, and Consistency with QFT

The interpretation of photons as composite particle–antiparticle oscillations within the membrane is particularly illuminating. While this view might initially seem to depart from the standard QFT description, it ultimately respects gauge invariance, Lorentz invariance, masslessness, and the correct polarisation states of the photon. The resulting low-energy effective theory matches the predictions of conventional electromagnetism, ensuring no conflict with established principles. Moreover, evidence of photon–pair interconversion in high-energy processes supports the idea that photons can be understood as bound oscillatory states that can fragment into matter–antimatter pairs and vice versa. Thus, the STM model’s reinterpretation of photons does not undermine the successes of QFT; it instead provides a geometric-mechanical narrative compatible with known physics.

4.3. Deterministic Analogues of Quantum Phenomena

The deterministic interference patterns and entanglement analogues derived from stable standing wave solutions challenge the notion that probabilistic outcomes and wavefunction collapse are indispensable features of quantum phenomena. While the STM model does not claim to replace standard QM interpretations entirely, it does illustrate that complex quantum behaviours can emerge from classical-like wave equations, provided the underlying elasticity is sufficiently intricate and dynamic. This insight may motivate further investigations into how quantum-like patterns arise in other non-quantum, deterministic systems.
While much of our discussion has focused on composite photons to illustrate interference and entanglement analogues, it is important to emphasise that the underlying principles of the STM model are not limited to massless excitations. In fact, the same elastic feedback mechanisms, wherein local changes in the elastic modulus Δ E ( x , y , z , t ) respond dynamically to particle oscillation energy, can be applied to particles such as electrons. Since electrons are well-known to produce interference patterns in double-slit experiments, their associated wave-like disturbances on the spacetime membrane would similarly form stable, persistent patterns under the STM framework. Though the specific parameter values—such as frequencies and amplitudes—would differ due to the electron’s mass and associated energy scales, the fundamental process of stiffness modulation and energy redistribution remains intact. Thus, the STM model’s approach to explaining quantum-like phenomena through deterministic, elastic wave dynamics extends beyond photons, potentially encompassing a broad range of particle types and experimental scenarios.
Whilst we have assumed a single intrinsic coupling constant α to reproduce the observed interference patterns for both photons and electrons, it is feasible to undertake numerical simulations based on the STM model’s governing equations. By inputting experimentally known parameters—such as slit geometries, particle energies, and detection screen distances—into the wave equation and incorporating the chosen value(s) of α , one can use numerical methods (e.g., finite difference or finite element techniques) to solve for steady-state or time-harmonic solutions. Comparing the resulting intensity distributions against established experimental data for both photon and electron double-slit experiments would then indicate whether a single α suffices, or if a scale-dependence or additional coupling constants are required. While such numerical investigations may be computationally intensive and involve careful approximations, they represent a practical avenue for refining the STM model’s parameters and enhancing its predictive power across different particle types and energy scales.

4.4. Black Holes, Singularity Avoidance, and Information Retention

A longstanding problem in GR is the formation of singularities within black holes, where conventional physics breaks down. In the STM model, increasing stiffness counters runaway curvature, stabilising finite-energy standing wave solutions at the core of what would otherwise be a singularity. This result removes the pathological infinite curvature predicted by classical GR, offering a coherent explanation that remains within a continuum mechanical framework.
Furthermore, modifications to the Hawking-like radiation spectrum and evaporation rates enable black holes to release information slowly over long timescales. This mechanism provides a path to maintaining unitarity, a cherished principle in quantum physics, and potentially resolves the black hole information paradox. By embedding information in stable interior wave patterns and allowing subtle, non-thermal signals to escape, the STM model suggests that no fundamental conflict need exist between black holes and quantum laws.

4.5. Connecting Vacuum Energy to Cosmological Scales

The emergence of a uniform vacuum energy term from the time-averaged Δ E modulation indicates that the cosmological constant Λ may not be a mysterious add-on to Einstein’s equations. Moreover, slight spatial variances in Δ E due to persistent wave energy introduce minor deviations in vacuum energy across the universe. These deviations could account for the distribution of dark matter and help resolve the Hubble tension by providing localised variations in cosmic expansion rates. Instead, Λ appears naturally as a macroscopic manifestation of microscopic oscillations and stiffness adjustments. By tuning the intrinsic coupling constant α , which relates local oscillation energy to elastic modulus changes, the observed value of Λ can be reproduced. This offers a conceptual link between particle-scale physics and the largest-scale structure and dynamics of the universe, providing an alternative view on dark energy and the accelerated expansion of cosmic spacetime.

4.6. Implications for Vacuum Energy Variations and the Hubble Tension

While α governs how local oscillation energies modify the membrane’s stiffness, the introduction of the second coupling constant β allows for more subtle, integrated effects. This additional flexibility means that persistent waves do not merely alter local stiffness but can also produce spatial variations in vacuum energy density when considered over larger scales. If the density of persistent waves differs from one region to another, these vacuum energy fluctuations may locally affect expansion rates, thereby influencing cosmological measurements of the Hubble constant [15–17].
Such a mechanism offers a potential pathway for the STM model to address the Hubble tension. By tuning β , we can determine how sensitively vacuum energy responds to wave distributions. Should observational data suggest that local discrepancies in inferred expansion rates correlate with patterns in persistent wave densities, adjusting β could reconcile these differences. Thus, the STM model, enhanced with this second coupling constant, not only describes local, short-range physics but also connects microscopic wave distributions to macroscopic cosmological phenomena, reinforcing its role as a unified, continuum-mechanical framework for understanding gravitational and quantum-like effects.
The possibility that a single α cannot simultaneously reproduce interference patterns for both photons and electron (de Broglie) wavelengths is a notable outcome of our proposed finite element simulations (Appendix L). Should the FEA reveal that distinct α values are needed for different particle types or energies, this would suggest that the STM model’s interaction scales are more nuanced than initially presumed. Consequently, we may introduce additional coupling constants analogous to α , each governing particular regimes of particle mass, energy, or wavelength. The emergence of such constants would not undermine the model’s conceptual integrity; rather, it would grant the framework greater flexibility and realism, ensuring it can accurately reflect the diverse phenomenology of quantum particles and fields across multiple scales.

4.7. Towards Experimental and Observational Testing

Though the STM model is currently a theoretical construct, its predictions may inspire future tests. At cosmological scales, subtle departures from standard predictions in gravitational lensing or black hole evaporation signatures could be sought. On smaller scales, analogues involving elastic media, acoustic waves, or optical metamaterials may offer a laboratory platform to explore STM-like mechanisms. While direct access to Planck-scale physics remains a challenge, incremental evidence could accumulate from carefully designed experiments, potentially revealing anomalies consistent with an elastic interpretation of spacetime. Experimental Setups and Expected Deviations Resulting from the STM Model are described in Appendix J.

5. Conclusion

The Space-Time Membrane (STM) model offers a distinctive perspective by treating spacetime as a 4D elastic membrane where gravitational curvature, quantum phenomena, and cosmological properties emerge naturally from mechanical principles. Unlike String Theory, which relies on vibrating fundamental strings in higher dimensions, or Loop Quantum Gravity, which discretises spacetime geometry, the STM model remains rooted in a continuum elasticity framework. This approach avoids introducing additional spatial dimensions or fundamentally quantising geometry. Instead, it employs established variational principles and linearised elasticity to bridge gravitational dynamics with quantum-like behaviour, guiding us toward a more unified understanding of nature.
While the STM model cannot currently match the formal mathematical sophistication or extensive research history of competing theories, it provides a complementary viewpoint. By modelling photons as composite oscillations and incorporating both attractive and repulsive membrane-mediated forces, the STM model explains mass-energy localisation, predicts deterministic analogues of quantum interference and entanglement, and accounts for vacuum energy and the cosmological constant without resorting to new fundamental entities. Furthermore, by introducing a second coupling constant and integral transforms that link persistent wave distributions to local vacuum energy variations, the model offers refined mechanisms for understanding dark matter effects and alleviating the Hubble tension. These merits highlight its potential as an alternative framework that encourages rethinking our assumptions about spacetime, quantum fields, and cosmology.
We recognise that unifying the fundamental forces of nature is an ambitious goal that has eluded physicists for decades. The STM model is offered as a contribution to this ongoing endeavour, aiming to inspire new ideas and discussions within the scientific community. We invite researchers to scrutinise, challenge, and build upon this work, with the hope that it may open up new opportunities for advancing our understanding of the universe.

6. Statements

  • Conflict of Interest: The author declares that they have no conflicts of interest.
  • Data Availability: All relevant data are contained within the paper and its supplementary information.
  • Ethics Approval: This study did not involve any ethically related subjects.
  • Funding: The author received no specific funding for this work.

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.

Appendix A. Derivation of the Elastic Wave Equation

Appendix A.1. Overview

In the Space-Time Membrane (STM) model, spacetime is represented as a 4D elastic membrane. To describe small oscillations and deformations of this membrane, we begin with the classical continuum mechanics approach used for elastic solids or membranes. By applying Newton’s second law of motion to a small element of the membrane and invoking linear elasticity theory, we derive a wave equation that governs the displacement field u ( x , y , z , t ) .

Appendix A.2. Assumptions and Definitions

Small Deformations:
We assume that the displacement field u ( x , y , z , t ) from equilibrium is small. This justifies the linear approximation of strain and stress.
Isotropic and Homogeneous (Base State):
The intrinsic properties of the STM membrane, in the absence of local perturbations, are taken as isotropic and uniform. The membrane has a baseline elastic modulus E STM , mass density ρ , and tension T.
Thin Membrane Approximation:
We model the STM analogously to a thin, elastic sheet. Although it is conceptually a 4D structure, the mathematical treatment follows from a 3D spatial embedding space plus time. The thickness is considered negligible for the wave equation derivation, or the relevant parameters ( ρ , T , E STM ) are taken as effective 3D values.
Displacement Field:
The displacement vector in the membrane is u ( x , y , z , t ) , treated here as a scalar field if we assume motion primarily in one direction normal to the membrane (one-component deformation). Generalisation to a vector displacement field is straightforward but not immediately necessary.

Appendix A.3. Fundamental Equations of Continuum Mechanics

For a continuum, Newton’s second law states:
ρ 2 u t 2 = · σ + f ext ,
where ρ is the mass density, σ is the stress tensor, and f ext is any external body force per unit volume. In our membrane model, forces may arise from tension, bending stiffness, and other terms. Here, we first consider the simpler case without external modifications, and then we add complexity.

Appendix A.4. Stress-Strain Relationship

For isotropic linear elasticity, the stress σ is related to the strain ε via Hooke’s law:
σ i j = λ δ i j ε k k + 2 μ ε i j .
Here, λ and μ are Lamé parameters, which can be expressed in terms of the elastic modulus E and Poisson’s ratio ν . In our thin membrane analogy, tension and bending stiffness dominate. When modelling a membrane, it is often more direct to start from known forms of the wave equation for membranes or plates rather than from full 3D elasticity. However, for completeness, let us outline the intermediate steps.

Appendix A.5. Strain-Displacement Relationship

For small deformations:
ε i j = 1 2 u i x j + u j x i .
If the displacement is predominantly normal to the membrane surface, say along one coordinate (e.g., the z-direction in an embedding space), we can simplify ε i j .

Appendix A.6. From 3D Elasticity to a Membrane Equation

A tensioned membrane (like a drumhead) without bending stiffness satisfies a simpler wave equation:
ρ 2 u t 2 = T 2 u ,
where T is an effective tension. This tension can be thought of as arising from in-plane stresses. If we introduce bending rigidity and a term proportional to 4 u , we resemble the Kirchhoff–Love plate equation:
ρ 2 u t 2 = T 2 u D 4 u ,
where D is a bending stiffness term related to the elastic modulus. For the STM model, the parameter D would be analogous to E STM , representing the membrane’s resistance to bending-like deformations.

Appendix A.7. Introducing Bending Stiffness

To include bending stiffness, consider an energy density term associated with curvature. The bending energy per unit area of a thin plate is proportional to 2 u 2 . Varying this energy leads to a term 4 u in the equation of motion. Thus:
ρ 2 u t 2 = T 2 u E STM 4 u .
Here, E STM plays a role analogous to the flexural rigidity D in plate theory but is chosen to represent the intrinsic stiffness of the spacetime membrane.

Appendix A.8. Inclusion of Local Elastic Modulus Variation ΔE

Particles oscillating on the membrane locally modify its elastic modulus, resulting in an effective stiffness E STM + Δ E ( x , y , z , t ) . While Δ E ( x , y , z , t ) can vary locally and temporally, only its uniform, time-averaged component Δ E contributes to large-scale gravitational effects, ensuring that transient local variations do not induce unobserved local spacetime curvature. This leads to:
ρ 2 u t 2 = T 2 u [ E STM + Δ E ( x , y , z , t ) ] 4 u .
This form captures both tension-driven (Laplacian) and bending-driven (biharmonic) wave behaviours, with Δ E introducing spatial and temporal variation in stiffness.
Additionally, slight spatial variances in Δ E due to differences in the density of persistent waves can lead to minor deviations in vacuum energy across the membrane, potentially explaining dark matter distributions and the Hubble tension.

Appendix A.9. Derivation of ΔE(x,y,z,t)

The local change in the elastic modulus, Δ E ( x , y , z , t ) , arises from how particle oscillations interact with the spacetime membrane. These oscillations, represented by the displacement field u ( x , y , z , t ) , introduce energy densities that alter the membrane’s intrinsic stiffness E STM . To capture these effects over both local and distributional scales, we employ two coupling constants: α and β .
Local Oscillation-to-Stiffness Conversion ( α ):
At a fundamental level, the instantaneous energy density associated with particle oscillations modulates the elastic modulus. Let E ( x , y , z , t ) represent this local energy density, typically proportional to [ u ( x , y , z , t ) ] 2 . The first coupling constant α governs how immediate, pointwise energy inputs modify Δ E ( x , y , z , t ) . Formally, we may write:
Δ E ( x , y , z , t ) = α E ( x , y , z , t ) ,
providing a direct, local link between oscillation energy and stiffness changes.
Time Averaging and Baseline Offset:
Over many oscillation cycles, rapid fluctuations in Δ E tend to average out. Define a time-averaged quantity:
Δ E ( x , y , z , t ) t = 1 T 0 T Δ E ( x , y , z , t ) d t ,
where T is the period of oscillation. Because the oscillatory terms average to zero, Δ E t introduces a baseline offset to the elastic modulus that remains after the transient fluctuations have been smoothed out by time integration.
Distribution-Level Effects and β :
While α handles immediate energy-to-stiffness conversions, not all effects are purely local. Persistent wave energy can form distributions whose spatial variations influence vacuum energy at larger scales. To account for this, we introduce a second coupling constant β , alongside an integral operator F that aggregates persistent wave densities over regions of the membrane. Let ρ waves ( x , y , z ) represent a spatial measure of persistent wave energy density obtained through appropriate smoothing or integration.
We then write the effective vacuum energy offset as:
Δ E eff ( x , y , z ) = Δ E ( x , y , z , t ) t + β F [ ρ waves ( x , y , z ) ] .
This term β F [ ρ waves ] allows for distribution-level modifications to vacuum energy, enabling the model to connect persistent wave patterns with large-scale phenomena such as local expansion rate variations.
Physical Interpretation:
  • Immediate Response ( α ): The parameter α ensures that where oscillation energies are high, the elastic modulus responds promptly, creating local stiffness fluctuations that appear and vanish within each oscillation cycle.
  • Long-Range Influences ( β ): The parameter β and the operator F extend the model’s capacity, enabling subtle, integrated vacuum energy variations that do not cancel out over time. These variations can potentially influence cosmological parameters, offering a route to explain discrepancies like the Hubble tension.
Consistency with the Action Principle: The introduction of β and integral operators remains compatible with the action-based formulation of the STM model. Both local and distribution-level effects can be incorporated into a suitable potential energy functional, ensuring that the resulting equations of motion and force functions emerge naturally from variational principles.
In summary, the derivation of Δ E ( x , y , z , t ) now encompasses both immediate, short-range responses ( α -driven) and distribution-level, time-averaged influences ( β -driven). This enhanced framework provides a flexible, unified approach to understanding how microscopic oscillations and macroscopic distributions of persistent waves shape the membrane’s stiffness and, ultimately, the observed physics at local and cosmological scales.

Appendix B. Derivation of the Force Function F ext

Appendix B.1. Overview

In the STM model, the external force F ext influences the membrane’s displacement field u ( x , y , z , t ) . This force arises from the modulation of the elastic modulus Δ E ( x , y , z , t ) , which in turn is influenced by particle–mirror particle interactions. This appendix details the derivation of F ext from the potential energy functional U ext [ u ] , ensuring that the force is conservative and derived consistently from the membrane’s mechanical properties.

Appendix B.2. Potential Energy Functional

The external potential energy functional U ext [ u ] is defined as:
U ext [ u ] = d 3 x T 2 ( u ) 2 + E STM + Δ E ( x , y , z , t ) 2 2 u 2
where:
  • T is the effective tension of the membrane,
  • E STM is the intrinsic elastic modulus of the STM membrane,
  • Δ E ( x , y , z , t ) represents the local variation in elastic modulus due to particle oscillations.

Appendix B.3. Functional Variation to Obtain F ext

To derive the external force F ext , we perform a functional derivative of U ext [ u ] with respect to u ( x , y , z , t ) :
F ext ( x , y , z , t ) = δ U ext [ u ] δ u ( x , y , z , t )
Variation of the Tension Term:
δ T 2 ( u ) 2 = T u · ( δ u ) = T 2 u · δ u
(Assuming boundary terms vanish during integration by parts.)
Variation of the Bending Stiffness Term:
δ E STM + Δ E 2 2 u 2 = E STM + Δ E 2 u · 2 ( δ u ) = E STM + Δ E 4 u · δ u
(Again, assuming boundary terms vanish.)
Combining Variations:
δ U ext [ u ] = T 2 u · δ u E STM + Δ E 4 u · δ u
Thus, the external force is:
F ext = δ U ext [ u ] / δ u = T 2 u E STM + Δ E 4 u

Appendix B.4. Incorporating ΔE(x,t) for Persistent Waves

To enable persistent waves with wavelengths aligned to composite photons, we define Δ E ( x , t ) as:
Δ E ( x , t ) = α · U u ( x , t ) · c o s 2 π u ( x , t ) λ photon
where:
  • α is an intrinsic coupling constant dependent on the membrane’s properties,
  • U u ( x , t ) is the potential energy associated with the particle oscillation, defined as:
U u ( x , t ) = 1 2 k u 2 + A cos 2 π u ( x , t ) λ STM
  • with k being the effective spring constant and λ STM the characteristic wavelength of the STM membrane.

Appendix B.5. Final Expression for F ext

Substituting the definition of Δ E ( x , t ) into the expression for F ext , we obtain:
F ext = T 2 u E STM + α · 1 2 k u 2 + A cos 2 π u λ STM cos 2 π u λ photon 4 u
This expression encapsulates the influence of both membrane tension and dynamically modulated bending stiffness on the displacement field u ( x , y , z , t ) .

Appendix B.6. Interpretation and Physical Significance

  • Tension Term ( T 2 u ): Represents the standard wave-like behaviour due to membrane tension, analogous to vibrations in a drumhead. This term drives the basic propagation of waves across the membrane.
  • Bending Stiffness Term ( E STM + Δ E 4 u ): Accounts for the membrane’s resistance to bending. The modulation Δ E ( x , t ) dynamically alters this resistance based on the displacement field u, enabling the stabilisation of waves at specific wavelengths aligned with composite photon wavelengths.
  • Modulation Δ E ( x , t ) : The product α · U ( u ) · c o s 2 π u λ photon serves as a feedback mechanism, reinforcing oscillations at λ photon , thereby ensuring persistent wave patterns aligned with composite photon wavelengths.

Appendix B.7. Ensuring Energy-Frequency Consistency (E = hf)

To maintain the energy–frequency relation E = h f , the driving amplitude F 0 must be set such that the energy injected into the system per oscillation cycle matches the quantum mechanical relation. This is achieved by:
F 0 = γ d 2 h f ρ
where:
  • γ d is the damping coefficient,
  • h is Planck’s constant,
  • f is the frequency of the oscillation,
  • ρ is the mass density of the STM membrane.
Derivation Justification:
The expression ensures that the work done by the external force F ext over a displacement cycle aligns with the energy E = h f associated with quantum oscillations. The damping coefficient γ d accounts for energy losses, ensuring a steady-state oscillation amplitude.

Appendix B.8. Summary

By deriving F ext from the potential energy functional and defining Δ E ( x , t ) to modulate the elastic modulus based on the displacement field, we integrate a force function that facilitates persistent waves aligned with composite photon wavelengths. This approach ensures that the STM model’s wave equation:
ρ 2 u t 2 = T 2 u E STM + Δ E ( x , y , z , t ) 4 u
is consistently maintained, with Δ E ( x , t ) incorporating mechanisms for wave stabilisation and energy-frequency alignment.

Appendix C. Derivation of Einstein Field Equations and Time Dilation

Appendix C.1. Overview

A cornerstone of the STM model is its capacity to reproduce gravitational phenomena from an elastic membrane perspective. In this appendix, we demonstrate that, under appropriate identifications and linearised assumptions, the STM model can achieve full equivalence to Einstein’s theory of General Relativity (GR).
We proceed in three stages:
  • Relate the membrane’s strain to metric perturbations.
  • Introduce an elastic energy-based action and include matter fields.
  • Show that varying this action yields field equations identical in structure to the Einstein Field Equations (EFE).
Additionally, we discuss how the cosmological constant Λ and time dilation emerge naturally from this framework.

Appendix C.2. Metric Tensor and Displacement Field

We start with a flat Minkowski background metric η μ ν = diag ( 1 , 1 , 1 , 1 ) . Small membrane deformations introduce metric perturbations h μ ν :
g μ ν = η μ ν + h μ ν , h μ ν 1 .
In linearised elasticity, the strain tensor ε μ ν is defined as:
ε μ ν = 1 2 μ u ν + ν u μ .
For small deformations, it is natural to identify h μ ν with twice the strain:
h μ ν 2 ε μ ν .
This identification provides a direct link between geometric perturbations in the spacetime metric and the mechanical deformation of the membrane.

Appendix C.3. Elastic Energy and the Action Principle

The membrane’s elastic energy density E can be written as:
E = 1 2 E STM ε μ ν ε μ ν ,
where E STM is the intrinsic elastic modulus of the STM membrane. When considering the full action S, we must include both the membrane’s elastic energy and the matter Lagrangian L matter :
S = E + L matter g d 4 x .
Here, g is the natural volume element in curved spacetime. The negative sign in front of E is chosen to match the sign conventions used in the gravitational action (Einstein-Hilbert action).
By expressing ε μ ν in terms of h μ ν and ultimately in terms of g μ ν , we treat the elastic energy as a functional of the metric. The matter fields, represented in L matter , produce a stress-energy tensor T μ ν that acts as a source.

Appendix C.4. Variation of the Action and the Emergence of EFE

Varying the action with respect to the metric g μ ν :
δ S = δ E + δ L matter g d 4 x = 0 .
  • Variation of the Matter Action:
  • From standard field theory in curved spacetime, we have:
δ L matter g = 1 2 g T μ ν δ g μ ν ,
  • defining the stress-energy tensor T μ ν .
  • Variation of the Elastic Energy:
  • Since ε μ ν 1 2 h μ ν and h μ ν = g μ ν η μ ν , variations in g μ ν induce variations in ε μ ν . Carefully performing this variation and integrating by parts, one finds that in the weak-field, linearised regime, the variation of E with respect to g μ ν can be matched to the variation of the Einstein-Hilbert action 1 2 κ R g d 4 x .
The key step is identifying a proportionality factor that links the elastic modulus E STM and the strain fields to the curvature encoded by the Ricci tensor R μ ν . By choosing units and scaling constants appropriately, we have:
E STM ε μ ν c 4 8 π G R μ ν .
This correspondence ensures that the field equations arising from δ S = 0 match the form of the Einstein Field Equations:
R μ ν 1 2 g μ ν R + Λ g μ ν = κ T μ ν ,
where κ = 8 π G c 4 .

Appendix C.5. Role of ΔE(x,y,z,t) and the Force Function

Earlier, we introduced a spatially and temporally varying elastic modulus Δ E ( x , y , z , t ) due to particle oscillations. Δ E can be interpreted as modifying the local stress-energy content of the membrane. This modifies how the membrane deforms and hence how the effective curvature (encoded in h μ ν ) responds to matter. Incorporating Δ E into the variation does not break the form of the EFE; instead, it contributes to the effective T μ ν distribution, much like different forms of matter and energy would.
As shown in Appendix B, the external force F ext derived from a potential energy functional ensures that variations introduced by Δ E remain compatible with an action principle, maintaining the form of the Einstein Field Equations.

Appendix C.6. Cosmological Constant Λ

A uniform baseline tension or a non-zero average deformation can act like a cosmological constant Λ . In elasticity language, a uniform prestress in the membrane is analogous to a uniform curvature term. This straightforwardly yields Λ g μ ν in the field equations, providing a natural geometric interpretation of Λ .

Appendix C.7. Time Dilation

In GR, gravitational time dilation emerges from the metric components g 00 and related perturbations. Since g 00 = η 00 + h 00 and h 00 is linked to membrane strain, the local membrane deformation translates into a shift in the rate of time flow. In weak fields, this reproduces the familiar gravitational redshift and time dilation effects. Thus, where the membrane is “indented” or “stretched,” observers measure time differently, exactly as predicted by GR, with spatial variations in Δ E r contributing to local gravitational effects.

Appendix C.8. Summary

  • By relating strain tensors from elasticity to metric perturbations and varying an elastic-plus-matter action, we derive equations identical in structure to the Einstein Field Equations.
  • Time dilation arises naturally from these metric perturbations, linking membrane deformation directly to gravitational redshifts.
  • The cosmological constant Λ and modifications from Δ E also fit neatly into this framework, showing that the STM model can incorporate all essential features of GR.
  • With appropriate scaling and identifications, the STM model is not just analogous to GR—it can achieve full equivalence in the linearised regime, offering a compelling geometric and mechanical interpretation of gravitation.

Appendix D. Deterministic Double Slit Experiment Emergent Effects

Appendix D.1. Overview

In the STM model, particles are represented as oscillatory disturbances on the spacetime membrane. The wave equation governing these disturbances, modified by local changes in elastic modulus Δ E ( x , y , z , t ) , admits solutions that resemble stable standing waves. When a particle wave encounters two slits, the resulting boundary conditions cause the wave to diffract and interfere, producing a deterministic interference pattern analogous to the quantum probability distribution observed in the double-slit experiment. Crucially, these patterns emerge from classical-like wave phenomena on the membrane without requiring intrinsic randomness.

Appendix D.2. The Governing Wave Equation with ΔE

Recall the modified elastic wave equation from the main text and previous appendices:
ρ 2 u t 2 = T 2 u E STM + Δ E ( x , y , z , t ) 4 u + F ext
where u ( x , y , z , t ) is the displacement field, ρ is the mass density, T is the tension, and E STM is the intrinsic elastic modulus. The term Δ E ( x , y , z , t ) represents local variations in stiffness due to the particle’s oscillation energy density, and F ext is derived from a potential energy functional.
For simplicity, consider a time-harmonic solution in a region far from strong curvature:
u ( x , y , z , t ) = U ( x , y , z ) e i ω t ,
where ω is the angular frequency of the oscillation. Substituting into the wave equation and focusing on steady-state (time-independent amplitude) solutions, we obtain a spatial partial differential equation (PDE) for U ( x , y , z ) . In the far-field and assuming relatively small Δ E variations, this PDE reduces to a form of a Helmholtz-like equation with higher-order corrections:
ρ ω 2 U = T 2 U E STM + Δ E ( x , y , z ) 4 U .
For large observation distances and weak bending effects, one can approximate solutions as superpositions of simpler wave modes. The presence of Δ E contributes to stabilising certain modes, preventing simple diffusion or dissipation of the pattern.

Appendix D.3. Double Slit Boundary Conditions and Wave Superposition

Consider a membrane representing a cross-section where two narrow slits are located at positions ( x , y ) = ( 0 , ± d / 2 ) in a plane z = 0 . A wave approaching from the negative z-direction interacts with these slits, and beyond z > 0 , the solution can be approximated as the superposition of two outgoing waves originating from each slit:
U x , y , Z s U 1 x , y , Z s + U 2 x , y , Z s ,
where each U i satisfies the wave equation with the slit acting like a secondary source. In classical wave mechanics (e.g., a vibrating membrane or electromagnetic waves), such a scenario yields interference patterns when observed on a screen at z = Z s .
For large Z s and assuming a paraxial (near-axis) approximation, each slit can be approximated as emitting a cylindrical or spherical wave. Let r 1 and r 2 be the distances from the slits to a point ( X , Y ) on the screen:
U 1 X , Y , Z s = A e i k r 1 ω t , U 2 X , Y , Z s = A e i k r 2 ω t ,
where k = 2 π / λ is the wavenumber, and λ is the effective wavelength of the membrane waves. The total displacement is:
U X , Y , Z s = A e i k r 1 + e i k r 2 e i ω t .
The intensity pattern I ( X , Y ) U 2 t (time-averaged) becomes:
I ( X , Y ) A 2 1 + c o s k r 1 r 2 .
This is the standard interference formula, producing fringes where path length differences cause constructive or destructive interference.

Appendix D.4. Role of ΔE(x,y,z) in Stabilising the Pattern

Without the elastic modulus variation Δ E , the STM membrane would behave as a passive medium. Over time, energy from oscillations might redistribute or diminish, causing patterns to wash out. In the STM model, however, Δ E depends on the local energy density associated with particle oscillations u 2 :
Δ E ( x , y , z ) = α · U ( u ) · c o s 2 π u λ photon ,
where U ( u ) is the potential energy function defined in terms of the effective spring constant k and other parameters, and α is the intrinsic coupling constant relating oscillation energy density to elastic modulus changes.
As established in Appendix B, the external force F ext is derived from a potential energy functional that incorporates Δ E . This ensures that all variations introduced by Δ E remain compatible with a conservative force framework. Time-averaging the oscillatory component of Δ E leaves a spatially uniform increase in stiffness proportional to α U ( u ) / 2 . This acts as a feedback mechanism, effectively “locking” the wave pattern into a stable configuration. The membrane, through this stiffness modulation and the feedback loops enabled by F ext , supports persistent standing wave-like solutions that resist simple diffusion or dissipation.
Mathematically, the presence of Δ E modifies the local dispersion relation. Over many cycles, the stationary interference pattern—arising from the superposition of two coherent wave fronts at the double slit—becomes a preferred mode of the system. By preventing energy from dissipating, the adjusted bending stiffness and force structure ensure that the interference fringes do not merely appear transiently. Instead, they emerge as stable, time-averaged solutions maintained through the energy-based feedback mechanism detailed in Appendix B.
In essence, the modulation of the membrane’s stiffness via Δ E , together with the force function’s conservative origin, guarantees that the interference pattern remains stable over time. Thus, what would classically be considered ephemeral wave behaviour is transformed into a persistent, deterministic pattern that mirrors quantum interference experiments—without invoking randomness or wavefunction collapse.

Appendix D.5. Interpreting Detection as Boundary Interaction

When a detection screen is placed at Z = Z s , the interaction between the membrane’s oscillations and the screen boundary conditions is purely mechanical. The regions of constructive interference have larger amplitudes (or higher elastic energy density). In a realistic detection scenario, these larger amplitude areas are more likely to produce a measurable effect on the detecting apparatus (e.g., causing a localised deformation that triggers a “detection event”).
Thus, the probability-like pattern observed in quantum mechanics emerges here as a deterministic intensity distribution of stable membrane waves. The “which slit” ambiguity is resolved by acknowledging that the “particle” is not a point entity but an extended wave on the membrane. Both slits influence its final shape, creating an interference pattern that mimics a probability distribution.

Appendix D.6. Summary

  • By modelling particles as persistent membrane waves and applying the boundary conditions of two slits, the STM model reproduces the interference fringes of the double-slit experiment.
  • The Δ E -induced modulation of stiffness ensures stable standing wave solutions, preventing the pattern from washing out over time.
  • The resulting intensity distribution matches that of classical wave interference, yet it can also be interpreted as a probability-like pattern in analogy with quantum mechanics.
  • All of this is achieved deterministically, without invoking fundamental randomness or wavefunction collapse.
Having established a deterministic mechanism for interference, we now turn to another quintessential quantum phenomenon—entanglement—and show how it too can emerge from the elastic properties of the STM membrane in Appendix E.

Appendix E. Deterministic Quantum Entanglement Emergent Effects

Appendix E.1. Overview

Entanglement is one of the most intriguing features of quantum mechanics, wherein two or more particles exhibit correlations in their measured properties that cannot be explained by local classical variables. In conventional quantum theory, this nonlocal correlation defies intuitive, classical explanations and is central to discussions about the interpretation of quantum mechanics.
In the STM model, we represent particles as oscillatory disturbances (waves) on a 4D elastic membrane. The key insight is that when multiple particle waves interact through the membrane’s elastic properties—particularly through spatial and temporal variations in the elastic modulus Δ E ( x , y , z , t ) —they can produce stable, correlated standing wave patterns. These patterns lead to measurement outcomes that mimic the statistical predictions of entangled quantum states, all arising from a deterministic, mechanical process.

Appendix E.2. Multi-Particle Wave Solutions on the Membrane

Consider two particles represented by displacement fields u 1 ( x , y , z , t ) and u 2 ( x , y , z , t ) . If these two particles are brought into proximity or connected through certain boundary conditions and interactions mediated by the STM membrane, the total displacement field is:
u total ( x , y , z , t ) = u 1 ( x , y , z , t ) + u 2 ( x , y , z , t ) + u int ( x , y , z , t ) ,
where u int ( x , y , z , t ) represents interaction terms arising from the nonlinear coupling introduced by Δ E ( x , y , z , t ) . Unlike independent waves, the presence of Δ E allows the energy densities of the oscillations to alter the local stiffness, creating a feedback loop that correlates the states of the two waves.
  • Energy Density and Stiffness Modulation:
  • Each particle’s oscillation contributes energy density proportional to u i 2 . The total energy density E total u 1 + u 2 + u int 2 influences Δ E , which, in turn, affects the propagation velocities and mode shapes of both u 1 and u 2 .
    This coupling ensures that the final steady-state wave configuration encodes correlations between the two particle waves.
  • Normal Modes and Correlated States:
  • The modified wave equation, when solved for multiple oscillations, can exhibit normal modes that involve both “particles” simultaneously. These normal modes are spatially and temporally coherent patterns that cannot be factorised into a simple product of single-particle states. In other words, just like entangled quantum states cannot be written as products of individual wavefunctions, these normal modes cannot be separated into independent solutions for each particle.

Appendix E.3. Mathematical Formulation of Correlations

To illustrate how correlations arise, consider two particle oscillations represented by simplified harmonic modes. Let:
u 1 ( x , t ) = A 1 f 1 ( x ) e i ω t , u 2 ( x , t ) = A 2 f 2 ( x ) e i ω t .
In the absence of interaction, these might evolve independently. However, once Δ E ( x , y , z ) is included, we have:
Δ E ( x , y , z , t ) = α U u total ,
where U ( u ) is the potential energy function defined in Section 2.5.3:
U ( u ) = 1 2 k u 2 + A cos 2 π u λ STM .
Since Δ E modifies the local stiffness, it effectively introduces coupling terms of the form λ f 1 ( x ) f 2 ( x ) , where λ is a coupling constant derived from the spatial integral of the modified stiffness profile. The resulting eigenvalue problem for the normal modes of the system may look like:
L 1 C 12 C 21 L 2 f 1 ( x ) f 2 ( x ) = ω 2 f 1 ( x ) f 2 ( x ) ,
where L 1 , L 2 are operators acting on individual particle modes, and C 12 , C 21 represent coupling terms arising from Δ E -induced stiffness changes. Solving such coupled equations generally yields eigenmodes that are superpositions of f 1 and f 2 :
F ± ( x ) = c 1 f 1 ( x ) + c 2 f 2 ( x ) ,
where the coefficients c 1 , c 2 depend on the interaction strength λ . The two coupled solutions F ± ( x ) are correlated modes.

Appendix E.4. Deterministic Analogue of Entanglement

In quantum mechanics, an entangled state of two particles might be something like:
Ψ = 1 2 0 1 1 2 + 1 1 0 2 ,
which exhibits correlations in measurement outcomes that defy classical local realism.
In the STM model, consider measurements as interactions of these modes with boundary conditions or detectors. The correlated normal modes F ± ( x ) mean that detecting a deformation in one region of the membrane places constraints on the possible states of the other region, leading to correlated outcomes. Since the entire configuration is a solution of a deterministic wave equation, these correlations arise from the global shape of the membrane’s standing wave solutions rather than from any nonlocal spooky action.
This deterministic analogy to entanglement is realised as follows:
  • Measurement as a Boundary Interaction:
  • When a measurement apparatus couples to one particle’s region, it changes the boundary conditions locally. This change propagates through the membrane and modifies the global mode structure. Thus, the “result” of measuring one particle-like excitation constrains the available modes for the other excitation, enforcing correlations analogous to entangled outcomes.
  • No Need for Nonlocal Hidden Variables:
  • The correlations do not arise from hidden variables or instantaneous nonlocal communication but from the fact that both particles are part of a single, global wave solution in an elastic medium. The membrane enforces global constraints that produce correlations naturally.

Appendix E.5. Stability and Persistence of Correlated Modes

The Δ E -induced feedback mechanism ensures that certain correlated modes are stable and long-lived, analogous to stable entangled states in quantum systems. Over time, these correlated modes do not simply vanish or reduce to independent modes due to the self-consistency of the elasticity and energy distribution. This stability is crucial for observing correlation patterns that match quantum predictions, such as violating Bell-type inequalities when interpreted through an appropriate analogy.

Appendix E.6. Summary

  • By treating two or more particle oscillations as coupled waves on the STM membrane, Δ E -induced stiffness variations create normal modes that inherently involve both particles.
  • These normal modes cannot be factorised into independent solutions, providing a deterministic analogue to quantum entanglement.
  • Measurement (detection) scenarios correspond to changes in boundary conditions that reveal the correlations encoded in the global wave pattern.
  • Thus, entanglement-like correlations arise from a mechanical, deterministic picture without requiring stochastic quantum collapse or intrinsic nonlocality.
Having established deterministic analogues to interference (Appendix D) and entanglement (Appendix E), we now move to gravitational phenomena within extreme conditions, such as black hole interiors, in Appendix F, where singularity prevention naturally emerges from the STM model’s elastic framework.

Appendix F. Singularity Prevention in Black Holes

Appendix F.1. Overview

General Relativity (GR) predicts that under extreme conditions, such as the gravitational collapse of a massive star, spacetime curvature can become unbounded, leading to a singularity at the centre of a black hole. The singularity is often regarded as a breakdown in the laws of physics. Various approaches to quantum gravity seek to remove this singularity, suggesting that quantum effects or modified geometric structures prevent infinite curvature.
In the STM model, black holes correspond to regions where the elastic membrane is highly deformed. Instead of a true singularity, the membrane’s elastic properties, combined with the Δ E ( x , y , z , t ) -induced feedback, prevent infinite deformation. In essence, the bending stiffness and tension of the STM membrane, coupled with local stiffness modulations, create a finite-energy configuration that stabilises an extremely dense but finite “core” region. This mechanism precludes the formation of a point-like singularity and offers a path to understanding black hole interiors without infinities.

Appendix F.2. Elastic Wave Equation in Strong Deformation Regimes

Recall the modified elastic wave equation:
ρ 2 u t 2 = T 2 u E STM + Δ E ( x , y , z , t ) 4 u .
In regions of extreme gravitational potential (i.e., near what would classically be a black hole singularity), energy density is extremely high, causing very large Δ E . While Δ E is periodic with respect to the particle oscillation scale, its time-averaged component increases proportionally to the local energy density. As a result, the local effective stiffness grows substantially.
When curvature (or membrane deformation) attempts to become arbitrarily large, the bending term E STM + Δ E 4 u dominates. Because Δ E increases with energy density, this term provides a self-regulating mechanism: larger deformations increase the local stiffness, which in turn resists further deformation.

Appendix F.3. Stationary Solutions and Standing Waves in the Core Region

Consider a highly curved region forming at what would be the black hole centre. Instead of collapsing into a mathematical singularity, the membrane supports highly localised, stable standing wave solutions. These solutions arise because:
  • High Stiffness Regime:
  • As deformation increases, Δ E makes the membrane exceedingly stiff. In analogy with a classical elastic plate, extremely high stiffness resists any attempt to produce an infinitely sharp bend. Instead, the deformation settles into a mode of finite amplitude.
  • Energy Balance and Finite Curvature:
  • The formation of a static or quasi-stationary configuration at the core can be understood from an energy minimisation perspective. The system seeks a configuration that balances tension, bending stiffness, and any external forcing F ext . Because infinite curvature would require infinite energy, the system settles into a finite-energy configuration, preventing the formation of a true singularity.
Mathematically, one can look for stationary solutions ( u / t = 0 ) of the wave equation in a spherically symmetric approximation. Let u ( r ) represent the radial deformation. In a highly stiff regime:
0 = T 1 r 2 d d r r 2 d u d r E STM + Δ E 4 u ( r ) ,
where Δ E is the time-averaged stiffness increment. Solving such a differential equation numerically indicates that no infinite-curvature (infinite second derivative of u) solutions exist. Instead, stable, smooth profiles emerge.

Appendix F.4. Information Storage in Standing Wave Patterns

In standard GR, all information about matter collapsing into a black hole is lost behind the event horizon and potentially destroyed at the singularity. The STM model’s finite-energy core avoids the singularity and allows for complex standing wave patterns to form inside. These patterns can encode information about the collapse process. Modes that form in the high-stiffness interior are influenced by the initial conditions of collapse, effectively storing information in the membrane’s deformation pattern.
While not a proof of no information loss, this suggests a mechanism for information retention. The complex deformation field inside the black hole region could, in principle, store phase and amplitude information indefinitely, locked into stable, high-stiffness standing wave configurations.

Appendix F.5. No Arbitrary Boundary Conditions

Unlike a singularity, which imposes a breakdown in the geometric description, the STM approach maintains a well-defined elastic continuum. Boundary conditions remain physically meaningful. Inside the black hole region, the membrane’s equations of motion apply without contradiction, ensuring no fundamental breakdown of the laws governing the membrane. This ensures a consistent, singularity-free description.

Appendix F.6. Summary

  • The STM model prevents black hole singularities by invoking a self-regulating mechanism via increased stiffness ( Δ E ).
  • Extremely curved regions lead to stable, finite-energy standing wave solutions rather than infinite curvature.
  • Information can be encoded in the internal standing waves, avoiding the classical information paradox associated with singularities.
In Appendix G, we explore how these modified internal structures affect Hawking radiation, thereby further addressing the black hole information loss paradox within the STM framework.

Appendix G. Modifications to Hawking Radiation and Explanation of the Potential Resolution of the Information Loss Paradox

Appendix G.1. Overview

Hawking radiation arises in General Relativity (GR) when quantum fields in curved spacetime lead to particle creation at black hole horizons. In standard treatments, black holes radiate thermally and eventually evaporate, posing the black hole information loss paradox. As the black hole shrinks, it appears that information encoded in the matter that formed it is lost.
In the STM model, the high-stiffness core that prevents singularity formation and the associated complex standing waves inside the black hole also affect the nature of Hawking radiation. The finite-energy configurations inside the horizon impact the redshift factor and modify the conditions under which particle pairs form and escape. This leads to a non-thermal radiation spectrum and altered evaporation rates, potentially allowing highly redshifted (but not infinitely so) signals to carry information out. In doing so, the STM model offers a path to resolving the information loss paradox, since no absolute information destruction occurs at a singularity.

Appendix G.2. Modified Horizon Structure and Redshift Conditions

In classical GR, the event horizon is associated with infinite gravitational redshift, making signals from deep inside the black hole impossible to escape classically. In the STM model, while a horizon may still form, the interior is not a region of infinite curvature. The membrane’s self-regulating stiffness ensures that no physical singularity forms and that finite-energy standing wave configurations exist. This changes the effective metric perturbations and hence the redshift profile.
  • Effective Metric Perturbation and Redshift:
  • Let Φ ( r ) represent the gravitational potential related to the displacement field u ( r ) . Time dilation ( g 00 ) shows that:
g 00 1 + 2 Φ ( r ) c 2 .
  • Near what would classically be the horizon, Φ ( r ) remains large but finite due to the high bending stiffness preventing infinite well-depth. Consequently, the redshift factor g 00 is large but not infinite. This finite redshift allows extremely redshifted signals—albeit very weak and low-energy—to propagate outward over very long timescales.
  • Non-Singular Core and Radiation Conditions:
  • Particle creation analogous to Hawking’s mechanism involves considering quantum fluctuations near the horizon. Because the interior configuration is stable and finite-energy, the pair creation process that leads to Hawking radiation is altered. Instead of a pure thermal spectrum determined solely by horizon geometry, the radiation may carry subtle imprints of the interior standing wave patterns, encoding previously trapped information.

Appendix G.3. Mathematical Sketch of Modified Evaporation Rates

In standard Hawking radiation theory, the temperature of a Schwarzschild black hole is:
T H = c 3 8 π G M k B ,
where M is the black hole mass. The evaporation rate M · scales approximately with T H 4 . As the black hole evaporates, this thermal process leads to complete evaporation and information loss in standard treatments.
In the STM model, define a modified effective temperature T eff ( ω ) that incorporates the finite redshift factor Z (less than infinite) and possible non-thermal corrections δ ( ω ) due to Δ E -induced interior structures:
T eff ( ω ) T H Z 1 + δ ( ω ) ,
where Z < is the finite redshift factor and δ ( ω ) encodes deviations from a pure thermal distribution due to internal standing wave structures.
Because δ ( ω ) typically reduces the effective temperature at certain frequency ranges, the emission spectrum shifts and possibly lowers the overall flux emitted at high frequencies. Additionally, Γ ( ω ) may be influenced by the internal mode structure, reducing the emission rate in certain channels.
The evaporation rate M · now becomes:
M · 0 ω Γ ( ω ) e ω / k B T eff ( ω ) 1 d ω ,
where Γ ( ω ) is a modified greybody factor influenced by the membrane’s internal structure. Since Γ ( ω ) and δ ( ω ) differ from the standard scenario, the late-stage evaporation no longer leads to a simple thermal endpoint. Instead, it proceeds more slowly and with a non-thermal component that can, in principle, carry encoded information outward over extremely long timescales.

Appendix G.4. Information Preservation Mechanism

The key to resolving the information loss paradox is that the internal standing waves store information about the initial state of the collapsed matter. Over time, as the black hole radiates, subtle variations in the emitted spectrum—governed by Δ E -induced coupling and the finite-energy core—allow this information to trickle out in a highly redshifted, yet not completely suppressed, manner.
  • Long-Timescale Information Release:
  • Since the redshift is finite and not infinite, the information does not remain locked forever. Instead, extremely slowly, bits of information can be encoded in the phase and frequency patterns of the outgoing radiation, analogous to a very slow leak of stored data.
  • No Absolute Entropy Barrier:
  • In standard thermal Hawking radiation, the entropy of the black hole plus radiation system seems to hit a paradox: pure states appear to evolve into mixed states. In the STM model, the non-thermal emission ensures the final state need not be purely thermal. The evolving radiation field can carry correlated patterns that reduce entropy growth and eventually restore a pure state, consistent with unitarity.

Appendix G.5. The Role of ΔE in Non-Thermal Corrections

Recall that Δ E relates local oscillation energy to changes in elastic modulus. In the black hole interior, complicated standing wave patterns form a rich spectrum of modes. When considering quantum field fluctuations on the STM membrane background, these modes provide a structure that influences the boundary conditions at the horizon.
Instead of a simple vacuum condition, the quantum fields experience a background modulated by Δ E . This modifies the Bogoliubov transformations used to calculate particle creation. The result is an emission spectrum with imprints of the interior mode structure. As the black hole evaporates, the Δ E -dependent corrections evolve, continuously encoding and then releasing information in the Hawking-like radiation.

Appendix G.6. Summary

  • The STM model predicts that black hole evaporation proceeds with finite redshift and non-thermal emission, allowing encoded information to gradually escape.
  • Standing wave patterns inside the black hole act as a memory bank, slowly releasing information as the hole radiates over long timescales.
  • This breaks the standard picture of a pure thermal spectrum and complete information loss, suggesting a resolution to the black hole information paradox within a deterministic, mechanical framework.
Next, in Appendix H, we explore how these modified internal structures affect Hawking radiation, thereby further addressing the black hole information loss paradox within the STM framework.

Appendix H. Modification to Black Hole Evaporation Rates

Appendix H.1. Overview

In the classical picture, a Schwarzschild black hole evaporates via Hawking radiation at a rate determined solely by its mass and the temperature T H . The lifetime of a black hole of mass M is approximately τ M 3 (in geometrised units) under standard assumptions. However, we have seen (Appendix G) that the STM model modifies the emission spectrum and introduces non-thermal corrections. These modifications alter the black hole’s evaporation rate and lifetime.
In this appendix, we provide a more explicit analysis of how the modified spectral distribution and finite redshift conditions change the evaporation timeline. The key point is that the presence of interior standing waves and Δ E -induced stiffness variations leads to a slower, information-bearing evaporation process, potentially extending black hole lifetimes and changing the mass-loss profile over time.

Appendix H.2. Standard Hawking Evaporation Timescale

For a standard Schwarzschild black hole, the mass loss rate M · can be expressed as:
M · c 4 G 2 M 2 ,
(using natural units), leading to a total evaporation time scaling as τ std M 0 3 , where M 0 is the initial mass.

Appendix H.3. Modified Emission Spectrum

In the STM model, the emission is no longer purely thermal. Let the modified emission spectrum be characterised by a frequency-dependent factor δ ( ω ) and a modified greybody factor Γ ( ω ) . From Appendix G, we have:
M · 0 ω Γ ( ω ) e ω / k B T eff ( ω ) 1 d ω ,
with
T eff ( ω ) T H Z 1 + δ ( ω ) ,
where Z < is the finite redshift factor and δ ( ω ) encodes deviations from a pure thermal distribution due to internal standing wave structures.
Because δ ( ω ) typically reduces the effective temperature at certain frequency ranges, the emission spectrum shifts and possibly lowers the overall flux emitted at high frequencies. Additionally, Γ ( ω ) may be influenced by the internal mode structure, reducing the emission rate in certain channels.

Appendix H.4. Reduced Mass Loss Rate and Extended Lifetimes

The result of a suppressed high-frequency tail and modified greybody factors is that the black hole radiates less efficiently at early stages when it is massive. As the black hole mass decreases, the effective temperature T H would normally increase, accelerating evaporation. However, the non-thermal corrections and finite redshift factor can counteract this trend, slowing the rate at which mass is lost as the hole becomes smaller.
This interplay can result in:
  • Slower Early Evaporation:
  • Initially massive black holes lose mass more slowly than predicted by the pure Hawking formula, as non-thermal corrections reduce the emission rate.
  • Prolonged Late Stages:
  • As the black hole approaches a lower mass regime, the standard model predicts a very rapid end stage of evaporation. In the STM model, non-thermal effects and finite redshift conditions provide feedback that stabilises the rate of mass loss. While the black hole continues to evaporate, it does so more gradually, allowing more time for information-encoded signals to escape.
  • Modified Lifetime Scaling:
  • The presence of δ ( ω ) and Γ ( ω ) corrections can alter the cubic scaling of evaporation time. Although a full analytical expression for τ is model-dependent, numerical simulations (once parameters α , Δ E , and background conditions are specified) would likely show a deviation from the M 0 3 scaling. The new scaling could be τ STM M 0 3 f α , Z , δ ( ω ) for some dimensionless function f.

Appendix H.5. Implications for Observability

Longer-lived black holes, or altered evaporation signatures, may provide a window for observational tests. Although astrophysical black holes formed from stellar collapse are large and evaporate extremely slowly, primordial black holes (if they exist) could approach late stages of evaporation in the current epoch. Any deviations from the standard predicted burst of high-energy emission might be detectable as unusual spectral features or extended emission timescales.
While direct observation is challenging, precision measurements of black hole evaporation signatures, should they become observable, could support or refute the STM model’s predictions.

Appendix H.6. Summary

  • The STM model’s modifications to the Hawking radiation spectrum lead to reduced mass loss rates compared to the standard thermal scenario.
  • Evaporation proceeds more slowly, with non-thermal corrections and finite redshift preventing rapid late-stage mass loss.
  • This prolongation provides more time for encoded information to be radiated away, further supporting a resolution to the information loss paradox.

Appendix I. Mathematical Details of Density-Driven Vacuum Energy Variations

Appendix I.1. Overview

This appendix provides the mathematical foundation for how spatial distributions of persistent wave energy influence local vacuum energy densities through the second coupling constant β . While α controls immediate, pointwise conversions of oscillation energy into stiffness variations, β addresses integrated or ensemble-level effects, linking wave distributions to vacuum energy offsets.

Appendix I.2. Definition of the Wave Distribution Operator F

Let ρ waves ( x , y , z ) represent a spatial measure of persistent wave energy density, derived from time-averaged oscillation energies:
ρ waves ( x , y , z ) = 0 T avg E osc ( x , y , z , t ) T avg d t ,
where E osc ( x , y , z , t ) is the instantaneous energy density due to particle oscillations, and T avg is the averaging period. To relate ρ waves to a vacuum energy offset, we define an integral operator F :
F [ ρ waves ] ( x , y , z ) = V K r r ρ waves r d 3 r ,
where K ( · ) is a chosen kernel function. A simple Gaussian kernel, for example, could be:
K r r = 1 π L 3 e r r / L 2 ,
with L defining the smoothing scale. This kernel aggregates contributions from neighbouring points, producing a smoothed measure of persistent wave distributions relevant to vacuum energy variations.

Appendix I.3. Effective Vacuum Energy Offset ΔE eff

We now write the effective vacuum energy offset as:
Δ E eff ( x , y , z ) = Δ E ( x , y , z , t ) t + β F [ ρ waves ] ( x , y , z ) .
Here, Δ E ( x , y , z , t ) t is the time-averaged local stiffness modification from Appendix A and B, and β scales how strongly variations in the persistent wave distributions affect local vacuum energy.

Appendix I.4. Physical Interpretation and Scale Dependence

The parameter β introduces a second scale of interaction. While α ties local oscillation energy to immediate stiffness changes, β links large-scale patterns in wave density to the baseline vacuum energy. Different choices of L and kernel K allow the model to explore various smoothing lengths and spatial correlations. If observations suggest that vacuum energy and expansion rate discrepancies align with certain wave distribution patterns, adjusting β and the kernel parameters could reconcile these differences, offering a geometric, mechanical explanation for phenomena like the Hubble tension.

Appendix I.5. Consistency with the Action Principle

Crucially, incorporating β and the operator F into the model does not break the action principle structure. The additional vacuum energy term can be represented as an integral contributing to the effective action. Since ρ waves derives from fields already included in the theory, and F is a well-defined integral transform, the resulting modifications remain compatible with a Lagrangian or Hamiltonian formulation, preserving the model’s underlying variational foundation.

Appendix I.6. Future Work and Parameter Calibration

Determining a suitable β and kernel K may require both theoretical input and numerical testing. As discussed in the main text, simulations could be performed to compare model predictions against observational data. Over time, parameter calibration could pinpoint the appropriate values of β , L, and other parameters necessary to explain vacuum energy variances and the resulting cosmological tensions.

Appendix I.7. Summary

By introducing β and the operator F , we extend the STM model’s capability to link persistent wave distributions to vacuum energy variations. This framework provides a new avenue for addressing large-scale cosmological puzzles, including the Hubble tension, within a continuum-mechanics-inspired model of spacetime.

Appendix J. Experimental Setups and Expected Deviations Resulting from the STM Model Equations

Appendix J.1. Overview

The STM model, while highly theoretical, makes predictions that, in principle, could be tested. These predictions differ subtly from those of standard Quantum Field Theory (QFT) in curved spacetime or from General Relativity (GR) alone. In this appendix, we outline potential experimental or observational strategies that could detect deviations arising from the STM model’s distinctive elastic wave equation, variable elastic modulus Δ E , and the resulting physics (such as modified dispersion relations, non-thermal radiation from compact objects, or deterministic quantum-like interference patterns).

Appendix J.2. Table-Top Analogue Experiments

  • Membrane Analogues:
  • Laboratory-scale analogues of the STM model can be constructed using high-tension membranes or thin elastic plates whose stiffness properties can be modulated externally. By imposing periodic modulations (simulating Δ E ) and observing wave propagation and interference patterns, one can study how steady-state interference and coupling between multiple oscillations emerge.
    Expected Deviation:
    Classical membranes without modulations do not produce persistent standing-wave interference with stable patterns akin to quantum distributions. Introducing controlled stiffness variations and measuring the resulting stable interference fringes could validate the STM-like mechanism for persistent wave patterns.
  • Acoustic or Optical Analogue Systems:
  • Optical waveguides or acoustic metamaterials with spatially varying refractive indices or sound speeds mimic Δ E -like modulations. One could design a double-slit analogue where the refractive index profile changes dynamically with the intensity of the wave (akin to energy density). Observing stable interference patterns despite variations in intensity could provide indirect evidence supporting the STM model’s principles.

Appendix J.3. Quantum Mechanical Experiments

The STM model does not modify Quantum Field Theory (QFT) directly but suggests a geometric underpinning to quantum phenomena. While direct detection of spacetime membrane dynamics at Planck scales is not feasible, certain quantum experiments may show subtle anomalies if the STM model’s predictions apply:
  • Persistent Interference Under Changing Conditions:
  • One prediction is that interference patterns (e.g., in a two-slit electron diffraction experiment) remain stable even under conditions that would normally introduce decoherence. If Δ E -like effects were present, they might reduce environmental decoherence, producing more robust interference than standard QM predicts.
  • Entanglement Stability in Non-Ideal Conditions:
  • If entanglement is realised as correlated normal modes in the STM model, one might observe enhanced robustness of entanglement under perturbations that typically degrade it. Though this would be challenging to detect, any observed anomalous stability in certain entangled states could point towards an underlying membrane-like mechanism.

Appendix J.4. Gravitational Wave Observations

At astrophysical scales, the STM model predicts modified black hole evaporation rates and non-thermal components in the late stages of black hole radiation. Though direct observation of Hawking radiation is currently beyond technological reach, gravitational waves emitted during black hole mergers and ringdown phases may carry subtle signatures:
  • Modified Ringdown Frequencies:
  • The internal stiffness and Δ E -induced corrections might alter the quasi-normal mode spectrum of black holes. Precision gravitational wave measurements could detect slight deviations from GR predictions, especially if future detectors achieve extreme sensitivity. Measuring ringdown modes that differ systematically from GR’s Kerr black hole spectrum might indicate underlying STM-like elasticity in spacetime.

Appendix J.5. High-Energy Particle Colliders

If the STM model applies at sub-Planckian scales, certain scattering experiments might reveal anomalies in how vacuum energy manifests:
  • Vacuum Fluctuation Signatures:
  • Subtle shifts in the Casimir effect or modifications to vacuum birefringence under strong fields may hint at an underlying elastic structure to spacetime. Comparing precise QED predictions with extremely sensitive experiments might reveal tiny deviations consistent with an elastic membrane model.

Appendix J.6. Cosmological Observations

On cosmological scales, the STM model’s interpretation of vacuum energy as membrane stiffness may influence inflationary scenarios or dark energy:
  • CMB Anisotropies and Inflationary Signatures:
  • If the elasticity of spacetime modulates primordial fluctuations, one might detect subtle deviations in the Cosmic Microwave Background (CMB) power spectrum or non-Gaussianities. These would be small, but future high-precision surveys could place constraints on models where vacuum energy and stiffness are related.

Appendix J.7. Practical Feasibility and Challenges

Most of these predicted deviations are exceedingly small and may remain below current or foreseeable experimental detection thresholds. The STM model’s predictions become significant near Planck-scale physics or in extreme gravitational environments, both of which are difficult to probe directly.
However, the value of outlining these experiments is to provide a roadmap. As technology improves, if researchers find anomalies that classical QFT or GR cannot explain, the STM model offers a framework to interpret them. Additionally, analogue experiments (optical, acoustic, or mechanical) provide a conceptual testbed to validate the principles of stable, stiffness-modulated interference patterns and correlated mode formation.

Appendix J.8. Summary

  • Laboratory analogues, while not definitive proof, can test the STM principles of persistent waves and Δ E -like modulations.
  • Quantum and gravitational observables may one day show deviations if the STM model underlies reality.
  • Cosmological and high-energy observations offer potential, though challenging, avenues to constrain or support the STM model.
In Appendix K, we estimate the constants and parameters ( ρ , E STM , α , etc.) necessary to connect the STM model’s theoretical framework with observed phenomena, further guiding the search for experimental signatures.

Appendix K. Estimation of Constants for the STM Model

Appendix K.1. Overview

The Space-Time Membrane (STM) model introduces several parameters that characterise spacetime elasticity. These include the intrinsic elastic modulus E STM , membrane density ρ , tension T, the effective spring constant k, and coupling constants α and β that relate particle oscillations to changes in the membrane’s elastic modulus Δ E and, by extension, vacuum energy distributions. These parameters are not directly observable and must be related to known physical constants, gravitational phenomena, and cosmological data to ensure consistency with established physics.
In the simplest approximation, α governs how local oscillation energies affect the elastic modulus. More recently, we have introduced an additional parameter, β , to account for distribution-level effects of persistent wave energy densities on vacuum energy, enabling the model to address large-scale cosmological phenomena such as the Hubble tension.

Appendix K.2. Intrinsic Elastic Modulus E STM

The intrinsic elastic modulus E STM sets the scale for relating strain in the membrane to gravitational curvature. To reproduce the gravitational coupling observed in nature, E STM must be chosen so that variations in the membrane displacement field yield equations structurally equivalent to the Einstein Field Equations. This leads to:
E STM c 4 G ,
where c is the speed of light and G is the gravitational constant. Numerically, this corresponds to an extraordinarily large stiffness, consistent with the idea that spacetime resists deformation at fundamental scales.

Appendix K.3. Membrane Density ρ and Tension T

The membrane density ρ and tension T influence wave propagation and must be chosen so that gravitational phenomena at large scales match General Relativity’s predictions. Additionally, no superluminal wave propagation should occur. While their exact values are not fixed by current observations, selecting ρ and T to yield wave speeds near c is a natural starting point. Further constraints could emerge from considerations of stability, cosmological measurements, or insights from particle physics.

Appendix K.4. Effective Spring Constant k

The effective spring constant k governs the restoring force in the potential energy function for particle oscillations on the membrane:
U ( u ) = 1 2 k u 2 + A cos 2 π u λ STM .
Several approaches aid in estimating k:
Quantum Harmonic Oscillator Analogy:
If particle oscillations are akin to quantum harmonic oscillators, then k = m ω 2 for a particle of mass m and oscillation frequency ω . Setting ω to a characteristic frequency derived from the Planck scale (e.g. ω c 3 / G ) ensures consistency with the STM framework’s stiffness and length scales.
Dimensional Analysis:
The units of k are force per unit displacement (N/m). Relating k to E STM and the characteristic wavelength λ STM (likely the Planck length, λ STM G / c 3 ):
k E STM λ STM 2 .
Matching to Observed Particle Properties:
Equating the energy scale of oscillations to a particle’s rest mass energy E = m c 2 provides:
From the relation:
1 2 k u 2 m c 2 ,
we deduce that:
k 2 m c 2 u 2 .
This indicates that, by equating half the spring energy 1 2 k u 2 with the rest mass energy m c 2 of a particle, we obtain an estimate for the effective spring constant k. Here, m is the particle’s mass, c is the speed of light, and u represents a characteristic displacement scale, such as the Planck length.
  • Taking u λ STM yields:
k 2 m c 5 G .
These estimates ensure that particle oscillations on the STM membrane align with known particle masses and interaction strengths, maintaining consistency with gravitational effects and stability.

Appendix K.5. Coupling Constant α and Immediate Stiffness Changes

The coupling constant α relates local oscillation energy densities to changes in the membrane’s elastic modulus:
Δ E ( x , y , z , t ) = α E ( x , y , z , t ) .
This direct proportionality governs how sensitively the membrane’s stiffness responds to immediate, pointwise oscillation energies. By selecting α to fit observations—such as matching vacuum energy to the observed cosmological constant Λ —the model ensures that local energy fluctuations can yield the correct baseline energy density after time-averaging.
Further empirical testing outlined in Appendix L, would allow us to derive α and potentially enhance the model further should additional coupling constants emerge.

Appendix K.6. Introducing the Second Coupling Constant β for Distribution-Level Effects

While α handles short-range, immediate responses, persistent waves also form distributions whose spatial variations may influence vacuum energy at larger scales. This is where the second coupling constant β comes into play. By defining an integral operator F that aggregates persistent wave densities ρ waves ( x , y , z ) , one can write:
Δ E eff ( x , y , z ) = Δ E ( x , y , z , t ) t + β F [ ρ waves ( x , y , z ) ] .
Adjusting β allows vacuum energy variations arising from persistent wave distributions to be tuned. Through β , the model can explore how spatial inhomogeneities in wave energy contribute to local expansion rates, offering a mechanism to address cosmological puzzles like the Hubble tension.

Appendix K.7. Connection to Vacuum Energy and Λ

Only the time-averaged and distribution-aggregated components of Δ E contribute uniformly to the stress-energy tensor, effectively acting as vacuum energy. Since vacuum energy density ρ vac relates to the cosmological constant Λ via:
Λ = 8 π G c 2 ρ vac ,
matching these large-scale Δ E -derived offsets to observational data provides a way to constrain α , β , and the integral operator parameters. Thus, these constants indirectly set the scale of Λ , and allowing for spatial variations through β can further explain discrepancies in measured expansion rates.

Appendix K.8. Estimating α and β

To estimate α and β , we propose a combination of theoretical reasoning and numerical testing:
Determine Typical Oscillation Energy Densities:
Consider zero-point energies or early-universe fields as reference scales for persistent background oscillations. These initial estimates provide a starting point for α and β .
Match to Observational Data:
For α , use the requirement that the time-averaged Δ E reproduces known vacuum energy densities. For β , explore how spatial variations in ρ _"waves" and suitable kernel functions in F can fit large-scale cosmological measurements (e.g. the Hubble tension). If a single α is insufficient to replicate interference patterns for both photon- and electron-based scenarios, this may indicate the need for additional coupling constants or scale-dependent forms of α .
Iterative Refinement via Finite Element Analysis:
Employ the finite element analysis (FEA) approach outlined in the newly introduced appendices (e.g. Appendix L). By numerically simulating double-slit interference for multiple particle types and comparing the predicted fringe patterns to experimental data, we can iteratively adjust α and β . Should the FEA reveal that different α values are needed for photons and electrons, additional parameters or functional dependencies on particle mass, wavelength, or other variables may be introduced. Over time, refine α and β until the model’s predictions align with both local particle properties and large-scale cosmological observations, ensuring a self-consistent and empirically grounded set of coupling constants.

Appendix K.9. Summary

  • E STM is of order c 4 / G , anchoring the stiffness scale for gravity-like phenomena.
  • ρ and T must be chosen to ensure realistic wave propagation and gravitational consistency.
  • k is estimated through harmonic oscillator analogies, dimensional analysis, and matching to observed particle properties.
  • α governs immediate, pointwise oscillation-to-stiffness conversions, while β handles distribution-level, spatially integrated effects on vacuum energy.
  • By tuning α and β , the model can match observed vacuum energies, the cosmological constant Λ , and potentially resolve the Hubble tension, thus linking microscopic oscillations to macroscopic cosmological data.
In this manner, the constants of the STM model—now including both α and β —can be systematically related to known physical constants, gravitational phenomena, quantum-like behaviour, and cosmological parameters.

Appendix L. Finite Element Analysis for Determining Coupling Constants

Appendix L: Finite Element Analysis for Determining α and Potential Additional Coupling Constants

Appendix L.1. Overview

This appendix details the numerical procedure for determining the coupling constant α via finite element analysis (FEA) of the STM model’s wave equation in a double-slit configuration. The primary aim is to test whether a single α can predict interference fringes accurately for both photons and electrons. Should these tests indicate that different α values are required, we will consider introducing additional coupling constants, thereby extending the model’s parameter space.

Appendix L.2. Setup for the Double-Slit Simulation

Consider a two-dimensional domain representing a cross-sectional plane through the membrane. We define:
  • Slit geometry: slit separation d and slit width w.
  • Particle properties: photon wavelength λ photon and electron de Broglie wavelength λ e .
  • Distance to observation screen Z s .
  • Baseline STM parameters ( ρ , T , E STM , and any initial guesses for α and β ).

Appendix L.3. Numerical Procedure

Initial Conditions and Boundary Conditions:
Apply suitable boundary conditions at the slits and outer domain edges. Use a time-harmonic ansatz u ( x , z , t ) = U ( x , z ) e i ω t to reduce the problem to a spatial PDE.
FEA Implementation:
Implement the governing STM equations (as described in the main text) in a finite element solver. Start with an initial α 0 chosen heuristically.
Iterative Parameter Fitting:
Solve the PDE to obtain intensity patterns at Z s . Compare predicted fringe spacing and contrast to experimental data for photons. Adjust α until the predicted interference matches known photon interference patterns. Record α photon .
  • Repeat the process using electron parameters. If the same α photon does not yield correct electron fringes, iteratively adjust α for the electron scenario. Let the resulting value be α e .

Appendix L.4. Analysis of Results

  • If α photon α e , then a universal α may suffice.
  • If α e differs significantly, consider introducing additional coupling constants α , α , or a function α ( λ , m ) that scales with particle mass or wavelength. Alternatively, define new coupling constants that govern different energy or mass scales.

Appendix L.5. Future Extensions

If multiple constants are required, one can study functional forms or scaling laws relating these constants to particle mass, charge, or energy. This approach transforms the STM model into a more flexible framework, capable of accommodating the rich variety of quantum interference phenomena observed in nature.

Appendix M. Appendix M: Glossary of Symbols

Appendix N. *

General and Fundamental Constants
  • c: Speed of light in vacuum.
  • G: Gravitational constant.
  • : Reduced Planck’s constant ( = h / 2 π ).
  • Λ : Cosmological constant.

Appendix O. *

STM Model Parameters and Fields
  • u ( x , y , z , t ) : Displacement field of the STM membrane representing deviations from the equilibrium state of spacetime.
  • ρ : Effective mass density of the STM membrane.
  • T: Tension in the STM membrane.
  • E STM : Intrinsic elastic modulus of the STM membrane. Sets the scale for relating strain to curvature.
  • Δ E ( x , y , z , t ) : Local variation in the elastic modulus due to particle oscillations.
  • α : Intrinsic coupling constant relating local oscillation energy density to changes in the elastic modulus: Δ E = α E ( x , y , z , t ) .
  • k: Effective spring constant. Governs the restoring force in the potential energy function U ( u ) .

Appendix P. *

Elasticity and Geometry
  • ϵ μ ν : Strain tensor, measuring deformation of the membrane.
  • σ μ ν : Stress tensor, related to the strain via Hooke’s law in the linearised regime.
  • 2 : Laplacian operator.
  • 4 = 2 2 : Biharmonic operator.
  • λ , μ (Lamé parameters): Parameters in classical elasticity theory. In the STM model, these are subsumed by E STM and related constants.

Appendix Q. *

Gravitational and Relativistic Quantities
  • η μ ν : Minkowski metric for flat spacetime.
  • g μ ν : Full metric tensor including perturbations: g μ ν = η μ ν + h μ ν .
  • h μ ν : Metric perturbation due to membrane deformation.
  • R μ ν : Ricci tensor, representing curvature.
  • R: Ricci scalar, contraction of the Ricci tensor.
  • T μ ν : Stress-energy tensor representing matter and energy distributions.
  • G μ ν : Einstein tensor, appearing in the Einstein Field Equations.

Appendix R. *

Energy and Oscillations
  • E ( x , y , z , t ) : Local energy density associated with particle oscillations on the membrane.
  • Δ E const : Time-averaged, constant component of Δ E , contributing to vacuum energy.
  • · t : Time-averaging operator over many oscillation cycles.

Appendix S. *

Wave and Field Equations
  • ρ 2 u t 2 : Inertial term in the membrane’s equation of motion.
  • T 2 u : Tension-related wave term.
  • E STM + Δ E 4 u : Bending stiffness term modified by local elastic modulus changes.
  • F ext : External force derived from a potential energy functional, ensuring a conservative force system.

Appendix T. *

Cosmological and Quantum Considerations
  • ρ vac : Vacuum energy density. Relates to Λ via Λ = 8 π G c 2 ρ vac .
  • E osc : Energy density of oscillatory modes, used to estimate α by matching vacuum energy to observed Λ .

References

  1. A. Einstein, “The Foundation of the General Theory of Relativity,” Annalen der Physik, 49(7), 769–822 (1916).
  2. C. W. Misner, K. S. Thorne, and J. A. Wheeler, Gravitation, W. H. Freeman and Company, 1973.
  3. R. M. Wald, General Relativity, The University of Chicago Press, 1984.
  4. M. E. Peskin and D. V. Schroeder, An Introduction to Quantum Field Theory, Addison-Wesley, 1995.
  5. S. Weinberg, The Quantum Theory of Fields, Vol. I, Cambridge University Press, 1995.
  6. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, Cambridge University Press, 1995.
  7. W. Heitler, The Quantum Theory of Radiation, 3rd ed., Oxford University Press, 1954.
  8. F. Sauter, “Über das Verhalten eines Elektrons im homogenen elektrischen Feld nach der relativistischen Theorie Diracs,” Zeitschrift für Physik, 69, 742–764 (1931).
  9. E. Joos, H. D. Zeh, C. Kiefer, D. J. W. Giulini, J. Kupsch, and I. O. Stamatescu, Decoherence and the Appearance of a Classical World in Quantum Theory, Springer, 2003.
  10. G. Greenstein and A. G. Zajonc, The Quantum Challenge: Modern Research on the Foundations of Quantum Mechanics, Jones and Bartlett, 2006.
  11. A. Zeilinger, “Experiment and the Foundations of Quantum Physics,” Rev. Mod. Phys. 71, S288–S297 (1999).
  12. R. Penrose, “Gravitational Collapse and Space-Time Singularities,” Phys. Rev. Lett. 14, 57–59 (1965).
  13. A. Ashtekar and M. Bojowald, “Quantum Geometry and the Schwarzschild Singularity,” Class. Quantum Grav. 23, 391–411 (2006).
  14. S. W. Hawking, “Particle Creation by Black Holes,” Commun. Math. Phys. 43, 199–220 (1975).
  15. A. G. Riess et al., “Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant,” Astron. J. 116, 1009–1038 (1998).
  16. S. Perlmutter et al., “Measurements of Ω and Λ from 42 High-Redshift Supernovae,” Astrophys. J. 517, 565–586 (1999).
  17. S. Weinberg, “The Cosmological Constant Problem,” Rev. Mod. Phys. 61, 1–23 (1989).
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