Codex Totalis Icon The Caldwell Legacy | Codex Totalis Printer Friendly PDF Printer Friendly (PDF)

The High-Dimensional Phase Orbiter (HDPO) Model:

Mathematical Formalism and Quantitative Predictions

Arthur Caldwell, Ph.D.1, Alani Sharma, Ph.D.2, and Elias Martel, Ph.D.¹

¹Institute for Advanced Theoretical Studies, New Chicago Quantum Dynamics Lab

²Department of Computational Mathematics, New Chicago University

*acaldwell@iats.edu.nc

August 2325

Abstract

This paper provides the detailed mathematical formalism that underpins the conceptual framework of the High-Dimensional Phase Orbiter (HDPO) model, a deterministic theory introduced in a previous work [1]. We address the key open questions of the initial proposal by: (1) providing a formal mathematical construction for the Governing Principle of Minimal Information-Action; (2) explicitly constructing and solving a minimal model for the 1D Quantum Harmonic Oscillator (QHO), including the derivation of the required projection map \(\pi\) that reproduces the Gaussian probability distribution; and (3) using this solved minimal model to calculate quantitative, experimentally testable predictions for the wavefunction collapse timescale and for deviations from the Born rule at attosecond resolutions. Finally, we outline a formal pathway for developing a fully Lorentz-covariant formulation and demonstrating how the postulate of a compact manifold provides a natural regularization for Quantum Field Theory.

1. Introduction

In a preceding paper [1], we introduced the conceptual framework for the High-Dimensional Phase Orbiter (HDPO) model, a deterministic sub-quantum theory from which the phenomena of Quantum Field Theory (QFT) are proposed to emerge as statistical approximations. While that work outlined the core postulates and philosophical underpinnings of the model, it deliberately left several critical mathematical constructions and quantitative calculations as open research challenges.

The purpose of this document is to address those challenges directly. We provide the rigorous mathematical derivations that elevate the HDPO model from a conceptual framework to a quantitative, falsifiable scientific theory. Our objective is twofold: first, to demonstrate the internal mathematical consistency of the model's core postulates; and second, to utilize this formalism to derive concrete, experimentally testable predictions that distinguish the HDPO model from standard quantum theory.

This work will proceed as follows. In Section 2, we provide a formal definition for the Governing Principle of Minimal Information-Action. In Section 3, we construct and explicitly solve a minimal model for the 1D Quantum Harmonic Oscillator, deriving the required projection map. In Section 4, we use this solved model to calculate numerical predictions for wavefunction collapse and Born rule deviations. Finally, in Section 5, we outline a path toward a fully Lorentz-covariant formulation and discuss the model's inherent regularization of QFT. Through these steps, we will demonstrate that the HDPO model has successfully transitioned from a conceptual framework to a quantitative, testable scientific program.

2. Formalization of the Governing Principle

The foundational axiom of the HDPO model is the Principle of Minimal Information-Action, which posits that the laws of physics are emergent consequences of a deeper optimization principle. In [1], this was presented conceptually. Here, we provide a formal mathematical construction for the information functional \(\mathcal{I}\) whose minimization is conjectured to determine the structure of physical reality.

2.1 The Information Functional \(\mathcal{I}\)

We propose that the functional \(\mathcal{I}\) is composed of two primary terms: a complexity term, \(K(\mathcal{M})\), which quantifies the information required to specify the geometric and topological structure of the hidden manifold \(\mathcal{M}\), and an entropy term, \(H(S)\), which quantifies the information content of the physical dynamics that unfold upon that manifold.

\[ \mathcal{I}[\mathcal{M}, H] = K(\mathcal{M}) + H(S) \tag{2.1} \]

The universe realizes the specific manifold \(\mathcal{M}\) and Hamiltonian \(H\) that together minimize this functional.

2.2 The Complexity Term \(K(\mathcal{M})\)

The complexity term, \(K(\mathcal{M})\), represents the information required to specify the geometric "stage" itself. A simple, highly symmetric manifold is informationally "cheaper" than a complex, baroque one. While a formal definition of algorithmic complexity for a continuous manifold is an open problem, a natural measure for geometric complexity is the total action of the geometry itself. We therefore propose that \(K(\mathcal{M})\) is proportional to the integrated curvature of the manifold, a formulation directly inspired by the Einstein-Hilbert action from General Relativity:

\[ K(\mathcal{M}) = \frac{1}{16\pi G_M} \int_{\mathcal{M}} R \, dV \tag{2.2} \]

where \(R\) is the Ricci scalar curvature of the hidden manifold \(\mathcal{M}\) and \(G_M\) is a fundamental constant analogous to the gravitational constant, setting the scale of the manifold's intrinsic "stiffness." This term is minimized for manifolds that are maximally "flat" or symmetric (e.g., spheres, tori).

2.3 The Entropy Term \(H(S)\)

The entropy term, \(H(S)\), represents the "cost" of describing the physical laws and the set of all possible histories that can occur. We define this term using the Shannon entropy formalism applied to the path integral of standard QFT. Let \(S[\phi]\) be the classical action for a given field configuration \(\phi(x)\). The probability measure on the space of all possible field configurations is given by the Feynman path integral:

\[ P[\phi] = \frac{1}{Z} e^{iS[\phi]/\hbar} \tag{2.3} \]

where \(Z\) is the partition function, \(Z = \int \mathcal{D}[\phi] \, e^{iS[\phi]/\hbar}\). The informational entropy of the dynamics is then the expectation value of the information content, \(-\log P[\phi]\):

\[ H(S) = - \int \mathcal{D}[\phi] \, P[\phi] \log P[\phi] \label{eq:entropy_S} \tag{2.4} \]

This term is minimized when the dynamics are simple, symmetric, and predictable.

2.4 The Variational Principle

The final step is to apply the variational principle, \(\delta\mathcal{I} = 0\), to the combined functional from Eq. (2.1). This principle creates a fundamental tension: a manifold that is too simple (low \(K(\mathcal{M})\)) may require very complex physical laws to reproduce observed phenomena (high \(H(S)\)). Conversely, a very complex manifold might allow for simpler physical laws. The observed laws of nature, including the specific symmetries of the Standard Model and the geometry of the HDPO manifold, are conjectured to be the result of the unique solution that finds the optimal balance between these two competing costs.

While an analytical solution to this variational problem is likely intractable, it is a well-defined problem amenable to modern computational methods. The "forward problem" methodology is predicated on using numerical techniques such as Geometric Monte Carlo to search the discretized space of possible manifold geometries and Hamiltonians for the configuration that minimizes \(\mathcal{I}\). This transforms the origin of physical law from a philosophical question into a computationally intensive but ultimately answerable one.

3. A Solved Minimal Model: The 1D Quantum Harmonic Oscillator

The "forward problem" methodology requires the construction of explicit, calculable models for simple physical systems. To demonstrate the viability of the HDPO framework, we present here a solved minimal model for the ground state of the one-dimensional Quantum Harmonic Oscillator (QHO). The QHO is an ideal test case due to its ubiquity in physics and its well-understood properties, including its Gaussian ground state probability distribution.

3.1 The Problem Statement

The objective is to define a minimal HDPO system—a compact manifold \(\mathcal{M}\), a Hamiltonian \(H\), and a projection map \(\pi\)—that precisely reproduces the known physics of the QHO ground state. Per the "minimal model" approach outlined in [1], we postulate the simplest possible structures that satisfy the necessary constraints.

  • The Manifold (\(\mathcal{M}\)): We postulate a 2-torus, \(\mathcal{M} = \mathcal{T}^2 = S^1 \times S^1\), parameterized by angular coordinates \((\theta_1, \theta_2)\). Its compactness ensures the stability of the resonant mode.
  • The Hamiltonian (\(H\)): The Hamiltonian for the hidden motion, \(H_{\text{internal}}\), must be chosen such that its energy matches the QHO's zero-point energy, \(E_0 = \frac{1}{2}\hbar\omega\). For uniform, ergodic motion on the torus, a simple Hamiltonian is of the form \(H_{\text{internal}} = \frac{p_1^2}{2I_1} + \frac{p_2^2}{2I_2}\), where the moments of inertia \(I_1, I_2\) are set to produce the correct energy.

The most significant challenge, and the focus of this section, is the construction of the projection map, \(\pi\).

3.2 The Geometric Constraint: Preservation of the Fisher Information Metric

The projection map \(\pi: \mathcal{T}^2 \to \mathbb{R}\) cannot be arbitrary. It must connect the uniform probability distribution on the hidden manifold (due to ergodic motion) to the specific, non-uniform Gaussian probability distribution of the QHO ground state in observable space, \(P(x) \propto e^{-m\omega x^2/\hbar}\).

We propose that the map is determined by a profound geometric constraint: it must be an information-preserving map. Specifically, it must preserve the Fisher information metric, which is a way of measuring the "distance" between probability distributions. This ensures that the information about the system's state is not lost during the projection. This constraint transforms the problem of finding \(\pi\) into a well-defined problem in differential geometry: solving a specific partial differential equation for the map \(x(\theta_1, \theta_2)\).

3.3 Derivation of the Projection Map from the Geometric Constraint

The geometric constraint established in Section 3.2—that the map \(\pi\) must preserve the Fisher information metric—imposes a powerful condition on the function \(x(\theta_1, \theta_2)\). The transformation of the probability measure from the uniform distribution on the torus, \(P_{\mathcal{T}^2}(\theta_1, \theta_2) = 1/(4\pi^2)\), to the Gaussian distribution in the observable space, \(P_{\text{QHO}}(x)\), requires that the area elements be related by the probability density:

\[ P_{\mathcal{T}^2} \, d\theta_1 d\theta_2 = P_{\text{QHO}}(x) \, dx \tag{3.1} \]

Since \(P_{\mathcal{T}^2}\) is a constant, this simplifies to \(C \cdot d\theta_1 d\theta_2 = e^{-m\omega x^2/\hbar} dx\), where \(C\) is a normalization constant. The differential \(dx\) can be expressed in terms of the partial derivatives of the map \(x(\theta_1, \theta_2)\) and the differentials \(d\theta_1, d\theta_2\). A full treatment requires the Jacobian of the transformation. This leads to the following non-linear, second-order partial differential equation for the projection map \(x(\theta_1, \theta_2)\):

\[ \left( \frac{\partial^2}{\partial \theta_1^2} + \frac{\partial^2}{\partial \theta_2^2} \right) x - \frac{2m\omega}{\hbar} x \left( \left(\frac{\partial x}{\partial \theta_1}\right)^2 + \left(\frac{\partial x}{\partial \theta_2}\right)^2 \right) = 0 \label{eq:pde_map} \tag{3.2} \]

This equation, a variant of the sinh-Gordon equation, is a formidable challenge. While a general closed-form solution is not known, we have found that a separable solution of the form \(x(\theta_1, \theta_2) = f(\theta_1)g(\theta_2)\) exists.

3.4 The Elliptic Function Solution

The substitution of a separable solution into Eq. (3.2) reveals that the functions \(f\) and \(g\) must satisfy the differential equation for the Jacobi elliptic functions. We report the specific solution here.

A simple projection like \(x = \cos(\theta_1)\) is known to produce an arcsine distribution and is therefore incorrect. The correct map that solves Eq. (3.2) and transforms the area element of the torus into the Gaussian-weighted area element of the line is a product of the sine amplitude (sn) and cosine amplitude (cn) elliptic functions:

\[ x(\theta_1, \theta_2) = A \cdot \text{sn}(\theta_1; k) \cdot \text{cn}(\theta_2; k) \label{eq:projection_map_final} \tag{3.3} \]

where \(A\) is a normalization constant. The crucial parameter is the elliptic modulus, \(k\), which determines the "shape" of the elliptic functions. For the projection to yield a perfect Gaussian distribution, the modulus \(k\) cannot be arbitrary. We find that it must be precisely determined by the physical parameters of the QHO system itself, relating the periods of the elliptic functions to the ratio of the hidden motion's frequency to the observable oscillator frequency, \(\omega\).

The existence of this elegant, analytic solution for the projection map is a stunning confirmation of the HDPO model's internal consistency. It demonstrates that the probabilistic rules of quantum mechanics can, in principle, emerge from a deeper, deterministic geometric structure.

3.5 Excited States and the Resonance Condition

The minimal model presented thus far has focused exclusively on reproducing the QHO ground state. A complete theory must also account for the full discrete spectrum of excited states, \(E_n = \hbar\omega(n + 1/2)\) for \(n \in \{1, 2, 3, ...\}\).

Within the HDPO framework, these excited states are not different types of motion, but rather higher harmonics of the fundamental resonant trajectory on the hidden torus. The quantization condition, first introduced conceptually in [1], requires that any stable, stationary state must correspond to a trajectory that forms a standing wave on the compact manifold. This means the total phase accumulated over one full period must be an integer multiple of \(2\pi\).

For the 2-torus, this condition applies to each independent cycle. We propose that the energy of a stable resonant mode is determined by two integer quantum numbers, \((n_1, n_2)\), corresponding to the number of wavelengths that fit along each dimension of the torus:

\[ E_{n_1, n_2} = \hbar\omega_1 (n_1 + 1/2) + \hbar\omega_2 (n_2 + 1/2) \tag{3.4} \]

where \(\omega_1\) and \(\omega_2\) are the fundamental frequencies of motion along the \(\theta_1\) and \(\theta_2\) coordinates, and the \(1/2\) terms represent the zero-point energy of the ground state.

The observed energy levels of the 1D QHO are then recovered by a specific projection and degeneracy condition. We find that for the specific projection map \(\pi\) derived, all states with the same sum \(n = n_1 + n_2\) are degenerate and project to the same observable energy level, \(E_n\). This naturally reproduces the known energy spectrum of the QHO. The Hermite polynomial shape of the excited state wavefunctions is likewise reproduced by the more complex projection of these higher-harmonic trajectories. A full derivation of these projections is reserved for a future work, but the mechanism is clear: the discrete energy ladder of quantum mechanics is a direct reflection of the allowed discrete harmonics of a hidden, deterministic motion.

4. Quantitative Predictions from the QHO Minimal Model

With a fully defined and solved minimal model for the 1D Quantum Harmonic Oscillator, we can now move beyond the qualitative predictions of [1] and derive quantitative, experimentally testable results. The model, being fully deterministic, allows for the direct numerical simulation of processes that are treated as instantaneous and probabilistic in standard quantum theory. We focus here on two key predictions that clearly distinguish the HDPO model from the Copenhagen interpretation.

4.1 Derivation of the Wavefunction Collapse Timescale

Simulation Methodology:

We model a strong position measurement by adding a steep quadratic potential term, \(V(x) = \frac{1}{2} k_{\text{meas}}(x-x_0)^2\), to the system's Hamiltonian at \(t=0\). This term, representing the interaction with a detector, perturbs the dynamics on the hidden 2-torus. We then numerically integrate the equations of motion for the trajectory on the torus under this new Hamiltonian and track the time-evolution of the projected, observable position \(x(t) = \pi(\theta_1(t), \theta_2(t))\).

Results:

Our simulations show that the trajectory is pulled from its wide, space-filling orbit and spirals into a new, stable limit cycle corresponding to the measurement outcome \(x \approx x_0\). The projected probability distribution, \(P(x,t)\), evolves smoothly from its initial Gaussian profile to a final, sharply peaked distribution. A representative evolution is shown in Figure 1.

Simulated time-evolution of the projected probability distribution P(x, t) for a QHO ground state during a strong position measurement.
Figure 1: Simulated time-evolution of the projected probability distribution \(P(x, t)\) for a QHO ground state during a strong position measurement. The initial Gaussian profile (\(t = 0\)) is shown to narrow into a final, localized state over the calculated finite duration, \(\tau_{\text{collapse}}\).

The time it takes for the system's variance to contract to within 1% of the final state's variance is defined as the collapse time, \(\tau_{\text{collapse}}\). We find this time is a calculable function of the oscillator's natural frequency, \(\omega\), and the measurement strength, \(k_{\text{meas}}\). For a strong measurement, we derive the quantitative prediction:

\[ \tau_{\text{collapse}} \approx \frac{2\pi}{\omega} \tag{4.1} \]

This finite, non-zero duration is a direct, falsifiable prediction. An attosecond pump-probe spectroscopy experiment should be able to resolve this transient, "mid-collapse" state.

Concrete Experimental Target:

For a typical quantum dot system used in spectroscopy, with a characteristic oscillation frequency of \(\omega \approx 10^{14}\) Hz, our model predicts a collapse timescale of \(\tau_{\text{collapse}} \approx 2\pi / \omega \approx 60\) femtoseconds. While challenging, this duration is well within the resolution of modern attosecond pump-probe laser systems. The primary experimental challenge will be to distinguish the deterministic evolution of the probability distribution from environmental decoherence. We propose that by performing the experiment in a cryogenically cooled, high-vacuum environment, the environmental decoherence time can be extended significantly beyond the predicted collapse time, allowing for a clear measurement of the intrinsic HDPO collapse process.

4.2 Derivation of Born Rule Deviations

The Born rule, in the HDPO model, is an emergent statistical law that arises from time-averaging the hidden trajectory over a measurement time, \(T_{\text{meas}}\), that is long compared to the trajectory's characteristic orbital period on the manifold, \(T_{\text{orb}}\). If a measurement can be performed fast enough (\(T_{\text{meas}} \ll T_{\text{orb}}\)), this approximation breaks down.

Simulation Methodology:

Using our solved QHO model, we can calculate the period of the hidden motion, which is determined by its energy \(E_0 = \frac{1}{2}\hbar\omega\). This gives \(T_{\text{orb}} = h/E_0 = 4\pi/\omega\). We then simulate an ensemble of measurements, each with a very short duration \(T_{\text{meas}}\), and construct the resulting probability distribution.

Results:

When \(T_{\text{meas}} \ll T_{\text{orb}}\), the apparatus only samples a small arc of the total hidden orbit. The resulting probability distribution is not a smooth Gaussian. Instead, as shown in Figure 2, it appears "lumpy" or structured, reflecting the specific regions of the manifold that were being transited during the brief measurement windows.

The magnitude of the root-mean-square (RMS) deviation from the Born rule's prediction, \(\Delta P\), is predicted to scale inversely with the measurement time. To observe a 1% deviation (\(\Delta P/P \approx 0.01\)) from the standard quantum prediction, an experiment would need to achieve a temporal resolution of:

\[ T_{\text{meas}} \approx 0.01 \cdot T_{\text{orb}} = 0.01 \cdot \frac{4\pi}{\omega} \approx \frac{0.04\pi}{\omega} \tag{4.2} \]
Simulated probability distribution showing lumpy deviations from the smooth Born rule prediction.
Figure 2: Simulated probability distribution for a QHO measurement performed with a temporal resolution of \(T_{\text{meas}} = 0.01 \cdot T_{\text{orb}}\). The resulting distribution (solid line) exhibits significant structured deviations from the smooth, time-averaged Gaussian profile predicted by the Born rule (dashed line).

This provides a concrete, quantitative target for experimentalists in the field of attosecond physics. A confirmed observation of such structured, time-dependent probability distributions would provide powerful evidence for an underlying, time-evolving hidden reality.

Concrete Experimental Target:

Using the same quantum dot system (\(\omega \approx 10^{14}\) Hz), the hidden orbital period is \(T_{\text{orb}} = 4\pi/\omega \approx 120\) fs. To observe a 1% deviation from the Born rule, a measurement with a temporal resolution of \(T_{\text{meas}} \approx 0.01 \cdot T_{\text{orb}} \approx 1.2\) fs is required. This is an extremely demanding but achievable target for state-of-the-art attosecond physics. The primary challenge will be accumulating sufficient statistics from these ultra-short measurements to construct a high-fidelity probability distribution. We propose that high-repetition-rate laser systems and advanced statistical filtering techniques will be necessary to distinguish the predicted "lumpy" structure from shot noise and detector inefficiencies.

5. Towards a Relativistic Formulation and Full QFT

The minimal model presented for the non-relativistic QHO is a crucial proof of principle. However, a complete foundation for modern physics requires a fully Lorentz-covariant formulation and a direct engagement with the machinery of Quantum Field Theory, including gauge symmetries and regularization. This section outlines the proposed pathway for this extension.

5.1 The Problem of Lorentz Covariance

A significant challenge for any hidden-variable theory is reconciling its dynamics, which often seem to rely on a preferred reference frame or absolute time parameter, with the principles of Special Relativity. The time parameter \(t\) used to describe the evolution of the hidden trajectory \(\Phi(t)\) appears to be such a preferred parameter.

We propose that this is a misinterpretation arising from the non-relativistic model. We conjecture that the hidden manifold \(\mathcal{M}\) is not a Riemannian manifold, but is in fact a Lorentzian manifold with a signature of \((+, -, -, -, ...)\). The hidden trajectory \(\Phi(t)\) is always a light-like (null) geodesic on this higher-dimensional manifold. We posit that the observed Lorentz invariance of our 3+1 dimensional spacetime is an emergent symmetry, a consequence of the projection map \(\pi\) from this fundamentally Lorentzian hidden reality. The proper time of the hidden trajectory is always zero; the "evolution" we perceive is a feature of the projection.

To ensure compatibility with observed relativistic causality, the projection map \(\pi\) must satisfy a crucial causal projection constraint. While the reconfiguration of the state \(\Phi(t)\) on the manifold is instantaneous, the expectation value of any observable operator \(\hat{O}\) in our spacetime must evolve in a way that respects the light cone. We propose that the projection of the system's commutator for two spacelike separated operators, \(\hat{O}_1(x_1)\) and \(\hat{O}_2(x_2)\), must be zero.

\[ \pi \left( [\hat{O}_1(\Phi(x_1)), \hat{O}_2(\Phi(x_2))] \right) = 0, \quad \text{for } (x_1 - x_2)^2 < 0 \label{eq:no_signaling} \tag{5.1} \]

This condition, known as microcausality, ensures that although the hidden variables are non-locally correlated, these correlations cannot be used to transmit information faster than light. The statistical outcomes of measurements at \(x_1\) and \(x_2\) will show the "spooky" correlations of entanglement, but the expectation value of any measurement at \(x_2\) cannot be controllably influenced by a choice of measurement at \(x_1\). The structure of the projection map \(\pi\) thus acts as a "causal filter," laundering the underlying non-locality to produce a causally well-behaved observable reality.

5.2 Inherent Regularization and Gauge Symmetries

As established in Postulate 6 [1], the compactness of \(\mathcal{M}\) provides a natural, non-perturbative ultraviolet (UV) cutoff, eliminating the infinities that plague standard QFT. A one-loop Feynman diagram, which would normally require regularization and renormalization, becomes a finite integral over a compact momentum subspace, \(\mathcal{K}\), defined by the geometry of \(\mathcal{M}\). For example, a divergent loop integral in SQT is replaced by:

\[ \int^{\infty} d^4k \, f(k) \quad \xrightarrow{\text{HDPO}} \quad \int_{\mathcal{K} \subset \mathcal{M}} d\mu(k) \, f(k) < \infty \tag{5.2} \]

This transforms divergent quantities into calculable, finite values.

Furthermore, the gauge symmetries of the Standard Model are conjectured to arise from the geometric symmetries (the isometry groups) of the hidden manifold itself. As a proof of concept, the U(1) gauge symmetry of electromagnetism can be shown to emerge from a simple \(S^1\) rotational symmetry in a subspace of the manifold. Reproducing the SU(2) and SU(3) symmetries of the electroweak and strong forces is a primary objective of the ongoing "forward problem" research program.

6. Conclusion

We have provided the rigorous mathematical formalism that was absent in the initial conceptual presentation of the High-Dimensional Phase Orbiter model. We have demonstrated the model's internal consistency by explicitly constructing and solving a minimal model for the 1D Quantum Harmonic Oscillator, including a derivation of the non-trivial projection map required to reproduce the correct quantum statistics.

Using this solved model, we have calculated concrete, quantitative, and falsifiable predictions for the duration of wavefunction collapse and for deviations from the Born rule at experimentally accessible attosecond timescales. This work elevates the HDPO model from a philosophical interpretation to a testable, scientific research program. The path forward—towards a full relativistic formulation and the reproduction of the Standard Model—is monumentally challenging, but the results presented herein provide a solid foundation and a clear methodology for that endeavor.

🖨 Download Printer Friendly (PDF) Version

References

  1. [1] A. Caldwell, A. Sharma, and E. Martel, "The High-Dimensional Phase Orbiter (HDPO) Model: A Deterministic and Geometric Foundation for Quantum Field Theory," Preprint Archive, Inst. for Adv. Theo. Studies, August 2325.
  2. [2] A. Einstein, B. Podolsky, and N. Rosen, "Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?" Physical Review, vol. 47, no. 10, pp. 777–780, 1935.
  3. [3] J. S. Bell, "On the Einstein Podolsky Rosen Paradox," Physics Physique Fizika, vol. 1, no. 3, pp. 195–200, 1964.
  4. [4] S. L. Adler, Quantum Theory as an Emergent Phenomenon. Cambridge University Press, 2004.
  5. [5] J. Li, "On the Emergence of Physical Law from Computationally Irreducible Systems," Journal of AI Metaphysics, vol. 14, no. 2, pp. 210-245, 2188.
  6. [6] E. Petrova, An Introduction to Sub-Quantum Kinematics. Cambridge University Press, 2275.
  7. [7] H. Vance, "Cognitive Throughput as a Bottleneck in Augmented Reality Overlays for Complex System Management,” Journal of Posthuman Engineering, vol. 34, no. 1, pp. 88-102, 2305.
  8. [8] L. Kowalski, "Heuristic Stability and Predictive Error Correction in Non-Linear, Complex Adaptive Systems," OmniResource Corporate Technical Journal, vol. 88, no. 4, pp. 602-618, 2323.
  9. [9] K. Shaw, "A Critical Analysis of the Thermodynamic Limits and Material Dependencies of Atmospheric Fabrication," Journal of Applied Industrial Physics, vol. 102, no. 1, pp. 45-58, 2324.
  10. [10] F. Krausz and M. Ivanov, "Attosecond physics,” Reviews of Modern Physics, vol. 81, no. 1, pp. 163-234, 2009.