A Multi-Framework Analysis of the Effective Potential Transforming Single-Slit to Double-Slit Diffraction Statistics
Part I: Foundational Analysis of Slit-Experiment Wave Mechanics
The bedrock of any analysis concerning quantum interference lies in a precise mathematical description of the phenomena. Before delving into the speculative realms of quantum potentials, classical analogues, and emergent gravity, it is imperative to establish the exact wavefunctions and their corresponding probability densities for both the single-slit and double-slit experiments. This section will derive these foundational quantities under the Fraunhofer (far-field) approximation, providing the mathematical basis—the initial and final states of our transformation—for the entirety of the subsequent report. The rigor applied here is not merely procedural; it is essential, as the subtle mathematical distinctions between these two scenarios are precisely what the sought-after "force" or "potential" must account for.
Section 1.1: The Wavefunction in Fraunhofer Diffraction
In the Fraunhofer regime, the distance to the observation screen is assumed to be much larger than the dimensions of the diffracting aperture, allowing for the approximation that waves arriving at any given point on the screen are effectively parallel.[1] This simplifies the analysis considerably, enabling the use of Fourier optics to relate the aperture function to the far-field diffraction pattern.
The Single-Slit Wavefunction
We begin by modeling a single slit of width $a$ centered at the origin of a coordinate system. The aperture can be described by a simple rectangular function, where the wavefunction immediately after the slit, $\psi_{aperture}(y')$, is constant for positions $|y'| \le a/2$ and zero elsewhere. This represents a uniform illumination of the slit by a coherent, monochromatic plane wave of wavelength $\lambda$.[2]
The far-field wavefunction, $\Psi_S(y)$, observed on a screen at a large distance $L$, is given by the Fourier transform of this aperture function. The position $y$ on the screen corresponds to a diffraction angle $\theta$, where for small angles, $\theta \approx y/L$. The path difference between a wavelet from the center of the slit and one from a point $y'$ within the slit is $y' \sin\theta$. This introduces a phase factor of $\exp(-ik(y' \sin\theta))$, where $k = 2\pi/\lambda$ is the wavenumber. The total amplitude at angle $\theta$ is the coherent sum (integral) of all these wavelets across the aperture.[3, 4]
The derivation proceeds by evaluating the following integral:
$$ \Psi_S(\theta) \propto \int_{-a/2}^{+a/2} e^{-ik y' \sin\theta} \, dy' $$This is a standard Fourier transform of a rectangular function. The integration yields:
$$ \Psi_S(\theta) \propto \left[ \frac{e^{-ik y' \sin\theta}}{-ik \sin\theta} \right]_{-a/2}^{+a/2} = \frac{e^{-ika \sin\theta / 2} - e^{+ika \sin\theta / 2}}{-ik \sin\theta} $$Using Euler's formula, $e^{i\phi} - e^{-i\phi} = 2i \sin\phi$, this simplifies to:
$$ \Psi_S(\theta) \propto \frac{2i \sin(ka \sin\theta / 2)}{ik \sin\theta} = a \frac{\sin(ka \sin\theta / 2)}{ka \sin\theta / 2} $$To standardize this expression, we introduce the dimensionless variable $\beta$, which encapsulates the experimental geometry and wavelength:
$$ \beta = \frac{ka \sin\theta}{2} = \frac{\pi a \sin\theta}{\lambda} $$The wavefunction for the single-slit experiment can then be written in its canonical form, using the normalized sinc function, where $\text{sinc}(x) = \sin(x)/x$:
$$ \Psi_S(\theta) = A_0 \text{sinc}(\beta) $$Here, $A_0$ is a normalization constant representing the amplitude at the central maximum ($\theta=0$, $\beta=0$).[2, 4, 5]
The Double-Slit Wavefunction
The analysis for the double-slit experiment builds directly upon the single-slit case. We model the apparatus as two identical slits of width $a$, with their centers separated by a distance $d$. The aperture function is now a superposition of two rectangular functions. The total wavefunction at the screen, $\Psi_D(\theta)$, is the coherent sum of the wavefunctions from each slit.[6, 7]
Let the two slits be centered at $y' = +d/2$ and $y' = -d/2$. The wavefunction from the slit at $+d/2$ is simply the single-slit wavefunction with a phase shift corresponding to its offset from the origin, and similarly for the slit at $-d/2$. The total wavefunction is:
$$ \Psi_D(\theta) = \Psi_{\text{slit at }-d/2}(\theta) + \Psi_{\text{slit at }+d/2}(\theta) $$The Fourier transform now involves an integral over two disjoint regions. A more elegant approach is to use the shift theorem of Fourier transforms. The transform of a function shifted by $c$ is the original transform multiplied by a phase factor $e^{-ikc\sin\theta}$. Therefore, the total wavefunction is the sum of two phase-shifted single-slit wavefunctions:
$$ \Psi_D(\theta) \propto \text{sinc}(\beta) \left( e^{-ik(-d/2)\sin\theta} + e^{-ik(d/2)\sin\theta} \right) $$ $$ \Psi_D(\theta) \propto \text{sinc}(\beta) \left( e^{ikd\sin\theta/2} + e^{-ikd\sin\theta/2} \right) $$Using Euler's formula again, $e^{i\phi} + e^{-i\phi} = 2 \cos\phi$, this becomes:
$$ \Psi_D(\theta) \propto \text{sinc}(\beta) \cos\left(\frac{kd\sin\theta}{2}\right) $$We define a second dimensionless variable, $\gamma$, to represent the interference term:
$$ \gamma = \frac{kd\sin\theta}{2} = \frac{\pi d \sin\theta}{\lambda} $$The final form of the double-slit wavefunction, normalized to the same single-slit amplitude $A_0$, is:
$$ \Psi_D(\theta) = 2 A_0 \text{sinc}(\beta) \cos(\gamma) $$This elegant result shows that the double-slit wavefunction is composed of two parts: a rapidly oscillating interference term, $\cos(\gamma)$, which depends on the slit separation $d$, modulated by a slowly varying diffraction envelope, $\text{sinc}(\beta)$, which is identical to the single-slit pattern and depends only on the slit width $a$.[6, 7, 8]
Section 1.2: Probability Densities and Intensity Patterns
The connection between the abstract wavefunction and the observable pattern on a detector screen is made through the Born rule, which states that the probability density, $P(\theta)$, of detecting a particle (in this case, a photon) at a given location is proportional to the modulus squared of the wavefunction at that point: $P(\theta) \propto |\Psi(\theta)|^2$.[9] In optics, this probability density is synonymous with the classical intensity of the light pattern.
Single-Slit Intensity
For the single-slit experiment, the probability density is:
$$ P_S(\theta) = |\Psi_S(\theta)|^2 = |A_0|^2 [\text{sinc}(\beta)]^2 $$Letting $I_S = |A_0|^2$ be the intensity at the central maximum, we have:
$$ P_S(\theta) = I_S \left[ \frac{\sin(\beta)}{\beta} \right]^2 \quad \text{where} \quad \beta = \frac{\pi a \sin\theta}{\lambda} $$This function describes the characteristic single-slit diffraction pattern: a broad, bright central maximum flanked by a series of progressively dimmer secondary maxima.[10, 11] The minima (points of zero intensity) occur when $\sin(\beta) = 0$ but $\beta \ne 0$. This condition translates to $\beta = m\pi$ for any non-zero integer $m$, which gives the well-known condition for destructive interference in a single slit [12, 13]:
$$ a \sin\theta = m\lambda, \quad m = \pm 1, \pm 2, \pm 3, \dots \quad (\text{Minima}) $$The secondary maxima are located approximately halfway between the minima, at $\beta \approx (m+1/2)\pi$.[14]
Double-Slit Intensity
For the double-slit experiment, the probability density is:
$$ P_D(\theta) = |\Psi_D(\theta)|^2 = |2 A_0|^2 [\text{sinc}(\beta)]^2 [\cos(\gamma)]^2 $$Letting $I_D = 4|A_0|^2 = 4I_S$ be the intensity at the central interference peak (where $\beta=0$ and $\gamma=0$), the expression becomes:
$$ P_D(\theta) = I_D \left[ \frac{\sin(\beta)}{\beta} \right]^2 \cos^2(\gamma) \quad \text{where} \quad \beta = \frac{\pi a \sin\theta}{\lambda}, \quad \gamma = \frac{\pi d \sin\theta}{\lambda} $$This equation mathematically captures the visual appearance of the double-slit pattern: a set of fine interference fringes described by the $\cos^2(\gamma)$ term, whose overall intensity is modulated by the single-slit diffraction envelope described by the $[\text{sinc}(\beta)]^2$ term.[8, 15, 16, 17]
The key features are:
1. Interference Maxima (Bright Fringes): Occur when $\cos^2(\gamma) = 1$, which means $\gamma = m\pi$ for integer $m$. This gives the condition for constructive interference [18, 19, 20]:
$$ d \sin\theta = m\lambda, \quad m = 0, \pm 1, \pm 2, \dots \quad (\text{Maxima}) $$2. Interference Minima (Dark Fringes): Occur when $\cos^2(\gamma) = 0$, which means $\gamma = (m+1/2)\pi$. This gives the condition for destructive interference [19]:
$$ d \sin\theta = \left(m + \frac{1}{2}\right)\lambda, \quad m = 0, \pm 1, \pm 2, \dots \quad (\text{Minima}) $$3. Diffraction Minima ("Missing Orders"): The overall pattern has zero intensity wherever the diffraction envelope is zero, i.e., at the same locations as the single-slit minima ($a \sin\theta = m'\lambda$). If an interference maximum is predicted to occur at the same angle as a diffraction minimum, it will be "missing" from the pattern.[16]
The central challenge of this report is now mathematically precise: we must find a potential or equivalent construct that can transform the probability distribution $P_S(\theta)$ into $P_D(\theta)$. The core of this transformation involves generating the highly oscillatory $\cos^2(\gamma)$ factor from the smooth, single-peaked structure of the single-slit pattern. This immediately suggests that any such "force" cannot be a simple, monolithic deflecting agent. Instead, it must be a complex, spatially varying field that imposes a fine structure onto the probability distribution, creating regions of high probability (constructive interference) and low probability (destructive interference). The spatial frequency of this potential must be directly related to the parameter $\gamma$, and thus to the slit separation $d$ and wavelength $\lambda$. This fundamental requirement will serve as a guiding principle and a benchmark for evaluating the validity of the potentials derived in the subsequent parts of this analysis.
| Property | Single-Slit Experiment (width $a$) | Double-Slit Experiment (width $a$, separation $d$) |
|---|---|---|
| Wavefunction $\Psi(\theta)$ | $A_0 \text{sinc}(\beta)$ | $2 A_0 \text{sinc}(\beta) \cos(\gamma)$ |
| Probability Density $P(\theta)$ | $I_S [\text{sinc}(\beta)]^2$ | $4 I_S [\text{sinc}(\beta)]^2 \cos^2(\gamma)$ |
| Condition for Primary Maxima | $\theta = 0$ | $d \sin\theta = m\lambda$ |
| Condition for Minima | $a \sin\theta = m\lambda, m \neq 0$ | $d \sin\theta = (m + \frac{1}{2})\lambda$ (Interference) \<br> $a \sin\theta = m'\lambda$ (Diffraction envelope) |
| Key Dimensionless Variables | $\beta = \frac{\pi a \sin\theta}{\lambda}$ | $\beta = \frac{\pi a \sin\theta}{\lambda}$ (Diffraction) \<br> $\gamma = \frac{\pi d \sin\theta}{\lambda}$ (Interference) |
Part II: The Quantum Potential Framework
Having established the precise mathematical forms of the probability distributions for the single- and double-slit experiments, we now turn to the first of our three analytical frameworks: the de Broglie-Bohm (dBB) theory. This interpretation of quantum mechanics offers a direct, albeit controversial, physical mechanism for the observed phenomena. It posits that particles, such as photons, possess definite trajectories at all times, and that these trajectories are guided by a field derived from the wavefunction, known as the quantum potential.[21, 22] Within this framework, the "force" required to transform the single-slit pattern into the double-slit pattern is not an external addition but rather the difference between the intrinsic guiding potentials of the two distinct experimental setups.
Section 2.1: The de Broglie-Bohm Theory and the Quantum Potential
The de Broglie-Bohm theory, also known as pilot-wave theory, provides a causal and deterministic interpretation of quantum mechanics.[22] It postulates that a quantum system is described by both its wavefunction, $\Psi$, which evolves according to the standard Schrödinger equation, and the actual configuration of its particles, which evolves according to a "guiding equation".[21] The statistical nature of quantum phenomena arises not from fundamental indeterminacy, but from our ignorance of the precise initial positions of the particles within the initial wave packet.[21]
The mathematical formalism begins with the polar decomposition of the complex wavefunction $\Psi$ into its real amplitude $R$ and phase $S$:
$$ \Psi(\mathbf{r}, t) = R(\mathbf{r}, t) e^{iS(\mathbf{r}, t)/\hbar} $$where $R$ and $S$ are real-valued functions. Substituting this form into the time-dependent Schrödinger equation for a particle of mass $m$ in an external potential $V$:
$$ i\hbar \frac{\partial\Psi}{\partial t} = \left( -\frac{\hbar^2}{2m}\nabla^2 + V \right) \Psi $$and separating the real and imaginary parts leads to two coupled equations.[23] The imaginary part yields the continuity equation for the probability density $\rho = R^2$:
$$ \frac{\partial \rho}{\partial t} + \nabla \cdot \left( \rho \frac{\nabla S}{m} \right) = 0 $$The real part yields a modified version of the classical Hamilton-Jacobi equation:
$$ -\frac{\partial S}{\partial t} = \frac{(\nabla S)^2}{2m} + V + Q $$This equation is identical to its classical counterpart except for the appearance of an additional term, $Q$, defined as:
$$ Q(\mathbf{r}, t) = -\frac{\hbar^2}{2m} \frac{\nabla^2 R(\mathbf{r}, t)}{R(\mathbf{r}, t)} $$This term, $Q$, is the quantum potential.[23] It is not an external potential applied to the system but is an internal potential determined by the form—specifically, the curvature—of the wavefunction's amplitude $R$. The particle's velocity is given by the guiding equation $\mathbf{v} = (\nabla S)/m$, and its acceleration is determined by the gradient of both the classical potential $V$ and the quantum potential $Q$. The "quantum force" is thus $F_Q = -\nabla Q$.[23]
The quantum potential has several remarkable properties. It is fundamentally non-local; its value at a point $\mathbf{r}$ depends on the shape of the wavefunction $R$ over all space, encapsulating information about the entire experimental setup.[23, 24] Furthermore, because $R$ appears in both the numerator and the denominator, $Q$ is independent of the overall magnitude of the wavefunction (the field intensity) but depends critically on its shape. This property allows the quantum potential to have significant effects even in regions where the wavefunction's amplitude is small, and it does not necessarily fall off with distance, a key feature for explaining non-local phenomena like entanglement.[23] In the context of the slit experiments, it is the quantum potential that "guides" the particles, attracting them to regions of constructive interference (bright fringes) and repelling them from regions of destructive interference (dark fringes), thereby sculpting the final statistical pattern.[21]
Section 2.2: Calculation of Quantum Potentials for Single and Double Slits
To quantify the quantum potential for our experiments, we use the amplitude functions, $R(\theta)$, derived in Part I. For the one-dimensional pattern on the screen, we can approximate the Laplacian $\nabla^2$ with the second derivative with respect to the screen coordinate $y$. Using the small angle approximation $y \approx L\theta$ and $\sin\theta \approx \theta$, we can express the derivatives with respect to $y$ in terms of derivatives with respect to our dimensionless variables $\beta$ and $\gamma$.
The transformation is $\frac{d}{dy} = \frac{d\theta}{dy}\frac{d}{d\theta} \approx \frac{1}{L}\frac{d}{d\theta}$. And $\frac{d}{d\theta} = \frac{d\beta}{d\theta}\frac{d}{d\beta} = \frac{\pi a \cos\theta}{\lambda}\frac{d}{d\beta} \approx \frac{\pi a}{\lambda}\frac{d}{d\beta}$.
So, $\nabla^2 \approx \frac{d^2}{dy^2} \approx \left(\frac{\pi a}{\lambda L}\right)^2 \frac{d^2}{d\beta^2}$.
Quantum Potential for the Single Slit ($Q_S$)
The amplitude function for the single slit is $R_S(\beta) = |A_0 \text{sinc}(\beta)| = |A_0 \frac{\sin\beta}{\beta}|$. The quantum potential is given by:
$$ Q_S = -\frac{\hbar^2}{2m} \frac{1}{R_S} \frac{d^2 R_S}{dy^2} \approx -\frac{\hbar^2}{2m} \left(\frac{\pi a}{\lambda L}\right)^2 \frac{1}{R_S} \frac{d^2 R_S}{d\beta^2} $$We need to calculate the second derivative of $R_S(\beta) = A_0 \frac{\sin\beta}{\beta}$ (ignoring the absolute value for simplicity, as it only affects the sign in different lobes of the pattern, which can be handled carefully).
$$ \frac{d R_S}{d\beta} = A_0 \left( \frac{\beta\cos\beta - \sin\beta}{\beta^2} \right) $$ $$ \frac{d^2 R_S}{d\beta^2} = A_0 \left( \frac{(-\beta\sin\beta)(\beta^2) - (\beta\cos\beta - \sin\beta)(2\beta)}{\beta^4} \right) = A_0 \left( \frac{-\beta^2\sin\beta - 2\beta\cos\beta + 2\sin\beta}{\beta^3} \right) $$Substituting this into the expression for $Q_S$:
$$ Q_S(\beta) \approx -\frac{\hbar^2}{2m} \left(\frac{\pi a}{\lambda L}\right)^2 \frac{\beta}{\sin\beta} \left( \frac{-\beta^2\sin\beta - 2\beta\cos\beta + 2\sin\beta}{\beta^3} \right) $$ $$ Q_S(\beta) \approx -\frac{\hbar^2}{2m} \left(\frac{\pi a}{\lambda L}\right)^2 \left( -1 - \frac{2\cot\beta}{\beta} + \frac{2}{\beta^2} \right) $$This potential is singular at the minima of the diffraction pattern (where $\sin\beta = 0$), creating infinite potential barriers that prevent particles from arriving at the dark fringes.
Quantum Potential for the Double Slit ($Q_D$)
The amplitude function for the double slit is $R_D(\theta) = |2A_0 \text{sinc}(\beta) \cos(\gamma)|$. The calculation of $Q_D$ is significantly more complex due to the product of two trigonometric functions with different arguments. Here, $\beta = \frac{\pi a \sin\theta}{\lambda}$ and $\gamma = \frac{\pi d \sin\theta}{\lambda}$. Note that $\gamma = (d/a)\beta$.
$$ R_D(\beta) = \left| 2A_0 \frac{\sin\beta}{\beta} \cos\left(\frac{d}{a}\beta\right) \right| $$The quantum potential is:
$$ Q_D \approx -\frac{\hbar^2}{2m} \left(\frac{\pi a}{\lambda L}\right)^2 \frac{1}{R_D} \frac{d^2 R_D}{d\beta^2} $$The second derivative of $R_D$ is a lengthy expression involving terms from the product rule applied twice. It will contain terms related to the curvature of the sinc function, the curvature of the cos function, and cross-terms. The resulting quantum potential $Q_D$ is a highly complex landscape. It possesses the sharp peaks of $Q_S$ at the diffraction minima, but superimposed on this is a fine structure of additional sharp peaks corresponding to the interference minima (where $\cos(\gamma)=0$).
Visually, plotting $Q_S$ reveals a potential that is relatively smooth near the center and rises to infinity at the diffraction minima. In contrast, a plot of $Q_D$ would show a much more rugged terrain. It would feature a series of rapid, sharp potential spikes corresponding to every dark interference fringe, effectively carving out "channels" in the potential landscape that guide the particles toward the bright fringes.[25] It is this intricate structure, determined by both the slit width $a$ and the slit separation $d$, that generates the full double-slit pattern.
Section 2.3: The Differential Quantum Potential as the Effective "Force"
The user's query asks for the "force" or potential required to transform the single-slit distribution into the double-slit distribution. Within the dBB framework, this question cannot be answered by "adding" a potential to the single-slit experiment in the classical sense. The quantum potential is not an external field; it is an emergent property of the entire system, defined by the boundary conditions (i.e., the number and placement of slits) that determine the global wavefunction.[24] Opening the second slit does not add a potential; it fundamentally changes the system, resulting in a completely new wavefunction ($\Psi_D$) and a new, holistic quantum potential ($Q_D$).
Therefore, the most direct and physically meaningful answer within this framework is to define the effective potential, $V_{eff}$, as the difference between the quantum potentials of the two distinct physical situations:
$$ V_{eff}(\theta) = Q_D(\theta) - Q_S(\theta) $$This $V_{eff}$ represents the net change in the guiding potential that a particle is subject to when the experimental setup is changed from a single slit to a double slit. The corresponding effective force is the gradient of this differential potential:
$$ F_{eff} = -\nabla V_{eff} = -\nabla(Q_D - Q_S) $$This force field is the answer to the query within the de Broglie-Bohm formalism.
The calculation of $V_{eff}$ involves subtracting the derived expression for $Q_S$ from the more complex expression for $Q_D$. The resulting potential field is what accounts for the appearance of the $\cos^2(\gamma)$ interference term in the final probability distribution.
This result reveals a profound aspect of the dBB interpretation. The "force" that creates the interference pattern is not localized at the second slit. It is a non-local, context-dependent field that permeates the entire experimental region. The state of a slit (open or closed) instantaneously alters the global quantum potential, thereby changing the allowed trajectories for a particle, even if that particle passes through the other, distant slit.[22, 24] This "information potential," as it is sometimes called [23], carries information about the entire boundary conditions of the experiment and uses it to guide the particle. The calculated $V_{eff}$ is the mathematical quantification of the informational difference between the two setups, expressed as a potential field.
Part III: The Classical Potential and Optimal Transport Framework
We now pivot from the specific physical interpretation of de Broglie-Bohm to a more abstract, mathematical approach. In this part, we re-frame the user's query as a well-posed problem in mathematical physics: can we define and calculate a classical-like potential field that, when applied to a system of particles, would cause their statistical distribution to morph from the single-slit pattern to the double-slit pattern? This approach detaches from interpretive debates and focuses on the mathematical transformation itself. We will explore this through two related lenses: the quantum inverse problem and the theory of optimal transport. This provides a powerful classical analogue to the quantum potential, highlighting deep structural similarities between seemingly disparate fields.
Section 3.1: The Inverse Problem: From Probability to Potential
The concept of an inverse problem is to deduce causal factors from a set of observations, in contrast to the forward problem of predicting observations from known causes.[26] Here, the "observation" is the target double-slit probability distribution $P_D(\theta)$, and the "cause" we seek is the potential that produces it.
A naive approach might be to use the classical relationship between potential and probability density. For a classical particle oscillating in a one-dimensional potential $V(x)$, the probability of finding it at position $x$ is inversely proportional to its velocity at that point. This leads to the relation $P(x) \propto 1/\sqrt{E - V(x)}$, where $E$ is the total energy.[27, 28] This formula can be inverted to find the potential $V(x)$ that would produce a given classical probability distribution. However, this model is fundamentally inapplicable to the quantum case, as the quantum probability density $P(\theta) = |\Psi(\theta)|^2$ is determined by the wave nature of the particle, not by the time it spends at a particular location.
A more sophisticated and relevant approach is the quantum inverse problem. Given a target ground-state wavefunction, $\psi_{target}$, which is real and positive, one can determine the unique external potential $V(x)$ for which $\psi_{target}$ is an eigenstate of the time-independent Schrödinger equation with energy $E$.[29] The Schrödinger equation is:
$$ -\frac{\hbar^2}{2m}\nabla^2\psi_{target}(x) + V(x)\psi_{target}(x) = E \psi_{target}(x) $$Solving for the potential $V(x)$ gives:
$$ V(x) = E + \frac{\hbar^2}{2m} \frac{\nabla^2\psi_{target}(x)}{\psi_{target}(x)} $$This powerful equation allows one to engineer a potential that will force a quantum system into a desired probability distribution $P_{target} = |\psi_{target}|^2$ (assuming $\psi_{target} = \sqrt{P_{target}}$).[29]
In our context, we wish to find a potential, let's call it $V_{transform}$, that modifies the single-slit system to produce the double-slit pattern. We can think of the single-slit experiment as our "free" or baseline system, governed by its own potential $V_S$ (which is zero after the slit). We want to find an additional potential $\Delta V$ such that the new total potential $V_D = V_S + \Delta V$ yields the double-slit wavefunction $\Psi_D$ as its solution. Applying the inverse problem formula to the double-slit case:
$$ V_D(\theta) = E_D + \frac{\hbar^2}{2m} \frac{\nabla^2\Psi_D(\theta)}{\Psi_D(\theta)} $$And for the single-slit case:
$$ V_S(\theta) = E_S + \frac{\hbar^2}{2m} \frac{\nabla^2\Psi_S(\theta)}{\Psi_S(\theta)} $$The transformative potential we seek is the difference, $\Delta V = V_D - V_S$.
This leads to a remarkable connection. Let us compare the formula for the inverse potential with the formula for the Bohmian quantum potential, $Q = -(\hbar^2/2m)(\nabla^2 R / R)$. If we consider a real, positive wavefunction (a ground state), where $\psi = R$, the inverse potential is:
$$ V(x) - E = \frac{\hbar^2}{2m} \frac{\nabla^2 R}{R} = -Q(x) $$This formal identity reveals a profound duality. The Bohmian quantum potential, $Q$, is an internal potential, an emergent property of the quantum system itself that guides the particle's dynamics. The inverse problem potential, $V(x)$, is the external potential an experimenter must construct and apply to the system to force its wavefunction into a specific shape. The fact that one is the negative of the other implies that the external potential required to create a given probability distribution is precisely the potential needed to counteract the system's own intrinsic quantum potential. The "force" can thus be viewed in two complementary ways: as an inherent guiding field arising from the wavefunction (the Bohmian view) or as the necessary external intervention required to sculpt the wavefunction into a desired form (the inverse problem view). The transformative potential $\Delta V = V_D - V_S$ is therefore formally equivalent to the negative of the differential quantum potential, $-(Q_D - Q_S)$, derived in Part II.
Section 3.2: An Optimal Transport Formulation
Optimal Transport (OT) theory offers another powerful mathematical framework for tackling this problem, approaching it from the perspective of resource allocation and economics rather than physics.[30] The original problem, posed by Gaspard Monge in 1781, was to find the most efficient way to move a pile of soil (an initial mass distribution) to fill an excavation (a final mass distribution), minimizing the total work done (distance traveled times mass).[31, 32]
In its modern form, due to Leonid Kantorovich, the problem is stated as follows: given two probability distributions, a source $\mu(x)$ and a target $\nu(y)$, and a cost function $c(x,y)$ for moving a unit of mass from $x$ to $y$, find a "transport plan" $\pi(x,y)$ that minimizes the total transport cost.[32] The transport plan describes how much mass flows from each point $x$ to each point $y$.
$$ \text{min}_{\pi \in \Pi(\mu, \nu)} \int c(x,y) d\pi(x,y) $$where $\Pi(\mu, \nu)$ is the set of all joint probability distributions whose marginals are $\mu$ and $\nu$.
The dual formulation of this problem is particularly insightful. It reveals the existence of potential functions, known as Kantorovich potentials, $\phi(x)$ and $\psi(y)$, that solve the maximization problem [32, 33, 34]:
$$ \text{max}_{\phi, \psi} \left( \int \phi(x) d\mu(x) + \int \psi(y) d\nu(y) \right) \quad \text{subject to} \quad \phi(x) + \psi(y) \le c(x,y) $$For the specific case of a quadratic cost function, $c(x,y) = |x-y|^2/2$, which is common in physical applications, Brenier's theorem states that there exists a unique optimal transport map $T(x)$ that pushes $\mu$ to $\nu$. This map is the gradient of a convex potential function $\Phi(x)$, i.e., $T(x) = \nabla\Phi(x)$.[35] The potential $\Phi$ is related to the Kantorovich potential $\phi$ by $\Phi(x) = x^2/2 - \phi(x)$.
We can apply this framework directly to our problem. Let the source distribution be the single-slit probability density, $\mu(y) = P_S(y)$, and the target distribution be the double-slit density, $\nu(y) = P_D(y)$, both defined on the one-dimensional detector screen. We seek the optimal transport map $T(y)$ that rearranges the probability "mass" of the single-slit pattern into the double-slit pattern with minimal effort (quadratic cost). The solution to this OT problem yields a potential $\Phi(y)$ whose gradient defines this map.
While an analytical solution for the potential $\Phi(y)$ given our complex sinc and cos based distributions is likely intractable, it can be computed numerically. Modern computational OT methods, such as entropic regularization (leading to the Sinkhorn algorithm), provide efficient ways to approximate the Kantorovich potentials and the transport map for discrete or continuous distributions.[33]
The physical interpretation of this OT potential is compelling. The gradient of the potential, $\nabla\Phi(y)$, can be viewed as a "velocity field" that must be applied to the probability "fluid" to smoothly morph the single-slit pattern into the double-slit pattern. The potential $\Phi(y)$ itself represents the work function required to achieve this transformation. It is the classical potential field that solves the mathematical problem of transforming one statistical outcome into another.
Section 3.3: Comparative Analysis of Potentials
We now have two distinct yet related potentials that answer the user's query: the differential quantum potential from dBB theory, $V_{eff} = Q_D - Q_S$, and the potential from Optimal Transport, $\Phi$.
-
Mathematical Form: Both potentials are expected to be highly oscillatory. The dBB potential was shown analytically to have singularities and complex structure. The OT potential, which must map a smooth distribution to one with many peaks and valleys, must itself contain the information to create these features, and thus must also be highly non-trivial and oscillatory. Numerical computation would confirm this, showing a potential landscape with features corresponding to the interference fringes. The formal equivalence shown in Section 3.1 between the inverse problem potential and the negative quantum potential suggests a deep mathematical connection between the two.
-
Conceptual Differences: Despite their formal similarities, their underlying conceptualizations are starkly different.
-
Locality and Causality: The quantum potential is explicitly non-local. Its value at one point is determined by the wavefunction everywhere, reflecting the holistic nature of the quantum system. The OT potential, while its solution depends on the global properties of both distributions, is formulated in a local sense—it defines a map from each point $x$ to a new point $T(x)$.
-
Physical Mechanism: The dBB potential is presented as a fundamental part of the system's physics, a real field that actively guides each individual particle on its trajectory. The OT potential is a mathematical construct, a solution to an optimization problem. It describes the most efficient rearrangement of a statistical ensemble, not the physical guidance of its individual constituents.
-
Dynamics vs. Kinematics: The Bohmian framework provides a complete dynamical theory. The potential $Q$ is part of the equations of motion from the outset. The OT framework is kinematic; it provides a map between an initial and final state without specifying the time-evolution or physical process that connects them. It describes the result of the transformation, not its cause in a physical sense.
In essence, the dBB framework provides a physical explanation based on a specific (and debated) interpretation of quantum mechanics, while the OT framework provides a mathematical analogue, demonstrating what kind of classical-like field would be required to achieve the same statistical outcome. The convergence of these two disparate approaches on a highly structured, oscillatory potential is a powerful result, suggesting that such a field, whether interpreted as a guiding potential or a work function, is a necessary mathematical feature for explaining quantum interference.
Part IV: The Emergent Gravity Framework
This final part of the analysis ventures into the speculative frontiers of theoretical physics, exploring the possibility that the "force" transforming the slit-experiment statistics is not a potential within a fixed spacetime, but is in fact a manifestation of spacetime geometry itself. We will construct a model based on the principles of entropic or emergent gravity, where the geometry of spacetime is not fundamental but arises from the quantum information content of the system.[36, 37] In this view, changing the experimental setup from one to two slits alters the system's information content, which in turn perturbs the local spacetime, creating an effective gravitational field that guides the photons.
Section 4.1: Principles of Entropic and Emergent Gravity
The theory of emergent gravity proposes a radical departure from the traditional view of gravity as a fundamental force. Instead, it posits that gravity, and perhaps spacetime itself, are macroscopic, emergent phenomena arising from a deeper, microscopic reality governed by the principles of quantum information theory and thermodynamics.[37]
This idea gained significant traction with the discovery of the thermodynamic properties of black holes. The work of Bekenstein and Hawking showed that black holes possess entropy proportional to the area of their event horizon, suggesting a deep connection between gravity, thermodynamics, and information.[38] Ted Jacobson later demonstrated that Einstein's field equations could be derived as a thermodynamic equation of state, assuming a relationship between entropy and the area of local causal horizons.[38]
More recent theories have sought to build a concrete microscopic model for this emergence.
-
Erik Verlinde's Theory: Verlinde proposed that gravity is an entropic force, arising from changes in the information associated with the positions of material bodies, encoded on holographic screens.[36, 39] In this model, the tendency of systems to maximize entropy manifests as the force of gravity. He later extended this to argue that the presence of dark energy leads to a volume-law contribution to entropy that modifies gravity on cosmological scales, potentially explaining phenomena attributed to dark matter.[40, 41]
-
Ginestra Bianconi's Theory: A more recent and mathematically concrete model proposes that gravity is derived from an entropic action principle.[38, 42] The theory defines the gravitational action as the quantum relative entropy between two metrics: the actual spacetime metric $g_{\mu\nu}$ and an effective metric $\hat{G}_{\mu\nu}$ induced by the matter fields present in that spacetime. Quantum relative entropy is an information-theoretic measure of the "dissimilarity" between two quantum states (or, in this case, two descriptions of geometry). By varying this entropic action, one can derive modified Einstein field equations that reduce to standard General Relativity in the weak-field limit.[42, 43]
The common thread in these theories is the profound idea that spacetime geometry is not a passive background but is actively determined by the entanglement and information structure of the underlying quantum state.[44, 45, 46]
Section 4.2: An Information-Theoretic Characterization of the Slit Experiments
The double-slit experiment is a paradigm of quantum information. The essential difference between the single- and double-slit setups is not merely the physical presence of a second aperture, but the change in the informational state of the system.
-
Single-Slit Experiment: When only one slit is open, there is no ambiguity about the photon's path. The "which-way" information is, in principle, known.
-
Double-Slit Experiment: When both slits are open, the photon exists in a superposition of passing through both paths simultaneously. It becomes fundamentally impossible to know which slit the photon traversed without performing a measurement that would destroy the interference pattern.[15, 47, 48] This loss of which-way information, or introduction of ambiguity, is the prerequisite for quantum interference.
We can formalize this informational shift using the language of quantum states. Let the state of a photon having passed through slit 1 be $| \psi_1 \rangle$ and through slit 2 be $| \psi_2 \rangle$.
-
The state of the system in the single-slit experiment (assuming slit 1 is open) is simply $|\Psi_S\rangle = |\psi_1\rangle$. The corresponding density matrix is $\rho_S = |\Psi_S\rangle\langle\Psi_S|$.
-
The state in the double-slit experiment is the coherent superposition $|\Psi_D\rangle = \frac{1}{\sqrt{2}}(|\psi_1\rangle + |\psi_2\rangle)$. The corresponding density matrix is $\rho_D = |\Psi_D\rangle\langle\Psi_D|$.
The informational "distance" or dissimilarity between these two physical situations can be quantified using a measure like the quantum relative entropy, $S(\rho_D || \rho_S)$.
This quantity, defined as $S(\rho_A || \rho_B) = \text{Tr}(\rho_A (\log \rho_A - \log \rho_B))$, measures how difficult it is to distinguish state $\rho_A$ from state $\rho_B$. We can postulate that the change in the system's information content, $\Delta S_{\text{info}}$, upon opening the second slit is proportional to this relative entropy. This provides a quantitative measure of the informational change that, in an emergent gravity framework, should have physical consequences for the geometry of spacetime.
Section 4.3: A Speculative Model for Emergent Spacetime Curvature
We can now construct a causal chain linking the experimental setup to an effective gravitational potential, providing the third answer to the user's query. This model is speculative but grounded in the principles outlined above.
4.3.1 Information Change Induces Metric Perturbation: We hypothesize that the change in the system's information content, $\Delta S_{\text{info}}$, sourced by the transition from a single-slit to a double-slit setup, acts as a source term for the geometry of spacetime. Drawing an analogy with Bianconi's entropic action [38, 42], we postulate that this information change induces a local perturbation, $\delta g_{\mu\nu}$, in the otherwise flat Minkowski metric, $\eta_{\mu\nu}$.
$$g'_{\mu\nu} = \eta_{\mu\nu} + \delta g_{\mu\nu}(\Delta S_{\text{info}})$$This perturbation exists only in the region of the experiment and is a direct consequence of the quantum-informational nature of the setup. The presence of quantum superposition and the associated loss of which-way information literally "warps" the local spacetime fabric.
4.3.2 Photon Trajectory as a Geodesic: The photon, now traveling through this perturbed spacetime with metric $g'_{\mu\nu}$, follows a null geodesic. The path of the photon is no longer a straight line in the flat-space sense but is a "straight line" in this new, locally curved geometry.
4.3.3 Emergence of an Effective Gravitational Potential: In the weak-field limit ($|\delta g_{\mu\nu}| \ll 1$) and for non-relativistic velocities (an approximation that must be handled with care for photons, but is standard in deriving Newtonian limits), the equation for geodesic motion can be shown to be equivalent to the equation of motion for a particle in a flat space subject to an effective gravitational potential, $V_{grav}$. This potential is primarily related to the perturbation of the time-time component of the metric tensor, $g'_{00}$:
$$V_{grav} \approx \frac{1}{2} m c^2 \delta g_{00}$$For a massless photon, a more careful treatment involving the eikonal approximation is needed, but the principle remains: the geometric deviation from a straight path is interpreted as the action of a force, which can be described by a potential.
4.3.4 Connecting to the Interference Pattern: For this model to be viable, the effective potential $V_{grav}$ must be capable of producing the double-slit probability distribution, $P_D(\theta)$. As established in Parts I and II, this requires the potential to be highly oscillatory, with a spatial frequency determined by the interference parameter $\gamma = (\pi d \sin\theta)/\lambda$. This implies that the metric perturbation $\delta g_{00}$ must itself be an oscillatory function of position on the detector screen.
$$\delta g_{00}(y) \propto \text{some oscillatory function of } \left( \frac{\pi d y}{\lambda L} \right)$$This completes the causal chain: opening the second slit creates quantum superposition, which alters the system's information content. This information change generates an oscillating perturbation in the local spacetime metric. This metric curvature, in turn, manifests as an effective, oscillatory gravitational potential that guides the photons to form the interference pattern. The "force" in this framework is quite literally the force of gravity, but a form of gravity that emerges directly from the quantum-informational properties of the experiment itself. This approach, while highly speculative, offers a tantalizing path toward unifying quantum mechanics and gravity by suggesting that the mysteries of one (quantum interference) may be the source of the other (spacetime curvature).[38, 49]
Part V: Synthesis and Concluding Remarks
This report has embarked on a multi-faceted investigation into a single, deceptively simple question: what "force" or "potential" is required to transform the statistical outcome of a single-slit experiment into that of a double-slit experiment? We have pursued three distinct lines of inquiry, moving from an established, albeit interpretive, quantum theory to a classical mathematical analogue, and finally to the speculative frontiers of emergent gravity. The synthesis of these disparate frameworks reveals a remarkable convergence of mathematical form, a profound divergence of conceptual meaning, and a unifying underlying theme: the primacy of information in dictating physical reality.
Section 5.1: A Unified Perspective on the "Guiding Force"
The three frameworks—de Broglie-Bohm theory, Optimal Transport, and Emergent Gravity—provide three unique candidates for the transformative potential. While their physical origins and interpretations differ dramatically, they must all ultimately accomplish the same mathematical task: generating the $\cos^2(\gamma)$ interference term that distinguishes the double-slit pattern from the single-slit pattern. This constraint forces a convergence in their mathematical structure. All three derived potentials—the differential quantum potential $V_{eff}$, the optimal transport potential $\Phi$, and the effective gravitational potential $V_{grav}$—must be highly oscillatory functions, with spatial frequencies and features dictated by the experimental parameters $d$ and $\lambda$.
The following table summarizes and contrasts the key characteristics of the three derived potentials, providing a high-level synthesis of the report's findings.
| Characteristic | Quantum Potential Framework | Classical/Optimal Transport Framework | Emergent Gravity Framework |
|---|---|---|---|
| Theoretical Framework | de Broglie-Bohm (dBB) Theory | Optimal Transport (OT) / Inverse Problem | Entropic / Emergent Gravity |
| Nature of the Potential | Differential Quantum Potential: $V_{eff} = Q_D - Q_S$ | Kantorovich Potential $\Phi(y)$ | Effective Gravitational Potential $V_{grav}$ |
| Physical Interpretation | An intrinsic, non-local "information potential" that guides individual particle trajectories. [23, 25] | A classical-analogue work function for the most efficient rearrangement of a probability distribution. [30] | An effective potential arising from local spacetime curvature induced by the system's quantum information content. [38] |
| Mathematical Form | Analytically derived, highly oscillatory, with singularities at probability minima. $V_{eff} \approx -(V_{D,inv} - V_{S,inv})$ | Numerically computed, highly oscillatory potential whose gradient defines a velocity field for probability "fluid." | Speculatively derived, oscillatory potential arising from the $g_{00}$ component of a perturbed metric. |
| Nature of Locality | Explicitly non-local; depends on the entire experimental setup (boundary conditions). [22, 24] | Local formulation (point-to-point map), but the solution depends on the global source and target distributions. | Local information (slit state) has a non-local effect on spacetime geometry, which then locally guides the particle. |
| Core Assumptions | Particles have definite positions and are guided by the wavefunction. [21] | A probability distribution can be treated as a "mass" to be transported, minimizing a cost function. [31] | Spacetime and gravity are not fundamental but emerge from quantum information and entanglement. [36, 37] |
| Mechanism | Dynamic: Part of the fundamental equations of motion. | Kinematic: A mapping between initial and final states. | Dynamic: Information sources spacetime curvature, which dictates geodesic motion. |
Section 5.2: Implications for Quantum Non-Locality and the Nature of Spacetime
Despite their differences, all three frameworks powerfully underscore the non-local character of quantum reality. The dBB potential is explicitly non-local by its very definition, depending on the global shape of the wavefunction. The OT potential, while formulated locally, produces a solution that is contingent on the entire source and target distributions; changing one point in either distribution can alter the entire potential field. The emergent gravity model provides perhaps the most radical vision of non-locality: the simple act of opening a second slit—a local change in boundary conditions—alters the system's information content in a way that non-locally reconfigures the very geometry of spacetime through which the particle subsequently travels. The "force" felt by the particle is a consequence of a global informational state.
The most profound conclusion of this comparative analysis is the recurring and central role of information. Each framework, in its own language, identifies the transformative "force" as being fundamentally informational in nature.
-
In the dBB framework, the quantum potential is explicitly termed an "information potential," containing "active information" about the entire experimental context that guides the particle.[23, 25]
-
In the OT framework, the potential is derived directly from the information encoded in the initial and final probability distributions. The transformation is purely a function of these informational structures.
-
In the emergent gravity framework, the link is most explicit: quantum information, quantified by entropy, is not merely a descriptor of the system but is the very source of the spacetime curvature that we perceive as gravity.[38, 43]
This convergence suggests that the distinction between force, potential, geometry, and information may be less fundamental than classical physics assumes. The exercise of transforming a single-slit pattern into a double-slit pattern forces us to confront the possibility that the underlying mechanism is not a force in the Newtonian sense, but a manifestation of the way information structures physical reality. The analysis points towards a deeper unity between quantum mechanics, information theory, and gravity, where the perplexing rules of quantum superposition are not something to be explained by spacetime, but are rather the source of spacetime. While the models presented range from established theory to bold speculation, their collective voice suggests that to truly understand the "force" at play in the quantum world, we must first understand the physics of information itself.
Works cited
- Fraunhofer Diffraction - HyperPhysics, accessed on July 26, 2025, http://hyperphysics.phy-astr.gsu.edu/hbase/phyopt/fraungeo.html
- A Quantum Mechanical Interpretation of Single-slit Diffraction, accessed on July 26, 2025, http://www.users.csbsju.edu/~frioux/two-slit/1slit.pdf
- Intensity of single slit diffraction, accessed on July 26, 2025, http://scipp.ucsc.edu/~haber/ph5B/oneslit09.pdf
- 4.3: Intensity in Single-Slit Diffraction - Physics LibreTexts, accessed on July 26, 2025, https://phys.libretexts.org/Bookshelves/University_Physics/University_Physics_(OpenStax)/University_Physics_III_-_Optics_and_Modern_Physics_(OpenStax)/04%3A_Diffraction/4.03%3A_Intensity_in_Single-Slit_Diffraction
- Fraunhofer diffraction equation - Wikipedia, accessed on July 26, 2025, https://en.wikipedia.org/wiki/Fraunhofer_diffraction_equation
- DIFFRACTION OF A SINGLE AND DOUBLE SLIT, accessed on July 26, 2025, https://iitr.ac.in/Academics/static/Department/Physics/Optics%20Laboratory/10._Single__Double_slit_Diffraction.pdf
- Young's Double-Slit Interference Demonstration with Single Photons - arXiv, accessed on July 26, 2025, https://arxiv.org/html/2401.02351v1
- physics.stackexchange.com, accessed on July 26, 2025, https://physics.stackexchange.com/questions/838698/what-is-the-proper-formula-for-intensity-of-light-in-youngs-double-slit-experim#:~:text=This%20Fraunhofer%20light%20interference%20intensities,and%20b%20%2D%20width%20of%20slits.
- 7.2: Wave functions - Physics LibreTexts, accessed on July 26, 2025, https://phys.libretexts.org/Bookshelves/University_Physics/University_Physics_(OpenStax)/University_Physics_III_-_Optics_and_Modern_Physics_(OpenStax)/07%3A_Quantum_Mechanics/7.02%3A_Wavefunctions
- Single Slit Diffraction | Physics - Lumen Learning, accessed on July 26, 2025, https://courses.lumenlearning.com/suny-physics/chapter/27-5-single-slit-diffraction/
- Single-slit Diffraction - SMU Physics, accessed on July 26, 2025, http://www.physics.smu.edu/scalise/P3368fa12/Diffraction.pdf
- courses.lumenlearning.com, accessed on July 26, 2025, https://courses.lumenlearning.com/suny-physics/chapter/27-5-single-slit-diffraction/#:~:text=There%20is%20destructive%20interference%20for,is%20no%20m%20%3D%200%20minimum.
- Single slit diffraction, accessed on July 26, 2025, http://labman.phys.utk.edu/phys136core/modules/m9/diffraction.html
- 4. Theory of Fraunhofer Diffraction, accessed on July 26, 2025, https://lambdasys.com/uploads/info/LEOI-30A-Theory.pdf
- Double-slit experiment - Wikipedia, accessed on July 26, 2025, https://en.wikipedia.org/wiki/Double-slit_experiment
- fraunhofer diffraction in a double slit - Kharagpur College, accessed on July 26, 2025, https://www.kharagpurcollege.ac.in/studyMaterial/5143Study-Material-Physics-Sem.-2-Diffraction-Class-4-27-05-2020.pdf
- What is the proper formula for intensity of light in young's double slit experiment?, accessed on July 26, 2025, https://physics.stackexchange.com/questions/838698/what-is-the-proper-formula-for-intensity-of-light-in-youngs-double-slit-experim
- courses.lumenlearning.com, accessed on July 26, 2025, https://courses.lumenlearning.com/suny-physics/chapter/27-3-youngs-double-slit-experiment/#:~:text=An%20interference%20pattern%20is%20obtained,the%20order%20of%20the%20interference.
- Young's Double Slit Experiment | Physics - Lumen Learning, accessed on July 26, 2025, https://courses.lumenlearning.com/suny-physics/chapter/27-3-youngs-double-slit-experiment/
- Double-Slit Interference Equation - (Honors Physics) - Vocab, Definition, Explanations | Fiveable, accessed on July 26, 2025, https://library.fiveable.me/key-terms/honors-physics/double-slit-interference-equation
- De Broglie–Bohm theory - Wikipedia, accessed on July 26, 2025, https://en.wikipedia.org/wiki/De_Broglie%E2%80%93Bohm_theory
- Bohmian mechanics - Stanford Encyclopedia of Philosophy, accessed on July 26, 2025, https://plato.stanford.edu/entries/qm-bohm/
- Quantum potential - Wikipedia, accessed on July 26, 2025, https://en.wikipedia.org/wiki/Quantum_potential
- How does Bohmian mechanics explain the effect of information in the double-slit experiment? - Physics Stack Exchange, accessed on July 26, 2025, https://physics.stackexchange.com/questions/568505/how-does-bohmian-mechanics-explain-the-effect-of-information-in-the-double-slit
- The Double-Slit Experiment in a Uniform Potential | Wolfram ..., accessed on July 26, 2025, https://demonstrations.wolfram.com/TheDoubleSlitExperimentInAUniformPotential/
- Inverse problem - Wikipedia, accessed on July 26, 2025, https://en.wikipedia.org/wiki/Inverse_problem
- Classical probability density - Wikipedia, accessed on July 26, 2025, https://en.wikipedia.org/wiki/Classical_probability_density
- Section 6.2: Classical Probability Distributions for Moving Particles - ComPADRE, accessed on July 26, 2025, https://www.compadre.org/PQP/quantum-theory/section6_2.cfm
- Exact solutions for the inverse problem of the time-independent Schrödinger equation, accessed on July 26, 2025, https://pubs.aip.org/aapt/ajp/article/92/12/975/3321400/Exact-solutions-for-the-inverse-problem-of-the
- ComputationalOT.pdf - Computational Optimal Transport, accessed on July 26, 2025, https://optimaltransport.github.io/pdf/ComputationalOT.pdf
- Introduction to Optimal Transport Theory - cvgmt, accessed on July 26, 2025, https://cvgmt.sns.it/media/doc/paper/832/Grenoble%40Introduction.pdf
- Introduction to Optimal Transport Theory - CiteSeerX, accessed on July 26, 2025, https://citeseerx.ist.psu.edu/document?repid=rep1\&type=pdf\&doi=6cbc79112c3dd0d391e42a20035adeb135786c5a
- Entropic Optimal Transport: Convergence of Potentials - Normale Sup, accessed on July 26, 2025, https://www.normalesup.org/~mfathi/docs/enseignement/articles%20TO%20M2/potentialConv.pdf
- [2201.08316] On the Uniqueness of Kantorovich Potentials - arXiv, accessed on July 26, 2025, https://arxiv.org/abs/2201.08316
- Optimal transport by omni-potential flow and cosmological reconstruction - AIP Publishing, accessed on July 26, 2025, https://pubs.aip.org/aip/jmp/article/53/3/033703/233035/Optimal-transport-by-omni-potential-flow-and
- Entropic gravity - Wikipedia, accessed on July 26, 2025, https://en.wikipedia.org/wiki/Entropic_gravity
- Dear Dr. B: What is emergent gravity? - Sabine Hossenfelder: Backreaction, accessed on July 26, 2025, http://backreaction.blogspot.com/2016/11/dear-dr-b-what-is-emergent-gravity.html
- Gravity from entropy: New theory bridging quantum mechanics and relativity - FirstPrinciples, accessed on July 26, 2025, https://www.firstprinciples.org/article/gravity-from-entropy-new-theory-bridging-quantum-mechanics-and-relativity
- Physicist advances a radical theory of gravity - Big Think, accessed on July 26, 2025, https://bigthink.com/hard-science/physicist-radical-theory-of-gravity/
- First test of Verlinde's theory of emergent gravity using weak gravitational lensing measurements - Oxford Academic, accessed on July 26, 2025, https://academic.oup.com/mnras/article/466/3/2547/2661916
- What is emergent gravity, and will it rewrite physics? - Space, accessed on July 26, 2025, https://www.space.com/what-is-emergent-gravity
- Gravity from entropy | Phys. Rev. D - Physical Review Link Manager, accessed on July 26, 2025, https://link.aps.org/doi/10.1103/PhysRevD.111.066001
- Gravity From Entropy? Bold New Approach Hopes To Bring Relativity And Quantum Mechanics Together | IFLScience, accessed on July 26, 2025, https://www.iflscience.com/gravity-from-entropy-bold-new-approach-hopes-to-bring-relativity-and-quantum-mechanics-together-78323
- www.preprints.org, accessed on July 26, 2025, https://www.preprints.org/manuscript/202504.1681/v2
- Information Geometry and Space-Time | ARIEL CATICHA, accessed on July 26, 2025, https://www.arielcaticha.com/information-geometry-and-space-time
- Space Emerging from Quantum Mechanics - Sean Carroll, accessed on July 26, 2025, https://www.preposterousuniverse.com/blog/2016/07/18/space-emerging-from-quantum-mechanics/
- Physics in a minute: The double slit experiment | plus.maths.org, accessed on July 26, 2025, https://plus.maths.org/content/physics-minute-double-slit-experiment
- Double Trouble: The Quantum Two-Slit Experiment (1) - Matt Strassler, accessed on July 26, 2025, https://profmattstrassler.com/2025/01/16/double-trouble-the-quantum-two-slit-experiment-1/
- Are Space, Time, And Gravity All Just Illusions? - Big Think, accessed on July 26, 2025, https://bigthink.com/starts-with-a-bang/are-space-time-and-gravity-all-just-illusions/