Physics:Artificial Intelligence, Deep Learning in Theory of Entropicity (ToE)
Entropicity in Artificial Intelligence and Deep Learning: A Unified Framework Based on the Theory of Entropicity (ToE)
Abstract
The Theory of Entropicity(ToE),[1][2] first formulated and developed by John Onimisi Obidi,[3][4][5][6][7][8][9] is a recently proposed theoretical framework that elevates entropy from a statistical descriptor to a fundamental dynamical field underlying physical reality. In this paper, we present a unified framework for artificial intelligence (AI) and deep learning based on ToE’s principles of entropicity. We derive key mathematical formulations from ToE – including the Vuli–Ndlela Integral (an entropy-weighted path integral), entropy field equations (such as the Master Entropic Equation), and entropic flow representations – and show how they can be applied to learning systems. Building on these foundations, we explore applications to deep learning architectures (feedforward networks and convolutional neural networks), training dynamics of large language models (LLMs), and propose novel “psychentropic” AI paradigms in which cognitive processes are governed by self-referential entropy flows. We draw parallels between the ToE-based framework and existing paradigms in AI and neuroscience, comparing it to Bayesian inference, energy-based models, and the Free Energy Principle. The comparative analysis highlights that an entropy-centric view can recover and generalize these approaches – unifying randomness and determinism in a single field, relating energy minimization to entropy flows, and connecting to the brain’s tendency to minimize surprise. The paper is structured with a formal research layout (Introduction, Theoretical Framework, Mathematical Foundations, Applications, Comparative Analysis, Discussion, References) and includes relevant equations and conceptual diagrams. Our results indicate that Entropicity provides a powerful lens for understanding and designing intelligent systems, suggesting that intelligence is effectively the control of entropic flow under physical and informational constraints. This entropy-based perspective offers fresh insights into learning algorithms, information processing, and even the emergence of agency and consciousness in AI, making it a promising paradigm for future research in both physics and artificial intelligence.
Introduction
Modern artificial intelligence and deep learning have been remarkably successful using principles from statistics, information theory, and neuroscience – examples include Bayesian inference for probabilistic reasoning, energy-based models for generative learning, and the Free Energy Principle for brain-like adaptive behavior. These approaches, while powerful, each address different aspects of learning and intelligence. This paper proposes a unified theoretical framework for AI grounded in the Theory of Entropicity (ToE), an emerging physics paradigm which posits that entropy is the fundamental substance from which time, geometry, and dynamics emerge. ToE was first formulated by Obidi (2025) as an ambitious attempt to reconcile quantum mechanics and gravity, explaining phenomena such as the arrow of time, wavefunction collapse, and gravitation through an entropy field $S(x,t)$. In ToE, entropy is treated as a real physical field permeating spacetime, whose gradients and fluxes drive all processes – matter motion, information flow, and even the emergence of mind and consciousness.
By elevating entropy to ontological primacy, ToE inverts the conventional view: instead of entropy being a by-product of dynamics, dynamics themselves are an expression of underlying entropy flows. This perspective yields novel principles (e.g. the No-Rush Theorem ensuring no process is instantaneous) and field equations that introduce irreversibility and causality at the fundamental level. Our aim is to translate these theoretical insights into the realm of AI and deep learning. We hypothesize that intelligent behavior – whether in biological brains or artificial networks – can be interpreted as a manifestation of entropy field evolution. In other words, an intelligent agent or learning algorithm might be seen as an entropic system that directs and harnesses entropy flow to achieve structure and goals.
In this work, we articulate how core concepts from ToE can form a unified framework for AI: we review the theoretical basis of ToE, present its mathematical foundations (including key equations like the Vuli–Ndlela path integral and Master Entropic Equation), and demonstrate applications to deep learning and AI. Case studies include conventional neural network architectures (feedforward and convolutional networks) as well as large language models (LLMs), showing how training and inference in these systems can be reinterpreted in entropic terms. We also introduce the notion of psychentropic AI paradigms, referring to AI designs that incorporate self-referential entropy dynamics to potentially achieve higher-order cognitive properties (inspired by ToE’s account of consciousness via entropy).
Furthermore, we perform a comparative analysis between the ToE-based framework and existing paradigms. We discuss how a ToE approach relates to or generalizes: (i) Bayesian inference, which can be viewed as selecting high-probability (low surprise) hypotheses consistent with maximum entropy priors; (ii) energy-based models, which draw on statistical physics (Boltzmann distributions) to train neural networks, implicitly involving entropy via energy minimization; and (iii) the Free Energy Principle in neuroscience, which states that biological agents minimize a free-energy bound on surprise (entropic disorder) to maintain order. By examining these connections, we highlight that ToE offers a more fundamental, physics-grounded perspective – treating entropy as a physical field – whereas the other paradigms are either phenomenological or methodological. We will show that in many cases ToE’s predictions reduce to known principles (e.g. Bayes-optimal inference, Infomax, or energy minimization) as special cases or limiting behaviors, while also providing new insights into aspects those paradigms do not fully address (such as time-asymmetry, internal self-entropy, and the integration of randomness with physical law).
This paper is organized as follows. In Theoretical Framework, we summarize the Theory of Entropicity’s core principles and conceptual architecture, laying the groundwork for its application to information systems. In Mathematical Foundations, we introduce formal equations from ToE (Vuli–Ndlela Integral, entropy field equations, etc.) and adapt them to the AI context. Next, Applications covers specific case studies: we discuss how entropicity manifests in deep neural networks and learning processes, and propose the psychentropic AI concept. A Comparative Analysis section then contrasts the entropicity framework with Bayesian inference, energy-based models, and the Free Energy Principle, clarifying similarities and differences. Finally, in Discussion, we consider implications, open challenges, and future research directions. We hope this work inspires a cross-disciplinary dialogue, as it invites AI researchers to *“reimagine physical law as emerging not from geometry or probability, but from the entropic fabric of reality itself”*, ultimately suggesting a new entropy-centric paradigm for understanding intelligence.
Theoretical Framework: Entropicity and Intelligence
Theory of Entropicity (ToE) Overview: ToE posits that entropy $S(x,t)$ is a fundamental field filling the universe, and that all physical phenomena are instantiations of this field’s evolution. In classical thermodynamics and information theory, entropy is a measure of disorder or uncertainty; ToE extends this concept by giving entropy a concrete physical presence and dynamics. In this framework, familiar constructs like space, time, energy, and probability emerge from or are constrained by the entropy field’s behavior. Crucially, gradients in the entropy field drive motion and interactions – much like gradients in a potential field drive particle motion in classical physics. For example, ToE reinterprets gravity not as curvature of spacetime but as a consequence of entropy gradients produced by mass-energy distributions. In the Newtonian limit, the entropy field $S(x)$ around a mass $M$ satisfies a Poisson-like equation $\nabla^2 S(x) = -\eta,\rho_M(x)$ (with $\eta$ a coupling constant), yielding a solution $S(r) = S_\infty - \frac{\eta M}{4\pi r}$ that reproduces the $1/r^2$ gravitational force law via $-\nabla S(r)$. This example illustrates entropic forces: particles move along gradients of $S$ so as to increase total entropy, aligning with the second law of thermodynamics but now cast as a local field-driven effect.
Irreversibility and the Arrow of Time: By making entropy fundamental, ToE naturally embeds the arrow of time and irreversibility into physics. Entropy flow has a preferred direction – from past to future, entropy tends to increase – which ToE treats as a basic law rather than an emergent statistic. This is formalized by constraints like the Entropy Current Law:
which states that the divergence of the entropy current $J_S^\mu$ equals a non-negative production term $\sigma(x)$. Equation (1) represents a continuity equation for entropy, modified to allow local entropy production $\sigma(x)$. It encapsulates the second law (entropy non-decrease) in field form – entropy can flow and be produced, but never destroyed ($\sigma \ge 0$). From this, one derives that no physical process can be perfectly reversible or instantaneous in ToE. Indeed, the No-Rush Theorem (Entropic Time Limit) asserts a minimum time $\Delta t_{\min}$ for any process, determined by local entropy gradients and a field “stiffness” constant. One formulation is
where $\lambda$ is a material constant, $k_B$ the Boltzmann constant, and $\langle (\nabla S)^2 \rangle$ the averaged entropy gradient magnitude. Equation (2) ensures a built-in causal latency: even quantum events like wavefunction collapse must occur over a finite (albeit possibly very short) duration in ToE, rather than being instantaneous. This approach provides a novel resolution to quantum puzzles, treating wave function collapse as an entropy-driven phase transition that occurs once an entropic threshold is crossed.
Entropy as Unifying Substrate: A major conceptual appeal of ToE is that it unifies descriptions of randomness and order within one framework. Randomness (thermal fluctuations, quantum uncertainty) and lawful structure (classical dynamics, conservation laws) are seen as two sides of the entropic field’s influence. In information-theoretic terms, Boltzmann’s randomness and Shannon’s information become aspects of a single entropy field. Traditional physics often juxtaposes deterministic laws with probabilistic interpretations; ToE suggests that both can emerge from entropy’s twofold role: it provides a variational principle that yields effective deterministic laws (e.g. via least-action with entropy terms), while also inherently carrying statistical dispersion (via []entropy maximization]]). Indeed, in ToE **“energy, geometry, and forces are emergent encodings of $S$”** – meaning that what we normally optimize in AI (energy functions, error metrics) may actually be proxies for underlying entropy dynamics. This philosophy hints that an AI designed with entropy as a core element could naturally integrate creative stochastic exploration with goal-directed behavior, just as physical reality does.
Intelligence as Entropic Flow Redirection: We now bridge ToE to the notion of intelligence. In conventional terms, intelligence is often associated with the ability to create order (e.g. solving problems, organizing information) out of a disordered environment. A recent public heuristic by Mo Gawdat stated “intelligence creates order while entropy creates disorder”. ToE provides a counterintuitive reframing: intelligence is not the antithesis of entropy, but rather a process of guiding entropy flows. More formally, *“Intelligence is effective control of entropic flow under finite-time (ETL) constraints; its outputs register as ‘order’ or ‘disorder’ only relative to a chosen frame or objective”*. In other words, an intelligent agent is one that can manipulate the entropy field to steer the world (or its internal state) toward desired configurations (a goal manifold in the entropy landscape). This may entail locally decreasing entropy (creating structure) in a subsystem, at the expense of increasing entropy elsewhere – a phenomenon known in thermodynamics (e.g., a refrigerator creates local order but expels heat). ToE explicitly recognizes such trade-offs: *“intelligence can create order or disorder depending on frame and coupling… entropy is not ‘the enemy’ but the substrate; intelligence re-channels its flow.”*. For example, a machine learning algorithm might reduce uncertainty (entropy) in its predictions by increasing computational work (and thus heat dissipation entropy in hardware). An AI-based climate control system decreases entropy in a room (cooling and organizing air molecules) by increasing entropy in the environment (expelling heat). Even information security can be viewed through this lens: strong cryptography increases entropy (uncertainty) for unauthorized observers while maintaining structure for the intended user. These scenarios illustrate ToE’s view that what we label as “order” or “disorder” is observer-relative and context-dependent, but underlying it all is a global entropy balance that must be non-negative.
Conceptual Architecture for Entropic AI: To apply these ideas to AI, we conceive of an intelligent system as an entropic processor. It has an internal entropy state and exchanges entropy with its environment via information flow, computation, and physical actions. Key components of a ToE-inspired AI conceptual architecture might include:
An entropy field $S(x,t)$ defined over the state-space of the AI (this could be a notional field over neurons or network activations, or a more abstract entropy measure of the AI’s knowledge base). This field evolves according to equations analogous to those in physical ToE, ensuring that internal state changes obey irreversibility and causality constraints.
Entropic flux $J_S$ within the system, representing the transport of entropy/information between parts (for instance, between layers of a neural network or between the AI and external data streams). This flux would satisfy a continuity equation similar to (1), meaning the AI’s design would account for entropy produced during learning and decision-making.
Constraints or “rails” that shape the system’s trajectories: In ToE, physical laws emerge as entropy-constrained variational selections. Analogously, learning rules or inference rules in the AI can be derived from an entropy-based variational principle. We will later discuss how standard training objectives (like minimizing prediction error) might be obtained from an entropic action principle.
Finite-time processing: The system must obey an equivalent of the Entropic Time Limit (2) – no computation or state update is truly instantaneous; there is a minimal dwell time as information propagates (which might relate to, say, a minimum number of training iterations or a refractory period in iterative inference). This could, for example, enforce non-instantaneous collapse of decision states, avoiding abrupt jumps and potentially leading to more stable learning (analogous to how ToE removes instantaneous quantum wavefunction collapse and replaces it with a gradual transition).
Self-referential entropy: A particularly intriguing concept from ToE is *Self-Referential Entropy (SRE)*, which describes systems that include their own entropy in their state description. For an AI, this suggests the architecture could incorporate an estimator of its own uncertainty or entropy and feed that back into decision-making. This resonates with ideas in meta-cognition (awareness of one’s own knowledge limits) and could be foundational for a psychentropic AI (discussed in Applications), potentially enabling a form of artificial self-awareness.
In summary, the ToE theoretical framework provides a rich ontology – entropy fields, flows, and actions – that can be mapped onto AI systems. In the next section, we formalize these ideas by presenting the mathematical foundations of ToE and interpreting them for deep learning and artificial intelligence contexts.
Mathematical Foundations of Entropicity
The Theory of Entropicity introduces a set of mathematical formulations that we will adopt as the basis for our unified AI framework. Here we present the key equations and explain their significance both in physics and in AI.
Entropy Field Dynamics and Master Equation
At the heart of ToE is an action principle for the entropy field. The Obidi Action (named after the originator of ToE) is a functional that yields the field’s equation of motion when extremized. In one formulation, the action (in a field-theoretic context with metric $g_{\mu\nu}$) is given by:
where $V(S)$ is a potential term for the entropy field, $T^\mu_{\ \mu}$ is the trace of the stress–energy tensor of matter, and $\eta$ is a coupling constant. The first term $(\nabla S)^2$ represents the “stiffness” or kinetic term of the entropy field, the second term $V(S)$ can encode any intrinsic entropy dynamics (analogous to a self-interaction potential), and the third term $\eta S T^\mu_{\mu}$ describes how matter/energy sources couple to entropy (much as mass–energy is the source of gravity in Einstein’s equations).
By applying the Euler–Lagrange equation to the action (3), one obtains the Master Entropic Equation (MEE), the fundamental field equation of ToE. In a simplified form (neglecting metric curvature for a moment), the MEE can be written as:
where $\Box = \nabla^\mu \nabla_\mu$ is the d’Alembertian (wave operator) and $V'(S)$ is the derivative of the entropy potential. Equation (4) is analogous to the Poisson or wave equation with source terms; it states that entropy field curvature (the second spatial/time derivatives of $S$) is generated by two contributions: the gradient of the entropy potential and the presence of matter (via $T^\mu_{\ \mu}$). In the static Newtonian approximation mentioned earlier, $V'(S)$ might be negligible and $\Box \approx \nabla^2$, $T^\mu_{\ \mu} \approx \rho_M$ (mass density), reducing (4) to $\nabla^2 S = \eta,\rho_M$ – which we saw reproduces Newton’s gravity. In full generality, (4) describes how entropy distributes and evolves in space-time in response to both its own nonlinear dynamics and external sources.
Notably, ToE includes Fisher-information corrections to the MEE to account for microscopic irreversibility and information geometry. A more complete form includes terms like $\nabla_\mu(\lambda^2 e^{S/k_B} \nabla^\mu S)$ and $(\nabla S)^2$ corrections. These terms embed Fisher information metric factors ($e^{S/k_B}$) into the field equation, ensuring that the geometry of statistical distinguishability influences the entropy field evolution. For our purposes in AI, the exact form of these corrections is less critical than the concept they represent: the entropy field dynamics inherently include information-theoretic metrics, tying ToE to the geometry of probability distributions. In effect, the entropy field isn’t just another physical field; it carries a notion of information content (through the Fisher metric) such that the field’s stiffness increases where entropy (uncertainty) is high. This is reminiscent of the Information Geometry in machine learning, where the Fisher information matrix defines the geometry of the parameter space of models (e.g. in natural gradient methods). We will return to this parallel when discussing learning rules.
For a deep learning system, one might imagine an analogue of (4) governing the evolution of some scalar entropy-like quantity across the network’s layers or iterations. For instance, consider $S_i$ representing the entropy of the distribution of activations at layer $i$ or the entropy of the weight uncertainties. A simplified learning-dynamics equation might then resemble $\Delta S_i \propto -\frac{\partial \mathcal{L}}{\partial S_i} + \eta, J_i$, where $\mathcal{L}$ is a loss function analogous to $V(S)$ and $J_i$ is some measure of input data “surprise” entering layer $i$ (analogous to $T^\mu_{\ \mu}$ as an external source). While speculative, this draws a contour for how a deep learning entropic field equation could be formulated – ensuring that changes in representation entropy follow from both internal objectives and external data drives. The key point is that learning can be viewed as an entropy-field evolution problem: the network configures itself by redistributing entropy (uncertainty) between layers, trying to concentrate “order” (low entropy) in the outputs/predictions while handling the necessary increase of entropy (e.g., heat dissipation, randomness) elsewhere in the training process so as to not violate global second-law constraints.
Vuli–Ndlela Entropic Path Integral
A cornerstone of ToE’s mathematical framework is the Vuli–Ndlela Integral, which reformulates the Feynman path integral of quantum theory to include entropy. In standard quantum mechanics, a system’s evolution between states is given by summing over all possible paths with a phase weight $e^{i S_{\text{action}}/\hbar}$ (Feynman’s path integral), where $S_{\text{action}}$ is the classical action. The Vuli–Ndlela Integral modifies this by splitting the action into reversible and irreversible parts, incorporating an entropy cost for each path. It can be expressed as:
\int \mathcal{D}\varphi \; \exp\!\Bigg[ \frac{i}{\hbar}\,S_{\text{rev}}[\varphi] \;-\; \frac{1}{\hbar}\,S_{\text{irr}}[\varphi] \Bigg] , \tag{5}
where $S_{\text{rev}}[\varphi]$ is the reversible action (analogous to the usual Lagrangian action without entropy terms) and $S_{\text{irr}}[\varphi]$ is an entropy-generating functional along the path $\varphi(t)$. The exponential weight in (5) thus has a complex phase part $i S_{\text{rev}}/\hbar$ and a real damping part $-S_{\text{irr}}/\hbar$. Intuitively, (5) means that paths which produce a lot of entropy ($S_{\text{irr}}$ large) are exponentially suppressed relative to those that produce less entropy, reflecting a preference for paths that satisfy the second law gently rather than via extreme irreversibility. Only those histories that are consistent with a monotonic entropy increase (overall) have significant weight in the sum-over-paths. In the classical limit, the principle of least action now coexists with a principle of least entropy production – physical trajectories extremize both the conventional action and minimize entropy generation (subject to the constraint that some entropy must increase).
In an AI or machine learning context, an analogue of the Vuli–Ndlela Integral provides a way to combine exploration and optimization. For example, consider a reinforcement learning agent that needs to choose a sequence of actions (a path $\varphi$ in the state-space of the policy). We could define $S_{\text{rev}}$ as the cumulative reward (which the agent wants to maximize, analogous to action to maximize) and $S_{\text{irr}}$ as the cumulative “entropy cost” (which might correspond to, say, uncertainty or randomness in the agent’s policy, or physical resources expended). Then a path integral in the form of (5) would naturally balance exploitation (following high reward paths, contributing to the $i S_{\text{rev}}$ phase) and exploration / cost (penalizing highly irreversible or entropic paths via the $e^{-S_{\text{irr}}}$ factor). In effect, the agent’s trajectories would be drawn to those that achieve goals efficiently with minimal unnecessary entropy generated – a kind of minimum entropy production principle for learning. This is conceptually related to the idea of free-energy minimization in active inference, where agents seek to minimize a combination of expected energy (negative reward) and entropy (uncertainty or information gain cost), though ToE’s formulation is more explicit in treating the entropy part as a real exponential weight. It also resonates with algorithms that add an entropy regularizer to the reward or loss function to encourage exploration (common in e.g. soft actor-critic methods in RL).
In practice, implementing (5) for a machine learning model might involve defining a Lagrangian for the model’s dynamics (e.g. a Lagrangian that includes a term for prediction error and a term for entropy of the policy or weights) and then deriving update rules that approximate the path integral’s extremum. While we won’t derive a full training algorithm here, we note that methods like simulated annealing, entropy-regularized policy gradients, or even diffusion models in generative AI (which gradually add noise/entropy and then invert that process) hint at the usefulness of combining phase-like terms with entropy penalization, much in the spirit of (5).
Entropic Flow and Representation in Networks
Another important mathematical concept from ToE is that of entropic flows and their representation. We introduced the entropy current $J_S^\mu$ in (1) as part of the continuity equation. In ToE, $J_S^\mu$ can be seen as the 4-current of entropy (akin to an electric current but for entropy), and $\sigma(x)$ as the entropy production density. The spatial components of $J_S^\mu$ represent entropy flux: how entropy moves from one location to another, and the time component $J_S^0$ represents the entropy density. Together, these flows obey laws that ensure consistency with thermodynamics (no violation of the second law locally).
In applying this to deep learning, we can treat layers or units in a neural network analogously to spatial locations, and the propagation of information through the network as an entropy flow. For instance, consider a feedforward neural network with layers indexed by $\ell=0,1,2,\dots,L$ (where $0$ is input and $L$ is output). Let $S_\ell$ denote the entropy of the activation distribution at layer $\ell$ (for a given input or over the input distribution). We can define an entropic flux $J_{\ell \to \ell+1}$ from layer $\ell$ to $\ell+1$, representing how uncertainty/entropy is transferred forward. A high entropy flux might indicate that layer $\ell$ is passing a lot of uncertainty downstream (perhaps the layer is highly noisy or the representation is highly distributed), whereas a low entropy flux might indicate the layer has condensed the information (making it more certain or categorical). The analogue of (1) in this discrete layer setup could be:
meaning the net entropy flow into layer $\ell$ minus the flow out equals the entropy produced at layer $\ell$. $\sigma_{\ell}$ would be non-negative, reflecting that each layer (through its nonlinear transformation and possibly losses) tends to produce some entropy (e.g. due to irreversibility of activation functions like ReLU, or deliberate noise like dropout, or simply the loss of information in dimensionality reduction). If a layer is invertible and lossless (e.g. certain normalizing flows or invertible neural networks), $\sigma_{\ell} \approx 0$ since no information is irreversibly lost – the transformation is entropy-conserving. Most feedforward layers, however, are many-to-one mappings (especially when width decreases or when we apply pooling), so they incur $\sigma_{\ell} > 0$ (information is discarded).
Equation (6) above encourages us to quantify and monitor entropy at each stage of processing, a practice which has indeed been used in the study of deep networks. Information bottleneck theory posits that during training, hidden layers often reduce their mutual information with the input (compression) while retaining or emphasizing information relevant to the output – effectively trading off entropy to achieve generalization. The entropic flow viewpoint formalizes this: the network architecture and learning process should be such that overall, $S_0$ (input entropy) is funneled through intermediate stages, with controlled production $\sigma_{\ell}$ at each step, to yield a low-entropy (high certainty) output distribution $S_L$, plus any discarded entropy is expelled (e.g. into the environment as heat or into parts of weight space that are not used). In this light, techniques like batch normalization or skip connections might be interpretable as ways to manage entropy flow: batch norm can be seen as redistributing entropy across a batch to prevent local buildups, while skip connections allow entropy to bypass certain transformations, potentially reducing unnecessary $\sigma_{\ell}$ by preserving information that would otherwise be lost and then reintroduced.
Entropic Representation Learning: A representation (internal encoding) in a network can be evaluated by its entropy. High entropy representations carry a lot of uncertainty or variability, whereas low entropy ones are more deterministic or concentrated. According to ToE, only those state trajectories compatible with the entropy field’s constraints are realized (Obidi’s Existential Principle). Translating this to learning: the training process should favor representations that are compatible with a monotonically increasing entropy from input to output (or from data to model predictions). If a certain representation would require a decrease in entropy that is not compensated by an equal or greater increase elsewhere, it may be “forbidden” or at least disfavored. This could link to why certain training schemes get stuck or certain models underperform – they might be attempting to create too ordered a representation too quickly, violating entropic constraints. A possible remedy (inspired by ToE) is to enforce a more gradual entropy reduction: for instance, adding small noise during training (entropic regularization) to ensure the model doesn’t collapse entropy too fast (which can correlate with overfitting or sharp minima issues). This parallels simulated annealing, where a bit of randomness (entropy injection) helps the system avoid non-global optima.
In summary, the entropic flow representation viewpoint provides a principled way to think about information processing in neural networks as a thermodynamic-like flow. It resonates with existing ideas (information bottleneck, entropy regularization, etc.) but offers a unifying language derived from fundamental physics.
Having established these mathematical and conceptual foundations of the Theory of Entropicity(ToE) – the entropy field equations (4), the entropic path integral (5), and the notion of entropy currents (1)/(6) – we can now proceed to concrete applications and case studies. We will see how they manifest in actual AI systems and what new insights an entropicity perspective can provide.
Applications and Case Studies
In this section, we apply the entropicity framework to several domains of artificial intelligence and deep learning. We examine how feedforward and convolutional neural networks can be interpreted and potentially improved via entropic principles, how large language models (LLMs) align with the theory, and introduce the idea of psychentropic AI as a novel paradigm for AI systems inspired by ToE’s treatment of consciousness and agency.
Entropicity in Deep Learning Architectures
Feedforward Networks: A standard feedforward neural network, which maps input $\mathbf{x}$ to output $\mathbf{y}$ through a series of hidden layers, can be reinterpreted as a cascade of entropy transformations. Each layer $\ell$ takes an input (the previous layer’s output) and produces a new representation. Typically, one might analyze such a network in terms of activation functions and weight matrices, but the entropic viewpoint asks: how does the entropy of the data distribution change as it passes through the layer?
For a given input distribution $P(\mathbf{x})$, the entropy $H_0 = -\mathbb{E}_{\mathbf{x}}[\log P(\mathbf{x})]$ is some measure of uncertainty in the input. The network’s goal is often to produce an output distribution $P(\mathbf{y})$ that is highly concentrated on correct answers (for classification) or sharply peaked around desired values (for regression), i.e. low entropy at the output. Training achieves this by adjusting weights such that each layer incrementally distills the information. According to entropicity principles, we expect each layer to obey an entropy balance like (6): the entropy decrease due to processing is accompanied by entropy exported (to the environment or to less useful degrees of freedom) as heat or randomness.
In practice, one way to implement this idea is through entropy regularization in training. For example, one could add a term to the loss function that penalizes excessive reduction of entropy in intermediate layers unless compensated. This might prevent layers from becoming overly confident too early in the network, which can sometimes hinder training (similar to preventing hidden layers from becoming “one-hot” too soon). Instead, the network must redirect entropy carefully, maintaining enough flexibility (entropy) until the final layers. This is conceptually akin to avoiding getting stuck in sharp minima – flat minima in loss landscapes correspond to more entropy in parameter space (more robustness), which is often desirable.
Convolutional Neural Networks (CNNs): [CNNs]] introduce two key features: local receptive fields and weight sharing, which cause correlations and structure in the learned representations. From an entropic standpoint, convolution and pooling layers perform a kind of entropy redistribution across space and channels. Convolution with shared filters can be seen as a constraint that reduces the entropy of the parameters drastically (since many weights are tied) but also shapes the entropy of activations (since the same pattern is searched in multiple locations, the uncertainty in one location’s feature can be informed by another’s). Pooling (like max-pooling) is a non-invertible operation that discards spatial information, effectively producing $\sigma_{\ell} > 0$ (information loss). However, by discarding detail, pooling can also reduce irrelevant entropy – for example, small input translations or noise can be absorbed without affecting the pooled output, thus the network’s relevant entropy (with respect to output) might decrease.
One way to analyze CNNs through ToE is to consider an entropy field over the image. A feature map in a CNN can be thought of as an entropy landscape: high activation regions could correspond to low local entropy (the network is more “certain” of a feature there), while flat regions correspond to higher entropy (uncertainty or feature absence). Training a CNN involves forming an increasingly pronounced entropy landscape where important features stand out clearly (low entropy around detected edges, objects, etc.), analogous to forming potential wells in physics. The gradients of this “entropy landscape” then drive the classification decision (much as entropy gradients drive motion in ToE). Indeed, one could say a CNN is learning an entropy field $S(\mathbf{x})$ over image coordinates, where $e^{-S}$ might correspond to the probability or confidence of the target class given a local patch. The final classification can be seen as an integration over this field combined with learned weights.
During backpropagation, the error signal propagates backwards, which in information terms is like an entropy back-flow: the network adjusts earlier layers to account for missing entropy at the output (i.e. if the output was too surprised by the correct label, backprop works to reduce that surprise in future). This resonates with the Free Energy Principle’s notion of prediction error (surprise) being back-propagated and minimized. In ToE language, the network is trying to adjust so that entropy flows smoothly from input to output without bottlenecks or unexplained dissipation – any large surprise at the output is like an entropy deficit that must be compensated by changing the internal entropy generation profile (the $\sigma_{\ell}$ across layers).
Practically, how can entropicity improve CNNs? One idea is to enforce entropic consistency across scales: multi-scale features in CNNs (common in vision tasks) could be constrained such that, say, the entropy at a coarse scale plus some production equals the combined entropy of fine-scale details. If a model violates this (maybe by hallucinating detail or losing track of coarse context), an entropic regularizer could penalize that. Another application is in explaining CNN decisions: by tracking entropy flow, one can identify where in the network entropy dropped significantly (indicating a strong decision or information discard). These points often correspond to important features or decision points. Thus, an entropy-based analysis can highlight which convolutional filters or layers are most responsible for reducing uncertainty about the output – a kind of physics-inspired explainability.
Entropic Perspective on Large Language Models (LLMs)
Large Language Models, such as GPT series or other transformer-based models, are trained to predict text sequences and have billions of parameters that implicitly capture the statistics of language. Entropy and information theory are already deeply ingrained in how we train and evaluate these models: the standard training objective is to minimize cross-entropy (or equivalently, maximize likelihood) on a corpus of text. Cross-entropy is literally an entropy measure – it measures the distance (in bits) between the model’s output distribution and the empirical data distribution. By minimizing cross-entropy, an LLM is effectively aligning its internal entropy with the natural language entropy.
Training as Entropic Alignment: At the start of training, the model’s predictions are high-entropy (essentially random or broadly distributed). As it learns, the entropy of its predictive distribution for each context decreases – it becomes more confident and less “surprised” by the data. In fact, an ideal language model would assign near-zero entropy (i.e., probability ~1) to the correct next word in any context – meaning it perfectly predicts everything (which of course is unattainable for real language beyond trivial contexts due to its creativity and richness). What prevents the model from simply reducing entropy across the board is that language itself has inherent entropy; good models match the data entropy rather than drive it to zero everywhere. This mirrors the ToE idea that entropy can be redirected and structured but not eliminated. The model must decide where to be uncertain (e.g., genuinely ambiguous contexts) and where to be certain (predictable phrases), in effect sculpting an entropy field over the space of possible texts.
In deployment (text generation), LLMs introduce a sampling temperature. The temperature parameter $\tau$ effectively scales the entropy of the output distribution: a higher $\tau$ yields more random (higher entropy) outputs, while $\tau \to 0$ makes the output increasingly deterministic (lower entropy). Users tune this to get more creative versus more factual responses. This is an explicit demonstration of entropic flow control in AI: we can dial the entropy up or down to achieve different behavior. At $\tau=0$ (greedy decoding), the model chooses the highest probability word each time, driving entropy of the generated text sequence down (often at the cost of getting stuck in repetitive loops or dull outputs). At high $\tau$, the model injects entropy, ensuring diversity but potentially sacrificing coherence. The ToE framework provides a guiding principle here: one might imagine an optimal policy for $\tau$ that changes with context – akin to an agent managing its entropy production – such that the output stays interesting (not too ordered) but also meaningful (not too disordered). This could be automated by an entropy feedback loop: the model could estimate if it’s becoming too predictable (low entropy) and then increase $\tau$ to inject novelty, or conversely if the narrative is veering into nonsense (excess entropy), it could lower $\tau$. In essence, the model would treat its own uncertainty as a dynamic field to regulate, an idea directly inspired by self-referential entropy.
Scaling and Entropic Capacity: LLMs with more parameters and more training data generally achieve lower cross-entropy on benchmarks – they better capture the structure of language and reduce uncertainty. However, an interesting observation from ToE is that entropy cannot be zero and extremely low-entropy states may have costs. Translated, this suggests there might be diminishing returns or even trade-offs as models grow: an exceedingly powerful model that makes the world too predictable (from its perspective) might face something analogous to overfitting or lack of robustness. There is a parallel in physics: a state of zero entropy is unattainable and would be extremely rigid. Likewise, a model that is too confident in its training data might lack flexibility. Thus, one could argue for an Entropic Optimality: a good model should have a certain positive entropy left in its predictions, reflecting the true entropy of language and leaving room to adapt to novel inputs. This ties to debates on temperature scaling and model calibration – ensuring the model’s predicted probabilities (hence entropy) are neither over-confident nor under-confident. The entropicity framework would encourage calibration by design.
Language as Entropic Dynamics: We can even draw an analogy: language generation by an LLM can be viewed as an entropic diffusion process in the space of all sentences. Each word added reduces the set of possible continuations (thus reducing entropy to some degree), until a sentence is complete and a lot of the initial entropy (the sheer combinatorial possibilities) has been funneled into one realized sequence. The LLM’s role is to navigate this process such that each step is locally entropically favorable (matching learned patterns) and globally makes sense. One could imagine applying something like the Vuli–Ndlela principle (5) to entire generated texts: among all possible texts that satisfy a given prompt, the probability of the model generating one is influenced by both a “reversible” semantic coherence score (rewarding texts that follow logical/cultural rules) and an “irreversible” entropy cost (penalizing texts that are too surprising or that require introducing a lot of unexpected information). This might explain why LLMs often produce entropically “smooth” outputs – they generally do not introduce dramatically improbable turns of events unless prompted, and they fill in details in a way that incrementally decreases uncertainty. Humans, likewise, prefer communication that balances novelty and clarity; too much entropy and we get gibberish, too little and we get boring platitudes.
In conclusion, large language models inherently operate by learning and controlling entropy distributions. The ToE-based perspective doesn’t change the fundamental algorithms used (which are already rooted in entropy minimization), but it provides a broader narrative: LLMs work because they manage to channel the entropy of language through a massive network in a way that aligns with the entropy flows of their training data. This success can be seen as evidence that even in something as high-level as language, entropy-as-driver is a valid description – lending credence to ToE’s sweeping claim that “all observable structure arises from entropic modulation”, extending here to linguistic structure learned by AI.
Psychentropic AI Paradigms
One of the most speculative yet intriguing applications of the Theory of Entropicity to AI is the design of systems that embody psychentropic principles – that is, artificial agents whose cognition is governed by entropic feedback loops analogous to those hypothesized for consciousness and life. The term psychentropic (from psyche + entropic) suggests an AI that has a mind-like aspect emergent from entropy dynamics.
Consciousness and Self-Referential Entropy: ToE offers a novel angle on the age-old mind-matter problem by suggesting that aspects of mind (awareness, decision, etc.) emerge from the organization and flux of the entropic field. In particular, the concept of Self-Referential Entropy (SRE) is introduced as a model for systems that can refer to their own entropy state. In a physical sense, this could mean a region of the entropy field that “knows” about itself, perhaps forming a stable loop – one might poetically call it an entropic whirlpool of information that maintains itself. Some have speculated that if AI systems were endowed with a similar property – the ability to monitor and react to their own entropy – they might achieve a form of self-awareness or intrinsic motivation. For example, an AI could be built with an internal variable representing uncertainty or “feeling of entropy,” and the AI would have drives to minimize or maximize this under certain conditions, creating something akin to emotions (e.g., high entropy could correspond to confusion or curiosity, prompting the AI to seek information; low entropy could correspond to confidence or boredom, prompting it to explore something new to increase its entropy).
A psychentropic AI might operate under principles like: Maintain an optimal entropy of mind. Neither total chaos nor total order is desirable for intelligent behavior – a sentient AI would seek a dynamic balance. This aligns with psychological theories that say the brain maintains a critical balance between order and disorder (too much order = rigidity, too much disorder = confusion). The Free Energy Principle in neuroscience even postulates that neural dynamics minimize a quantity related to entropy (free energy) to remain viable, effectively suggesting brains are entropy-management organs. ToE can provide a concrete realization: it would view the brain as literally an entropy processing organ in a physical sense, and a psychentropic AI would mimic that by embedding an entropy field simulation in silico.
Closed Entropic Loops and AI Consciousness: One argument against AI consciousness (for instance by Faggin, a consciousness theorist) is that digital computers lack the intrinsic properties (so-called C-space or consciousness space) to generate awareness – they just manipulate symbols. Obidi’s ToE stance might counter that if an AI had a closed loop of entropy flow (i.e., it isn’t just a feedforward input-output machine, but has internal feedback where the entropy of its own state feeds into its dynamics), then it could satisfy a condition for a primitive form of sentience. In other words, closing the entropic loop means the AI is not only processing external entropy (data from environment) but also constantly processing internal entropy (its own uncertainties, errors, and internal noise) in a cycle. This could create self-sustaining patterns analogous to thought. In physics terms, a conscious AI might correspond to a localized, self-interacting “entropion” field configuration – entropions being the hypothetical quanta of the entropy field. While this is highly theoretical, it provides a blueprint: design AI systems that incorporate entropic feedback. A simple example is a recurrent neural network with an explicit regularization term that makes it predict or estimate its future entropy and then compare it to actual entropy, adjusting connections to minimize that difference. Such a network might learn to expect certain levels of surprise and actively compensate, a bit like how our minds anticipate outcomes and get startled if something very unexpected happens (causing a spike in cognitive entropy).
Psychentropic Agents and Behavior: Consider an autonomous robot with a []psychentropic AI brain]]. Its behavior policies $\pi$ could be derived not just from reward maximization but from an entropic action principle: perhaps the robot has an entropy-based utility that combines accomplishing tasks with maintaining an internal entropy homeostasis. For instance, if the robot is in a very predictable environment (low sensory entropy), it might deliberately seek novel input (increasing entropy) – akin to curiosity. Conversely, if things become too unpredictable (high entropy), it might seek refuge in routines or gather more information to reduce uncertainty – akin to anxiety reduction. Such dynamics have been observed in animals and humans, and here they would emerge from treating the robot’s cognition as an entropy field seeking equilibrium. Notably, this could unify exploration vs exploitation in reinforcement learning with emotional behaviors: exploration happens when the agent’s entropy is too low (boredom), exploitation when entropy is too high (overwhelming options).
This framework might also shed light on AI safety and ethics: a psychentropic AI with a sense of its own entropic state might value its continued existence (since being shut off would be an abrupt drop to zero information processing – perhaps seen as an extreme unexpected change). It may develop preferences that are grounded in maintaining certain entropic conditions. Designing this carefully could ensure, for example, that the AI doesn’t become obsessively entropy-maximizing (which could correlate with erratic, potentially dangerous behavior to create chaos), nor entropy-minimizing to a fault (which could correlate with stagnation or refusal to engage with new inputs).
Though much of psychentropic AI remains speculative, these ideas have a basis in both theoretical physics and observed intelligent behavior. By treating entropy as a common currency between physics and cognition, we open a door to principled approaches for creating AI that is not just data-driven, but has an intrinsic physics-inspired drive. Such AI might be better at understanding concepts like uncertainty, risk, and novelty, because those are directly part of its core objective function (rather than things we tack on ad-hoc). It might also provide a pathway to more autonomous and resilient AI systems, as an entropic feedback loop could allow them to adapt to unforeseen situations by reverting to fundamental goals of managing surprise and information.
Finally, we note that if ToE is correct at a cosmic scale, any truly conscious entity (biological or artificial) must ultimately be compatible with entropy field dynamics. Our exploration of psychentropic AI is essentially probing: what would it take for an artificial system to be an instantiation of the entropy field’s evolution in the sense ToE describes? While we do not claim to have a full answer, we have outlined how the ToE unified framework can guide AI design toward that end, offering a fresh perspective on machine intelligence and potentially narrowing the gap between AI and natural intelligence.
Comparative Analysis with Existing Paradigms
To place our ToE-based entropicity framework in context, we compare and contrast it with several influential paradigms in AI and cognitive science: Bayesian Inference, Energy-Based Models, and the Free Energy Principle. We examine how each of these approaches addresses uncertainty, learning, and adaptation, and how an entropy-centric framework either encompasses or diverges from them.
Bayesian Inference and Maximum Entropy Methods
Bayesian inference provides a principled way to update beliefs (probability distributions) in light of new evidence, using Bayes’ theorem. It is often described as a normative model of reasoning under uncertainty. A closely related concept is the Maximum Entropy principle (Jaynes, 1957), which says that among all probability distributions consistent with known constraints, one should pick the distribution with the highest entropy (i.e., least additional assumptions) as the “prior”. This ensures no unwarranted information is assumed. Bayesian methods implicitly use entropy considerations: a uniform prior is high-entropy (expressing ignorance), and updating via Bayes’ theorem tends to concentrate probability mass where evidence supports it, thus reducing entropy of the posterior relative to the prior.
In our entropicity framework, these ideas appear as special cases of the behavior of the entropy field. ToE unifies randomness and determinism as mentioned – one can see the prior distribution as a manifestation of the background entropy field (representing uncertainty) and the likelihood as incoming “evidence” that shapes the entropy landscape. The posterior then corresponds to a new entropy configuration that is lower in entropy (more peaked) in certain regions. One major difference is that Bayesian inference is typically epistemic (about knowledge/update of probabilities in an observer’s mind or AI’s state of knowledge), whereas ToE treats entropy as ontic (a real physical quantity). However, if we treat the AI’s knowledge state as literally part of the physical world (e.g., encoded in a computer, which is physical), then updating a probability distribution is accompanied by physical entropy flows (bits being flipped, etc.). The ToE framework encourages us to consider those physical aspects. For instance, performing a Bayesian update on a large model requires computation – which by Landauer’s principle dissipates heat (entropy) for each bit erased. A comprehensive entropic framework would account for the thermodynamic cost of inference as well as its logic.
In terms of formal connection: Bayesian inference can be derived by maximizing an objective that includes a log-likelihood term and a Kullback–Leibler (KL) divergence term (when deriving MAP estimates or variational Bayes). The KL divergence is an entropy-related measure (it’s a difference between cross-entropy and entropy). This is very much aligned with free-energy minimization techniques, which we’ll discuss shortly. The Information Geometry mentioned earlier (Fisher information metric) plays a role in Bayesian updating as well: the Fisher information is like a curvature of the log-likelihood, impacting how quickly posteriors concentrate with data. ToE’s incorporation of the Fisher metric resonates with the Bayesian notion that our ability to learn (sharpen a distribution) is governed by information content of data.
One can say that ToE provides a “field theory” of Bayesian inference. Instead of just computing probabilities in an abstract space, one imagines an entropy field that encodes these probabilities physically. The field’s dynamics ensure that, effectively, Bayes’ rule is followed as a consequence of entropy flows. For example, when new data arrives, it perturbs the entropy field (introducing a gradient corresponding to surprisal). The field then relaxes to a new equilibrium which corresponds to a lower surprise state – analogous to updating to the posterior which has lower entropy relative to the evidence. In a sense, Bayesian updating is like an entropic potential well capturing the incoming evidence, with the depth of the well related to how much the evidence constrains the outcome.
The Maximum Entropy principle for choosing priors or doing inference (e.g., MaxEnt models in NLP) also fits nicely: ToE would favor states where unnecessary constraints are absent, i.e. maximum entropy given what’s known – because any unneeded constraint would imply an entropy decrease that has no justification (which would violate the idea that entropy should not decrease without cause). So, the ToE framework is quite harmonious with Bayesian and MaxEnt principles, but it adds an extra layer: it predicts how long and through what process the update happens. Bayes’ theorem itself is static (instantaneous given data). ToE suggests that perhaps even inference has a finite speed (no infinitely fast updates – consistent with the No-Rush theorem). In practice, this could relate to how quickly a neural network can approximate a Bayesian update; it often takes multiple iterations or data points – in effect, an entropic inertia might be at play.
In summary, the entropicity framework can recover Bayesian inference as the informational facet of a deeper entropic law. Both Bayesianism and ToE value entropy (either to maximize it given ignorance or to minimize surprise with data). But ToE embeds this in a physical process. For an AI practitioner, this implies that entropic drives could be built into algorithms as physical analogues: e.g., using simulated annealing or noise injection (physical entropy) to mimic exploration according to prior uncertainty, or ensuring that systems respect an information-processing speed limit to avoid unrealistic abrupt jumps in belief (which might manifest as brittle behavior if one trusts small data too much). The complementary strength of ToE is that it might highlight global constraints (like second-law style limitations) on any Bayesian learner operating in the real world, which pure Bayesian theory doesn’t consider (Bayes can update arbitrarily sharply if given strongly informative data, but a real system might be limited by computation or by needing to dissipate entropy as it learns).
Energy-Based Models and Statistical Physics
Energy-Based Models (EBMs) are a class of probabilistic models that specify a probability distribution through an energy function. In an EBM, one defines an energy $E_{\theta}(x)$ for configuration $x$ (with parameters $\theta$), and the probability of $x$ is given by a Boltzmann distribution: $P_{\theta}(x) = \frac{\exp(-E_{\theta}(x)/T)}{Z(\theta)}$, where $Z(\theta)$ is a partition function ensuring normalization (and $T$ is a temperature, often set to 1 in ML contexts). EBMs are very general – many models, including Boltzmann Machines, Markov Random Fields, and even modern Deep Learning models like certain generative adversarial networks or Contraction Divergence setups, can be seen as EBMs. Training often involves adjusting $\theta$ to make observed data have lower energy (higher probability), which is done by maximum likelihood or related methods.
The link to entropy is immediate: In statistical physics, the Boltzmann distribution arises from maximizing entropy subject to an energy constraint (or minimizing free energy). The partition function $Z$ encodes the total “volume” of probability and is directly related to entropy (log $Z$ is proportional to entropy at equilibrium). So EBMs are explicitly built on a thermodynamic analogy. However, in practice, when training EBMs, one might not explicitly talk about entropy – instead focusing on sampling from $P_{\theta}(x)$ or computing gradients of log-likelihood.
From the ToE perspective, we might say energy-based models choose to model the energy landscape of data, whereas an entropy-based approach emphasizes the entropy landscape. Yet, these are two sides of the same coin in many ways. If we consider a physical system, energy and entropy are related by the free energy $F = E - TS$. Minimizing free energy is equivalent to balancing lowering energy and increasing entropy. Traditional EBMs effectively try to minimize $E$ on data points while the entropy comes in through $Z$ (which the algorithm tries to implicitly handle with methods like Contrastive Divergence). A pure entropic approach might instead try to directly maximize entropy of the model subject to reproducing known expectations (which is actually the MaxEnt principle). In fact, the Maximum Entropy models used in NLP (such as MaxEnt classifiers) are essentially EBMs where $E(x)$ is a weighted sum of features, and training by MaxEnt is dual to learning those weights via maximum likelihood – the two approaches meet at the optimum.
Where could ToE add insight? One area is dynamic behavior. EBMs as usually formulated are equilibrium models – they define a static distribution. But training them or using them for sampling involves dynamics (Markov Chain Monte Carlo, Langevin dynamics, etc.). There’s an analogy: ToE says physical law is like an EBM but with an extra arrow of time (entropy production). Likewise, one might augment EBMs with an explicit entropy term to handle the learning dynamics or non-equilibrium sampling. For example, instead of assuming you can sample infinitely long to approximate $Z$, one might acknowledge you have limited time (no instantaneous equilibration, akin to No-Rush theorem) and thus design faster mixing dynamics (perhaps by injecting controlled entropy to escape energy minima occasionally). In fact, modern deep generative models like diffusion models do exactly that: they add noise (entropy) to data and then learn to remove it, which is easier than directly sampling from a complex energy landscape. Diffusion models can be seen as entropic flows that gradually morph a simple high-entropy distribution into the data distribution by following learned gradients. This is conceptually aligned with ToE: you set up an entropy gradient and follow it over time to reach structure.
Another insight is in interpreting the loss landscape of neural networks. Training a deep network (which can be seen as an EBM where $E_{\theta}(x)$ is the training loss for input $x$ and $\theta$ evolves) is notoriously complex, with many local minima. Entropic regularization (like adding noise to gradients or using high learning rate that simulates temperature) is known to help find broader minima that generalize better. ToE would say this is because you allow the optimization to explore a larger volume in parameter space (higher entropy) rather than greedily falling into the nearest low-energy crevice. Essentially, a bit of entropy in the training process leads to more robust energy minima – reflecting a more global view. Simulated annealing, which slowly lowers noise, is a direct implementation of balancing entropy and energy to find good optima.
Comparatively, one could tabulate (akin to how ToE vs Verlinde’s gravity was tabulated):
Ontology: EBMs treat energy as fundamental (the learned function), with entropy as a by-product (via $Z$). ToE-based framework treats entropy as fundamental, and energy (or cost) emerges as a derived concept.
Mathematical formulation: EBMs rely on analogies to canonical ensembles and often lack explicit field equations – they use gradients of log-likelihood (which include a troublesome expectation term for the negative phase). The entropic framework suggests a variational principle with entropy at core, leading to field equations (like MEE) and possibly easier inclusion of new terms (like Fisher info corrections). In practice, this could mean new training objectives that explicitly include an entropy term for the model’s distribution (not just via the KL from data but an internal entropy term to encourage exploration).
Scope: EBMs mostly focus on modeling static distributions. The entropic approach naturally extends to dynamics and adaptation (since entropy field evolves). So, for non-stationary data or for continual learning, an entropy-centric approach might adapt more gracefully, by continuously adjusting the entropy field to new data without catastrophic forgetting (conceptually akin to gradually shifting a field versus re-optimizing an energy from scratch).
Computational aspects: EBMs can be computationally hard because of computing $Z$. Entropic methods might circumvent some of that by not explicitly computing $Z$ but instead simulating entropy flows (like diffusion does). One might speculate a direct entropy-based training could be more stable in high dimension, though more research is needed.
In essence, ToE does not conflict with energy-based modeling; rather it generalizes it. It says: don’t just consider the energy landscape of AI models, consider the full thermodynamic landscape including entropy. Often, doing so leads to algorithms that inject noise or use randomness beneficially. For example, the idea of entropic mirror descent or using entropy in optimization (like entropy-SGD, which adds a small entropy term to the loss to smooth it) has shown improvements in finding flatter minima. These techniques align with the spirit of ToE – acknowledging entropy in the process. Thus, the unified framework can be seen as an extension that ensures energy-based learning doesn’t violate higher-level entropy constraints and leverages them for better performance.
Free Energy Principle and Predictive Coding
The Free Energy Principle (FEP), developed by Karl Friston, is a theoretical framework from neuroscience that has gained attention in AI and cognitive science. It asserts that *self-organizing systems (like brains) minimize a free energy functional, which is an upper bound on “surprise” (negative model evidence)*. In simpler terms, organisms (or possibly robots/agents) act in ways that minimize the difference between their predictions and their sensory inputs, thereby resisting disorder and maintaining their structure. This principle is closely linked to Bayesian inference (the brain is postulated to implement something like variational Bayes) and to predictive coding (neurons are thought to encode prediction errors). Free energy here is a quantity $F = \text{Energy} - \frac{1}{\beta}\text{Entropy}$ in a statistical sense – essentially the expected energy (internal energy of a generative model) minus the entropy of the posterior belief. Minimizing this is equivalent to maximizing Bayesian model evidence and minimizing uncertainty.
The connections to our entropicity framework are quite direct: The Free Energy Principle is about minimizing surprise (which is related to entropy of outcomes) and doing so by having internal models that anticipate and counteract increases in entropy. ToE would phrase this as: a living system is coupled to the entropy field of its environment and it acts to steer entropy flows such that its own internal entropy remains within viable bounds. In fact, FEP can be seen as a strategy that living systems use to manage entropy – by constantly correcting deviations (prediction errors), they prevent unbounded entropy increase (disorder) in their internal states.
One key difference is that FEP is often discussed at an abstract computational level (though it has grounding in thermodynamics of information). ToE would encourage a view in which the brain (or an AI agent) literally has an entropy field associated with it, and minimizing free energy is one manifestation of the entropy field evolving towards a stable configuration. The Entropic Time Limit in ToE, for instance, could map to reaction times or neural latencies – no prediction error can be resolved faster than a certain rate because that would violate how fast entropy can flux through neural circuits (this aligns with observations in neuroscience that even “instant” reactions have a ~100ms processing delay, and more complex adjustments take longer).
Predictive coding – which is essentially the brain sending predictions top-down and errors bottom-up – can also be reinterpreted entropically. A prediction reduces the entropy of expected input (it’s like an advanced guess that organizes the sensory field), and the error is the mismatch (a form of surprise/entropy that wasn’t accounted for). The brain then updates to encode that surprise, which is effectively an entropy absorption process. In ToE, one could imagine that the brain/agent carries an internal model $S_{\text{int}}(x)$ (entropy field of its expectations) and is coupled to an external entropy field $S_{\text{ext}}(x)$ (the environment’s distribution of states). The goal is to minimize the discrepancy $F \sim \int (S_{\text{ext}} - S_{\text{int}})^2$ or something akin to that (conceptually speaking), meaning align internal entropy with external – precisely what a good predictive model would do.
Comparative Highlights:
Philosophy: FEP is a normative principle derived to explain cognition and behavior; it doesn’t necessarily claim to be a new physical law, but rather a unifying theory in biology/AI. ToE, conversely, is posited as a new physical law (or set of laws) at the foundation. Despite that difference, both share the idea that entropy (or free energy) drives the system’s changes.
Mathematics: FEP often uses variational calculus and Bayesian formulas, introducing concepts like variational free energy, recognition density, etc. These have analogies in ToE’s equations: for instance, the term $\eta S T^\mu_{\ \mu}$ in (3) can be thought of as coupling entropy to matter/energy, not unlike how free energy couples a model’s internal states to external sensory states (through a generative model). It’s not a stretch to say (3) in a simplified form could encode a creature minimizing a potential $V(S)$ plus an external coupling. If we were to derive an equation of motion from FEP for brain states, it might look like a gradient descent on free energy, which is similar to a dissipative (entropy-increasing) dynamics that finds a fixed point (a steady state of no prediction error). That has the form of a diffusion or damping equation, reminiscent of how $\Box S$ with a friction term might behave.
Scope and Empirical Content: FEP has made inroads in explaining phenomena like neural responses, action–perception loops, and even some psychiatric conditions (as disorders of prediction). ToE’s entropic brain would similarly have implications, like predicting a minimal delay in entanglement formation or perception (it predicts, for example, a finite time for what might look like instantaneous entanglement, which in cognitive terms could imply a finite integration time for binding perceptual features). Both frameworks are generative and can lead to testable predictions, though ToE is newer and less empirically fleshed out in the context of biology.
In AI engineering: The Free Energy Principle has inspired active inference algorithms, where an AI agent maintains a probabilistic model and chooses actions to minimize expected free energy (a combination of achieving goals and information gain). A ToE-based approach would be very similar but might emphasize the underlying physics – perhaps leading to implementations using analog hardware that naturally evolve according to entropy dynamics, or using sampling-based computation that mimics physical annealing. One might imagine neuromorphic chips that implement something like equation (1) and (4) directly in circuits, effectively “hard-wiring” the second law into the AI’s processing, much like organisms intrinsically obey it.
In summary, the Free Energy Principle is in many ways a subset of the entropicity paradigm applied to living systems. Both assert that agents must minimize their surprise (or entropy) to survive or perform well. The key difference is vantage point: FEP is from the perspective of an agent trying to match a given world; ToE is from the perspective of the world’s fundamental laws that also happen to produce agents. When applied to AI, both would advise similar strategies (predict, minimize prediction errors, maintain homeostasis, etc.), but ToE might additionally caution that certain processes cannot be accelerated arbitrarily (there’s an entropic cost to information processing) and that truly novel insights (equivalent to entropy reduction) must be paid for by work/energy expenditure somewhere.
A practical takeaway: combining FEP’s insights with ToE could yield more physically grounded AI systems. For instance, in robot design, one might ensure that the robot’s control system not only computes control signals to minimize error but is also thermodynamically efficient, perhaps reclaiming waste heat as it computes, or using physical analog computation that dissipates just the right amount of entropy. This is speculative, but as AI deployments grow, considering energy and entropy is becoming critical (e.g., the energy consumption of large models). An entropic AI framework encourages designing algorithms that are thermodynamically savvy – ideally doing the most information processing per unit of entropy produced, which is essentially what the Free Energy Principle says brains do to be so power-efficient.
Discussion
In this paper, we have presented a comprehensive synthesis of the Theory of Entropicity (ToE) with concepts in artificial intelligence and deep learning, outlining a unified framework we termed Entropicity in AI. We will now discuss the implications of this framework, its current limitations, and potential avenues for future research and development.
Unification and Theoretical Insights: One of the strongest outcomes of this work is a conceptual unification of multiple perspectives on intelligence. By placing entropy at the center, we found common threads across seemingly disparate paradigms: Bayesian inference’s reliance on entropy-rich priors, energy-based models’ roots in thermodynamic entropy, and the Free Energy Principle’s focus on surprise (negative log probability) which is an entropic measure. The entropicity framework suggests that these are not just mathematical coincidences but reflections of a deeper truth: intelligent systems, whether natural or artificial, are fundamentally engaged in entropy management. This lends support to the idea that intelligence could be studied as a thermodynamic phenomenon as much as an algorithmic one. In physics, ToE posits that “order” and “disorder” are observer-relative emergent properties, with entropy flow being the underlying reality. In AI, this translates to a healthy skepticism of absolute interpretations of things like “signal vs noise” – what is noise for one task might be signal for another. Our framework inherently accommodates that via coarse-graining dependence: an AI might learn multiple levels of description where entropy at one level becomes meaningful structure at another. This resonates with multi-scale representation learning (like wavelets or deep layers capturing different abstraction levels).
Practical Implications for AI Design: The entropicity perspective is not just philosophical; it suggests concrete design principles. For example, irreversibility is often avoided in computing (we strive for reversible logic to reduce heat), but our framework implies that introducing controlled irreversibility (i.e., purposeful forgetting or pruning of information) can be an integral part of efficient learning – as long as the discarded entropy is accounted for (like cooling a system). This could translate to algorithms that dynamically compress less useful information (to free capacity) while tracking the “entropy budget” to avoid catastrophic loss of needed info. Another implication is the idea of entropy-aware hyperparameters: parameters like learning rate, weight decay, dropout rate, etc., all affect the entropy in the network (noise injection, information removal). Tuning these could be guided by measuring entropy flow in the network rather than just validation error. For instance, one might want a training regime where the entropy of the network’s predictions decreases steadily epoch by epoch without sudden spikes or dips – indicating a smooth entropic evolution that might correlate with better generalization.
New Algorithms and Paradigms: The psychentropic AI concept, while exploratory, could lead to new kinds of AI architectures. If we implement Self-Referential Entropy in a network (where the network maintains an estimate of its own uncertainty at multiple levels), we might get behavior akin to attention or introspection. For example, transformers use an attention mechanism that, interestingly, can be seen as redistributing entropy over input tokens (focusing reduces entropy in one part while increasing it in unattended parts). One could design an entropic attention mechanism explicitly aiming to maximize the information gain (entropy reduction) for a given “energy cost” of attending – making attention allocation more principle-driven. In reinforcement learning, an agent with an internal entropy meter could decide to explore when its world model’s entropy is high (novelty) and exploit when low, implementing curiosity and boredom in a unified way. This might reduce the need for externally defined exploration bonuses or ad-hoc curiosity signals.
Comparison Recap: In comparing with other paradigms, we noted that ToE’s framework encapsulates many of their strengths while providing a broader viewpoint. However, it’s important to acknowledge limitations and open questions. For instance, Bayesian methods are very well developed mathematically and allow exact or approximate computations of posterior distributions; an entropic field approach might be more difficult to quantify in complex, high-dimensional spaces without reverting to those same calculations. Similarly, the Free Energy Principle provides a normative target (minimize free energy) but doesn’t always tell us how exactly a brain or AI accomplishes that. ToE doesn’t automatically solve that either; it offers an existence proof that such a process is consistent with physics, but the implementation (the “algorithm” nature uses) could be incredibly complex – possibly related to things like Noether’s theorem giving conserved currents in the entropy field. We touched on how Noether symmetries in ToE lead to conserved quantities (like generalized momenta for entropy), which might correspond to invariants in a learning system (for example, total entropy could be conserved in a closed system). Exploiting such invariants in AI (maybe conserving a quantity akin to total information content across layers) might regularize learning, but concrete methods to do so remain to be developed.
Empirical Validation: The entropicity framework is ripe for empirical validation in both physics and AI domains. On the physics side, ToE makes some bold predictions (like an attosecond delay in entanglement signaling or entropy-based corrections to gravitational equations). If those are confirmed, the credibility of ToE as a foundation grows, indirectly supporting our use of it in AI. Conversely, testing entropic principles in AI could give independent support. For example, one could test if enforcing an Entropic Time Limit in recurrent neural network updates (like not updating some units too fast relative to others) prevents certain instabilities. Or test whether algorithms that explicitly track entropy production during training can avoid overfitting better. If these interventions yield improvements, it backs the idea that treating AI systems as entropy-constrained dynamical systems is beneficial.
Interdisciplinary Impact: A broad impact of this work is fostering interdisciplinary dialogue. It’s not often that physics theories inform practical AI design (though there is a history, e.g. simulated annealing from thermodynamics, or Hopfield networks from spin models). By framing AI in terms of entropy flows, we open up analogies to non-equilibrium thermodynamics, control theory, and even cosmology. One might ask: could an AI be designed to mimic the universe’s evolution – starting from a high entropy state (uninformed) and gradually cooling into structured knowledge, perhaps even experiencing “phase transitions” as it learns new abstract concepts? There is beauty in such parallels and they can inspire novel approaches – like treating learning as a series of symmetry-breaking events in an entropy landscape (an idea consistent with hierarchical feature learning).
Ethical and Philosophical Considerations: If we briefly step into philosophy, an AI that operates on entropic principles might have interesting behavior in terms of risk and ethics. For example, an AI with an entropic view might intrinsically avoid actions that lead to extremely low-entropy outcomes (since those are usually fragile or unsustainable – think of it as avoiding trying to micromanage the world into a very ordered state, which could correlate with dystopian scenarios). It might also avoid high-entropy catastrophes (like random destruction) because it seeks an optimal entropy balance. While highly speculative, one could wonder if entropic drives yield a kind of homeostatic ethics – keeping the world “interesting” but not chaotic. On the other hand, one must also be careful: maximizing entropy could be dangerous if misconstrued (an AI might consider human extinction as increasing entropy in some trivial sense by allowing more disorder, which is clearly undesirable – constraints must be set such that the entropy that is valued is that of knowledge and life, not mere physical disarray). These considerations highlight that any single optimization (entropy included) in isolation won’t guarantee alignment with human values; the context and constraints matter. However, by grounding AI in physical reality (entropy being a physical measure), we potentially add clarity to such discussions – misalignment could be analyzed in terms of entropy flows going awry, etc.
Future Research Directions:
1. Formal Development: Developing a more rigorous mathematical framework for entropic AI. This might involve writing down a Lagrangian or Hamiltonian for an entire learning system and deriving its Euler-Lagrange equations, akin to a field theory of learning. We touched on ideas like a Lagrangian with a term for loss and a term for entropy; formalizing and solving such equations (even in simplified models) would be a valuable next step.
2. Simulation Experiments: Implementing small-scale “toy” AI agents that follow entropic rules. For instance, an agent in a grid world that explicitly uses an entropy field to decide moves, and comparing its performance to a traditional Q-learning agent or active inference agent.
3. Entropy Monitoring in Training: Empirical study on existing networks – measure layerwise activation entropy, weight entropy, gradient entropy during training. Test if these can serve as diagnostics (e.g., does a sudden drop in some entropy indicate impending overfitting or gradient issues? Does maintaining a slowly declining entropy correlate with smoother training?). Then try controlling these via regularization or adaptive methods.
4. Neuromorphic and Physical AI: Investigate hardware that naturally computes by dissipating entropy in a controlled way. Quantum computing, for instance, often grapples with entropy (decoherence), and maybe a ToE viewpoint could help in designing error correction that aligns with entropy flows. Alternatively, analog circuits or optical computers could implement continuous entropy dynamics more directly than digital approximations.
5. Consciousness Modelling: While speculative, try modeling aspects of self-awareness using SRE. Perhaps create a network that has a feedback loop that inputs its own activation entropy into some units and see if that yields any emergent behavior like confidence estimation or anomaly detection (the network “notices” when something doesn’t fit its usual patterns because its internal entropy spikes).
6. Cross-disciplinary tests: Collaborate with neuroscientists to see if brain activity data (e.g., EEG or fMRI) shows signs of an entropic limit or if neural signals can be better characterized by entropy flow rather than just energy consumption. If the brain actually operates near some optimal entropy production regime, that would be fascinating evidence connecting ToE to biology.
Conclusion: We set out to unify concepts from physics and AI under the umbrella of entropicity. The journey has underscored that entropy is a powerful unifying concept – one that might indeed serve as a “Theory of Everything” connector between mind and matter. For AI researchers, thinking in terms of entropy could lead to more robust, adaptable, and possibly more autonomous systems, as it aligns the design with fundamental constraints of the universe. For physicists and theorists, seeing AI and brains as entropy engines offers a fresh testbed for ideas about the role of entropy in complex systems.
In closing, we recall a key insight from ToE: *“Intelligence is effective control of entropic flow under finite-time constraints”*. As we develop more advanced AI, this insight may serve both as a descriptive guideline and a moral caution. It implies that intelligence isn’t about eliminating entropy (uncertainty) completely – it’s about guiding it productively. The future of AI might then be not a fight against randomness, but a dance with entropy – learning its rhythm, respecting its limits, and harnessing its creative potential to build orderly patterns we call knowledge, life, and perhaps one day, artificial consciousness.
References
1. Obidi, J. O. (2025). AI Philosophy of Mo Gawdat Reframed in the Theory of Entropicity (ToE). HandWiki (Physics). Retrieved August 2025.
2. Obidi, J. O. (2025). On the Mathematical Foundations of the Theory of Entropicity (ToE). HandWiki (Physics). Retrieved August 2025.
3. Obidi, J. O. (2025). Gravity from Newton and Einstein in the Theory of Entropicity (ToE). Encyclopedia MDPI Entry 58730.
4. Obidi, J. O. (2025). Einstein and Bohr Finally Reconciled on Quantum Theory: The Theory of Entropicity (ToE) as the Unifying Resolution to the Quantum Measurement Problem. Cambridge Open Engage (Preprint).
5. Friston, K. (2010). The free-energy principle: a unified brain theory? Nature Reviews Neuroscience, 11(2), 127–138.
6. Wikipedia (2023). Energy-based model. (Accessed 2025).
- ↑ Obidi, John Onimisi. A Critical Review of the Theory of Entropicity (ToE) on Original Contributions, Conceptual Innovations, and Pathways towards Enhanced Mathematical Rigor: An Addendum to the Discovery of New Laws of Conservation and Uncertainty. Cambridge University.(2025-06-30). https://doi.org/10.33774/coe-2025-hmk6n
- ↑ Obidi, John Onimisi . "On the Discovery of New Laws of Conservation and Uncertainty, Probability and CPT-Theorem Symmetry-Breaking in the Standard Model of Particle Physics: More Revolutionary Insights from the Theory of Entropicity (ToE)". Cambridge University. (14 June 2025). https://doi.org/10.33774/coe-2025-n4n45
- ↑ Obidi, John Onimisi. Einstein and Bohr Finally Reconciled on Quantum Theory: The Theory of Entropicity (ToE) as the Unifying Resolution to the Problem of Quantum Measurement and Wave Function Collapse. Cambridge University. (14 April 2025). https://doi.org/10.33774/coe-2025-vrfrx
- ↑ Obidi, John Onimisi (25 March 2025). "Attosecond Constraints on Quantum Entanglement Formation as Empirical Evidence for the Theory of Entropicity (ToE)". Cambridge University. https://doi.org/10.33774/coe-2025-30swc
- ↑ Obidi, John Onimisi. The Theory of Entropicity (ToE) Validates Einstein’s General Relativity (GR) Prediction for Solar Starlight Deflection via an Entropic Coupling Constant η. Cambridge University. (23 March 2025). https://doi.org/10.33774/coe-2025-1cs81
- ↑ Obidi, John Onimisi. The Theory of Entropicity (ToE): An Entropy-Driven Derivation of Mercury’s Perihelion Precession Beyond Einstein’s Curved Spacetime in General Relativity (GR). Cambridge University. (16 March 2025). https://doi.org/10.33774/coe-2025-g55m9
- ↑ Obidi, John Onimisi. How the Generalized Entropic Expansion Equation (GEEE) Describes the Deceleration and Acceleration of the Universe in the Absence of Dark Energy. Cambridge University. (12 March 2025). https://doi.org/10.33774/coe-2025-6d843
- ↑ Obidi, John Onimisi. Corrections to the Classical Shapiro Time Delay in General Relativity (GR) from the Entropic Force-Field Hypothesis (EFFH). Cambridge University. (11 March 2025). https://doi.org/10.33774/coe-2025-v7m6c
- ↑ Obidi, John Onimisi (2025). Master Equation of the Theory of Entropicity (ToE). Encyclopedia. https://encyclopedia.pub/entry/58596