Fact-checked by Grok 2 weeks ago

Entropic force

An entropic force is an emergent macroscopic force in thermodynamic systems that originates from the statistical tendency to maximize , rather than from direct interactions of . It drives the system toward configurations with the greatest number of accessible microstates, appearing as an effective force that depends on temperature and the constraints of the system's . In , entropic forces arise when constraints limit the available , leading to a reduction in that manifests as a restoring or repulsive effect. A foundational example is the across a , where solute particles' random thermal motions create an effective force balancing external pressures, as derived by in his 1905 analysis of . This pressure follows the and exemplifies how microscopic fluctuations yield macroscopic behavior without energetic potentials. Entropic forces play a central role in physics, such as in chains where stretching reduces conformational , producing an elastic recoil force proportional to temperature, as seen in the elasticity of rubber. In colloidal suspensions, they cause depletion attractions between particles due to osmotic imbalances from crowding, influencing and phase transitions. Biological systems also harness these forces, for instance, in DNA stretching under confinement or , where maximization guides molecular interactions. Beyond classical applications, entropic forces have inspired theoretical extensions, including proposals that emerges as an entropic effect from gradients on holographic screens, as formulated by Erik Verlinde in 2011. Such ideas link to fundamental interactions, though they remain subjects of debate regarding their consistency with established physics. Overall, entropic forces underscore 's role as a driving principle across scales, from microscopic fluctuations to cosmological phenomena.

Fundamentals

Definition

An entropic is a macroscopic, emergent in thermodynamic systems that arises from the tendency of microscopic constituents to maximize the system's , rather than from direct pairwise interactions or gradients. In , quantifies the number of accessible microstates corresponding to a given macrostate; when constraints alter this number, the system responds by adjusting configurations to restore higher , manifesting as an effective . This contrasts sharply with conservative forces, such as gravitational or electrostatic forces, which derive from explicit functions and are independent of or configurational multiplicity. The conceptual foundations of entropic forces trace back to Ludwig Boltzmann's pioneering work in the 1870s, where he established the statistical interpretation of as S = k \ln W, with k as Boltzmann's constant and W the number of microstates, laying the groundwork for understanding how probabilistic distributions drive macroscopic behavior. The specific notion of entropic forces gained prominence in the mid-20th century within , where researchers like Werner Kuhn applied to explain the elasticity of rubber-like materials as arising from entropy rather than energetic bonds; Kuhn's 1934 and 1946 contributions formalized this by modeling polymer chains as random walks whose retraction maximizes configurational freedom. At its core, the principle governing entropic forces aligns with the second law of thermodynamics: systems spontaneously evolve toward equilibrium states of maximum entropy, producing force-like effects without underlying energy minima. For example, imposing positional constraints on particles reduces the multiplicity of available states, thereby decreasing entropy and generating a restorative "force" that drives expansion to alleviate the confinement and restore probabilistic diversity. This entropic drive is inherently temperature-dependent, scaling with thermal energy, and underscores how apparent macroscopic forces can emerge purely from statistical imperatives.

Thermodynamic Principles

The second law of thermodynamics states that for any spontaneous process in an isolated system, the total entropy S must increase or remain constant, expressed as dS \geq 0, which drives systems toward states of maximum disorder or equilibrium. This principle underpins the emergence of entropic forces, as configurations that maximize entropy are favored, leading to apparent forces that counteract constraints on disorder. A foundational concept for quantifying is Boltzmann's formula, S = k \ln \Omega, where k is Boltzmann's and \Omega represents the number of accessible microstates corresponding to a macrostate. This statistical definition links microscopic multiplicity to macroscopic , providing the basis for understanding how changes in configuration space generate entropic effects in thermodynamic systems. In isothermal processes, the F = U - T S, with U as and T as , serves as the relevant potential, minimized at under constant volume and temperature. The force arising from entropy variations relates to the term -T \nabla S, where spatial gradients in contribute to the effective force alongside internal energy changes. Entropic forces differ from enthalpic forces, which stem primarily from changes in U (or in constant-pressure contexts), by originating from gradients rather than energetic interactions. Entropic contributions become dominant in systems at high temperatures or those involving flexible structures, such as polymers, where thermal motion amplifies the drive toward configurational disorder, often making the force increase with temperature.

Mathematical Formulation

General Expression

The entropic force is mathematically expressed in its general form as \mathbf{F} = T \nabla S, where T denotes the absolute temperature, S = S(\mathbf{X}) is the total entropy of the system as a function of the position coordinate \mathbf{X}, and \nabla S represents the spatial gradient of the entropy. This formulation implies that the force acts in the direction of the entropy gradient, thereby driving the system toward configurations that maximize entropy and resisting those that would decrease it; its magnitude scales linearly with temperature, emphasizing the purely statistical origin of the effect in thermal equilibrium. Dimensionally, the expression yields force in newtons (N), since temperature T is in kelvins (K), entropy S in joules per kelvin (J/K), and \nabla S in J/(K \cdot m), resulting in T \nabla S equivalent to energy per unit length (J/m = N). The relation holds for quasi-static processes in large-scale systems with many particles (N \gg 1), where fluctuations are negligible and the entropy can be treated as a smooth function of macroscopic variables. It emerges from the thermodynamic basis of minimizing the A = U - T S under conditions where the U is independent of position.

Derivation from

In , the equilibrium for the configuration of a at constant temperature T is given by the form P(\mathbf{X}) \propto \exp(-\Delta F / k_B T), where \mathbf{X} represents the system's coordinates, \Delta F is the change in , k_B is Boltzmann's constant, and the is ensured by the partition function Z = \int d\mathbf{X} \exp(-\Delta F / k_B T). This distribution arises from the principle of maximizing the subject to constraints on the average and , reflecting the most probable state consistent with the available . The decomposes as F = U - T S, where U is the average and S is the , defined microscopically as S = -k_B \sum_i P_i \ln P_i (or in the continuum limit, S = -k_B \int P(\mathbf{X}) \ln P(\mathbf{X}) \, d\mathbf{X}). The total on the system is the negative of the , \mathbf{F}_\text{total} = -\nabla F. Substituting the decomposition yields \mathbf{F}_\text{total} = -\nabla U + T \nabla S, isolating the entropic term T \nabla S when the internal energy gradient \nabla U is negligible or separately accounted for, such as in systems dominated by configurational . This entropic emerges as the macroscopic manifestation of the microscopic tendency to maximize the number of accessible microstates. To derive the equilibrium distribution probabilistically, one maximizes the S under the constraints of fixed \sum_i P_i = 1 and fixed average \sum_i P_i \varepsilon_i = \langle \varepsilon \rangle, where \varepsilon_i are the energies. This is achieved using Lagrange multipliers \lambda_0 and \lambda_1, forming the functional \mathcal{L} = S + \lambda_0 (\sum_i P_i - 1) + \lambda_1 (\sum_i P_i \varepsilon_i - \langle \varepsilon \rangle). Taking the variation \delta \mathcal{L} / \delta P_i = 0 leads to P_i = \exp(-\lambda_0 - \lambda_1 \varepsilon_i), or equivalently P_i \propto \exp(-\beta \varepsilon_i) with \beta = 1 / k_B T, confirming the Boltzmann form. Edwin Jaynes framed this derivation within , interpreting the maximum principle as a method of from incomplete : the constraints represent partial about the , and the resulting distribution is the least biased (maximum ) estimate consistent with that , avoiding unfounded assumptions about unobserved details. In this view, entropic forces arise not from direct interactions but from the inferential drive to increase under informational constraints, providing a foundational justification for their in complex .

Physical Examples

Ideal Gas Pressure

In the context of an confined within a , the observed P results from the collisions of gas particles with the container walls, a phenomenon traditionally explained through . However, from the perspective of , this pressure manifests as an entropic force, driven by the system's tendency to maximize its by expanding the available volume V, which increases the number of accessible microstates \Omega \propto V^N for N . Thermodynamically, the pressure relates to the F via the relation P = -\left( \frac{\partial F}{\partial V} \right)_T, where F = U - T S and U is the . For an , U depends solely on temperature and is independent of volume, so the volume dependence of F arises entirely from the entropic term -T S. The S includes a configurational contribution S = N k \ln \left( \frac{V}{N} \right) + f(T, N), where k is Boltzmann's constant and f(T, N) encapsulates temperature- and particle-number-dependent terms; differentiating yields \left( \frac{\partial S}{\partial V} \right)_T = \frac{N k}{V}, confirming P = \frac{N k T}{V}. This formulation highlights the key insight that, at constant , the pressure of an is purely in origin, with no contribution from interparticle potential energies, as the particles are non-interacting.

Polymer Elasticity

In the freely jointed chain model, introduced by Werner Kuhn in , a is represented as a chain of N rigid segments, each of length l, joined by frictionless hinges that permit uncorrelated orientations. This idealization captures the random coil configuration of flexible in solution or melt, where thermal motion drives the chain to adopt numerous conformations to maximize . Fixing the end-to-end \mathbf{R} restricts these configurations, reducing the configurational and generating an entropic restoring upon stretching. The entropy S for a fixed \mathbf{R} in the Gaussian chain approximation, valid for large N and R \ll N l, follows from the probability distribution of end-to-end distances, which resembles a three-dimensional : S(\mathbf{R}) = S_0 - \frac{3 k_B R^2}{2 N l^2}, where S_0 is the maximum at R = 0, k_B is Boltzmann's constant, and the quadratic term arises from the applied to the segment vectors. This loss reflects fewer accessible microstates under extension, analogous to but distinct from the isotropic confinement in pressure. The resulting entropic force \mathbf{F} is derived from the A = U - T S, assuming negligible change U (), so \mathbf{F} = -(\partial A / \partial \mathbf{R})_T = T (\partial S / \partial \mathbf{R})_T: \mathbf{F} = \frac{3 k_B T}{N l^2} \mathbf{R}. This mimics a classical , with effective spring constant $3 k_B T / N l^2 that scales with T, explaining why rubber stiffens upon heating—a signature of entropic dominance over energetic contributions. Kuhn's theoretical predictions aligned with early experiments on , confirming the model's validity for crosslinked networks where chain uncoiling provides reversible deformation.

Brownian Motion

Brownian motion describes the random, diffusive of a particle suspended in a fluid, arising from incessant collisions with surrounding molecules at . When the particle is subjected to an external potential V, the equilibrium probability density \rho of its position follows the \rho \propto \exp(-V / kT), where k is Boltzmann's constant and T is . This distribution implies an effective entropic force \mathbf{F}_\text{ent} = -kT \nabla \ln \rho, which emerges from the tendency to maximize configurational and drives the particle toward regions of higher probability density, counterbalancing the external potential. The entropic force connects directly to the Einstein relation, which relates the particle's mobility \mu—defined as the ratio of its average drift velocity to an applied force—to the diffusion coefficient D via \mu = D / kT. In steady state, this force balances the viscous drag \gamma \mathbf{v}, where \gamma = 1/\mu is the friction coefficient and \mathbf{v} is velocity; the relation ensures that diffusive spreading due to thermal fluctuations is inversely proportional to the dissipative drag. introduced this in his 1905 analysis of , interpreting the irregular particle paths as evidence of atomic-scale diffusion driven by osmotic (entropic) pressures, thereby providing a microscopic foundation for the kinetic theory of matter. A rigorous derivation of the entropic force follows from the Fokker-Planck equation, which governs the time evolution of \rho(\mathbf{r}, t): \partial_t \rho = -\nabla \cdot \mathbf{J}, where the is \mathbf{J} = \mu \mathbf{F} \rho - D \nabla \rho in the overdamped limit, with \mathbf{F} the total force. At , \mathbf{J} = 0, yielding \mu \mathbf{F} \rho = D \nabla \rho, or \mathbf{F} = kT \nabla \ln \rho; the entropic component then appears as \mathbf{F}_\text{ent} = -kT \nabla \ln \rho when isolating the diffusive contribution that mimics an balancing external influences. This framework links entropic forces to the , as the Einstein relation equates fluctuation strength (diffusion) to dissipation (drag), ensuring thermodynamic consistency in stochastic dynamics.

Biological and Chemical Examples

Hydrophobic Effect

The hydrophobic effect exemplifies an that drives the aggregation of non-polar molecules or moieties in , primarily through changes in the solvent's configurational . When hydrophobic solutes are introduced into , the solvent molecules reorganize to form a more structured layer around the solute, minimizing the disruption to their hydrogen-bonding network. This structuring imposes constraints on 's degrees of , resulting in a negative change for . Upon association of hydrophobic groups, the structured is liberated, allowing it to revert to a higher- bulk state, yielding a positive overall change (ΔS > 0) that favors aggregation. This mechanism was first articulated in the 1940s by Frank and Evans, who proposed the "caged " or "" model, wherein molecules encase hydrophobic solutes in clathrate-like structures resembling ice, with reduced mobility and . Experimental and studies have since validated this model, showing enhanced tetrahedral ordering and slower dynamics in the hydration shell of non-polar solutes like or alkanes, with the penalty arising from the increased structural ordering and reduced dynamics of molecules in the hydration shell, showing enhanced tetrahedral ordering. The release of these caged waters during hydrophobic association provides the entropic driving force, often dominating over any enthalpic contributions from van der Waals interactions between the solutes. The strength of this force displays a pronounced temperature dependence, peaking around 300-400 due to optimal structuring at ambient conditions, beyond which disrupts the cages and weakens the effect. At lower temperatures, the effect diminishes as 's hydrogen-bond network becomes more rigid overall. In applications to , the hydrophobic effect stabilizes the native structure by promoting the burial of non-polar side chains, reducing solvent-accessible surface area and releasing structured ; this contributes significantly to the folding in many proteins. Calorimetric measurements of model systems, such as transfer from to organic phases or association, reveal an entropy gain upon desolvation, reflecting the release of structured molecules. These values underscore the effect's role in biomolecular recognition, though modulated by specific enthalpic factors in complex systems.

Colloidal Systems

In colloidal systems, entropic forces arise prominently through depletion interactions, where smaller particles or chains—known as depletants—are sterically excluded from the thin layer surrounding larger colloidal particles, generating an effective attractive between the colloids due to an imbalance in . This exclusion reduces the available volume for depletants when two colloids approach closely, increasing their configurational elsewhere in the and driving the colloids together to maximize overall system . Such forces are purely entropic, with no enthalpic contribution from direct interactions between the colloids or depletants, and they play a key role in promoting in suspensions like polymer-colloid mixtures or binary hard-sphere systems. The foundational description of this phenomenon is provided by the Asakura-Oosawa model, which treats both colloids and depletants as in the limit of low depletant concentration. The effective depletion potential U(r) between two spherical colloids of diameter \sigma separated by center-to-center distance r (where \sigma < r < \sigma + 2q, and q is the depletant radius) is given by U(r) = -\frac{\pi \rho}{6} (\sigma + 2q - r)^3 k_B T, where \rho is the of depletants, k_B is Boltzmann's constant, and T is the ; outside this range, U(r) = 0. This quadratic-well potential leads to short-range , with the depth scaling as \rho q^3 k_B T, enabling tunable control over assembly by varying depletant size and concentration. The model, originally developed in the 1950s, has been validated through simulations and experiments showing its accuracy for ideal depletants but requires extensions for polydispersity or higher concentrations. These depletion-driven entropic forces can induce in colloidal suspensions, transitioning from disordered fluids to ordered phases solely to increase , such as fluid-crystal coexistence or demixing into colloid-rich and depletant-rich regions. A notable example occurs in mixtures of semiflexible polymers and spherical colloids, where entropic attractions promote liquid-liquid alongside nematic ordering of the polymers, resulting in thermodynamically stable multiphase structures with colloidal particles partitioned into distinct domains. This entropy-driven ordering has been observed in experiments using , highlighting applications in designing anisotropic materials. Recent advancements have leveraged engineered depletion interactions to achieve precise control over colloidal , including dynamical effects akin to backaction in responsive suspensions. For instance, a 2023 study on engineered entropic forces in optomechanical setups with fluid interfaces, analogous to colloidal dynamics, demonstrated ultrastrong dynamical backaction, amplifying interactions by orders of magnitude.

In the cytoskeleton, entropic forces arise prominently in the overlap regions between and filaments, where diffusible crosslinkers or agents become confined, generating pressure that drives filament sliding and contraction. These forces stem from the statistical tendency of confined molecules to maximize by expanding the available volume, effectively pushing filaments together to increase overlap lengths. For instance, in microtubule networks, crosslinkers like Ase1 confine to overlap zones, producing a directed that stabilizes structures during . The magnitude of this entropic force can be approximated as F = \frac{n k_B T}{L}, where n is the number of confined crosslinkers, k_B is Boltzmann's constant, T is , and L is the overlap length; experimental measurements using report forces up to approximately 3.7 pN, sufficient to counterbalance motor-driven sliding. Theoretical models demonstrate that such entropic mechanisms can produce contractile forces in cytoskeletal networks without ATP hydrolysis, relying solely on thermal fluctuations and confinement. In these passive systems, crowding agents or diffusible binders induce network shrinkage by favoring configurations with greater filament overlap, mimicking active contractility observed in cells. This ATP-independent contraction highlights how entropic effects supplement or even substitute for motor proteins in remodeling dynamic structures like the actin cortex or microtubule asters. Entropic contributions from filament fluctuations play a key role in cell motility, where thermal bending and orientational disorder in actin networks reduce entropy upon polymerization, generating propulsive forces at the leading edge. In lamellipodia, for example, the confinement of fluctuating filaments against the plasma membrane creates an entropic spring-like resistance, enabling directed protrusion and retrograde flow without dominant enthalpic contributions from filament stretching. These fluctuations, inherent to semiflexible polymers like actin, amplify force output during assembly, facilitating efficient crawling on substrates. Studies from 2015 revealed that entropic loads from confined crosslinkers can induce buckling in active networks, where compressive sliding forces exceed the filament's rigidity, leading to and reorganization. This buckling under entropic provides a for adaptive reshaping in cellular processes, such as , and underscores the interplay between and mechanical stability . These phenomena build on foundational principles of polymer elasticity, where scales dictate the entropic response to confinement.

Theoretical Extensions

Entropic Gravity

Entropic gravity proposes that the gravitational force arises not as a but as an emergent phenomenon driven by gradients in the underlying microstructure of . This perspective draws from thermodynamic principles and , suggesting that emerges from the tendency of systems to maximize , much like in a gas arises from molecular disorder. In this framework, the familiar of Newtonian is recast as a statistical effect, where the displacement of masses alters the associated with informational on holographic screens. The seminal formulation of entropic gravity was introduced by Erik Verlinde in 2010, who argued that gravity originates from changes in the entropy of a holographic screen enclosing a mass. Verlinde posited that the entropic force F on a test mass m displaced by \Delta x near a spherical screen of radius r and temperature T is given by F = T \frac{\Delta S}{\Delta x}, where the entropy change is \Delta S = \frac{2\pi k m c}{\hbar} \Delta x, with k Boltzmann's constant, c the speed of light, and \hbar reduced Planck's constant. This leads to the emergent Newtonian force F = \frac{G M m}{r^2}, where G is the gravitational constant and M the enclosed mass, with the screen's Unruh temperature T = \frac{\hbar a}{2\pi k c} and acceleration a = \frac{G M}{r^2}. Verlinde's theory extends to general relativity by incorporating relativistic effects and has been explored for its implications in resolving dark matter puzzles through modified entropic contributions at galactic scales. A significant modification appeared in 2025 with Ginestra Bianconi's "Gravity from Entropy," which derives gravitational dynamics from the quantum relative between matter fields and the metric. Bianconi proposes an entropic action based on the quantum relative S(\rho \| \sigma), generalizing the Araki entropy to curved spacetimes, and shows that varying this action yields Einstein's field equations, potentially unifying and without introducing new fundamental forces. This approach treats as arising from the minimization of relative , offering a quantum information-theoretic foundation that addresses some limitations of classical entropic models. Criticisms of entropic gravity, including Verlinde's original proposal, center on its failure to fully resolve issues, such as the divergences in holographic calculations and the lack of a complete microscopic theory for the underlying bits. Skeptics argue that while the thermodynamic is intriguing, it does not explain why gravity appears classical at macroscopic scales or how it integrates with without ad hoc assumptions. Analyses of rotation curves in 2024 have explored entropic modifications as alternatives to , showing some alignments but also discrepancies, such as in clusters. While intriguing, empirical support remains limited, with ongoing theoretical scrutiny. A 2025 Quanta Magazine article highlights renewed interest in by exploring how the universe's increase could drive attractive forces between masses, framing as a "push" from rising disorder rather than a pull from , based on a February 2025 model by predicting testable effects in quantum superpositions. This perspective suggests cosmological implications, such as entropic contributions to acceleration, where expanding universes maximize through gravitational clustering, potentially offering a thermodynamic for the observed cosmic expansion without invoking a . A July 2025 New Scientist article further discusses how could address and energy puzzles. Physicists remain divided, viewing it as a provocative that inspires new models but requires empirical validation.

Causal Entropic Forces

Causal entropic forces represent a generalization of traditional entropic forces to non-equilibrium systems, where the force arises from maximizing the diversity of future paths in configuration space rather than equilibrium configurations. Introduced by Wissner-Gross and Freer in 2013, these forces are defined as \mathbf{F}(\mathbf{X}_0, \tau) = T_c \nabla_{\mathbf{X}} S_c(\mathbf{X}, \tau) \big|_{\mathbf{X}_0}, where T_c is a causal temperature parameter, \mathbf{X}_0 is the initial configuration, \tau is a future time horizon, and S_c(\mathbf{X}, \tau) is the causal path entropy measuring the uncertainty in accessible future paths starting from \mathbf{X}. The entropy S_c is computed via path integrals as S_c(\mathbf{X}, \tau) = -k_B \int_{\mathbf{x}(0)=\mathbf{X}}^{\mathbf{x}(\tau) \in \Omega} \Pr[\mathbf{x}(t) \mid \mathbf{x}(0)] \ln \Pr[\mathbf{x}(t) \mid \mathbf{x}(0)] \, D\mathbf{x}(t), with k_B Boltzmann's constant and \Pr[\cdot] the conditional probability over paths \mathbf{x}(t) constrained to an accessible region \Omega. This formulation incorporates time-asymmetric causality by weighting paths forward in time, distinguishing it from equilibrium entropic forces that rely on symmetric statistical ensembles. In applications, causal entropic forces drive adaptive behaviors in simple physical agents, such as tool use, where one disk manipulates another to access a confined object within a , emerging solely from maximization over future paths. Similarly, social cooperation arises when two disks synchronize to pull a string and retrieve an object, demonstrating how these forces can induce collective actions without explicit programming. These examples link causal entropic forces to , positing it as a process that maximizes causal flow—accelerating the diversity of possible future histories—in open, non-equilibrium systems. Speculatively, such forces exhibit gravity-like attraction in adaptive , where agents are drawn toward configurations that enhance their future path diversity, analogous to how emerges from information gradients in . Recent work extends this to , modeling populations of entropically driven agents on lattices that self-organize into polymer-like structures through local interactions, mimicking collective adaptation without centralized control. Further 2025 extensions include studies on effects in entropic systems and links to maximum caliber models for emergent intuition in critical dynamics. This underscores the role of causal entropic forces in predicting goal-directed behaviors across scales, from individual agents to emergent groups.

Modern Applications

Engineered Entropic Systems

Engineered entropic systems involve the deliberate manipulation of entropic forces to achieve desired material properties or device functionalities, often leveraging thermodynamic principles to enhance stability, assembly, or dynamical responses in physical setups. These approaches draw from fundamental concepts in , where entropy gradients drive emergent behaviors without relying on direct energetic interactions. Recent advancements have focused on optomechanical, , colloidal, and quantum platforms, enabling applications in sensing, , and quantum technologies. In optomechanical systems, entropic forces have been engineered to produce ultrastrong dynamical backaction, surpassing traditional effects by orders of magnitude. A study demonstrated this by using a superfluid third-sound , where generates that induces entropic forces, optimized when the thermal response time matches the driving frequency (Ωτ ≈ 1). This mechanism facilitates multiphonon scattering, leading to lasing with a threshold power of 1–3.4 —a reduction by a factor of over 1000 compared to previous optically driven systems. The approach, detailed in experiments by Schliesser et al., highlights entropic forces' potential for quantum-limited sensing and coherent generation in cryogenic environments. High-entropy alloys represent another frontier, where entropic contributions stabilize complex atomic structures against . In a 2025 investigation of CuBiI4 crystals, suitable for optoelectronic applications like solar cells, configurational was shown to favor a site-disordered cubic amid over 10^13 possible atomic arrangements. Using density-functional theory and cluster-energy expansion models, Tuttle et al. identified low-energy configurations via simulations, confirming that minimization—dominated by at elevated temperatures—stabilizes the disordered structure over ordered alternatives. This entropy-driven stabilization enhances mechanical robustness and electronic properties, such as bandgap tunability, in multi-component perovskites. Colloidal engineering exploits tunable depletion forces to direct in , mimicking biological organization on micron scales. Depletion interactions, arising from imbalances caused by smaller depletant particles or polymers, can be precisely controlled to induce attractive potentials between larger colloids. For instance, a 2024 experiment by Onuh and Harries used 1 μm particles on substrates with depletants like 0.5% (w/v) nanoparticles or poly() (PAA) polymers, tuning assembly via (4–9) to alter PAA's and effective size. At neutral with nanoparticles, ordered multilayer packings formed; at 9 with PAA, irregular aggregates covered up to 9000 μm², demonstrating how depletion strength scales with depletant concentration and enables programmable patterns for photonic or rheological materials. Proposals for quantum entropic forces extend these ideas to microscopic scales, suggesting gravity-like effects emerge from entropy in entangled systems. A 2025 theoretical framework by Carney constructs fully quantum-mechanical models where Newton's law arises from extremizing the of qubit or oscillator arrays, rather than virtual particle exchange. These local and non-local entropic models predict distinguishable signatures from perturbative , such as modified noise spectra in interferometers. Near-future experiments, including optomechanical table-top setups and entanglement witnesses proposed since 2023, aim to detect these effects through precision measurements of decoherence rates or gravitational analogs in ultracold atomic arrays.

Machine Learning and AI

In , entropic forces have been analogized to the dynamics of training, where (SGD) acts as a diffusive process that maximizes representational to drive toward optimal solutions. A 2025 study formalizes this by proposing an entropic-force theory for deep representation learning, showing that SGD induces an effective force proportional to the gradient of in the parameter space, promoting exploration in high-dimensional loss landscapes and accelerating universal approximation capabilities in networks. This framework reveals how thermodynamic-like principles govern the irreversibility of training trajectories, with empirical demonstrations on and models illustrating maximization as a key driver of . Building on these ideas, causal entropic forces have been applied to challenges, particularly in ensuring safe by constraining agent behaviors through entropy-based objectives. In a analysis of deep alignment, researchers model value alignment as an entropy maximization problem under causal constraints, where entropic forces emerge to bound misaligned actions, preventing catastrophic deviations in superintelligent systems. This approach draws from theoretical causal entropic forces, which posit that agents infer and pursue long-term goals by maximizing causal in processes. Such methods have shown promise in simulations of multi-agent environments. Entropic causal inference extends these concepts to structure learning, enabling the identification of causal graphs from observational data by minimizing information-theoretic divergences akin to entropic potentials. A 2025 paper on graph identifiability demonstrates that entropic measures can uniquely recover directed acyclic graphs (DAGs) under mild distributional assumptions, with finite-sample guarantees for identifiability in high-dimensional settings. This technique links directly to biological causal networks, such as gene regulatory pathways, by treating inference as an entropic force that pulls toward sparse, biologically plausible structures, outperforming classical methods like PC algorithm in accuracy on synthetic benchmarks with 20-30 nodes. Finally, swarm AI systems leverage entropic to mimic biological , where decentralized agents coordinate via entropy-driven interactions to achieve . In a 2025 agent-based model, entropy metrics quantify swarming states, revealing how local entropic forces—analogous to those in physical colloids—induce global patterns like or without central control. These systems bridge computational and biological domains, with applications in where swarm efficiency rivals centralized planners in dynamic environments.

References

  1. [1]
    [PDF] Entropic forces in Brownian motion - arXiv
    I. INTRODUCTION. An entropic force is an emergent phenomenon resulting from the tendency of a thermodynamic system to maximize its entropy. The system, or the ...
  2. [2]
    [PDF] The concept of an “entropic force” can be introduced by
    Then the heat transfer would be reversible and there would be no damping.) The actual force on the piston is due to the gauge pressure of the gas. However, ...
  3. [3]
    [PDF] the brownian movement - DAMTP
    The last equation states that equilibrium with the force is brought about by osmotic pressure forces. Equation (I) can be used to find the coefficient of ...
  4. [4]
    Chemical potential formalism for polymer entropic forces - Nature
    Mar 1, 2019 · The entropic force is one of the most elusive interactions in macromolecules. It is clearly defined theoretically, but practically difficult ...
  5. [5]
    [1001.0785] On the Origin of Gravity and the Laws of Newton - arXiv
    Jan 6, 2010 · Newton's law of gravitation is shown to arise naturally and unavoidably in a theory in which space is emergent through a holographic scenario.
  6. [6]
    [PDF] Is Gravity an Entropic Force? - arXiv
    An entropic force can be defined as an effective macroscopic force that originates in ... is an entropic force in physics. Besides, it is far from clear whether ...
  7. [7]
    [PDF] Statistical mechanics of entropic forces: disassembling a toy - LPTMS
    Sep 23, 2010 · The so-called entropic forces appearing in the description of such systems are the statistical mechanics counterparts of the reaction forces.
  8. [8]
    It Ain't Necessarily So: Ludwig Boltzmann's Darwinian Notion ... - MDPI
    Ludwig Boltzmann's move in his seminal paper of 1877, introducing a statistical understanding of entropy, was a watershed moment in the history of physics.
  9. [9]
    The Theories of Rubber Elasticity and the Goodness of Their ... - MDPI
    Aug 22, 2024 · Later, in 1946, Kuhn used a statistical macromolecule model to develop the first theory of the entropic elasticity of rubber [3]. To understand ...
  10. [10]
    Theory of the Elasticity of Rubber - AIP Publishing
    It is the presence of the intermolecular bonds, which link the molecules into a network and thus control its form, which differentiates rubberlike materials ...
  11. [11]
    Entropic forces in Brownian motion - AIP Publishing
    Dec 1, 2014 · On a macroscopic scale this increase in entropy can be described as an (emerging) entropic force using Boltzmann's law. Such forces are just as ...
  12. [12]
    [PDF] Entropy and entropic forces - UiO
    • Entropy and equilibrium conditions. • Thermodynamic identity. • Mixtures and chemical potential. • Helmholtz and Gibbs free energy. • Boltzmann statistics.
  13. [13]
    [PDF] 1 Thermodynamics of rubbers - INFN Roma
    The Helmholtz free energy is a thermodynamic state function of variables T,V ... the entropic force can be rewritten (as one of the Maxwell relations) as.
  14. [14]
    [PDF] On the Origin of Gravity and the Laws of Newton arXiv:1001.0785v1 ...
    Jan 6, 2010 · In this case the correspondence rules must be such that F = T∇S translates in to the inertial force. F = ma. Again the formulas work out as in ...<|control11|><|separator|>
  15. [15]
  16. [16]
  17. [17]
    Entropy of an Ideal Gas - HyperPhysics
    Entropy of an Ideal Gas · N = number of atoms · k = Boltzmann's constant · V = volume · U = internal energy · h = Planck's constant.
  18. [18]
    [PDF] Lecture 8: Free energy
    The Helmholtz free energy is one of the most useful quantities in thermodynamics. Its usefulness stems from the fact that dV , dN and dT are readily measurable ...
  19. [19]
    [PDF] Ideal Chain Statistics, Free Energy and Chain Deformation
    The entropy as a function of the end to end distance R is provided by Boltzmann's relation S = k ln Ω(R) where Ω(R) is simply the number of ways (conformations ...Missing: jointed formula
  20. [20]
    Impact of branching on the elasticity of actin networks - PMC
    Cross-linked actin networks exhibit this kind of entropic elasticity with a strong stress-stiffening (9, 10). Enthalpic elasticity comes from the mechanical ...Missing: relevance | Show results with:relevance
  21. [21]
    [PDF] brownian movement - albert einstein, ph.d.
    The last equation states that equilibrium with the force K is brought about by osmotic pressure forces. Equation (I) can be used to find the coefficient of ...
  22. [22]
    The Hydrophobic Effects: Our Current Understanding - PMC - NIH
    In 1945, Frank and Evans [8] suggested that the observed loss of entropy was related to the structural changes of liquid water as hydrophobic solutes were ...
  23. [23]
    Revealing the Frank–Evans “Iceberg” Structures within the Solvation ...
    Feb 4, 2021 · The existence of the ordered aggregates around the hydrophobic solutes complies with the concept of “icebergs” proposed by Frank and Evans.Methods · Analysis · Parameters δ And τ
  24. [24]
    Recent progress in understanding hydrophobic interactions - PNAS
    Oct 24, 2006 · We present here a brief review of direct force measurements between hydrophobic surfaces in aqueous solutions.
  25. [25]
    The Hydrophobic Temperature Dependence of Amino Acids Directly ...
    May 22, 2015 · From such measurements, models and theory we know that the hydrophobic force peaks between 30–80°C and becomes weaker at both lower and higher ...Missing: 400 | Show results with:400
  26. [26]
    A View of the Hydrophobic Effect | The Journal of Physical Chemistry B
    Hydrophobicity is entropic in cold water and enthalpic in hot water. A.Iceberg Model. The iceberg model of Frank and Evans explains the large heat capacity of ...
  27. [27]
    Mechanism of the hydrophobic effect in the biomolecular recognition ...
    The hydrophobic effect—a rationalization of the insolubility of nonpolar molecules in water—is centrally important to biomolecular recognition.Abstract · Sign Up For Pnas Alerts · Results And Discussion
  28. [28]
    Depletion attraction in colloidal and bacterial systems - Frontiers
    Sep 5, 2023 · Depletion attraction is a common entropy force observed in colloidal systems. As a common phenomenon in colloidal and bacterial systems, studying the mechanism ...
  29. [29]
    Thermodynamically controlled multiphase separation of ... - Nature
    Aug 29, 2023 · We develop a series of heterogeneous colloidal suspensions that exhibit both liquid-liquid phase separation of semiflexible binary polymers and liquid crystal ...
  30. [30]
  31. [31]
  32. [32]
  33. [33]
    [2408.14391] Gravity from entropy - arXiv
    Aug 26, 2024 · Abstract page for arXiv paper 2408.14391: Gravity from entropy. ... The proposed entropic action is the quantum relative entropy between the ...
  34. [34]
    Gravity from entropy | Phys. Rev. D - Physical Review Link Manager
    Mar 3, 2025 · Gravity from entropy. Ginestra Bianconi*. School of Mathematical ... Here we have established the relation of the adopted quantum relative action ...
  35. [35]
    Is Gravity Just Entropy Rising? Long-Shot Idea Gets Another Look.
    Jun 13, 2025 · A new argument explores how the growth of disorder could cause massive objects to move toward one another. Physicists are both interested and skeptical.What Is Entropy? A Measure of... · Why Gravity Is Not Like the...
  36. [36]
    Emergent gravity may be a dead idea, but it's not a bad one
    Nov 19, 2024 · Emergent gravity is a bold idea. It claims that the force of gravity is a mere illusion, more akin to friction or heat—a property that emerges ...
  37. [37]
    (PDF) Gravity is not an entropic force - ResearchGate
    Aug 7, 2025 · Our arguments reinforce other objections concerning the entropic gravity proposal, such as the irreversibility effects of entropic forces ...
  38. [38]
    Causal Entropic Forces | Phys. Rev. Lett.
    Apr 19, 2013 · Our results suggest a potentially general thermodynamic model of adaptive behavior as a nonequilibrium process in open systems.
  39. [39]
    Entropically Driven Agents
    ### Summary: Relation of Causal Entropic Forces to Swarm Intelligence or Adaptive Evolution
  40. [40]
    Engineered entropic forces allow ultrastrong dynamical backaction
    May 24, 2023 · We develop a framework to engineer the dynamical backaction from entropic forces, applying it to achieve phonon lasing with a threshold three orders of ...
  41. [41]
  42. [42]
  43. [43]
    On the quantum mechanics of entropic forces
    ### Summary of Proposals for Quantum Entropic Forces
  44. [44]
    Entropic Forces in Deep and Universal Representation Learning
    May 18, 2025 · Neural Thermodynamics: Entropic Forces in Deep and Universal Representation Learning. Authors:Liu Ziyin, Yizhou Xu, Isaac Chuang.
  45. [45]
    [2509.16463] Entropic Causal Inference: Graph Identifiability - arXiv
    Entropic causal inference is a recent framework for learning the causal graph between two variables from observational data by finding the ...