Fact-checked by Grok 2 weeks ago

Microcanonical ensemble

The microcanonical ensemble is a fundamental concept in that describes the statistical behavior of an characterized by fixed values of energy E, volume V, and number of particles N, where all accessible microstates consistent with these constraints are assumed to be equally probable. This ensemble provides a framework for connecting microscopic dynamics to macroscopic thermodynamic properties, particularly for systems in without exchange of energy or matter with their surroundings. Introduced by J. Willard Gibbs in his foundational work on the elements of statistical mechanics around 1902, the microcanonical ensemble forms one of the three primary ensembles in Gibbsian statistical mechanics, alongside the canonical and grand canonical ensembles. It emphasizes its role in treating isolated systems directly through the postulate of equal a priori probabilities for microstates within a narrow energy shell. In classical mechanics, the ensemble is represented by a uniform distribution over the phase space volume \Gamma(E, V, N) corresponding to energies between E and E + \Delta, where \Delta is infinitesimally small in the thermodynamic limit; in quantum mechanics, it corresponds to the number of states \Omega(E, V, N) with exact energy E. Key properties of the microcanonical ensemble include its definition of as S = k \ln \Omega, where k is Boltzmann's constant, providing a measure of the system's microscopic degeneracy and linking directly to the second law of through \Delta S \geq 0 for isolated processes. emerges statistically as $1/T = (\partial S / \partial E)_{V,N}, and the ensemble resolves issues like the by incorporating the indistinguishability of particles via factors such as N! in the measure. It assumes , meaning time averages equal ensemble averages, and is particularly applicable to gases, systems, and other models where exact is enforced. In practice, the microcanonical ensemble underpins derivations of equations of state, such as PV = NkT for ideal gases, and heat capacities, while its entropy formulation facilitates the thermodynamic limit where fluctuations become negligible. Though computationally challenging for complex systems due to the need to enumerate or integrate over vast phase spaces, it remains theoretically central for understanding equilibrium in closed systems and serves as a benchmark for more approximate ensembles in open environments.

Fundamentals

Definition

The microcanonical ensemble is a statistical ensemble in that describes the possible states of an with fixed total E, volume V, and number of particles N. It consists of all accessible compatible with these macroscopic constraints, where each is assigned equal probability. The fundamental postulate underlying this ensemble asserts that, in , an is equally likely to occupy any consistent with its given , volume, and particle number. This framework originated from Ludwig Boltzmann's investigations into during the 1860s and 1870s, which laid the groundwork for equating statistical probabilities to dynamical behaviors in isolated systems. It was later formalized by in his 1902 treatise Elementary Principles in Statistical Mechanics, where he explicitly defined the microcanonical distribution as one with uniform phase-space density for systems of constant energy. The ensemble's conceptual foundation relies on the , which states that for sufficiently large and ergodic systems, the time average of a physical equals its ensemble average across the microstates, thereby linking dynamical evolution to statistical equilibrium.

Applicability

The microcanonical ensemble is ideally applicable to isolated physical systems in which external interactions, such as or particle exchange, are negligible, thereby maintaining fixed values of total E, volume V, and particle number N. This ensemble assumes the system evolves ergodically, sampling all accessible microstates uniformly within a narrow shell, which requires the system to be sufficiently large for statistical fluctuations to average out effectively. Suitable examples include systems confined in perfect insulators or adiabatic containers that prevent leakage, as well as computationally simulated isolated atomic or molecular clusters in studies. The validity of the microcanonical ensemble aligns with the , where the number of particles N \to \infty, ensuring equivalence to macroscopic thermodynamic descriptions while various definitions of thermodynamic quantities, such as , converge. However, finite-size effects become prominent in smaller systems, leading to deviations like non-equivalence between ensembles or anomalous behaviors in specific heat, which must be accounted for in precise applications. Limitations arise when systems are open or in with reservoirs, where energy or particle exchange occurs, making the or grand canonical ensembles more appropriate instead. Additionally, challenges emerge for small systems where fluctuations dominate or in quantum regimes with significant , as the ensemble's assumptions may not fully capture non-ergodic dynamics or finite-dimensional constraints.

Formal Descriptions

Classical Formulation

In classical statistical mechanics, the microcanonical ensemble describes an isolated system of N particles in a volume V with fixed total E, where the probability is uniformly distributed over the in defined by the constant E. The is parameterized by the \mathbf{q} = (q_1, \dots, q_{3N}) and momenta \mathbf{p} = (p_1, \dots, p_{3N}), with the dynamics governed by the H(\mathbf{q}, \mathbf{p}). The ensemble average of an A(\mathbf{q}, \mathbf{p}) is computed as an integral over this energy shell, ensuring that all accessible microstates consistent with the constraints are equally likely. The probability density \rho(\mathbf{q}, \mathbf{p}) for the microcanonical ensemble is given by \rho(\mathbf{q}, \mathbf{p}) = \frac{\delta \bigl( H(\mathbf{q}, \mathbf{p}) - E \bigr)}{\omega(E, V, N)}, where \delta is the Dirac delta function that enforces the energy constraint, and \omega(E, V, N) is the normalization factor representing the surface measure of the energy shell. This formulation, introduced by Gibbs, ensures that the density is zero outside the hypersurface H = E and uniform on it, with the delta function providing an infinitesimal thickness to the shell for practical computation. The cumulative phase space volume \Omega(E, V, N) enclosed by the energy surface H \leq E, accounting for indistinguishability of particles and the via Planck's constant h, \Omega(E, V, N) = \frac{1}{N! \, h^{3N}} \int d^{3N}q \, d^{3N}p \, \theta \bigl( E - H(\mathbf{q}, \mathbf{p}) \bigr), where \theta is the that restricts the integral to the region below energy E. For systems where the energy shell is thin compared to variations in H (valid for large N), \Omega(E, V, N) can be approximated by differentiating the cumulative , leading to the surface measure concentrated on H = E. This integral is evaluated over the accessible configuration , typically a of V for the coordinates \mathbf{q}. The , or the derivative of the volume with respect to , provides the surface measure directly: \omega(E) = \frac{d \Omega}{dE} = \frac{1}{N! \, h^{3N}} \int d^{3N}q \, d^{3N}p \, \delta \bigl( H(\mathbf{q}, \mathbf{p}) - E \bigr). This \omega(E) quantifies the number of states per unit interval at E and serves as the normalization for the delta-function density, enabling explicit calculations for solvable Hamiltonians like the . Under the , which posits that the system's trajectory densely explores the energy surface, the time average of an equals the average over the microcanonical measure. Specifically, for a flow generating the dynamics, \langle A \rangle = \lim_{T \to \infty} \frac{1}{T} \int_0^T A(t) \, dt = \int A(\mathbf{q}, \mathbf{p}) \, \rho(\mathbf{q}, \mathbf{p}) \, d\Gamma, where d\Gamma = d^{3N}q \, d^{3N}p / (N! \, h^{3N}) is the volume element. This equivalence justifies using the for properties in ergodic classical systems.

Quantum Formulation

In , the microcanonical ensemble characterizes an with precisely fixed E, volume V, and number of particles N. Unlike the classical case, it involves a set of energy eigenstates, focusing on the degenerate corresponding to the exact eigenvalue E of the \hat{H}. This formulation assumes the system is in a pure energy without broadening, though practical implementations may use a narrow energy for finite systems. The ensemble is represented by the density \hat{\rho}, which uniformly distributes over the energy eigenspace: \hat{\rho} = \frac{\hat{P}_E}{\operatorname{Tr}(\hat{P}_E)}, where \hat{P}_E = \sum_{n: E_n = E} |n\rangle\langle n| is the orthogonal onto all eigen |n\rangle with energy E_n = E. This ensures \hat{\rho} is idempotent (\hat{\rho}^2 = \hat{\rho}) and traces to unity, embodying a maximally mixed within the . The degeneracy, or number of accessible states, is given by \Omega(E, V, N) = \operatorname{Tr}(\hat{P}_E), which equals the dimension of the Hilbert subspace at energy E. This quantity plays the role of the classical phase space volume, scaled appropriately for quantum statistics. The microcanonical average of an observable \hat{A} is then the trace over this subspace: \langle \hat{A} \rangle = \operatorname{Tr}(\hat{\rho} \hat{A}) = \frac{1}{\Omega} \sum_{n: E_n = E} \langle n | \hat{A} | n \rangle. This expectation value is diagonal in the energy basis, reflecting the equal weighting of degenerate states. In the semiclassical limit of large systems (N \to \infty), the quantum g(E) = \frac{d\Omega}{dE} connects to the classical formulation via Weyl's law: g(E) \approx \omega(E), where \omega(E) is the classical \frac{1}{N! h^{3N}} \int d^{3N}q \, d^{3N}p \, \delta \bigl( H(\mathbf{q}, \mathbf{p}) - E \bigr). This approximation captures the leading-order smoothing of quantum fluctuations, aligning quantum degeneracy with classical counts while accounting for indistinguishability and quantization.

Thermodynamic Properties

Key Quantities

In the microcanonical ensemble, the U is fixed and equals the total E of the system by construction, as the ensemble consists of all microstates with precisely that value for given V and particle number N. The T emerges as a derived quantity from the S, defined as the inverse of the of with respect to at constant and particle number: \frac{1}{T} = \left( \frac{\partial S}{\partial E} \right)_{V,N}. This relation interprets temperature as the inverse slope of the entropy versus energy curve, indicating how rapidly the number of accessible microstates grows with energy; a steeper slope corresponds to a lower temperature. The pressure P is similarly obtained from the entropy as \frac{P}{T} = \left( \frac{\partial S}{\partial V} \right)_{E,N}, which connects the ensemble's fixed-energy description to mechanical work through volume changes, reflecting the system's tendency to expand into available . For systems where particle number can vary, the chemical potential \mu is given by \frac{\mu}{T} = -\left( \frac{\partial S}{\partial N} \right)_{E,V}, quantifying the change in entropy per added particle at fixed energy and volume, and thus the cost of incorporating additional particles into the system. Response functions, such as the heat capacity at constant volume C_V, follow from these definitions; specifically, C_V = \left( \frac{\partial E}{\partial T} \right)_V = -\frac{ \left( \frac{\partial S}{\partial E} \right)_{V,N}^2 }{ \left( \frac{\partial^2 S}{\partial E^2} \right)_{V,N} }, revealing that C_V depends on the curvature of the entropy-energy relation, which can yield negative values in finite systems like self-gravitating clusters where energy addition decreases temperature.

Entropy

In the microcanonical ensemble, the entropy S(E, V, N) is defined as S(E, V, N) = k \ln \Omega(E, V, N), where k is Boltzmann's constant and \Omega(E, V, N) denotes the number of accessible microstates consistent with the fixed total E, volume V, and particle number N. This expression connects the macroscopic thermodynamic to the microscopic multiplicity of states, establishing a foundational link between and . The logarithm ensures that is additive for independent systems, as the total multiplicity is the product of individual multiplicities, yielding S = S_1 + S_2 exactly for non-interacting subsystems. In the thermodynamic limit of large system size, the exhibits extensivity, approximating S \approx s(E/V, N/V) \, V, where s represents the density per unit volume. This property arises because the multiplicity \Omega grows exponentially with system size, \Omega \sim e^{S/k}, allowing subextensive corrections like surface effects or logarithmic terms to become negligible relative to the dominant extensive term. The third law of thermodynamics emerges naturally from this framework: as approaches zero, the system confines to the unique , where \Omega \to 1 and thus S \to 0. Fluctuations in the microcanonical ensemble are inherently constrained by the fixed , but considering a thin energy shell of width \Delta E around E, the relative variance satisfies \Delta E / E \sim 1/\sqrt{\Omega}. This leads to corresponding fluctuations \delta S / S \sim 1/\sqrt{\Omega}, which vanish exponentially in the since \Omega grows superexponentially with system size, ensuring the of thermodynamic quantities. From an information-theoretic perspective, the microcanonical ensemble corresponds to the that maximizes the Shannon H = -\sum_i p_i \ln p_i under the constraint of fixed total energy, resulting in uniform probabilities p_i = 1/\Omega over the accessible microstates. This maximization yields H = \ln \Omega = S/k, highlighting the equivalence between statistical and the measure of uncertainty or missing information in the ensemble.

Phase Transitions

In the microcanonical ensemble, phase transitions are identified through non-analyticities in the S(E) = k \ln \Omega(E), where \Omega(E) is the , leading to discontinuities or singularities in derived thermodynamic quantities such as T(E) = \left( \frac{\partial S}{\partial E} \right)^{-1} and specific heat C_V(E). Unlike the , where phase transitions manifest as smoother discontinuities in derivatives of the , the microcanonical formulation reveals sharper signatures due to the fixed energy constraint, including potential back-bending in caloric curves and negative heat capacities in finite systems. These features arise from the geometry of the surface, where convex intrusions signal transitions and kinks indicate second-order ones. For phase transitions, the microcanonical ensemble exhibits a characteristic back-bending in the T(E) curve, corresponding to a dip in S(E) that reflects the associated with coexistence between phases. This back-bending implies regions of negative specific heat, where increasing energy decreases temperature, analogous to the Clausius-Clapeyron equation through the equality of areas under the T(E) curve on either side of the transition. In finite systems, this behavior underscores ensemble inequivalence with the , where the transition appears as a bimodal rather than a direct caloric . However, in the thermodynamic limit, the ensembles become equivalent, and the microcanonical signatures align with ones. Second-order phase transitions in the microcanonical ensemble are marked by singularities in \Omega(E), such as essential singularities or branch points, which produce divergences in response functions like C_V(E) or without . These singularities stem from where the entropy's higher derivatives become infinite, leading to power-law behaviors in thermodynamic quantities near the transition energy. As C_V(E) = \left( \frac{\partial T}{\partial E} \right)_V^{-1}, such divergences occur when \left( \frac{\partial T}{\partial E} \right)_V \to 0, highlighting . A key mechanism for these signatures in finite isolated systems is the emergence of a bimodal energy distribution, which introduces multimodality in \Omega(E) and can yield negative specific heat, particularly in self-bound systems like atomic clusters or gravitational aggregates. For instance, in gravitational systems, the long-range attractive forces amplify this effect, allowing energy addition to cool the core while heating the halo, resulting in C_V < 0 over certain energy ranges. This phenomenon resolves in the thermodynamic limit, where infinite-volume constraints eliminate finite-size artifacts and restore positive definiteness.

Connections to Thermodynamics

Analogies

The microcanonical ensemble establishes a direct analogy to classical thermodynamics through the fundamental relation of thermodynamics, which emerges from the partial derivatives of the entropy function S(E, V, N). Specifically, the temperature T, pressure P, and chemical potential \mu are defined as $1/T = (\partial S / \partial E)_{V,N}, P/T = (\partial S / \partial V)_{E,N}, and -\mu/T = (\partial S / \partial N)_{E,V}, leading to the differential form dS = (1/T) dE + (P/T) dV - (\mu/T) dN. Rearranging this yields the standard thermodynamic identity dE = T dS - P dV + \mu dN, which encapsulates the first and second laws in the energy representation and mirrors the macroscopic thermodynamic potential for isolated systems. From this entropy-based formulation, Maxwell relations arise naturally as consequences of the exact differential nature of dS, analogous to those in phenomenological thermodynamics. For instance, the mixed second partial derivatives imply (\partial T / \partial V)_S = - (\partial P / \partial S)_V, providing symmetry relations between thermodynamic response functions that ensure consistency across ensembles. These relations highlight how the microcanonical description reproduces the interconnectedness of thermodynamic variables observed empirically. The Gibbs-Duhem relation further bridges the microcanonical framework to intensive thermodynamic variables. At constant , the relation integrates to S dT = V dP - N d\mu, expressing the interdependence of T, P, and \mu in states, much like in the Euler homogeneous function approach to extensive properties. This form underscores the ensemble's ability to capture constraints without invoking fluctuations. Historically, Ludwig Boltzmann's H-theorem provides a foundational analogy linking the microcanonical ensemble to the second law of thermodynamics. The theorem demonstrates that, under the assumption of molecular chaos, the H-function (related to the negative ) decreases toward its minimum, corresponding to the over accessible states in the microcanonical ensemble, thereby maximizing S at and justifying the in isolated systems. This probabilistic interpretation resolves the apparent reversibility paradox in dynamics by positing the microcanonical distribution as the most probable state. The concavity of the function S(E, V, N) in the microcanonical ensemble ensures thermodynamic stability, analogous to the stability criteria in classical . This property implies that small perturbations in extensive variables lead to restorative responses, preventing or instabilities, and supports the use of Legendre transforms to generate other thermodynamic potentials, such as the F(T, V, N) = E - T S, for systems at fixed temperature.

Equivalence with Other Ensembles

In the thermodynamic limit, where the number of particles N and volume V approach infinity while keeping the density N/V fixed, the microcanonical ensemble becomes equivalent to the canonical ensemble. In this regime, energy fluctuations in the canonical ensemble vanish relative to the mean energy, such that the average energy \langle E \rangle_{\text{canonical}} approximates the fixed energy E_{\text{micro}} of the microcanonical ensemble, and the entropies satisfy S_{\text{canonical}} \approx S_{\text{micro}}. This equivalence extends to thermodynamic observables like pressure and chemical potential, ensuring consistent predictions for large systems. The connection between the ensembles is formalized through the saddle-point approximation to the partition function Z(\beta), which is expressed as an integral over : Z(\beta) \approx \int dE \, \omega(E) e^{-\beta E}, where \omega(E) is the microcanonical and \beta = 1/T. This integral is dominated by contributions near the energy E^* satisfying \partial \ln \omega / \partial E = \beta, yielding the canonical free energy and thermodynamics that match the microcanonical results in the large-system limit. However, equivalence does not hold universally; inequivalence arises in finite systems or those with long-range interactions, such as gravitational systems. In these cases, the microcanonical ensemble can exhibit negative —where temperature decreases as energy increases—while the cannot, due to the absence of such fluctuations in the latter. For example, in self-gravitating systems, this leads to distinct phase diagrams between ensembles. The grand canonical ensemble, which allows particle number fluctuations, also equivalents to the microcanonical in the thermodynamic limit when chemical potential \mu fluctuations are negligible, typically for large N. This holds for systems where particle exchange does not dominate, ensuring matching averages for density and energy. In modern contexts, particularly quantum many-body simulations since the , the microcanonical ensemble remains foundational for studying isolated systems, such as in lattice models and strongly correlated electrons, where exact dynamics reveal phase transitions inaccessible in other ensembles. Advances in numerical methods, like free-cumulant approaches, highlight its role in probing thermalization and inequivalence in quantum settings.

Examples

Ideal Gas

The microcanonical ensemble provides an exact framework for computing thermodynamic properties of a classical ideal gas consisting of N indistinguishable monatomic particles of mass m confined to a volume V with fixed total energy E. The density of states \Omega(E, V, N) represents the phase-space volume of the energy shell of infinitesimal thickness at energy E, normalized by the classical phase-space factor h^{3N} N! to account for indistinguishability and units. For the ideal gas Hamiltonian H = \sum_{i=1}^N \mathbf{p}_i^2 / (2m), the position integrals yield V^N, while the momentum integrals correspond to the surface area of a $3N-dimensional hypersphere of radius \sqrt{2mE}. This evaluation gives the explicit expression \Omega(E, V, N) = \frac{V^N}{N! \, h^{3N}} \frac{(2 \pi m)^{3N/2} E^{3N/2 - 1}}{\Gamma(3N/2)}. This formula follows from the general definition of \Omega in the classical microcanonical ensemble applied to non-interacting particles. The entropy S is given by S = k \ln \Omega(E, V, N), where k is Boltzmann's constant. For large N, Stirling's approximation \ln N! \approx N \ln N - N and the asymptotic form of the gamma function \ln \Gamma(3N/2) \approx (3N/2 - 1) \ln(3N/2) - 3N/2 simplify \ln \Omega to its leading terms, yielding the Sackur-Tetrode equation in microcanonical form: S = k N \left[ \ln \left( \frac{V}{N} \left( \frac{4 \pi m E}{3 N h^2} \right)^{3/2} \right) + \frac{5}{2} \right]. This seminal expression for the absolute entropy of a monatomic ideal gas incorporates quantum considerations via Planck's constant h while remaining classical, and it was independently derived by Sackur and Tetrode to resolve the entropy constant in the ideal gas law. The temperature T emerges from the thermodynamic relation $1/T = (\partial S / \partial E)_{V,N} = (1/k) (\partial \ln \Omega / \partial E)_{V,N}. Differentiating \ln \Omega with respect to E gives (\partial \ln \Omega / \partial E)_{V,N} = (3N/2 - 1)/E \approx 3N/(2E) in the thermodynamic limit N \to \infty, leading to T = 2E / (3 N k) or equivalently E = (3/2) N k T. This confirms the equipartition theorem, assigning (1/2) k T per quadratic degree of freedom for the $3N translational modes. The pressure P follows from P = T (\partial S / \partial V)_{E,N} = [k](/page/K) T (\partial \ln \Omega / \partial V)_{E,N}. Since \ln \Omega \propto N \ln V, the derivative yields P = N [k](/page/K) T / V, recovering the . Substituting E = (3/2) N [k](/page/K) T further implies P V = (2/3) E, which aligns with the for a non-interacting gas where dominates. These results hold under the assumptions of non-interacting point particles with no internal , in the classical high-temperature regime where quantum degeneracy effects (such as Bose-Einstein condensation or Fermi degeneracy) are negligible—specifically, when the thermal de Broglie wavelength \lambda = h / \sqrt{2 \pi m [k](/page/K) T} satisfies \lambda^3 \ll V/N.

Ideal Gas in Gravitational Field

The in a uniform serves as an extension of the homogeneous case, introducing spatial inhomogeneity due to the external potential while maintaining the microcanonical framework of fixed E, particle number N, and container dimensions. The is confined to a cylindrical volume with L along the z-direction (where $0 < z < L) and cross-sectional area A, such that the volume V = A L. The for N non-interacting particles of m is given by H = \sum_{i=1}^N \left( \frac{\mathbf{p}_i^2}{2m} + m g z_i \right), where \mathbf{p}_i is the of the i-th particle, g is the , and z_i is its coordinate. The microcanonical \Omega(E, V, N) can be approximated in the using the method or saddle-point integration over the volume, accounting for the contribution. This yields single-particle distributions that, for large N, approach forms: the distribution becomes Maxwellian, and the height distribution follows the \rho(z) \propto e^{-m g z / kT}, where T is the emerging from the saddle-point . Unlike the uniform without , the effective volume available to particles is reduced by the weighting, leading to a gradient that increases toward the bottom of the container. The entropy S = k \ln \Omega in this setup is analogous to the Sackur-Tetrode expression for the homogeneous but incorporates a gravitational correction that diminishes the effective phase space volume, roughly S \approx N k \left[ \ln \left( \frac{V_{\text{eff}}}{N} \left( \frac{4\pi m E_k}{3 N h^2} \right)^{3/2} \right) + \frac{5}{2} \right], where V_{\text{eff}} accounts for the gravitational compression and E_k is the total . The remains defined via the as T = \frac{2}{3} \frac{\langle K \rangle}{N k}, with \langle K \rangle the average , preserving the despite the inhomogeneity; numerical studies confirm equivalence to the in the large-N limit. In the self-gravitating limit for large N, where the external field is replaced by mutual gravitational interactions, the microcanonical ensemble reveals potential for and the gravothermal catastrophe, characterized by negative specific heat where energy loss leads to core heating and collapse. This arises because the implies $2K + W = 3 P V (with W < 0 the ), allowing dE/dT < 0 for bound systems with E < -0.335 G M^2 / R. Such behavior highlights ensemble inequivalence: the predicts a first-order phase transition to a dilute, nearly uniform gaseous phase coexisting with a dense , whereas the microcanonical prohibits stable uniform due to the fixed-energy .

Paramagnetic Spin System

A paradigmatic quantum example of the microcanonical ensemble is provided by a of N non-interacting particles, each with \mu, subjected to an external B directed along the z-axis. The energy levels for each spin are \pm \mu B, corresponding to alignment (down, lower energy -\mu B) or anti-alignment (up, higher energy +\mu B) with the field. The total energy E of the system is fixed and given by E = -\mu B M, where M is the net number of spins aligned down (i.e., M = N_- - N_+, with N_- and N_+ the numbers of down and up spins, respectively, satisfying N_- + N_+ = N). Thus, N_- = (N + M)/2 and N_+ = (N - M)/2, with |M| \leq N. The degeneracy \Omega(E), or number of microstates consistent with the fixed energy E, is the number of ways to choose N_- spins out of N to be down, yielding the \Omega(E) = \binom{N}{\frac{N + M}{2}} = \frac{N!}{\left( \frac{N + M}{2} \right)! \left( \frac{N - M}{2} \right)!}. This exact expression counts the accessible configurations in the energy shell at E. For large N, simplifies the S = k \ln \Omega(E), where k is Boltzmann's constant, to \frac{S}{k} \approx N h(p), \quad h(p) = -p \ln p - (1 - p) \ln (1 - p), with p = (N + M)/(2N) = (1 + m)/2 the fraction of down and m = M/N the reduced (|m| \leq 1). This h(p) captures the configurational disorder, maximized at p = 1/2 (m = 0) where S/k \approx N \ln 2, corresponding to maximum degeneracy. The T emerges from the thermodynamic $1/T = (\partial S / \partial E)_{N}. With E = -\mu B M and m = M/N, \partial E / \partial m = -\mu B N, so \frac{1}{T} = \frac{\partial S / \partial m}{\partial E / \partial m} = -\frac{1}{\mu B N} \frac{\partial S}{\partial m}. Using S \approx -N k \left[ \frac{1+m}{2} \ln \frac{1+m}{2} + \frac{1-m}{2} \ln \frac{1-m}{2} \right], the is \partial S / \partial m = -k N \artanh(m), yielding (with k = 1 for simplicity) \frac{1}{T} = \frac{1}{\mu B} \artanh\left( \frac{M}{N} \right). Here, \artanh(x) = \frac{1}{2} \ln \left( \frac{1 + x}{1 - x} \right). As M \to 0 (zero net , high disorder), T \to \infty; conversely, as |M| \to N (full alignment, low disorder), T \to 0^+. This inverts the result m = \tanh(\mu B / T), highlighting ensemble equivalence in the . In the high-temperature limit (T \gg \mu B / k, small |m|), expanding \artanh(m) \approx m yields m \approx (\mu B)/T, or total magnetization M \approx N \mu (B / T), mimicking the Curie law M = (C B)/T with Curie constant C = N \mu^2 / k. However, the microcanonical formulation reveals fluctuations absent in mean-field approximations: for fixed E (fixed m), are suppressed, but the sharp variation of m with T—from near-zero at high T to near-unity at low T—exhibits transition-like behavior, analogous to a paramagnetic response without true due to non-interacting spins. This toy model illustrates how fixed-energy constraints amplify alignment at low effective temperatures, contrasting continuous classical systems.

References

  1. [1]
    11. The Microcanonical Ensemble - Chemistry LibreTexts
    Jan 29, 2023 · We first consider the microcanonical ensemble because it is the one directly defined in postulate II of statistical mechanics.
  2. [2]
    [PDF] 8.044s13 Notes on the Microcanonical Ensemble
    It is a logarithmic measure of the microscopic degeneracy associated with a macroscopic. (that is, thermodynamic) state of the system. • k is Boltzmann's ...
  3. [3]
    [PDF] Chapter 8 Microcanonical ensemble
    By definition, such a system exchanges neither particles nor energy with the surroundings. Uniform distribution of microstates. The assumption, that thermal ...
  4. [4]
    Microcanonical Ensemble - an overview | ScienceDirect Topics
    Even though Gibbs first considered the canonical ensemble and then derived the microcanonical one from it, the following formal development of statistical ...
  5. [5]
    Elementary Principles in Statistical Mechanics/Chapter X - Wikisource
    Sep 17, 2023 · Chapter X. On a distribution in phase called microcanonical in which all the systems have the same energy. Chapter XI. Maximum and minimum ...<|separator|>
  6. [6]
    Microcanonical Ensemble
    Abstract In this chapter the microcanonical ensemble, which describes an isolated system, is introduced. In this ensemble the entropy and the temperature, i.e. ...Missing: applicability | Show results with:applicability<|control11|><|separator|>
  7. [7]
    [PDF] Thermodynamic laws in isolated systems - arXiv
    After recalling basic assumptions and definitions for the microcanonical ensemble (Sec. IIA), we will sum- marize the most commonly considered microcanonical.
  8. [8]
    [PDF] Statistics of non-interacting bosons and fermions in micro-canonical ...
    There are circumstances, however, where the system under consideration may be considered as being isolated (micro-canonical ensemble). This paper first reviews ...
  9. [9]
    Inhomogeneity of Local Temperature in Small Clusters in ...
    Jul 6, 2007 · ... dynamics simulations, the local temperatures of atoms in small, isolated crystalline clusters in microcanonical equilibrium are not uniform.
  10. [10]
    [PDF] Finite-size behaviour of the microcanonical specific heat - arXiv
    In the microcanonical ensemble various definitions of the temperature are possible, the different expressions becoming equivalent in the thermodynamic limit.
  11. [11]
    [PDF] The microcanonical thermodynamics of finite systems - arXiv
    Otherwise, in a canonical ensemble, the system would fluctuate between the two pure phases separated by an energy difference, ∆E ≈ Elatent, on the order of the ...
  12. [12]
    Large Finite-Size Effects of Discrete Systems in Microcanonical ...
    Abstract. We describe a microcanonical ensemble theory with a rigorous statistical mechanies basis which can be used for exact analytic calculations or.
  13. [13]
    [PDF] arXiv:quant-ph/0511162v3 18 Apr 2006
    The quantum microcanonical ensemble introduced here is applicable to any isolated finite-dimensional quantum system for which the ideal gas approximation is ...
  14. [14]
    [PDF] Exact results in the microcanonical formalism - arXiv
    Sep 4, 2015 · The microcanonical formalism allows the application of. Statistical Mechanics to isolated small systems, but there are fundamental principles of.
  15. [15]
    [PDF] Elementray Principles in Statistical Mechanics. - Project Gutenberg
    Jan 22, 2016 · ... microcanonical ... Mechanics, by Josiah Willard Gibbs. *** END OF THIS PROJECT GUTENBERG EBOOK ELEMENTARY PRINCIPLES STATISTICAL MECHANICS ***.
  16. [16]
    [PDF] landau - lifshitz
    Statistical physics, often called for brevity simply statistics, consists in the study of the special laws which govern the behaviour and properties of mac-.Missing: microcanonical | Show results with:microcanonical
  17. [17]
    Microcanonical Ensemble in Quantum Statistical Mechanics
    Canonical and microcanonical Monte Carlo simulations of lattice-gas mixtures. J. Chem. Phys. (October 2006). Online ISSN 1089-7658; Print ISSN 0022-2488 ...
  18. [18]
    [PDF] On quantum microcanonical equilibrium - arXiv
    Dec 14, 2006 · Abstract. A quantum microcanonical postulate is proposed as a basis for the equilibrium properties of small quantum systems.
  19. [19]
    [PDF] Statistical Mechanics Contents 1 Thermodynamics
    Mar 10, 2011 · 4.1 Quantum mechanics, Weyl's law . . . . . 22. 4.2 Spin and ... g(α) is the density of states (effectively Ω(α)). In the case of ...
  20. [20]
    [PDF] Lecture 7: Ensembles
    Ensembles are the possible microstates a system could be in, sharing a macroscopic property. Examples include microcanonical, canonical, Gibbs, and grand ...
  21. [21]
    [PDF] Statistical Physics - DAMTP
    definition of the heat capacity (1.9) in the microcanonical ensemble. In the canonical ensemble, where the energy is not fixed, the corresponding definition is.
  22. [22]
    Quantum Statistical Mechanics
    The microcanonical ensemble will then involve S(E,N). If we connect our system to a heat and particle bath, we find pi=e−βϵi+βμniZ,. where βμ=−1kB∂S∂Nβ=1 ...
  23. [23]
    [PDF] Statistical Physics
    ... ∂S/∂N = −µ/T and introduced the grand canonical potential which can be expressed through the grand partition func- tion. Ω(T,V,µ) = −T ln. X. N exp(µN/T). X.
  24. [24]
    [PDF] Statistical Mechanics of Ideal Gases - JILA
    In statistical mechanics it is convenient to work with internal energy U ... (c) Pressure. 34. Page 35. According to the standard relation ∂S/∂V = P/T, eq.
  25. [25]
    [PDF] Quantum confinement and negative heat capacity - Purdue Chemistry
    ... specific heat of a system is strictly non-negative. ... which are described by the Canonical ensemble, may not have a negative heat capacity. ... ing to the ...
  26. [26]
    Translation of Ludwig Boltzmann's Paper “On the Relationship ...
    Confusingly, Planck later chose to write Boltzmann's equation for entropy as S=klnW+constant [6]. Crucially, Ω is an extensive quantity, which leads to ...
  27. [27]
    [PDF] Ensembles - UMD Physics
    In the microcanonical ensemble temperature measures the energy dependence of the multiplicity function for isolated systems. For isolated systems, you specify ...Missing: adiabatic | Show results with:adiabatic
  28. [28]
    Maximum Entropy Principle for the Microcanonical Ensemble - arXiv
    Sep 7, 2007 · We derive the microcanonical ensemble from the Maximum Entropy Principle (MEP) using the phase space volume entropy of P. Hertz.
  29. [29]
  30. [30]
    First order phase transitions in the canonical and the microcanonical ...
    On the other hand, the distinction between first and a second order phase transition is rather direct if based on the microcanonical density of states which is ...
  31. [31]
    On the negative specific heat paradox - Astrophysics Data System
    A system which can show negative heat capacity can not be divided into equal weakly coupled parts. Imagine a system of negative heat capacity C1 = - I C1 I ...
  32. [32]
    [cond-mat/9911257] Phase Transitions in "Small" systems - arXiv
    Nov 17, 1999 · Authors:D.H.E.Gross, E.Votyakov. View a PDF of the paper titled Phase Transitions in "Small" systems, by D.H.E.Gross and E.Votyakov. View PDF.
  33. [33]
    [PDF] Untitled
    the microcanonical ensemble ... properties of the system described by S(E, V, N). ... This will lead to physically interesting relations, called Maxwell relations.
  34. [34]
    [PDF] Lecture 8: Free energy
    The Euler equation at the Gibbs-Duhem equation hold for almost all statistical mechanical systems. Keep in mind however that the extensivity of entropy is ...
  35. [35]
    Boltzmann's Work in Statistical Physics
    Nov 17, 2004 · Boltzmann claimed that the H-theorem provided the desired theorem from mechanics corresponding to the second law. However, this claim came ...
  36. [36]
    Thermodynamic equilibrium and its stability for microcanonical ...
    Aug 19, 2005 · It is generally assumed that the thermodynamic stability of equilibrium states is reflected by the concavity of entropy.Missing: ensemble | Show results with:ensemble
  37. [37]
    [1403.6608] Equivalence and nonequivalence of ensembles - arXiv
    Mar 26, 2014 · We present general and rigorous results showing that the microcanonical and canonical ensembles are equivalent at all three levels of description considered in ...
  38. [38]
    [PDF] Conditional Equilibrium and the Equivalence of Microcanonical and ...
    C.E. and Gibbs states; it was a fundamental belief of the founders of statistical mechanics that in the thermodynamic limit the various ensembles yield the same.
  39. [39]
    [PDF] Chapter 9 Canonical ensemble
    Equivalence of ensembles. In the thermodynamic limit the average value of an ob- servable is in general independent of the ensemble (microcanonical or canonical) ...
  40. [40]
    The equivalence between the canonical and microcanonical ...
    Aug 1, 2007 · A straightforward technique is suggested that demonstrates that a microcanonical ensemble and canonical ensemble behave in exactly the same way
  41. [41]
    [PDF] Derivation of Canonical Distribution from Microcanonical
    If we use the above asymptotic analysis, the saddle points for numerator and denominator are different. But in fact, for N ≫ 1, the two saddle points are nearly ...
  42. [42]
    [PDF] Statistical Mechanics Lecture set 3: Canonical Ensemble
    - Microcanonical ensemble: fixed E,V,N → S(E,V,N). - Canonical ensemble: fixed T,V,N → F(T,V,N) = E − TS. - Grandcanonical ensemble: fixed T,V,µ → Π(T,V,µ) ...
  43. [43]
    Inequivalence of Ensembles in a System with Long-Range Interactions
    Jun 29, 2001 · In this region the microcanonical ensemble exhibits energy ranges with negative specific heat and temperature jumps at transition energies.Missing: capacity | Show results with:capacity
  44. [44]
    Ensemble inequivalence in systems with long-range interactions
    Dec 7, 2001 · In particular it has been recently shown that negative specific heat can arise in the microcanonical ensemble in the thermodynamic limit for ...Missing: capacity | Show results with:capacity
  45. [45]
    Ensemble inequivalence in systems with long-range interactions
    Ensemble inequivalence has been observed in several systems. In particular it has been recently shown that negative specific heat can arise in the ...
  46. [46]
    Microcanonical free cumulants in lattice systems | Phys. Rev. B
    Feb 7, 2025 · In this paper, we present a detailed discussion of the free-cumulants approach to many-body dynamics within the microcanonical ensemble.
  47. [47]
    Revealing microcanonical phases and phase transitions of strongly ...
    Dec 13, 2023 · We propose a method to study the quantum thermodynamics of strongly correlated electrons from quantum dynamics.
  48. [48]
    Microcanonical single-particle distributions for an ideal gas in a ...
    The single-particle energy, velocity and height probability distributions for an ideal gas in a gravitational field are obtained within the microcanonical ...
  49. [49]
    Microcanonical single-particle distributions for an ideal gas in a ...
    We have derived, in the microcanonical ensemble framework, the single-particle height, velocity and energy distributions of an ideal gas in a gravitational.
  50. [50]
    Microcanonical ensemble study of a gas column under gravity
    The study of a classical ideal gas column of finite height H in a uniform gravitational field g is made by the microcanonical ensemble at energy E. The primary ...
  51. [51]
    The thermodynamics of self-gravitating systems is a fascinating ...
    This "isothermal collapse'' is the counterpart of the "gravothermal catastrophe'' in the microcanonical ensemble (see Figs. 2, 3). In Sect. 3, we study the ...
  52. [52]
    None
    Nothing is retrieved...<|control11|><|separator|>
  53. [53]
    [PDF] Quantum microcanonical ensemble 1 Macrostate vs. microstates
    It turns out that because of this spin, when particles are put in a magnetic field they act like magnetic dipoles: it is energetically favorable to align ...
  54. [54]
    [PDF] Statistical Mechanics Lecture set 2: Microcanonical Ensemble Abstract
    May 2, 2021 · In this set of lectures we will introduce and discuss the microcanonical ensemble description of quantum and classical statistical mechanics. We ...