Fact-checked by Grok 2 weeks ago

Virial theorem

The Virial theorem is a fundamental principle in classical and that establishes a relationship between the time-averaged and a corresponding "virial" quantity derived from the system's , applicable to stable, bounded systems of interacting particles bound by conservative forces. Formulated by in 1870 as part of his work on the kinetic theory of heat, the theorem provides that for a system in or periodic motion, the time average of the \langle T \rangle satisfies $2 \langle T \rangle = \langle \sum_i \mathbf{r}_i \cdot \mathbf{F}_i \rangle, where \mathbf{r}_i are position vectors and \mathbf{F}_i are the forces acting on the particles; for potentials homogeneous of degree n (such as n = -1 for gravitational or interactions), this simplifies to $2 \langle T \rangle = n \langle V \rangle. Originally derived by Clausius to connect molecular motions to macroscopic thermodynamic properties like in ideal gases—yielding the relation PV = \frac{2}{3} \langle T \rangle when internal forces vanish—the theorem extends to diverse fields beyond its thermodynamic roots. In , it arises from averaging the time derivative of the virial function over long periods, assuming or bounded orbits, and holds for both discrete particle systems and continua under suitable boundary conditions. The theorem's applications span , where it estimates masses of star clusters and galaxies by equating observed velocities ( proxies) to energies, as first applied by in 1933 to infer unseen "" in cluster; in , it relates internal thermal support to gravitational binding for in stars; and in , an analogous quantum version links expectation values of kinetic and potential energies in bound states, underpinning scaling relations in and molecular bonding. These uses highlight its role in predicting stability, energy partitioning, and macroscopic behavior from microscopic dynamics across scales, from gases to cosmological structures.

Historical Development

Origins and Early Formulations

The virial theorem emerged in the context of 19th-century and the burgeoning theory of , where efforts focused on bridging microscopic particle motions with macroscopic thermodynamic phenomena. In 1870, presented the foundational ideas in his lecture "On a Theorem Applicable to " to the Niederrheinische Gesellschaft für Natur- und Heilkunde. This work built on his earlier 1862 on the theory of , aiming to relate as internal (twice the modern ) to mechanical work through force-position correlations in stable systems. Clausius introduced the term "virial" to denote the time-averaged quantity \sum (X x + Y y + Z z), where X, Y, Z are the components on a particle at coordinates x, y, z, derived from the Latin vis for . For a of particles interacting via pairwise central forces in steady-state motion—characterized by periodic or quasi-periodic trajectories without overall or —the theorem asserts that the time average of the (twice the modern ) equals this virial. Clausius stated the core result as: "The mean of the is equal to its virial," expressed mathematically as \bar{v} = \sum (X x + Y y + Z z). This formulation resonated with contemporary mechanics, including Lagrangian approaches to constrained systems, by providing a scalar relation amenable to averaging over long times in ergodic-like motions. In early statistical mechanics, it offered a key tool for gases, linking internal heat to the internal virial from pairwise potentials \phi(r) (involving terms like r \frac{d\phi}{dr}) and external pressure-volume work, with the pressure contribution equaling $3 p v in three dimensions for ideal gases without internal forces—thus illuminating equipartition principles for molecular collisions without requiring full probability distributions.

Key Contributions and Evolution

In 1911, applied foundational ideas related to the virial theorem through his investigations of periodic systems and in self-gravitating configurations, laying groundwork for its application to dynamic equilibria. Building on Rudolf Clausius's original 1870 concept of the virial as a measure of molecular forces in gases, Poincaré's work emphasized the balance of forces in periodic motions, influencing later formulations for bounded systems. A significant 20th-century advancement came in 1919 when James Jeans clarified the theorem's application to gravitational systems by employing time averages, enabling precise relations between kinetic and potential energies in . This refinement, detailed in Jeans's Problems of Cosmogony and Stellar Dynamics, addressed limitations in earlier scalar approaches and facilitated broader use in . By the 1920s, the theorem gained widespread adoption in , particularly through Jeans's and Arthur Eddington's extensions to stellar structures and galactic , marking its transition from theoretical to observational analysis. The theorem's integration into statistical mechanics evolved prominently through and , who incorporated it to support the , justifying the equivalence of time averages and averages for thermodynamic properties. This development, evident in Gibbs's foundational texts on theory, allowed the virial theorem to underpin derivations of equations of state from partition functions, solidifying its role in justifying macroscopic behavior from microscopic dynamics. Post-1950s refinements extended the theorem to plasma physics, where Andrei Shafranov's 1966 virial analysis demonstrated that self-generated magnetic fields alone cannot confine plasmas, necessitating external supports for equilibrium—a key insight for fusion research. Concurrently, in the 1960s, numerical validations emerged through early N-body simulations, such as those by Sebastian von Hoerner (1960) and Sverre Aarseth (1963), which tested the theorem's predictions in computational models of gravitational systems, confirming its utility for verifying relaxation and virialization in particle ensembles.

Classical Statement and Derivation

General Theorem for Bounded Systems

The classical virial theorem, first formulated by Rudolf Clausius in 1870, applies to stable systems of discrete particles interacting via internal forces, relating the time-averaged kinetic energy to the forces acting on the particles. For such a system in equilibrium, the theorem states that twice the average total kinetic energy equals minus the average of the virial, defined as the sum over all particles of the scalar product of their position vectors and the total forces on them. In , this is expressed as $2 \langle T \rangle = -\left\langle \sum_i \mathbf{r}_i \cdot \mathbf{F}_i \right\rangle, where T = \sum_i \frac{1}{2} m_i |\mathbf{v}_i|^2 is the of the , \mathbf{r}_i is the position vector of the i-th particle relative to the center of mass, \mathbf{F}_i is the on that particle (typically from pairwise interactions), and \langle \cdot \rangle denotes the time average taken over a long interval or the period of bounded motion. This relation holds under the assumptions of bounded motion, where particle positions and velocities remain confined within finite limits to ensure the averages converge, and time-independent potentials that derive from conservative, pairwise forces between particles. The theorem is grounded in Newtonian mechanics through the , though contemporary derivations frequently employ formulations for broader applicability to constrained or generalized coordinate systems. When the potential energy V is a of the coordinates of degree n (satisfying V(\lambda \mathbf{r}) = \lambda^n V(\mathbf{r}) for scalar \lambda > 0), the virial theorem connects directly to the as $2 \langle T \rangle = n \langle V \rangle. For inverse-square law forces, such as those in gravitational or interactions, the potential scales as n = -1, yielding $2 \langle T \rangle + \langle V \rangle = 0; thus, the magnitude of the average equals half that of the average , providing a key relation for bound systems. A representative application arises in the two-body central problem with an inverse-square potential, as in the Keplerian description of planetary . For elliptical bound , the virial theorem confirms $2 \langle T \rangle + \langle V \rangle = 0, implying a negative total E = \langle T \rangle + \langle V \rangle = -\langle T \rangle that characterizes the stability of the .

via Scalar Virial

The scalar virial quantity for a of N particles is defined as G = \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{p}_i, where \mathbf{r}_i is the position vector and \mathbf{p}_i is the momentum vector of the i-th particle. To derive the time evolution of G, differentiate with respect to time using , \dot{\mathbf{r}}_i = \frac{\partial H}{\partial \mathbf{p}_i} = \mathbf{v}_i and \dot{\mathbf{p}}_i = -\frac{\partial H}{\partial \mathbf{r}_i} = \mathbf{F}_i, where \mathbf{v}_i is the and \mathbf{F}_i is the total on the i-th particle. This yields \frac{dG}{dt} = \sum_{i=1}^N \left( \dot{\mathbf{r}}_i \cdot \mathbf{p}_i + \mathbf{r}_i \cdot \dot{\mathbf{p}}_i \right) = \sum_{i=1}^N \left( \mathbf{v}_i \cdot \mathbf{p}_i + \mathbf{r}_i \cdot \mathbf{F}_i \right). The first term simplifies to twice the T = \frac{1}{2} \sum_{i=1}^N m_i v_i^2 = \frac{1}{2} \sum_{i=1}^N \frac{p_i^2}{m_i}, since \mathbf{v}_i \cdot \mathbf{p}_i = m_i v_i^2, giving \sum \mathbf{v}_i \cdot \mathbf{p}_i = 2T. Thus, \frac{dG}{dt} = 2T + \sum_{i=1}^N \mathbf{F}_i \cdot \mathbf{r}_i.[13][1] For a bounded in , where the particles are confined to a finite region and the motion is (e.g., periodic or quasi-periodic), consider the time over a long interval \tau: \left\langle \frac{dG}{dt} \right\rangle = \lim_{\tau \to \infty} \frac{1}{\tau} \int_0^\tau \frac{dG}{dt} \, dt = \lim_{\tau \to \infty} \frac{G(\tau) - G(0)}{\tau}. Since the system is bounded, G remains finite and does not grow indefinitely with time, so the limit vanishes: \left\langle \frac{dG}{dt} \right\rangle = 0. Therefore, $2 \langle T \rangle + \left\langle \sum_{i=1}^N \mathbf{F}_i \cdot \mathbf{r}_i \right\rangle = 0, or equivalently, $2 \langle T \rangle = - \left\langle \sum_{i=1}^N \mathbf{F}_i \cdot \mathbf{r}_i \right\rangle.[13][1][14] This relation holds under the assumption of bounded motion, which ensures the surface terms associated with any confining potential (such as walls of a ) vanish in the averaging process, as the particles do not escape to and contributions from the boundaries average to zero over long times. In , the time average can be equated to an ensemble average over via the ergodic theorem, which posits that for an ergodic system, the time spent in any accessible region is proportional to its volume, allowing \langle \cdot \rangle to represent either temporal or statistical averaging equivalently.

Power-Law Interaction Cases

A significant class of interaction potentials to which the virial theorem applies straightforwardly consists of power-law forms, where the potential energy scales as V \propto r^n with r the interparticle separation and n the homogeneity degree. Such potentials are homogeneous functions, satisfying V(\lambda \mathbf{r}) = \lambda^n V(\mathbf{r}) for scaling factor \lambda > 0. Euler's theorem for homogeneous functions then implies that \sum_i \mathbf{r}_i \cdot \nabla_{\mathbf{r}_i} V = n V, providing a direct link to the force term in the virial expression. In the classical virial theorem for bounded systems in , the time average of the virial G = \sum_i \mathbf{p}_i \cdot \mathbf{r}_i yields $2 \langle T \rangle + \langle \sum_i \mathbf{r}_i \cdot \mathbf{F}_i \rangle = 0, where T is the total and \mathbf{F}_i = -\nabla_{\mathbf{r}_i} V the forces. For power-law potentials, \sum_i \mathbf{r}_i \cdot \mathbf{F}_i = - \sum_i \mathbf{r}_i \cdot \nabla_{\mathbf{r}_i} V = -n V, so the theorem simplifies to the general $2 \langle T \rangle = n \langle V \rangle, where averages are over sufficiently long times for ergodic systems. This scaling holds for both - and multi-particle systems, revealing how kinetic and potential energies according to the exponent. For multi-particle systems with pairwise interactions, the total potential is V = \frac{1}{2} \sum_{i \neq j} v(|\mathbf{r}_i - \mathbf{r}_j|), where each pair potential v(r_{ij}) \propto r_{ij}^n is homogeneous of degree n. Under a uniform scaling \mathbf{r}_k \to \lambda \mathbf{r}_k for all particles, V \to \lambda^n V, confirming the overall homogeneity and thus the same virial relation $2 \langle T \rangle = n \langle V \rangle. The detailed expansion follows from applying Euler's theorem to each pairwise term: \sum_i \mathbf{r}_i \cdot \nabla_{\mathbf{r}_i} V = \sum_{i < j} \left[ (\mathbf{r}_i - \mathbf{r}_j) \cdot \nabla_{r_{ij}} v(r_{ij}) \right] = n V, leading to the force virial \langle \sum_i \mathbf{r}_i \cdot \mathbf{F}_i \rangle = -n \langle V \rangle. This form is particularly useful for systems like self-gravitating clusters or molecular gases with inverse-power interactions. Specific cases illustrate the theorem's implications. For gravitational or electrostatic interactions following an inverse-square force law, the potential scales as n = -1 (V \propto -1/r), yielding $2 \langle T \rangle = -\langle V \rangle and total energy \langle E \rangle = \langle T \rangle + \langle V \rangle = -\langle T \rangle, which bounds the system for negative energies. For harmonic oscillators, where V \propto r^2 so n = 2, the relation becomes $2 \langle T \rangle = 2 \langle V \rangle or \langle T \rangle = \langle V \rangle, equi-partitioning the energy. Linear force laws, with V \propto r and n = 1, give $2 \langle T \rangle = \langle V \rangle, as seen in uniform field problems like charged particles in constant electric fields. The ratio $2 \langle T \rangle / (n \langle V \rangle) = 1 serves as a dimensionless virial coefficient, quantifying departure from equilibrium in dynamical systems; values near unity indicate stability, while deviations signal collapse or expansion, aiding analysis in contexts like stellar dynamics or cluster stability.

Applications in Classical Systems

Equilibrium in Gases and Oscillators

The virial theorem finds a fundamental application in the equilibrium of an ideal gas confined to a volume V under pressure P. For a system of non-interacting particles, the theorem relates the time-averaged total kinetic energy \langle T \rangle to the pressure through the equation $2 \langle T \rangle = 3 P V. This relation arises from the scalar virial, where the internal forces between particles vanish, and the only contributions come from the impulsive forces during molecular collisions with the container walls. Rearranging yields the ideal gas law in terms of kinetic energy: P V = \frac{2}{3} \langle T \rangle, establishing a direct link between macroscopic pressure and microscopic motion. In classical statistical mechanics, this result extends to the microcanonical ensemble for N particles in a fixed volume, where ensemble averages replace time averages for ergodic systems. The virial theorem then equates $2 \langle T \rangle to the sum of wall pressure contributions, confirming the equipartition of kinetic energy across three dimensions per particle. For a classical undamped harmonic oscillator governed by the potential V = \frac{1}{2} k y^2, the virial theorem predicts equal partitioning between kinetic and potential energies in the time average: \langle T \rangle = \langle V \rangle. This equality holds because the potential scales quadratically with position (n=2 in the power-law form), leading to a virial of \langle \mathbf{r} \cdot \nabla V \rangle = 2 \langle V \rangle. The theorem also applies to driven damped oscillators, providing insight into steady-state energy balance. Consider the one-dimensional equation of motion m \ddot{y} = -k y + f(t) - \gamma \dot{y}, where f(t) is the force and \gamma > 0 introduces linear . In the long-time limit, the virial theorem yields $2 \langle T \rangle = 2 \langle V \rangle - \langle y f(t) \rangle, where the quadratic potential contributes $2 \langle V \rangle, and the driving term -\langle y f(t) \rangle (with sign depending on the between y and f(t)) compensates for to maintain steady oscillations. This form highlights how external input sustains the average against frictional losses.

Gravitational Bound Structures

In self-gravitating systems, such as star clusters or gaseous spheres bound by their own , the virial theorem simplifies to a specific form due to the inverse-square nature of the gravitational . For a , configuration, twice the time-averaged equals the magnitude of the time-averaged gravitational potential energy: $2 \langle T \rangle + \langle V \rangle = 0, where T represents the total associated with the random motions of particles or the motions in a , and V is the gravitational self-potential energy, which is negative for bound systems. This relation holds for systems where external forces are negligible and the configuration is steady-state, often after relaxation on the dynamical timescale. A key consequence of this equation is the partitioning of the total energy E = \langle T \rangle + \langle V \rangle. Substituting the virial relation yields E = -\langle T \rangle, or equivalently E = \frac{1}{2} \langle V \rangle, indicating that bound self-gravitating systems possess negative total energy, with kinetic energy providing the positive contribution that balances half the potential depth. This negative energy signature distinguishes bound structures from unbound ones and underscores the stability condition: any perturbation increasing the total energy to zero or positive would lead to dispersal. The virial theorem enables estimation of the total mass M within a characteristic R, known as the virial radius, using observed . Specifically, M \propto \frac{\sigma^2 R}{[G](/page/G)}, where \sigma is the one-dimensional velocity dispersion measuring the typical random speeds, and G is the ; this proportionality arises directly from expressing \langle T \rangle \approx \frac{3}{2} M \sigma^2 for an isotropic velocity distribution and relating \langle V \rangle \approx -\frac{3}{5} \frac{[G](/page/G) M^2}{R} for a uniform sphere, though the exact factor depends on the density profile. The \langle V \rangle itself connects to the mass distribution via , \nabla^2 \Phi = 4\pi [G](/page/G) \rho, where \Phi is the and \rho the density, allowing V = \frac{1}{2} \int \rho \Phi \, dV without requiring a full solution for the theorem's application. Numerical N-body simulations, which model collisionless self-gravitating systems as ensembles of particles interacting via Newtonian gravity, routinely validate virial equilibrium. After an initial relaxation phase, these simulations demonstrate that the system satisfies $2 \langle T \rangle + \langle V \rangle \approx 0 to high precision, confirming the theorem's for long-term structural stability in such idealized, dissipationless environments. The virial theorem also provides insight into the distinction between bound and unbound states through the . For a particle at the virial radius, the speed v_{\rm esc} = \sqrt{-2 \Phi(R)} = \sqrt{10} \sigma for the uniform , derived by comparing the specific required to reach against the virial-predicted depth of the ; velocities below this threshold keep particles bound, while exceeding it allows , thus defining the energetic for system retention.

Quantum Mechanical Version

Operator Formulation and Expectation Values

In , the virial theorem is adapted to the operator formalism, where values are taken with respect to stationary states of the . For a system described by the H = T + V, with operator T = \sum_i \frac{\mathbf{p}_i^2}{2m_i} and operator V depending on positions \{\mathbf{r}_i\}, the theorem relates the value of the to that of a specific form involving the potential. This formulation arises naturally from the Heisenberg equation of motion and applies to bound states, providing a quantum analog to the classical result. The derivation begins with Ehrenfest's theorem, which governs the time evolution of expectation values: for any operator \hat{A} without explicit time dependence, \frac{d}{dt} \langle \hat{A} \rangle = \frac{i}{\hbar} \langle [H, \hat{A}] \rangle. To obtain the virial theorem, the dilation generator (or virial) operator is introduced as G = \frac{1}{2} \sum_i (\mathbf{r}_i \cdot \mathbf{p}_i + \mathbf{p}_i \cdot \mathbf{r}_i), which is Hermitian and scales coordinates and momenta appropriately. For a stationary state |\psi\rangle (an energy eigenstate of H), the time average of \frac{d}{dt} \langle G \rangle vanishes, yielding \langle [H, G] \rangle = 0. Computing the commutator explicitly gives [H, G] = [T, G] + [V, G]. The kinetic part evaluates to [T, G] = i \hbar \sum_i \frac{\mathbf{p}_i^2}{m_i} = 2 i \hbar T, while the potential part is [V, G] = -i \hbar \sum_i \mathbf{r}_i \cdot \nabla_i V, assuming V is homogeneous or can be treated via its gradient. Thus, \langle [H, G] \rangle = 2 i \hbar \langle T \rangle - i \hbar \left\langle \sum_i \mathbf{r}_i \cdot \nabla_i V \right\rangle = 0, leading to the quantum virial theorem: $2 \langle T \rangle = \left\langle \sum_i \mathbf{r}_i \cdot \nabla_i V \right\rangle, where expectation values are \langle \hat{O} \rangle = \langle \psi | \hat{O} | \psi \rangle. This holds for time-independent potentials and eigenstates of H. For bound states, the theorem can also be derived directly from the time-independent H |\psi\rangle = E |\psi\rangle. Acting with [G, H] and taking the expectation value confirms the same relation, as the structure enforces the balance between kinetic and potential contributions in stationary states. This operator approach extends the classical scalar virial by incorporating quantum commutators, ensuring the result is precise for wavefunctions rather than trajectories. A canonical example is the , where the potential is V = -\frac{e^2}{4\pi \epsilon_0 r} (in , V = -1/r). Here, \mathbf{r} \cdot \nabla V = -V (since V is homogeneous of degree -1), so \sum \mathbf{r}_i \cdot \nabla_i V = -V, yielding $2 \langle T \rangle = -\langle V \rangle. Combined with the total energy E = \langle T \rangle + \langle V \rangle, this implies \langle T \rangle = -E and \langle V \rangle = 2E, mirroring the classical virial theorem for inverse-square forces and validating the quantum treatment for systems.

Pokhozhaev Identity for Nonlinear Equations

The Pokhozhaev identity arises as an extension of the quantum virial theorem to nonlinear elliptic equations, particularly the stationary , providing a between kinetic, linear potential, and nonlinear energies for localized solutions. It is instrumental in analyzing the and of solutions in quantum field theories with self-interaction terms, where the nonlinearity captures mean-field effects. Originally derived for equations of the form −Δu + λ f(u) = 0, the identity has been generalized to include external potentials and is widely used to derive nonexistence results for supercritical nonlinearities and stability criteria for bound states. The derivation begins with the stationary equation in n dimensions: -\Delta u + V(x) u = \lambda |u|^{p-1} u, where u ∈ H^1(ℝ^n) is a real-valued with sufficient at , V is a smooth potential, λ is the nonlinearity strength, and p >1 is the power. Multiply the equation by the test function r · ∇u, where r = x = (x_1, ..., x_n), and integrate over ℝ^n. The terms vanish due to the decay of u and ∇u. For the kinetic term, yields: \int_{\mathbb{R}^n} (r \cdot \nabla u) (-\Delta u) \, dx = \frac{n-2}{2} \int_{\mathbb{R}^n} | \nabla u |^2 \, dx. For the linear potential term, \int_{\mathbb{R}^n} (r \cdot \nabla u) V u \, dx = -\frac{1}{2} \int_{\mathbb{R}^n} (r \cdot \nabla V) |u|^2 \, dx - \frac{n}{2} \int_{\mathbb{R}^n} V |u|^2 \, dx. For the nonlinear term, note that r · ∇ (|u|^{p+1}) = (p+1) |u|^{p-1} u (r · ∇ u), so \int_{\mathbb{R}^n} (r \cdot \nabla u) \lambda |u|^{p-1} u \, dx = \frac{\lambda}{p+1} \int_{\mathbb{R}^n} r \cdot \nabla (|u|^{p+1}) \, dx = -\frac{n \lambda}{p+1} \int_{\mathbb{R}^n} |u|^{p+1} \, dx, using the and decay. Combining these, the Pokhozhaev identity is \frac{n-2}{2} \int_{\mathbb{R}^n} |\nabla u|^2 \, dx + \frac{1}{2} \int_{\mathbb{R}^n} (r \cdot \nabla V) |u|^2 \, dx + \frac{n}{2} \int_{\mathbb{R}^n} V |u|^2 \, dx = -\frac{n \lambda}{p+1} \int_{\mathbb{R}^n} |u|^{p+1} \, dx. This relation equates twice the kinetic energy (scaled by (n-2)/2) to contributions from the potential and interaction, analogous to the classical virial theorem but adapted for field-theoretic nonlinearities. In applications to Bose-Einstein condensates (BECs), the governs the ground and excited states of the Gross-Pitaevskii equation, modeling dilute atomic gases with s-wave scattering interactions. For trapped BECs with V(x) = (1/2)|x|^2 (where r · ∇V = 2V), the identity simplifies, and combined with the equation multiplied by u, yields relations like 2K = 3P + 2I in for cubic nonlinearity (p=3, λ = g >0 for repulsive interactions), where K is the , P the trap energy, and I the interaction , ensuring stability below a critical atom number N_c for attractive cases (g<0). Exceeding N_c leads to collapse, as the identity shows the total energy has no positive minimum. Similarly, in nonlinear optics, the identity applies to stationary solitons in the nonlinear modeling light propagation in Kerr media, where the nonlinearity represents self-focusing (λ <0). For 3D with p=5/3 (subcritical power |u|^{2/3} u, often arising in effective models for higher-order effects or reduced dimensions), the identity reveals blow-up risks when λ drives the interaction term negative and dominant. This case highlights scaling invariance near thresholds, aiding analysis of soliton stability and pulse compression. The identity also provides stability criteria: for supercritical nonlinearities (p > (n+2)/(n-2)), it implies nonexistence of nontrivial H^1 solutions in star-shaped domains or whole , as the or decay terms force inconsistency between positive and negative interaction scaling. In subcritical cases (p < (n+2)/(n-2)), solutions exist, but the identity quantifies collapse thresholds by showing when the effective potential well cannot support bound states without infinite energy. These criteria are pivotal for predicting dynamical instability in BECs and optical solitons.

Relativistic Generalizations

Special Relativistic Form

In special relativity, the virial theorem is generalized to account for the relativistic dynamics of particles moving at speeds comparable to the speed of light, where the Newtonian form no longer holds due to the nonlinear dependence of momentum on velocity. The relativistic kinetic energy of a system of N particles is defined as T = \sum_{i=1}^N (\gamma_i - 1) m_i c^2, where m_i is the rest mass of the i-th particle, c is the speed of light, and \gamma_i = [1 - v_i^2/c^2]^{-1/2} is the Lorentz factor, with v_i = |\mathbf{v}_i| the speed of the particle. The theorem is derived by considering the relativistic equations of motion, \frac{d\mathbf{p}_i}{dt} = \mathbf{F}_i, where \mathbf{p}_i = \gamma_i m_i \mathbf{v}_i is the relativistic momentum and \mathbf{F}_i is the three-force acting on the particle. Define the virial quantity G = \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{p}_i. Its time derivative is \frac{dG}{dt} = \sum_{i=1}^N \mathbf{v}_i \cdot \mathbf{p}_i + \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{F}_i = \sum_{i=1}^N \gamma_i m_i v_i^2 + \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{F}_i. For a stable, bound system where the particles are confined within a finite volume and the configuration is time-independent on average (e.g., ergodic motion), the long-time average \left\langle \frac{dG}{dt} \right\rangle = 0. This yields the special relativistic virial theorem: \left\langle \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{F}_i \right\rangle + \left\langle \sum_{i=1}^N \gamma_i m_i v_i^2 \right\rangle = 0. The second term can be expressed in terms of the kinetic energy: \gamma_i m_i v_i^2 = m_i c^2 (\gamma_i - 1/\gamma_i), so \sum_i \gamma_i m_i v_i^2 = T + \sum_i m_i c^2 (1 - 1/\gamma_i). Thus, the theorem becomes \left\langle \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{F}_i \right\rangle + \left\langle T \right\rangle + \left\langle \sum_{i=1}^N m_i c^2 \left(1 - \frac{1}{\gamma_i}\right) \right\rangle = 0. Approximate forms arise in specific regimes; for instance, expanding for mildly relativistic speeds (v_i \ll c) recovers the classical result $2 \left\langle K \right\rangle + \left\langle \sum_i \mathbf{r}_i \cdot \mathbf{F}_i \right\rangle = 0, where K = \sum_i \frac{1}{2} m_i v_i^2. In the ultra-relativistic limit (\gamma_i \gg 1, v_i \approx c), the term $1/\gamma_i \approx 0, simplifying the theorem to \left\langle \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{F}_i \right\rangle + \left\langle T \right\rangle \approx 0. This limit is crucial for high-energy systems like relativistic plasmas or radiation-dominated configurations, where the interaction terms (e.g., from electromagnetic or gravitational forces) balance the total kinetic energy directly, contrasting with the factor of 2 in the non-relativistic case. For such systems, if the particles behave like a fluid with isotropic pressure P, the kinetic term relates to $3 \left\langle P \right\rangle V \approx \left\langle T \right\rangle, leading to $3 \left\langle P \right\rangle V + \left\langle \sum_i \mathbf{r}_i \cdot \mathbf{F}_i \right\rangle \approx 0. A fully Lorentz-invariant formulation of the scalar virial theorem emerges from four-vector considerations in field theory, using the conservation of the stress-energy tensor T^{\mu\nu} and the dilatation (scale) current. For a system invariant under scale transformations, the theorem relates the total energy E to the volume integral of the trace of the stress-energy tensor: E = -\int T^\mu_\mu \, dV. This holds in special relativity for interacting fields (e.g., electromagnetic) supporting particle-like bound states and reduces to the particle form in the appropriate limit.

Uniform Systems in Relativity

In general relativity, the virial theorem for uniform-density systems, such as spherically symmetric stars, relates the integrated pressure support to the gravitational binding energy and relativistic internal energy contributions. For a uniform sphere, this takes the approximate form $3 \int P \, dV + \langle V_{\rm grav} \rangle = -\langle T_{\rm rel} \rangle, where \int P \, dV represents the total pressure integral over the volume, \langle V_{\rm grav} \rangle is the average gravitational potential energy (negative), and \langle T_{\rm rel} \rangle denotes the relativistic kinetic and internal energy terms, including rest mass contributions. This relation arises from integrating the equations of hydrostatic equilibrium, accounting for both special relativistic effects in the matter and general relativistic curvature in the gravitational field. The Tolman-Oppenheimer-Volkoff (TOV) equation provides the foundational hydrostatic balance for such systems, given by \frac{dP}{dr} = -\frac{G m(r)}{r^2} \left( \epsilon + P \right) \left( 1 + \frac{P}{\epsilon} \right) \left( 1 + \frac{4\pi r^3 P}{m(r) c^2} \right) \left( 1 - \frac{2 G m(r)}{r c^2} \right)^{-1}, where P is pressure, \epsilon is total energy density (including rest mass), m(r) is the enclosed mass within radius r, G is the gravitational constant, and c is the speed of light. For uniform density, the TOV equation can be solved analytically, yielding pressure and energy density profiles that, when volume-integrated, directly yield the relativistic virial relation. This integration ties the pressure term $3 \int P \, dV to the relativistic enhancements in both the internal energy density and the effective gravitational potential, ensuring global equilibrium. A key application arises in white dwarfs modeled as uniform-density spheres supported by relativistic degenerate electron gas pressure against gravity. In the ultra-relativistic limit, where electron velocities approach c, the degeneracy pressure follows P = \frac{1}{3} (\epsilon_e - \rho_e c^2), with \epsilon_e the electron energy density and \rho_e the electron rest-mass density, leading to an effective equation of state P \approx K \rho^{4/3} (with K = \frac{(3\pi^2)^{1/3} \hbar c}{4} \left( \frac{1}{m_H \mu_e} \right)^{4/3}, \hbar the reduced Planck's constant, m_H the hydrogen mass, and \mu_e the mean molecular weight per electron). Substituting into the virial theorem balances this pressure integral against the gravitational energy, resulting in a unique maximum mass for stable configurations, M \approx 1.44 M_\odot (solar masses), beyond which the total energy minimum disappears, leading to dynamical instability and collapse. This limit emerges because the \gamma = 4/3 adiabatic index (from P \propto \rho^\gamma) is marginal for stability in the relativistic regime, violating the virial condition for bounded equilibria above the critical mass. In the non-relativistic limit, where typical particle speeds v \ll c, the degeneracy pressure shifts to P \propto \rho^{5/3} (\gamma = 5/3), and the relativistic terms in T_{\rm rel} and the TOV corrections vanish, recovering the classical virial theorem $3 \int P \, dV + \langle V_{\rm grav} \rangle = 0. This allows stable white dwarfs of arbitrarily low mass without a collapse threshold, highlighting how relativity introduces the upper mass bound.

Extensions and Advanced Topics

Incorporation of Electromagnetic Fields

The virial theorem extends naturally to systems incorporating electromagnetic fields by including the Lorentz force acting on charged particles and the self-energy of the fields. For a collection of charged particles in external or self-generated electric \mathbf{E} and magnetic \mathbf{B} fields, the force term in the virial expression becomes \sum_i \mathbf{r}_i \cdot q_i (\mathbf{E} + \mathbf{v}_i \times \mathbf{B}), where the time average over a steady-state configuration replaces the standard gravitational or interaction potentials. This contribution accounts for both electrostatic interactions via the electric field and magnetic effects via the velocity-dependent term, leading to a modified relation $2 \langle T \rangle + \langle V_\text{grav} + V_\text{em} \rangle = 0 for bound systems, where V_\text{em} denotes the electromagnetic potential energy among particles. To incorporate the energy of the electromagnetic fields themselves in a self-consistent manner, the derivation employs the Maxwell stress tensor, which describes the momentum flux and stress due to the fields. The scalar virial theorem arises from integrating the equation of motion multiplied by position, yielding volume and surface terms from the divergence of the stress tensor \overleftrightarrow{T}^\text{em}_{jk} = \frac{1}{4\pi} \left( E_j E_k + B_j B_k - \frac{1}{2} \delta_{jk} (E^2 + B^2) \right) (in cgs units). For a confined volume where surface contributions vanish, this adds a field energy term to the virial balance: $2 \langle T \rangle + \langle V_\text{grav} + V_\text{em} \rangle + \int \frac{E^2 + B^2}{8\pi} \, dV = 0. Both the electric and magnetic field energy terms \int E^2 / 8\pi \, dV and \int B^2 / 8\pi \, dV act as supportive contributions opposing binding forces through field stress, with the interaction nature (repulsive or attractive) reflected in V_\text{em}. This form can also be derived via integration of the Poynting theorem, which conserves electromagnetic energy and links field dynamics to mechanical work. In plasma systems, where collective behavior dominates and electric fields are often screened, the electromagnetic contribution simplifies to focus on magnetic effects. The Lorentz force \mathbf{J} \times \mathbf{B}/c (in cgs) leads to a magnetic tension term in the virial integral, \int \mathbf{r} \cdot (\mathbf{J} \times \mathbf{B}) \, dV, which, through vector identity and integration by parts, evaluates to -\int B^2 / 8\pi \, dV for closed configurations. Thus, the equilibrium virial theorem becomes $2 \langle T \rangle + \langle V_\text{grav} \rangle + 3 (\gamma - 1) \langle U \rangle + \int \frac{B^2}{8\pi} \, dV = 0, where U is the internal thermal energy and \gamma is the adiabatic index; the magnetic energy provides stability against collapse. A representative application occurs in solar coronal loops, where arched magnetic structures confine hot plasma against gravity and thermal pressure. Here, the virial theorem implies that the magnetic energy \int B^2 / 8\pi \, dV balances the gravitational potential energy, with estimates showing magnetic energies on the order of $10^{28}–$10^{30} erg for typical loops, sufficient to maintain equilibrium over scales of $10^4–$10^5 km. This balance highlights how magnetic tension supports the loop geometry, preventing sagging under solar gravity.

Non-Equilibrium and Driven Systems

In non-equilibrium systems, the virial theorem generalizes beyond the steady-state condition where the time average of the virial's rate of change vanishes. For a classical system of N particles, the virial is defined as G = \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{p}_i, and its exact time derivative satisfies \frac{dG}{dt} = 2T + \sum_{i=1}^N \mathbf{r}_i \cdot \mathbf{F}_i + \oint \mathbf{r} \cdot (\mathbf{v} \cdot d\mathbf{A}), where T is the total kinetic energy, \mathbf{F}_i are the forces on particle i, and the surface integral accounts for boundary fluxes in continuous systems. Unlike equilibrium cases, the ensemble or time average \left\langle \frac{dG}{dt} \right\rangle \neq 0, reflecting net energy injection or loss, with \left\langle \frac{dG}{dt} \right\rangle = 2 \langle T \rangle + \left\langle \sum \mathbf{r}_i \cdot \mathbf{F}_i \right\rangle + boundary terms. This form captures transient dynamics or steady non-equilibrium states driven by external influences. In driven systems, external non-conservative forces introduce power input that sustains the system against dissipation. The average power injected by external forces is P_{\rm drive} = -\left\langle \sum_{i=1}^N \mathbf{F}_{\rm ext,i} \cdot \mathbf{v}_i \right\rangle, which in a steady non-equilibrium state balances dissipative losses to maintain constant energy flux. For instance, in Langevin dynamics with damping coefficient \gamma_i for particle i, the mesoscopic virial equation relates local kinetic energy to forces and heat flow: \left\langle \frac{p_i^2}{m_i} \right\rangle = \left\langle \left( \partial_{q_i} H - f_i \right) q_i \right\rangle, where H is the Hamiltonian, f_i includes driving, and dissipation appears via \dot{Q}_i = \gamma_i \left( \left\langle \frac{p_i^2}{m_i} \right\rangle - k_B T_i \right), linking to non-zero entropy production. Summing over particles yields a macroscopic relation akin to a non-equilibrium equation of state, \frac{1}{\gamma} \dot{Q} + \frac{N k_B T}{d} = \frac{\bar{P} V}{d} + C_{\rm int}, where d is dimensionality, \bar{P} is pressure, and C_{\rm int} captures internal constraints. A representative example arises in viscous fluids, where internal friction modifies the force term. The tensor virial theorem extends to include the viscous stress tensor \sigma^v_{jk}, contributing a dissipation term \int_V \sigma^v_{jk} \frac{\partial v_k}{\partial x_j} dV to the virial balance, representing frictional power loss \gamma \int v^2 dV in simplified models. In steady state for such driven dissipative fluids, the relation becomes \langle 2T + W \rangle + \langle D \rangle = P_{\rm drive}, where W is the potential energy (virial \sum \mathbf{r} \cdot \nabla W), and D is the average dissipation. This framework applies to sheared or flowing fluids under external forcing. For damped systems under continuous driving, such as a classical oscillator with friction \mu \dot{x} and external force F(t), the steady-state virial theorem adjusts to \langle m \dot{x}^2 \rangle = \langle k x^2 \rangle + \mu \langle x \dot{x} \rangle - \langle x F(t) \rangle, where the cross-term \mu \langle x \dot{x} \rangle quantifies dissipation (vanishing in equilibrium but non-zero here), balanced by driving input \langle x F(t) \rangle. In stochastic contexts, recent molecular dynamics simulations (post-2020) employing for non-equilibrium states adapt this via mesoscopic virial equations to compute pressures and stresses, incorporating random forces to model driven fluctuations in complex fluids.

Astrophysical and Cosmological Contexts

Stellar Interiors and Dynamics

In stellar interiors, the virial theorem provides a fundamental relation for systems in hydrostatic equilibrium, balancing thermal kinetic energy against gravitational potential energy. For a star modeled as an ideal gas, the theorem takes the form $2 \langle T_{\text{thermal}} \rangle + 3 \int P \, dV + \langle V_{\text{grav}} \rangle = 0, where \langle T_{\text{thermal}} \rangle is the volume-averaged thermal kinetic energy, \int P \, dV is the pressure integral over the stellar volume, and \langle V_{\text{grav}} \rangle is the gravitational potential energy (typically negative). This equation arises from integrating the hydrostatic equilibrium condition \frac{dP}{dr} = -\rho \frac{G m(r)}{r^2} and applying the virial theorem, linking the star's global energy content to its structural stability. For polytropic models of stellar interiors, which approximate the equation of state as P = K \rho^{1 + 1/n} with polytropic index n, the virial theorem yields scaling relations that connect the polytropic constant K to the star's mass M. In the case of an ideal gas with adiabatic index \gamma = 5/3 (corresponding to n = 3/2), dimensional analysis from hydrostatic balance and the virial relation gives K \propto M^{2/3}, assuming homologous structures with roughly constant mean density across masses; this scaling highlights how higher-mass stars require stiffer equations of state to maintain equilibrium. Such relations are crucial for modeling convective cores in main-sequence stars, where the polytropic approximation captures the transport of energy without detailed radiative transfer. During stellar evolution, the virial theorem governs the energy release in contracting phases, such as protostellar collapse or post-main-sequence adjustments. As a star contracts quasi-statically, the decrease in gravitational potential energy \Delta V_{\text{grav}} (negative change) leads to an increase in thermal energy by half that amount, with the other half radiated away to maintain equilibrium; for a bound configuration, this implies that half the magnitude of the change in gravitational potential energy is added as thermal energy, per the virial relation E = \langle T_{\text{thermal}} \rangle + \langle V_{\text{grav}} \rangle = \langle V_{\text{grav}} \rangle / 2. This process powers the , providing luminosity on timescales of $10^7 years for solar-mass stars before nuclear ignition dominates. The virial theorem's form varies between radiative and convective stars due to differences in energy transport and internal energy composition. In radiative stars, where energy is carried by photon diffusion, the theorem includes contributions from both gas and radiation pressure: the full relation becomes $3 \int P_{\text{gas}} \, dV + U_{\text{rad}} + \langle V_{\text{grav}} \rangle = 0, with radiation internal energy U_{\text{rad}} = \frac{4\sigma}{3c} \int T^4 \, dV, reflecting the stiffer support from radiation in massive stars; this modifies the effective polytropic index toward n=3. Convective stars, in contrast, follow near-adiabatic gradients with \nabla = \nabla_{\text{ad}}, closely approximating n=3/2 polytropes and the ideal gas scaling above, as mixing enforces uniform entropy. Nuclear energy generation acts as a small perturbation in both cases, contributing less than 1% to the total energy budget compared to gravitational and thermal terms, but it sustains long-term equilibrium by offsetting radiative losses. Recent observations of white dwarf cooling have incorporated neutrino effects into virial-based models, particularly for degenerate interiors where thermal energy loss influences slight structural adjustments. Post-2010 surveys, such as those using the hot white dwarf luminosity function, constrain neutrino emission rates (e.g., from plasmon decay), which accelerate early cooling by up to 20% in white dwarfs of mass \sim 0.6 M_\odot; in the virial framework for degenerate support, $3 \int P_{\text{deg}} \, dV + \langle V_{\text{grav}} \rangle \approx 0, neutrino losses perturb the thermal component minimally but refine age estimates for these gravitational bound structures.

Galaxies, Clusters, and Virial Mass Estimation

The virial theorem provides a fundamental tool for estimating the total mass of galaxies and galaxy clusters by relating their observed velocity dispersions to the gravitational potential. For self-gravitating systems in virial equilibrium with isotropic velocity dispersion \sigma, the virial mass M_{\text{vir}} within a characteristic radius R is given by M_{\text{vir}} = \frac{5 \sigma^2 R}{G}, where G is the gravitational constant; this estimator assumes that the kinetic energy is \frac{3}{2} M \sigma^2 and balances half the magnitude of the potential energy for a uniform sphere (a common approximation uses 3 instead for other density profiles). In galaxies, this approach is applied to stellar or gas kinematics, but observed rotation curves often deviate from predictions based on visible matter alone, remaining flat at large radii rather than declining as expected under Keplerian motion. These discrepancies, first systematically observed in spiral galaxies like Andromeda and M31, indicate that additional unseen mass—attributed to dark matter—is required to maintain the high orbital velocities, with dark matter halos providing the necessary gravitational binding. Vera Rubin's spectroscopic measurements in the 1970s confirmed this for dozens of spirals, showing velocities around 200–250 km/s extending to 20 kpc or more, implying dark matter contributions exceeding visible mass by factors of 5–10. In galaxy clusters, the virial theorem is similarly employed using the motions of member galaxies or the hot intracluster medium (ICM). The ICM, observed via X-ray emission, reaches temperatures T \sim 10^7–10^8 K, corresponding to thermal velocities that trace the cluster potential. For an ideal gas, the ICM temperature relates to the galaxy velocity dispersion \sigma through kT = \mu m_p \sigma^2, where k is Boltzmann's constant, m_p is the proton mass, and \mu \approx 0.6 is the mean molecular weight for ionized gas; this equipartition allows inference of the total cluster mass by combining X-ray data with optical velocity measurements, yielding M_{\text{vir}} \sim 10^{14}–10^{15} M_\odot within the virial radius r_{\text{vir}}, dominated by dark matter. Pioneering applications, such as to the , revealed mass-to-light ratios M/L \sim 100–300 in solar units, far exceeding stellar values and underscoring the dark matter's role in cluster dynamics. For systems with anisotropic velocity distributions, where radial and tangential dispersions differ, the simple isotropic virial estimator can bias masses by up to 20–50%; instead, the is used to model the density and velocity profiles. The steady-state in spherical coordinates relates the radial velocity dispersion \sigma_r to the gravitational potential \Phi via \frac{d(\nu \sigma_r^2)}{dR} + 2 \beta \frac{\nu \sigma_r^2}{R} = -\nu \frac{d\Phi}{dR}, where \nu(R) is the tracer density and \beta = 1 - (\sigma_\theta^2 + \sigma_\phi^2)/(2 \sigma_r^2) parameterizes anisotropy; solving this with observed profiles yields the mass profile M(R) = -R^2 \nu^{-1} d(\nu \sigma_r^2)/dR + \dots, enabling more accurate M_{\text{vir}} for clusters like . This method has been validated against N-body simulations, reducing systematic errors in anisotropy assumptions to below 10%. Observed velocities are line-of-sight projections, requiring deprojection to apply the accurately. The projected accounts for this by integrating the 3D quantities along the line of sight, yielding an estimator M_{\text{vir}} = \frac{3\pi}{G} \frac{\sum_i \sigma_{\text{los},i}^2}{\sum_i 1/R_i}, where \sigma_{\text{los}} is the line-of-sight velocity dispersion and R_i are projected separations (using the harmonic mean radius); corrections for surface pressure terms and incompleteness are essential for incomplete sampling within r_{\text{vir}}. This projected form has been applied to samples of hundreds of clusters, confirming consistency between optical and X-ray masses within 15–20%. Recent James Webb Space Telescope (JWST) observations in the 2020s have refined virial radius measurements for galaxies, particularly in the early universe at z > 6, by resolving extended structures and kinematics inaccessible to prior telescopes. For instance, JWST/NIRCam imaging of low-mass galaxies like those in the Field of Frontier Fields has detected accretion shelves—overdense gas features—extending to the virial radius r_{\text{vir}} \sim 10–20 kpc, allowing precise calibration of M_{\text{vir}} and revealing more compact halos than previously modeled, with implications for reionization-era dynamics. These data, combined with , suggest virial radii 20–30% smaller than local analogs, refining mass estimates and highlighting rapid evolution in halo binding.

References

  1. [1]
    [PDF] Physics 5153 Classical Mechanics The Virial Theorem and The ...
    The first, the Virial Theorem, applies to systems that are bounded. It allows us to determine statistical properties of systems. The other is the Poisson ...
  2. [2]
    [PDF] Virial Theorem
    Clausius in 1870.[1] chmy564-17. Mon 23jan17. Page 2. Rudolf Claussius (Wikipedia) ... In 1870 he introduced the virial theorem which applied to heat.
  3. [3]
    The virial theorem in stellar astrophysics - NASA ADS
    Some applications of the virial theorem to problems involving the pulsational stability of white dwarfs, effects of rotation and magnetic fields on the white- ...
  4. [4]
    The Virial Theorem and Covalent Bonding - ACS Publications
    Aug 30, 2018 · Within this framework, the virial theorem guarantees the correct atomic orbital scaling in a molecule and the correct relation between the ...
  5. [5]
    Virial Theorem and Its Applications in Instability of Two-Phase Water ...
    In mathematics, virial theorem serves as an important method in the blowup problem of PDEs. In the 1970s, Zakharov–Sobolev–Synakh [66] and Glassey [20] proved ...
  6. [6]
    [PDF] 122 Prof. R. Clausius on a Mechanical Theorem 1 be the extent of ...
    On a Mechanical Theorem applicable to Heat. By R. CLAUSIUS *. IN. N a treatise which appeared in 1862, on the mechanical theory of heatt, I advanced a ...
  7. [7]
    The Vital Virial of Rudolph Clausius: From Stat Mech to Quantum Mech
    Jul 24, 2024 · The Virial Theorem makes a simple statement about the balance between kinetic energy and potential energy (in a conservative mechanical system).Missing: definition | Show results with:definition
  8. [8]
    James Hopwood Jeans. 1877-1946 - jstor
    He also delivered the Rouse. Ball Lecture at Cambridge in 1925 on 'Atomicity and quanta'. He received the Gold Medal of the Royal Astronomical Society in 1922, ...Missing: historical key
  9. [9]
    Timeline of thermodynamics - Wikipedia
    1870 – Clausius proves the scalar virial theorem; 1872 – Ludwig Boltzmann states the Boltzmann equation for the temporal development of distribution ...Missing: original | Show results with:original
  10. [10]
    Application of the virial theorem to some astronomical problems
    Applications of the virial theorem to astronomical problems were first made by H. Poincaré' and A. S. Eddington.2 Most recently P. Ledoux has successfully ...Missing: history review
  11. [11]
    [PDF] I. Development of the Virial Theorem - Astrophysics Data System
    This is effectively a statement of system boundary conditions uniquely determining the system's past and future. It is the essence of the Louisville theorem of ...
  12. [12]
    Shafranov's virial theorem and magnetic plasma confinement - arXiv
    Sep 21, 2000 · Shafranov's virial theorem implies that nontrivial magnetohydrodynamical equilibrium configurations must be supported by externally supplied ...
  13. [13]
    [PDF] SIMULATIONS OF STRUCTURE FORMATION IN THE UNIVERSE
    The first astronomical N-body computations using digital computers were made in the early 1960s (von Hoerner 1960, 1963, Aarseth 1963). These early simula ...
  14. [14]
    Potential analogue - Neo-classical physics
    70 (1870), 1314-. 1319. On a quantity that is analogous to potential and a theorem that relates to it. By R. CLAUSIUS. Translated by D. H. Delphenich. ______.
  15. [15]
    2.11: Virial Theorem - Physics LibreTexts
    Jan 22, 2021 · The Virial theorem deals with average properties and has applications to statistical mechanics. Consider an ideal gas. According to the ...Missing: early | Show results with:early
  16. [16]
    5.3: The Virial Theorem - Physics LibreTexts
    May 11, 2024 · For a potential energy homogeneous in the coordinates, of degree. , say, and spatially bounded motion, there is a simple relation between the ...Missing: n | Show results with:n
  17. [17]
    11.8: Inverse-square, two-body, central force - Physics LibreTexts
    Jun 28, 2021 · Also note that, as predicted by the Virial Theorem, the kinetic energy always is half the potential energy for the inverse square law force.
  18. [18]
    [PDF] 1 The virial theorem
    ˙pi = Fi. (1). Write. G = ∑i pi · ri. (2). Then. ˙G = ∑i. ˙pi · ri + ∑i pi · vi = ∑i. Fi · ri + 2T. (3). Let us compute the time average of each quantity over ...
  19. [19]
    [PDF] Lecture 10 - A Document With An Image - University of Washington
    The scalar virial theorem is useful for estimating global av- erage properties, such as total mass, escape velocity and relaxation time, while the tensor virial ...
  20. [20]
    Temperature in the classical microcanonical ensemble
    Mar 1, 2001 · Special cases of the new formula include a number of familiar results including the virial theorem.Missing: paper | Show results with:paper<|control11|><|separator|>
  21. [21]
    Using the virial theorem | American Journal of Physics | AIP Publishing
    Dec 1, 1986 · Elementary applications of the virial theorem are presented. The driven harmonic oscillator and the anharmonic oscillator are considered, along ...
  22. [22]
    [PDF] Virial Theorem Simplest
    In general, the virial theorem can be applied to any gravitating system after one dynamical timescale has elapsed.
  23. [23]
    Virial Theorem - an overview | ScienceDirect Topics
    The virial theorem is defined as a relation that connects the time averages of kinetic energy and potential energy for a system of particles, ...
  24. [24]
    [PDF] Systematics in virial mass estimators for pressure-supported systems
    May 1, 2018 · For the case of self-gravitating systems, Merritt (1987) de- scribes how to infer M(< r0) from measurements of 〈σ2 los〉 under the ...
  25. [25]
  26. [26]
    ASTR 5630/40 8 : THEORY II : STELLAR DYNAMICS (Whittle)
    (5) The Virial Theorem. This fundamental result describes how the total energy (E) of a self-gravitating system is shared between kinetic energy (K) and ...
  27. [27]
    [PDF] arXiv:0901.1076v6 [quant-ph] 12 Dec 2012
    Dec 12, 2012 · Quantum mechanical virial theorem is derived from the time derivative of the expectation value of operator G, i.e. from the expectation value of ...
  28. [28]
    [PDF] Quantum mechanical approaches to the virial
    Jun 30, 2015 · the fact the quantum virial theorem is in direct correspondence with the classical virial theorem ... in 1870 in ”On a Mechanical Theorem ...
  29. [29]
    An elementary derivation of the quantum virial theorem from ...
    The virial theorem is an important relation used for computing certain averages in statistical, classical and quantum mechanics [1]. It is also a powerful ...<|separator|>
  30. [30]
    A generalized Pohozaev identity and its applications - Project Euclid
    Introduction. In this paper we establish a generalized Pohozaev identity and its variant for the radial solutions of the following quasilinear elliptic equation ...Missing: derivation | Show results with:derivation
  31. [31]
    Some applications of the Pohozaev identity - AIP Publishing
    Apr 23, 2009 · The Pohozaev-type identity is a basic tool to show the nonexistence of nontrivial solution (see, e.g., Cazenave,2 Willem,13 and Kuzin and ...
  32. [32]
    [PDF] Local Uniqueness of Ground States for Rotating Bose-Einstein ...
    Dec 27, 2021 · Keywords: Bose-Einstein condensates; rotational velocity; local uniqueness; Pohozaev identity. 1 Introduction. A Bose-Einstein condensate (BEC) ...
  33. [33]
    The Pohozaev-type inequalities and their applications for a kind of ...
    Dec 9, 2020 · In this paper, we first derive a new kind of Pohozaev-type inequalities for p-Laplacian equations in a more general class of non-star-shaped ...
  34. [34]
    Pohozaev identities and their applications to nonlinear elliptic ...
    Aug 10, 2025 · For nonlinear elliptic equations, Pohozaev identity that arises from Pohozaev 27 plays a very important role in the study of the existence ...Missing: derivation | Show results with:derivation
  35. [35]
    [PDF] THE RELATIVISTIC VIRIAL THEOREM Wolfgang LUCHA - INIS-IAEA
    In this note we derive a relation which may be regarded as the relativistic generalization of the virial theorem well known in nonrelativistic quantum mechanics ...
  36. [36]
    [PDF] The virial theorem and the kinetic energy of particles of a ... - arXiv
    The virial theorem is considered for a system of randomly moving particles that are tightly bound to each other by the gravitational and electromagnetic ...<|separator|>
  37. [37]
    [PDF] II Contemporary Aspects of the Virial Theorem
    May 2, 2016 · The tensor representation of the virial theorem is an attempt to restore some of the information lost in reducing the full vector equations ...Missing: plasma refinements
  38. [38]
    [PDF] The relativistic virial theorem and scale invariance - arXiv
    Jun 4, 2013 · In this approximation, both the vector and scalar potentials are homogeneous functions of degree n = −1, so the virial theorem (5), expanded to ...
  39. [39]
    [PDF] arXiv:1206.0618v1 [astro-ph.CO] 4 Jun 2012
    Jun 4, 2012 · In the present paper, a fully generally-relativistic virial theorem based on the Tolman-Oppenheimer-Volkoff (TOV) so- lution for homogeneous ...
  40. [40]
    On Massive Neutron Cores | Phys. Rev.
    In this paper we study the gravitational equilibrium of masses of neutrons, using the equation of state for a cold Fermi gas, and general relativity.
  41. [41]
    Virial theorem for a system of charged particles - IOPscience
    A refined formulation of the virial theorem for a system of charged particles is given. The renormalization of the total electromagnetic field energy is taken ...<|control11|><|separator|>
  42. [42]
    [PDF] The energy-momentum tensor of the electromagnetic field ... - arXiv
    Dec 3, 2023 · The virial theorem for a system of charged particles uses traces of EMT [1] of the components of a physical system and describing the stationary ...
  43. [43]
    [PDF] The Virial Theorem, MHD Equilibria, and Force-Free Fields
    The Virial Theorem helps understand equilibrium properties of a system in terms of energies, and provides broad insight into relaxed systems.Missing: electromagnetic | Show results with:electromagnetic
  44. [44]
    (PDF) Study of Solar Magnetic and Gravitational Energies Through ...
    Feb 9, 2015 · Virial theorem is important for understanding stellar structures. It produces an interesting connection between the magnetic energy and the ...
  45. [45]
    Mesoscopic virial equation for nonequilibrium statistical mechanics
    Dec 5, 2015 · Abstract page for arXiv paper 1512.01687: Mesoscopic virial equation for nonequilibrium statistical mechanics. ... View PDF · TeX Source · view ...
  46. [46]
    Not found - ADS
    **Summary:**
  47. [47]
    [2302.12008] Quantum dissipation and the virial theorem - arXiv
    Feb 23, 2023 · In this note, we study the celebrated virial theorem for dissipative systems, both classical and quantum. The classical formulation is discussed ...Missing: driven | Show results with:driven
  48. [48]
    Dynamical Nonequilibrium Molecular Dynamics Simulations Identify ...
    Jun 7, 2023 · We apply dynamical nonequilibrium molecular dynamics (D-NEMD) simulations to evaluate structural and dynamical responses of M pro to the presence and absence ...Introduction · Results and Discussion · Methods · Supporting Information
  49. [49]
    [PDF] Equations of stellar structure
    SIMPLE ENERGETICS OF STARS: VIRIAL THEOREM, etc. Gravitational energy and hydrostatic equilibrium. We shall consider stars in a hydrostatic equilibrium, but not ...
  50. [50]
    [PDF] The Virial Theorem 1
    The virial theorem relates total gravitational and thermal energies of stars in hydrostatic equilibrium, linking internal and gravitational potential energies.
  51. [51]
    [PDF] Lecture 7a Polytropes
    To the extent that any star is supported by ideal gas pressure, the Virial theorem holds and the total energy of the star. E tot. = U + Ω= Ω. 2. =−. 3. 10− 2n.
  52. [52]
    [PDF] 9. Planetary Models - CalTech GPS
    The Virial theorem also shows up in other areas of physics (e.g the expected velocity dispersion of a self-gravitating star cluster).
  53. [53]
    [PDF] 10. Timescales in Stellar Interiors
    This allows us to apply the. Virial Theorem to a broad diversity of systems in motion, from atoms swirling within a star to stars orbiting in a globular cluster ...<|control11|><|separator|>
  54. [54]
    [PDF] The Virial Theorem - PHYS 633: Introduction to Stellar Astrophysics
    The Virial Theorem. Consider the equation of hydrostatic equilibrium,. dP dm. = −. Gm. 4πr4 . (1). Multiply both sides by 4πr3 and integrate with respect to the ...
  55. [55]
    [PDF] 5 – Stellar Structure I
    These two forces play the principal role in determining stellar structure – ... hydrostatic equilibrium and the Virial theorem. We have derived a minimum.
  56. [56]
    CONSTRAINING NEUTRINO COOLING USING THE HOT WHITE ...
    yr of the white dwarf's existence, the rate of cooling is dominated by neutrino emission. Thus, if we can amass a large enough sample of white dwarfs with ...INTRODUCTION · FITTING ATMOSPHERIC... · MODELS · RESULTSMissing: post- | Show results with:post-
  57. [57]
    [PDF] arXiv:1007.2659v1 [astro-ph.SR] 15 Jul 2010
    Jul 15, 2010 · In young white dwarfs, neutrinos increase the cooling rate and produce a temperature inversion. Neutrinos result from pure leptonic ...
  58. [58]
    Virial theorem and the dynamics of clusters of galaxies in the brane ...
    Aug 16, 2007 · ... The virial mass can also be expressed in terms of the characteristic velocity dispersion σ 1 as [28]. M V = 3 G σ 1 2 R V . (91).
  59. [59]
    The Rotation of Spiral Galaxies - Science
    The gravitational force of this dark matter is presumed to be responsible for the high rotational velocities of stars and gas in the disks of spiral galaxies.<|control11|><|separator|>
  60. [60]
    June 1980: Vera Rubin Publishes Paper Hinting at Dark Matter
    May 11, 2023 · An astronomer named Vera Rubin had accumulated a convincing body of evidence that something unseen in the universe was causing galaxies to behave in unexpected ...
  61. [61]
    X-ray emission from clusters and groups of galaxies - PMC
    The x-ray temperature agrees extremely well (on average) with the optical velocity dispersion (β = μmpσ2/kT ≈ 1) (14); however, there is a real variance in the ...
  62. [62]
    X-ray Emission from Clusters of Galaxies - C.L. Sarazin
    The application of the virial theorem to clusters and the discovery that most of their mass was nonluminous was first made by Zwicky (1933) and Smith (1936).
  63. [63]
    Dynamical Equilibrium of Galaxy Clusters - IOP Science
    Consequently, the virial mass calculated from the blue galaxies is 3.5 ± 1.3 times larger than from the red galaxies. However, applying the Jeans equation of ...
  64. [64]
    Modified virial formulae and the theory of mass estimators
    Abstract. We show how to estimate the enclosed mass from the observed motions of an ensemble of test particles. Traditionally, this problem has been attack.Missing: RG | Show results with:RG
  65. [65]
    [PDF] Dynamical mass measurements of clusters of galaxies
    Affects estimates of the velocity distribution of galaxies, like the line-of-sight velocity dispersion (virial theorem, Jeans equation). Solutions ...
  66. [66]
    OPTICAL MASS ESTIMATES OF GALAXY CLUSTERS - IOP Science
    The virial theorem and the Jeans equation should give consistent mass estimates. ... constant velocity dispersion pro–le and isotropic orbits. On these ...
  67. [67]
    Detection of Accretion Shelves Out to the Virial Radius of a Low ...
    Jun 20, 2024 · Detection of Accretion Shelves Out to the Virial Radius of a Low-mass Galaxy with JWST. Charlie Conroy, Benjamin D. Johnson, ...
  68. [68]
    [2310.13048] Detection of Accretion Shelves Out to the Virial Radius ...
    Oct 19, 2023 · Abstract page for arXiv paper 2310.13048: Detection of Accretion Shelves Out to the Virial Radius of a Low-Mass Galaxy with JWST.