Fact-checked by Grok 2 weeks ago

Perfect fluid

A perfect fluid, also known as an ideal fluid, is a theoretical construct in physics used to model fluids in relativistic hydrodynamics and , characterized by the absence of , , and thermal conductivity, with its behavior fully determined by an isotropic pressure and in the fluid's . This idealization assumes isentropic flow, where is conserved along streamlines, and no dissipative effects such as or occur, making it a simplified yet powerful for many astrophysical and cosmological scenarios. The mathematical description of a perfect fluid is encapsulated in its energy-momentum tensor, given by
T^{\mu\nu} = (\epsilon + P) u^\mu u^\nu + P g^{\mu\nu},
where \epsilon is the proper , P is the isotropic , u^\mu is the normalized such that u^\mu u_\mu = -1 (in units where c=1), and g^{\mu\nu} is the . In the fluid's local , this tensor simplifies to a diagonal form with \epsilon along the time component and P along the spatial components, reflecting the lack of momentum flux or anisotropic stresses. The dynamics are governed by the conservation laws \nabla_\mu T^{\mu\nu} = 0 and, for a single conserved particle number, \nabla_\mu (n u^\mu) = 0, where n is the proper .
Perfect fluids are typically supplemented with an equation of state P = P(\epsilon), which relates to and dictates the fluid's thermodynamic behavior; common examples include (P = 0), (P = \epsilon / 3), and stiff matter (P = \epsilon). For barotropic fluids, the equation of state is a function of density alone, enabling analytical solutions, while polytropic forms P = K \epsilon^\gamma (with constant K and adiabatic index \gamma) model more complex scenarios like stellar interiors. These relations ensure thermodynamic consistency, often assuming local and constant per particle. In applications, perfect fluids serve as foundational models in for describing matter distributions in —such as the Friedmann-Lemaître-Robertson-Walker universe filled with matter, radiation, or —and in for compact objects like neutron stars or the interiors of black holes. They also appear in special relativistic contexts, like high-energy particle collisions, and provide the zeroth-order approximation in hydrodynamic expansions that include for more realistic fluids. Despite their simplifications, perfect fluid solutions have yielded key insights, such as the Tolman-Oppenheimer-Volkoff equation for in stars.

Fundamental Concepts

Definition

A perfect fluid, also known as an ideal fluid, represents an idealized model in where dissipative effects are absent, originating from the foundational work in 18th-century hydrodynamics by Leonhard Euler, who formulated equations for in 1757. This concept was further developed in the through contributions to non-viscous fluid theories, and it gained prominence in the early 20th century with the advent of , where it was formalized around 1915–1916 by and to describe matter distributions in curved , such as stellar interiors. Precisely, a perfect fluid is defined as a fluid exhibiting zero (η = 0) and zero (κ = 0), resulting in a stress tensor that is isotropic and solely dependent on scalar thermodynamic variables, such as the fluid's and pressure. In this model, the absence of stresses and thermal gradients ensures that and transport occur without frictional losses or diffusive . Unlike real fluids, which exhibit dissipative phenomena such as viscosity-induced and conduction that lead to and , perfect fluids simplify analysis by neglecting these effects, allowing for exact solvability in many theoretical scenarios. In the of the , where the bulk velocity vanishes, the state is fully characterized by the mass- density ρ and the isotropic p, encapsulating all necessary thermodynamic information without additional complexities.

Physical Properties

A perfect fluid is characterized by isotropic , meaning that in its local , the p exerts equal force in all directions, resulting in the absence of stresses or anisotropic components in the . This property arises from the assumption of no internal or , ensuring that transport occurs solely through gradients rather than diffusive processes. The thermodynamic state of a perfect fluid is described by key variables that depend on the context: in non-relativistic settings, the proper mass \rho_m, T, and specific s; in relativistic cases, the \rho, T, and density s. These variables are interrelated through an , which dictates how responds to changes in density or energy, and they evolve without dissipative effects due to the fluid's nature. The flow is inviscid, implying no via , which permits solutions featuring irrotational or potential flows where velocity fields derive from a . Perfect fluids are often modeled under adiabatic conditions, where is conserved along flow lines (ds = 0), reflecting the lack of conduction or exchange with the surroundings, though non-conducting addition can be considered in some formulations. This isentropic behavior simplifies the dynamics, as the fluid's evolution follows reversible processes. Regarding , perfect fluids can exhibit either incompressible behavior, with constant , or compressible responses, where varies according to the equation of state, enabling phenomena like with speed c_s = \sqrt{dp/d\rho}.

Classical Fluid Dynamics

Governing Equations

In the non-relativistic regime, the dynamics of a perfect fluid, characterized by negligible and thermal conductivity, are governed by the and Euler's equation, which express and , respectively. The is given by \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{v}) = 0, where \rho is the fluid density and \mathbf{v} is the velocity field; this equation ensures that the rate of change of mass in a equals the net flux through its surface. Euler's equation for takes the form \rho \left( \frac{\partial \mathbf{v}}{\partial t} + (\mathbf{v} \cdot \nabla) \mathbf{v} \right) = -\nabla p, where p is the pressure; this arises from applying Newton's second law to a fluid element, balancing inertial forces with the pressure gradient while neglecting viscous stresses as in the Navier-Stokes equations. These equations can be derived using variational principles from a formulation for ideal s. The is constructed as S = \int dt \, d^3x \left[ \frac{1}{2} \rho |\mathbf{v}|^2 + \phi \nabla \cdot \mathbf{u} - \beta_i \left( \frac{\partial \alpha_i}{\partial t} + \mathbf{u} \cdot \nabla \alpha_i \right) \right], where \phi is a , \mathbf{u} is the , \alpha_i are embedding coordinates for the particles, and \beta_i are Lagrange multipliers enforcing the \det(\partial \alpha_i / \partial x_j) = 1 for incompressibility; varying this with respect to the fields yields the and Euler equations. For steady flows, this variational approach leads to Bernoulli's along streamlines: \int \frac{dp}{\rho} + \frac{1}{2} v^2 + \Phi = \text{constant}, where \Phi is the gravitational potential; the pressure p emerges as p = -\frac{\partial \phi}{\partial t} - \beta_j \frac{\partial \alpha_j}{\partial t} - \frac{1}{2} \rho |\mathbf{v}|^2 + \text{constant}. Bernoulli's principle, a direct consequence of this integral for steady, incompressible flows, describes the trade-off between pressure and velocity: as the fluid speed v increases, the pressure p decreases to conserve energy, assuming constant density and no external work. This relation holds along streamlines and is fundamental for analyzing irrotational flows without shocks. A key characteristic speed in perfect fluid perturbations is the adiabatic sound speed c_s = \sqrt{ \left( \frac{\partial p}{\partial \rho} \right)_{\text{adiabatic}} }, which quantifies the of small disturbances under adiabatic conditions, derived from linearizing the Euler and equations around equilibrium while assuming an linking p and \rho. This speed sets the scale for effects in the flow.

Applications

In classical , the perfect fluid model finds significant application in , where it describes the equilibrium state of fluids under gravitational influence. The pressure variation with depth in a static is governed by the equation \frac{dP}{dz} = -\rho g, where P is , z is the vertical coordinate, \rho is the constant , and g is . This relation enables calculations of pressure distributions in oceans and planetary atmospheres, assuming incompressibility and the absence of motion, as seen in models for depth profiles or atmospheric layering. Potential flow theory extends the perfect fluid approximation to irrotational, incompressible flows, characterized by \nabla \times \mathbf{v} = 0, where \mathbf{v} is the . By introducing a \phi such that \mathbf{v} = \nabla \phi, the governing equation simplifies to \nabla^2 \phi = 0, which is solved to model flow around obstacles like airfoils in or ship hulls in . These solutions provide insights into generation and estimation in low-viscosity regimes, such as over wings or steady-state water flow past vessels. The perfect fluid model also underpins the analysis of water waves, particularly linearized surface waves on or shallow water bodies. For small-amplitude perturbations, the \omega^2 = g k \tanh(k h) relates \omega, k, g, and water depth h, predicting wave propagation speeds and stability in harbors, coastal , and oceanographic forecasting. This approximation facilitates the design of breakwaters and the study of tidal dynamics by neglecting viscous dissipation. Historically, 18th-century developments by Leonhard Euler and applied perfect fluid concepts to , deriving principles for steady flow in channels and pipes that influenced early designs like aqueducts and waterwheels. Euler's generalization of Bernoulli's for inviscid flows provided foundational tools for analyzing efflux from orifices and flow in conduits, marking a shift toward rational hydrodynamic theory. Despite these successes, the perfect fluid model has notable limitations in practical scenarios. It breaks down at high Reynolds numbers, where inertial forces dominate and arises due to unmodeled , as in layers over surfaces or pipe flows exceeding laminar thresholds. Additionally, in compressible regimes, the assumption fails near shocks, where abrupt changes occur, necessitating more advanced models for supersonic flows or blast waves.

Relativistic Fluid Dynamics

Stress-Energy Tensor

In relativistic fluid dynamics, the stress-energy-momentum tensor serves as the relativistic generalization of the momentum flux density, encapsulating the , density, and contributions of a perfect fluid. It acts as the source term in the , coupling the fluid's dynamics to curvature in . The contravariant form of the tensor is T^{\mu\nu} = (\rho + p) u^\mu u^\nu + p g^{\mu\nu}, where \rho denotes the total proper (including rest mass, , and contributions from fields), p is the isotropic , u^\mu is the normalized such that u^\mu u_\mu = -1 in the mostly-plus , and g^{\mu\nu} is the inverse . In the local of the fluid, where u^\mu = (1, 0, 0, 0), the components reduce to T^{00} = \rho, T^{0i} = 0, and T^{ij} = p \delta^{ij}, reflecting the along the time direction and uniform as the spatial . This expression arises from relativistic kinetic theory through the integration of the particle second moments over the phase-space f(x, p) in the fluid , specifically T^{\mu\nu} = \int \frac{d^3 p}{p^0} p^\mu p^\nu f(x, p), under the idealization of no anisotropic stresses or dissipative effects. For a local like the Jüttner distribution, this yields the perfect fluid form with isotropic . As a second-rank tensor, T^{\mu\nu} transforms covariantly under Lorentz transformations, preserving its physical interpretation of energy-momentum flux across inertial frames—unlike the non-relativistic stress tensor, which lacks explicit energy components and relativistic invariance. A canonical example is the photon gas, treated as a perfect fluid of massless particles with p = \rho/3, derived from the isotropic averaging of momenta in , which contributes significantly to early-universe dynamics. The divergence-free condition \nabla_\mu T^{\mu\nu} = 0 encodes the laws governing fluid evolution.

Hydrodynamic Equations

The hydrodynamic equations for relativistic perfect fluids are derived from the local laws of , , and particle number, expressed covariantly in curved . The fundamental principle is the of the stress-energy tensor, given by \nabla_\mu T^{\mu\nu} = 0, where \nabla_\mu denotes the compatible with the g_{\mu\nu}, and T^{\mu\nu} is the stress-energy tensor for the perfect fluid. This equation encodes both the evolution of and the dynamics of the fluid's u^\mu, with u^\mu u_\mu = -1 in the mostly plus . To separate the components parallel and perpendicular to the flow, the tensor \Delta^{\mu\nu} = g^{\mu\nu} + u^\mu u^\nu is employed, which projects tensors orthogonal to u^\mu. Contracting \nabla_\mu T^{\mu\nu} = 0 with u_\nu yields the equation along the fluid worldlines: u_\nu \nabla_\mu T^{\mu\nu} = -(\rho + p) \nabla_\mu u^\mu - u^\mu \nabla_\mu \rho = 0, where \rho is the proper and p is the pressure, both measured in the . The orthogonal , \Delta^\nu_\lambda \nabla_\mu T^{\mu\lambda} = 0, leads to the relativistic Euler equation, describing the acceleration of the : (\rho + p) u^\lambda \nabla_\lambda u^\nu = \Delta^{\nu\sigma} \nabla_\sigma p, where the right-hand side represents the projected perpendicular to u^\nu, and the left-hand side is the relativistic convective of the (the ). This form highlights the balance between inertial forces and pressure gradients in curved . In addition to energy-momentum , perfect fluids often incorporate the of , expressed as \nabla_\mu (n u^\mu) = 0, where n is the proper of in the fluid . This ensures the preservation of particle number along the , serving as a in relativistic terms. For adiabatic processes without dissipation, is imposed via \nabla_\mu (s n u^\mu) = 0, where s is the specific per baryon, constant along streamlines in isentropic flows. These scalar laws close the system when combined with an relating \rho, p, n, and s. In the special relativistic limit, applicable in flat Minkowski , the covariant derivatives reduce to partial derivatives, so the energy-momentum conservation simplifies to \partial_\mu T^{\mu\nu} = 0, while the and equations become \partial_\mu (n u^\mu) = 0 and \partial_\mu (s n u^\mu) = 0, respectively. This flat-space case recovers the standard form of relativistic hydrodynamics without gravitational effects, providing a foundation for the more general curved- formulation.

Applications in Astrophysics and Cosmology

Cosmological Models

In cosmological models, the perfect fluid approximation is widely used to describe the large-scale dynamics of the within the Friedmann-Lemaître-Robertson-Walker (FLRW) , which assumes spatial homogeneity and . The 's content—such as , , and —is modeled as a perfect fluid with \rho and isotropic p, neglecting and conduction on cosmic scales. This simplification arises from the observed uniformity of the and the success of the \LambdaCDM model in fitting observations. The perfect fluid stress-energy tensor T^{\mu\nu} = (\rho + p/c^2) u^\mu u^\nu + p g^{\mu\nu}, where u^\mu is the of comoving observers and g^{\mu\nu} is the , sources the in . The dynamics of the FLRW universe are governed by the Friedmann equations, derived by substituting the FLRW metric into Einstein's equations with the perfect fluid stress-energy tensor. The first Friedmann equation relates the Hubble parameter H = \dot{a}/a (where a(t) is the scale factor and dot denotes time derivative) to the energy density and curvature: \left( \frac{\dot{a}}{a} \right)^2 = \frac{8\pi G}{3} \rho - \frac{k c^2}{a^2}, where G is the gravitational constant, c is the speed of light, and k is the spatial curvature parameter (k = -1, 0, +1). The second equation describes the acceleration: \frac{\ddot{a}}{a} = -\frac{4\pi G}{3} \left( \rho + \frac{3p}{c^2} \right). These equations capture the expansion history, with \rho and p evolving according to the conservation law \dot{\rho} + 3H(\rho + p/c^2) = 0, which follows from the Bianchi identities. Different epochs of the universe are dominated by perfect fluids with distinct equations of state p = w \rho c^2, where the parameter w determines the sign and magnitude of cosmic acceleration or deceleration via the second Friedmann equation. In the matter-dominated era, non-relativistic baryonic and dark matter behave as a pressureless perfect fluid (w = 0), leading to a scale factor evolution a \propto t^{2/3} in a flat universe, consistent with structure formation observations from redshift z \approx 1100 to z \approx 0.3. The radiation-dominated era, prevalent in the early universe (z > 3000), features relativistic particles like photons and neutrinos as a perfect fluid with w = 1/3, yielding a \propto t^{1/2} and a decelerating expansion. For dark energy, modeled as a cosmological constant (\Lambda) in the \LambdaCDM framework, w = -1 (so p = -\rho c^2), driving exponential acceleration a \propto e^{H t} at late times (z < 0.5), as evidenced by Type Ia supernovae and cosmic microwave background data. As of 2025, constraints from DESI and other surveys are consistent with w \approx -1 within uncertainties, but recent analyses suggest possible evolution in the dark energy equation of state. Recent DESI results (as of 2025) refine the baryon density \eta and suggest possible time-varying w, tightening tests of perfect fluid assumptions in \LambdaCDM. Values of w > -1/3 cause deceleration, while w < -1/3 leads to acceleration. In the context of (BBN), occurring roughly 1–20 minutes after the at temperatures T \sim 0.1–1 MeV, baryonic matter is treated as a non-relativistic perfect fluid (w \approx 0) interacting with the radiation-dominated . The low baryon-to-photon ratio (\eta \approx 6 \times 10^{-10}) ensures the remains nearly radiation-dominated, allowing weak interactions to freeze out and set the neutron-to-proton ratio (\approx 1/6) before begins. This perfect fluid model predicts primordial abundances of light elements—such as ^4He (Y_p \approx 0.247), ^2H (\approx 2.53 \times 10^{-5}), ^3He (\approx 10^{-5}), and ^7Li (\approx 5 \times 10^{-10})—in good agreement with observations for ^4He, D, and ^3He from absorption lines and stellar spectra, though the ^7Li prediction exceeds observations by (the ), providing a key test of the standard cosmological model. Deviations from perfect fluid assumptions, such as viscosity, would alter freeze-out dynamics and element yields.

Stellar Structure

In stellar structure, the perfect fluid approximation is widely used to model the internal of and compact objects, where is treated as a continuous distribution without or heat conduction, allowing focus on hydrostatic balance between and . This idealization simplifies the equations governing spherically symmetric configurations, enabling analytical and numerical solutions for and pressure profiles. In the Newtonian limit, valid for low-mass stars where gravitational fields are weak, hydrostatic equilibrium requires that the pressure gradient balances the gravitational force per unit volume. The governing equation is \frac{dp}{dr} = -\rho \frac{G m(r)}{r^2}, where p is pressure, \rho is density, G is the gravitational constant, r is radial distance, and m(r) is the mass enclosed within radius r. This is coupled with the mass continuity equation \frac{dm}{dr} = 4\pi r^2 \rho, derived from the divergence of the in spherical symmetry. These equations, solved iteratively from the stellar outward, yield structure models consistent with observed luminosities and radii for main-sequence stars. For relativistic stars, where strong gravity necessitates , the Tolman-Oppenheimer-Volkoff (TOV) equation generalizes for a perfect fluid. It reads \frac{dp}{dr} = -(\rho + \frac{p}{c^2}) \frac{G m(r)}{r^2} \left(1 + \frac{p}{\rho c^2}\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}, with c the ; this incorporates corrections from curvature and relativistic fluid . The TOV equation, paired with mass and an p = p(\rho), determines the internal structure of compact objects like neutron stars, predicting maximum masses around 2-3 solar masses depending on the nuclear . Polytropic models provide tractable solutions by assuming an p = K \rho^{1 + 1/n}, where K is a constant and n is the polytropic index. Substituting into the Newtonian equations yields the Lane-Emden equation, \frac{1}{\xi^2} \frac{d}{d\xi} \left( \xi^2 \frac{d\theta}{d\xi} \right) = -\theta^n, in dimensionless variables \xi (scaled radius) and \theta (scaled potential, with \rho \propto \theta^n). Solutions to this boundary-value problem (\theta(0) = 1, \theta'(0) = 0) define density profiles; for example, n=1 yields an analytic sinc function, while numerical integration is needed for other n. These models approximate convective or radiative zones in stars, with n \approx 3/2 for radiative envelopes and n=3 for convective cores. White dwarfs exemplify a perfect fluid where dominates, modeled as a non-relativistic with p \propto \rho^{5/3} (corresponding to n=3/2). This polytropic relation supports stars up to about 1.4 solar masses against gravity, beyond which relativistic effects soften the equation of state (p \propto \rho^{4/3}, n=3), leading to instability. Observations of Sirius B confirm this structure, with radii around Earth's size and surface gravities exceeding 10^8 times solar. Neutron stars treat baryonic as a perfect fluid, with pressure arising from neutron degeneracy and strong interactions constrained by (QCD). The equation of state transitions from soft at low densities to stiff quark-gluon phases at high densities (above ~2-3 times saturation), enabling support for masses up to ~2.3-2.5 masses, as observed in pulsars like PSR J0952-0607. Microscopic QCD calculations, including lattice simulations, inform these models, ensuring consistency with gravitational wave signals from mergers. The Oppenheimer-Snyder model demonstrates the role of perfect fluids in gravitational collapse, using pressureless dust (p=0) for a homogeneous sphere. Solving Einstein's equations for this fluid matched to a Schwarzschild exterior shows inevitable formation of an event horizon, marking the birth of a black hole as the star's radius shrinks below $2GM/c^2. This seminal solution highlights how perfect fluid dynamics under general relativity preclude stable equilibria for sufficiently massive configurations.

Special Cases and Extensions

Superfluidity

Superfluids represent a quantum mechanical realization of perfect fluids, exhibiting zero viscosity and frictionless flow under specific conditions, such as low temperatures. In liquid helium-4 (^4He), superfluidity emerges below the λ-transition temperature of approximately 2.17 K, where the fluid transitions from the normal helium I phase to the superfluid helium II phase, allowing it to flow without resistance through narrow channels. This behavior was first experimentally observed in 1938 through independent measurements: Pyotr Kapitza in Moscow demonstrated the absence of viscosity by observing unrestricted flow under pressure gradients, while John F. Allen and Donald Misener at the University of Toronto confirmed zero viscous drag in capillary tubes. These findings established superfluidity as a macroscopic quantum phenomenon, distinct from classical inviscid flows due to its underlying quantum coherence. The theoretical framework for superfluid hydrodynamics is provided by Lev Landau's two-fluid model, introduced in 1941, which describes helium II as a of two interpenetrating components: a normal fluid with ρ_n that carries all and , and a superfluid component with ρ_s that flows without dissipation. The total is given by ρ = ρ_n + ρ_s, where ρ_n vanishes as temperature approaches , and ρ_s approaches the total . The superfluid v_s is irrotational, satisfying ∇ × v_s = 0, enabling descriptions, while the normal component v_n experiences viscous effects and thermal gradients. This model successfully predicts phenomena like the fountain effect and , where temperature waves propagate due to counterflow between the components. A hallmark of superfluidity is the quantization of circulation around vortices, arising from the single-valuedness of the wavefunction in the Bose-Einstein description proposed by and . The circulation ∮ v · dl is quantized in units of κ = h/m, where h is Planck's constant and m is the mass of the (approximately 6.65 × 10^{-27} kg for ^4He), yielding κ ≈ 9.97 × 10^{-8} m²/s. In rotating superfluids, these singly quantized vortices arrange into lattices, mimicking classical solid-body rotation but with discrete , observable in experiments with rotating buckets of . Vortex dynamics, including reconnection and , govern dissipation in turbulent superfluids. Superfluid flow remains dissipationless only below a critical velocity v_c, beyond which quantized vortices proliferate, leading to onset of . Theoretically, v_c ≈ (ħ/m) / ξ, where ħ = h/2π is the reduced Planck's constant and ξ is the healing length (or ), typically on the order of 10^{-8} m near the λ-point, decreasing to scales at lower temperatures. Experimental measurements in narrow channels confirm this scaling, with v_c reaching up to ~60 m/s in pure conditions, limited by roton creation or vortex . In relativistic contexts, manifests in the cores of , where degenerate neutron matter pairs via BCS-like mechanisms, forming a with gaps Δ on the order of 0.1–1 MeV. These gaps suppress emission and enable phenomena in spin-down, modeled using relativistic two-fluid hydrodynamics that incorporate the superfluid order parameter. Observations of cooling and rotation irregularities provide indirect evidence for such superfluid states.

Equations of State

In perfect fluid dynamics, the equation of state relates the isotropic p to the \rho and density s, generally taking the form p = p(\rho, s). This functional dependence arises from the requirement of local , where the fluid's state is fully specified by these variables. The associated relativistic thermodynamic identity, derived from of in the fluid's , is d\rho = T \, ds + \mu \, dn, where T is the , \mu the , and n the particle number density; equivalently, the pressure follows from the Gibbs-Duhem relation dp = s \, dT + n \, d\mu. A common specific case is the ideal gas equation of state, applicable to non-degenerate gases in relativistic regimes. Here, p = (\gamma - 1)(\rho - \rho_m c^2), where \gamma = C_p / C_v is the adiabatic index (ratio of specific heats at constant pressure and volume), \rho_m is the rest-mass density, and c the speed of light; this form separates the rest-mass contribution from the internal (thermal) energy density. Polytropic equations of state provide a versatile for self-gravitating systems, expressed as p = K \rho^\gamma, where K is a constant and \gamma the polytropic index (often related to the adiabatic index). These are widely used in modeling stellar interiors and cosmological expansions due to their analytic tractability in solving . For radiation-dominated fluids, such as those consisting of photons or ultra-relativistic particles, the equation of state simplifies to p = \frac{1}{3} \rho, reflecting the equality of pressure and contributions in the relativistic limit. In degenerate fermionic matter, relevant to compact objects like white dwarfs, the pressure arises from the rather than thermal motion. For non-relativistic electrons, p \propto \rho^{5/3}; in the ultra-relativistic limit, p \propto \rho^{4/3}. These forms support stellar structures against up to a maximum . The cosmological constant represents as a perfect fluid with p = -\rho, leading to accelerated expansion in cosmological models. The speed of sound c_s in a perfect fluid is given by c_s^2 = \left( \frac{\partial p}{\partial \rho} \right)_s, evaluated at constant entropy; for physical consistency in relativistic theories, c_s < c ensures causality and stability.

References

  1. [1]
    [PDF] VI.3 Perfect relativistic fluid
    VI.3 Perfect relativistic fluid. By definition, a fluid is perfect when there is no dissipative current in it, see definition (III.16a).
  2. [2]
  3. [3]
  4. [4]
    [PDF] Perfect fluids - SciPost
    Jul 19, 2018 · We do this for perfect fluids, i.e. maintaining local equilib- rium, whose energy-momentum tensor we define in the rest or laboratory frame. In ...
  5. [5]
    [PDF] Notes on the History of the General Equations of Hydrodynamics
    Oct 26, 2001 · In fact, Euler's equations refer to fluids without dissipative mechanisms, sometimes called "perfect" or. "ideal" fluids. While such fluids ...
  6. [6]
    Solution generating theorems for perfect fluid spheres - IOP Science
    The first static spherically symmetric perfect fluid solution with constant density was found by Schwarzschild in 1918. Generically, perfect fluid spheres are ...<|control11|><|separator|>
  7. [7]
    [PDF] Introduction to 3+1 Numerical Relativity
    Jun 5, 2025 · ... perfect fluid, that is, a fluid with zero viscosity and zero heat conduction. Such a fluid is characterized by the energy density ρ and ...
  8. [8]
    [PDF] Perfect Fluids arXiv:1710.04708v3 [hep-th] 14 May 2018
    May 14, 2018 · It involves introducing a fluid variable, the kinetic mass density, which is needed to define the most gen- eral energy-momentum tensor for ...
  9. [9]
    [PDF] Fluid Mechanics - DAMTP
    equation of state that relates the pressure to the density,. P = P(ρ). For example, for the ideal gas undergoing adiabatic deformations, we showed that the.
  10. [10]
    [PDF] Relativistic thermodynamics of perfect fluids - arXiv
    Oct 11, 2022 · The internal energy density and the pressure of a relativistic perfect fluid carry inertia, which leads to a relativistic coupling between heat ...
  11. [11]
    [PDF] Chapter 6 Fluid Mechanics - MIT OpenCourseWare
    In particular, the continuity equation re- mains unchanged, while the Euler equations along with the conservation laws for momentum, energy, and entropy must be ...Missing: perfect governing
  12. [12]
    [PDF] The Euler equations - Principles of Fluid Dynamics
    The same fluid can be considered either ideal or not, depending on the case. The continuity and Euler equations were derived by Leonhard Euler in 1755 [16] ...
  13. [13]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · The speed of sound, c, in a substance is the speed at which infinitesimal pressure disturbances propagate through the surrounding substance.Missing: c_s drho
  14. [14]
    [PDF] I Fundamentals
    This first chapter reviews the fundamental principles of fluid mechanics, emphasizing the relationship between the underlying microscopic description of.
  15. [15]
    Potential Flow Theory – Introduction to Aerospace Flight Vehicles
    A potential flow assumes the aggregate of an incompressible, irrotational, and inviscid fluid motion, ie, which is called an “ideal” flow.Missing: perfect | Show results with:perfect
  16. [16]
    [PDF] 3 IRROTATIONAL FLOWS, aka POTENTIAL FLOWS - DAMTP
    condition for Laplace's equation.* If the flow is irrotational, this is the only boundary condition we need. We can now apply techniques from Part IB Methods.
  17. [17]
  18. [18]
    [PDF] Chapter 1 Lecture: Review of Linear Surface Gravity Waves
    ω2 = gk tanh(kh). (1.13). The pressure under the fluid is can also be solved for now with the linearized Bernoulli's equation: p ... This means that kh 1, which ...Missing: perfect | Show results with:perfect
  19. [19]
    [PDF] A History of Hydrodynamics from the Bernoullis to Prandtl
    The first attempt at applying a general dynamical principle to fluid motion occurred in Daniel Bernoulli's Hydrodynamica of 1738. The principle was the ...
  20. [20]
    [PDF] Brief History of Fluid Mechanics - Joseph Majdalani
    Euler developed both the differential equations of motion and their integrated form, now called the Bernoulli equation. D'Alembert used them to show his famous ...<|control11|><|separator|>
  21. [21]
    Types of Fluid Flows – Introduction to Aerospace Flight Vehicles
    A fluid that is not subjected to the action of viscosity or compressibility and does not flow in a turbulent manner is called an ideal or inviscid fluid.
  22. [22]
    The Reynolds Number: A Journey from Its Origin to Modern ... - MDPI
    While the Reynolds number is essential to understanding fluid dynamics, its limitations in highly turbulent flows or complex geometries can complicate ...
  23. [23]
    Stress-energy tensor - Cosmological Dynamics - E. Bertschinger
    The stress-energy tensor provides the source for metric variables. It includes energy density, energy flux density, and spatial stress tensor.<|separator|>
  24. [24]
  25. [25]
    [PDF] General Relativity
    3.2 Stress-Energy-Momentum Tensor for a Perfect Fluid and for. Dust ... In General Relativity, we must allow for the definition of a tensor related to.<|control11|><|separator|>
  26. [26]
    [PDF] The stress-energy (energy-momentum) tensor - No contents here
    The stress-energy tensor is a symmetric tensor both in the dust (and perfect fluid) formulation, and in the EM case. This is a general property, associated with ...
  27. [27]
    [PDF] AN INTRODUCTION TO RELATIVISTIC HYDRODYNAMICS
    In this section, we illustrate the power of the Carter-Lichnerowicz equation by deriving from it various conservation laws in a very easy way. We consider a.
  28. [28]
    [PDF] Relativistic fluid dynamics - Nikhef
    Relativistic fluid dynamics involves the energy-momentum tensor, conservation laws, and the Euler equation, with the fluid described by a density current.
  29. [29]
    [PDF] 22. Big-Bang Cosmology - Particle Data Group
    May 31, 2024 · The Big-Bang model proposes the early Universe was very hot and dense, and has since expanded and cooled. It began in the 1940s with Gamow, ...
  30. [30]
    [PDF] Updated Cosmology - UNAM
    Aug 12, 2017 · ... fluid is at rest. This is the stress-energy tensor of a perfect-fluid as seen by a comoving observer. More generally. Tµν = p c2. + ρ. uµuν ...
  31. [31]
    Dark Energy and the Accelerating Universe - J.A. Frieman et al.
    During the radiation-dominated era, a(t) propto t1/2; during the matter-dominated era, a(t) ... A striking success of the consensus cosmology is its ability to ...Missing: ∝ | Show results with:∝
  32. [32]
    [PDF] 28. Dark Energy - Particle Data Group
    Dec 1, 2023 · typically specified in terms of the equation-of-state-parameter w = p/ρ (= −1 for a cosmological constant). Furthermore, the conclusion ...
  33. [33]
    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.
  34. [34]
    [PDF] Equations of Stellar Structure
    If the time derivative in the equation (1.11a) vanishes then the star is in hydrostatic equilibrium. If the time derivative in equation (1.11d) vanishes then ...
  35. [35]
    Static Solutions of Einstein's Field Equations for Spheres of Fluid
    A method is developed for treating Einstein's field equations, applied to static spheres of fluid, in such a manner as to provide explicit solutions.
  36. [36]
    On Continued Gravitational Contraction | Phys. Rev.
    In the present paper we study the solutions of the gravitational field equations which describe this process.Missing: dust | Show results with:dust
  37. [37]
    January 1938: Discovery of Superfluidity - American Physical Society
    Jan 1, 2006 · When helium-4 is chilled to below about 2.2 K, it starts to behave in some very weird ways. The fluid passes through narrow tubes with ...
  38. [38]
    Critical velocity in superfluid helium
    It is found that for ~/d << 1, where ~ is the coherence length of the system, the superfluid critical velocity Ve OC d -1/2.<|control11|><|separator|>
  39. [39]
    Quantized Vortex Rings in Superfluid Helium | Phys. Rev.
    The circulation of these vortex rings can be determined by measuring their energy and velocity; it is found to be equal to one quantum ℎ m , where ℎ is Planck's ...
  40. [40]
    Critical velocities in flows of superfluid 4 He - AIP Publishing
    Mar 21, 2025 · We review and discuss the underlying physics of critical velocities in flows of superfluid 4He (for historical reasons known as He II).
  41. [41]
    Critical velocity in superfluid helium | Journal of Low Temperature ...
    Dec 13, 1976 · It is found that for ξ c /d≪1, where ξ c is the coherence length of the system, the superfluid critical velocityv c ∝d −1/2.
  42. [42]
    [astro-ph/0008161] Superfluidity in relativistic neutron stars - arXiv
    Aug 10, 2000 · The purpose of these notes is to give a brief review of superfluidity in neutron stars. After a short presentation explaining why and how ...Missing: gaps | Show results with:gaps
  43. [43]
    Superfluidity in Relativistic Neutron Stars - SpringerLink
    The purpose of this chapter is to give a brief review of superfluidity in neutron stars. After a short presentation explaining why and how superfluidity is ...Missing: gaps | Show results with:gaps
  44. [44]
    [PDF] Relativistic fluid dynamics: physics for many different scales
    Jun 1, 2020 · Abstract. The relativistic fluid is a highly successful model used to describe the dynam- ics of many-particle systems moving at high ...
  45. [45]
    [PDF] 1. White Dwarfs and Neutron Stars
    We saw that the pressure of degenerate electrons has two limiting cases, both of the form. P ∝ ργ , where in the non-relativistic limit γ = 5/3 , while in the ...