Fact-checked by Grok 2 weeks ago

Inviscid flow

Inviscid flow, also referred to as ideal flow, describes the motion of a in which the effects of and other dissipative , such as and , are completely neglected, treating the as perfectly frictionless and non-viscous. This approximation simplifies the mathematical modeling of , allowing for the analysis of flows where viscous forces are negligible compared to inertial and forces, such as in high-Reynolds-number regimes. The core governing equations for inviscid flow are the Euler equations, which include the expressing mass conservation and the equations capturing the balance of inertial, , and body forces without viscous terms. For incompressible inviscid flows—applicable to scenarios where speeds are much lower than the , such as —these reduce to the incompressible Euler equations: \nabla \cdot \mathbf{u} = 0 for incompressibility and \partial_t \mathbf{u} + (\mathbf{u} \cdot \nabla) \mathbf{u} + \nabla p = 0 for , where \mathbf{u} is the velocity field and p is (with normalized to unity). A key feature of many inviscid flow analyses is the assumption of irrotationality, where the \nabla \times \mathbf{u} = 0, enabling the use of theory to represent the velocity as the gradient of a \phi, i.e., \mathbf{u} = \nabla \phi. This irrotational condition is particularly valid for external flows around streamlined bodies, such as airfoils or wings, where rotational effects from are confined to thin s. From the incompressible Euler equations under irrotationality follows Bernoulli's equation, p + \frac{1}{2} |\mathbf{u}|^2 = \text{constant}, which relates pressure and velocity throughout the flow field and underpins much of inviscid flow prediction for and pressure distributions in steady flows. Inviscid flow models are widely applied in to compute inviscid pressure loads on components, rotor blades, and launch vehicles, often serving as the outer flow solution coupled with boundary layer corrections for more realistic viscous effects. However, these approximations break down near surfaces where no-slip conditions cannot hold without , leading to phenomena like , where inviscid theory predicts zero for steady flow past a body despite real-world drag.

Basic Concepts

Definition and Assumptions

Inviscid flow describes the idealized motion of a in which viscous effects are entirely neglected, treating the as possessing zero dynamic (μ = 0), thereby eliminating internal and stresses within the . This model assumes that the behaves without dissipative , such as , mass , or thermal conductivity, focusing solely on inertial forces in the conservation of momentum. The approximation is valid in scenarios where the approaches infinity, indicating that viscous forces are negligible relative to inertial ones, such as in high-speed external flows like air over aircraft wings or low- fluids at large scales. The primary assumptions underpinning inviscid flow include the absence of any frictional resistance between fluid layers, which simplifies the governing equations by removing viscous terms, and the conservation of without accounting for energy dissipation due to . These assumptions enable the treatment of the fluid as frictionless and often irrotational for further simplification in analyses, though the core inviscid condition stands alone. Physically, this leads to streamlined predictions for pressure distributions and forces in external flows, facilitating analytical solutions that are computationally efficient and insightful for preliminary design in . However, the implications of these assumptions reveal limitations: inviscid flow ignores phenomena like and , which arise from real viscous effects, as exemplified by where steady, inviscid flow around a predicts zero net drag despite experimental evidence to the contrary. This idealization thus provides a foundational understanding but requires viscous corrections for accurate real-world predictions, particularly near surfaces. Historically, the concept of inviscid flow originated in 18th-century hydrodynamics, with introducing principles relating pressure and velocity in frictionless fluids and Leonhard Euler deriving the fundamental for such flows in 1757. These early contributions laid the groundwork for inviscid theory, which was later expanded in the through formalizations like methods by researchers such as .

Reynolds Number and Flow Regimes

The Reynolds number, denoted as Re, serves as a dimensionless parameter that quantifies the ratio of inertial forces to viscous forces within a fluid flow, providing a critical criterion for determining when the inviscid flow approximation is applicable. It is defined mathematically as Re = \frac{\rho U L}{\mu}, where \rho is the fluid density, U is a characteristic velocity, L is a characteristic length scale, and \mu is the dynamic viscosity of the fluid. This formulation, originally introduced in the context of pipe flows, allows engineers and scientists to scale flow behaviors across different systems without dimensional analysis complications. At low Reynolds numbers (typically Re < 1), viscous forces dominate, leading to creeping or Stokes flow where diffusion of momentum prevails over convection, and the flow remains highly ordered and predictable. In contrast, high Reynolds numbers (often Re > 10^3 or higher) indicate inertia-dominated regimes, where viscous effects become negligible except in thin layers near solid surfaces, justifying the use of inviscid models for the bulk flow. As Re \to \infty, the flow approaches the ideal inviscid limit, enabling simplifications like the Euler equations for analyzing external or free-stream conditions. Flow regimes are broadly classified based on Reynolds number transitions, with laminar flows occurring at moderate Re where streamlines remain parallel and smooth, and turbulent flows emerging at higher Re (e.g., Re > 2300 in pipes) characterized by chaotic eddies and enhanced mixing. However, the inviscid approximation is particularly relevant in the high-Re turbulent regime away from walls, as turbulence models often incorporate inviscid cores. This classification was experimentally validated through pioneering pipe flow studies by Osborne Reynolds in 1883, who observed dye streaks transitioning from straight laminar paths to erratic turbulent dispersion as flow speed increased, establishing Re as the governing parameter for regime shifts. In practical applications, such as , Reynolds numbers exceeding $10^6—common for at cruise—permit inviscid simulations to accurately predict distributions and with minimal viscous corrections, as demonstrated in scaling and computational validations. These thresholds underscore the Reynolds number's role in bridging theoretical inviscid ideals with real-world viscous influences.

Governing Equations

Euler Equations

The Euler equations constitute the fundamental system of partial differential equations that govern the motion of an inviscid, compressible fluid, assuming no viscosity and often adiabatic conditions. Originally derived by Leonhard Euler in his 1757 memoir "Principes généraux du mouvement des fluides," these equations extend earlier work on incompressible flows to the compressible case, marking one of the earliest formulations of partial differential equations in . Euler's derivation emphasized the conservation principles for fluids under forces, laying the groundwork for modern and gas dynamics. In conservative form, the Euler equations express the conservation of , , and total across control volumes, making them suitable for numerical simulations involving discontinuities. For a three-dimensional flow, they are written as \frac{\partial}{\partial t} \begin{pmatrix} \rho \\ \rho \mathbf{u} \\ \rho E \end{pmatrix} + \nabla \cdot \begin{pmatrix} \rho \mathbf{u} \\ \rho \mathbf{u} \otimes \mathbf{u} + p \mathbf{I} \\ (\rho E + p) \mathbf{u} \end{pmatrix} = 0, where \rho is the fluid density, \mathbf{u} is the velocity vector, p is the , \mathbf{I} is the tensor, and E = e + \frac{1}{2} |\mathbf{u}|^2 is the total energy per unit mass with e denoting the . This form ensures that integrals over arbitrary volumes remain conserved in the absence of sources, preserving , , and in the inviscid limit. To close the system, an is required, such as the p = (\gamma - 1) \rho e for a polytropic gas with constant specific heat ratio \gamma. The non-conservative form, also known as the primitive form, expresses the equations in terms of and , which is often more intuitive for analytical solutions in smooth flows. The remains \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{u}) = [0](/page/0), while the simplifies to \frac{D \mathbf{u}}{Dt} = -\frac{[1](/page/1)}{\rho} \nabla p, where \frac{D}{Dt} = \frac{\partial}{\partial t} + \mathbf{u} \cdot \nabla is the . The , devoid of dissipative terms, takes the form \frac{D E}{Dt} = -\frac{[1](/page/1)}{\rho} \nabla \cdot (p \mathbf{u}), ensuring no viscous heating or friction losses. As a system of hyperbolic partial differential equations, the Euler equations support wave propagation at finite speeds, characterized by real eigenvalues corresponding to and particle velocities. This hyperbolic nature permits the formation of shocks and contact discontinuities in solutions, where smooth initial data can develop singularities in finite time, necessitating weak solutions that satisfy jump conditions like the Rankine-Hugoniot relations. In the inviscid regime, these equations rigorously conserve mass, momentum, and total energy globally, distinguishing them from viscous counterparts by excluding and mechanisms. The Euler equations represent the zero-viscosity limit of the Navier-Stokes equations, capturing high-Reynolds-number behaviors without effects.

Derivation from Navier-Stokes Equations

The Navier-Stokes equations govern the motion of viscous, Newtonian fluids, incorporating both inertial and viscous effects. The momentum equation takes the form \rho \left( \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} \right) = -\nabla p + \nabla \cdot \boldsymbol{\tau} + \rho \mathbf{g}, where \rho is the fluid density, \mathbf{u} is the velocity vector, p is the pressure, \mathbf{g} is the body force per unit mass (e.g., gravity), and \boldsymbol{\tau} is the viscous stress tensor. For Newtonian fluids under the Stokes hypothesis, the deviatoric part of the stress tensor is expressed as \boldsymbol{\tau} = \mu \left[ \nabla \mathbf{u} + (\nabla \mathbf{u})^T - \frac{2}{3} (\nabla \cdot \mathbf{u}) \mathbf{I} \right], with \mu as the dynamic viscosity and \mathbf{I} the identity tensor; this form assumes the bulk viscosity is zero, reducing the general second viscosity coefficient \lambda = -\frac{2}{3} \mu. The continuity equation for mass conservation, \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{u}) = 0, remains unaffected by viscosity. The energy equation in the Navier-Stokes system includes terms for heat conduction and viscous dissipation, \rho \frac{D h}{Dt} = \frac{D p}{Dt} + \nabla \cdot (k \nabla T) + \Phi, where h is the specific enthalpy, k is the thermal conductivity, T is temperature, \frac{D p}{Dt} = \frac{\partial p}{\partial t} + \mathbf{u} \cdot \nabla p, and \Phi is the dissipation function proportional to \mu. To derive the inviscid Euler equations, viscous effects are neglected by setting \mu = 0, which eliminates the \nabla \cdot \boldsymbol{\tau} term in the momentum equation, leaving only the inertial, pressure gradient, and body force contributions: \rho \left( \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} \right) = -\nabla p + \rho \mathbf{g}. This simplification retains the continuity equation unchanged. For the energy equation, setting \mu = 0 removes the dissipation \Phi, and neglecting heat conduction by taking k = 0 yields the inviscid form \rho \frac{D h}{Dt} = \frac{D p}{Dt}. These steps apply similarly to compressible or incompressible cases, with the approximation justified in flows where the Reynolds number \mathrm{Re} = \frac{\rho U L}{\mu} \gg 1, rendering viscous terms negligible relative to inertial ones. The derivation assumes an isotropic stress state reduced to hydrostatic pressure alone, eliminating shear and normal viscous stresses, and ignores thermal conduction, focusing on adiabatic processes without diffusive heat transfer. This is valid for high-Reynolds-number regimes, such as external or geophysical flows away from boundaries. However, the inviscid model loses the at solid surfaces, as viscous forces no longer enforce zero tangential . Furthermore, the Navier-Stokes equations exhibit parabolic character due to viscous , enabling smoothing of irregularities, whereas the Euler equations are hyperbolic, permitting sharp discontinuities like shocks without inherent dissipation.

Boundary Conditions

Slip Condition at Solid Boundaries

In inviscid flow, the boundary condition at a solid surface enforces impermeability, meaning the normal component of the must vanish to prevent through the wall. This is expressed mathematically as \mathbf{u} \cdot \mathbf{n} = 0, where \mathbf{u} is the and \mathbf{n} is the unit to the surface. Unlike viscous flows, there is no enforcement of a on the tangential component; instead, \mathbf{u} \cdot \mathbf{t} \neq 0, where \mathbf{t} is the unit , allowing the to slip freely along the . This slip condition corresponds to an infinite slip length, reflecting the absence of to generate stresses at the . The slip boundary condition arises naturally from the Euler equations governing inviscid flow, which are first-order partial differential equations requiring only a single per surface, typically the normal component. In contrast to the Navier-Stokes equations for viscous fluids, where both normal and tangential velocities are zero at solid walls (), the inviscid approximation simplifies modeling by eliminating frictional effects at . This leads to solutions where the fluid can be discontinuous or slip indefinitely parallel to the wall, which is physically idealized but useful for high-Reynolds-number approximations far from surfaces. For applications involving in steady, incompressible , the slip condition alone is insufficient to uniquely determine the flow field around bodies with sharp edges, such as trailing edges. The Kutta-Joukowski condition addresses this by stipulating that the rear coincides with the trailing edge, ensuring smooth flow departure and finite velocity there; this fixes the circulation \Gamma around the airfoil as \Gamma = \pi c U_\infty \sin \alpha, where c is the length, U_\infty is the freestream velocity, and \alpha is the angle of attack. The resulting per unit span is then L' = \rho_\infty U_\infty \Gamma, per the Kutta-Joukowski theorem, enabling prediction of aerodynamic without viscous effects. A key implication of the slip condition is the absence of skin friction drag, as no shear stresses act tangentially on the surface to resist motion. For steady inviscid flow past a closed body, this culminates in , which predicts zero net drag force despite intuitive expectations of resistance. This paradox, resolved in real flows by thin viscous boundary layers near the wall that introduce drag, underscores the limitations of pure inviscid models for drag prediction while highlighting their utility for pressure-dominated phenomena like .

Prandtl's Boundary Layer Hypothesis

introduced the boundary layer hypothesis in 1904 during his presentation at the Third in , , offering a groundbreaking reconciliation between ideal inviscid flow models and the real-world viscous effects dominant near solid boundaries. In this work, titled "Über Flüssigkeitsbewegung bei sehr kleiner Reibung" (On Fluid Motion with Very Small Friction), Prandtl argued that friction confines its influence to a narrow region adjacent to the surface, allowing the bulk of the flow to be treated as effectively inviscid while satisfying the at the wall. This conceptual division addressed longstanding paradoxes in , such as the d'Alembert paradox, where inviscid theory predicted zero drag on bodies despite experimental evidence to the contrary. At the heart of the hypothesis lies the idea of a thin where plays a crucial role, with its thickness scaling as \delta \sim L / \sqrt{\mathrm{Re}}, where L is a of the body and \mathrm{Re} is the based on that length and the free-stream velocity. As \mathrm{Re} becomes large, this layer thins relative to L, enabling the outer flow to approximate inviscid conditions with a slip boundary consistent with the Euler equations. Inside the , the flow exhibits rapid variations normal to the surface, such that normal derivatives dominate over streamwise ones by a factor of \sqrt{\mathrm{Re}}, which justifies neglecting streamwise diffusion of compared to and normal viscous diffusion. These scalings lead to the Prandtl boundary layer equations, a reduced set of the Navier-Stokes equations that retain essential viscous terms while simplifying the problem to a more tractable parabolic form. The hypothesis's historical significance stems from its role in establishing matched asymptotic expansions, where the inviscid outer solution provides the pressure distribution driving the viscous inner layer, fundamentally advancing high-Reynolds-number flow analysis. In applications, it elucidates on by quantifying from velocity profiles within the layer, typically contributing the majority of total in streamlined bodies at high \mathrm{Re}. It also predicts boundary layer separation under adverse pressure gradients, where reverse flow forms separation bubbles, sharply increasing pressure and leading to phenomena like airfoil stall. The approach is asymptotically valid for flows with \mathrm{Re} \gg 1, providing a cornerstone for modern and engineering predictions of and .

Applications and Examples

Potential Flow Theory

Potential flow theory describes a class of inviscid flows that are also irrotational, meaning the vorticity vector \nabla \times \mathbf{u} = 0 everywhere in the flow field. This assumption simplifies the governing equations significantly, allowing analytical solutions for idealized fluid motions. Such flows are particularly useful for modeling external away from viscous boundary layers, where rotational effects are negligible. Under the irrotational condition, the velocity field \mathbf{u} can be expressed as the gradient of a scalar \phi, so \mathbf{u} = \nabla \phi. For incompressible flows, the \nabla \cdot \mathbf{u} = 0 then reduces to \nabla^2 \phi = 0. Solutions to this are functions, which possess properties like the mean-value theorem and , facilitating the use of superposition and boundary integral methods. Along streamlines in unsteady irrotational flow, Bernoulli's equation holds: \frac{\partial \phi}{\partial t} + \frac{1}{2} |\nabla \phi|^2 + \frac{[p](/page/Pressure)}{\rho} + [g](/page/Gravity)z = F(t), where F(t) is an arbitrary function of time, [p](/page/Pressure) is , \rho is , [g](/page/Gravity) is , and z is . For steady flow, this simplifies to \frac{1}{2} |\mathbf{u}|^2 + \frac{[p](/page/Pressure)}{\rho} + [g](/page/Gravity)z = \text{constant}. This integral of the Euler equations provides pressure distributions directly from the potential. Basic potential flows serve as building blocks for more complex configurations via linear superposition, as solutions to form a . Uniform flow has potential \phi = U x in Cartesian coordinates, where U is the free-stream speed. A two-dimensional (or for negative strength) emits radially with \phi = \frac{m}{2\pi} \ln r, where m is the strength related to volume flow rate. The , a limiting case of source-sink pair, has \phi = -\frac{\mu}{2\pi r} \cos \theta in polar coordinates, with \mu as doublet strength. A line vortex features circulatory motion with \phi = -\frac{\Gamma}{2\pi} \theta, where \Gamma is circulation. Superposition of uniform flow and a doublet yields flow around a circular cylinder: \phi = U \left( r + \frac{a^2}{r} \right) \cos \theta, where a is cylinder radius. Adding circulation from a vortex produces lift via the Kutta-Joukowski theorem, though drag remains zero. For a sphere, uniform flow plus a dipole gives \phi = U \cos \theta \left( r - \frac{a^3}{2 r^2} \right). This zero net drag on closed bodies in steady potential flow, first noted by d'Alembert in 1752, is known as d'Alembert's paradox, highlighting the idealization's limitation in predicting form drag. For compressible flows, the irrotational assumption leads to a nonlinear potential equation derived from the Euler equations, but linearized approximations enable analytical progress. In subsonic flow, the Prandtl-Glauert transformation scales coordinates by \beta = \sqrt{1 - M_\infty^2}, where M_\infty is free-stream , reducing the problem to incompressible form. For supersonic flow, the linearized equation becomes the wave equation \beta^2 \phi_{xx} + \phi_{yy} + \phi_{zz} = 0 with \beta = \sqrt{M_\infty^2 - 1}, solved using characteristics. These extensions, developed by Prandtl and Glauert in the , apply to slender bodies at moderate numbers.

Superfluids

Superfluids represent a class of quantum fluids that achieve true through macroscopic quantum mechanical effects, exhibiting zero below a critical . Liquid helium-4 (⁴He) transitions to the superfluid phase, known as He II, at the of approximately 2.17 K, where thermal excitations are minimized and the fluid behaves as a coherent . Similarly, liquid helium-3 (³He) becomes superfluid at much lower temperatures around 2.5 mK under specific pairing conditions, though ⁴He is the primary example due to its Bose-Einstein statistics facilitating . In these states, the superfluid component flows without dissipation, embodying the ideal inviscid assumptions of classical but governed by quantum principles. The behavior of superfluids is captured by the two-fluid model, proposed by in 1941, which decomposes the fluid into an inviscid superfluid component with ρ_s and v_s, and a viscous normal fluid component with ρ_n and v_n, where the total ρ = ρ_s + ρ_n. The superfluid fraction ρ_s vanishes above the critical temperature and approaches the total as temperature decreases toward , while the normal component carries and . This model explains phenomena like the absence of in the superfluid part, allowing it to permeate narrow channels without resistance, as observed in experiments with capillaries and porous media. A hallmark property is the frictionless flow manifested in Rollin films, thin superfluid layers (typically 100–300 nm thick) that creep along surfaces due to van der Waals forces, enabling helium to flow out of containers against without measurable loss. These films demonstrate superfluidity's , with transfer rates up to several cm/s in narrow gaps, far exceeding classical viscous limits. Another key feature is the formation of quantized vortices, where circulation around vortex lines is restricted to integer multiples of the quantum κ = h/m, with h as Planck's constant and m the mass (κ ≈ 9.97 × 10^{-4} cm²/s for ⁴He); this quantization arises from the single-valuedness of the wavefunction in the superfluid's order parameter. Superflow remains stable below a critical velocity v_c (typically 10–100 cm/s depending on ), but exceeds it leads to breakdown via vortex and , marking the onset of . Theoretically, Landau's phenomenological two-fluid hydrodynamics provides the foundation, with equations for mass and momentum conservation separating the components: ∂ρ/∂t + ∇·(ρ_s v_s + ρ_n v_n) = 0 and ρ_s (∂v_s/∂t + (v_s·∇)v_s) = -∇μ, where μ is the , highlighting the irrotational, potential-like flow of the superfluid. For weakly interacting Bose systems like dilute Bose-Einstein condensates (BECs) in ultracold atomic gases, the Gross-Pitaevskii equation iℏ ∂ψ/∂t = [-ℏ²/2m ∇² + V + g|ψ|²] ψ approximates the superfluid dynamics, treating the condensate wavefunction ψ as a macroscopic order parameter whose phase gradients yield the inviscid field v_s = (ℏ/m) ∇θ, with θ the phase. This mean-field approach captures vortex quantization and superflow stability, bridging quantum superfluids to classical inviscid models, though it neglects thermal fluctuations present in . Superfluidity was experimentally discovered in late 1937 through independent observations of helium's anomalous flow properties, with publications appearing in 1938: Pyotr Kapitza in reported zero below 2.2 K via measurements in narrow tubes, while John F. Allen and Donald Misener in observed frictionless film flow and fountain effects. These milestones, confirmed by subsequent and thermal conductivity experiments, established superfluids as real-world inviscid fluids. Applications leverage these properties in cryogenic cooling systems for maintaining low temperatures in particle accelerators and telescopes, and in precision measurements such as quantum gyroscopes using persistent currents around quantized vortices for rotation sensing with minimal drift. Recent experiments as of 2025 have further explored superfluid dynamics, including the collision of two superfluids revealing hybrid classical-quantum patterns, confirmation of molecular superfluidity in hydrogen clusters at nanoscale, and discovery of a universal law governing quantum vortex dynamics in superfluid helium-4.

References

  1. [1]
    [PDF] Governing Equations of Fluid Dynamics
    2.8.2 Equations for Inviscid Flow. Inviscid flow is, by definition, a flow where the dissipative, transport phenomena of viscosity, mass diffusion and ...
  2. [2]
    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.
  3. [3]
    [PDF] An Introduction to the Incompressible Euler Equations - UC Davis Math
    Sep 25, 2006 · This system models the flow of an inviscid, incompressible fluid with con- stant density. We have set the density equal to one without loss ...
  4. [4]
    [PDF] Incompressible, Inviscid, Irrotational Flow
    As described earlier, irrotational flow is defined as a flow in which the vorticity, ω, is zero and since ω = ∇ × u. (Bga1) it follows that the condition, ...
  5. [5]
    [PDF] Potential Flow Theory - MIT
    We can treat external flows around bodies as invicid (i.e. frictionless) and irrotational. (i.e. the fluid particles are not rotating).
  6. [6]
    Bernoulli's Equation
    In the derivation of the equation, the dissipation of energy by viscous forces was neglected (inviscid flow), and there was no introduction of energy allowed ( ...Missing: definition | Show results with:definition
  7. [7]
    Application of a quasi-3D inviscid flow and boundary layer analysis ...
    The objective was to improve the rotor aerodynamic characteristics by hub-shroud contouring. The approach relies on available 2D inviscid methods coupled with ...
  8. [8]
    Inviscid Flow - an overview | ScienceDirect Topics
    Theoretically, inviscid flow is approached in the limit as the Reynolds number goes to infinity. However, for practical problems, many flows with high but ...
  9. [9]
    Potential Flow Theory – Introduction to Aerospace Flight Vehicles
    History. The origins of the potential flow theory can be traced back to the 18th century, with contributions from Daniel Bernoulli and Leonhard Euler. Bernoulli ...
  10. [10]
    d'Alembert's Paradox - an overview | ScienceDirect Topics
    There is no drag on an arbitrary-cross-section object in steady two-dimensional, irrotational constant-density flow, a more general statement of d'Alembert's ...
  11. [11]
    [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 ...
  12. [12]
    "Principes généraux du mouvement des fluides" by Leonhard Euler
    Sep 25, 2018 · Euler proves that solutions of the equations of motion can exist even when the forces are such that equilibrium is impossible. He shows that the ...
  13. [13]
    The Euler equations of gas dynamics - Clawpack
    In this chapter we study the system of hyperbolic PDEs that governs the motions of a compressible gas in the absence of viscosity.
  14. [14]
    [PDF] Chapter 3 The Stress Tensor for a Fluid and the Navier Stokes ...
    A key step in formulating the equations of motion for a fluid requires specifying the stress tensor in terms of the properties of the flow, in particular ...
  15. [15]
    [PDF] CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION
    Apr 1, 2013 · nal energy due to dissipation of kinetic energy by viscous friction. ... THE ENERGY EQUATION FOR INVISCID, NON-HEAT CONDUCTING FLOW. With ...<|control11|><|separator|>
  16. [16]
    [PDF] Chapter 3 – Ideal Fluid Flow +
    – Setting ν = 0 the viscous term in the Navier-Stokes equations drops out and we obtain the Euler equations. The Euler equations are 1st order PDE's in space, ...
  17. [17]
    [PDF] SLMath - 2 The Navier-Stokes equations
    Unlike the Euler equations, the Navier-Stokes equations are semi-linear and at the highest order in deriva- tives in space and time, parabolic-elliptic (though ...
  18. [18]
    Ludwig Prandtl, “Motion of Fluids with Very Little Viscosity” (1904)
    In this article, engineer Ludwig Prandtl discusses the “motion of fluids with very little viscosity,” that is the flow of water or air.
  19. [19]
    [PDF] Prandtl's Boundary Layer Theory - UC Davis Math
    Prandtl's idea was that when the viscosity is small a thin boundary layer forms near the boundary in which the tangential velocity component ad j usts rapidly ...
  20. [20]
    [PDF] Ludwig Prandtl and Boundary Layers in Fluid Flow
    In 1904, Prandtl proposed the concept of bound- ary layers that revolutionised the study of fluid mechanics. In this article we present the basic ideas of ...
  21. [21]
    [PDF] 3 IRROTATIONAL FLOWS, aka POTENTIAL FLOWS - DAMTP
    Once again, we argue that the flow starts from rest, hence remains irrotational (in this inviscid fluid model), hence may be described by a velocity potential φ ...Missing: seminal | Show results with:seminal
  22. [22]
    Genesis of d'Alembert's paradox and analytical elaboration of ... - arXiv
    Jan 19, 2008 · After having at first refused the possibility of a vanishing drag, d'Alembert in 1768 established the paradox, but only for bodies with a head-tail symmetry.
  23. [23]
    [PDF] Compressible Potential Flow
    Compressible potential flow refers to a methodology used to solve steady, irrotational, compressible flows in which the direction of the velocity vector is ...Missing: linearized seminal
  24. [24]
    [PDF] sak12e5.tmp - NASA Technical Reports Server (NTRS)
    The Prandtl-Glauert method for subsonic potential flow of a compressible fluid has generally been believed to lead to an increase in the pressures over a ...Missing: seminal | Show results with:seminal
  25. [25]
    Theory of the Superfluidity of Helium II | Phys. Rev.
    Theory of the Superfluidity of Helium II, L. Landau Institute of Physical Problems, Academy of Sciences USSR, Moscow, USSR
  26. [26]
    Anderson's considerations on the flow of superfluid helium: Some ...
    Aug 17, 2015 · Superfluids display quantum properties over large distances. Superfluid currents may persist indefinitely unlike those of ordinary fluids (
  27. [27]
    Liquid Helium II: The Hydrodynamics of the Two-Fluid Model
    Exact non-linear equations of motion for reversible processes in the two-fluid model of liquid He II are obtained from a variational principle.Missing: superfluid | Show results with:superfluid
  28. [28]
    Flow of Superfluid Helium in a Porous Medium | Phys. Rev.
    Flow of superfluid liquid helium has been studied experimentally in a porous medium formed by packing fine powder into a hollow glass sphere.<|separator|>
  29. [29]
    Rollin Film Rates in Liquid Helium | Phys. Rev.
    It is found that the film transfer rate is somewhat greater than that observed for isothermal flow under a gravitational potential difference. Evidence is ...Missing: superfluid | Show results with:superfluid
  30. [30]
    Quantized vortices in superfluid | Rev. Mod. Phys.
    Jul 1, 1987 · Quantized vortices in superfluid 3He have unique phenomena, including coreless vortices with two flow quanta and half-integer circulation ...
  31. [31]
    Critical velocities in flows of superfluid 4 He - AIP Publishing
    Mar 21, 2025 · The superfluid streamlines are concentric circles, and the superfluid velocity is given by v s ( r ) = κ / ( 2 π r ) ⁠, where r is the distance ...