Fact-checked by Grok 2 weeks ago

Incompressible flow

Incompressible flow is a type of fluid motion in which the density of the fluid remains constant throughout the flow field, implying that the volume of any fluid element does not change as it moves. This approximation holds for liquids, where density variations are inherently small due to high bulk modulus, and for gases at low speeds where the Mach number is less than 0.3, resulting in density changes of less than 5% from pressure variations. The mathematical foundation of incompressible flow is provided by the continuity equation and the Navier-Stokes momentum equations under the assumption of constant density. The continuity equation, \nabla \cdot \mathbf{v} = 0, enforces mass conservation by ensuring that the velocity field is divergence-free, meaning no net flow into or out of any volume element. The incompressible Navier-Stokes equations, derived from Newton's second law applied to fluid elements, take the form \rho \left( \frac{\partial \mathbf{v}}{\partial t} + (\mathbf{v} \cdot \nabla) \mathbf{v} \right) = -\nabla P + \mu \nabla^2 \mathbf{v} + \rho \mathbf{g}, where \rho is the constant density, \mathbf{v} is the velocity vector, P is pressure, \mu is dynamic viscosity, and \mathbf{g} is the gravitational acceleration vector. These equations couple the transport of momentum with viscous effects and external forces, but the constant density simplifies the system compared to compressible flows by eliminating density as a variable. For inviscid cases (\mu = 0), the equations reduce to the Euler equations, \frac{\partial \mathbf{u}}{\partial t} + \mathbf{u} \cdot \nabla \mathbf{u} + \nabla p = 0 with \nabla \cdot \mathbf{u} = 0, highlighting the role of pressure in balancing inertial forces. Incompressible flow models are essential in engineering applications involving low-speed fluid dynamics, such as subsonic aerodynamics for aircraft design, hydrodynamics for ship hulls and pipelines, and environmental flows like river currents or ocean circulation. They also underpin biomedical simulations of blood flow in arteries and industrial processes like lubrication and cooling systems, where computational fluid dynamics (CFD) solvers enforce the divergence-free condition to predict pressure distributions and flow patterns accurately. Despite the idealization, these models capture dominant physics in regimes where compressibility effects are minimal, enabling efficient analysis and design optimization.

Definition and Fundamentals

Definition

Incompressible flow refers to the motion of a fluid in which the density remains constant throughout the flow field, independent of position and time. This assumption simplifies the analysis of fluid behavior, particularly for scenarios where variations in density are negligible, such as in the flow of liquids or gases at sufficiently low velocities. The concept of incompressible flow originated in the development of classical fluid mechanics during the 18th and 19th centuries. Leonhard Euler laid foundational work in 1755 by deriving equations governing the motion of inviscid, incompressible fluids, building on principles of conservation of momentum. Later, Claude-Louis Navier extended this framework in 1822 by incorporating viscosity, presenting the Navier-Stokes equations specifically for incompressible viscous fluids, which became central to modern fluid dynamics. Incompressible flow is applicable to physical situations where the flow speed is much less than the speed of sound, characterized by a low Mach number (Ma ≪ 1), ensuring that compressibility effects are minimal. This regime encompasses the flow of water through pipes, where density constancy arises from the fluid's inherent properties, and subsonic air flows around structures at low speeds. Another representative example is the circulation of blood in vessels, modeled as an incompressible fluid to analyze hemodynamic patterns without significant density changes.

Key Assumptions

The primary assumption underlying incompressible flow is that the fluid density remains constant (ρ = constant) throughout the flow field, implying negligible changes in volume due to pressure variations. This holds for liquids and certain gases where the bulk modulus is high relative to typical pressure differences, ensuring that density perturbations are much smaller than unity (δρ/ρ ≪ 1). Secondary assumptions include isothermal conditions, where temperature variations are minimal or constant, and negligible viscous heating, which disregards the dissipation of mechanical energy into heat that could otherwise alter density. Additionally, the flow speed must be much lower than the speed of sound, typically corresponding to a Mach number (Ma) less than 0.3, to prevent significant compressibility effects from pressure waves. These conditions justify simplifying the governing equations by treating density as uniform and independent of pressure or temperature gradients except in specific contexts like buoyancy. The incompressible assumption is invalid for high-speed flows, such as supersonic jets or those involving shock waves, where density changes become substantial and acoustic wave propagation cannot be neglected. For justification in buoyancy-driven flows, the Boussinesq approximation is employed, assuming small density variations confined to the gravitational term while maintaining constant reference density elsewhere, which is suitable for low-amplitude thermal perturbations. Similarly, low-Mach-number expansions provide a rigorous limit as Ma approaches zero, deriving the incompressible Navier-Stokes equations from compressible ones by imposing a divergence-free velocity constraint.

Mathematical Formulation

Derivation from Continuity Equation

The continuity equation in fluid dynamics expresses the conservation of mass for a fluid element and is fundamental to deriving the conditions for incompressible flow. In its general form for a compressible fluid, it is given by \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{v}) = 0, where \rho is the fluid density, t is time, \mathbf{v} is the velocity vector, and \nabla \cdot denotes the divergence operator. This equation arises from applying the principle of mass conservation to an infinitesimal control volume, balancing the rate of change of mass within the volume against the net mass flux across its surfaces. To derive the incompressible flow condition, assume the fluid density \rho is constant in both space and time, which is a key approximation for liquids or low-speed gas flows where density variations are negligible. Under this assumption, the partial derivative \frac{\partial \rho}{\partial t} = 0, and the density can be factored out of the divergence term since \rho is uniform. Substituting these into the general continuity equation yields \nabla \cdot \mathbf{v} = 0. This simplified form indicates that the velocity field is divergence-free, or solenoidal, meaning there is no net flux of volume through any closed surface in the flow. The step-by-step process begins with the material derivative of density: for incompressibility, \frac{D\rho}{Dt} = 0, which expands to \frac{\partial \rho}{\partial t} + (\mathbf{v} \cdot \nabla) \rho = 0. Combining this with the continuity equation eliminates density variations, directly leading to the divergence-free condition. The implication of \nabla \cdot \mathbf{v} = 0 is the conservation of volume for fluid elements, ensuring that the fluid does not expand or contract during motion, which aligns with the physical behavior of incompressible substances like water under typical engineering conditions. This constraint enforces that the flow preserves the volume of material parcels, facilitating simplified analyses in many hydrodynamic problems.

Integration into Navier-Stokes Equations

The full compressible Navier-Stokes equations describe the motion of fluids where density variations are significant, incorporating these changes in both the momentum and energy conservation terms. The momentum equation takes the form \rho \left( \frac{\partial \mathbf{U}}{\partial t} + (\mathbf{U} \cdot \nabla) \mathbf{U} \right) + \nabla p - \delta \Delta \mathbf{U} - (\eta + \frac{1}{3} \mu) \nabla (\nabla \cdot \mathbf{U}) = \mathbf{G}, where \rho is the variable density, \mathbf{U} is the velocity vector, p is pressure, \delta and \eta relate to viscosities, \mu is the dynamic viscosity, and \mathbf{G} represents external forces. The continuity equation \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{U}) = 0 enforces mass conservation with density fluctuations, while the energy equation \frac{\partial E}{\partial t} + \nabla \cdot ((E + p) \mathbf{U}) = \nabla \cdot (\boldsymbol{\tau} \cdot \mathbf{U} - \mathbf{q}), with E as total energy, \boldsymbol{\tau} the viscous stress tensor, and \mathbf{q} = -k \nabla T the heat flux (with k thermal conductivity), couples density to thermodynamic states via equations of state like p = (\gamma - 1)(E - \frac{1}{2} \rho \mathbf{U} \cdot \mathbf{U}) for ideal gases. Under the incompressible flow assumption, density \rho is treated as constant, leading to the simplified continuity equation \nabla \cdot \mathbf{v} = 0, which constrains the velocity field \mathbf{v} to be divergence-free. The momentum equation reduces to \rho \left( \frac{\partial \mathbf{v}}{\partial t} + (\mathbf{v} \cdot \nabla) \mathbf{v} \right) = -\nabla p + \mu \nabla^2 \mathbf{v} + \rho \mathbf{g}, where the left side uses the material derivative \frac{D\mathbf{v}}{Dt} = \frac{\partial \mathbf{v}}{\partial t} + (\mathbf{v} \cdot \nabla) \mathbf{v} to represent acceleration, p is pressure, \mu is constant viscosity, and \mathbf{g} is gravitational acceleration. This form arises from applying Newton's second law to an incompressible Newtonian fluid, assuming constant viscosity and neglecting second viscosity effects. In this framework, pressure p functions as a Lagrange multiplier to enforce the incompressibility constraint \nabla \cdot \mathbf{v} = 0. Mathematically, the weak formulation of the equations seeks \mathbf{v} \in V (divergence-free space) and p \in L^2_0 such that (\nabla \mathbf{v}, \nabla \mathbf{w}) - (p, \nabla \cdot \mathbf{w}) = (\mathbf{f}, \mathbf{w}) for test functions \mathbf{w}, where the pressure term projects onto the constraint gradient, ensuring the velocity minimizes the energy functional subject to \text{div} \mathbf{v} = 0. This interpretation holds in both steady Stokes and full Navier-Stokes limits for incompressible flows. The incompressible assumption introduces key simplifications by decoupling density from pressure and temperature, thereby eliminating the need for an energy equation and associated equations of state. This reduces the system to four equations (continuity and three momentum components) for four unknowns (velocity components and pressure), compared to the five or more in compressible cases, significantly lowering computational complexity in simulations. The pressure is then recovered via a Poisson equation derived from the divergence of the momentum equation, further streamlining numerical treatment without density-energy coupling.

Physical Properties and Relations

Relation to Compressibility

Compressibility in fluid flow refers to the change in fluid density in response to variations in pressure, which is fundamentally tied to the speed of sound c = \sqrt{\gamma R T}, where \gamma is the specific heat ratio, R is the gas constant, and T is the absolute temperature. In compressible flows, these density changes become significant when the flow velocity approaches or exceeds the local speed of sound, altering the flow behavior compared to incompressible approximations. The incompressible flow approximation holds when the flow speed v is much less than the speed of sound c (typically v \ll c, or Mach number \mathrm{Ma} < 0.3), resulting in nearly constant density and isobaric processes where pressure variations do not substantially affect density. This low Mach number regime aligns with the key assumptions of incompressible flow, allowing simplification of the governing equations by neglecting density variations. In contrast, compressibility effects dominate at higher Mach numbers (\mathrm{Ma} > 0.3), leading to phenomena such as shock waves, where abrupt density, pressure, and velocity changes occur across a discontinuity, and expansion fans, which are continuous regions of rarefaction in supersonic flows. These features are prominent in high-speed aerodynamics, such as transonic flows around aircraft where local acceleration over wings can trigger shock formation. A practical example of the transition is airflow over aircraft wings: at low speeds (e.g., takeoff, \mathrm{Ma} \approx 0.2), the flow is treated as incompressible with minimal density change, yielding predictable lift via the circulation theorem; however, at high speeds (e.g., cruise near \mathrm{Ma} = 0.8), compressibility induces shock waves on the upper surface, increasing drag and potentially causing buffet or loss of control. This shift necessitates compressible flow analysis to accurately predict performance and structural loads in aviation design.

Connection to Solenoidal Fields

In incompressible flow, the velocity field \mathbf{v} satisfies the condition \nabla \cdot \mathbf{v} = 0, rendering it a solenoidal vector field. This divergence-free property ensures that the net flux through any closed surface is zero, consistent with the absence of volume changes in the fluid. Solenoidal vector fields draw an analogy to in , where \nabla \cdot \mathbf{B} = 0 due to the nonexistence of magnetic monopoles. Geometrically, this implies no sources or sinks in the : field lines must either close upon themselves or extend infinitely, maintaining a constant "density" of lines without creation or . The Helmholtz decomposition provides a foundational for understanding this , stating that any sufficiently in a can be uniquely expressed as the sum of a solenoidal (divergence-free) component and an irrotational (curl-free) component, conditions. In fluid dynamics, the solenoidal part captures the incompressible nature of the velocity, separating it from any compressible or rotational effects. This solenoidal character finds direct application in potential flow theory, where the incompressible velocity field is both divergence-free and irrotational, enabling its representation as the gradient of a scalar potential \phi such that \mathbf{v} = \nabla \phi.

Distinction from Material Incompressibility

Material incompressibility refers to an intrinsic constitutive property of a fluid where its density remains constant regardless of changes in pressure or temperature, corresponding to an infinite bulk modulus that prevents any volume change under applied stress. This idealization is well-approximated by liquids such as water, which exhibit extremely low compressibility due to their high bulk modulus, on the order of 2.2 × 10^9 Pa, resulting in negligible density variations even under significant pressure changes. In contrast, flow incompressibility is a kinematic approximation applied to the flow field, where the material derivative of density is zero (Dρ/Dt = 0), implying that fluid elements do not change volume as they move through the domain, even if the fluid itself is slightly compressible. This condition holds when density variations are small enough not to significantly influence the flow dynamics, typically in scenarios with low Mach numbers (M < 0.3). The key distinction lies in their scope: material incompressibility is an absolute property of the fluid medium, demanding uniform constant density (∇ρ = 0 and ∂ρ/∂t = 0), whereas flow incompressibility is a flow-specific assumption that permits minor spatial density gradients as long as they do not affect the Lagrangian evolution of individual fluid parcels. Consequently, a compressible fluid like air can exhibit incompressible flow under low-speed conditions, such as in aerodynamics below 100 m/s, while an incompressible fluid like water inherently produces incompressible flow. This separation allows the incompressible flow model to be applied more broadly, beyond strictly incompressible materials, for computational and analytical simplicity in engineering contexts.

Associated Flow Constraints

Inviscid and Irrotational Flows

In the context of incompressible flow, the inviscid assumption neglects viscous effects by setting the dynamic viscosity μ to zero, simplifying the governing equations to the Euler equations. These equations describe the motion of an ideal fluid with constant density ρ, combining the continuity equation ∇ · v = 0 with the momentum balance. The resulting momentum equation takes the form \frac{D\mathbf{v}}{Dt} = -\frac{1}{\rho} \nabla p + \mathbf{g}, where v is the velocity vector, p is the pressure, and g is the body force per unit mass, typically gravity. This formulation captures the acceleration of fluid particles due to pressure gradients and external forces without dissipative terms. A further simplification arises under the irrotational flow assumption, where the vorticity vanishes, satisfying ∇ × v = 0 throughout the domain. Helmholtz's theorems establish that in inviscid, barotropic flows, vortex lines are transported with the fluid, enabling irrotational conditions for flows without initial vorticity or boundaries that introduce it. Consequently, the velocity field can be represented using a scalar potential φ such that v = ∇φ, reducing the problem to solving Laplace's equation ∇²φ = 0 for incompressible conditions. This potential flow approach is widely applied in external aerodynamics, such as modeling airflow over aircraft wings or slender bodies, where rotational effects are minimal away from surfaces. The divergence-free nature of incompressible velocity fields supports such irrotational representations when boundary conditions permit a single-valued potential. For steady, incompressible, inviscid, and irrotational flows, the Euler equations integrate to Bernoulli's equation along streamlines. Starting from the steady Euler equation ∇ · (v v) = -(1/ρ) ∇p + g and assuming irrotationality (ω = 0), dotting with v yields v · ∇(v²/2) = -(1/ρ) v · ∇p + v · g. Integrating along a streamline, with conservative gravity g = -∇(gh) and constant ρ, results in p + \frac{1}{2} \rho v^2 + \rho g h = \constant. This relation equates pressure energy, kinetic energy per unit volume, and gravitational potential energy. In fully irrotational flows, the constant holds uniformly across the entire domain, not just streamlines, enhancing its utility for global flow analysis. The principle originates from energy conservation considerations in Bernoulli's 1738 work, with the inviscid form refined by Euler. Potential flow models under these assumptions reveal key limitations, exemplified by D'Alembert's paradox. In steady, incompressible, inviscid, irrotational flow past a closed body moving at constant velocity, the theory predicts zero net drag force, as pressure forces symmetrize fore and aft. This counterintuitive result, first derived by d'Alembert in 1752 while analyzing fluid resistance, arises because the ideal fluid slips tangentially without skin friction or flow separation. The paradox underscores the necessity of viscosity for realistic drag prediction in aerodynamics and hydrodynamics.

Steady and Unsteady Constraints

In steady incompressible flow, the partial derivative with respect to time vanishes (∂/∂t = 0), simplifying the governing equations to the time-independent divergence-free condition ∇ · v = 0 from the continuity equation and the steady Navier-Stokes equations for momentum conservation. This constraint implies that the velocity field does not vary temporally, allowing for solutions where flow properties remain constant at fixed points in space, which is particularly useful for analyzing persistent configurations like uniform channel flows. For unsteady incompressible flows, time derivatives are retained in the momentum equation, such as the ∂v/∂t term, while the continuity equation continues to enforce ∇ · v = 0 instantaneously at every time step. This maintains the solenoidal nature of the velocity field despite temporal variations, often arising from external forcings or boundary changes. An additional constraint in unsteady cases is the negligible role of acoustic waves, as the incompressible approximation assumes an infinite speed of sound, filtering out pressure perturbations that would propagate as compressible disturbances. Representative examples include oscillating flows in hydraulic pumps, where small-amplitude perturbations around steady operation are modeled under incompressible assumptions for frequencies below about 0.02 times the shaft rotation rate. However, these constraints have limitations in unsteady high-speed flows, where phenomena like cavitation can violate incompressibility by introducing local density variations through phase changes and bubble dynamics. In such scenarios, the steady-state incompressible solution breaks down when cavitation parameters reach regimes where density fluctuations become significant, necessitating compressible models.

Numerical and Computational Approaches

Approximation Techniques

Perturbation methods provide a framework for approximating incompressible flows by treating compressibility effects as small perturbations around a low Mach number regime. In this approach, the compressible Navier-Stokes equations are expanded asymptotically in powers of the Mach number M, where M = u / c with u the flow speed and c the sound speed. The leading-order terms yield the incompressible limit, with density variations appearing at higher orders, such as O(M) for density perturbations and O(M^2) for pressure fluctuations relative to the background pressure. This expansion is particularly useful for flows where acoustic waves are fast compared to convective timescales, allowing the elimination of stiff pressure terms while retaining essential physics. The low-Mach number expansion begins by nondimensionalizing the equations and assuming M \ll 1, leading to a decomposition of pressure into a thermodynamic part (uniform to leading order) and a dynamic part (governing the flow). Seminal derivations employ singular perturbation techniques, ensuring the solution remains well-posed in the incompressible limit without initial acoustic disturbances. For instance, the method derives modified equations like the Heat-Fluctuation-Dominated Hydrodynamics model, valid for subsonic homogeneous turbulence. These approximations are foundational for analyzing weakly compressible effects in otherwise incompressible regimes, such as in astrophysical or geophysical flows. Boundary layer theory, introduced by Ludwig Prandtl, offers a semi-analytical approximation for incompressible viscous flows at high Reynolds numbers, where Re = UL / \nu \gg 1 with U the free-stream velocity, L a characteristic length, and \nu the kinematic viscosity. The core idea is to separate the flow into a thin inner viscous layer near solid boundaries, where friction effects are confined, and an outer inviscid region treated by potential flow theory. This reduces the full three-dimensional Navier-Stokes equations to a more tractable two-dimensional parabolic system in the boundary layer coordinates, enabling marching solutions downstream from the leading edge. Prandtl's 1904 formulation assumes negligible viscosity outside the layer, with the layer thickness scaling as \delta \sim L / \sqrt{Re}, justifying the neglect of streamwise diffusion. The boundary layer equations for incompressible flow are: \frac{\partial u}{\partial x} + \frac{\partial v}{\partial y} = 0, u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = -\frac{1}{\rho} \frac{\partial p}{\partial x} + \nu \frac{\partial^2 u}{\partial y^2}, with pressure p imposed from the outer inviscid solution and no-slip conditions at the wall. This approximation revolutionized drag prediction in aerodynamics by quantifying skin friction while avoiding the full viscous computation. It holds for attached flows but breaks down at separation points, where adverse pressure gradients cause reversal. Similarity solutions exploit self-similar profiles to simplify boundary layer equations for specific geometries, reducing partial differential equations to ordinary ones. A canonical example is the Blasius solution for laminar flow over a semi-infinite flat plate in incompressible conditions, assuming zero pressure gradient and uniform free-stream velocity U. By introducing the similarity variable \eta = y \sqrt{U / (\nu x)} and stream function \psi = \sqrt{\nu U x} \, f(\eta), the equations collapse to the nonlinear ODE: f''' + \frac{1}{2} f f'' = 0, with boundary conditions f(0) = f'(0) = 0 and f'(\infty) = 1. This third-order equation, solved numerically via shooting methods, yields the velocity profile u / U = f'(\eta), revealing a boundary layer thickness \delta \approx 5 \sqrt{\nu x / U}. The solution, first obtained by Heinrich Blasius in 1908, provides exact skin friction \tau_w = 0.332 \rho U^2 / \sqrt{Re_x} and underpins similarity analyses for wedges or stagnation points. It is valid for high Re_x far from the leading edge, where end effects are negligible. Error analysis in these approximations quantifies their range of validity, particularly for perturbation methods where compressibility introduces inaccuracies scaling with Mach number. In low-Mach expansions, pressure predictions deviate from incompressibility by O(M^2), as dynamic pressure perturbations are quadratic in velocity while thermodynamic pressure is nearly constant; this ensures accurate velocity fields for M < 0.3, but acoustic modes may require higher-order corrections for O(M) density errors. For boundary layer theory, errors arise from neglected streamwise diffusion, bounded by O(1 / Re) in the inner layer, with outer flow assumptions failing near separation (e.g., when \partial p / \partial x > 0). Similarity solutions like Blasius inherit these, plus numerical integration errors typically below 0.1% for the wall shear. Overall, these techniques maintain fidelity for Re > 10^3 and M \ll 1, with breakdowns signaling the need for full compressible or turbulent models.

Simulation Methods

Simulation methods for incompressible flows primarily address the challenge of enforcing the divergence-free velocity condition (∇·v = 0) while solving the coupled Navier-Stokes equations numerically. These techniques are essential for time-dependent simulations where direct enforcement of incompressibility leads to an elliptic pressure Poisson equation, complicating iterative solutions. Projection methods, finite volume and finite difference discretizations, artificial compressibility approaches, and modern mesoscopic methods like lattice Boltzmann form the core of these strategies, each balancing accuracy, stability, and computational efficiency for various flow regimes. Projection methods decouple the velocity and pressure computations to satisfy the incompressibility constraint efficiently. In Chorin's original algorithm, developed in 1968, an intermediate velocity field \mathbf{u}^* is first obtained by advancing the momentum equation without the pressure gradient term over a time step \Delta t: \frac{\mathbf{u}^* - \mathbf{u}^n}{\Delta t} = -\nabla \cdot (\mathbf{u}^n \otimes \mathbf{u}^n) + \nu \nabla^2 \mathbf{u}^n, where \mathbf{u}^n is the velocity at time n\Delta t and \nu is the kinematic viscosity. This intermediate field is then projected onto the divergence-free space by solving a Poisson equation for a potential \phi: \nabla^2 \phi = \frac{\nabla \cdot \mathbf{u}^*}{\Delta t}, followed by the correction \mathbf{u}^{n+1} = \mathbf{u}^* - \nabla \phi and pressure update p^{n+1} = p^n + \phi. This fractional-step approach enforces \nabla \cdot \mathbf{u}^{n+1} = 0 exactly in discrete form, making it first-order accurate in time but widely adopted for its simplicity and robustness in viscous incompressible simulations. Subsequent improvements, such as second-order variants, have enhanced temporal accuracy while retaining the projection framework. Finite volume and finite difference methods on staggered grids mitigate issues in velocity-pressure coupling, particularly the odd-even decoupling or checkerboard oscillations that plague colocated grids. Introduced by Harlow and Welch in 1965, the staggered grid arrangement places velocity components at cell faces and scalar fields like pressure at cell centers, ensuring that pressure gradients act directly on velocities without interpolation artifacts. This setup naturally conserves momentum and mass in incompressible flows, as demonstrated in simulations of time-dependent viscous flows with free surfaces. For instance, in a finite volume discretization, the momentum flux across faces incorporates the staggered velocities, while the continuity equation integrates over volumes to yield a pressure correction equation, avoiding non-physical pressure modes. These methods are particularly effective for structured grids in engineering applications like aerodynamics and heat transfer. Artificial compressibility methods transform the incompressible system into a hyperbolic one by introducing a pseudo-time derivative for pressure, allowing steady-state solutions via time-marching. Chorin proposed this approach in 1967, modifying the continuity equation to: \frac{\partial p}{\partial \tau} + \beta^2 \nabla \cdot \mathbf{v} = 0, where \tau is artificial time and \beta is an artificial compressibility parameter tuned to the flow speed (typically \beta \approx U_{\max}, the maximum velocity). Coupled with the momentum equations, the system is solved iteratively until \partial p / \partial \tau \to 0, recovering the divergence-free condition. This method converges faster than purely elliptic solvers for steady flows and extends to unsteady cases with dual time-stepping, though it requires careful choice of \beta to minimize acoustic wave artifacts. It has been extensively validated for low-Mach-number flows in complex geometries. Modern approaches like the lattice Boltzmann method (LBM) offer advantages for incompressible flows in the low-Mach-number limit, especially with complex boundaries. Derived from kinetic theory, LBM simulates fluid particles on a discrete lattice via collision and streaming steps, recovering the Navier-Stokes equations through Chapman-Enskog expansion. For incompressible regimes, models such as the double-distribution function or incompressible LBM variants enforce density constancy while handling viscosity via relaxation parameters. A key scheme by He, Chen, and Zhang (1999) adapts LBM for multiphase incompressible flows but applies to single-phase cases by recovering the pressure from a non-equilibrium distribution tensor, enabling accurate simulations of Rayleigh-Taylor instability with errors below 1% in growth rates. LBM excels in parallelization and boundary treatments, such as immersed boundary methods for irregular geometries, making it suitable for microfluidics and porous media flows.

References

  1. [1]
    Types of Fluid Flows – Introduction to Aerospace Flight Vehicles
    For an incompressible flow, any temperature changes are generally small. However, in a compressible flow, significant temperature changes may occur. Check Your ...
  2. [2]
    [PDF] Outline Mach Number - Sites
    • Speed of Sound & Mach Number. ♢ Compressible ... • Incompressible Flow, M<0.3. • Subsonic Flow, 0.3<M<0.8. • Transonic Flow, 0.8<M< 1.2. • Supersonic Flow, 1.2 ...
  3. [3]
    [PDF] THE NAVIER-STOKES EQUATION In this lesson, we will
    The result is the famous Navier-Stokes equation, shown here for incompressible flow. To solve fluid flow problems, we need both the continuity equation and the ...
  4. [4]
    [PDF] An Introduction to the Incompressible Euler Equations - UC Davis Math
    Sep 25, 2006 · The incompressible Euler equations model the flow of an inviscid, incompressible fluid, with the equations: ut + u · ∇u + ∇p = 0, ∇ · u = 0.
  5. [5]
    [PDF] Computational Challenges of Viscous Incompressible Flows
    Incompressible flow is an approximation of flow where flow speed is insignificant everywhere compared to the speed of sound of the medium. If incomprcssible ...
  6. [6]
    [PDF] Chapter 13: Foundations of Fluid Dynamics [version 1013.1.K]
    As we shall shortly see, approximating a flow as incompressible is a good approximation when the flow speed is much less than the speed of sound and the fluid ...
  7. [7]
    [PDF] Chapter 12 Fluid Mechanics 1 Density
    Fluid flow is incompressible if the density ρ is a constant, independent of x, y, z, and t. 3. Fluid flow can be viscous or nonviscous. Viscosity in fluid ...
  8. [8]
    [PDF] 1 Short History
    Navier, in 1822, in a memoir (report) for the French. Academy, presented the following equations for incompressible fluids. ∂u. ∂x. +. ∂v. ∂y. +. ∂w. ∂z. = 0. 1 ...
  9. [9]
    C:\SAVE ME\AOE3104\preliminaries.wpd
    Flows at low Mach numbers can be treated as incompressible, while those at high. Mach numbers must be treated as compressible. What we mean by high or low is ...
  10. [10]
    [PDF] Chapter 6 Fluid Mechanics - MIT OpenCourseWare
    Example for a horizontal pipe filled with an incompressible fluid (which is approximately true of water at room temperature) of constant density ρ. Lets ...
  11. [11]
    Incompressible Flow - an overview | ScienceDirect Topics
    Incompressible flow is where fluid density change due to pressure is negligible, typically when Mach number is less than 0.3, or where density is constant ...
  12. [12]
    The Incompressibility Assumption | CFD-101 - FLOW-3D
    Thus, the assumption of incompressibility requires that fluid speed be small compared to the speed of sound, (1)Missing: dynamics | Show results with:dynamics
  13. [13]
    [PDF] LESSON 04H: THE BOUSSINESQ APPROXIMATION JM Cimbala
    But the viscous dissipation term φ is usually negligibly small in the Boussinesq approximation, just as for incompressible flow.
  14. [14]
    The mathematical theory of low Mach number flows
    Incompressible fluid flow differs from compressible fluid flow in that one of the equations of evolution is replaced by the constraint that the flow be ...
  15. [15]
    [PDF] Generalizing the Boussinesq Approximation to Stratified ...
    May 14, 2007 · The simplifications required to apply the Boussinesq approximation to compressible flow are compared with those in an incompressible fluid.
  16. [16]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics 5.2 ...
    Dec 15, 2021 · Thus, an incompressible fluid has the same Continuity Equation as a fluid with constant and uniform density. (3) Another useful form of the ...
  17. [17]
    [PDF] INTRODUCTORY LECTURES ON FLUID DYNAMICS
    We say that any incompressible flow satisfying the continuity equation ∇ · u =0isa kinematically possible motion. The Navier-Stokes equation plus ...
  18. [18]
    [PDF] Introduction to Compressible Computational Fluid Dynamics
    The Navier-Stokes equations are commonly expressed in one of two forms. One form is known as the incompressible flow equations and the other is known as the.
  19. [19]
    [PDF] Chapter 30 Navier-Stokes Equations - MIT OpenCourseWare
    29.6.3 Irrotational Incompressible Steady Flow: Bernoulli Equation ............... 13. 29.10 Navier-Stokes Equations for an Incompressible Newtonian Fluid .....
  20. [20]
    [PDF] arXiv:1703.08216v1 [math.AP] 23 Mar 2017
    Mar 23, 2017 · For the equations modelling incompressible fluid flows it is frequently remarked that the pressure term acts as a Lagrange multiplier enforcing ...
  21. [21]
    [PDF] Solution methods for the Incompressible Navier-Stokes Equations
    Solution methods include pressure-based and density-based approaches, pressure correction schemes, and explicit/implicit schemes. Pressure is derived from mass ...
  22. [22]
    Isentropic Flow Equations
    However, as the speed of the flow approaches the speed of sound we must consider compressibility effects on the gas. ... a = sqrt(gam * p / r) = sqrt (gam * R * T).Missing: c = | Show results with:c =
  23. [23]
    [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.
  24. [24]
    Compressible Flow vs Incompressible Flow in Fluid Mechanics
    Aug 11, 2023 · In compressible flow, fluid density is not constant, while in incompressible flow, density remains constant. Mach number and density change are ...Difference between... · Compressibility and Mach... · Continuity EquationMissing: key | Show results with:key
  25. [25]
    Incompressible vs. Compressible Flow - CFD Land
    For incompressible flows, especially at low Mach numbers (Mach < 0.3), density is constant in an incompressible flow, as variations in pressure and temperature ...
  26. [26]
    The Lift Coefficient
    For very low speeds (< 200 mph) the compressibility effects are negligible. At higher speeds, it becomes important to match Mach numbers between the two cases.
  27. [27]
    Compressible Flow - an overview | ScienceDirect Topics
    Shock waves do not appear in the flow. In engineering practice, subsonic flows for which M < 0.3 are often treated as being incompressible. (iii). Transonic ...
  28. [28]
    Centered Expansion Fan - NASA Glenn Research Center
    Across a shock wave, the Mach number decreases, the static pressure increases, and there is a loss of total pressure because the process is irreversible.
  29. [29]
    [PDF] Theoretical Study of Mach number and Compressibility Effect on the ...
    If the aircraft passes at a low speed, (< 250 mph) the density of air remains constant. But for the higher speed of an aircraft, some of the energy of the ...
  30. [30]
    Critical Mach Number - an overview | ScienceDirect Topics
    We need to discuss some aspects of the effect of compressibility on the drag of aircraft. As the Mach number is increased from a low value the speed of the air ...<|control11|><|separator|>
  31. [31]
    Incompressible flow - Galileo
    Aug 11, 1997 · Incompressible flow. ... This is often expressed in terms of the Mach number Ma, defined as.
  32. [32]
    None
    ### Definition of Solenoidal Field in Context of Incompressible Flow
  33. [33]
    Divergence - Richard Fitzpatrick
    Such a field is called a solenoidal vector field. The simplest example of a solenoidal vector field is one in which the lines of force all form closed loops.
  34. [34]
    [PDF] The Helmholtz-Hodge Decomposition—A Survey
    Abstract—The Helmholtz-Hodge Decomposition (HHD) describes the decomposition of a flow field into its divergence-free and curl- free components.
  35. [35]
    Helmholtz's decomposition for compressible flows and its ... - NIH
    Helmholtz's decomposition does not compute fluid dynamics but separates the compressible and vortical effects. Hence, it is self-evident that this ...
  36. [36]
    [PDF] Fluids - Duke Physics
    (In effect, it has infinite bulk modulus.) In this case we can easily integrate the above equation to find an important formula: Incompressible fluid at rest.<|control11|><|separator|>
  37. [37]
    [PDF] Fluids University of Oxford Third Year, Part B1
    we need to distinguish between an incompressible fluid and an incompressible flow as, under some circumstances, air in motion for example can be ...
  38. [38]
    [PDF] CHAPTER 5 Continuity & Momentum Equations
    important to distinguish an incompressible fluid from an incompressible flow, which is defined by dρ/dt = 0. The continuity equation shows that an.
  39. [39]
    Derivation of the Euler equation of motion (conservation of momentum)
    Aug 10, 2020 · The Euler equation of motion describes inviscid, unsteady flows of compressible or incompressible fluids. Pressure forces on a fluid element.
  40. [40]
    [0804.4802] Principles of the motion of fluids - arXiv
    Apr 30, 2008 · Principles of the motion of fluids. Authors:Leonhard Euler. View a PDF of the paper titled Principles of the motion of fluids, by Leonhard Euler.Missing: 1757 | Show results with:1757
  41. [41]
    [PDF] Helmholtz (1858)
    Helmholtz's paper on vortex motions was first published in 1858 in the Journal fur die reine and angewandte Mathematik*. It was the first in a series of ground ...Missing: citation | Show results with:citation
  42. [42]
    [PDF] VORTEX DYNAMICS: THE LEGACY OF HELMHOLTZ AND KELVIN
    (1867) Translation of (Helmholtz 1858): On the integrals of the hy- drodynamical equations, which express vortex-motion, Phil.Mag., 33, 485–512. 18. Tait ...
  43. [43]
    [PDF] Bernoulli's Equation
    This is Bernoulli's equation for steady, incompressible, inviscid, irrotational flow with conservative body forces due to gravity. It is one of the most ...Missing: original source
  44. [44]
    Genesis of d'Alembert's paradox and analytical elaboration of the ...
    Aug 15, 2008 · The first hint of d'Alembert's paradox–the vanishing of the drag for a solid body surrounded by a steadily moving ideal incompressible fluid ...
  45. [45]
    [PDF] A physically motivated approach for filtering acoustic waves from the ...
    acoustic waves from the equations governing compressible stratified flow ... Thus, in adiabatic incompressible flow,. Dρ. Dt. = 0. (1.1). Here D/Dt denotes ...
  46. [46]
    Chapter 9 - Hydrodynamics of Pumps - Christopher E. Brennen
    This chapter is devoted to a description of the methods available for the analysis of unsteady flows in pumps and their associated hydraulic systems.
  47. [47]
  48. [48]
    [PDF] The equations of nearly incompressible fluids. I. Hydrodynamics ...
    nearly incompressible fluid equations as applied to the equa- tions of ... that leads to incompressible flow. The term '“nearly incom- pressible” is ...Missing: distinction | Show results with:distinction
  49. [49]
    [PDF] A low Mach correction able to deal with low Mach acoustics - Hal-Inria
    Abstract. This article deals with acoustic computations in low Mach number flows with density based solvers. For ensuring a good resolution of the low Mach ...Missing: Ma²) | Show results with:Ma²)
  50. [50]
    None
    ### Summary of Prandtl's 1904 Boundary Layer Theory
  51. [51]
    [PDF] Blasius Solution for a Flat Plate Boundary Layer
    The first exact solution to the laminar boundary layer equations, discovered by Blasius (1908), was for a simple constant value of U(s) and pertains to the ...
  52. [52]
    [PDF] Numerical solution of the Navier-Stokes equations - UC Berkeley math
    A. J. CHORIN, "The numerical solution of the Navier-Stokes equations for incompressible fluid," AEC Research and Development Report No. NYO-1480-82, New ...
  53. [53]
    [PDF] Numerical calculation of time-dependent viscous incompressible ...
    A new technique is described for the numerical investigation of the time-dependent flow of an incompressible fluid, the boundary of which is partially confined ...Missing: staggered | Show results with:staggered
  54. [54]
    Comparison of finite-volume numerical methods with staggered and ...
    The paper presents a detailed comparison of two finite-volume solution methods for two-dimensional incompressible fluid flows, one with staggered and the other ...
  55. [55]
    [PDF] A numerical method for solving incompressible viscous flow problems
    (2). THE METHOD OF ARTIFICIAL COMPRESSIBILITY. We introduce the auxiliary system of equations ə¸u¡ + Rə،(µ؟u¡) = -dip + Au; + Fi,. + apaju = 0, p = p/8. ρίδ ...
  56. [56]
    [PDF] Comparison of Artificial Compressibility Methods
    Various artificial compressibility methods for calculating the three-dimensional incompressible Navier-Stokes equations are compared. Each method is de-.
  57. [57]
    A Lattice Boltzmann Scheme for Incompressible Multiphase Flow ...
    In this paper, we propose a new lattice Boltzmann scheme for simulation of multiphase flow in the nearly incompressible limit.
  58. [58]
    LATTICE BOLTZMANN METHOD FOR FLUID FLOWS
    ABSTRACT. We present an overview of the lattice Boltzmann method (LBM), a parallel and efficient algorithm for simulating single-phase and multiphase fluid ...