fbpx
Wikipedia

Navier–Stokes equations

The Navier–Stokes equations (/nævˈj stks/ nav-YAY STOHKS) are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842-1850 (Stokes).

The Navier–Stokes equations mathematically express momentum balance and conservation of mass for Newtonian fluids. They are sometimes accompanied by an equation of state relating pressure, temperature and density.[1] They arise from applying Isaac Newton's second law to fluid motion, together with the assumption that the stress in the fluid is the sum of a diffusing viscous term (proportional to the gradient of velocity) and a pressure term—hence describing viscous flow. The difference between them and the closely related Euler equations is that Navier–Stokes equations take viscosity into account while the Euler equations model only inviscid flow. As a result, the Navier–Stokes are a parabolic equation and therefore have better analytic properties, at the expense of having less mathematical structure (e.g. they are never completely integrable).

The Navier–Stokes equations are useful because they describe the physics of many phenomena of scientific and engineering interest. They may be used to model the weather, ocean currents, water flow in a pipe and air flow around a wing. The Navier–Stokes equations, in their full and simplified forms, help with the design of aircraft and cars, the study of blood flow, the design of power stations, the analysis of pollution, and many other things. Coupled with Maxwell's equations, they can be used to model and study magnetohydrodynamics.

The Navier–Stokes equations are also of great interest in a purely mathematical sense. Despite their wide range of practical uses, it has not yet been proven whether smooth solutions always exist in three dimensions—i.e., whether they are infinitely differentiable (or even just bounded) at all points in the domain. This is called the Navier–Stokes existence and smoothness problem. The Clay Mathematics Institute has called this one of the seven most important open problems in mathematics and has offered a US$1 million prize for a solution or a counterexample.[2][3]

Flow velocity

The solution of the equations is a flow velocity. It is a vector field—to every point in a fluid, at any moment in a time interval, it gives a vector whose direction and magnitude are those of the velocity of the fluid at that point in space and at that moment in time. It is usually studied in three spatial dimensions and one time dimension, although two (spatial) dimensional and steady-state cases are often used as models, and higher-dimensional analogues are studied in both pure and applied mathematics. Once the velocity field is calculated, other quantities of interest such as pressure or temperature may be found using dynamical equations and relations. This is different from what one normally sees in classical mechanics, where solutions are typically trajectories of position of a particle or deflection of a continuum. Studying velocity instead of position makes more sense for a fluid, although for visualization purposes one can compute various trajectories. In particular, the streamlines of a vector field, interpreted as flow velocity, are the paths along which a massless fluid particle would travel. These paths are the integral curves whose derivative at each point is equal to the vector field, and they can represent visually the behavior of the vector field at a point in time.

General continuum equations

The Navier–Stokes momentum equation can be derived as a particular form of the Cauchy momentum equation, whose general convective form is

 
By setting the Cauchy stress tensor   to be the sum of a viscosity term   (the deviatoric stress) and a pressure term   (volumetric stress), we arrive at
Cauchy momentum equation (convective form)

 

where

In this form, it is apparent that in the assumption of an inviscid fluid – no deviatoric stress – Cauchy equations reduce to the Euler equations.

Assuming conservation of mass we can use the mass continuity equation (or simply continuity equation),

 
to arrive at the conservation form of the equations of motion. This is often written:[4]
Cauchy momentum equation (conservation form)

 

where   is the outer product:

 

The left side of the equation describes acceleration, and may be composed of time-dependent and convective components (also the effects of non-inertial coordinates if present). The right side of the equation is in effect a summation of hydrostatic effects, the divergence of deviatoric stress and body forces (such as gravity).

All non-relativistic balance equations, such as the Navier–Stokes equations, can be derived by beginning with the Cauchy equations and specifying the stress tensor through a constitutive relation. By expressing the deviatoric (shear) stress tensor in terms of viscosity and the fluid velocity gradient, and assuming constant viscosity, the above Cauchy equations will lead to the Navier–Stokes equations below.

Convective acceleration

 
An example of convection. Though the flow may be steady (time-independent), the fluid decelerates as it moves down the diverging duct (assuming incompressible or subsonic compressible flow), hence there is an acceleration happening over position.

A significant feature of the Cauchy equation and consequently all other continuum equations (including Euler and Navier–Stokes) is the presence of convective acceleration: the effect of acceleration of a flow with respect to space. While individual fluid particles indeed experience time-dependent acceleration, the convective acceleration of the flow field is a spatial effect, one example being fluid speeding up in a nozzle.

Compressible flow

Remark: here, the deviatoric stress tensor is denoted   (instead of   as it was in the general continuum equations and in the incompressible flow section).

The compressible momentum Navier–Stokes equation results from the following assumptions on the Cauchy stress tensor:[5]

  • the stress is Galilean invariant: it does not depend directly on the flow velocity, but only on spatial derivatives of the flow velocity. So the stress variable is the tensor gradient  .
  • the stress is linear in this variable:  , where   is the fourth-order tensor representing the constant of proportionality, called the viscosity or elasticity tensor, and : is the double-dot product.
  • the fluid is assumed to be isotropic, as with gases and simple liquids, and consequently   is an isotropic tensor; furthermore, since the stress tensor is symmetric, by Helmholtz decomposition it can be expressed in terms of two scalar Lamé parameters, the second viscosity   and the dynamic viscosity  , as it is usual in linear elasticity:
    Linear stress constitutive equation (expression used for elastic solid)

     

    where   is the identity tensor,   is the rate-of-strain tensor and   is the divergence (i.e. rate of expansion) of the flow. So this decomposition can be explicitly defined as:

     

Since the trace of the rate-of-strain tensor in three dimensions is:

 

The trace of the stress tensor in three dimensions becomes:

 

So by alternatively decomposing the stress tensor into isotropic and deviatoric parts, as usual in fluid dynamics:[6]

 

Introducing the bulk viscosity  ,

 

we arrive to the linear constitutive equation in the form usually employed in thermal hydraulics:[5]

Linear stress constitutive equation (expression used for fluids)

 

Both second viscosity   and dynamic viscosity   need not be constant – in general, they depend on two thermodynamics variables if the fluid contains a single chemical species, say for example, pressure and temperature. Any equation that makes explicit one of these transport coefficient in the conservation variables is called an equation of state.[7]

The most general of the Navier–Stokes equations become

Navier–Stokes momentum equation (convective form)

 

Apart from its dependence of pressure and temperature, the second viscosity coefficient also depends on the process, that is to say, the second viscosity coefficient is not just a material property. For instance, in the case of a sound wave with a definitive frequency that alternatively compresses and expands a fluid element, the second viscosity coefficient depends on the frequency of the wave. This dependence is called the dispersion. In some cases, the second viscosity   can be assumed to be constant in which case, the effect of the volume viscosity   is that the mechanical pressure is not equivalent to the thermodynamic pressure:[8] as demonstrated below.

 
 
However, this difference is usually neglected most of the time (that is whenever we are not dealing with processes such as sound absorption and attenuation of shock waves,[9] where second viscosity coefficient becomes important) by explicitly assuming  . The assumption of setting   is called as the Stokes hypothesis.[10] The validity of Stokes hypothesis can be demonstrated for monoatomic gas both experimentally and from the kinetic theory,;[11] for other gases and liquids, Stokes hypothesis is generally incorrect. With the Stokes hypothesis, the Navier–Stokes equations become
Navier–Stokes momentum equation (convective form)

 

If the dynamic viscosity μ is also assumed to be constant, the equations can be simplified further. By computing the divergence of the stress tensor, since the divergence of tensor   is   and the divergence of tensor   is  , one finally arrives to the compressible (most general) Navier–Stokes momentum equation:[12]

Navier–Stokes momentum equation (convective form)

 

where   is the material derivative. The left-hand side changes in the conservation form of the Navier–Stokes momentum equation:

Navier–Stokes momentum equation (conservation form)

 

Bulk viscosity is assumed to be constant, otherwise it should not be taken out of the last derivative. The convective acceleration term can also be written as

 
where the vector   is known as the Lamb vector.

For the special case of an incompressible flow, the pressure constrains the flow so that the volume of fluid elements is constant: isochoric flow resulting in a solenoidal velocity field with  .[13]

Incompressible flow

The incompressible momentum Navier–Stokes equation results from the following assumptions on the Cauchy stress tensor:[5]

  • the stress is Galilean invariant: it does not depend directly on the flow velocity, but only on spatial derivatives of the flow velocity. So the stress variable is the tensor gradient  .
  • the fluid is assumed to be isotropic, as with gases and simple liquids, and consequently   is an isotropic tensor; furthermore, since the deviatoric stress tensor can be expressed in terms of the dynamic viscosity  :
    Stokes' stress constitutive equation (expression used for incompressible elastic solids)
     

    where

     
    is the rate-of-strain tensor. So this decomposition can be made explicit as:[5]
    Stokes's stress constitutive equation (expression used for incompressible viscous fluids)
     

Dynamic viscosity μ need not be constant – in incompressible flows it can depend on density and on pressure. Any equation that makes explicit one of these transport coefficient in the conservative variables is called an equation of state.[7]

The divergence of the deviatoric stress is given by:

 
because   for an incompressible fluid.

Incompressibility rules out density and pressure waves like sound or shock waves, so this simplification is not useful if these phenomena are of interest. The incompressible flow assumption typically holds well with all fluids at low Mach numbers (say up to about Mach 0.3), such as for modelling air winds at normal temperatures.[14] the incompressible Navier–Stokes equations are best visualized by dividing for the density:[15]

Incompressible Navier–Stokes equations (convective form)

 

If the density is constant throughout the fluid domain, or, in other words, if all fluid elements have the same density,  , then we have

Incompressible Navier–Stokes equations (convective form)

 

where   is called the kinematic viscosity.

A laminar flow example

Velocity profile (laminar flow):

 
for the x-direction, simplify the Navier–Stokes equation:
 

Integrate twice to find the velocity profile with boundary conditions y = h, u = 0, y = −h, u = 0:

 

From this equation, substitute in the two boundary conditions to get two equations:

 

Add and solve for B:

 

Substitute and solve for A:

 

Finally this gives the velocity profile:

 

It is well worth observing the meaning of each term (compare to the Cauchy momentum equation):

 

The higher-order term, namely the shear stress divergence  , has simply reduced to the vector Laplacian term  .[16] This Laplacian term can be interpreted as the difference between the velocity at a point and the mean velocity in a small surrounding volume. This implies that – for a Newtonian fluid – viscosity operates as a diffusion of momentum, in much the same way as the heat conduction. In fact neglecting the convection term, incompressible Navier–Stokes equations lead to a vector diffusion equation (namely Stokes equations), but in general the convection term is present, so incompressible Navier–Stokes equations belong to the class of convection–diffusion equations.

In the usual case of an external field being a conservative field:

 
by defining the hydraulic head:
 

one can finally condense the whole source in one term, arriving to the incompressible Navier–Stokes equation with conservative external field:

 

The incompressible Navier–Stokes equations with conservative external field is the fundamental equation of hydraulics. The domain for these equations is commonly a 3 or less dimensional Euclidean space, for which an orthogonal coordinate reference frame is usually set to explicit the system of scalar partial differential equations to be solved. In 3-dimensional orthogonal coordinate systems are 3: Cartesian, cylindrical, and spherical. Expressing the Navier–Stokes vector equation in Cartesian coordinates is quite straightforward and not much influenced by the number of dimensions of the euclidean space employed, and this is the case also for the first-order terms (like the variation and convection ones) also in non-cartesian orthogonal coordinate systems. But for the higher order terms (the two coming from the divergence of the deviatoric stress that distinguish Navier–Stokes equations from Euler equations) some tensor calculus is required for deducing an expression in non-cartesian orthogonal coordinate systems.

The incompressible Navier–Stokes equation is composite, the sum of two orthogonal equations,

 
where   and   are solenoidal and irrotational projection operators satisfying   and   and   are the non-conservative and conservative parts of the body force. This result follows from the Helmholtz theorem (also known as the fundamental theorem of vector calculus). The first equation is a pressureless governing equation for the velocity, while the second equation for the pressure is a functional of the velocity and is related to the pressure Poisson equation.

The explicit functional form of the projection operator in 3D is found from the Helmholtz Theorem:

 
with a similar structure in 2D. Thus the governing equation is an integro-differential equation similar to Coulomb and Biot–Savart law, not convenient for numerical computation.

An equivalent weak or variational form of the equation, proved to produce the same velocity solution as the Navier–Stokes equation,[17] is given by,

 

for divergence-free test functions   satisfying appropriate boundary conditions. Here, the projections are accomplished by the orthogonality of the solenoidal and irrotational function spaces. The discrete form of this is eminently suited to finite element computation of divergence-free flow, as we shall see in the next section. There one will be able to address the question "How does one specify pressure-driven (Poiseuille) problems with a pressureless governing equation?".

The absence of pressure forces from the governing velocity equation demonstrates that the equation is not a dynamic one, but rather a kinematic equation where the divergence-free condition serves the role of a conservation equation. This all would seem to refute the frequent statements that the incompressible pressure enforces the divergence-free condition.

Weak form of the incompressible Navier–Stokes equations

Strong form

Consider the incompressible Navier–Stokes equations for a Newtonian fluid of constant density   in a domain

 
with boundary
 
being   and   portions of the boundary where respectively a Dirichlet and a Neumann boundary condition is applied ( ):[18]
 
  is the fluid velocity,   the fluid pressure,   a given forcing term,   the outward directed unit normal vector to  , and   the viscous stress tensor defined as:[18]
 
Let   be the dynamic viscosity of the fluid,   the second-order identity tensor and   the strain-rate tensor defined as:[18]
 
The functions   and   are given Dirichlet and Neumann boundary data, while   is the initial condition. The first equation is the momentum balance equation, while the second represents the mass conservation, namely the continuity equation. Assuming constant dynamic viscosity, using the vectorial identity
 
and exploiting mass conservation, the divergence of the total stress tensor in the momentum equation can also be expressed as:[18]
 
Moreover, note that the Neumann boundary conditions can be rearranged as:[18]
 

Weak form

In order to find the weak form of the Navier–Stokes equations, firstly, consider the momentum equation[18]

 
multiply it for a test function  , defined in a suitable space  , and integrate both members with respect to the domain  :[18]
 
Counter-integrating by parts the diffusive and the pressure terms and by using the Gauss' theorem:[18]
 

Using these relations, one gets:[18]

 
In the same fashion, the continuity equation is multiplied for a test function q belonging to a space   and integrated in the domain  :[18]
 
The space functions are chosen as follows:
 
Considering that the test function v vanishes on the Dirichlet boundary and considering the Neumann condition, the integral on the boundary can be rearranged as:[18]
 
Having this in mind, the weak formulation of the Navier–Stokes equations is expressed as:[18]
 

Discrete velocity

With partitioning of the problem domain and defining basis functions on the partitioned domain, the discrete form of the governing equation is

 

It is desirable to choose basis functions that reflect the essential feature of incompressible flow – the elements must be divergence-free. While the velocity is the variable of interest, the existence of the stream function or vector potential is necessary by the Helmholtz theorem. Further, to determine fluid flow in the absence of a pressure gradient, one can specify the difference of stream function values across a 2D channel, or the line integral of the tangential component of the vector potential around the channel in 3D, the flow being given by Stokes' theorem. Discussion will be restricted to 2D in the following.

We further restrict discussion to continuous Hermite finite elements which have at least first-derivative degrees-of-freedom. With this, one can draw a large number of candidate triangular and rectangular elements from the plate-bending literature. These elements have derivatives as components of the gradient. In 2D, the gradient and curl of a scalar are clearly orthogonal, given by the expressions,

 

Adopting continuous plate-bending elements, interchanging the derivative degrees-of-freedom and changing the sign of the appropriate one gives many families of stream function elements.

Taking the curl of the scalar stream function elements gives divergence-free velocity elements.[19][20] The requirement that the stream function elements be continuous assures that the normal component of the velocity is continuous across element interfaces, all that is necessary for vanishing divergence on these interfaces.

Boundary conditions are simple to apply. The stream function is constant on no-flow surfaces, with no-slip velocity conditions on surfaces. Stream function differences across open channels determine the flow. No boundary conditions are necessary on open boundaries, though consistent values may be used with some problems. These are all Dirichlet conditions.

The algebraic equations to be solved are simple to set up, but of course are non-linear, requiring iteration of the linearized equations.

Similar considerations apply to three-dimensions, but extension from 2D is not immediate because of the vector nature of the potential, and there exists no simple relation between the gradient and the curl as was the case in 2D.

Pressure recovery

Recovering pressure from the velocity field is easy. The discrete weak equation for the pressure gradient is,

 

where the test/weight functions are irrotational. Any conforming scalar finite element may be used. However, the pressure gradient field may also be of interest. In this case, one can use scalar Hermite elements for the pressure. For the test/weight functions   one would choose the irrotational vector elements obtained from the gradient of the pressure element.

Non-inertial frame of reference

The rotating frame of reference introduces some interesting pseudo-forces into the equations through the material derivative term. Consider a stationary inertial frame of reference   , and a non-inertial frame of reference  , which is translating with velocity   and rotating with angular velocity   with respect to the stationary frame. The Navier–Stokes equation observed from the non-inertial frame then becomes

Navier–Stokes momentum equation in non-inertial frame

 

Here   and   are measured in the non-inertial frame. The first term in the parenthesis represents Coriolis acceleration, the second term is due to centrifugal acceleration, the third is due to the linear acceleration of   with respect to   and the fourth term is due to the angular acceleration of   with respect to  .

Other equations

The Navier–Stokes equations are strictly a statement of the balance of momentum. To fully describe fluid flow, more information is needed, how much depending on the assumptions made. This additional information may include boundary data (no-slip, capillary surface, etc.), conservation of mass, balance of energy, and/or an equation of state.

Continuity equation for incompressible fluid

Regardless of the flow assumptions, a statement of the conservation of mass is generally necessary. This is achieved through the mass continuity equation, given in its most general form as:

 
or, using the substantive derivative:
 

For incompressible fluid, density along the line of flow remains constant over time,

 

Therefore divergence of velocity is always zero:

 

Stream function for incompressible 2D fluid

Taking the curl of the incompressible Navier–Stokes equation results in the elimination of pressure. This is especially easy to see if 2D Cartesian flow is assumed (like in the degenerate 3D case with   and no dependence of anything on  ), where the equations reduce to:

 

Differentiating the first with respect to  , the second with respect to   and subtracting the resulting equations will eliminate pressure and any conservative force. For incompressible flow, defining the stream function   through

 
results in mass continuity being unconditionally satisfied (given the stream function is continuous), and then incompressible Newtonian 2D momentum and mass conservation condense into one equation:
 

where   is the 2D biharmonic operator and   is the kinematic viscosity,  . We can also express this compactly using the Jacobian determinant:

 

This single equation together with appropriate boundary conditions describes 2D fluid flow, taking only kinematic viscosity as a parameter. Note that the equation for creeping flow results when the left side is assumed zero.

In axisymmetric flow another stream function formulation, called the Stokes stream function, can be used to describe the velocity components of an incompressible flow with one scalar function.

The incompressible Navier–Stokes equation is a differential algebraic equation, having the inconvenient feature that there is no explicit mechanism for advancing the pressure in time. Consequently, much effort has been expended to eliminate the pressure from all or part of the computational process. The stream function formulation eliminates the pressure but only in two dimensions and at the expense of introducing higher derivatives and elimination of the velocity, which is the primary variable of interest.

Properties

Nonlinearity

The Navier–Stokes equations are nonlinear partial differential equations in the general case and so remain in almost every real situation.[21][22] In some cases, such as one-dimensional flow and Stokes flow (or creeping flow), the equations can be simplified to linear equations. The nonlinearity makes most problems difficult or impossible to solve and is the main contributor to the turbulence that the equations model.

The nonlinearity is due to convective acceleration, which is an acceleration associated with the change in velocity over position. Hence, any convective flow, whether turbulent or not, will involve nonlinearity. An example of convective but laminar (nonturbulent) flow would be the passage of a viscous fluid (for example, oil) through a small converging nozzle. Such flows, whether exactly solvable or not, can often be thoroughly studied and understood.[23]

Turbulence

Turbulence is the time-dependent chaotic behaviour seen in many fluid flows. It is generally believed that it is due to the inertia of the fluid as a whole: the culmination of time-dependent and convective acceleration; hence flows where inertial effects are small tend to be laminar (the Reynolds number quantifies how much the flow is affected by inertia). It is believed, though not known with certainty, that the Navier–Stokes equations describe turbulence properly.[24]

The numerical solution of the Navier–Stokes equations for turbulent flow is extremely difficult, and due to the significantly different mixing-length scales that are involved in turbulent flow, the stable solution of this requires such a fine mesh resolution that the computational time becomes significantly infeasible for calculation or direct numerical simulation. Attempts to solve turbulent flow using a laminar solver typically result in a time-unsteady solution, which fails to converge appropriately. To counter this, time-averaged equations such as the Reynolds-averaged Navier–Stokes equations (RANS), supplemented with turbulence models, are used in practical computational fluid dynamics (CFD) applications when modeling turbulent flows. Some models include the Spalart–Allmaras, kω, kε, and SST models, which add a variety of additional equations to bring closure to the RANS equations. Large eddy simulation (LES) can also be used to solve these equations numerically. This approach is computationally more expensive—in time and in computer memory—than RANS, but produces better results because it explicitly resolves the larger turbulent scales.

Applicability

Together with supplemental equations (for example, conservation of mass) and well-formulated boundary conditions, the Navier–Stokes equations seem to model fluid motion accurately; even turbulent flows seem (on average) to agree with real world observations.

The Navier–Stokes equations assume that the fluid being studied is a continuum (it is infinitely divisible and not composed of particles such as atoms or molecules), and is not moving at relativistic velocities. At very small scales or under extreme conditions, real fluids made out of discrete molecules will produce results different from the continuous fluids modeled by the Navier–Stokes equations. For example, capillarity of internal layers in fluids appears for flow with high gradients.[25] For large Knudsen number of the problem, the Boltzmann equation may be a suitable replacement.[26] Failing that, one may have to resort to molecular dynamics or various hybrid methods.[27]

Another limitation is simply the complicated nature of the equations. Time-tested formulations exist for common fluid families, but the application of the Navier–Stokes equations to less common families tends to result in very complicated formulations and often to open research problems. For this reason, these equations are usually written for Newtonian fluids where the viscosity model is linear; truly general models for the flow of other kinds of fluids (such as blood) do not exist.[28]

Application to specific problems

The Navier–Stokes equations, even when written explicitly for specific fluids, are rather generic in nature and their proper application to specific problems can be very diverse. This is partly because there is an enormous variety of problems that may be modeled, ranging from as simple as the distribution of static pressure to as complicated as multiphase flow driven by surface tension.

Generally, application to specific problems begins with some flow assumptions and initial/boundary condition formulation, this may be followed by scale analysis to further simplify the problem.

 
Visualization of (a) parallel flow and (b) radial flow.

Parallel flow

Assume steady, parallel, one-dimensional, non-convective pressure-driven flow between parallel plates, the resulting scaled (dimensionless) boundary value problem is:

 

The boundary condition is the no slip condition. This problem is easily solved for the flow field:

 

From this point onward, more quantities of interest can be easily obtained, such as viscous drag force or net flow rate.

Radial flow

Difficulties may arise when the problem becomes slightly more complicated. A seemingly modest twist on the parallel flow above would be the radial flow between parallel plates; this involves convection and thus non-linearity. The velocity field may be represented by a function f(z) that must satisfy:

 

This ordinary differential equation is what is obtained when the Navier–Stokes equations are written and the flow assumptions applied (additionally, the pressure gradient is solved for). The nonlinear term makes this a very difficult problem to solve analytically (a lengthy implicit solution may be found which involves elliptic integrals and roots of cubic polynomials). Issues with the actual existence of solutions arise for   (approximately; this is not 2), the parameter   being the Reynolds number with appropriately chosen scales.[29] This is an example of flow assumptions losing their applicability, and an example of the difficulty in "high" Reynolds number flows.[29]

Convection

A type of natural convection that can be described by the Navier–Stokes equation is the Rayleigh–Bénard convection. It is one of the most commonly studied convection phenomena because of its analytical and experimental accessibility.

Exact solutions of the Navier–Stokes equations

Some exact solutions to the Navier–Stokes equations exist. Examples of degenerate cases—with the non-linear terms in the Navier–Stokes equations equal to zero—are Poiseuille flow, Couette flow and the oscillatory Stokes boundary layer. But also, more interesting examples, solutions to the full non-linear equations, exist, such as Jeffery–Hamel flow, Von Kármán swirling flow, stagnation point flow, Landau–Squire jet, and Taylor–Green vortex.[30][31][32] Note that the existence of these exact solutions does not imply they are stable: turbulence may develop at higher Reynolds numbers.

Under additional assumptions, the component parts can be separated.[33]

A two-dimensional example

For example, in the case of an unbounded planar domain with two-dimensional — incompressible and stationary — flow in polar coordinates (r,φ), the velocity components (ur,uφ) and pressure p are:[34]

 

where A and B are arbitrary constants. This solution is valid in the domain r ≥ 1 and for A < −2ν.

In Cartesian coordinates, when the viscosity is zero (ν = 0), this is:

 
A three-dimensional example

For example, in the case of an unbounded Euclidean domain with three-dimensional — incompressible, stationary and with zero viscosity (ν = 0) — radial flow in Cartesian coordinates (x,y,z), the velocity vector v and pressure p are:[citation needed]

 

There is a singularity at x = y = z = 0.

A three-dimensional steady-state vortex solution

 
Wire model of flow lines along a Hopf fibration.

A steady-state example with no singularities comes from considering the flow along the lines of a Hopf fibration. Let   be a constant radius of the inner coil. One set of solutions is given by:[35]

 

for arbitrary constants   and  . This is a solution in a non-viscous gas (compressible fluid) whose density, velocities and pressure goes to zero far from the origin. (Note this is not a solution to the Clay Millennium problem because that refers to incompressible fluids where   is a constant, and neither does it deal with the uniqueness of the Navier–Stokes equations with respect to any turbulence properties.) It is also worth pointing out that the components of the velocity vector are exactly those from the Pythagorean quadruple parametrization. Other choices of density and pressure are possible with the same velocity field:

Other choices of density and pressure

Another choice of pressure and density with the same velocity vector above is one where the pressure and density fall to zero at the origin and are highest in the central loop at z = 0, x2 + y2 = r2:

 

In fact in general there are simple solutions for any polynomial function f where the density is:

 

Viscous three-dimensional periodic solutions

Two examples of periodic fully-three-dimensional viscous solutions are described in.[36] These solutions are defined on a three-dimensional torus   and are characterized by positive and negative helicity respectively. The solution with positive helicity is given by:

 
where   is the wave number and the velocity components are normalized so that the average kinetic energy per unit of mass is   at  . The pressure field is obtained from the velocity field as   (where
navier, stokes, equations, stohks, partial, differential, equations, which, describe, motion, viscous, fluid, substances, named, after, french, engineer, physicist, claude, louis, navier, anglo, irish, physicist, mathematician, george, gabriel, stokes, they, w. The Navier Stokes equations n ae v ˈ j eɪ s t oʊ k s nav YAY STOHKS are partial differential equations which describe the motion of viscous fluid substances named after French engineer and physicist Claude Louis Navier and Anglo Irish physicist and mathematician George Gabriel Stokes They were developed over several decades of progressively building the theories from 1822 Navier to 1842 1850 Stokes Claude Louis NavierGeorge Gabriel Stokes The Navier Stokes equations mathematically express momentum balance and conservation of mass for Newtonian fluids They are sometimes accompanied by an equation of state relating pressure temperature and density 1 They arise from applying Isaac Newton s second law to fluid motion together with the assumption that the stress in the fluid is the sum of a diffusing viscous term proportional to the gradient of velocity and a pressure term hence describing viscous flow The difference between them and the closely related Euler equations is that Navier Stokes equations take viscosity into account while the Euler equations model only inviscid flow As a result the Navier Stokes are a parabolic equation and therefore have better analytic properties at the expense of having less mathematical structure e g they are never completely integrable The Navier Stokes equations are useful because they describe the physics of many phenomena of scientific and engineering interest They may be used to model the weather ocean currents water flow in a pipe and air flow around a wing The Navier Stokes equations in their full and simplified forms help with the design of aircraft and cars the study of blood flow the design of power stations the analysis of pollution and many other things Coupled with Maxwell s equations they can be used to model and study magnetohydrodynamics The Navier Stokes equations are also of great interest in a purely mathematical sense Despite their wide range of practical uses it has not yet been proven whether smooth solutions always exist in three dimensions i e whether they are infinitely differentiable or even just bounded at all points in the domain This is called the Navier Stokes existence and smoothness problem The Clay Mathematics Institute has called this one of the seven most important open problems in mathematics and has offered a US 1 million prize for a solution or a counterexample 2 3 Contents 1 Flow velocity 2 General continuum equations 2 1 Convective acceleration 3 Compressible flow 4 Incompressible flow 4 1 Weak form of the incompressible Navier Stokes equations 4 1 1 Strong form 4 1 2 Weak form 4 2 Discrete velocity 4 3 Pressure recovery 5 Non inertial frame of reference 6 Other equations 6 1 Continuity equation for incompressible fluid 7 Stream function for incompressible 2D fluid 8 Properties 8 1 Nonlinearity 8 2 Turbulence 8 3 Applicability 9 Application to specific problems 9 1 Parallel flow 9 2 Radial flow 9 3 Convection 10 Exact solutions of the Navier Stokes equations 10 1 A three dimensional steady state vortex solution 10 2 Viscous three dimensional periodic solutions 11 Wyld diagrams 12 Representations in 3D 12 1 Cartesian coordinates 12 2 Cylindrical coordinates 12 3 Spherical coordinates 13 Navier Stokes equations use in games 14 See also 15 Citations 16 General references 17 External linksFlow velocity EditThe solution of the equations is a flow velocity It is a vector field to every point in a fluid at any moment in a time interval it gives a vector whose direction and magnitude are those of the velocity of the fluid at that point in space and at that moment in time It is usually studied in three spatial dimensions and one time dimension although two spatial dimensional and steady state cases are often used as models and higher dimensional analogues are studied in both pure and applied mathematics Once the velocity field is calculated other quantities of interest such as pressure or temperature may be found using dynamical equations and relations This is different from what one normally sees in classical mechanics where solutions are typically trajectories of position of a particle or deflection of a continuum Studying velocity instead of position makes more sense for a fluid although for visualization purposes one can compute various trajectories In particular the streamlines of a vector field interpreted as flow velocity are the paths along which a massless fluid particle would travel These paths are the integral curves whose derivative at each point is equal to the vector field and they can represent visually the behavior of the vector field at a point in time General continuum equations EditMain article Derivation of the Navier Stokes equations See also Cauchy momentum equation Conservation form The Navier Stokes momentum equation can be derived as a particular form of the Cauchy momentum equation whose general convective form isD u D t 1 r s g displaystyle frac mathrm D mathbf u mathrm D t frac 1 rho nabla cdot boldsymbol sigma mathbf g By setting the Cauchy stress tensor s textstyle boldsymbol sigma to be the sum of a viscosity term t textstyle boldsymbol tau the deviatoric stress and a pressure term p I textstyle p mathbf I volumetric stress we arrive at Cauchy momentum equation convective form r D u D t p t r g displaystyle rho frac mathrm D mathbf u mathrm D t nabla p nabla cdot boldsymbol tau rho mathbf g where D D t textstyle frac mathrm D mathrm D t is the material derivative defined as t u textstyle frac partial partial t mathbf u cdot nabla r textstyle rho is the mass density u textstyle mathbf u is the flow velocity textstyle nabla cdot is the divergence p textstyle p is the pressure t textstyle t is time t textstyle boldsymbol tau is the deviatoric stress tensor which has order 2 g textstyle mathbf g represents body accelerations acting on the continuum for example gravity inertial accelerations electrostatic accelerations and so on In this form it is apparent that in the assumption of an inviscid fluid no deviatoric stress Cauchy equations reduce to the Euler equations Assuming conservation of mass we can use the mass continuity equation or simply continuity equation r t r u 0 displaystyle frac partial rho partial t nabla cdot rho mathbf u 0 to arrive at the conservation form of the equations of motion This is often written 4 Cauchy momentum equation conservation form t r u r u u p t r g displaystyle frac partial partial t rho mathbf u nabla cdot rho mathbf u otimes mathbf u nabla p nabla cdot boldsymbol tau rho mathbf g where textstyle otimes is the outer product u v u v T displaystyle mathbf u otimes mathbf v mathbf u mathbf v mathrm T The left side of the equation describes acceleration and may be composed of time dependent and convective components also the effects of non inertial coordinates if present The right side of the equation is in effect a summation of hydrostatic effects the divergence of deviatoric stress and body forces such as gravity All non relativistic balance equations such as the Navier Stokes equations can be derived by beginning with the Cauchy equations and specifying the stress tensor through a constitutive relation By expressing the deviatoric shear stress tensor in terms of viscosity and the fluid velocity gradient and assuming constant viscosity the above Cauchy equations will lead to the Navier Stokes equations below Convective acceleration Edit See also Cauchy momentum equation Convective acceleration An example of convection Though the flow may be steady time independent the fluid decelerates as it moves down the diverging duct assuming incompressible or subsonic compressible flow hence there is an acceleration happening over position A significant feature of the Cauchy equation and consequently all other continuum equations including Euler and Navier Stokes is the presence of convective acceleration the effect of acceleration of a flow with respect to space While individual fluid particles indeed experience time dependent acceleration the convective acceleration of the flow field is a spatial effect one example being fluid speeding up in a nozzle Compressible flow EditRemark here the deviatoric stress tensor is denoted s textstyle boldsymbol sigma instead of t textstyle boldsymbol tau as it was in the general continuum equations and in the incompressible flow section The compressible momentum Navier Stokes equation results from the following assumptions on the Cauchy stress tensor 5 the stress is Galilean invariant it does not depend directly on the flow velocity but only on spatial derivatives of the flow velocity So the stress variable is the tensor gradient u textstyle nabla mathbf u the stress is linear in this variable s u C u textstyle boldsymbol sigma left nabla mathbf u right mathbf C left nabla mathbf u right where C textstyle mathbf C is the fourth order tensor representing the constant of proportionality called the viscosity or elasticity tensor and is the double dot product the fluid is assumed to be isotropic as with gases and simple liquids and consequently V textstyle mathbf V is an isotropic tensor furthermore since the stress tensor is symmetric by Helmholtz decomposition it can be expressed in terms of two scalar Lame parameters the second viscosity l textstyle lambda and the dynamic viscosity m textstyle mu as it is usual in linear elasticity Linear stress constitutive equation expression used for elastic solid s l u I 2 m e displaystyle boldsymbol sigma lambda nabla cdot mathbf u mathbf I 2 mu boldsymbol varepsilon where I textstyle mathbf I is the identity tensor e u 1 2 u 1 2 u T textstyle boldsymbol varepsilon left nabla mathbf u right equiv frac 1 2 nabla mathbf u frac 1 2 left nabla mathbf u right T is the rate of strain tensor and u textstyle nabla cdot mathbf u is the divergence i e rate of expansion of the flow So this decomposition can be explicitly defined as s l u I m u u T displaystyle boldsymbol sigma lambda nabla cdot mathbf u mathbf I mu left nabla mathbf u nabla mathbf u mathrm T right Since the trace of the rate of strain tensor in three dimensions is tr e u displaystyle operatorname tr boldsymbol varepsilon nabla cdot mathbf u The trace of the stress tensor in three dimensions becomes tr s 3 l 2 m u displaystyle operatorname tr boldsymbol sigma 3 lambda 2 mu nabla cdot mathbf u So by alternatively decomposing the stress tensor into isotropic and deviatoric parts as usual in fluid dynamics 6 s l 2 3 m u I m u u T 2 3 u I displaystyle boldsymbol sigma left lambda tfrac 2 3 mu right left nabla cdot mathbf u right mathbf I mu left nabla mathbf u left nabla mathbf u right mathrm T tfrac 2 3 left nabla cdot mathbf u right mathbf I right Introducing the bulk viscosity z textstyle zeta z l 2 3 m displaystyle zeta equiv lambda tfrac 2 3 mu we arrive to the linear constitutive equation in the form usually employed in thermal hydraulics 5 Linear stress constitutive equation expression used for fluids s z u I m u u T 2 3 u I displaystyle boldsymbol sigma zeta nabla cdot mathbf u mathbf I mu left nabla mathbf u nabla mathbf u mathrm T tfrac 2 3 nabla cdot mathbf u mathbf I right Both second viscosity z textstyle zeta and dynamic viscosity m textstyle mu need not be constant in general they depend on two thermodynamics variables if the fluid contains a single chemical species say for example pressure and temperature Any equation that makes explicit one of these transport coefficient in the conservation variables is called an equation of state 7 The most general of the Navier Stokes equations become Navier Stokes momentum equation convective form r D u D t r u t u u p m u u T 2 3 u I z u I r g displaystyle rho frac mathrm D mathbf u mathrm D t rho left frac partial mathbf u partial t mathbf u cdot nabla mathbf u right nabla p nabla cdot left mu left nabla mathbf u nabla mathbf u mathrm T tfrac 2 3 nabla cdot mathbf u mathbf I right zeta nabla cdot mathbf u mathbf I right rho mathbf g Apart from its dependence of pressure and temperature the second viscosity coefficient also depends on the process that is to say the second viscosity coefficient is not just a material property For instance in the case of a sound wave with a definitive frequency that alternatively compresses and expands a fluid element the second viscosity coefficient depends on the frequency of the wave This dependence is called the dispersion In some cases the second viscosity z textstyle zeta can be assumed to be constant in which case the effect of the volume viscosity z textstyle zeta is that the mechanical pressure is not equivalent to the thermodynamic pressure 8 as demonstrated below u I u displaystyle nabla cdot nabla cdot mathbf u mathbf I nabla nabla cdot mathbf u p p z u displaystyle bar p equiv p zeta nabla cdot mathbf u However this difference is usually neglected most of the time that is whenever we are not dealing with processes such as sound absorption and attenuation of shock waves 9 where second viscosity coefficient becomes important by explicitly assuming z 0 textstyle zeta 0 The assumption of setting z 0 textstyle zeta 0 is called as the Stokes hypothesis 10 The validity of Stokes hypothesis can be demonstrated for monoatomic gas both experimentally and from the kinetic theory 11 for other gases and liquids Stokes hypothesis is generally incorrect With the Stokes hypothesis the Navier Stokes equations become Navier Stokes momentum equation convective form r D u D t r u t u u p m u u T 2 3 u I r g displaystyle rho frac mathrm D mathbf u mathrm D t rho left frac partial mathbf u partial t mathbf u cdot nabla mathbf u right nabla p nabla cdot left mu left nabla mathbf u nabla mathbf u mathrm T tfrac 2 3 nabla cdot mathbf u mathbf I right right rho mathbf g If the dynamic viscosity m is also assumed to be constant the equations can be simplified further By computing the divergence of the stress tensor since the divergence of tensor u textstyle nabla mathbf u is 2 u textstyle nabla 2 mathbf u and the divergence of tensor u T textstyle left nabla mathbf u right mathrm T is u textstyle nabla left nabla cdot mathbf u right one finally arrives to the compressible most general Navier Stokes momentum equation 12 Navier Stokes momentum equation convective form r D u D t r u t u u p m 2 u 1 3 m u r g displaystyle rho frac mathrm D mathbf u mathrm D t rho left frac partial mathbf u partial t mathbf u cdot nabla mathbf u right nabla p mu nabla 2 mathbf u tfrac 1 3 mu nabla nabla cdot mathbf u rho mathbf g where D D t textstyle frac mathrm D mathrm D t is the material derivative The left hand side changes in the conservation form of the Navier Stokes momentum equation Navier Stokes momentum equation conservation form t r u r u u p m 2 u 1 3 m u r g displaystyle frac partial partial t rho mathbf u nabla cdot rho mathbf u otimes mathbf u nabla p mu nabla 2 mathbf u tfrac 1 3 mu nabla nabla cdot mathbf u rho mathbf g Bulk viscosity is assumed to be constant otherwise it should not be taken out of the last derivative The convective acceleration term can also be written asu u u u 1 2 u 2 displaystyle mathbf u cdot nabla mathbf u nabla times mathbf u times mathbf u tfrac 1 2 nabla mathbf u 2 where the vector u u textstyle nabla times mathbf u times mathbf u is known as the Lamb vector For the special case of an incompressible flow the pressure constrains the flow so that the volume of fluid elements is constant isochoric flow resulting in a solenoidal velocity field with u 0 textstyle nabla cdot mathbf u 0 13 Incompressible flow EditThe incompressible momentum Navier Stokes equation results from the following assumptions on the Cauchy stress tensor 5 the stress is Galilean invariant it does not depend directly on the flow velocity but only on spatial derivatives of the flow velocity So the stress variable is the tensor gradient u textstyle nabla mathbf u the fluid is assumed to be isotropic as with gases and simple liquids and consequently t textstyle boldsymbol tau is an isotropic tensor furthermore since the deviatoric stress tensor can be expressed in terms of the dynamic viscosity m textstyle mu Stokes stress constitutive equation expression used for incompressible elastic solids t 2 m e displaystyle boldsymbol tau 2 mu boldsymbol varepsilon wheree 1 2 u u T displaystyle boldsymbol varepsilon tfrac 1 2 left mathbf nabla u mathbf nabla u mathrm T right is the rate of strain tensor So this decomposition can be made explicit as 5 Stokes s stress constitutive equation expression used for incompressible viscous fluids t m u u T displaystyle boldsymbol tau mu left nabla mathbf u nabla mathbf u mathrm T right Dynamic viscosity m need not be constant in incompressible flows it can depend on density and on pressure Any equation that makes explicit one of these transport coefficient in the conservative variables is called an equation of state 7 The divergence of the deviatoric stress is given by t 2 m e m u u T m 2 u displaystyle nabla cdot boldsymbol tau 2 mu nabla cdot boldsymbol varepsilon mu nabla cdot left nabla mathbf u nabla mathbf u mathrm T right mu nabla 2 mathbf u because u 0 textstyle nabla cdot mathbf u 0 for an incompressible fluid Incompressibility rules out density and pressure waves like sound or shock waves so this simplification is not useful if these phenomena are of interest The incompressible flow assumption typically holds well with all fluids at low Mach numbers say up to about Mach 0 3 such as for modelling air winds at normal temperatures 14 the incompressible Navier Stokes equations are best visualized by dividing for the density 15 Incompressible Navier Stokes equations convective form u t u u n 2 u 1 r p g displaystyle frac partial mathbf u partial t mathbf u cdot nabla mathbf u nu nabla 2 mathbf u frac 1 rho nabla p mathbf g If the density is constant throughout the fluid domain or in other words if all fluid elements have the same density r r 0 textstyle rho rho 0 then we have Incompressible Navier Stokes equations convective form u t u u n 2 u p r 0 g displaystyle frac partial mathbf u partial t mathbf u cdot nabla mathbf u nu nabla 2 mathbf u nabla left frac p rho 0 right mathbf g where n m r 0 textstyle nu frac mu rho 0 is called the kinematic viscosity A laminar flow exampleVelocity profile laminar flow u x u y u y 0 u z 0 displaystyle u x u y quad u y 0 quad u z 0 for the x direction simplify the Navier Stokes equation 0 d P d x m d 2 u d y 2 displaystyle 0 frac mathrm d P mathrm d x mu left frac mathrm d 2 u mathrm d y 2 right Integrate twice to find the velocity profile with boundary conditions y h u 0 y h u 0 u 1 2 m d P d x y 2 A y B displaystyle u frac 1 2 mu frac mathrm d P mathrm d x y 2 Ay B From this equation substitute in the two boundary conditions to get two equations 0 1 2 m d P d x h 2 A h B 0 1 2 m d P d x h 2 A h B displaystyle begin aligned 0 amp frac 1 2 mu frac mathrm d P mathrm d x h 2 Ah B 0 amp frac 1 2 mu frac mathrm d P mathrm d x h 2 Ah B end aligned Add and solve for B B 1 2 m d P d x h 2 displaystyle B frac 1 2 mu frac mathrm d P mathrm d x h 2 Substitute and solve for A A 0 displaystyle A 0 Finally this gives the velocity profile u 1 2 m d P d x y 2 h 2 displaystyle u frac 1 2 mu frac mathrm d P mathrm d x left y 2 h 2 right It is well worth observing the meaning of each term compare to the Cauchy momentum equation u t Variation u u Divergence Inertia per volume w Internal source n 2 u Diffusion Divergence of stress g External source displaystyle overbrace vphantom frac underbrace frac partial mathbf u partial t text Variation underbrace vphantom frac mathbf u cdot nabla mathbf u text Divergence text Inertia per volume overbrace vphantom frac partial partial underbrace vphantom frac nabla w begin smallmatrix text Internal text source end smallmatrix underbrace vphantom frac nu nabla 2 mathbf u text Diffusion text Divergence of stress underbrace vphantom frac mathbf g begin smallmatrix text External text source end smallmatrix The higher order term namely the shear stress divergence t textstyle nabla cdot boldsymbol tau has simply reduced to the vector Laplacian term m 2 u textstyle mu nabla 2 mathbf u 16 This Laplacian term can be interpreted as the difference between the velocity at a point and the mean velocity in a small surrounding volume This implies that for a Newtonian fluid viscosity operates as a diffusion of momentum in much the same way as the heat conduction In fact neglecting the convection term incompressible Navier Stokes equations lead to a vector diffusion equation namely Stokes equations but in general the convection term is present so incompressible Navier Stokes equations belong to the class of convection diffusion equations In the usual case of an external field being a conservative field g f displaystyle mathbf g nabla varphi by defining the hydraulic head h w f displaystyle h equiv w varphi one can finally condense the whole source in one term arriving to the incompressible Navier Stokes equation with conservative external field u t u u n 2 u h displaystyle frac partial mathbf u partial t mathbf u cdot nabla mathbf u nu nabla 2 mathbf u nabla h The incompressible Navier Stokes equations with conservative external field is the fundamental equation of hydraulics The domain for these equations is commonly a 3 or less dimensional Euclidean space for which an orthogonal coordinate reference frame is usually set to explicit the system of scalar partial differential equations to be solved In 3 dimensional orthogonal coordinate systems are 3 Cartesian cylindrical and spherical Expressing the Navier Stokes vector equation in Cartesian coordinates is quite straightforward and not much influenced by the number of dimensions of the euclidean space employed and this is the case also for the first order terms like the variation and convection ones also in non cartesian orthogonal coordinate systems But for the higher order terms the two coming from the divergence of the deviatoric stress that distinguish Navier Stokes equations from Euler equations some tensor calculus is required for deducing an expression in non cartesian orthogonal coordinate systems The incompressible Navier Stokes equation is composite the sum of two orthogonal equations u t P S u u n 2 u f S r 1 p P I u u n 2 u f I displaystyle begin aligned frac partial mathbf u partial t amp Pi S left mathbf u cdot nabla mathbf u nu nabla 2 mathbf u right mathbf f S rho 1 nabla p amp Pi I left mathbf u cdot nabla mathbf u nu nabla 2 mathbf u right mathbf f I end aligned where P S textstyle Pi S and P I textstyle Pi I are solenoidal and irrotational projection operators satisfying P S P I 1 textstyle Pi S Pi I 1 and f S textstyle mathbf f S and f I textstyle mathbf f I are the non conservative and conservative parts of the body force This result follows from the Helmholtz theorem also known as the fundamental theorem of vector calculus The first equation is a pressureless governing equation for the velocity while the second equation for the pressure is a functional of the velocity and is related to the pressure Poisson equation The explicit functional form of the projection operator in 3D is found from the Helmholtz Theorem P S F r 1 4 p F r r r d V P I 1 P S displaystyle Pi S mathbf F mathbf r frac 1 4 pi nabla times int frac nabla prime times mathbf F mathbf r mathbf r mathbf r mathrm d V quad Pi I 1 Pi S with a similar structure in 2D Thus the governing equation is an integro differential equation similar to Coulomb and Biot Savart law not convenient for numerical computation An equivalent weak or variational form of the equation proved to produce the same velocity solution as the Navier Stokes equation 17 is given by w u t w u u n w u w f S displaystyle left mathbf w frac partial mathbf u partial t right bigl mathbf w left mathbf u cdot nabla right mathbf u bigr nu left nabla mathbf w nabla mathbf u right left mathbf w mathbf f S right for divergence free test functions w textstyle mathbf w satisfying appropriate boundary conditions Here the projections are accomplished by the orthogonality of the solenoidal and irrotational function spaces The discrete form of this is eminently suited to finite element computation of divergence free flow as we shall see in the next section There one will be able to address the question How does one specify pressure driven Poiseuille problems with a pressureless governing equation The absence of pressure forces from the governing velocity equation demonstrates that the equation is not a dynamic one but rather a kinematic equation where the divergence free condition serves the role of a conservation equation This all would seem to refute the frequent statements that the incompressible pressure enforces the divergence free condition Weak form of the incompressible Navier Stokes equations Edit Strong form Edit Consider the incompressible Navier Stokes equations for a Newtonian fluid of constant density r textstyle rho in a domainW R d d 2 3 displaystyle Omega subset mathbb R d quad d 2 3 with boundary W G D G N displaystyle partial Omega Gamma D cup Gamma N being G D textstyle Gamma D and G N textstyle Gamma N portions of the boundary where respectively a Dirichlet and a Neumann boundary condition is applied G D G N textstyle Gamma D cap Gamma N emptyset 18 r u t r u u s u p f in W 0 T u 0 in W 0 T u g on G D 0 T s u p n h on G N 0 T u 0 u 0 in W 0 displaystyle begin cases rho dfrac partial mathbf u partial t rho mathbf u cdot nabla mathbf u nabla cdot boldsymbol sigma mathbf u p mathbf f amp text in Omega times 0 T nabla cdot mathbf u 0 amp text in Omega times 0 T mathbf u mathbf g amp text on Gamma D times 0 T boldsymbol sigma mathbf u p hat mathbf n mathbf h amp text on Gamma N times 0 T mathbf u 0 mathbf u 0 amp text in Omega times 0 end cases u textstyle mathbf u is the fluid velocity p textstyle p the fluid pressure f textstyle mathbf f a given forcing term n displaystyle hat mathbf n the outward directed unit normal vector to G N textstyle Gamma N and s u p textstyle boldsymbol sigma mathbf u p the viscous stress tensor defined as 18 s u p p I 2 m e u displaystyle boldsymbol sigma mathbf u p p mathbf I 2 mu boldsymbol varepsilon mathbf u Let m textstyle mu be the dynamic viscosity of the fluid I textstyle mathbf I the second order identity tensor and e u textstyle boldsymbol varepsilon mathbf u the strain rate tensor defined as 18 e u 1 2 u u T displaystyle boldsymbol varepsilon mathbf u frac 1 2 left left nabla mathbf u right left nabla mathbf u right mathrm T right The functions g textstyle mathbf g and h textstyle mathbf h are given Dirichlet and Neumann boundary data while u 0 textstyle mathbf u 0 is the initial condition The first equation is the momentum balance equation while the second represents the mass conservation namely the continuity equation Assuming constant dynamic viscosity using the vectorial identity f T f displaystyle nabla cdot left nabla mathbf f right mathrm T nabla nabla cdot mathbf f and exploiting mass conservation the divergence of the total stress tensor in the momentum equation can also be expressed as 18 s u p p I 2 m e u p 2 m e u p 2 m 1 2 u u T p m D u u T p m D u u 0 p m D u displaystyle begin aligned nabla cdot boldsymbol sigma mathbf u p amp nabla cdot left p mathbf I 2 mu boldsymbol varepsilon mathbf u right amp nabla p 2 mu nabla cdot boldsymbol varepsilon mathbf u amp nabla p 2 mu nabla cdot left tfrac 1 2 left left nabla mathbf u right left nabla mathbf u right mathrm T right right amp nabla p mu left Delta mathbf u nabla cdot left nabla mathbf u right mathrm T right amp nabla p mu bigl Delta mathbf u nabla underbrace nabla cdot mathbf u 0 bigr nabla p mu Delta mathbf u end aligned Moreover note that the Neumann boundary conditions can be rearranged as 18 s u p n p I 2 m e u n p n m u n displaystyle boldsymbol sigma mathbf u p hat mathbf n left p mathbf I 2 mu boldsymbol varepsilon mathbf u right hat mathbf n p hat mathbf n mu frac partial boldsymbol u partial hat mathbf n Weak form Edit In order to find the weak form of the Navier Stokes equations firstly consider the momentum equation 18 r u t m D u r u u p f displaystyle rho frac partial mathbf u partial t mu Delta mathbf u rho mathbf u cdot nabla mathbf u nabla p mathbf f multiply it for a test function v textstyle mathbf v defined in a suitable space V textstyle V and integrate both members with respect to the domain W textstyle Omega 18 W r u t v W m D u v W r u u v W p v W f v displaystyle int limits Omega rho frac partial mathbf u partial t cdot mathbf v int limits Omega mu Delta mathbf u cdot mathbf v int limits Omega rho mathbf u cdot nabla mathbf u cdot mathbf v int limits Omega nabla p cdot mathbf v int limits Omega mathbf f cdot mathbf v Counter integrating by parts the diffusive and the pressure terms and by using the Gauss theorem 18 W m D u v W m u v W m u n v W p v W p v W p v n displaystyle begin aligned int limits Omega mu Delta mathbf u cdot mathbf v amp int Omega mu nabla mathbf u cdot nabla mathbf v int limits partial Omega mu frac partial mathbf u partial hat mathbf n cdot mathbf v int limits Omega nabla p cdot mathbf v amp int limits Omega p nabla cdot mathbf v int limits partial Omega p mathbf v cdot hat mathbf n end aligned Using these relations one gets 18 W r u t v W m u v W r u u v W p v W f v W m u n p n v v V displaystyle int limits Omega rho dfrac partial mathbf u partial t cdot mathbf v int limits Omega mu nabla mathbf u cdot nabla mathbf v int limits Omega rho mathbf u cdot nabla mathbf u cdot mathbf v int limits Omega p nabla cdot mathbf v int limits Omega mathbf f cdot mathbf v int limits partial Omega left mu frac partial mathbf u partial hat mathbf n p hat mathbf n right cdot mathbf v quad forall mathbf v in V In the same fashion the continuity equation is multiplied for a test function q belonging to a space Q textstyle Q and integrated in the domain W textstyle Omega 18 W q u 0 q Q displaystyle int limits Omega q nabla cdot mathbf u 0 quad forall q in Q The space functions are chosen as follows V H 0 1 W d v H 1 W d v 0 on G D Q L 2 W displaystyle begin aligned V left H 0 1 Omega right d amp left mathbf v in left H 1 Omega right d quad mathbf v mathbf 0 text on Gamma D right Q amp L 2 Omega end aligned Considering that the test function v vanishes on the Dirichlet boundary and considering the Neumann condition the integral on the boundary can be rearranged as 18 W m u n p n v G D m u n p n v v 0 on G D G N G N m u n p n h on G N v G N h v displaystyle int limits partial Omega left mu frac partial mathbf u partial hat mathbf n p hat mathbf n right cdot mathbf v underbrace int limits Gamma D left mu frac partial mathbf u partial hat mathbf n p hat mathbf n right cdot mathbf v mathbf v mathbf 0 text on Gamma D int limits Gamma N underbrace vphantom int limits Gamma N left mu frac partial mathbf u partial hat mathbf n p hat mathbf n right mathbf h text on Gamma N cdot mathbf v int limits Gamma N mathbf h cdot mathbf v Having this in mind the weak formulation of the Navier Stokes equations is expressed as 18 find u L 2 R H 1 W d C 0 R L 2 W d such that W r u t v W m u v W r u u v W p v W f v G N h v v V W q u 0 q Q displaystyle begin aligned amp text find mathbf u in L 2 left mathbb R left H 1 Omega right d right cap C 0 left mathbb R left L 2 Omega right d right text such that 5pt amp quad begin cases displaystyle int limits Omega rho dfrac partial mathbf u partial t cdot mathbf v int limits Omega mu nabla mathbf u cdot nabla mathbf v int limits Omega rho mathbf u cdot nabla mathbf u cdot mathbf v int limits Omega p nabla cdot mathbf v int limits Omega mathbf f cdot mathbf v int limits Gamma N mathbf h cdot mathbf v quad forall mathbf v in V displaystyle int limits Omega q nabla cdot mathbf u 0 quad forall q in Q end cases end aligned Discrete velocity Edit With partitioning of the problem domain and defining basis functions on the partitioned domain the discrete form of the governing equation is w i u j t w i u u j n w i u j w i f S displaystyle left mathbf w i frac partial mathbf u j partial t right bigl mathbf w i left mathbf u cdot nabla right mathbf u j bigr nu left nabla mathbf w i nabla mathbf u j right left mathbf w i mathbf f S right It is desirable to choose basis functions that reflect the essential feature of incompressible flow the elements must be divergence free While the velocity is the variable of interest the existence of the stream function or vector potential is necessary by the Helmholtz theorem Further to determine fluid flow in the absence of a pressure gradient one can specify the difference of stream function values across a 2D channel or the line integral of the tangential component of the vector potential around the channel in 3D the flow being given by Stokes theorem Discussion will be restricted to 2D in the following We further restrict discussion to continuous Hermite finite elements which have at least first derivative degrees of freedom With this one can draw a large number of candidate triangular and rectangular elements from the plate bending literature These elements have derivatives as components of the gradient In 2D the gradient and curl of a scalar are clearly orthogonal given by the expressions f f x f y T f f y f x T displaystyle begin aligned nabla varphi amp left frac partial varphi partial x frac partial varphi partial y right mathrm T 5pt nabla times varphi amp left frac partial varphi partial y frac partial varphi partial x right mathrm T end aligned Adopting continuous plate bending elements interchanging the derivative degrees of freedom and changing the sign of the appropriate one gives many families of stream function elements Taking the curl of the scalar stream function elements gives divergence free velocity elements 19 20 The requirement that the stream function elements be continuous assures that the normal component of the velocity is continuous across element interfaces all that is necessary for vanishing divergence on these interfaces Boundary conditions are simple to apply The stream function is constant on no flow surfaces with no slip velocity conditions on surfaces Stream function differences across open channels determine the flow No boundary conditions are necessary on open boundaries though consistent values may be used with some problems These are all Dirichlet conditions The algebraic equations to be solved are simple to set up but of course are non linear requiring iteration of the linearized equations Similar considerations apply to three dimensions but extension from 2D is not immediate because of the vector nature of the potential and there exists no simple relation between the gradient and the curl as was the case in 2D Pressure recovery Edit Recovering pressure from the velocity field is easy The discrete weak equation for the pressure gradient is g i p g i u u j n g i u j g i f I displaystyle mathbf g i nabla p left mathbf g i left mathbf u cdot nabla right mathbf u j right nu left nabla mathbf g i nabla mathbf u j right left mathbf g i mathbf f I right where the test weight functions are irrotational Any conforming scalar finite element may be used However the pressure gradient field may also be of interest In this case one can use scalar Hermite elements for the pressure For the test weight functions g i textstyle mathbf g i one would choose the irrotational vector elements obtained from the gradient of the pressure element Non inertial frame of reference EditThe rotating frame of reference introduces some interesting pseudo forces into the equations through the material derivative term Consider a stationary inertial frame of reference K textstyle K and a non inertial frame of reference K textstyle K which is translating with velocity U t textstyle mathbf U t and rotating with angular velocity W t textstyle Omega t with respect to the stationary frame The Navier Stokes equation observed from the non inertial frame then becomes Navier Stokes momentum equation in non inertial frame r D u D t p m 2 u 1 3 m u r g r 2 W u W W x d U d t d W d t x displaystyle rho frac mathrm D mathbf u mathrm D t nabla bar p mu nabla 2 mathbf u tfrac 1 3 mu nabla nabla cdot mathbf u rho mathbf g rho left 2 mathbf Omega times mathbf u mathbf Omega times mathbf Omega times mathbf x frac mathrm d mathbf U mathrm d t frac mathrm d mathbf Omega mathrm d t times mathbf x right Here x textstyle mathbf x and u textstyle mathbf u are measured in the non inertial frame The first term in the parenthesis represents Coriolis acceleration the second term is due to centrifugal acceleration the third is due to the linear acceleration of K textstyle K with respect to K textstyle K and the fourth term is due to the angular acceleration of K textstyle K with respect to K textstyle K Other equations EditThe Navier Stokes equations are strictly a statement of the balance of momentum To fully describe fluid flow more information is needed how much depending on the assumptions made This additional information may include boundary data no slip capillary surface etc conservation of mass balance of energy and or an equation of state Continuity equation for incompressible fluid Edit Main article Continuity equation Regardless of the flow assumptions a statement of the conservation of mass is generally necessary This is achieved through the mass continuity equation given in its most general form as r t r u 0 displaystyle frac partial rho partial t nabla cdot rho mathbf u 0 or using the substantive derivative D r D t r u 0 displaystyle frac mathrm D rho mathrm D t rho nabla cdot mathbf u 0 For incompressible fluid density along the line of flow remains constant over time D r D t 0 displaystyle frac mathrm D rho mathrm D t 0 Therefore divergence of velocity is always zero u 0 displaystyle nabla cdot mathbf u 0 Stream function for incompressible 2D fluid EditTaking the curl of the incompressible Navier Stokes equation results in the elimination of pressure This is especially easy to see if 2D Cartesian flow is assumed like in the degenerate 3D case with u z 0 textstyle u z 0 and no dependence of anything on z textstyle z where the equations reduce to r u x t u x u x x u y u x y p x m 2 u x x 2 2 u x y 2 r g x r u y t u x u y x u y u y y p y m 2 u y x 2 2 u y y 2 r g y displaystyle begin aligned rho left frac partial u x partial t u x frac partial u x partial x u y frac partial u x partial y right amp frac partial p partial x mu left frac partial 2 u x partial x 2 frac partial 2 u x partial y 2 right rho g x rho left frac partial u y partial t u x frac partial u y partial x u y frac partial u y partial y right amp frac partial p partial y mu left frac partial 2 u y partial x 2 frac partial 2 u y partial y 2 right rho g y end aligned Differentiating the first with respect to y textstyle y the second with respect to x textstyle x and subtracting the resulting equations will eliminate pressure and any conservative force For incompressible flow defining the stream function ps textstyle psi throughu x ps y u y ps x displaystyle u x frac partial psi partial y quad u y frac partial psi partial x results in mass continuity being unconditionally satisfied given the stream function is continuous and then incompressible Newtonian 2D momentum and mass conservation condense into one equation t 2 ps ps y x 2 ps ps x y 2 ps n 4 ps displaystyle frac partial partial t left nabla 2 psi right frac partial psi partial y frac partial partial x left nabla 2 psi right frac partial psi partial x frac partial partial y left nabla 2 psi right nu nabla 4 psi where 4 textstyle nabla 4 is the 2D biharmonic operator and n textstyle nu is the kinematic viscosity n m p textstyle nu frac mu p We can also express this compactly using the Jacobian determinant t 2 ps ps 2 ps y x n 4 ps displaystyle frac partial partial t left nabla 2 psi right frac partial left psi nabla 2 psi right partial y x nu nabla 4 psi This single equation together with appropriate boundary conditions describes 2D fluid flow taking only kinematic viscosity as a parameter Note that the equation for creeping flow results when the left side is assumed zero In axisymmetric flow another stream function formulation called the Stokes stream function can be used to describe the velocity components of an incompressible flow with one scalar function The incompressible Navier Stokes equation is a differential algebraic equation having the inconvenient feature that there is no explicit mechanism for advancing the pressure in time Consequently much effort has been expended to eliminate the pressure from all or part of the computational process The stream function formulation eliminates the pressure but only in two dimensions and at the expense of introducing higher derivatives and elimination of the velocity which is the primary variable of interest Properties EditNonlinearity Edit The Navier Stokes equations are nonlinear partial differential equations in the general case and so remain in almost every real situation 21 22 In some cases such as one dimensional flow and Stokes flow or creeping flow the equations can be simplified to linear equations The nonlinearity makes most problems difficult or impossible to solve and is the main contributor to the turbulence that the equations model The nonlinearity is due to convective acceleration which is an acceleration associated with the change in velocity over position Hence any convective flow whether turbulent or not will involve nonlinearity An example of convective but laminar nonturbulent flow would be the passage of a viscous fluid for example oil through a small converging nozzle Such flows whether exactly solvable or not can often be thoroughly studied and understood 23 Turbulence Edit Turbulence is the time dependent chaotic behaviour seen in many fluid flows It is generally believed that it is due to the inertia of the fluid as a whole the culmination of time dependent and convective acceleration hence flows where inertial effects are small tend to be laminar the Reynolds number quantifies how much the flow is affected by inertia It is believed though not known with certainty that the Navier Stokes equations describe turbulence properly 24 The numerical solution of the Navier Stokes equations for turbulent flow is extremely difficult and due to the significantly different mixing length scales that are involved in turbulent flow the stable solution of this requires such a fine mesh resolution that the computational time becomes significantly infeasible for calculation or direct numerical simulation Attempts to solve turbulent flow using a laminar solver typically result in a time unsteady solution which fails to converge appropriately To counter this time averaged equations such as the Reynolds averaged Navier Stokes equations RANS supplemented with turbulence models are used in practical computational fluid dynamics CFD applications when modeling turbulent flows Some models include the Spalart Allmaras k w k e and SST models which add a variety of additional equations to bring closure to the RANS equations Large eddy simulation LES can also be used to solve these equations numerically This approach is computationally more expensive in time and in computer memory than RANS but produces better results because it explicitly resolves the larger turbulent scales Applicability Edit Further information Discretization of Navier Stokes equations Together with supplemental equations for example conservation of mass and well formulated boundary conditions the Navier Stokes equations seem to model fluid motion accurately even turbulent flows seem on average to agree with real world observations The Navier Stokes equations assume that the fluid being studied is a continuum it is infinitely divisible and not composed of particles such as atoms or molecules and is not moving at relativistic velocities At very small scales or under extreme conditions real fluids made out of discrete molecules will produce results different from the continuous fluids modeled by the Navier Stokes equations For example capillarity of internal layers in fluids appears for flow with high gradients 25 For large Knudsen number of the problem the Boltzmann equation may be a suitable replacement 26 Failing that one may have to resort to molecular dynamics or various hybrid methods 27 Another limitation is simply the complicated nature of the equations Time tested formulations exist for common fluid families but the application of the Navier Stokes equations to less common families tends to result in very complicated formulations and often to open research problems For this reason these equations are usually written for Newtonian fluids where the viscosity model is linear truly general models for the flow of other kinds of fluids such as blood do not exist 28 Application to specific problems EditThe Navier Stokes equations even when written explicitly for specific fluids are rather generic in nature and their proper application to specific problems can be very diverse This is partly because there is an enormous variety of problems that may be modeled ranging from as simple as the distribution of static pressure to as complicated as multiphase flow driven by surface tension Generally application to specific problems begins with some flow assumptions and initial boundary condition formulation this may be followed by scale analysis to further simplify the problem Visualization of a parallel flow and b radial flow Parallel flow Edit Assume steady parallel one dimensional non convective pressure driven flow between parallel plates the resulting scaled dimensionless boundary value problem is d 2 u d y 2 1 u 0 u 1 0 displaystyle frac mathrm d 2 u mathrm d y 2 1 quad u 0 u 1 0 The boundary condition is the no slip condition This problem is easily solved for the flow field u y y y 2 2 displaystyle u y frac y y 2 2 From this point onward more quantities of interest can be easily obtained such as viscous drag force or net flow rate Radial flow Edit Difficulties may arise when the problem becomes slightly more complicated A seemingly modest twist on the parallel flow above would be the radial flow between parallel plates this involves convection and thus non linearity The velocity field may be represented by a function f z that must satisfy d 2 f d z 2 R f 2 1 f 1 f 1 0 displaystyle frac mathrm d 2 f mathrm d z 2 Rf 2 1 quad f 1 f 1 0 This ordinary differential equation is what is obtained when the Navier Stokes equations are written and the flow assumptions applied additionally the pressure gradient is solved for The nonlinear term makes this a very difficult problem to solve analytically a lengthy implicit solution may be found which involves elliptic integrals and roots of cubic polynomials Issues with the actual existence of solutions arise for R gt 1 41 textstyle R gt 1 41 approximately this is not 2 the parameter R textstyle R being the Reynolds number with appropriately chosen scales 29 This is an example of flow assumptions losing their applicability and an example of the difficulty in high Reynolds number flows 29 Convection Edit A type of natural convection that can be described by the Navier Stokes equation is the Rayleigh Benard convection It is one of the most commonly studied convection phenomena because of its analytical and experimental accessibility Exact solutions of the Navier Stokes equations EditSome exact solutions to the Navier Stokes equations exist Examples of degenerate cases with the non linear terms in the Navier Stokes equations equal to zero are Poiseuille flow Couette flow and the oscillatory Stokes boundary layer But also more interesting examples solutions to the full non linear equations exist such as Jeffery Hamel flow Von Karman swirling flow stagnation point flow Landau Squire jet and Taylor Green vortex 30 31 32 Note that the existence of these exact solutions does not imply they are stable turbulence may develop at higher Reynolds numbers Under additional assumptions the component parts can be separated 33 A two dimensional exampleFor example in the case of an unbounded planar domain with two dimensional incompressible and stationary flow in polar coordinates r f the velocity components ur uf and pressure p are 34 u r A r u f B 1 r r A n 1 p A 2 B 2 2 r 2 2 B 2 n r A n A B 2 r 2 A n 2 2 A n 2 displaystyle begin aligned u r amp frac A r u varphi amp B left frac 1 r r frac A nu 1 right p amp frac A 2 B 2 2r 2 frac 2B 2 nu r frac A nu A frac B 2 r left frac 2A nu 2 right frac 2A nu 2 end aligned where A and B are arbitrary constants This solution is valid in the domain r 1 and for A lt 2n In Cartesian coordinates when the viscosity is zero n 0 this is v x y 1 x 2 y 2 A x B y A y B x p x y A 2 B 2 2 x 2 y 2 displaystyle begin aligned mathbf v x y amp frac 1 x 2 y 2 begin pmatrix Ax By Ay Bx end pmatrix p x y amp frac A 2 B 2 2 left x 2 y 2 right end aligned A three dimensional exampleFor example in the case of an unbounded Euclidean domain with three dimensional incompressible stationary and with zero viscosity n 0 radial flow in Cartesian coordinates x y z the velocity vector v and pressure p are citation needed v x y z A x 2 y 2 z 2 x y z p x y z A 2 2 x 2 y 2 z 2 displaystyle begin aligned mathbf v x y z amp frac A x 2 y 2 z 2 begin pmatrix x y z end pmatrix p x y z amp frac A 2 2 left x 2 y 2 z 2 right end aligned There is a singularity at x y z 0 A three dimensional steady state vortex solution Edit Wire model of flow lines along a Hopf fibration A steady state example with no singularities comes from considering the flow along the lines of a Hopf fibration Let r textstyle r be a constant radius of the inner coil One set of solutions is given by 35 r x y z 3 B r 2 x 2 y 2 z 2 p x y z A 2 B r 2 x 2 y 2 z 2 3 u x y z A r 2 x 2 y 2 z 2 2 2 r y x z 2 r x y z r 2 x 2 y 2 z 2 g 0 m 0 displaystyle begin aligned rho x y z amp frac 3B r 2 x 2 y 2 z 2 p x y z amp frac A 2 B left r 2 x 2 y 2 z 2 right 3 mathbf u x y z amp frac A left r 2 x 2 y 2 z 2 right 2 begin pmatrix 2 ry xz 2 rx yz r 2 x 2 y 2 z 2 end pmatrix g amp 0 mu amp 0 end aligned for arbitrary constants A textstyle A and B textstyle B This is a solution in a non viscous gas compressible fluid whose density velocities and pressure goes to zero far from the origin Note this is not a solution to the Clay Millennium problem because that refers to incompressible fluids where r textstyle rho is a constant and neither does it deal with the uniqueness of the Navier Stokes equations with respect to any turbulence properties It is also worth pointing out that the components of the velocity vector are exactly those from the Pythagorean quadruple parametrization Other choices of density and pressure are possible with the same velocity field Other choices of density and pressureAnother choice of pressure and density with the same velocity vector above is one where the pressure and density fall to zero at the origin and are highest in the central loop at z 0 x2 y2 r2 r x y z 20 B x 2 y 2 r 2 x 2 y 2 z 2 3 p x y z A 2 B r 2 x 2 y 2 z 2 4 4 A 2 B x 2 y 2 r 2 x 2 y 2 z 2 5 displaystyle begin aligned rho x y z amp frac 20B left x 2 y 2 right left r 2 x 2 y 2 z 2 right 3 p x y z amp frac A 2 B left r 2 x 2 y 2 z 2 right 4 frac 4A 2 B left x 2 y 2 right left r 2 x 2 y 2 z 2 right 5 end aligned In fact in general there are simple solutions for any polynomial function f where the density is r x y z 1 r 2 x 2 y 2 z 2 f x 2 y 2 r 2 x 2 y 2 z 2 2 displaystyle rho x y z frac 1 r 2 x 2 y 2 z 2 f left frac x 2 y 2 left r 2 x 2 y 2 z 2 right 2 right Viscous three dimensional periodic solutions Edit Two examples of periodic fully three dimensional viscous solutions are described in 36 These solutions are defined on a three dimensional torus T 3 0 L 3 displaystyle mathbb T 3 0 L 3 and are characterized by positive and negative helicity respectively The solution with positive helicity is given by u x 4 2 3 3 U 0 sin k x p 3 cos k y p 3 sin k z p 2 cos k z p 3 sin k x p 3 sin k y p 2 e 3 n k t u y 4 2 3 3 U 0 sin k y p 3 cos k z p 3 sin k x p 2 cos k x p 3 sin k y p 3 sin k z p 2 e 3 n k t u z 4 2 3 3 U 0 sin k z p 3 cos k x p 3 sin k y p 2 cos k y p 3 sin k z p 3 sin k x p 2 e 3 n k t displaystyle begin aligned u x amp frac 4 sqrt 2 3 sqrt 3 U 0 left sin kx pi 3 cos ky pi 3 sin kz pi 2 cos kz pi 3 sin kx pi 3 sin ky pi 2 right e 3 nu kt u y amp frac 4 sqrt 2 3 sqrt 3 U 0 left sin ky pi 3 cos kz pi 3 sin kx pi 2 cos kx pi 3 sin ky pi 3 sin kz pi 2 right e 3 nu kt u z amp frac 4 sqrt 2 3 sqrt 3 U 0 left sin kz pi 3 cos kx pi 3 sin ky pi 2 cos ky pi 3 sin kz pi 3 sin kx pi 2 right e 3 nu kt end aligned where k 2 p L displaystyle k 2 pi L is the wave number and the velocity components are normalized so that the average kinetic energy per unit of mass is U 0 2 2 displaystyle U 0 2 2 at t 0 displaystyle t 0 The pressure field is obtained from the velocity field as p p 0 r 0 u 2 2 displaystyle p p 0 rho 0 boldsymbol u 2 2 where p 0 displaystyle p 0 img, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.