Você está na página 1de 33

VII.

Derivation of the Navier-Stokes Equations and Solutions In this chapter, we will derive the equations governing 2-D, unsteady, compressible viscous flows. These equations (and their 3-D form) are called the Navier-Stokes equations. They were developed by Navier in 1831, and more rigorously be Stokes in 1845. Now, over 150 years later, these equations still stand with no modifications, and form the basis of all simpler forms of equations such as the potential flow equations that were derived in Chapter I. In two dimensions, we have five flow properties that are unknowns: the two velocity components u,v; density r, temperature T and pressure p. Therefore, we need 5 equations linking them. One of these 5 equations is the equation of state, given by

At moderate temperatures that arise in subsonic and supersonic flows without chemical reactions, this equation of state may be simplified to the following form:

Here R is a gas constant, given by R/M, where R is the universal gas constant, and M is the molecular weight of the gas (or the gas mixture). For air, the gas constant is given by R= 2817 Joules/kg/ K. The other four equations are: a) Conservation of mass, known as continuity, b) Conservation of u- momentum c) Conservation of v- momentum, and d) Conservation of energy. Conservation of Mass We consider a small control volume (CV) of height Dy, width Dx, and of depth unity perpendicular to the plane of the paper.

The principle of conservation of mass states that "The rate at which mass increases within the control volume = The rate at which mass enters the control volume through its four boundaries" Let r be the average density of the fluid within the control volume. Then,

Next, consider the rate at which mass enters through the four boundaries, one by one. Consider the boundary #1, first.

We can assume that the above flux is computed at the center of face #1. We can consider the other three boundaries in a similar manner. The rate at which mass enters through faces 2,3 and 4 are, respectively (-ruDy)2 , (+(rvDx)3 and (-rvDx)4. Here the subscripts refer to the face. Summing up the contributions from the four faces, and equating the result to the time rate of change of mass within the CV, we get

Now, consider the limits of the above equation as Dx and Dy goes to zero. From calculus, for any arbitrary function f(x,y),

Applying the above limits, and bringing all the terms to the left hand side, we get

The above equation, in vector form is given by:

The vector form is more useful than it would first appear. If we want to derive the continuity equation in another coordinate system such as the polar, cylindrical or spherical coordinate system, all we need to know is (a) look up the 'Del' operator in that system, (b) look up the rules for the dot product of 'Del' operator and a vector in that system, (c) perform the dot product. Conservation of u- Momentum Equation Before we can proceed any further, we need to get a firm understanding of terms such as viscosity, viscous stresses, conductivity, etc. Consider the left face of the control volume considered earlier. The air molecules to the left of this CV can interact with our CV in one of three ways: (i) organized motion from left to right. While the molecules are constantly moving about back and forth, over a small period of time, the majority of these molecules either enter the control volume (u >0) or leave the control volume. This "average" over a period of time is called the flow velocity component u, and is measured by probes such as LDVs and hot wires. (ii) Exchange of u- momentum between the molecules on the left and those on the right by collisions. In this case, there no net gain in mass, but there is a gain (or loss) in momentum. These collision effects may be averaged over a sufficiently small period of time, and may be viewed as a pressure force exerted by the fluid on the left on our CV. Again, only this average effect is felt or measured by pressure probes, and barometers. The individual collisions occur far too rapidly and far too frequently to be sensed by probes or measuring devices.

(iii) _Exchange of u- and v- momentum by random linear motion of molecules jumping in and out of our control volume, across the face 1. In this case, all the molecules that jumped in also jump out over a sufficiently small time period. Thus, this random motion does not add mass to our control volume (and was not considered in our "continuity" equation). They however bring uand v-momentum in or out (associated with their random motion). The time averages of these rates at which u- and vmomentum is brought into the CV across a face are called viscous forces. The forces per unit area are called viscous stresses. The viscous stresses that bring in/out u- momentum are called normal viscous stresses, while those that bring in v- momentum (by entering the face at an angle) are called tangential viscous stresses. By convention, pressure forces are considered positive, if they act towards the fluid element, or control volume. The normal viscous stresses (following solid mechanics conventions) are considered positive if they act away from the control volume, producing a tension.

These stresses (normal, and tangential or shear) are given the symbol t. They are identified by two subscripts. (i) The first subscript indicates the plane on which they act. For example, if a plane is normal to the x- axis, the first subscript will be x. (ii) The second subscript identifies the direction of the force associated with the force. For example, if a shear force is pointing in the y- direction, the second subscript will be y. Thus, viscous stress is a tensor quantity, and requires three pieces of information (its magnitude, its direction and the plane on which it acts) to completely specify it. This separates a tensor from a vector (magnitude and direction), and a scalar (magnitude only).

Newtonian Fluids Because our primary unknowns are the flow properties (u,v,p,r,T) there is a need to link the stresses t with these physical variables. In solid mechanics (Hooke's law) stress is set proportional to strain. This works for solids because a solid undergoes only a finite amount of deformation when a force or stress is applied to it. In fluid mechanics, this approach does not work because fluid continuously deforms when a shear stress is applied. It is this characteristic that distinguishes a fluid from a solid.

Newton came up with the idea of requiring the stress t to be linearly proportional to the time rate at which strain occurs. Specifically he studied the following problem. There are two flat plates separated by a distance 'h'. The top plate is moved at a velocity V, while the bottom plate is held fixed. Newton postulated (since then experimentally verified) that the shear force or shear stress needed to deform the fluid was linearly proportional to the velocity gradient:

The proportionality factor turned out to be a constant at moderate temperatures, and was called the coefficient of viscosity, m. Furthermore, for this particular case, the velocity profile is linear, giving V/h = u/y. Therefore, Newton postulated:

Fluids that have a linear relationship between stress and strain rate are called Newtonian fluids. This is a property of the fluid, not the flow. Water and air are examples of Newtonian fluids, while blood is a non-Newtonian fluid. Stokes Hypothesis:

Stokes extended Newton's idea from simple 1-D flows (where only one component of velocity is present) to multidimensional flows. here, the fluid element may experience a strain rate both due to gradients such as u/y as well as v/x. He developed the following relations, collectively known as Stokes relations.

These expressions hold for 3-D flows. For 2-D flows, somewhat simpler expressions are obtained if we set w, the z- component of velocity, to zero, and if we set all derivatives with respect to z to be zero. The quantity m is called the molecular viscosity, and is a weak function of temperature. For air viscosity increases with temperature, because viscous effects are associated with random molecular motion. The coefficient l was chosen by Stokes so that the sum of the normal stresses txx, tyy and tzz are zero. Then

The above equation, and the requirement that the three normal stresses add up to zero are called Stokes hypothesis. Returning Back to u- Momentum Equation.... We now return to the derivation of the u- momentum equation. This equation is a generalization of Newton's second law of motion, and may be verbally stated as: "The rate of change of u- momentum within a control volume is equal to the net rate at which u- momentum enters the control volume + Forces (pressure, viscous and body) acting on the control volume in the x- direction" As before we consider the control volume of height Dy and width Dx. We neglect body forces such as gravity, electrical and electromagnetic effects.

Summing up these contributions, dividing through DxDy and taking the limits as Dx and Dy go to zero, we get:

Derivation of v- Momentum Equation: The v- momentum equation may be derived using a logic identical to that used above, and is left as an exercise to the student. The final form is:

Derivation of the Energy Equation:

The energy equation is a generalized form of the first law of Thermodynamics (that you studied in ME3322 and AE 3004). The only difference here is that we are studying an open system (i.e. control volume) that can gain and lose mass. The classical form studied in courses on thermodynamics is applicable only for closed systems - i.e. fixed collection of particles. The first law of Thermodynamics states that "(A) The rate at which the total (i.e. internal + kinetic) energy increases within a control volume is equal to (B) the rate at which total energy enters the control volume + (C) the rate at which work is done on the control volume boundary by surface forces + (D) the rate at which work is done on the CV by body forces + (E) the rate at which heat is added to the control volume at the surfaces by heat conduction + (F) the rate at which heat is released is added within the CV due to chemical reactions." In our derivation, we will neglect the term (D) which corresponds to the work done by the body forces (such as gravity, electrical and electromagnetic forces) and (F) which corresponds to chemical reactions. In some applications, e.g. modeling of weather, term D is important, while in others (e.g. modeling combustors) term F is important. We define the specific total energy (i.e. total energy per unit mass) E as

In the above equation 'e' is specific internal energy. The word 'specific' means 'per unit mass'. Term A: Then, term A may be expressed as follows:

Term B: Term B involves mass entering and leaving the four boundaries of the CV, carrying with it, the total energy E. This term may be expressed as:

As in the continuity equation, as Dx and Dy go to zero, the terms within the square brackets will go to -(ruE)/x and -(rvE)/y, respectively. Term C: At the surfaces of the control volume, there are pressure and viscous forces that do work on the fluid, as the fluid particles cross the boundary. For examples, on face #1, the pressure forces and viscous forces may be expressed as

. The fluid velocity at this boundary is forces acting on face #1 is, then,

. The rate at which work is done by the surface

Summing up contributions from all the four faces, we get ,

Term D: The work done by body forces is neglected here. Term E: The heat conduction effects are associated with the random motion of gas molecules across the control volume. As they move in and out, they bring energy into and out of the control volume. When integrated over a small but finite period of time, a net exchange of heat energy occurs at the boundary, without any exchange in mass. Since this process is a random, chaotic process, it must somehow be empirically modeled. We adapt Fourier's law used to model conduction of heat through solids, which states "The rate at which heat flows across a surface of unit area is proportional to the negative of the temperature gradient normal to this surface". The constant of proportionality is called conductivity k, a property of the solid (or fluid, in the present context). Note that the heat flux is proportional to the negative of the gradient, because heat flows from hot to cold. We can sum up the heat conduction effects at all the four boundaries. The result is,

Summing up all the terms A through E, and dropping the common factor DxDy that appears everywhere, we get the energy equation.

Prandtl Number: The coefficient of conductivity k is a property of the fluid, like the molecular viscosity. It is also a weak function of temperature and so is the molecular viscosity. Both these empirical coefficients are associated with transport of energy and momentum by random molecular motion. It follows that these two terms should be, somehow, related. Indeed, the following simple relationship exists.

Here, Cp is the specific heat at constant pressure. The quantity Pr is called the Prandtl number, and is a property of the fluid. For air Pr is around 0.72 (close to unity). It is a measure of the ratio between the viscous effects and conduction effects. In some highly conductive, but only slightly viscous fluids such as mercury, Pr is very small. In other fluids such as honey which is highly viscous, but only slightly conductive, Pr is very high. Sutherland's Law: For air, the variation of viscosity (and hence conductivity) with temperature may be empirically described y the Sutherland's Law which states:

where m0 denotes the viscosity at the reference temperature T0, and S1 is a constant. For air, S1 assumes the value 110 degrees kelvin.

Summary The 2-D unsteady Navier-Stokes equations may be written in a number of forms. One common form of these equations is as follows: Continuity:

u- Momentum:

v-Momentum:

Energy Equation:

Here.

= density; u,v = Cartesian Components of velocity along x,y axes;


p = Pressure ; T = Temperature. Also, e = Specific Internal Energy = Internal energy per unit mass of the fluid = CvT , where Cv is the specific heat at constant volume. h0 = Specific Total Enthalpy = Total enthalpy per unit mass of the fluid = CpT+(u2+v2)/2 , where Cp is the specific heat at constant pressure. Finally, the viscous stresses are related to the velocity field by Stokes relations

The molecular viscosity m and conductivity k are properties of the fluid and are functions of temperature. These two quantities are related by the Prandtl number

For air, the Prandtl number is around 0.72 at room temperatures. Simplified Form for Steady, 2-D Incompressible Flows: In steady, incompressible flows, we can drop the time derivatives because the flow is steady. The density may be also assumed constant. Then, the first three equations in the full Navier-Stokes equations set become:

where n is called the kinematic viscosity= /

Nondimensionalization of the Viscous Flow Equations Consider the 2-D viscous flow past an airfoil shown below:

At first glance it appears that such a flow will depend on a large number of parameters such as (a) airfoil shape, including its chord c ,(b) its angle of attack, (c) the freestream temperature T (d) the freestream density r , (e) the freestream velocity V, (f) the freestream viscosity m, (g) the freestream conductivity k, and so on. We wish to know if these large number of physical and geometric parameters can be grouped into a handful of parameters that can be systematically varied to study their effect on the flow. There are two common ways of identifying these parameters: (i) Dimensional analysis: Here we attempt to combine the parameters listed about to arrive at a nondimensional form. For example, after some trial and error, we can show that the quantity rVc/m is a nondimensional quantity. Intuition and experience are needed to realize that this nondimensional parameter, called Reynolds number, is also a useful physical parameter. Dimensional analysis will also produce combinations such as nondimensional, is not a significant parameter in incompressible flows, as experience shows. . This parameter, while

(ii) Nondimensionalization of Governing Equations: This approach provides a formal manner by which nondimensional parameters of physical significance may be uncovered. This approach is usually used in combination with the dimensional analysis shown above. To illustrate how this approach works, let us consider the 2- flow over an airfoil shown above. Let us introduce nondimensional quantities identified with a prime (') as follows:

When these nondimensional quantities are used to replace the corresponding physical quantities in the continuity equation the following form results:

This form looks identical to the dimensional form given in Handout #1, and no new nondimensional parameter emerges. If we repeat this process with the u- momentum equation, and replace terms such as txy and p with terms such as

we get, after some minor algebra:

If we repeat this process with the v- momentum equation, again the Reynolds number emerges as the nondimensional parameter. Consider two different flows over the same airfoil, but of different chord. The present normalization shows that these two flows are governed by the same nondimensional form of the governing equations, and will result in the same nondimensional flow quantities such as u', v' etc. if the flows are geometrically similar (same airfoil shape, same a), and dynamically similar (same Reynolds number). The parameter Reynolds number may also be interpreted in another way. Consider a typical inviscid term such as p appearing on the left side of the u- momentum equation. This "inertial" force term is roughly of order . Consider a viscous stress

term such as txy. This term is of order . Then, Reynolds number is a measure of the ratio between the inertial forces and viscous forces acting on a fluid element. If this number is large, then inertial forces dominate over viscous forces and vice versa. In practical aeronautical applications, the Reynolds number is invariably large. For example, the airfoil over a helicopter rotor blade operates at a Reynolds number of 3 to 5 Million, while the airfoil in the Root section of a modern transport aircraft operates at a Reynolds number of 30 to 50 Million. In such flows, viscous forces are small and may be neglected over most of the flow, except in very small regions called boundary layers over the airfoil. Nondimensionalization of the Energy Equation: If we nondimensionalize the energy equation by the same principles, we get, in addition to the Reynolds number, the parameter . This parameter occurs due to the presence of the heat conduction terms and viscous stress work terms in the energy equation. This parameter is called the Prandtl number, and is a property of the fluid. For air, Prandtl number is around 0.72.

In compressible flows, the quantity is a nontrivial quanity and must be prescribed. Using equation of state, the definition of the speed of sound and the definition of Mach number, we can relate this quantity to the Mach number and the ratio of specific heats, g.

Thus, in compressible flows, two flows are identical if they are geometrically similar and if Reynolds number, Mach number, Prandtl number and ratio of specific heats all match.

In incompressible flows, this quantity is not a significant parameter. As hydraulics engineers will attest, raising the static pressure at some point in the flow simply raises the pressure level everywhere by the same constant level, and does not alter the flow behavior. Some Exact Solutions of the Navier-Stokes Equations We next turn our attention to the exact solutions of incompressible Navier-Stokes equations. Even though these equations were derived over a century ago, only a handful of exact solutions exist for some highly simplified situations. This is because of the nonlinear nature of these equations. The governing equations are:

In the above equations, we have assumed the conductivity k and viscosity m to be constant. We will further restrict ourselves to steady flows (/t = 0) , and incompressible flows (r = constant). The above form of equations apply only to 2-D planar flows. Similar forms of Navier-Stokes equations exist in other coordinate systems such as the cylindrical, polar and spherical coordinate systems. Parallel Flows The first class of flows to be considered is the flow between (or within) infinitely long parallel plates, and infinitely long tubes.

In such flows, it is reasonable to assume that the flow behavior will be independent of the x- location. That is, x-derivatives of all flow properties such as u/x, v/x, T/x etc. vanish. The pressure derivative p/x is assumed to be a constant, required to drive the flow. In such as situation, continuity equation becomes

Setting u/x=0 yields v/y = 0 or (rv)/r=0. The assumption v/x=0 implies that the velocity component v is not a function of x. Thus, continuity yields v = constant, for our flows. Since the boundary condition requires v to be zero at the walls of our parallel plates, and at the walls of the tube, continuity + boundary conditions together yield

In other words, the flow only has a u- component and is parallel to the x- axis. For this reason, the first class of flows we study are called parallel flows. Exact Solution # 1: Planar Couette Flow: In 1890 Couette performed experimental studies of flow between two concentric cylinders, where one of the cylinder is fixed, the other is spinning. This situation is similar to what happens in a ball bearing. here we study the 2-D analog of flow between two parallel plates. One of the plate is held fixed, while the other one is moving at a velocity V. The plates are separated by a distance h. A constant pressure gradient dp/dx is applied to this flow.

For this case, the u- momentum equation yields

Setting the x- derivatives to zero, and requiring v to ve zero, the above equation may be simplified to yield

Notice that we have begun to replace partial derivatives with respect to y with ordinary derivatives, because flow properties are only functions of y. Integrating the above equation twice, we get:

where A and B are constants of integration. These may be evaluated by requiring u=0 at y=0 and u= V at y=h.

Two special situations are of interest. In the first case, the pressure gradient dp/dx=0 is zero, and the flow motion is brought about by the motion of the top plate at the constant velocity V. In that case, the velocity profile is linear, giving u(y) = V y/h In the second situation, both the plates are at rest and V=0. the fluid is motion is caused by the application of pressure gradient dp/dx. In this case, the velocity profile is a parabola, given by

The shear stress txy(i.e. the force per unit area exerted by the fluid on the flat plates, and vice versa) may be computed from Stokes' relations. For example, for the case where dp/dx=0, we get

The non-dimensional "skin friction" coefficient Cf is then computed as

Temperature Distribution for Planar Couette Flow:

The energy equation, which is part of the Navier-Stokes equations is usually not solved in incompressible flow applications, unless we are interested in a heat transfer application. Let us consider the situation where dp/dx=0, the top plate is moving at the velocity V, and the pressure gradient dp/dx=0. Let the bottom and top plates be at two different temperatures T1 and T2 respectively. In this case, we can easily solve for the temperature distribution in the fluid between the plates, and compute the rate at which heat is transferred from the hot plate to the cold plate, as follows.

The energy equation for this case is (Verify for yourself, starting with the form given in Handout #1):

For our parallel flow, we can set /x =0 and v = 0. For the case where dp/dx=0, the velocity profile is linear, as derived earlier. Then,

When these expressions are substituted into the energy equation, and when all x-derivatives are set to zero, and when v is set to zero, the following form results.

Integrating the above equation twice, we get

where C and D are constants of integration. These constants may be found by applying the boundary condition T= T1 at y=0, and T = T2 at y= h. The final form is

The first two terms in the above expression are linear in y, and model the effect of conduction on the temperature distribution. The third term models the effect of heat geeration due to viscous work. Since m is small, this term is likely to be of significance only in high speed flows.

pressure term.

In physics, the NavierStokes equations, named after Claude-Louis Navier and George Gabriel Stokes, describe the motion of fluid substances. These equations arise from applying Newton's second law to fluid motion, together with the assumption that the fluid stress is the sum of a diffusing viscous term (proportional to the gradient of velocity), plus a

The equations are useful because they describe the physics of many things of academic and economic interest. They may be used to model the weather, ocean currents, water flow in a pipe and air flow around a wing. The NavierStokes 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 NavierStokes equations are also of great interest in a purely mathematical sense. Somewhat surprisingly, given their wide range of practical uses, mathematicians have not yet proven that in three dimensions solutions always exist (existence), or that if they do exist, then they do not contain any singularity (smoothness). These are called the Navier Stokes existence and smoothness problems. The Clay Mathematics Institute has called this one of the seven most important open problems in mathematics and has offered a US$1,000,000 prize for a solution or a counter-example.[1] The NavierStokes equations dictate not position but rather velocity. A solution of the NavierStokes equations is called a velocity field or flow field, which is a description of the velocity of the fluid at a given point in space and time. Once the velocity field is solved for, other quantities of interest (such as flow rate or drag force) may be found. 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; however for visualization purposes one can compute various trajectories.

[ [edit] Nonlinearity The NavierStokes equations are nonlinear partial differential equations in almost every real situation. 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. [edit] Turbulence Turbulence is the time dependent chaotic behavior 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 NavierStokes equations describe turbulence properly. The numerical solution of the NavierStokes 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 (see 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, timeaveraged equations such as the Reynolds-averaged NavierStokes equations (RANS), supplemented with turbulence models, are used in practical computational fluid dynamics (CFD) applications when modeling turbulent flows. Some models include the SpalartAllmaras, k- (k-omega), k- (k-epsilon), and SST models which add a variety of additional equations to bring closure to the RANS equations. Another technique for solving numerically the NavierStokes equation is the Large eddy simulation (LES). This approach is computationally more expensive than the RANS method (in time and computer memory), but produces better results since the larger turbulent scales are explicitly resolved. [edit] Applicability Together with supplemental equations (for example, conservation of mass) and well formulated boundary conditions, the NavierStokes equations seem to model fluid motion accurately; even turbulent flows seem (on average) to agree with real world observations. The NavierStokes 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 NavierStokes equations. Depending on the Knudsen number of the problem, statistical mechanics or possibly even molecular dynamics may be a more appropriate approach. Another limitation is simply the complicated nature of the equations. Time tested formulations exist for common fluid families, but the application of the NavierStokes equations to less common families tends to result in very complicated formulations which are an area of current research. For this reason, these equations are usually written for Newtonian fluids. Studying such fluids is "simple" because the viscosity model ends up being linear; truly general models for the flow of other kinds of fluids (such as blood) do not, as of 2011, exist. [edit] Derivation and description Main article: Derivation of the NavierStokes equations The derivation of the NavierStokes equations begins with an application of Newton's second law: conservation of momentum (often alongside mass and energy conservation) being written for an arbitrary portion of the fluid. In an inertial frame of reference, the general form of the equations of fluid motion is:[2]

where

is the flow velocity, is the fluid density, p is the pressure,

is the (deviatoric)

stress tensor, and represents body forces (per unit volume) acting on the fluid and is the del operator. This is a statement of the conservation of momentum in a fluid and it is an application of Newton's second law to a continuum; in fact this equation is applicable to any non-relativistic continuum and is known as the Cauchy momentum equation. This equation is often written using the material derivative Dv/Dt, making it more apparent that this is a statement of Newton's second law:

The left side of the equation describes acceleration, and may be composed of time dependent or convective effects (also the effects of non-inertial coordinates if present). The right side of the equation is in effect a summation of body forces (such as gravity) and divergence of stress (pressure and shear stress). [edit] 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 very significant feature of the NavierStokes equations is the presence of convective acceleration: the effect of time independent acceleration of a fluid with respect to space. While individual fluid particles are indeed experiencing time dependent acceleration, the convective acceleration of the flow field is a spatial effect, one example being fluid speeding up in a nozzle. Convective acceleration is represented by the nonlinear quantity:

which may be interpreted either as derivative of the velocity vector

or as

with

the tensor

Both interpretations give the same result, independent is interpreted as the covariant derivative.[3]

of the coordinate system provided [edit] Interpretation as (v)v The convection term is often written as

where the advection operator

is used. Usually this representation is preferred as it is [3]

simpler than the one in terms of the tensor derivative [edit] Interpretation as v(v)

Here is the tensor derivative of the velocity vector, equal in Cartesian coordinates to the component by component gradient. The convection term may, by a vector calculus identity, be expressed without a tensor derivative:[4][5]

The form has use in irrotational flow, where the curl of the velocity (called vorticity) is equal to zero. Regardless of what kind of fluid is being dealt with, convective acceleration is a nonlinear effect. Convective acceleration is present in most flows (exceptions include onedimensional incompressible flow), but its dynamic effect is disregarded in creeping flow (also called Stokes flow) . [edit] Stresses The effect of stress in the fluid is represented by the and terms; these are gradients

of surface forces, analogous to stresses in a solid. is called the pressure gradient and arises from the isotropic part of the stress tensor. This part is given by normal stresses that turn up in almost all situations, dynamic or not. The anisotropic part of the stress tensor gives rise to , which conventionally describes viscous forces; for incompressible is the deviatoric stress tensor, and the stress

flow, this is only a shear effect. Thus, tensor is equal to:[6]

where is the 33 identity matrix. Interestingly, only the gradient of pressure matters, not the pressure itself. The effect of the pressure gradient is that fluid flows from high pressure to low pressure. The stress terms p and are yet unknown, so the general form of the equations of motion is not usable to solve problems. Besides the equations of motionNewton's second lawa force model is needed relating the stresses to the fluid motion.[7] For this reason, assumptions on the specific behavior of a fluid are made (based on natural observations) and applied in order to specify the stresses in terms of the other flow variables, such as velocity and density. The NavierStokes equations result from the following assumptions on the deviatoric stress tensor :[8]

the deviatoric stress vanishes for a fluid at rest, and by Galilean invariance also does not depend directly on the flow velocity itself, but only on spatial derivatives of the flow velocity in the NavierStokes equations, the deviatoric stress is expressed as the product of the tensor gradient of the flow velocity with a viscosity tensor , i.e. :

the fluid is assumed to be isotropic, as valid for gases and simple liquids, and consequently is an isotropic tensor; furthermore, since the deviatoric stress tensor is symmetric, it turns out that it can be expressed in terms of two scalar dynamic viscosities and :

where and is the rate of expansion of the flow

is the rate-of-strain tensor

the deviatoric stress tensor has zero trace, so for a three-dimensional flow 2 + 3 = 0 As a result, in the NavierStokes equations the deviatoric stress tensor has the following form:[8]

with the quantity between brackets the non-isotropic part of the rate-of-strain tensor The dynamic viscosity does not need to be constant in general it depends on conditions like temperature and pressure, and in turbulence modelling the concept of eddy viscosity is used to approximate the average deviatoric stress. The pressure p is modelled by use of an equation of state.[9] For the special case of an incompressible flow, the pressure constrains the flow in such a way that the volume of fluid elements is constant: isochoric flow resulting in a solenoidal velocity field with [edit] Other forces The vector field represents body forces. Typically these consist of only gravity forces, but may include other types(such as electromagnetic forces). In a non-inertial coordinate system, other "forces" such as that associated with rotating coordinates may be inserted. Often, these forces may be represented as the gradient of some scalar quantity F, with . Gravity in the z direction, for example, is the gradient of gz. Since pressure shows up only as a gradient, this implies that solving a problem without any such body force can be mended to include the body force by using a modified pressure . The pressure and force terms on the right hand side of the Navier Stokes equation become [10]

[edit] Other equations The NavierStokes equations are strictly a statement of the conservation of momentum. In order to fully describe fluid flow, more information is needed (how much depends on the assumptions made). This additional information may include boundary data (no-slip, capillary surface, etc.), the conservation of mass, the conservation of energy, and/or an equation of state. 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:

[edit] Incompressible flow of Newtonian fluids A simplification of the resulting flow equations is obtained when considering an incompressible flow of a Newtonian fluid. The assumption of incompressibility rules out the possibility of sound or shock waves to occur; so this simplification is invalid if these phenomena are important. The incompressible flow assumption typically holds well even when dealing with a "compressible" fluid such as air at room temperature at low Mach numbers (even when flowing up to about Mach 0.3). Taking the incompressible flow assumption into account and assuming constant viscosity, the NavierStokes equations will read, in vector form:[11]

Here f represents "other" body forces (forces per unit volume), such as gravity or centrifugal force. The shear stress term becomes the useful quantity ( is the vector Laplacian) when the fluid is assumed incompressible, homogeneous and Newtonian, where is the (constant) dynamic viscosity.[12] It's well worth observing the meaning of each term (compare to the Cauchy momentum equation):

Note that only the convective terms are nonlinear for incompressible Newtonian flow. The convective acceleration is an acceleration caused by a (possibly steady) change in velocity over position, for example the speeding up of fluid entering a converging nozzle. Though individual fluid particles are being accelerated and thus are under unsteady motion, the flow field (a velocity distribution) will not necessarily be time dependent. Another important observation is that the viscosity is represented by the vector Laplacian of the velocity field (interpreted here as the difference between the velocity at a point and the mean velocity in a small volume around). This implies that Newtonian viscosity is diffusion of momentum, this works in much the same way as the diffusion of heat seen in the heat equation (which also involves the Laplacian). If temperature effects are also neglected, the only "other" equation (apart from initial/boundary conditions) needed is the mass continuity equation. Under the incompressible assumption, density is a constant and it follows that the equation will simplify to:

This is more specifically a statement of the conservation of volume (see divergence). These equations are commonly used in 3 coordinates systems: Cartesian, cylindrical, and spherical. While the Cartesian equations seem to follow directly from the vector equation above, the vector form of the NavierStokes equation involves some tensor calculus which means that writing it in other coordinate systems is not as simple as doing so for scalar equations (such as the heat equation). [edit] Cartesian coordinates Writing the vector equation explicitly,

Note that gravity has been accounted for as a body force, and the values of gx,gy,gz will depend on the orientation of gravity with respect to the chosen set of coordinates. The continuity equation reads:

When the flow is at steady-state, does not change with respect to time. The continuity equation is reduced to:

When the flow is incompressible, is constant and does not change with respect to space. The continuity equation is reduced to:

The velocity components (the dependent variables to be solved for) are typically named u, v, w. This system of four equations comprises the most commonly used and studied form. Though comparatively more compact than other representations, this is still a nonlinear system of partial differential equations for which solutions are difficult to obtain. [edit] Cylindrical coordinates A change of variables on the Cartesian equations will yield[11] the following momentum equations for r, , and z:

The gravity components will generally not be constants, however for most applications either the coordinates are chosen so that the gravity components are constant or else it is assumed that gravity is counteracted by a pressure field (for example, flow in horizontal pipe is treated normally without gravity and without a vertical pressure gradient). The continuity equation is:

This cylindrical representation of the incompressible NavierStokes equations is the second most commonly seen (the first being Cartesian above). Cylindrical coordinates are chosen to take advantage of symmetry, so that a velocity component can disappear. A very common case is axisymmetric flow with the assumption of no tangential velocity (u = 0), and the remaining quantities are independent of :

[edit] Spherical coordinates In spherical coordinates, the r, , and momentum equations are[11] (note the convention used: is polar angle, or colatitude,[13] 0 ):

Mass continuity will read:

These equations could be (slightly) compacted by, for example, factoring 1 / r2 from the viscous terms. However, doing so would undesirably alter the structure of the Laplacian and other quantities. [edit] Stream function formulation

Taking the curl of the NavierStokes equation results in the elimination of pressure. This is especially easy to see if 2D Cartesian flow is assumed (w = 0 and no dependence of anything on z), where the equations reduce to:

Differentiating the first with respect to y, the second with respect to x and subtracting the resulting equations will eliminate pressure and any conservative force. 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 degrade into one equation:

where is the (2D) biharmonic operator and is the kinematic viscosity, also express this compactly using the Jacobian determinant:

. We can

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. [edit] Pressure-free velocity formulation The incompressible NavierStokes 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 above eliminates the pressure (in 2D) at the expense of introducing higher derivatives and elimination of the velocity, which is the primary variable of interest. The incompressible NavierStokes equation is composite, the sum of two orthogonal equations,

, where S and I are solenoidal and irrotational projection operators satisfying S + I = 1 and and are the nonconservative 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 and not convenient for numerical computation. An equivalent weak or variational form of the equation, proved to produce the same velocity solution as the NavierStokes equation,[14] 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 imminently suited to finite element computation of divergence-free flow, as we shall see in the next section. There we 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 law. This all would seem to refute the frequent statements that the incompressible pressure enforces the divergence-free condition. [edit] 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 which 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 firstderivative 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-offreedom 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[15] .[16] 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. [edit] 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 obtainied from the gradient of the pressure element. [edit] Compressible flow of Newtonian fluids There are some phenomena that are closely linked with fluid compressibility. One of the obvious examples is sound. Description of such phenomena requires more general presentation of the NavierStokes equation that takes into account fluid compressibility. If viscosity is assumed a constant, one additional term appears, as shown here:[17][18]

where v is the volume viscosity coefficient, also known as bulk viscosity. This additional term disappears for an incompressible fluid, when the divergence of the flow equals zero. [edit] Application to specific problems The NavierStokes 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. For example, after assuming steady, parallel, one dimensional, nonconvective pressure driven flow between parallel plates, the resulting scaled (dimensionless) boundary value problem is:

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

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. 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 nonlinearity. The velocity field may be represented by a function f(z) that must satisfy:

This ordinary differential equation is what is obtained when the NavierStokes 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 > 1.41 (approximately; this is not the square root of 2), the parameter R being the Reynolds

number with appropriately chosen scales. This is an example of flow assumptions losing their applicability, and an example of the difficulty in "high" Reynolds number flows. [edit] Exact solutions of the NavierStokes equations Some exact solutions to the NavierStokes equations exist. Examples of degenerate cases with the non-linear terms in the NavierStokes 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; for example the TaylorGreen vortex.[19][20][21] Note that the existence of these exact solutions does not imply they are stable: turbulence may develop at higher Reynolds numbers. A two dimensional example[show] For example, in the case of an unbounded planar domain with two-dimensional incompressible and stationary flow in polar coordinates (r,), the velocity components (ur,u) and pressure p are:[22]

where A and B are arbitrary constants. This solution is valid in the domain r 1 and for A < 2. In Cartesian coordinates, when the viscosity is zero ( = 0), this is:

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

There is a singularity at x = y = z = 0. [edit] A three dimensional steady-state vortex solution

Some of the flow lines along a Hopf fibration. A nice steady-state example with no singularities comes from considering the flow along the lines of a Hopf fibration. Let r be a constant radius to the inner coil. One set of solutions is given by[23]:

g=0 =0 for arbitrary constants A and 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 is a constant.) 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[show] 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:

[edit] Wyld diagrams Wyld diagrams are bookkeeping graphs that correspond to the NavierStokes equations via a perturbation expansion of the fundamental continuum mechanics. Similar to the Feynman diagrams in quantum field theory, these diagrams are an extension of Keldysh's technique for nonequilibrium processes in fluid dynamics. In other words, these diagrams assign graphs to the (often) turbulent phenomena in turbulent fluids by allowing correlated and interacting fluid particles to obey stochastic processes associated to pseudo-random functions in probability distributions.[24] [edit] NavierStokes equations use in games The NavierStokes equations are used extensively in video games in order to model a wide variety of natural phenomena. These include simulations of effects such as water, fire, smoke etc. Many of the implementations used are based on the seminal paper "Real-Time Fluid Dynamics for Games"[25] by J. Stam. More recent implementations based upon this work run on the GPU as opposed to the CPU and achieve a much higher degree of performance.[26] [27]

Top of Form

Bottom Form

of

On this slide we show the three-dimensional unsteady form of the Navier-Stokes Equations. These equations describe how the velocity, pressure, temperature, and density of a moving fluid are related. The equations were derived independently by G.G. Stokes, in England, and M. Navier, in France, in the early 1800's. The equations are extensions of the Euler Equations and include the effects of viscosity on the flow. These equations are very complex, yet undergraduate engineering students are taught how to derive them in a process very similar to the derivation that we present on the conservation of momentum web page. The equations are a set of coupled differential equations and could, in theory, be solved for a given flow problem by using methods from calculus. But, in practice, these equations are too difficult to solve analytically. In the past, engineers made further approximations and simplifications to the equation set until they had a group of equations that they could solve. Recently, high speed computers have been used to solve approximations to the equations using a variety of techniques like finite difference, finite volume, finite element, and spectral methods. This area of study is called Computational Fluid Dynamics or CFD. The Navier-Stokes equations consists of a time-dependent continuity equation for conservation of mass, three time-dependent conservation of momentum equations and a time-dependent conservation of energy equation. There are four independent variables in the problem, the x, y, and z spatial coordinates of some domain, and the time t. There are six dependent variables; the pressure p, density r, and temperature T (which is contained in the energy equation through the total energy Et) and three components of the velocity vector; the u component is in the x direction, the v component is in the y direction, and the w component is in the z direction, All of the dependent variables are functions of all four independent variables. The differential equations are therefore partial differential equations and not the ordinary differential equations that you study in a beginning calculus class. You will notice that the differential symbol is different than the usual "d /dt" or "d /dx" that you see for ordinary differential equations. The symbol " " is is used to indicate partial derivatives. The symbol indicates that we are to hold all of the independent variables fixed, except the variable next to symbol, when computing a derivative. The set of equations are:

Continuity:

r/

t+

(r * u)/ t+

x+

(r * v)/ x+

y+

(r * w)/

z=0 y+ (r * u * w)/ z=p/ x

X - Momentum: + 1/Re * { tauxx/

(r * u)/ x+ (r * v)/ x+

(r * u^2)/ y+

(r * u * v)/ z} (r * v^2)/ z} (r * v * w)/ z} y+

tauxy/ t+

tauxz/ x+ tauyz/ x+

Y - Momentum: + 1/Re * { tauxy/

(r * u * v)/ y+

y+

(r * v * w)/

z=-

p/

tauyy/ t+ tauyz/ x+

Z - Momentum: + 1/Re * { Energy: tauxz/ Et/

(r * w)/ x+

(r * u * w)/ y+

y+

(r * w^2)/

z=-

p/

tauzz/

t+ qx/

(u * Et)/ x+ qy/

(v * Et)/ qz/ z}

(w * Et)/

z=-

(r * u)/

x-

(r * v)/

y-

(r * w)/

- 1/(Re*Pr) * { + 1/Re * {

y+

(u * tauxx + v * tauxy + w * tauxz)/

x+

(u * tauxy + v * tauyy + w * tauxz)/

y+

(u * tauxz + v * tauyz + w * tauzz)/

z}

where Re is the Reynolds number which is a similarity parameter that is the ratio of the scaling of the inertia of the flow to the viscous forces in the flow. The q variables are the heat flux components and Pr is the Prandtl number which is a similarity parameter that is the ratio of the viscous stresses to the thermal stresses. The tau variables are components of the stress tensor. A tensor is generated when you multiply two vectors in a certain way. Our velocity vector has three components; the stress tensor has nine components. Each component of the stress tensor is itself a second derivative of the velocity components. The terms on the left hand side of the momentum equations are called the convection terms of the equations. Convection is a physical process that occurs in a flow of gas in which some property is transported by the ordered motion of the flow. The terms on the right hand side of the momentum equations that are multiplied by the inverse Reynolds number are called the diffusion terms. Diffusion is a physical process that occurs in a flow of gas in which some property is transported by the random motion of the molecules of the gas. Diffusion is related to the stress tensor and to the viscosity of the gas. Turbulence, and the generation of boundary layers, are the result of diffusion in the flow. The Euler equations contain only the convection terms of the Navier-Stokes equations and can not, therefore, model boundary layers. There is a special simplification of the Navier-Stokes equations that describe boundary layer flows. Notice that all of the dependent variables appear in each equation. To solve a flow problem, you have to solve all five equations simultaneously; that is why we call this a coupled system of equations. There are actually some other equation that are required to solve this system. We only show five equations for six unknowns. An equation of state relates the pressure, temperature, and density of the gas. And we need to specify all of the terms of the stress tensor. In CFD the stress tensor terms are often approximated by a turbulence model.

Você também pode gostar