Você está na página 1de 10

Chapter 2 Shell Momentum Balances and Velocity Distributions in Laminar Flow 92.

1 Shell momentum balances and boundary conditions 92.2 Flow of a falling film 92.3 Flow through a circular tube 92.4 Flow through an annulus 92.5 Flow of two adjacent immiscible fluids 92.6 Creeping flow around a sphere In this chapter we show how to obtain the velocity profiles for laminar flows of fluids in simple flow systems. These derivations make use of the definition of viscosity, the ex- pressions for the molecular and convective momentum fluxes, and the concept of a mo- mentum balance. Once the velocity profiles have been obtained, we can then get other quantities such as the maximum velocity, the average velocity, or the shear stress at a surface. Often it is these latter quantities that are of interest in engineering problems. In the first section we make a few general remarks about how to set up differential momentum balances. In the sections that follow we work out in detail several classical examples of viscous flow patterns. These examples should be thoroughly understood, since we shall have frequent occasions to refer to them in subsequent chapters. Although these problems are rather simple and involve idealized systems, they are nonetheless often used in solving practical problems. The systems studied in this chapter are so arranged that the reader is gradually introduced to a variety of factors that arise in the solution of viscous flow problems. In 52.2 the falling film problem illustrates the role of gravity forces and the use of Cartesian co- ordinates; it also shows how to solve the problem when viscosity may be a function of position. In 52.3 the flow in a circular tube illustrates the role of pressure and gravity forces and the use of cylindrical coordinates; an approximate extension to compressible flow is given. In 52.4 the flow in a cylindrical annulus emphasizes the role played by the boundary conditions. Then in 52.5 the question of boundary conditions is pursued fur- ther in the discussion of the flow of two adjacent immiscible liquids. Finally, in 92.6 the flow around a sphere is discussed briefly to illustrate a problem in spherical coordinates and also to point out how both tangential and normal forces are handled. The methods and problems in this chapter apply only to steady flow.By "steady" we mean that the pressure, density, and velocity components at each point in the stream do not change with time. The general equations for unsteady flow are given in Chapter 3.

This chapter is concerned only with laminar flow. "Laminar flow" is the orderly flow that is observed, for example, in tube flow at velocities sufficiently low that tiny particles injected into the tube move along in a thin line. This is in sharp contrast with the wildly chaotic "turbulent flow" at sufficiently high velocities that the particles are flung apart and dispersed throughout the entire cross section of the tube. Turbulent flow is the sub- ject of Chapter 5. The sketches in Fig. 2.0-1 illustrate the difference between the two flow regimes. 2.1 SHELL MOMENTUM BALANCES AND BOUNDARY CONDITIONS The problems discussed in 52.2 through 52.5 are approached by setting up momentum balances over a thin "shell" of the fluid. For steady pow, the momentum balance is

This is a restricted statement of the law of conservation of momentum. In this chapter we apply this statement only to one component of the momentumnamely, the component in the direction of flow. To write the momentum balance we need the expressions for the convective momentum fluxes given in Table 1.7-1 and the molecular momentum fluxes given in Table 1.2-1; keep in mind that the molecular momentum flux includes both the pressure and the viscous contributions. In this chapter the momentum balance is applied only to systems in which there is just one velocity component, which depends on only one spatial variable; in addition, the flow must be rectilinear. In the next chapter the momentum balance concept is ex- tended to unsteady-state systems with curvilinear motion and more than one velocity component.

The procedure in this chapter for setting up and solving viscous flow problems is as follows: Identify the nonvanishing velocity component and the spatial variable on which it depends. Write a momentum balance of the form of Eq. 2.1-1 over a thin shell perpendicular to the relevant spatial variable. Let the thickness of the shell approach zero and make use of the definition of the first derivative to obtain the corresponding differential equation for the momentum flux. Integrate this equation to get the momentum-flux distribution. Insert Newton's law of viscosity and obtain a differential equation for the velocity. Integrate this equation to get the velocity distribution. Use the velocity distribution to get other quantities, such as the maximum velocity, average velocity, or force on solid surfaces.

In the integrations mentioned above, several constants of integration appear, and these are evaluated by using "boundary conditionsu-that is, statements about the velocity or stress at the boundaries of the system. The most commonly used boundary conditions are as follows: a) At solid-fluid interfaces the fluid velocity equals the velocity with which the solid surface is moving; this statement is applied to both the tangential and the normal component of the velocity vector. The equality of the tangential components is referred to as the "no-slip condition.'' b) At a liquid-liquid interfacial plane of constant x, the tangential velocity compo- nents v, and v,are continuous through the interface (the "no-slip condition") as are also the molecular stress-tensor components p + T,,, rxy,and T,,. c) At a liquid-gas interfacial plane of constant x, the stress-tensor components T,, and T,, are taken to be zero, provided that the gas-side velocity gradient is not too large. This is reasonable, since the viscosities of gases are much less than those of liquids. In all of these boundary conditions it is presumed that there is no material passing through the interface; that is, there is no adsorption, absorption, dissolution, evapora- tion, melting, or chemical reaction at the surface between the two phases. Boundary con- ditions incorporating such phenomena appear in Problems 3C.5 and llC.6, and 518.1. In this section we have presented some guidelines for solving simple viscous flow problems. For some problems slight variations on these guidelines may prove to be appropriate.

,92.2 FLOW OF A FALLING FILM The first example we discuss is that of the flow of a liquid down an inclined flat plate of length L and width W, as shown in Fig. 2.2-1. Such films have been studied in connection with wetted-wall towers, evaporation and gas-absorption experiments, and applications of coatings. We consider the viscosity and density of the fluid to be constant. A complete description of the liquid flow is difficult because of the disturbances at the edges of the system (z = 0, z= L, y = 0, y = W). An adequate description can often be

obtained by neglecting such disturbances, particularly if W and L are large compared to the film thickness 6. For small flow rates we expect that the viscous forces will prevent continued acceleration of the liquid down the wall, so that v, will become independent of z in a short distance down the plate. Therefore it seems reasonable to postulate that v, = v,(x), v, = 0, and v, = 0, and further that p = p(x). From Table B.l it is seen that the only nonvanishing components of Iare then T,, = T,, = -p(dv,/dx). We now select as the "system" a thin shell perpendicular to the x direction (see Fig. 2.2-2). Then we set up a z-momentum balance over this shell, which is a region of thick- ness Ax, bounded by the planes z = 0 and z = L, and extending a distance Win the y di-rection. The various contributions to the momentum balance are then obtained with the help of the quantities in the "z-component" columns of Tables 1.2-1 and 1.7-1. By using the components of the "combined momentumflux tensor" + defined in 1.7-1 to 3, we can include all the possible mechanisms for momentum transport at once:

By using the quantities +,, and +,, we account for the z-momentum transport by all mechanisms, convective and molecular. Note that we take the "in" and "out" directions in the direction of the positive x- and z-axes (in this problem these happen to coincide with the directions of z-momentum transport). The notation I,,,, means "evaluated at x + Ax," and g is the gravitational acceleration. When these terms are substituted into the z-momentum balance of Eq. 2.1-1, we get

Fig. 2.2-2 Shell of thickness Ax over which a z-momentum balance is made. Arrows show the momentum fluxes associated with the surfaces of the shell. Since v, and v, are both zero, pvxvz and pvp, are zero. Since v, does not depend on y and z, it follows from Table B.l that T,, = 0 and T,, = 0. Therefore, the dashed-underlined fluxes do not need to be considered. Both p and pv,v, are the same at z = 0 and z = L, and therefore do not appear in the final equation for the balance of zmomentum, Eq. 2.2-10. When this equation is divided by LW Ax, and the limit taken as Ax approaches zero, we get

The first term on the left side is exactly the definition of the derivative of 4,: with respect to x. Therefore Eq. 2.2-7 becomes

At this point we have to write out explicitly what the components +,, and 4,: are, mak- ing use of the definition of + in Eqs. 1.7-1 to 3 and the expressions for rxzand T,, in Ap- pendix B.1. This ensures that we do not miss out on any of the forms of momentum transport. Hence we get

In accordance with the postulates that v, = v,(x), v, = 0, v, = 0, and p = p(x), we see that (i) since v, = 0, the pup, term in Eq. 2.2-9a is zero; (ii) since v, = v,(x), the term -2,u(dv,/dz) in Eq. 2.2-9b is zero; (iii) since v, = v,(x), the term pv,v, is the same at z = 0 and z = L; and (iv) since p = p(x), the contribution p is the same at z = 0 and z = L. Hence T,, depends only on x, and Eq. 2.2-8 simplifies to

This is the differential equation for the momentum flux T,,. It may be integrated to give

The constant of integration may be-evaluated by using the boundary condition at the gas-liquid interface (see 2.1):

Substitution of this boundary condition into Eq. 2.2-11 shows that C, = 0. Therefore the momentum-flux distribution is

as shown in Fig. 2.2-3. Next we substitute Newton's law of viscosity

into the left side of Eq. 2.2-13 to obtain

which is the differential equation for the velocity distribution. It can be integrated to give

The constant of integration is evaluated by using the no-slip boundary condition at the solid surface:

B.C. 2 at x = 6, v, = 0 (2.2-17)

Substitution of this boundary condition into Eq. 2.2-16 shows that C2 = (pg cos P/24a2. Consequently, the velocity distribution is

This parabolic velocity distribution is shown in Fig. 2.2-3. It is consistent with the postu- lates made initially and must therefore be a possible solution. Other solutions might be possible, and experiments are normally required to tell whether other flow patterns can actually arise. We return to this point after Eq. 2.2-23. Once the velocity distribution is known, a number of quantities can be calculated: (i) The maximum velocity vZ,,,, is clearly the velocity at x = 0; that is,

a. (ii) The average velocity (v,) over a cross section of the film is obtained as follows:

The double integral in the denominator of the first line is the cross-sectional area of the film. The double integral in the numerator is the volume flow rate through a differential element of the cross section, v,dx dy, integrated over the entire cross section. (iii) The mass rate of flow w is obtained from the average velocity or by integration of the velocity distribution

(iii)The film thickness S may be given in terms of the average velocity or the mass rate of flow as follows:

(v) The force per unit area in the z direction on a surface element perpendicular to the x direction is +T,, evaluated at x = 6. This is the force exerted by the fluid (region of lesser x) on the wall (region of greater x). The z-component of the force F of the fluid on the solid surface is obtained by integrating the shear stress over the fluid-solid interface:

This is the z-component of the weight of the fluid in the entire film-as we would have expected. Experimental observations of falling films show that there are actually three "flow regimes," and that these may be classified according to the Reynolds number,' Re, for the flow. For falling films the Reynolds number is defined by Re = 4S(vz)p/p.The three flow regime are then: laminar flow with negligible rippling Re <20 laminar flow with pronounced rippling 20 <Re < 1500 turbulent flow Re > 1500

The analysis we have given above is valid only for the first regime, since the analysis was restricted by the postulates made at the outset. Ripples appear on the surface of the fluid at all Reynolds numbers. For Reynolds numbers less than about 20, the ripples are very long and grow rather slowly as they travel down the surface of the liquid; as a re- sult the formulas derived above are useful up to about Re = 20 for plates of moderate length. Above that value of Re, the ripple growth increases very rapidly, although the flow remains laminar. At about Re = 1500 the flow becomes irregular and chaotic, and the flow is said to be t~rbulent.~,~ At this point it is not clear why the value of the Reynolds number should be used to delineate the flow regimes. We shall have more to say about this in g3.7.

This discussion illustrates a very important point: theoretical analysis of flow systems is limited by the postulates that are made in setting up the problem. It is absolutely necessary to do experiments in order to establish the flow regimes so as to know when instabilities (spontaneous oscillations) occur and when the flow becomes turbulent. Some information about the onset of instability and the demarcation of the flow regimes can be obtained by theoretical analysis, but this is an extraordinarily difficult subject. This is a result of the inherent nonlinear nature of the governing equations of fluid dynamics, as will be explained in Chapter 3. Suffice it to say at this point that experiments play a very important role in the field of fluid dynamics.

Você também pode gostar