02 - THE BOUNDARY LAYER Cengel

You might also like

Download as pdf
Download as pdf
You are on page 1of 10
FIGURE 10-73 ‘Nondimensional axial velocity (bine curve) and pressure coefficient (green curve) along the floor below a vacuum leaner modeled as an irrotational region of flow. ‘Masia! FIGURE 10-74 Based on an irrotational flow approximation, the maximum speed along the floor beneath a vacuum cleaner nozzle occurs at x= tb. A stagnation point occurs directly below the nozzle. aaoo 1 2 3 [Normalized disiencealeag the Boor, x* a Ee eee ae a tatenis) Maximson speed along the floor. ‘We expect that the vacuum cleaner is most effective at sucking up ditt from the floce when the speed along the floor is greatest and the pressive along, the floor is lowest. Thus, contrary to what you may have thought, the best perfor nance is wot directly below the suction ints, but rather at x= b, as illustrated in Fig. 10-75. Discussion Notice that we never used the wid of the vacuum nozzle in our analysis, since a line sink kas no length scale. You can convince yourself that a ‘vacuum cleaner works best at x a 4b by performing a simple experiment with a ‘acum cleaner and some small granular material (ike sugar or salt) om a aed floor. It wus oot that tbe irotational approximation is quite realise for flow into the inlet ofa vacuum cleaver everywhere except very close (othe floor, because the flow is rotational there. 250m/s (16) ‘We conclude this section by emphasizing that although the irrotational flow approximation is mathematically simple, and velocity and pressure fields are easy to obtain, we must be very careful where we apply it, The irrotational flow approximation breaks down in regions of non-negligible vorticity, especially near solid walls, where fluid particles rotate because of viscous stresses caused by the no-slip condition at the wall. This leads us to the final section in this chapter (Section 10-6) in which we discuss the boundary layer approximation, 40-6 = THE BOUNDARY LAYER APPROXIMATION ‘As discussed in Sections 10-4 and 10-5, there are at least two flow situ- ations in which the viscous term in the Navier-Stokes equation can be neglected. The first occurs in high Reynolds number regions of flow where net viscous forces are known to be negligible compared to inertial and/or pressure forces; we call these inviscid regions of flow. The second situation occurs when the vorticity is negligibly small; we call these irrotational or potential regions of flow. In cither case, removal of the viscous terms from the Navier-Stokes equation yields the Euler equation (Eg. 10-13 and also Eq, 10-25). While the math is greatly simplified by dropping the viscous terms, there are some serious deficiencies associated with application of the Euler equation to practical engineering flow problems. High on the list of deficiencies is the inability to specify the no-slip condition at solid walls. This leads to unphysical results such as zero viscous shear forces on solid walls and zero aerodynamic drag on bodies immersed in a free stream, We can therefore think of the Euler equation and the Navier-Stokes equation as two mountains separated by a huge chasm (Fig. 10-75a). We make the fol- lowing statement about the boundary layer approximation: The boundary layer approximation bridges the gap between the Euler equation and the Navier-Stokes equation, and between the slip condition and the no-slip condition at solid walls (Fig. 10-756). From a historical perspective, by the mid-1800s, the Navier-Stokes equation was known, but couldn't be solved except for flows of very sim ple geometries. Meanwhile, mathematicians were able to obtain beautiful analytical solutions of the Euler equation and of the potential flow equa- tions for flows of complex geometry, but their results were often physi- cally meaningless. Hence, the only reliable way to study fluid flows was empirically, ic., with experiments. A major breakthrough in fluid mechan- ies occurred in 1904 when Ludwig Prandtl (1875-1953) introduced the boundary layer approximation. Prandtl’s idea was to divide the flow into two regions: an outer flow region that is inviscid and/or irrotational, and an inner flow region called a boundary layer—a very thin region of flow near a solid wall where viscous forces and rotationality cannot be ignored (Fig. 10-76). In the outer flow region, we use the continuity and Euler equations to obtain the outer flow velocity field, and the Bernoulli equa- tion to obtain the pressure field. Alternatively, if the outer flow region is irrotational, we may use the potential flow techniques discussed in Section 10-5 (eg., superposition) to obtain the outer flow velocity field. In either case, we solve for the outer flow region first, and then fit in a thin boundary layer in regions where rotationality and viscous forces cannot be neglected. Within the boundary layer we solve the boundary layer equations, to be discussed shortly. (Note that the boundary layer equations are themselves approximations of the full Navier-Stokes equation, as we will see.) The boundary layer approximation corrects some of the major deficien- cies of the Euler equation by providing a way to enforce the no-slip coi dition at sofid walls. Hence, viscous shear forces can exist along walls, bodies immersed in a free stream can experience aerodynamic drag, and flow separation in regions of adverse pressure gradient can be predicted more accurately. The boundary layer concept therefore became the workhorse of engineering fluid mechanics throughout most of the 1900s. However, the advent of fast, inexpensive computers and computational fluid dynamics (CFD) software in the latter part of the twentieth century enabled numeri- cal solution of the Navier-Stokes equation for flows of complex geome- try. Today, therefore, it is no longer necessary to split the flow into outer flow regions and boundary layer regions—we can use CFD to solve the 559 Eu hp :. FIGURE 10-75 (a) A lmge gap exists between the Baler equation (which allows slip at ‘walls) and the Navier-Stokes equation (which supports the no-slip condition); (8) the boundary layer approximation bridges that gap. Outer How (inviscid andlor inouational region of flow) Boundary layer (tational with, on-negligitle vizcoue force) WA FIGURE 10-76 Prandil’s boundary layer concept splits the flow into an outer flow region and a thin boundary layer region (not to scale. 560 zee Flow of a uniform stream parallel to a flat plate (deavings not to scale). (a)Re, ~ 10, (6) Re, ~ 10". The, larger the Reynolds nitober, the thinner the boundary layer along the plate st a given x-location FIGURE 10-78 Flow visualization of a laminar fat plete boundary layer profile, Photograph taken by F. X. Wortmann in 1953 as visvelized with the tellurionn method. Flow is from left to right, snd the leading edge of the flat plate is far to the left of the field of view, Wormnain PX, 1077 AGARD Cont, Prono mapper 12 full set of equations of motion (continuity plus Navier-Stokes) throughout the whole flow field. Nevertheless, boundary layer theory is still useful in some engineering applications, since it takes much less time to arrive at a solution. In addition, there is a lot we can learn about the behavior of flow- ing fluids by studying boundary layers. We siress again that boundary layer solutions are only approximations of full Navier-Stokes solutions, and we must be careful where we apply this ot any approximation, The key to successful application of the boundary layer approximation is the assumption that the boundary layer is very thin, The classic example is a uniform stream flowing parallel to a long flat plate aligned with the x-axis. Boundary layer thickness 6 at some location z along the plate is sketched in Fig. 16-77. By convention, 6 is usually defined as the distance away from the wall at which the velocity component parallel to the wall is 99 percent of the fluid speed outside the boundary layer. It turns out that for a given fluid and plate, the higher the free-stream speed V, the thinner the boundary layer (Fig, 10-77), In noncimensional terms, we define the Reynolds number based on distance x along the wall, Reynolds number along a flat plate: (10-60) Hence, ‘At s given x-location, the higher the Reynolds number the thinner the boundary layer Tn other words, the higher the Reynolds number, all else being equal, the more reliable the boundary layer epproximation. We are confident that the bound- ary layer is thin when 6 Re, .). Itis also com- ‘mon in heat transfer to use this value as the critical Re; in fact, relations for average friction and heat transfer coefficients are derived by assuming the flow to be laminar for Re, lower than Re, .«, and turbulent otherwise. ‘The logic here is to ignore transition by treating the first part of transition as laminar and the remaining part as turbulent. We follow this convention throughout the rest of the book unless noted otherwise. The transition process is unsteady as well and is difficult to predict, even with modern CFD codes. In some cases, engineers install rough sandpaper or wires called trip wires along the surface, in order to force transition at a desired location (Fig. 10-83). The eddies from the trip wire cause enhanced local mixing and create disturbances that very quickly lead to a turbulent boundary layer. Again, the vertical scale in Fig. 10-83 is greatly exaggerated for illustrative purposes. SS a eT Se An aluminum canoe moves horizontally along the surface of a lake at 3.5 mifh Fig. 10-84). The temperature of the lake water is 50°F. The bottom of the 1x cance is 20 ft Tong and is flat, Is the boundary layer on the canoe bottom laminar pr turbulent? SOLUTION We are to assess whether the boundary layer on the bottom of a canoe is laminar or turbulent. Assumptions 1 The flow is steady and incompressible. 2 Ridges, dings, and ther nononiformities in the bottom of the canoe are ignored—the bottom is assumed to be « smooth flat plate aligned exactly with the direction of flow. 3 From the frame of reference of the canoe, the water below the boundary layer under the feanoe moves at uniforin speed V = 3.5 map, Properties The kinematic viscosity of water at T= SOF is v= 1407 x 107 i, Analysis First, we caleulate the Reynolds number at the stern of the cance, Vx __G.5 mifh}20 fy) (= 8) th ) Fy 1407 x 105 RG Imi /\36008, Since Re, is much greater than Re, (5 x 10°), and is even greater than Re, iin (50 10°), the boundary layer is defintely turbulent by the back of the canoe, Discussion Since the canoe bottom is nether pecfeetly senacth nor perfeotly fist, and since we expect some disturbances in the lake water due to waves, the paddles, swimming fish, ete, transition (0 turbulence is expected to occur much eaulier and more rapidly than illustrated forthe ideal case in Big. 10-81. Hence we ae even more confident that this boundary layer is turbulent. = 7.30 x 10° The Boundary Layer Equations Now that we have a physical feel for boundary layers, we need the equations of motion to be used in boundary layer calculations—the boundary layer equations, For simplicity we consider only steady, two- dimensional flow in the x-plane in Cartesian coordinates. The methodology used here can be extended, however, to axisymmetric boundary layers or to three-dimensional boundary layers in any coordinate system. We neglect gravity since we are not dealing with free surfaces or with buoyancy- 6, the pressure gradient term is orders of magnitude greater than the advective terms on the left side of the equation. Thus, the only term left in Eq. 10-64 is the pressure term, Since no other term in the equation can balance that term, we have no choice but to set it equal to zero. Thus, the nondimensional j-momientum equation reduces to apt ay or, in terms of the physical variables, ter Bo Bound yer SS ow Soma pesan gio srghatomdaytons ad yy In words, although pressure may vary along the wall (in the x-direction), there is negligible change in pressure in the direction normal to the wall. This is illustrated in Fig. 10-88, At x = x,, P= P, at all values of y across the boundary layer from the wall to the outer flow. At some other location, X= xy the pressure may have changed, but P = P, at all values of y across that portion of the boundary layer The pressure actoss @ boundary layer (y-direction) is nearly constant. FIGURE 10-88 Pressure may change along a boundary layer (x-direction), but the Physically, because the boundary layer is so thin, streamfines within the change in pressure across a boundary boundary layer have negligible curvature when observed at the scale of the layer direction) is negligible. boundary layer thickness. Curved streamlines require a centripetal accelera- sion, which comes from a pressure gradient along the radius of curvature. Since the streamlines are not significantly curved in a thin boundary layer, there is no significant pressure gradient across the boundary layer. 566 ‘Outer flow, \ Boundary ayer FIGURE 10-89 ‘The pressure in the irrotational region of flow outside of a boundary layer ccan be measured by static pressure taps in the surface of the wall. Two such pressure taps ae sketched. 7 ta)=v =| wer ae FIGURE 10-90 ‘An order-of-magnitude analysis of the laminar boundary layer equations along a flat plate reveals that 5 grows like x (not to scale). One immediate consequence of Eq. 10-65 and the statement just presented is that at any x-location along the wall, the pressure at the outer edge of the boundary layer (y & 6) is the same as that at the wall (y = 0). This leads to 2 tremendous practical applicstion; namely, the pressure at the outer edge of a boundary layer can he measured experimentally by a static pressure tap at the wall directly beneath the boundary layer (Fig. 10-89), Experimental- ists routinely take advantage of this fortunsie situation, and countless airfoil shapes for airplane wings and turbomachinery blades were tested with such pressure taps over the past century. ‘The experimental pressure data shown in Fig. 10-64 for flow over a cir cular cylinder were measured with pressure taps at the cylinder’s surface, yet they are used to compare with the pressure calculated by the irrotational ‘outer flow approximation. Such a comparison is valid, because the pressure obtained outside of the boundary layer (from the Euler equation or poten- tial flow analysis coupled with the Bernoulli equation) applies all the way through the boundary layer to the wall. Returning to the development of the boundary layer equations, we use Eq, 10-65 to greatly simplify the x-component of the momentum equation. Specifically, since P is not a function of y, we replace dP/ax by dPldx, where P is the value of pressure calculated from our outer flow approxima- tion (using either continuity plus Euler, or the potential flow equations plus Bernoulli). The x-component of the Navier-Stokes equation becomes au ou Lap a Pe gatz a t Ya t “ey ew “Dosey 13 Pw After some algebra, and after multiplying each term by L/L, we get oH yd ae (ode ey out wet Senn aat (ar) amt lar)ls) je Comparing terms in Eq. 10-66, the middle term on the right side is clearly orders of magnitude smaller than the terms on the left side, since Rez = ULiv > 1. What about the last term on the right? If we neglect this term, we throw out all the viscous terms and are back to the Euler equatio Clearly this term must remain. Furthermore, since all the remaining terms in Eq. 10-66 are of order unity, the combination of parameters in parentheses in the last term on the right side of Eq. 10-66 must also be of order unity, iy “Again recognizing that Re, = U/L/v, we see immediately that 61 EO VRE, This confirms our previous statement that at a given streamwise location along the wall, the larger the Reynolds number, the thinner the boundary layer. If we substitute x for Z, in Fg. 10-67, we also conclude that for a lami- nar boundary layer on a flat plate, where U(x) constant, 5 grows like the square root of x (Fig. 10-90). (10-67) In terms of the original (physical) variables, Eq, 10-66 is written as Se ee at ay (10-68) Note that the last term in Eq. 10-68 is not negligible in the boundary layer, since the y-derivative of velocity gradient du/dy is sufficiently large to offset the (typically small) value of kinematic viscosity v. Finally, since we Know from our y-momentum equation analysis that the pressure across the boundary layer is the same as that outside the boundary layer (Eq, 10-65), wwe apply the Bernoulli equation to the outer flow region. Differentiating with respect to x we get Phere constant + LP ey (0. poe = pa =) where we note that both P and U are functions of x only, as illustrated in Fig. 10-91. Substitution of Eq. 10-69 into Eq. 10-68 yields ou ae, aU Ou wet ay Uae tos (10-70) and we have eliminated pressure from the boundary layer equations. We summarize the set of equations of motion for a steady, incompress- ‘ble, laminar boundary layer in the xj-plane without significant gravitational effects, ae, de ax * oy eg ye yy SH ae ay ae 7 OF 0 Boundary layer equations: (10-79 Mathematically, the full Navier-Stokes equation is elliptic in space, which means that boundary conditions are required over the entire boundary of the flow domain. Physically, flow information is passed in all directions, both upstream and downstream. On the other hand, the a-momentam boundary layer equation (the second equation of Eq. 10-71) is parabolie, This means that we need to specify boundary conditions on oaly three sides of the (two- dimensional) flow domain. Physically, flow information is not passed in the direction opposite to the flow (from downstream). This fact greatly reduces the level of difficulty in solving the boundary layer equations. Specifically, we don't need to specify boundary conditions downstream, only upstream and on the top and bottom of the flow domain (Fig. 10-92). For a typical boundary layer problem slong a wall, we specify the no-slip condition at the wall (w= 0 = 0 at y = 0), the outer flow condition st the edge of the bound- ary layer and beyond [u = U(x) as y + oo}, and a starting profile at some upstream location [1 = tjaing()) At X= Spies WHEE paring MAY OF MAY not be zero]. With these boundary conditions, we simply march downstream in the x-direction, solving the boundary layer equations as we go. This is particularly attractive for numerical boundary layer computations, because ‘once we know the profile at one x-location (:}, we can march to the next x-location (x;,,), and then use this newly calculated profile as the starting profile to march to the next x-location (2;,.), ete 367 P=PX,U=00) ay Y, Oo) Pay ae Pas FIGURE 10-91 Outer flow speed parallel to the wall is UG) and is obtained from the outer flow pressure, P(x). This speed appears in the x-component of the boundary layer momentum equation, Eq. 10-70. [No bousdary conditions on dowasteat edge of fw dotmain FIGURE 10-92 ‘The boundary layer equation set is par- abolic, 50 boundary conditions need to be specified on only three sides of the flow domain.

You might also like