Sunteți pe pagina 1din 6

ME 563 - Intermediate Fluid Dynamics - Su

Lecture 24 - Boundary layers: the boundary layer on a flat plate

Reading: Acheson, §8.2-8.5.

In Lecture 23 we developed the boundary layer equations for 2-D, steady flow, which were

∂u ∂u 1 dp ∂2u
u +v =− +ν 2 (1)
∂x ∂y ρ dx ∂y
∂u ∂v
+ = 0. (2)
∂x ∂y
To derive these, our key assumption was that u and v vary much more rapidly in the y-direction
than in the x-direction. In terms of derivatives, this meant
∂u ∂u
 and
∂y ∂x
∂v ∂v
 , (3)
∂y ∂x
which allowed us to toss out various terms in the Navier-Stokes equations. We can get some
interesting insight by looking further at (3). Let U0 be a characteristic value of u, and suppose that
u varies by order U0 over a distance L in the x-direction, and suppose also that u varies by order
U0 over a distance δ in the y-direction. Thus, L is a measure of the boundary layer length and δ
is a measure of the boundary layer thickness. In terms of derivatives, we can write
 
∂u U0
=O and
∂y δ
 
∂u U0
=O ,
∂x L

and by applying (3), we get

δ  L, (4)

which means that the boundary layer is much thinner than it is long, which we already anticipated
earlier in our definition of a boundary layer as a thin layer in where viscous effects are important.
Applying these estimates for the magnitude of u and its derivatives, we can also estimate the
magnitude of v. Start with (2):
 
∂v ∂u U0
=− =O . (5)
∂y ∂x L

We can approximate ∂v/∂y in the boundary layer as follows –

∂v v(y = δ) − v(y = 0)
≈ . (6)
∂y δ
But v(y = 0) = 0 (the boundary is at y = 0), so combining (5) and (6), we get
 
U0 δ
v=O ,
L
which quantifies our qualitative conclusion from the last lecture that v was much smaller than u.

1
Finally, we can use a similar analysis to estimate the thickness of the boundary layer as a
function of the Reynolds number. Looking at (1), we want to ensure that the viscous term on the
right side of the equation is comparable in magnitude to the terms on the left. Remember the
bigger picture; away from the boundary, the flow is inviscid, so the Euler equation pertains, which
means we only have to worry about the left side of (1) and the pressure term on the right. In the
boundary layer, viscosity is important, so we additionally have to consider the viscous term on the
right side of (1). Applying the same estimates as above, the terms on the left side are each of order
U02 /L, while the viscous term is of order νU0 /δ2 . Setting these magnitudes equal, we get

δ2 ν
≈ , so
L U0
 1/2
δ ν
≈ = Re−1/2 , (7)
L U0 L

where we have defined the Reynolds number as Re = U0 L/ν. Compare this with (4). Our initial
assumption, that u and v vary more quickly in y than in x, led to the condition δ  L. From (7),
we can see that that assumption is equivalent to requiring that the Reynolds number be large.

1 The boundary layer on a flat plate


Let’s look at the example of a boundary layer on a flat plate (Fig. 1). In this situation, a uniform,
parallel, incompressible flow with u = U encounters a flat plate, infinitely thin, aligned with the
flow direction. The plate is located at y = 0 and begins at x = 0. The viscosity of the fluid is

Figure 1: A uniform flow encountering a flat plate.

very small, so the flow is essentially inviscid except near the flat plate, where the no-slip condition
pertains, so u(y = 0) = 0 for x > 0. We will consider the steady-state situation, so time derivatives
are zero.
We can make one simplification to the boundary layer equations for this problem. Consider
a streamline that is located just ‘outside’ of the boundary layer. By ‘outside,’ we mean that the
streamline is in that part of the u(y) distribution where u ≈ U . Because this streamline is outside
of the boundary layer, we can treat the fluid as ideal on the streamline. Since the flow is steady, we
can apply the (first) Bernoulli theorem on the streamline, i.e. ρp + 12 u2 is constant on the streamline.
Now, remember from Lecture 23 that we can split the flow into two parts, the main (inviscid) flow
and the boundary layer, and get solutions for those parts separately. For this problem, the main
flow is uniform, and since the main flow is inviscid, there is no no-slip condition. Thus the main flow
remains uniform even in the presence of the plate, and in particular u is constant on streamlines
outside the boundary layer. Bernoulli’s theorem then tells us that the pressure, p, is constant on
those streamlines. Since the streamlines in the main flow are parallel, we have dp/dx = 0 in the main
flow. But – we saw in the last lecture that p is basically a function only of x in the boundary layer.
The pressure distribution inside the boundary layer, then, is the same as the pressure distribution

2
in the main flow just outside the boundary layer. For this problem, this means
dp
=0
dx
in the boundary layer.
With the pressure term out, then, the boundary layer equations become

∂u ∂u ∂2u
u +v =ν 2 (8)
∂x ∂y ∂y
∂u ∂v
+ = 0. (9)
∂x ∂y
As we’ve done in other problems, we’re going to look for a similarity solution for u. Acheson
justifies this approach somewhat in the text; remember (from the review notes for the first exam)
that if a similarity solution doesn’t exist, it will become obvious in the course of the calculation.
The similarity solution we seek is of the form

u = U h(η), (10)

where
y
η= ,
g(x)
and rather than assume a particular form for the normalizing function g, we’ll let it come out of
the calculation.
The first thing we’ll do is write u (and v) in terms of a stream function, ψ, which will guarantee
that our solutiion automatically satisfies (9). Thus
∂ψ ∂ψ
u= , v=− . (11)
∂y ∂x
By integrating (10) with respect to y, we then get

ψ = U g(x) h(s) ds + k(x). (12)
0

Remember now that lines of constant ψ are streamlines. We want the plate to be a streamline (it
would be bad if velocity vectors pointed into or out of the plate). We can arbitrarily set ψ = 0 on
the plate. The surface of the plate is located at y = 0, which also means η = 0. So (12) becomes

ψ = U g(x) h(s) ds,
0

which can also be written as

ψ = U g(x)f (η), where f (0) = 0, (13)

so f (η) is now the similarity function we’ll be looking for.


From this form of ψ, and using (11), we have for the velocity components
dg
u = U f 0 (η) and v = U (ηf 0 − f ) (14)
dx

3
where the primes mean d/dη. From (14), the velocity derivative terms in (8) are
 
∂u ∂η y dg
= U f 00 = U f 00 − 2 ,
∂x ∂x g dx
∂u ∂η 1
= U f 00 = U f 00
∂y ∂y g
∂2u 1
= U f 000 2 , (15)
∂y 2 g
and inserting into (8), we get

y dg dg f 00 f 000
−U 2 f 0 f 00 + U 2
(ηf 0
− f ) = νU ,
g2 dx dx g g2
which gives
1 dg 1 dg 1 dg f 000
−U 2 f 0 f 00 η + U 2 f 0 f 00 η − U 2 f f 00 = νU 2 ,
g dx g dx g dx g
which finally simplifies to
U dg 00
f 000 + g f f = 0. (16)
ν dx
Remember that in writing (10), we were making the assumption that the velocity, which was
originally a function of the two (independent) variables x and y, could instead be expressed as
the function of a single independent variable, the similarity variable, η, that combines x and y.
Equation (10) led to (13), where the similarity function was f (η). In order for the assumption of
similarity to be valid, the differential equation for f (η), (16), has to have η as the only independent
variable. But g(x), which is a function of the independent variable x, appears in (16). So for the
similarity assumption to hold, we have to eliminate x as an independent variable. The way to do
this is to set g · dg/dx constant. The choice of this constant is arbitrary. We will pick
dg ν
g = (17)
dx U
since this choice will also eliminate ν and U from (16).
(Why can we seemingly arbitrarily pick g(x) to satisfy the similarity assumptions? The answer
is that we are free to choose g anyway we like – in this case, we’re looking for a similarity solution
with the similarity variable y/g(x) – as long as the assumptions we make are consistent and our
solution for u and v satisfies the boundary layer equations. Basically, the boundary layer equations
have a solution. All we’re doing is trying to find that solution. To make the job of finding the
solution easier, we express the problem in a more convenient fashion, where in this case ‘convenient’
means in terms of just one independent variable. By doing this we don’t affect the solution is any
way. If the solution can’t actually be written in the form we want, we will know because the
equations we get will be inconsistent with our assumptions.)
To solve for g(x), first observe that
 
dg d 1 2
g = g .
dx dx 2

Substituting this in (17) and integrating, we get


1 2 νx
g = +d
2 U

4
Figure 2: Initial stage of the boundary layer.

where d is a constant. To specify the value of d, look at what happens when g = 0. If g = 0, then
by (15), ∂u/∂y becomes undefined. But it is quite reasonable for ∂u/∂y to be undefined at x = 0.
Consider Fig. 2. Immediately upstream of the plate, at x . 0, the velocity profile u(y) is uniform.
Then, starting at x = 0, the no-slip condition holds on the plate surface, y = 0. For x & 0, however,
the boundary layer is still vanishingly thin, so u has to vary from 0 to U over a distance δ ≈ 0.
But the derivative ∂u/∂y is of order U/δ, so at x = 0 we can say that ∂u/∂y is undefined.
We therefore set g = 0 at x = 0, which gives us
1 2 νx
g =
2 U
or
 1/2
2νx
g(x) = .
U

Finally, we have
y
ψ = (2νU x)1/2 f (η), where η = ,
(2νx/U )1/2 )

which satisfies the differential equation

f 000 + f f 00 = 0,

with boundary conditions

f (0) = f 0 (0)
f 0 (0) = 0
f 0 (∞) = 1.

The first boundary condition comes from setting v = 0 at y = 0 in (14), the second comes from
setting u = 0 at y = 0, and the third is from setting u = U outside the boundary layer.
There is no analytical solution for this problem. Instead, the problem can be solved numerically.
The resulting velocity profile, f (η) = u/U , is shown in Fig. 3.

5
Figure 3: The flat plate boundary layer profile, u/U = f (η) (Acheson’s Figure 8.8).

S-ar putea să vă placă și