You are on page 1of 15

Non-Dimensional Navier-Stokes Equations

Kiran Kumar
February 19, 2013
1 Navier-Stokes equation in 2D

Q
t
+

E
x
+

F
y
=


E
v
x
+

F
v
y

Q =
_

u
v
E
t
_

_
,

E =
_

_
u
u
2
+ p
uv
(E
t
+ p)u
_

_
,

F =
_

_
v
vu
v
2
+ p
(E
t
+ p)v
_

E
v
=
_

_
0

xx

xy
u
xx
+ v
xy
q
x
_

_
,

F
v
=
_

_
0

xy

yy
u
xy
+ v
yy
q
y
_

xx
=
2
3

_
2
u
x

v
y
_
,
xy
=
yx
=
_
u
y
+
v
x
_
,
xx
=
2
3

_
2
v
y

u
x
_
q
x
= k
T
x
, q
y
= k
T
y
p = ( 1)
_
E
t


2
_
u
2
+ v
2
_
_
2 Non dimensional Navier-Stokes equations
2.1 Case 1
L
r
= L, U
r
= a

,
r
=

, T
r
= T

,
r
=

=
ta
L
, x

=
x
L
, y

=
y
L
,

, k

=
k
k
u

=
u
a
, v

=
v
a
, p

=
p
a
2

, E

t
=
Et
a
2

1
2.1 Case 1 2 NON DIMENSIONAL NAVIER-STOKES EQUATIONS
2.1.1 Momentum Equation
On LHS,
(u)
t
=
_

a
2

L
_
(

)
t

On RHS,

xx
=
2
3

_
2
u
x

v
y
_
=
_

L
_
2
3

_
2
u

_
=
_

L
_
_
1

a
2

_
LHS
2
3

_
2
u

_
=
_

L
M

_
2
3

_
2
u

_
=
M

Re

2
3

_
2
u
x

v
y
_
2.1.2 Energy Equation
On LHS,
(E
t
)
t
=
_

a
3

L
_
(E

t
)
t

On RHS,
u
xx
=
_

a
2

L
_
u

=
_

a
2

L
__
1

a
3

_
LHS
u

=
M

Re
u
xy
q
x
= k
T
x
=
_
C
p
Pr
_
T
x
=
_
R
( 1)Pr
_
T
x
=
_

RT

( 1)PrL
_

=
_

a
2

( 1)L
_

Pr
T

=
_

a
2

( 1)L
__
1

a
3

_
LHS

Pr
T

=
_

L
__
M

( 1)
_

Pr
T
x
=
_
M

( 1)Re

_

Pr
T
x
2
3 VISCOSITY
3 Viscosity
The viscosity of a uid is a measure of its resistance to gradual deformation by shear stress
or tensile stress. For liquids, it corresponds to the informal notion of thickness. For
example, honey has a higher viscosity than water.
Viscosity is due to friction between neighboring parcels of the uid that are moving
at dierent velocities. When uid is forced through a tube, the uid generally moves
faster near the axis and very little near the walls, therefore some stress (such as a pressure
dierence between the two ends of the tube) is needed to overcome the friction between
layers and keep the uid moving. For the same velocity pattern, the stress is proportional
to the uids viscosity.
A uid that has no resistance to shear stress is known as an ideal uid or inviscid
uid. In the real world, zero viscosity is observed only at very low temperatures, in super
uids. Otherwise all uids have positive viscosity. If the viscosity is very high, such as in
pitch, the uid will seem to be a solid in the short term. In common usage, a liquid whose
viscosity is less than that of water is known as a mobile liquid, while a substance with a
viscosity substantially greater than water is simply called a viscous liquid.
When we deform a solid, so that it is strained, we know that the solid will exert a
restoring force which opposes the strain; for small strains the restoring force is proportional
to the strain, and we have the familiar Hookes Law. Real uids also oppose strains;
however, in a uid it is not the amount of strain which is important but the rate at which
the strain is produced. For instance, if you were to draw a utensil through your favorite
viscous uid (maple syrup, say), you would nd that it was easy for slow motions of the
utensil, but more dicult if the utensil is moved rapidly.
3.1 Shear Viscosity
The shear viscosity of a uid expresses its resistance to shearing ows, where adjacent
layers move parallel to each other with dierent speeds. It can be dened through the
idealized situation known as a Couette ow, where a layer of uid is trapped between two
horizontal plates, one xed and one moving horizontally at constant speed u. (The plates
are assumed to be very large, so that one need not consider what happens near their edges.)
If the speed of the top plate is small enough, the uid particles will move parallel to
it, and their speed will vary linearly from zero at the the bottom to u at the top. Each
layer of uid will move faster than the one just below it, and friction between them will
give rise to a force resisting their relative motion. In particular, the uid will apply on the
top plate a force in the direction opposite to its motion, and an equal but opposite to the
bottom plate. An external force is therefore required in order to keep the top plate moving
at constant speed.
The magnitude F of this force is found to be proportional to the speed u and the area
A of each plate, and inversely proportional to their separation y. That is,
F = A
u
y
(1)
3
3.1 Shear Viscosity 3 VISCOSITY
The proportionality factor in this formula is the viscosity (specically, the dynamicviscosity)
of the uid.
The ratio
u
y
is called the rate of shear deformation or shear velocity, and is the derivative
of the uid speed in the direction perpendicular to the plates. Isaac Newton expressed the
viscous forces by the dierential equation
=
u
y
(2)
where =
F
A
and
u
y
is the local shear velocity. This formula assumes that the ow is
moving along parallel lines and the y-axis, perpendicular to the ow, points in the direction
of maximum shear velocity. This equation can be used where the velocity does not vary
linearly with y, such as in uid owing through a pipe.
A uid which responds to a shear stress
F
A
in this manner is called a Newtonian uid ; it
has the property that the viscosity is independent of velocity. Most of the uids that we will
be interested in will be Newtonian (air, water). Non-Newtonian uids (Silly-Putty, paint,
polymeric uids) are more complicated and beyond the scope of the present description.
Some typical uid viscosities are given in Table- 3.3. Also listed in Table- 3.3 are the
kinematic viscosities . Recall that 1 pascal equals 1 kgm-1s-2 and note that the viscosity
of water is, conveniently, 1 mPas = 1 gm-1s-1; it is common to quote relative viscosities,
which are viscosities relative to water, or in mPas, and also to use the cgs unit of viscosity,
gcm-1s-1, called the poise (after Poiseuille) and equal to 0.1 Pa.s .
Fluid Viscosity () Kinematic viscosity ()
Pa.s m
2
s
1
Air (0

C) 1.7 10
5
1.32 10
5
Air (20

C) 1.8 10
5
1.32 10
5
Water (0

C) 1.8 10
3
1.32 10
5
Water (20

C) 1.0 10
3
1.32 10
5
Glycerin (20

C) 1.4 1.1 10
3
Blood (37

C) 4.0 10
3
Table 1: Viscosities of some common or interesting uids in SI units.
Before incorporating viscosity into the equations of uid mechanics, lets take a moment
to discuss some of the properties of viscosity. Viscosity does not depend upon pressure
in a signicant way, but you know from personal experience that it does depend upon
temperaturepancake syrup becomes less viscous after you take it out of the refrigerator
and let it warm up. Generally, the viscosity of liquids decreases with increasing tempera-
ture. Gases, on the other hand, have viscosities which generally increase with increasing
temperature. Why the dierence?
Viscosity is fundamentally a consequence of the intermolecular interactions in the uid.
In a dilute gas you can think in terms of binary collisions of pairs of molecules. Now
4
3.2 Kinematic viscosity 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
viscosity is the transfer of momentum from one part of a uid to an adjacent part, and in
a dilute gas this transfer will be more eective if the collisions are very energetic. Since
the mean kinetic energy of a molecule is (3/2)kBT, higher temperatures imply a larger
viscosity for a gas. A liquid, on the other hand, is a strongly interacting collection of
molecules, and the idea of a binary collision is meaningless. Microscopically the molecules
in a liquid do have some short range ordera given molecule is surrounded by a cage
of other molecules. Momentum transfer in the liquid will involve the movement of these
cages around one another. Increasing the temperature causes the cages to jiggle a bit
more, allowing them to slip more easily past one another, thereby reducing the momentum
transfer between adjacent bits of the uid, and thereby reducing the viscosity of the liquid.
More detail than this requires kinetic theory.
3.2 Kinematic viscosity
The kinematic viscosity is the dynamic viscosity divided by the density of the uid .
It is usually denoted by the Greek letter . It is a convenient concept when analyzing the
Reynolds number, that expresses the ratio of the inertial forces to the viscous forces:
Re =
uD

=
uD

(3)
3.3 Bulk viscosity
When a compressible uid is compressed or expanded evenly, without shear, it may still
exhibit a form of internal friction that resists its ow. These forces are related to the
rate of compression or expansion by a factor s, called the volume viscosity, bulk viscosity
or second viscosity. The bulk viscosity is important only when the uid is being rapidly
compressed or expanded, such as in sound and shock waves. Bulk viscosity explains the
loss of energy in those waves, as described by Stokes law of sound attenuation.
4 Viscosity eects at high Reynolds numbers
When the Reynolds number is large you might think that the viscosity could be ignored
altogether, in which case we return to non viscous uid mechanics. However, we once again
encounter dAlemberts paradoxthat a non viscous uid exerts no drag on a solid bodyso
we are at a loss when it comes to explaining aerodynamic drag. The important insight in
resolving this paradox is due to L. Prandtl, who in 1904 suggested that the viscosity could
be ignored everywhere except in a thin layer close to the surface of a body. Understanding
the behavior of this boundary layer has been crucial to the development of modern uid
mechanics and aerodynamics.
5
4.1 Boundary layers 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
4.1 Boundary layers
One of the simplest ow congurations which illustrates the boundary layer concept is the
ow of a uid parallel to a thin, at plate. If the uid were non viscous, the streamlines
would be parallel to the plate and nothing very interesting happens. For a viscous uid,
however, we must apply the no-slip boundary condition on the surface of the plate. The
thickness of the boundary layer, which will be denoted by , is the distance required for the
velocity prole to approach its free stream value. Recalling that the viscosity is a measure
of the diusion of velocity (or vorticity), the thickness of the boundary layer after a time
t is approximately given by

t (4)
Now in a time t an element of uid which begins at the leading edge of the plate will have
moved a distance x Ut , so that the boundary layer thickness a distance x from the
leading edge is

_
x
U
(5)
Therefore, the boundary layer thickness at the trailing edge of the plate, measured relative
to the length of the plate itself, is

x
l
x
Re

1
2
x
(6)
where we see that the boundary layer thickness decreases with increasing Reynolds number
(for an assumed laminar ow).
4.2 Skin friction
The boundary layer produces a drag on the plate due to the viscous stresses which are
developed at the wall. The viscous stress at the surface of the plate is

xy
(x, y = 0) =
v
x
y

y=0
(7)
Once this stress is known, we have only to integrate it over the surface of the plate to
obtain the total drag force D:
D =
_
lz
0
dz
_
lx
0
dx
xy
(x, y = 0) drag per side. (8)
To get an estimate of the velocity gradient
vx
y
near the wall, we note that by denition
the width of the boundary layer is the distance over which the velocity returns to its free
stream value, so
v
x
y

y=0

= U

U
x
(9)
6
4.3 Boundary layer separation and pressure drag 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
Performing the integral to obtain the drag, we nd
D l
z
_
lz
0
U

U
x
dx
= 2(l
x
l
z
)U
2
_

Ul
x
Dening the coecient of drag for the plate as C
d
=
D
rU
2
A/2
, with A = lxlz, we nd that
C
d
= 4Re

1
2
x
. A real calculation starting from the Navier-Stokes equation yields
C
d
= 1.33Re

1
2
x
(10)
not far from our simple estimate. This drag is often referred to as skin friction , and
is due to the viscous stresses acting on the surface of the plate. If the boundary layer
remains attached to the body (which it may not; see below), then this is the sole source
of aerodynamic drag on a body. At high Reynolds numbers, say 10
6
, this gives a drag
coecient of 10
3
, which is relatively small.
The previous analysis assumed that the ow in the boundary layer was laminar. How-
ever, in large Reynolds number ow we often encounter turbulent boundary layers , which
tend to produce a larger drag. The turbulent mixing of the uid near the surface of a
solid body leads to more ecient momentum transport away from the body, increasing the
gradient of the velocity prole at the surface and therefore the viscous stress on the plate.
For boundary layers which remain attached to a body the drag due to skin friction can be
reduced if the boundary layer can be persuaded to remain laminar.
4.3 Boundary layer separation and pressure drag
In most situations it is inevitable that the boundary layer becomes detached from a solid
body. This boundary layer separation results in a large increase in the drag on the body.
We can understand this by returning to the ow of a non viscous uid around a cylinder.
The pressure distribution is the same on the downstream side of the cylinder as on the
upstream side; thus, there were no unbalanced forces on the cylinder and therefore no drag
(dAlemberts paradox again). If the ow of a viscous uid about a body is such that the
boundary layer remains attached, then we have almost the same resultwell just have a
small drag due to the skin friction. However, if the boundary layer separates from the
cylinder, then the pressure on the downstream side of the cylinder is essentially constant,
and equal to the low pressure on the top and bottom points of the cylinder. This pressure is
much lower than the large pressure which occurs at the stagnation point on the upstream
side of the cylinder, leading to a pressure imbalance and a large pressure drag on the
cylinder. For instance, for a cylinder in a ow with a Reynolds number in the range
10
3
< Re < 10
5
, the boundary layer separates and the coecient of drag is C
d
1.2
, much larger that the coecient of drag due to skin friction, which we would estimate
7
4.3 Boundary layer separation and pressure drag 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
to be about 10
2
. A Reynolds number-independent drag coecient leads to a drag force
D
U
2
A
2
. More importantly, the power P required to maintain a constant speed in the
presence of this drag is P = DU =
U
3
A
2
, so that it increases with the cube of the speed.
To see the importance of this dependence, suppose that you ride a bicycle at 15 mph,
which is a respectable speed. Most of the resistance at this speed is due to aerodynamic
drag (there are other sources, such as mechanical friction, rolling friction, and so on, but
I dont think they dominate at this speed). Now suppose you want to get to class in half
the time in the morning, so you decide to ride at 30 mph. This requires 8 times as much
power!
Boundary layers tend to separate from a solid body when there is an increasing uid
pressure in the direction of the owthis is known as an adversepressuregradient in the
jargon of uid mechanics. Increasing the uid pressure is akin to increasing the potential
energy of the uid, leading to a reduced kinetic energy and a deceleration of the uid.
When this happens the boundary layer thickens (recall that
_
x
U
), leading to a
reduced gradient of the velocity prole (
vx
y
decreases), with a concomitant decrease in
the wall shear stress . For a large enough pressure gradient the shear stress can be reduced
to zero, and separation often occurs. The uid is no longer pulling on the wall, and
opposing ow can develop which eectively pushes the boundary layer o of the wall.
Separation is bound to occur in a suciently large adverse pressure gradient. On the other
hand, boundary layers like decreasing pressure gradients, which accelerate the uid and
cause the boundary layer to thin.
Given these considerations, we see that minimizing the pressure drag amounts to pre-
venting or delaying boundary layer separation. Since adverse pressure gradients are the
cause of separation, we want to avoid these or at least make the gradients small. Trailing
stagnation points are bound to cause problems, so separation can often be delayed by plac-
ing the trailing stagnation point at a cusp, so that the uid leaves the body smoothly. This
is known as streamlining , and is the preferred shape for airfoils, cars, and sh! Another
way of delaying separation is by forcing the boundary layer to become turbulent. The
more ecient mixing which occurs in a turbulent boundary layer reduces the boundary
layer thickness and increases the wall shear stress, often preventing the separation which
would occur for a laminar boundary layer under the same conditions. You can see that
there is a trade-o herethe turbulent boundary layer produces a greater drag due to skin
friction, but can often reduce the pressure drag by preventing, or reducing, boundary layer
separation. Since the latter is usually dominant at high Reynolds numbers, various schemes
have been invented for producing turbulent boundary layers. The dimples on a golf ball,
the fuzz on a tennis ball, or the seams on a baseball are good examples of this. Apparently
a dimpled golf ball has one-fth the drag of a smooth golf ball of the same size! Also, air-
plane wings are often engineered with vortex generators on the upper surface to produce
a turbulent boundary layer.
8
4.3 Boundary layer separation and pressure drag 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
4.3.1 Boundary-Layer Separation
The study of ow separation from the surface of a solid body, and the determination
of global changes in the ow eld that develop as a result of the separation, are among
the most fundamental and dicult problems of uid dynamics. It is well known that
most liquid and gas ows observed in nature and encountered in engineering applications
involve separation. This is because many of the common gases and liquids, such as
air and water, have extremely small viscosity and, therefore, most practical ows are
characterised by very large values of the Reynolds number; both theory and experiment
show that increasing Reynolds number almost invariably results in separation. In fact, to
achieve an unseparated form of the ow past a rigid body, rather severe restrictions must
be imposed on the shape of the body.
The dierence between a separated ow and its theoretical unseparated counterpart
(constructed solely on the basis of inviscid ow analysis) concerns not only the form of
trajectories of uid particles, but also the magnitudes of aerodynamic forces acting on the
body. For example, for blu bodies in an incompressible ow, it is known from experimental
observations that the drag force is never zero; furthermore, it does not approach zero as
the Reynolds number becomes large. On the other hand, one of the most famous results
of the inviscid ow theory is dAlemberts paradox which states that a rigid body does
not experience any drag in incompressible ow. It is well known that this contradiction is
associated with the assumption of a fully attached form of the ow; this situation almost
never happens in practice.
Separation imposes a considerable limitation on the operating characteristics of aircraft
wings, helicopter blades, turbines, etc., leading to a signicant degradation of their perfor-
mance. It is well known that the separation is normally accompanied by a loss of the lift
force, sharp increase of the drag, increase of the heat transfer at the reattachment region,
pulsations of pressure and, as a result, utter and buet onset.
It is hardly surprising that the problem of ow separation has attracted considerable
interest amongst researchers. The traditional approach of studying the separation phe-
nomenon is based on seeking possible simplications that may be introduced in the gov-
erning Navier-Stokes equations when the Reynolds number is large. The rst attempts
at describing separated ow past blunt bodies are due to Helmholtz (1868) and Kirchho
(1869) in the framework of the classical theory of inviscid uid ows, but there was no ad-
equate explanation as to why separation occurs. Prandtl (1904) was the rst to recognize
the physical cause of separation at high Reynolds numbers as being associated with the
separation of boundary layers that must form on all solid surfaces.
In accordance with the Prandtls theory, a high Reynolds number ow past a rigid body
has to be subdivided into two characteristic regions. The main part of the ow eld may be
treated as inviscid. However, for all Reynolds numbers, no matter how large, there always
exists a thin region near the wall where the ow is predominantly viscous. Prandtl termed
this region the boundary layer, and suggested that it is because of the specic behavior
of this layer that ow separation takes place. Flow development in the boundary layer
depends on the pressure distribution along the wall. If the pressure gradient is favorable,
9
4.3 Boundary layer separation and pressure drag 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
i.e. the pressure decreases downstream, then the boundary layer remains well attached to
the wall. However with adverse pressure gradient, when the pressure starts to rise in the
direction of the ow, the boundary layer tends to separate from the body surface. The
reason for separation was explained by Prandtl in the following way. Since the velocity in
the boundary layer drops towards the wall, the kinetic energy of uid particles inside the
boundary layer appears to be less than that at the outer edge of the boundary layer, in
fact the closer a uid particle is to the wall the smaller appears to be its kinetic energy.
This means that while the pressure rise in the outer ow may be quite signicant, the uid
particles inside the boundary layer may not be able to get over it. Even a small increase
of pressure may cause the uid particles near the wall to stop and then turn back to form
a recirculating ow region characteristic of separated ows.
It might seem surprising that the clear understanding of the physical processes leading
to the separation, could not be converted into a rational mathematical theory for more
than half a century. The fact is that the classical boundary-layer theory, which was in-
tended by Prandtl for predicting ow separation, was based on the so called hierarchical
approach when the outer inviscid ow should be calculated rst ignoring the existence of
the boundary layer, and only after that one can turn to the boundary layer analysis. By
the late forties it became obvious that such a strategy leads to a mathematical contra-
diction associated with so called Goldsteins singularity at the point of separation. The
form of this singularity was rst described by Landau & Lifshitz (1944) who demonstrated
that the shear stress in the body surface upstream of separation drops as the square root
of the distance from the separation, and the velocity component normal to the surface
tends to innity being inversely proportional to the shear stress. This result was later
conrmed based on more rigorous mathematical terms by Goldstein (1948). Goldstein also
proved and this result appeared to be of paramount importance for further development
of the boundary-layer separation theory that the singularity at separation precludes the
solution to be continued beyond the separation point into the region of reverse ow.
The Goldsteins theoretical discovery came at the time when an important development
was taking place in experimental investigation of separated ows. The most disputable was
the eect of upstream inuence through the boundary layer in supersonic ow prior to sep-
aration. It might be observed, for example, when a shock wave impinges the boundary layer
on a rigid body surface. Starting with Ferri (1939) a number of researches demonstrated
that instead of simple reection of the shock wave from the body surface, as it would
happen in a fully inviscid ow, more complicated shock structure, called the lambda-
structure, develops. It consists of the primary shock on impinging the boundary layer
at some point on the body surface, and the secondary shock which forms some distance
upstream of this point. The secondary shock is provoked by the thickening of the boundary
layer which, in its turn, is caused by propagation of disturbances through the boundary
layer from the region of higher pressure downstream of the main shock. This process, ob-
viously, can not be explained in the framework of classical boundary-layer theory. Indeed,
following the Prandtl hierarchical strategy, one has to consider the external inviscid ow
region rst and then the boundary layer. In the case of supersonic ow the external region
is governed by the hyperbolic equations. The boundary-layer equations, when solved with
10
4.3 Boundary layer separation and pressure drag 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
prescribed pressure gradient, as it is required by classical Prandtl theory, are known to be
of the parabolic type. Therefore neither the external ow nor the boundary layer allow
upstream propagation of disturbances.
Although the boundary-layer theory in its classical form was found to be insucient
for describing the separation phenomenon, Prandtls insight into physical processes leading
to separation and, even more so, the mathematical approach suggested by Prandtl for
analyzing high Reynolds number ows, laid a foundation for all subsequent studies in the
asymptotic theory of separation. In a broader sense Prandtls idea of subdividing the entire
ow eld into a number of regions with distinctively dierent ow properties, proved to be
a beginning of one of the most powerful tools in modern asymptotic analysis, the method
of matched asymptotic expansion.
Signicant progress in theoretical study of separated ows has been achieved in the
last thirty years, and a comprehensive description of the underlying ideas and the main
results of the theory may be found in a monograph by Sychev et al (1998). A key element
of the separation process, which was not fully appreciated in the classical Prandtls (1904)
description, is a mutual interaction between the boundary layer and the external inviscid
ow. Because of this interaction, a sharp pressure rise may develop spontaneously at a
location on the body surface where in accordance with the Prandtls theory the boundary
layer would be well attached. This pressure rise leads to a rapid deceleration of uid
particles near the wall and formation of the reverse ow downstream of the separation.
The interaction precludes development of the Goldstein singularity.
The asymptotic theory of viscous-inviscid interaction, known now as the triple-deck the-
ory, was formulated simultaneously by Neiland (1969) and Stewartson & Williams (1969)
for the self-induced separation in supersonic ow and by Messiter (1970) for incompress-
ible uid ow near a railing edge of a at plate. Based on the asymptotic analysis of the
Navier-Stokes equations they demonstrated the region of interaction is O(Re
3/8
) long and
has a three-tiered structure being composed of the viscous near-wall sublayer (region 1),
main part of the boundary layer (region 2) and inviscid potential ow region 3 situated
outside the boundary layer.
Characteristic thickness of the viscous sublayer is estimated an O(Re
5/8
) quantity,
i.e. it occupies O(Re
1/8
) portion of the boundary layer being comprised of the stream
laments immediately adjacent to the wall. The ow velocity in this region is O(Re
1/8
)
relative to the free-stream velocity, and due to the slow motion of gas here the ow exhibits
high sensitivity to pressure variations. Even small pressure rise along the wall may cause
signicant deceleration of uid particles there. This leads to thickening of ow laments,
and the streamlines change their shape being displaced from the wall.
The main part of the boundary layer, the middle tier of the interactive structure,
represents a continuation of the conventional boundary layer into the region of interaction.
Its thickness is estimated as O(Re
1/2
) and the velocity is an order one quantity. The ow
in this tier is signicantly less sensitive to the pressure variations. It does not produce any
noticeable contribution to the displacement eect of the boundary layer, which means that
all the stream lines in the middle tier are parallel to each other and carry the deformation
produced by the displacement eect of the viscous sublayer.
11
4.4 Laminar-Turbulent Transition 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
Finally, the upper tier is situated in the potential ow region outside the boundary layer.
It serves to convert the perturbations in the form of the stream lines into perturbations
of pressure. These are then transmitted through the main part of the boundary layer back
to the sublayer.
Later it became clear that the triple-deck interaction region, while being small, plays a
key role in many uid ows. It, for instance, governs upstream inuence in the supersonic
boundary layer, development of dierent modes of instabilities, bifurcation of the solution
and possible hysteresis in separated ows. As far as separation phenomena are concerned,
the theory has been extended to describe boundary-layer separation from a smooth body
surface in incompressible uid ow, supersonic ow separation provoked by a shock wave
impinging upon the boundary layer, incipient and large scale separations at angular points
of the body contour both in subsonic and supersonic ows, separation at the trailing edge
of a thin aerofoil appearing as a result of increase of the angle of attack or the aerofoil
thickness, leading-edge separation, separation of the boundary layer in hypersonic ow on
a hot or cold wall, separation provoked by a wall roughness, etc. (see Sychev et al. 1998
and references in this book.)
However, despite obvious progress in this eld, many aspects of the boundary-layer
separation theory remain unresolved. Most notably, the theory remains predominantly
restricted to incipient or small scale separations when the entire recirculating region to-
gether with the separation and reattachment points ts into the O(Re
3/8
) long region
of interaction. Even in the studies specically aimed at describing developed separations
(see Neiland 1969; Stewartson & Williams 1969; Sychev 1972; Ruban 1974), the analysis
is conned to the local ow behavior near the separation point.
Very little is still known about developed separations, separation in transonic ow,
three-dimensional boundary-layer separation, unsteady separation, etc. Future research in
these areas should be exciting.
4.4 Laminar-Turbulent Transition
Laminar-turbulent transition is an extraordinarily complicated process, consisting of a
great number of competing events. The initial process is the transformation of external
disturbances into internal instability oscillations of the boundary layer, taking the well-
known form of Tollmien-Schlichting waves. In relatively quiet ows, the initial amplitude
of these waves is insucient to provoke immediate transition. Tollmien-Schlichting waves
must rst amplify in the boundary layer to trigger non-linear eects, characteristic of
the transition process. Numerous experiments have clearly revealed that the extent of
the amplication region, and hence the location of the transition point on the body
surface, is strongly dependent not only on the amplitude and/or the spectrum of external
disturbances but also on their physical nature. Some of the disturbances easily penetrate
into the boundary layer and turn into Tollmien-Schlichting waves; others do not. To study
these dierences, the boundary-layer receptivity to external disturbances was proposed
by Morkovin (1969) as a key problem in the laminar-turbulent transition process. The
main objective of the investigations in this eld is the determination of the amplitudes
12
4.4 Laminar-Turbulent Transition 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
of generated Tollmien-Schlichting waves, and as a result the elucidation of which types of
external disturbances can more easily provoke Tollmien-Schlichting waves.
From the mathematical point of view, the receptivity issue appears to be even more
dicult than the stability problem. The latter is often associated with the solution of
the Orr-Sommereld equation, while the former involves the solution of a boundary-value
problem for the Navier-Stokes equations. To date, direct numerical simulation of the full
Navier-Stokes equations appears to be exceedingly dicult as far as unstable boundary-
layer ows at high Reynolds number are concerned. On the other hand, asymptotic meth-
ods are well-suited for this type of problem.
The rst paper where asymptotic techniques were used to solve a receptivity prob-
lem was published by Terentev (1981). His analysis was devoted to Tollmien-Schlichting
waves generated by a vibrator installed in the boundary layer on the body surface. The
classical experiments of Dryden (1956) and Schubauer & Skramsted (1948) were the rst
to generate Tollmien-Schlichting waves in a wind tunnel by means of a vibrating ribbon.
This technique is simple, and still used to provoke Tollmien-Schlichting waves in exper-
imental studies. In the theoretical analysis of Terentev (1981), the role of the vibrator
was modeled by a short, exible section of the body surface, and to describe the Tollmien-
Schlichting wave generation process, he used an unsteady version of triple-deck theory. It
was shown previously by Lin (1946), Smith (1979) and Zhuk & Ryzhov (1980) that this
provides the asymptotic description of Tollmien-Schlichting waves near the lower branch
of the boundary-layer neutral stability curve. To satisfy the restrictions of triple-deck
theory, the (non-dimensional) longitudinal extent l of the vibrating part of the body sur-
face was chosen to be l = O(Re
3/8
), while the (non-dimensional) frequency of oscillation
= O(Re
1/4
), where Re is the Reynolds number. As a result of this analysis, an explicit
formula may be obtained for the amplitude of a Tollmien-Schlichting wave propagating in
the boundary layer, downstream of the vibrator.
The generation of Tollmien-Schlichting waves via the interaction of an external acoustic
eld with a wall roughness element was investigated by Ruban (1984) and independently
by Goldstein (1985). Eective transformation of external disturbances into Tollmien-
Schlichting waves is possible if resonance conditions are satised. For boundary-layer ows,
the resonance between external disturbances and internal instability waves supposes coin-
cidence not only of frequencies, but also of wave numbers. This coincidence may easily be
achieved in the problem considered by Terentev (1981) since the frequency of the vibrat-
ing section of the surface and its extent may be chosen independently. If an acoustic wave
plays the role of the external perturbation, it is possible to choose its (non-dimensional)
frequency to be = O(Re
1/4
), but then the wavelength of the disturbance is much larger
than that appropriate to Tollmien-Schlichting waves. In order to introduce the necessary
length scale into the problem, Ruban (1984) and Goldstein (1985) supposed that the body
surface was not absolutely smooth, and the acoustic wave interacts with a stationary dis-
turbance provoked in the boundary layer by wall roughness. The eective generation of
Tollmien-Schlichting waves was shown to take place if the (non-dimensional) length l of
the roughness is O(Re
3/8
).
It is well known from experimental observations that the laminar-turbulent transi-
13
4.5 Law of the wall 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
tion process is sensitive not only to wall vibrations and/or acoustic elds, but is also
strongly aected by free-stream turbulence. The above mentioned paper by Duck, Ruban
& Zhikharev is devoted to the analysis of Tollmien-Slichting wave generation by external
vorticity perturbations interacting with a small wall roughness.
4.5 Law of the wall
In uid dynamics, the law of the wall states that the average velocity of a turbulent ow
at a certain point is proportional to the logarithm of the distance from that point to the
wall, or the boundary of the uid region. This law of the wall was rst published by
Theodore von Krmn, in 1930. It is only technically applicable to parts of the ow that are
close to the wall (20% of the height of the ow), though it is a good approximation for
the entire velocity prole of natural streams
4.5.1 General logarithmic formulation
The logarithmic law of the wall is a self similar solution for the mean velocity parallel
to the wall, and is valid for ows at high Reynolds numbers in an overlap region with
approximately constant shear stress and far enough from the wall for (direct) viscous eects
to be negligible:
u
+
=
1

lny
+
+ C
+
with y
+
=
u

, u

=
_

and u
+
=
u
u

(11)
where
y
+
is the wall coordinate: the distance y to the wall, made dimensionless with the friction
velocity u

and kinematic viscosity ,


u
+
is the dimensionless velocity: the velocity u parallel to the wall as a function of y
(distance from the wall), divided by the friction velocity u

is the wall shear stress,


is the uid density,
u

is called the friction velocity or shear velocity


is the Von Von Karm an constant,
C
+
is a constant, and
ln is the natural logarithm.
14
4.5 Law of the wall 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
From experiments, the Von K arman constant is found to be 0.41 and C
+
5.0 for a
smooth wall. With dimensions, the logarithmic law of the wall can be written as:
u =
u

log
y
y
0
(12)
where y
0
is the distance from the boundary at which the idealized velocity given by
the law of the wall goes to zero. This is necessarily nonzero because the turbulent velocity
prole dened by the law of the wall does not apply to the laminar sublayer. The distance
from the wall at which it reaches zero is determined by comparing the thickness of the
laminar sublayer with the roughness of the surface over which it is owing. For a near-wall
laminar sublayer of thickness and a characteristic roughness length-scale k
s
,
k
s
<

: hydraulically smooth flow,


k
s

: transitional flow,
k
s
>

: hydraulically rough flow,


Intuitively, this means that if the roughness elements are hidden within the laminar sub-
layer, they have a much dierent eect on the turbulent law of the wall velocity prole
than if they are sticking out into the main part of the ow. This is also often more formally
formulated in terms of a boundary Reynolds number, Re
w
, where
Re

=
u

k
s

(13)
The ow is hydraulically smooth for Re

< 3, hydraulically rough for Re

> 100, and


transitional for intermediate values. Values for y
0
are given by:
y
0
=

9u
for hydraulically smooth flow,
y
0
=
ks
30
hydraulically rough flow,
Intermediate values are generally given by the empirically-derived Nikuradse diagram,
though analytical methods for solving for this range have also been proposed. For channels
with a granular boundary, such as natural river systems,
k
s
3.5D
84
where D
84
is the average diameter of the 84
th
largest percentile of the grains of the bed
material.
15

You might also like