You are on page 1of 26

Non-Dimensional Navier-Stokes Equations

Kiran Kumar
September 19, 2014

1 NAVIER-STOKES EQUATION

1
1.1

Navier-Stokes equation
Differential form
~
~
~v F~v
~
E
F
E
Q
+
+
=
+
t
x
y
x
y

u
v
2

vu

~ = u , E
~ = u + p , F
~ =
Q
2
v

v + p
uv
Et
(Et + p)u
(Et + p)v

0
0

xx
xy
, F~v =

~v =
E

xy
yy
uxx + vxy qx
uxy + vyy qy

Stress vectors

yy





u v
u v
2

+
, xy = yx =
,
xx = 2
3
x y
y x


v u
T
Cp
R
2
T

, qy = k
, k=
=
= 2
qx = k
3
y x
x
y
Pr
( 1)P r
h
i

2
p = ( 1) Et
u + v2
2

1.2 Integral form

1.2

1 NAVIER-STOKES EQUATION

Integral form

~ d +
W

~c F
~ v )ds =
(F

~
Qd

V
0
u
uV + nx p
nx xx + ny xy + nz xz

~ = v , F
~ c = vV + ny p , F
~ v = nx yx + ny yy + nz yz
W

w
wV + nz p
nx zx + ny zy + nz zz
E
HV
nx x + ny y + nz z

0
,
x = uxx + vxy + wxz + k T
x

fe,x

,
~ =
fe,y
y = uyx + vyy + wyz + k T
,
Q
y

fe,z
~
f e ~
v + qh
z = uzx + vzy + wzz + k T
z

V =~
v~
n = nx u + ny v + nx w


(u2 +v2 +w2 )
p = ( 1) E
2
p
H = E + p , k = C
Pr


u v w
u
xx =
+
+
,
+ 2
x y
z
x


u v w
v
yy =
+
+
+ 2
,
x y
z
y


u v w
w
zz =
+
+
+ 2
,
x y
z
z

R
(1)P r


u v
xy = yx =
+
y x


u w
xz = zx =
+
z
x


v w
yz = zy =
+
z
y

2
=
3

1.3

Spatial Discretisation

Assuming a particular control volume does not change in time, the time derivative of the
~ can be cast in the form
conservative variables W
Z
~

~ d = W
W
t
t
Finally,
~
W
1
=
t

I

~c F
~ v )ds
(F

~
Qd

1.3 Spatial Discretisation

1 NAVIER-STOKES EQUATION

If we consider a particular volume I,J,K then


"N
#
F
X
~ I,J,K
dW
1
~c F
~ v )m Sm (Q)
~ I,J,K
=
(F
dt
I,J,K m=1
The indices in capital letters (I, J, K) reference the control volume, NF denotes the number
of control volume faces and Sm stands for the area of the face m. The term in square
brackets on the right-hand side of the above equation is generally termed as the residual.
~ I,J,K . Hence,
It is denoted here by R
~ I,J,K
1 ~
dW
=
RI,J,K
dt
I,J,K

In the framework of finite volume schemes control volume can be defined as cell centered
or vertex centred.

2 NON DIMENSIONAL NAVIER-STOKES EQUATIONS

Non dimensional Navier-Stokes equations

2.1

Case 1
Lr = L,

2.1.1

t =

ta
,
L

u =

u
,
a

Ur = a ,

r = ,
y = Ly ,

x = Lx ,
v =

v
,
a

p =

Tr = T ,
=

p
,
a2

r =
k =

E t =

k
k

Et
a2

Momentum Equation

On LHS,
(u)
t


=

a2
L

( u )
t

On RHS,
xx

2.1.2


 


2
u v
a  2 u v
=
2

2
=
3
x y
L
3
x
y


 a   1 
2 u v

2
=
L
a2 LHS 3
x
y





2 u v
M
2
=
U L
3
x
y


M 2
u v
=
2

Re 3
x y

Energy Equation

On LHS,
=

a3
L

a2
L

u x y

a2
L



(Et )
t

(Et )
t

On RHS,
uxx =
=
=

M
uxy
Re

1
a3

u x y
LHS

2.1 Case 1

qx =
=
=
=

2 NON DIMENSIONAL NAVIER-STOKES EQUATIONS





Cp T
R
T
T
=
=
k
x
P r x
( 1)P r x





T
RT
a2
T

=
( 1)P rL
x
( 1)L P r x






a2
1
T
M
T

=
3

( 1)L
a LHS P r x
U L
( 1) P r x


T
M
( 1)Re P r x

3 VISCOSITY

Viscosity

The viscosity of a fluid 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 fluid that are moving
at different velocities. When fluid is forced through a tube, the fluid generally moves
faster near the axis and very little near the walls, therefore some stress (such as a pressure
difference between the two ends of the tube) is needed to overcome the friction between
layers and keep the fluid moving. For the same velocity pattern, the stress is proportional
to the fluids viscosity.
A fluid that has no resistance to shear stress is known as an ideal fluid or inviscid
fluid. In the real world, zero viscosity is observed only at very low temperatures, in super
fluids. Otherwise all fluids have positive viscosity. If the viscosity is very high, such as in
pitch, the fluid 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 fluids also oppose strains;
however, in a fluid 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 fluid (maple syrup, say), you would find that it was easy for slow motions of the
utensil, but more difficult if the utensil is moved rapidly.

3.1

Shear Viscosity

The shear viscosity of a fluid expresses its resistance to shearing flows, where adjacent
layers move parallel to each other with different speeds. It can be defined through the
idealized situation known as a Couette flow, where a layer of fluid is trapped between two
horizontal plates, one fixed 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 fluid 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 fluid 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 fluid 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
7

u
y

(1)

3.1 Shear Viscosity

3 VISCOSITY

The proportionality factor in this formula is the viscosity (specifically, the dynamicviscosity)
of the fluid.
The ratio uy is called the rate of shear deformation or shear velocity, and is the derivative
of the fluid speed in the direction perpendicular to the plates. Isaac Newton expressed the
viscous forces by the differential equation
=

u
y

(2)

is the local shear velocity. This formula assumes that the flow is
where = FA and u
y
moving along parallel lines and the y-axis, perpendicular to the flow, 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 fluid flowing through a pipe.
A fluid which responds to a shear stress FA in this manner is called a Newtonian fluid ; it
has the property that the viscosity is independent of velocity. Most of the fluids that we will
be interested in will be Newtonian (air, water). Non-Newtonian fluids (Silly-Putty, paint,
polymeric fluids) are more complicated and beyond the scope of the present description.
Some typical fluid viscosities are given in Table- 3.1. Also listed in Table- 3.1 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
m2 s1
Air (0 C)
1.7 105
1.32 105
Air (20 C)
1.8 105
1.32 105

3
Water (0 C)
1.8 10
1.32 105
Water (20 C)
1.0 103
1.32 105

Glycerin (20 C) 1.4


1.1 103
Blood (37 C)
4.0 103
Table 1: Viscosities of some common or interesting fluids in SI units.

Before incorporating viscosity into the equations of fluid mechanics, lets take a moment
to discuss some of the properties of viscosity. Viscosity does not depend upon pressure
in a significant 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 temperature. Gases, on the other hand, have viscosities which generally increase with increasing
temperature. Why the difference?
Viscosity is fundamentally a consequence of the intermolecular interactions in the fluid.
In a dilute gas you can think in terms of binary collisions of pairs of molecules. Now
8

3.2 Kinematic viscosity 4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS


viscosity is the transfer of momentum from one part of a fluid to an adjacent part, and in
a dilute gas this transfer will be more effective 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 fluid, 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 fluid .
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 =

3.3

uD
uD
=

(3)

Bulk viscosity

When a compressible fluid is compressed or expanded evenly, without shear, it may still
exhibit a form of internal friction that resists its flow. 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 fluid 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.

Viscosity effects 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 fluid mechanics. However, we once again
encounter dAlemberts paradoxthat a non viscous fluid 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 fluid
mechanics and aerodynamics.

4.1 Boundary layers

4.1

4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS

Boundary layers

One of the simplest flow configurations which illustrates the boundary layer concept is the
flow of a fluid parallel to a thin, flat plate. If the fluid were non viscous, the streamlines
would be parallel to the plate and nothing very interesting happens. For a viscous fluid,
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 profile to approach its free stream value. Recalling that the viscosity is a measure
of the diffusion of velocity (or vorticity), the thickness of the boundary layer after a time
t is approximately given by

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

(5)
U
Therefore, the boundary layer thickness at the trailing edge of the plate, measured relative
to the length of the plate itself, is
x
1
Rex 2
(6)
lx
where we see that the boundary layer thickness decreases with increasing Reynolds number
(for an assumed laminar flow).

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

vx
(7)
xy (x, y = 0) =
y y=0
Once this stress is known, we have only to integrate it over the surface of the plate to
obtain the total drag force D:
Z lz Z lx
D=
dz
dx xy (x, y = 0) drag per side.
(8)
0

x
To get an estimate of the velocity gradient v
near the wall, we note that by definition
y
the width of the boundary layer is the distance over which the velocity returns to its free
stream value, so
s

vx
U
U

=U
(9)

y y=0

10

4.3 Boundary layer separation


4 VISCOSITY
and pressure
EFFECTS
drag AT HIGH REYNOLDS NUMBERS
Performing the integral to obtain the drag, we find
s
Z
lz

U
dx
x
0
r

2
= 2(lx lz )U
U lx

D lz

Defining the coefficient of drag for the plate as Cd =


12

Cd = 4Rex

D
,
rU 2 A/2

with A = lxlz, we find that

. A real calculation starting from the Navier-Stokes equation yields


1

Cd = 1.33Rex 2

(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 106 , this gives a drag
coefficient of 103 , which is relatively small.
The previous analysis assumed that the flow in the boundary layer was laminar. However, in large Reynolds number flow we often encounter turbulent boundary layers , which
tend to produce a larger drag. The turbulent mixing of the fluid near the surface of a
solid body leads to more efficient momentum transport away from the body, increasing the
gradient of the velocity profile 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 flow of a non viscous fluid 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 flow of a viscous fluid 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 flow with a Reynolds number in the range
103 < Re < 105 , the boundary layer separates and the coefficient of drag is Cd 1.2
, much larger that the coefficient of drag due to skin friction, which we would estimate
11

4.3 Boundary layer separation


4 VISCOSITY
and pressure
EFFECTS
drag AT HIGH REYNOLDS NUMBERS
to be about 102 . A Reynolds number-independent drag coefficient leads to a drag force
2
D U2 A . More importantly, the power P required to maintain a constant speed in the
3
presence of this drag is P = DU = U2 A , 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 fluid
pressure in the direction of the flowthis is known as an adversepressuregradient in the
jargon of fluid mechanics. Increasing the fluid pressure is akin to increasing the potential
energy of the fluid, leading to a reduced kinetic energy and a deceleration
of the fluid.
p x
),
leading
to a
When this happens the boundary layer thickens (recall that
U
vx
reduced gradient of the velocity profile ( 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 fluid is no longer pulling on the wall, and
opposing flow can develop which effectively pushes the boundary layer off of the wall.
Separation is bound to occur in a sufficiently large adverse pressure gradient. On the other
hand, boundary layers like decreasing pressure gradients, which accelerate the fluid and
cause the boundary layer to thin.
Given these considerations, we see that minimizing the pressure drag amounts to preventing 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 placing the trailing stagnation point at a cusp, so that the fluid leaves the body smoothly. This
is known as streamlining , and is the preferred shape for airfoils, cars, and fish! Another
way of delaying separation is by forcing the boundary layer to become turbulent. The
more efficient 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-off 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-fifth the drag of a smooth golf ball of the same size! Also, airplane wings are often engineered with vortex generators on the upper surface to produce
a turbulent boundary layer.

12

4.3 Boundary layer separation


4 VISCOSITY
and pressure
EFFECTS
drag AT HIGH REYNOLDS NUMBERS
4.3.1

Boundary-Layer Separation

The study of flow separation from the surface of a solid body, and the determination
of global changes in the flow field that develop as a result of the separation, are among
the most fundamental and difficult problems of fluid dynamics. It is well known that
most liquid and gas flows 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 flows are
characterized 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 flow past a rigid body, rather severe restrictions must
be imposed on the shape of the body.
The difference between a separated flow and its theoretical unseparated counterpart
(constructed solely on the basis of inviscid flow analysis) concerns not only the form of
trajectories of fluid particles, but also the magnitudes of aerodynamic forces acting on the
body. For example, for bluff bodies in an incompressible flow, 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 flow theory is dAlemberts paradox which states that a rigid body does
not experience any drag in incompressible flow. It is well known that this contradiction is
associated with the assumption of a fully attached form of the flow; 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 significant degradation of their performance. 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, flutter and buffet onset.
It is hardly surprising that the problem of flow separation has attracted considerable
interest amongst researchers. The traditional approach of studying the separation phenomenon is based on seeking possible simplifications that may be introduced in the governing Navier-Stokes equations when the Reynolds number is large. The first attempts
at describing separated flow past blunt bodies are due to Helmholtz (1868) and Kirchhoff
(1869) in the framework of the classical theory of inviscid fluid flows, but there was no adequate explanation as to why separation occurs. Prandtl (1904) was the first 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 flow past a rigid body
has to be subdivided into two characteristic regions. The main part of the flow field 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 flow is predominantly viscous. Prandtl termed
this region the boundary layer, and suggested that it is because of the specific behavior
of this layer that flow separation takes place. Flow development in the boundary layer
depends on the pressure distribution along the wall. If the pressure gradient is favorable,
13

4.3 Boundary layer separation


4 VISCOSITY
and pressure
EFFECTS
drag 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 flow, 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 fluid particles inside the
boundary layer appears to be less than that at the outer edge of the boundary layer, in
fact the closer a fluid particle is to the wall the smaller appears to be its kinetic energy.
This means that while the pressure rise in the outer flow may be quite significant, the fluid
particles inside the boundary layer may not be able to get over it. Even a small increase
of pressure may cause the fluid particles near the wall to stop and then turn back to form
a recirculating flow region characteristic of separated flows.
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 intended by Prandtl for predicting flow separation, was based on the so called hierarchical
approach when the outer inviscid flow should be calculated first 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 contradiction associated with so called Goldsteins singularity at the point of separation. The
form of this singularity was first 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 infinity being inversely proportional to the shear stress. This result was later
confirmed 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 flow.
The Goldsteins theoretical discovery came at the time when an important development
was taking place in experimental investigation of separated flows. The most disputable was
the effect of upstream influence through the boundary layer in supersonic flow prior to separation. 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 reflection of the shock wave from the body surface, as it would
happen in a fully inviscid flow, more complicated shock structure, called the lambdastructure, 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, obviously, 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 flow
region first and then the boundary layer. In the case of supersonic flow the external region
is governed by the hyperbolic equations. The boundary-layer equations, when solved with
14

4.3 Boundary layer separation


4 VISCOSITY
and pressure
EFFECTS
drag 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 flow nor the boundary layer allow
upstream propagation of disturbances.
Although the boundary-layer theory in its classical form was found to be insufficient
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 flows, laid a foundation for all subsequent studies in the
asymptotic theory of separation. In a broader sense Prandtls idea of subdividing the entire
flow field into a number of regions with distinctively different flow properties, proved to be
a beginning of one of the most powerful tools in modern asymptotic analysis, the method
of matched asymptotic expansion.
Significant progress in theoretical study of separated flows 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
flow. 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 fluid
particles near the wall and formation of the reverse flow 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 theory, was formulated simultaneously by Neiland (1969) and Stewartson & Williams (1969)
for the self-induced separation in supersonic flow and by Messiter (1970) for incompressible fluid flow near a railing edge of a flat plate. Based on the asymptotic analysis of the
Navier-Stokes equations they demonstrated the region of interaction is O(Re3/8 ) 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 flow region 3 situated
outside the boundary layer.
Characteristic thickness of the viscous sublayer is estimated an O(Re5/8 ) quantity,
i.e. it occupies O(Re1/8 ) portion of the boundary layer being comprised of the stream
filaments immediately adjacent to the wall. The flow velocity in this region is O(Re1/8 )
relative to the free-stream velocity, and due to the slow motion of gas here the flow exhibits
high sensitivity to pressure variations. Even small pressure rise along the wall may cause
significant deceleration of fluid particles there. This leads to thickening of flow filaments,
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(Re1/2 ) and the velocity is an order one quantity. The flow
in this tier is significantly less sensitive to the pressure variations. It does not produce any
noticeable contribution to the displacement effect 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 effect of the viscous sublayer.
15

4.4 Laminar-Turbulent Transition


4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
Finally, the upper tier is situated in the potential flow 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 fluid flows. It, for instance, governs upstream influence in the supersonic
boundary layer, development of different modes of instabilities, bifurcation of the solution
and possible hysteresis in separated flows. As far as separation phenomena are concerned,
the theory has been extended to describe boundary-layer separation from a smooth body
surface in incompressible fluid flow, supersonic flow 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 flows, 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 flow 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 field, 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 together with the separation and reattachment points fits into the O(Re3/8 ) long region of
interaction. Even in the studies specifically aimed at describing developed separations (see
Neiland 1969; Stewartson & Williams 1969; Sychev 1972; Ruban 1974), the analysis is confined to the local flow behavior near the separation point. Very little is still known about
developed separations, separation in transonic flow, 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 wellknown form of Tollmien-Schlichting waves. In relatively quiet flows, the initial amplitude
of these waves is insufficient to provoke immediate transition. Tollmien-Schlichting waves
must first amplify in the boundary layer to trigger non-linear effects, characteristic of
the transition process. Numerous experiments have clearly revealed that the extent of
the amplification 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 differences, 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 field is the determination of the amplitudes
of generated Tollmien-Schlichting waves, and as a result the elucidation of which types of
16

4.4 Laminar-Turbulent Transition


4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS
external disturbances can more easily provoke Tollmien-Schlichting waves.
From the mathematical point of view, the receptivity issue appears to be even more
difficult than the stability problem. The latter is often associated with the solution of
the Orr-Sommerfield 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 difficult as far as unstable boundarylayer flows at high Reynolds number are concerned. On the other hand, asymptotic methods are well-suited for this type of problem.
The first paper where asymptotic techniques were used to solve a receptivity problem 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 first
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 experimental studies. In the theoretical analysis of Terentev (1981), the role of the vibrator
was modeled by a short, flexible section of the body surface, and to describe the TollmienSchlichting 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 surface was chosen to be l = O(Re3/8 ), while the (non-dimensional) frequency of oscillation
= O(Re1/4 ), 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
field with a wall roughness element was investigated by Ruban (1984) and independently
by Goldstein (1985). Effective transformation of external disturbances into TollmienSchlichting waves is possible if resonance conditions are satisfied. For boundary-layer flows,
the resonance between external disturbances and internal instability waves supposes coincidence 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 vibrating 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(Re1/4 ), 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 disturbance provoked in the boundary layer by wall roughness. The effective generation of
Tollmien-Schlichting waves was shown to take place if the (non-dimensional) length l of
the roughness is O(Re3/8 ).
It is well known from experimental observations that the laminar-turbulent transition process is sensitive not only to wall vibrations and/or acoustic fields, but is also
17

4.5 Law of the wall

4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS

strongly affected 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 fluid dynamics, the law of the wall states that the average velocity of a turbulent flow
at a certain point is proportional to the logarithm of the distance from that point to the
wall, or the boundary of the fluid region. This law of the wall was first published by
Theodore von Krmn, in 1930. It is only technically applicable to parts of the flow that are
close to the wall (20% of the height of the flow), though it is a good approximation for
the entire velocity profile 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 flows at high Reynolds numbers in an overlap region with
approximately constant shear stress and far enough from the wall for (direct) viscous effects
to be negligible:
r
u

u
1
+
+
+
+
(11)
with y = , u =
and u+ =
u = lny + C

u
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 fluid density,
u is called the friction velocity or shear velocity
is the Von Von Karman constant,
C + is a constant, and
ln is the natural logarithm.
From experiments, the Von Karman 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=

log
18

y
y0

(12)

4.5 Law of the wall

4 VISCOSITY EFFECTS AT HIGH REYNOLDS NUMBERS

where y0 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
profile defined 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 flowing. For a near-wall
laminar sublayer of thickness and a characteristic roughness length-scale ks ,
ks < :
ks :
ks > :

hydraulically smooth f low,


transitional f low,
hydraulically rough f low,

Intuitively, this means that if the roughness elements are hidden within the laminar sublayer, they have a much different effect on the turbulent law of the wall velocity profile
than if they are sticking out into the main part of the flow. This is also often more formally
formulated in terms of a boundary Reynolds number, Rew , where
Re =

u ks

(13)

The flow is hydraulically smooth for Re < 3, hydraulically rough for Re > 100, and
transitional for intermediate values. Values for y0 are given by:

9u
ks
y0 = 30

y0 =

f or

hydraulically

hydraulically

smooth f low,

rough f low,

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,
ks 3.5D84
where D84 is the average diameter of the 84th largest percentile of the grains of the bed
material.

19

5 DIMENSIONLESS NUMBERS IN HEAT TRANSFER

Dimensionless Numbers in Heat Transfer

Dimensionless numbers described in this article are the most common numbers used in
heat transfer:
1. Reynolds Number
2. Nusselt Number
3. Prandtl Number
4. Grashof Number
5. Rayleigh Number

5.1
5.1.1

Physical properties of fluids


Density

Mass of fluid contained in a unit volume. Its units are Kg/m3 or slugs/f t3 . Typical
values: Water = 1000 kg/m3 , Mercury = 13546 kg/m3 , Air = 1.23 kg/m3 , Paraffin Oil =
800 kg/m3 . (at pressure = 1.013 e+5 Pascals and Temperature = 288.15 K.)
5.1.2

Viscosity

Viscosity, , is the property of a fluid, due to cohesion and interaction between molecules,
which offers resistance to sheer deformation of the fluid. Different fluids deform at different
rates under the same shear forces. Fluid with a high viscosity such as syrup, deforms more
slowly than fluid with a low viscosity such as water. All fluids are viscous, Newtonian
, where,
Fluids obey the linear relationship given by Newtons law of viscosity = du
dy
is the shear stress, is the coefficient of dynamic viscosity - The Coefficient of Dynamic
Viscosity, , is defined as the shear force, per unit area, (or shear stress ), required to
drag one layer of fluid with unit velocity past another layer a unit distance away. Units:
Newton seconds per square meter, or Kilograms per meter per second. (Although note
that is often expressed in Poise, P, where 10 P = 1 kgm1 s1 .) Typical values: Water
=1.14 e3 kgm1 s1 , Air =1.78 e5 kgm1 s1 , Mercury =1.552kgm1 s1 , Paraffin Oil
=1.9 kgm1 s1 . Kinematic Viscosity, , is defined as the ratio of dynamic viscosity to
mass density, = . Units: square meters per second, (Although note that is often
expressed in Stokes, St, where St = 1 e4 m2 /s .) Dimensions: . Typical values: Water
=1.14 e6 m2 /s , Air =1.46 e5 m2 /s , Mercury =1.145 e4 m2 /s , Paraffin Oil =2.375
e3 m2 /s .

20

5.1 Physical properties of fluids


5 DIMENSIONLESS NUMBERS IN HEAT TRANSFER
5.1.3

Thermal Conductivity

Thermal conductivity is a measure of the ability of a material to conduct heat. It is defined


using the Fouriers law of condution which, relates the rate of heat transfer by conduction
to the temperature gradient:
dT
(14)
Q = kA
dx
where k is the thermal conductivity. Using the Fouriers law we can define the thermal
conductivity as the rate of heat transfer through a unit thickness of a material per unit
area and per unit temperature difference. A good conductor of heat has a high value
of thermal conductivity. The thermal conductivity is expressed in the units of (energy
rate/(length.Temperature) ). In metric system, its unit is W/mK.
Thermaml conductivity of most materials vary with temperature. For example:
T (K)
100
200
300
400
600
800

Copper
482
413
401
393
379
366

Aluminium
302
237
237
240
231
218

For both cases the thermal conductivity decreases with temperature.Thermal conductivity of most liquids decrease with increasing temperature. Water is, however, an exception
to this rule. According to the kinetic theory of gases, the thermal conductivity of gases
is proportional to the square root of the absolute temperature and inversely proportional
to the square root of the molar mass. It is obvious that the thermal conductivity of a gas
increases with the increasing temperature.
5.1.4

Specific Heat

Specific heat is the amount of heat that is required to raise the temperature of a unit mass
of a substance by one degree. In a constant pressure process
Q = mC
p T

(15)

where Cp is the specific heat at constant pressure. The units for the specific heat are kJ/kg
K (or C).Typical values of Cp for various materials (at 300 K) are shown below:
Material
Aluminium (pure)
Copper (pure)
Gold
Silicon
Water
Air
21

Cp (kJ/kg.K)
903
385
129
712
4180
1005

5.2 Reynolds Number


5.1.5

5 DIMENSIONLESS NUMBERS IN HEAT TRANSFER

Coefficient of Thermal Expansion

This property is usually denoted by and is defined as the change in the density of a
substance as a function of temperature at constant pressure. It can be approximated as:
1
T

= T

(16)
(17)

For ideal gases P = RT, so


1
(18)
T
In other words, to find the change in density as a function of a change in temperature, we
just multiply the density by T.
=

5.1.6

Thermal Diffusivity

When a temperature gradient is applied to a material, the heat travels from the high
temperature region to the low temperature. A measure of how heat propagates through
a medium may be defined by the ratio of the heat conducted through the material to the
heat stored in the material. Heat capacity is defined as the product of density and specific
heat,Cp . The thermal diffusivity is defined as:
=

k
Cp

(19)

The thermal diffusivity is, therefore, the ratio of heat conducted through the material to the
heat stored per unit volume. The larger the thermal diffusivity the faster the propagation
of heat into the material. If the thermal diffusivity is small it means that a big part of the
heat is absorbed by the material and only a small portion is conducted through.Unit of
is m2 /s. Some typical value of thermal diffusivity:
Material
Liquid Metals
Gases
Water
Oils

5.2

Pr
0.004-0.03
0.7-1.0
1.7-13.7
50-100,000

Reynolds Number

Reynolds number defined as Re VL (where L is a characteristic length) may be interpreted as the ratio of two forces that influence the behavior of fluid flow in the boundary
layer. These two forces are the inertia forces and viscous forces:
Re =

InertiaF orces
V iscousF orces
22

5.3 Nusselt Number

5 DIMENSIONLESS NUMBERS IN HEAT TRANSFER

V 2
L
V
V iscousF orces 2
L
When the Reynolds number is large, the inertia forces are in command. Viscous forces
dominate the boundary layer when the Reynolds number is small. Now, how does this
relate to transition from laminar to turbulent flow?
Any real flow of fluid contains small disturbances that will grow given enough opportunities. as long as the viscous forces dominate these disturbances are under control. As
the inertia forces get bigger, the viscosity can no longer maintain order and these tiny
disturbances grow into trouble makers and we transition to turbulent flow.
Another important quantity of the boundary layer that is influenced by the Reynolds
number is its thickness. As the Reynolds number increases, the viscous layer gets squeezed
into a smaller distance from the surface.
The value of Reynolds number beyond which the flow is no longer considered laminar
is called the critical Reynolds number. For flow over a flat plate, the critical Reynolds
number is observed to vary between 1e+5 to 3e+6 depending on the turbulence level in
the free stream and the roughness of the surface. We normally use 5e+5 as the critical
Reynolds number for flow over flat plates.
Calculation of the Reynolds number is easy as long as you:
InertiaF orces

Identify the characteristic length


Pick the right velocity
Use a consistent set of units
For flow over a flat plate, the characteristic length is the length of the plate and the characteristic velocity is the free stream velocity. For pipes the characteristic length is the pipe
diameter and the characteristic velocity is the average velocity through the pipe obtained
by dividing the volumetric flow rate by the cross-sectional area (we are assuming that the
pipe is full, of course). For pipes with a non-circular cross-section, the characteristic length
is the Hydraulic Diameter defined as 4A/P , where A is the cross-sectional area of the duct
and P is the wetted perimeter. You can easily verify that for a circular pipe the hydraulic
diameter equals the pipe diameter. For non-cicular pipes the average velocity is used as
the characteristic velocity. The situation gets messy when you are dealing with a problem
that has many velocity and length scales such as the flow inside a computer cabinet. You
must decide, based on your design objectives, which length and velocity length scales make
sense for calculation of the Reynolds number.

5.3

Nusselt Number

Nusselt number is the dimensionless heat transfer coefficient and appears when you are
dealing with convection. It, therefore, provides a measure of the convection heat transfer
23

5.4 Prandtl Number

5 DIMENSIONLESS NUMBERS IN HEAT TRANSFER

at the surface. It is defined as hL/k where, h is the heat transfer coefficient, L is a


characteristic length and k is the thermal conductivity. But, what does this grouping
mean from a physical standpoint? Lets find out.
We have to look at the boundary layer in order to explain the concept of Nusselt
number. We will, of course, cover the basics of the boundary layer in a separate tutorial
but for now it suffices to say that when a fluid flows over a solid surface, the first layer of the
fluid stick to the boundary (we even have a name for this thing called, no slip condition).
This causes the flow to retard in the vicinity of the wall. As we move away from the wall
the effect of this no slip thing gets smaller and smaller up to a point where it is no longer
felt by the fluid. To get to this point, though, we have had to go through a layer of fluid
who still knows about the wall. This layer is called the boundary layer. This was the
effect of the wall on the velocity (or momentum). A similar argument applies when, for
example, a cold fluid flows over a hot surface. The first layer of the fluid (which is now
stuck to the surface) gets its heat from the surface through pure conduction. It then gives
its newly acquired energy to all of the other fluid molecules that it comes in contact with
as they pass by it (this is convection). As we move further and further away from the
wall, the effect of the hot wall is felt less and less (it, of course, depends on the thermal
conductivity of the fluid). Eventually, there comes a point where the fluid does not have
a clue about the hot wall. The layer of fluid between the wall and this point is called the
thermal boundary layer. It is where all of the action is taking place (as far as heat transfer
between the solid and fluid is concerned). Before continuing with the Nusselt number, let
us define another dimensionless property:

5.4

Prandtl Number

Heat transfer gurus have invented another dimensionless number called the Prandtl number
which is a grouping of the properties of the fluid but it has a significance to our discussion.
Prandtl number is defined as:
Cp

(20)
Pr = =

k
It is the ratio of momentum diffusivity (kinematic viscosity) to thermal diffusivity. It can
be related to the thickness of the thermal and velocity boundary layers. It is actually the
ratio of velocity boundary layer to thermal boundary layer. When P r = 1, the boundary
layers coincide.
Typical values of the Prandtl number are:
Material
Liquid Metals
Gases
Water
Oils

Pr
0.004-0.03
0.7-1.0
1.7-13.7
50-100,000

When Pr is small, it means that heat diffuses very quickly compared to the velocity (momentum). This means the thickness of the thermal boundary layer is much bigger than
24

5.5 Grashof Number

5 DIMENSIONLESS NUMBERS IN HEAT TRANSFER

the velocity boundary layer for liquid metals.


Now, back to the Nusselt number.
Remember that we explained how the first fluid layer stick to the solid surface and the
heat is transferred via conduction. Well, lets write this equation:
 
dT
(21)
qcond = kf luid
dy y=0
This heat (flux) is what will be taken away by convection by the moving fluid: qconv =
h(Ts T ) which is a simple definition of the heat transfer coefficient, h.
qcond = qconv
kf luid

dT
dy

(22)

i
y=0

(23)
(Ts T )
Note the dimensions of h from the above equation as k/L since the temparatures cancel
out. So, why not create a dimensionless heat transfer coefficient dividing both sides by
k/L. The Nusselt number is thus born:
h=

hL
(24)
k
In a boundary layer situation the characteristic length is the thickness of the boundary
layer. Consider a fluid layer of thickness L and a temperature difference of T across this
layer. Heat transfer by convection can be calculated as h while heat transfer by conduction
is k/L. Dividing the convection heat transfer to the conduction heat transfer, we get:
Nu =

hT
qconv
=
= Nu
qcond
kT /L

(25)

So, the Nusselt number may be viewed as the ratio of convection to conduction for a layer
of fluid. If N u = 1, we have pure conduction. Higher values of Nusselt mean that the heat
transfer is enhanced by convection.

5.5

Grashof Number

You see this number and you should think of natural or free convection. The Grashof
number is the ratio of buoyancy forces to the viscous forces.
gT 3
(26)
2
where is a relevant characteristic length. In natural convection the Grashof number plays
the same role the is played by the Reynolds number in forced convection. The buoyant
forces are fighting with viscous forces and at some point they overcome the viscous forces
and the flow is no longer nice and laminar. For a vertical plate, the flow transitions to
turbulent around a Grashof number of 109 .
Gr =

25

5.6 Rayleigh Number

5.6

5 DIMENSIONLESS NUMBERS IN HEAT TRANSFER

Rayleigh Number

The Rayleigh number is the product of Grashof and Prandtl numbers. It turns out that
in natural convection the Nusselt number scales with Rayleigh rather than just Grashof.
Most correlations in natural convection are of the form:

where
Ra =

N u = C(Ra)n

(27)

g(Ts Tinf ty ) 3
Pr
2

(28)

26

You might also like