You are on page 1of 62

Motion of Particles in Fluid

3.2. FLOW PAST A CYLINDER


AND A SPHERE
The flow of fluid relative to an infinitely long
cylinder, a spherical particle and a non-spherical
particle is considered, followed by a discussion of
the motion of particles in both gravitational and
centrifugal fields.
For a non-viscous fluid flowing past a cylinder,
as shown in Figure 3.1, the velocity and direction
of flow varies around the circumference. Thus at
A and D the fluid is brought to rest and at B and C
the velocity is at a maximum.
Since the fluid is non-viscous, there is no drag,
and an velocity gradient = exists at the surface
of the cylinder.

If the fluid is incompressible and the cylinder is


small, the sum of the kinetic energy and the
pressure energy is constant at all points on the
surface according to Bernoulli equation.

The kinetic energy is a maximum at B and C


and zero at A and D, so that the pressure falls
from A to B and from A to C and rises again
from B to D and from C to D; the pressure at A
and D being the same.
No net force is therefore exerted by the fluid on
the cylinder.
It is found that, although the predicted pressure
variation for a non-viscous fluid = results
obtained with a viscous fluid over the front
face, very considerable differences occur at the
rear face.

When a viscous fluid flows over a surface,


the fluid is retarded in the boundary layer which
is formed near the surface and that the
boundary layer increases in thickness with
increase in distance from the leading edge.
If the pressure in the direction of flow, the
retardation of the fluid and the boundary layer
is thinner in consequence.
If the pressure , however, the retardation
and the thickness of the boundary layer
increases more rapidly.

The force acting on the fluid at some point in


the boundary layer may then be sufficient to
cause flow in the reverse direction with the
result that an eddy current is set up.
A region of reverse flow then exists near the
surface where the boundary layer has
separated as shown in Figure 3.2.
At any position after separation point, the
velocity rises from zero at the surface to a
maximum negative value and falls again to
zero.

dp/dx < 0
dU/dx > 0

dp/dx = 0
dU/dx = 0

dp/dx > 0
dU/dx < 0

As if all shapes of
curves were
pushed upward

dp/dx > 0
dU/dx < 0

It then increases in the positive direction until it


reaches the main stream velocity at the edge of
the boundary layer, as shown in Figure 3.2.
At PQ the velocity in the X-direction is zero and
the direction of flow in the eddies is clockwise
For the flow of a viscous fluid past the cylinder,
the pressure decreases from A to B and also
from A to C so that the boundary layer is thin
and the flow is similar to that obtained with a
non-viscous fluid.

From B to D and from C to D the pressure is


rising and therefore the boundary layer rapidly
thickens with the result that it tends to separate
from the surface
If separation occurs, eddies are formed in the
wake of the cylinder and energy is thereby
dissipated and an additional force, known as
form drag, is set up.
In this way, on the forward surface of the
cylinder, the pressure distribution is similar to
that obtained with the fluid of zero viscosity.

On the other hands, on the rear surface, the


boundary layer is thickening rapidly and pressure
variations are very different in the 2 cases
Therefore, there are two force components: the
skin friction (or viscous drag) and the form drag
(due to the pressure distribution).
At low velocity, no separation of the boundary
layer takes place, although as the velocity is
increased, separation occurs and the skin friction
gradually decreases

If the velocity of the fluid is very high, however,


or if turbulence is artificially induced, the flow
within the boundary layer will change from
streamline to turbulent before separation takes
place.
Since the rate of transfer of momentum through
a fluid in turbulent motion >> that in a fluid
flowing under streamline conditions, separation
is less likely to occur, because the fast-moving
fluid outside the boundary layer is able to keep
the fluid within the boundary layer moving in the
forward direction.

If separation does occur, this takes place


nearer to D in Figure 3.1, the resulting eddies
are smaller.
Turbulence may arise either from an increased
fluid velocity or from artificial roughening of the
forward face of the immersed body.
Prandtl roughened the forward face of a sphere
in which sand particles have been stuck to the
front face, as shown in Figure 3.3, with the
result that the drag was considerably reduced.

positive and
negative relative
to Pinfinity (see next
slide)

For the case of creeping flow, that is flow at very


low velocities relative to the sphere (Re < 0.2),
the drag force F on the particle was obtained by
Stokes who solved the hydrodynamic equations
of motion, the NavierStokes equations, to give:
.
Equation 3.1, which is known as Stokes law is
applicable only at very low values of the particle
Reynolds number and deviations become
progressively greater as Re' increases.

Conditions of flow relative to a spherical


particle are similar to those relative to a
cylinder, except that the flow pattern is 3D.
The flow is characterised by the Reynolds
number Re'(= ud/) in which is the density
of the fluid, is the viscosity of the fluid, d is
the diameter of the sphere, and u is the
velocity of the fluid relative to the particle.

3.3. THE DRAG FORCE ON A


SPHERICAL PARTICLE
3.3.1. Drag coefficients
The most satisfactory way of representing the
relation between drag force and velocity
involves the use of two dimensionless groups.
The first group is the particle Reynolds number
Re' (= ud/).
The second is the group R'/u2, in which R' is
the force per unit projected area of particle in a
plane perpendicular to the direction of motion.

For a sphere, the projected area is that of a


circle of the same diameter as the sphere.

1. R'/u2 is a form of drag coefficient, often


denoted by the symbol CD'. Frequently, a drag
coefficient CD is defined as the ratio of R' to
1/2 u2.

When the force F is given by Stokes law


(equation 3.1), then:

Equations 3.1 and 3.5 are applicable only at


very low values of the Reynolds number Re'
(Re < 0.2) (region a in Figure 3.4)
Goldstein(2) has shown that, for values of Re'
up to about 2, the relation between R'/u2 and
Re' is given by an infinite series of which
equation 3.5 is just the first term.

Laminar BL and
Laminar BL and
slight separation
no separation

Laminar BL and
large separation

turbulent BL
and reduced
separation area

The correction factors for Stokes law from both


equation 3.6 and equation 3.7 are given in
Table 3.1. It is seen that the correction
becomes progressively greater as Re'
increases.

The relation between R'/u2 and Re' is


conveniently given in graphical form by means
of a logarithmic plot as shown in Figure 3.4.
The graph may be divided into four regions as
shown.
Region (a) (104 < Re' < 0.2)
In this region, the relationship between R'/u2
and Re' is a straight line of slope 1
represented by equation 3.5:
.

Region (b) (0.2 < Re' < 5001000)


In this region, the slope of the curve changes
progressively from 1 to 0 as Re' increases.
Several workers have suggested approximate
equations for flow in this intermediate region.
Dallavelle(6) proposed that R'/u2 may be
regarded as being composed of two component
parts, one due to Stokes law and the other, a
constant, due to additional non-viscous effects.

Schiller and Naumann(7) gave the following


simple equation which gives a reasonable
approximation for values of Re' up to about
1000:
.
Region (c) (5001000 < Re' < 2 105)
In this region, Newtons law is applicable and
the value of R'/u2 is approximately constant:
.

Region (d) (Re' > 2 105)


When Re' exceeds about 2 105, the flow in
the boundary layer changes from laminar to
turbulent and the separation takes place nearer
to the rear of the sphere.
The drag force is decreased considerably and:

A comprehensive review of the various


equations proposed to relate drag coefficient to
particle Reynolds number has been carried out
by Clift, Grace and Weber(8).
One of the earliest equations applicable over a
wide range of values of Re' is that due to
Wadell (9) which may be written as:
.

Subsequently, Khan and Richardson(10) have


examined the experimental data and suggest
that a very good correlation between R'/u2
and Re' , for values of Re' up to 105, is given
by:

3.3.2. Total force on a particle


The force on a spherical particle may be
expressed using eqs 3.5, 3.9, 3.10 and 3.11
for each of the regions a, b, c and d as follows.

The projected area of the particle is d2/4.


Thus the total force on the particle is given by:

This is the expression originally obtained by


Stokes(1) already given as equation 3.1.
In region (b), from equation 3.9:

This relation (in region c) is often known as


Newtons law.

Alternatively using equation 3.13, which is


applicable over the first three regions (a), (b)
and (c) gives:

3.3.3. Terminal falling velocities


If a spherical particle is allowed to settle in a
fluid under gravity, its velocity will increase until
the accelerating force is exactly balanced by the
resistance force.
Although this state is approached exponentially,
the effective acceleration period is generally of
short duration for very small particles.

If this terminal falling velocity is such that the


corresponding value of Re' < 0.2, the drag force
on the particle is given by equation 3.15.
If the corresponding value of 0.2 <Re' < 500,
the drag force is given approximately by
Schiller and Naumann in equation 3.17
Under terminal falling conditions, velocities
rarely correspond to Re' 105, with the small
particles generally used in industry.

The accelerating force due to gravity is given


by:
where s is the density of the solid.
The terminal falling velocity u0 corresponding to
region (a) is given by:

0: subscript for
terminal velocity

The terminal falling velocity corresponding to


region (c) is given by:

In the expressions given for the terminal falling


velocity, the following assumptions are held:
(a) That the settling is not affected by the
presence of other particles in the fluid. This
condition is known as free settling. When the
interference of other particles is appreciable, the
process is known as hindered settling.
(b) That the walls of the containing vessel do
not exert an appreciable retarding effect.

From equations 3.24 and 3.25, it is seen that


terminal falling velocity of a particle in a given
fluid becomes greater as both particle size
and density are increased.
If for a particle of material A of diameter dA and
density A, Stokes law is applicable, then the
terminal falling velocity u0A is given by equation
3.24 as:
.

Similarly, for a particle of material B:

The condition for u0A = u0B is then:

If Newtons law is applicable, equation 3.25


holds and:

For equal settling velocities:

In general, the relationship for equal settling


velocities is:

where S = 1/2 for the Stokes law region, S = 1


for Newtons law and, as an approximation, 1/2
< S < 1 for the intermediate region.

This method of calculating the terminal falling


velocity is satisfactory provided that it is known
which equation should be used for the
calculation of drag force or drag coefficient.
It has already been seen that the equations
give the drag coefficient in terms of the particle
Reynolds number Re'0 (= u0d/) which is itself
a function of the terminal falling velocity u0
which is to be determined.

The problem is most effectively solved by the


generation of a new dimensionless group
which is independent of the particle velocity.
The resistance force per unit projected area of
the particle under terminal falling conditions R'0
is given by (applicable for any Re'):

.
= drag force/(cross-sectional area) at
terminal velocity (using subscript 0)

Applicable for
terminal velocity
condition

=2/3 Ga

CD'0.Re'02 = 2/3 Ga

Using equations 3.5, 3.9 and 3.10 to express


R'/u2 in terms of Re' over the appropriate
range of Re', then:
Re0 < 0.2
0.2<Re0 <
1000
Re0 > 1000

(R'0 /u02)Re'02 = 2/3 Ga can be evaluated if the


properties of the fluid and the particle are
known.
Derived from
Stokes law

Derived from Schiller and


Naumann eq. law

Derived from
Newtons law

In Table 3.4, values of log Re' are given as a


function of log{(R'/u2)Re' 2} and the data taken
from tables given by Heywood (11), are
represented in graphical form in Figure 3.6.
In order to determine the terminal falling
velocity of a particle, (R'0/u02)Re'02 is
evaluated and the corresponding value of Re'0,
and hence of the terminal velocity, is found
either from Table 3.4 or from Figure 3.6.

To be
used

Example 3.1
What is the terminal velocity of a spherical
steel particle, 0.40 mm in diameter, settling in
an oil of density 820 kg/m3 and viscosity 10 mN
s/m2? The density of steel is 7870 kg/m 3.
Solution
For a sphere:

3.8. MOTION OF PARTICLES IN A


CENTRIFUGAL FIELD
In most practical cases where a particle is moving
in a fluid under the action of a centrifugal field,
gravitational effects << and may be neglected.
The equation of motion for the particles is similar
to that for motion in the gravitational field, except
that the gravitational acceleration g must be
replaced by the centrifugal acceleration r2,
where r is the radius of rotation and is the
angular velocity.

analogous to the particle


motion in gravitational field

g replaced by r2

For a spherical particle in a fluid, the equation


of motion for the Stokes law region is:
Particle acceleration

As the particle moves outwards, the


accelerating force increases and therefore it
never acquires an equilibrium velocity in the
fluid.
Centrifugal force

drag force

bouyancy force

It works normal the fluid


rotating flow and away
from the axis of rotating
fluid flow

The solution of equation 3.112 takes the form:

For Stokes flow

If the particle starts (t = 0) at a radius r1 and


at zero velocity (dr/dt) = 0, then by making
derivation to equation 3.114:

2 B.Cs. (r and
dr/dt at t=0, r=r1)
for 2 eqs to get
2 constants B1
and B2
3.114

Hence r/r1 may be directly calculated at any


value of t , although a numerical solution is
required to determine t for any particular value
of r/r1.

If the inertial term on the right-hand side of


equation 3.108 is neglected (there is force
balance), then:

Thus, the instantaneous velocity (dr/dt) is equal


to the terminal velocity u0 in the gravitational
field, increased by a factor of r2/g.

Equation 3.109 modifies to:

Thus the time taken for a particle to move to a


radius r from an initial radius r1 is given by:

For Stokes flow

For a suspension fed to a centrifuge, the time


taken for a particle initially situated in the liquid
surface (r1 = r0) to reach the wall of the bowl (r
= R) is given by:

analogous to the
particle motion in
gravitational field

You might also like