Sie sind auf Seite 1von 7

Lecture 2.20.

One Dimensional Flows 89

2.5 Simple Viscous Flows


Lecture 2.20 In one dimensional flows� the inertial term is absent and the problems are usually
amenable to analytical solutions.
In this lecture we will consider simple flow application of Navier-Stokes equation, on which it is pos-
sible to obtain a closed form analytical solution to the velocity fields. Many problems are discussed
in the introductory texts Wilkes (1999); Fox and McDonald (2001); Denn (1980), we will consider only
a few illustrative applications which have more of chemical engineering relevance.

A general procedure to solve the Navier-Stokes equation The Navier-Stokes Equations Equa-
tion (2.29) is a set of four coupled non-linear second order partial differential equations. Even if
the boundary conditions are known, obtaining a solution is not a straight forward mathematical
problem. For most engineering applications, at least if the problem can be simplified, it is possible
to reduce the complexity in the equations at the very beginning. This is achieved by making some
assumptions, and idealisations. Good approximations come with experience and practice. But there
is a general sequence of steps that may be followed to simplify the problem.

1. Guess which is the dominant velocity direction, if there could be one.

2. Guess the independent variables that this depends on.

3. Validate the guess by substituting it in the continuity equation. Remember there are no physical
parameters appearing in the continuity equation which could alter the balance. It could be
possible to obtain additional information about the velocity field. This approach was used in
the differential analysis of the boundary layer.

4. Substitute the guesses in the momentum equation and simplify the differentials. For many
problems, the concentrating on pressure terms usually leads to further simplifications.

We will illustrate this procedure in a few problems.

2.5.1 Poiseuille and Couette Flows


Poiseuille flow is a pressure driven motion of a fluid. Couette flows are those in which the pressure
gradients are absent, and the flow is driven by motion of the boundary; it is also called as a drag flow.
You have studied Poiseuille flows in a pipe and steady and unsteady Couette flows, in your earlier
course Transport Phenomena. Here we will consider pressure driven flow in a rectangular channel.
Consider a rectangular channel of length L (along the x direction), and cross-section of W × H
(along z and y directions respectively), through which there is a flow driven by a pressure gradient
acting in the x-direction. We will make certain assumptions and simplify the problem. If we obtain a
solution and it is consistent with the assumptions, then we take that we have found out one solution
to the non-linear differential equations. But we must admit there there could be other possibilities as
well, which can be verified only by experiments, or the complete solution to the equation.
Assumption-1: Steady flow. This immediately drops all the time derivatives, and we have al-
ready restricted our solution space. Assumption-2: The aspect ratio of the channel is very large. This
implies W � H. This seems a restrictive assumption, but again it is only an assumption. We can
take this to an extreme, which we call idealisation, and take W/H → ∞. This implies the wall does
not influence the velocity, or that there is “no wall” in the z direction. This means the velocities do
not change in the z-direction. Assumption-3 The flow is fully-developed. This means that regions of
influence of the entrance and exit are negligible in comparison with the length L. This is also to say
that the velocities are not a function of x. This reduces the problem statement in the velocities to the
following
ui = fi �y) (2.131)

c Sunthar� February 15� 2011


�P
90 Chapter 2. Differential Analysis of Fluid Flow

Substituting in the continuity equation we get uy = v = constant. From the no-slip condition, we
have that the velocity has to vanish at the boundaries. This implies v = 0 through out. The NSE in
Cartesian coordinates reduce to
∂p d2 u
0=− �µ 2 (2.132)
∂x dy
∂p
0=− (2.133)
∂y
∂p d2 w
0=− �µ 2 (2.134)
∂z dy
Note we have used complete derivative for the velocities, and partial for the pressure. Inspecting the
pressure terms, we find that the pressure is only a function of �x� z). In Equation (2.132), ∂p
∂x is not a
function of y where as the viscous term is, and the viscous term is not a function of x whereas the
pressure term is. This implies each of them must be equal to the same constant. A similar argument
applies for the z-momentum equation. This gives us the solutions (along with the no-slip conditions)
� � �2 
H 2 Δp 

2y 
u= − 1 −  (2.135)
8µ L  H 
 � �2 
H 2 Δp 
� �
2y 
w= − 1 −  (2.136)
8µ W H 

The average velocity in the z-direction is


�H/2
H2
� �
1 Δp
�w� ≡ dy w = − (2.137)
H 12 µ W
−H/2

Since there is no net-flow in the z-direction because of the presence of the walls, this implies the
pressure gradient is zero, or from Equation (2.136), w = 0. Equation (2.135), v = w = 0 is the solution
we obtain. This is consistent with our assumptions.

2.5.2 Torsional flow


Viscometers are used to determine the viscosity of fluids. They are usually employed to measure
viscosities of liquids. The commercial viscometers can typically measure only high viscosity liquids,
about 100 times that of water. A cylinder is rotated in a liquid and the torque required to generate
a constant rpm is measured. This is used to compute the viscosity of the fluid. Rheometers are
more sophisticated and have many more controls, and can measure many more properties of flow
for complex fluids that exhibit both viscous and elastic behaviours.
In reality the velocities around the shaft can be complex, but we make simplifications as before.
We assume that the length of the shaft L � R, the radius of the shaft, thereby neglecting the end
effects between the bottom of the vessel and the shaft. Further we assume that there is no axial or
radial motion, but motion only in the θ-direction. From the continuity equation we have that vθ = f �r)
(we have neglected the variation in the z direction due to the length assumption above). The NSE
reduce to
vθ ∂p
−ρ =− (2.138)
r ∂r � �
d 1 d�r vθ )
0=µ (2.139)
dr r dr
∂p
0=− (2.140)
∂z

c Sunthar� February 15� 2011


�P
Lecture 2.20. One Dimensional Flows 91

From Equation (2.139) we get


C2
vθ = C1 r � (2.141)
r
At the outer radius of the vessel the velocity is zero, so for any large r the velocity is finite, therefore
the term proportional to r term must vanish from the equation. No-slip at the shaft surface (vθ = R ω
at r = R,) gives
ω R2
vθ = (2.142)
r
The torque is proportional to the surface force exerted by the fluid on the shaft. We require to estimate
τrθ , which can be obtained from the tables for the velocity gradient in cylindrical coordinates

τ = µ�∇ v � ∇ vT )
(2.143)
� �
∂vθ 1 ∂vr vθ
τrθ = µ � −
∂r r ∂θ r
The stress at r = R is therefore
τrθ �r = R) = −2 µ ω (2.144)
The differential force on an element swept by dθ across the length is τrθ L R dθ and the corresponding
torque is τrθ L R2 dθ. Integrating over the angle, we get an expression for the total torque

G = −4 π L R2 µ ω

2.5.3 Free surface flows


So far we have considered a flow where there is only a single fluid. In the presence of more than
one fluid, proper interface condition for the force needs to be specified. The simplest condition when
the interface is stationary or moving with a constant velocity. In this case the forces (surface forces)
acting on either side of the interface must balance each other.

σ1 = σ2 or σ�1) �2)
i j = σi j (2.145)

We can also equate the normal and tangential component of the forces on either side. For the simplest
case of a stationary fluids with a curved interface this reduces to the normal stress balance:

p�1) = p�2) � (2.146)
R
at the interface, where γ is the surface tension and R is the radius of curvature. When two fluids flow
past each other, with a flat interface, the tangential force arises due to the viscous stress. A stress
balance gives
∂u1 ∂u2
τ�1) �2)
yx = τyx or µ1 = µ2 (2.147)
∂y ∂y
In the case of air-water interface, we rewrite this as
∂uw µa ∂ua
= (2.148)
∂y µw ∂y
In most cases the air region is unbounded (like in the case of a jet of liquid in air) therefore the velocity
gradient is also negligibly small. Even if the region of air is bounded by walls present in comparable
length scales, the viscosity of water is about 50 times that of air, making the RHS negligibly small for
the velocity gradient in water. This leads to a general condition that there is no shear stress acting
on the surface, or it is a stress-free boundary; the condition is also known as an inviscid boundary
condition. This is a very useful simplification in treating flows of air water interface. The stress free
condition leads to the conclusion that in the cases of jets the velocity profile in the liquid is flat.

c Sunthar� February 15� 2011


�P
92 Chapter 2. Differential Analysis of Fluid Flow

Die Wire Coating


Radius: Rd Rw Rc
Figure 2.16: Schematic diagram of wire coating.

2.5.4 Wire Coating


Wire coating is the process of coating a polymeric material on, usually, metallic wires. The process
is done in the liquid state of the polymer, and then it is let to dry. A wire is pulled through a die
connected to a reservoir of molten polymer, leaving a coat of constant thickness. The problem is
to determine the coating thickness as a function of rate of pulling and the other parameters of the
system. Also of importance is the force required to pull the wire, at a given velocity.
Consider the schematic shown in Figure 2.16. Inside the die the liquid is flowing in an annular
space created by the void between the wire and the die. At distances far away from the die exit, the
coating is uniform, implying the interface is flat and the free surface condition (stress-free) yields a
flat velocity profile. Performing a integral mass balance across a large control volume, that takes the
shape of the liquid we get at steady state

�Rd
− dr 2 π r ρ vz � Vw π �R2c − R2w ) = 0 (2.149)
Rw

where Vw is the velocity the wire is pulled with. In order to find Rc we need to know the velocity
profile vz �r) inside the die annulus. As before, we consider a fully developed velocity profile, as this
makes the equations simpler (Note that when we positioned the CV boundary in the die, we did
not require it to have a fully developed velocity profile, i.e., the Equation (2.149) is valid even if the
velocity is not fully developed, but the expression is easier to evaluate for fully developed flow). The
NSE reduce to p = constant and � � ��
1 ∂ ∂vz
0=µ r (2.150)
r ∂r ∂r
With the no-slip boundary conditions this gives
ln r/Rd
vz �r) = Vw (2.151)
ln Rw /Rd
Substituting in Equation (2.149) we get
� 2 1/2
 R − R2w 
Rc =  d  (2.152)
2 ln Rd /Rw

and the force required to pull the die is obtained by integrating the surface force on the wire over the
length of the die
2 π µ Vw L
Fw = (2.153)
ln Rw /Rd
assuming the same the fully developed velocity profile to be dominating the entire length. This also
ignores any radial variations in the velocity from the die exit to the regime of constant wire thickness.

c Sunthar� February 15� 2011


�P
Lecture 2.20. One Dimensional Flows 93

2.5.5 Tube flow of Power law fluid


Most of the fluid we have considered are Newtonian, that is the shear stress is related by a linear
relationship to the local velocity gradient as given in Equation (2.30). However, there are several
fluids which do not exhibit this behaviour. This is especially true of polymeric fluids, and other
fluids that have a constituent substance that cannot be considered to be simple spherical molecules
of same size. These fluids exhibit behaviours that are vastly different and rich in variety. Any fluid
that does not show the Newtonian behaviour, in the velocity gradients of practical interest, is called a
non-Newtonian fluid (Strictly speaking, there could be very high gradients of velocity where even the
so called Newtonian fluids, such as air and water, exhibit non-linear behaviour; but these are very
rare for practical considerations.)
The local velocity gradient (also called as the local shear rate) can be vastly different for various
processes. Below is a compilation of some of the processes.
Phenomenon Shear rate (s−1 ) Application
−6
Sedimentation of powders 10 Medicines, paints
Extruders 10 Polymers
Mixing 102
Rubbing 103 Application of creams
High speed Coating 105 Printing inks
Lubrication 106 Engines
For non-Newtonian fluids, the viscosity is not a constant. It could be dependent on the local
velocity gradient (or shear rate). It could be time-varying as well. It is convenient to define a viscosity
function
τ
η≡ (2.154)
γ̇
where τ is the local shear stress, in a simple Couette geometry, and γ̇ is the local shear rate (or velocity
gradient). Some of the common non-Newtonian behaviours are
� Shear-thinning materials. The shear stress is no longer linear with the shear rate, but increases
sub-linearly. Or in other words, the viscosity (which is the local slope in the stress vs rate curve)
decreases as the shear rate is increased. Most polymeric substances exhibit this behaviour. The
behaviour is also called as pseudo-plastic.
� Thickening. Slurries and suspensions of particles in a fluid (paint for example) show a increase
in the viscosity with shear rate, and is called as shear-thickening or dialatant materials.
� Yeild Stress materials: Some materials deform elastically upto a certain stress and begin to
flow only beyond a stress known as yeild stress. These materials may exhibit a Newtonian or
non-Newtonian behaviour beyond the yield point. They are also known as Bingham Plastic
� The three types discussed here have a steady viscosity function, so long as the applied shear
rate is kept constant. However some materials show a time-dependent (or history dependent)
viscosity. Thixotropic fluids show a viscosity thinning with time at a constant shear rate, Rheopec-
tic fluids show a viscosity thickening with time.
With that brief introduction to non-Newtonian behaviour, we take up a simple problem of a
Poiseuille flow of a power-law fluid. “Power law” fluids are a class of shear thinning materials that
show a power-law dependence of the viscosity function with the shear rate.

η = K �γ̇)n−1 (2.155)

where K is a empirical constant called as the consistency factor. For a Couette flow the shear stress is
given by
τ xy = η γ̇ = K γ̇n (2.156)

c Sunthar� February 15� 2011


�P
94 Chapter 2. Differential Analysis of Fluid Flow

A more general expression for the shear rate tensor is given by


� �
τ = η�II) ∇v � ∇vT (2.157)
�n−1)/2
η�II) = K |II| (2.158)
� �
where II is called as the second invariant of the velocity gradient tensor defined as, II = 12 �tr2�)2 − tr�2�)2 ,
� = 12 �∇v � ∇vT ) is the rate of deformation tensor, and tr denotes the trace of a matrix.
As with the Poiseuille flow we will assume the following (fully-developed, axis-symmetric, and
one-dimensional)
vz = vz �r)� vr = vθ = 0 (2.159)

This satisfies the continuity equation. To solve for the velocities, we have to go back to the general
momentum balance, not that for Newtonian fluids. We look up the cylindrical coordinate equations

∂p
0=−
∂r
∂p
0=−
∂θ
∂p 1 d
0=− � �rτrz )
∂z r dr

Solving for τrz gives

ΔP
τrz = r (2.160)
2L

Note that even for non-Newtonian fluids the stress is still a linear function of the distance from the
walls. It is only the velocity profiles that will be different. Inserting the expression for the shear stress
�� ��n−1
dvz dvz
τrz = K ��� ��� (2.161)
dr dr
and solving for the velocity and using the usual boundary conditions we get
� n�1 � ��1/n � � r ��n�1)/n �
n R ΔP
vz = − 1− (2.162)
n � 1 2K 2K L R

In terms of the mean velocity �vz � we get, after simplifications,


� � r ��n�1)/n �
3n � 1
vz = �vz � 1 − (2.163)
n�1 R

For shear thinning fluids, n < 1 (n is usually in the range 0.3 < n < 1), the profile is flatter towards the
centre in comparison with a parabolic velocity profile for Newtonian fluids.

Physical interpretation of the velocity profile The stress distribution is identical in the cases of
Newtonian and non-Newtonian fluids. For a given pressure drop, the non-Newtonian fluid experi-
ences an lower viscosity, therefore we can expect a higher flow rate, implying the average velocity
�vz � would be higher. This can be seen from Figure 2.17, where the area under the curve for the
non-Newtonian fluid is higher than that for the Newtonian fluid.

c Sunthar� February 15� 2011


�P
Lecture 2.20. One Dimensional Flows 95

0.8

0.6
ylabel

0.4

0.2

n = 1, Newtonian
n = 0.3
0
0 0.2 0.4 0.6 0.8 1
xlabel

Figure 2.17: Fully developed velocity profiles for a power law fluid with n = 0.3 compared to that of
a Newtonian fluid.

2.6 Creeping Flows


2.6.1 Flow past a sphere
Concepts You Must Know
1. Simplifications of NSE for low Reynolds Number flows
2. Creeping flow equations are linear differential equations: Unique solution exists
3. Uniform flow past a sphere is easily solved in spherical coordinates
4. The net drag force on the sphere is due to “form friction” (normal stresses) and “skin friction”
(tangential stresses) on its surface.
5. The total drag may also be obtained equating the energy supplied to move the sphere to the
total viscous dissipation in the fluid.

c Sunthar� February 15� 2011


�P

Das könnte Ihnen auch gefallen