0% found this document useful (0 votes)
42 views59 pages

CH8791 - Unit 1 Notes of Lessons

Notes

Uploaded by

jenish
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
42 views59 pages

CH8791 - Unit 1 Notes of Lessons

Notes

Uploaded by

jenish
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 59

Unit 1: TRANSPORT PHENOMENA BY MOLECULAR MOTION

1. Vectors/Tensors

2. Newton’s law of viscosity

3. Newtonian & Non-Newtonian fluids

4. Rheological models

5. Temperature, pressure and composition dependence of viscosity

6. Kinetic theory of viscosity

7. Fourier’s law of heat conduction, Temperature, pressure and composition

dependence of thermal conductivity

8. Kinetic theory of thermal conductivity, Fick’s law of diffusion

9. Temperature, pressure and composition dependence of diffusivity,

Kinetic theory of diffusivity.

10.Revision
Introduction to Vectors and Tensors
Some useful references for learning about vectors and tensors are the books listed as references at
the end.

Some Basics

We encounter physical entities such as position, velocity, momentum, stress, temperature, heat
flux, concentration, and mass flux in transport problems - there is a need to describe them in
mathematical terms and manipulate the representations in various ways. This requires the tools of
tensor analysis.

Scalars

An entity such as temperature or concentration that has a magnitude (and some units that need not
concern us right now), but no sense of direction, is represented by a scalar.

Vectors

In contrast, consider the velocity of a particle or element of fluid; to describe it fully, we need to
specify both its magnitude ( in some suitable units) and its instantaneous spatial direction. Other
examples are momentum, heat flux, and mass flux. These quantities are described by vectors. In
books, vectors are printed in boldface. In ordinary writing, we may represent a vector in different
ways.
 
v, v, v or vi

Gibbs notation index notation

The last requires comment. Whereas we represent the vectorial quantity with a symbol, we often
know it only via its components in some basis set. Note that the vector as an entity has an invariant
identity independent of the basis set in which we choose to represent it.

In index notation, the subscript “i” is a free index - that is, it is allowed to take on any of the three
values 1, 2, 3, in 3-dimensional space. Thus, vi really stands for the ordered set ( v1 , v2 , v3 ) .

Basis Sets

The most common basis set in three-dimensional space is the orthogonal triad ( i , j, k )
corresponding to a rectangular Cartesian coordinate system. i stands for a unit vector in the x −
direction and j and k represent unit vectors in the y and z − directions respectively. Note that
this is not a unique basis set. The directions of i , j , k depend on our choice of the coordinate
directions.

4
There is no reason for the basis set to be composed of orthogonal vectors. The only requirement
is that the three vectors chosen do not lie in a plane. Orthogonal sets are the most convenient,
however.

We find the components of a vector in the directions of the base vectors by taking inner (dot)
products.
vx =
v ⋅ i , vy =
v ⋅ j , vz = v ⋅k

Then, v = vx i + v y j + vz k

You can verify the consistency of the above by taking inner products of both sides of the equation
with the base vectors and recognizing that the base vectors are orthogonal.

i ⋅ j = j⋅k = k ⋅ i ≡ 0

The order of the vectors in the inner product is unimportant.

a⋅b = b⋅a

Scalar and Vector Fields

In practice the temperature, velocity, and concentration in a fluid vary from point to point (and
often with time). Thus, we think of fields - temperature field, velocity field, etc.

In the case of a vector field such as the velocity in a fluid, we need to represent the velocity at
every point in space in the domain of interest. The advantage of the rectangular Cartesian basis
( )
set i , j , k is that it is invariant as we translate the triad to any point in space. That is, not only
are these base vectors of unit length, but they never change direction as we move from one point
to another, once we have chosen our x, y, and z directions.

Vector Operations

The entity v has an identity of its own. Its length and spatial direction are independent of the
basis set we choose. As the vectors in the basis set change, the components of v change according
to standard rules.

Vectors can be added; the results are new vectors. If we use component representation, we simply
add each component. Subtraction works in a similar manner.

Vectors also can be multiplied, but there is more than one way to do it. We define the dot and
cross products, also known as inner (or scalar) and vector products, respectively, as shown below.

a ⋅ b= ax bx + a y by + az bz is a scalar. We commonly use a numerical subscript for the components;

5
( )
in this case, the basis set is the orthogonal triad e (1), e ( 2) , e (3) . Let

a = a1 e (1) + a2 e ( 2) + a3 e (3)

Then,

a ⋅ b= a1b1 + a2b2 + a3b3

3
= ∑ ai bi
i =1
3
In the above, we usually omit ∑
i=1
. When an index is repeated, summation over that index is

implied.

a⋅b =ai bi This is called the Einstein summation convention

a ⋅ a= ai ai= a 2 or a
2

where a is the length of a and is invariant; “invariant” means that the entity does not change as
the basis set is altered.

a × b is the vector product. As implied by the name, it is a vector; it is normal to the plane
containing a and b . ( a , b , a × b ) form a right-handed system (this is an arbitrary convention ,
but we have to choose one or the other, so we choose “right”). The order is important, for,

a × b =− b × a

that is, b × a points opposite to a × b .

We can write

e (1) e ( 2) e ( 3)
a × b = a1 a2 a3
b1 b2 b3

There is a compact representation of a determinant that helps us write

ε ijk ai b j
a×b =

(Note that k is a free index. The actual symbol chosen for it is not important; what matters is
that the right side has one free index, making it a vector)
6
ε ijk is called the permutation symbol

ε ijk = 0 if any two of the indices are the same

= +1 if i, j , k form an even permutation of 1, 2, 3 [example: 1,2,3]

= − 1 if i, j , k form an odd permutation of 1, 2, 3 [example: 2, 1, 3]

We can assign a geometric interpretation to a ⋅ b and a × b . If the angle between the two vectors
a and b is θ , then

a⋅b =a b cos θ

and the length of a × b is a b sin θ . You may recognize a b sin θ as the area of the parallelogram
formed by a and b as two adjacent sides. Given this, it is straightforward to see that

a⋅b× c =ε ijk ai b j ck

is the volume of the parallelepiped with sides a , b , and c . This is called the triple scalar product.

Second Order Tensors

Note that we did not define vector division. The closest we come is in the definition of second-
order tensors!

Imagine

a
= Τ
b

Instead, we write

a = Τ⋅b

A tensor (unless explicitly stated otherwise we’ll only be talking about “second-order” and shall
therefore omit saying it every time) “operates” on a vector to yield another vector. It is very useful
to think of tensors as operators as you’ll see later.

Note the “dot” product above. Using ideas from vectors, we can see how the above equation may
be written in index notation.

7
ai = Τij b j

It is important to note that b ⋅ Τ would be bi Τij and would be different from Τ ⋅ b in general.

The two underbars in Τ now take on a clear significance; we are referring to a doubly subscripted
entity. We can think of a tensor as a sum of components in the same way as a vector. For this, we
use the following result.

e (i )⋅ Τ ⋅ e ( j ) = Τij Scalar
We’re not using index notation on the left side of the above equality.

Thus, to get Τ23 we would find e ( 2) ⋅ Τ ⋅ e (3) . We can then think of T as a sum.

Τ = Τ11 e (1) e (1) + Τ12 e (1) e ( 2) + Τ13 e (1) e (3)

+Τ21 e ( 2) e (1) + Τ22 e ( 2) e ( 2) + Τ23 e ( 2) e (3)

+Τ31 e (3) e (1) + Τ32 e(3) e ( 2) + Τ33 e (3) e (3)

What are the quantities e (1) e ( 2) and others like them? They are called dyads. They are a basis
set for representing tensors. Each is a tensor that only has one component in this basis set. Note
that e(i ) e ( j ) ≠ e ( j ) e (i ) . A dyadic product is a third way in which we can multiply two vectors.

You can see that tensors and matrices have a lot in common!

In fact, we commonly write the components of a tensor as the elements of a 3 x 3 matrix.

 Τ11 Τ12 Τ13 


 
 Τ21 Τ22 Τ23 
Τ Τ Τ 
 31 32 33 

Naturally, as we change our basis set, the components of a given tensor will change, but the entity
itself does not change. Of course, unlike vectors, we cannot visualize tensors – we only “know”
them by what they do to vectors that we “feed” them!

A good example of a tensor in fluid mechanics is the stress at a point. To completely specify the
stress vector, we not only need to specify the point, but also the orientation of the area element.
At a given point, we can orient the area in infinitely many directions, and for each orientation, the
stress vector would, in general be different.

8
Force ← has magnitude and direction
Stress =
Area ← has magnitude and direction

In fact, we can show that stress is indeed a tensor (for proof, see Aris, p. 101). So, we get

t = n ⋅Τ

n
t

The symbol n represents the unit normal (vector) to the area element, and t is the stress vector
acting on that element. The second-order tensor Τ completely describes the state of stress at a
point. By convention, t is the stress exerted by the fluid into which n points on the fluid adjoining
it.

Just as a vector has one invariant (its length), a tensor has three invariants. They are defined as
follows.
Let A or Aij be the tensor.

↓ abbreviation
=
I A trace= { A} Aii
{ A} tr=
1 2
=
II A I A − II A  where
2 

II A tr { A ⋅ A}
= Note: A ⋅ A is the tensor Aij Ajk

III A = Determinant of A = Det { A}

= ε ijk A1i A2 j A3k

= ε ijk Ail Aj 2 Ak 3

9
As the basis set is changed, the invariants do not change even though the components of the tensor
may change. For more details, consult Aris, p. 26, 27 or Slattery, p. 47, 48.

A symmetric tensor Aij is one for which Aij = Aji . Thus, there are only six independent
components. Stress is a symmetric tensor (except in unusual fluids). Symmetric tensors with real
elements are self-adjoint operators, a concept about which you can learn more in advanced work.

A skew-symmetric tensor Aij is one for which Aij = − Aji . You can see immediately that the
diagonal elements must be zero ( because Aii = − Aii ) . Skew-symmetric tensors have only three
independent components. Vorticity is an example of a skew symmetric tensor.

If we write a skew-symmetric tensor Aij in the form

0 a3 − a2 
 
 −a3 0 a1 
a − a1 0 
 2

We can see that there is a vector a that can be formed using the elements of Aij . The two are
related by the following result, which is useful in the context of the physical significance of
vorticity.

A⋅ x = x × a

Any second order tensor can be decomposed into the sum of a symmetric tensor and a skew-
symmetric tensor.

( Aij + Aji ) + ( Aij − Aji ) or in Gibbs notation, A = ( A + AT ) + ( A − AT )


1 1 1 1
Aij =
2
 2
 2 
 2 

Symmetric Tensor Skew-Symmetric Tensor Symmetric Tensor Skew-Symmetric Tensor

Here, AT is the transpose of the tensor A . AT has components that form a matrix whose columns
are the rows of the matrix of components of A .

There is a special tensor that leaves a vector undisturbed. It is called the identity or unit tensor I
.
1 0 0 
  I ⋅x = x
I = 0 1 0
  for any x
 0 0 1

In index notation, we write I as δ ij , the Kronecker delta.

10
Symmetric tensors have a very special property. Remember that we define a tensor as the
representation of some field at a point; at that point, there are three special directions, orthogonal
to each other, associated with a symmetric tensor. When the tensor operates on a vector in one
of these directions, it returns another vector pointing in the same (or exactly opposite) direction!
The new vector, however, will have a different length in general. This multiplication factor in the
length is called the principal value or eigenvalue of the tensor. If the eigenvalue is negative, the
output vector from the operation will point opposite to the input vector.

Because there are, in general, three directions that are special, there are usually three distinct
principal values, one associated with each direction. Even for tensors that are not symmetric, there
are three principal values; however these need not all be real. Sometimes, two are complex. Even
when the principal values are real, the directions associated with them need not be orthogonal if
the tensor is not symmetric.

The problem for the principal or eigenvalues of A is

A⋅ x= λ x ≡ λ I ⋅ x

Therefore,

 A − λ I  ⋅ x =0

From linear algebra, for non-trivial solutions of the above system to exist, we must have

det  A − λ I  =
0

The resulting third degree equation for the eigenvalues is

−λ 3 + I Aλ 2 − II Aλ + III A =
0

and has three roots λ1 , λ2, and λ3 . When these roots are each used, in turn, and we solve for x ,
we obtain an eigenvector that is known only to within an arbitrary multiplicative constant.
Commonly, the eigenvector is normalized so that it has unit length.

From the above, you can see that corresponding to a symmetric tensor, there is a special rectangular
Cartesian set of basis vectors of unit length. If we choose this as the basis set, the tensor will have
a simple diagonal form with the diagonal components being the eigenvalues.

If you’re wondering what happens when two eigenvalues are identical, it is easy to show that any
vector in the plane normal to the third eigenvector (corresponding to the third eigenvalue) is
acceptable as an eigenvector. In other words, on that plane, the tensor operating on a vector in any
direction will yield a vector in the same direction with a magnification factor corresponding to the
repeated eigenvalue.

11
If all three eigenvalues are identical, then any direction in space will be acceptable as the direction
of the eigenvectors. Such a tensor is called isotropic for this reason. I is an isotropic tensor with
eigenvalues equal to unity. Any scalar multiple of I also is isotropic.

Vector Calculus

If we consider a scalar field such as temperature, we find the rate of change with distance in some
direction, x , by calculating ∂Τ ∂x . How can we represent the rate of change in three-dimensional
space without specifying a particular direction? We do this via the gradient operator. The entity
∇Τ [we call it “grad T”] is a vector field. In index notation we write the gradient operator as
∂ / ∂xi where i is a free index, so that ∇Τ = ∂T / ∂xi .

At a given point in space, the vector ∇Τ points in the direction of greatest change of T. To obtain
the rate of change of T at that point in any specified direction, n , we simply “project” ∇Τ in that
direction.

∂Τ
= ∇Τ ⋅ n
∂n
unit vector

Surfaces in space on which a field has the same value everywhere are level surfaces. In the case
of temperature fields, these surfaces are called isotherms. Along such a surface, the temperature
cannot change. Therefore, the ∇Τ vector is everywhere normal to isothermal surfaces since it
must yield a value of zero when projected onto such surfaces.

It is straightforward to establish from the definition that


∂ ∂ ∂
∇ ≡ e (1) + e ( 2) + e ( 3)
∂x1 ∂x2 ∂x3

in a rectangular Cartesian coordinate system ( x1 , x2 , x3 ) .

Note that ∇ is an operator and not a vector. So, you should exercise care in manipulating it.

The ∇ operator is the generalization of a derivative. We can differentiate vector fields in more
than one way.

Divergence

∇ ⋅ v or div v is called the divergence of the vector field v . If the rectangular Cartesian
components of v are v1, v2 , v3 , then
∂v1 ∂v2 ∂v3 ∂vi
∇ ⋅=
v + + = in index notation
∂x1 ∂x2 ∂x3 ∂xi

12
As you can see, the result is a scalar field. It can be shown that the divergence of a vector field at
any point represents the outflow of the field.

Curl

∇× v or curl v is a vector field. As the name implies, it measures the “rotation” of the vector v
. Again, in ( x1, x2 , x3 ) coordinates,
e(1) e( 2) e ( 3)
∂ ∂ ∂
∇× v = ∂vk
∂x1 ∂x2 ∂x3 = ε ijk in index notation.
∂x j
v1 v2 v3

The term “curl” implies rotation, and the curl of a vector field is indeed related to rotation as will
be evident when we discuss fluid mechanics later.

There are two important theorems you should know. They are simply stated here without proof.

Divergence Theorem

If the volume V in space is bounded by the surface S ,

∫ ∇ ⋅ a dV
V
= ∫ dS ⋅ a
S

dS

13
The vector field a should be continuous and differentiable. The symbol dS represents a vector
surface element. If n is the unit normal to the surface,

dS = n dS

The entity dV is a volume element. In the theorem, the left side is a volume integral and the right
side is an integral over the surface that bounds the volume. Finally, a need not be a vector field,
but can be a tensor field of any order.

The divergence theorem, also known as Green’s transformation, is a very useful result that
permits us to convert volume integrals into surface integrals. By applying it to an infinitesimal
volume, you can visualize the physical significance of the divergence of a vector field at a point
as the outward “flow” of the field from that point.

Stokes Theorem

This permits the conversion of integrals over a surface to those around a bounding curve. Imagine
a surface that does not completely enclose a volume, but rather is open, such as a baseball cap. Let
S be the surface and C, the curve that bounds it.

C
t
S

If a vector field a is defined everywhere necessary, and is continuous and differentiable, Stokes
theorem states:
∫ ( ∇ × a ) ⋅ dS
S
= ∫ a ⋅ t ds
C

where dS is a vector area element on the surface S , ds is a scalar line element on the bounding

curve C, and t is a unit tangent vector on C.

14
The integral on the right side is known as the circulation of a around the closed curve C. The
field a appearing in the theorem can be replaced by a tensor field of any desired order.

S
C

By imagining the surface S to lie completely on the plane of the paper as shown, you can visualize
the physical significance of ∇ × a . If you make S shrink to an infinitesimal area, the area integral
on the left side becomes the product of the component of ∇ × a normal to the plane of the paper
and the area. The line integral is still the circulation around an infinitesimal closed loop
surrounding the point. If a is the velocity field v , by making the boundary an infinitesimal
circle of radius ε , the right side can be seen to be approximately 2πε v , where v is the magnitude
of the velocity around the loop. The left side is approximately πε 2 (∇ × v) ⋅ n where n is the unit
1 v
normal to the plane of the paper. Therefore, (∇ × v) ⋅ n ≈ , which becomes the instantaneous
2 ε
angular velocity of the fluid at the point on the plane of the paper as ε → 0 . Because there is
1
nothing unique about the choice of the plane of the paper, we can see that (∇ × v) in fact
2
represents the instantaneous angular velocity vector of a fluid element at a given point, the
component of which in any direction is obtained by projecting in that direction.

15
The Gradient of a Vector Field

Just as we defined the gradient of a scalar field, it is possible to define the gradient of a vector or
tensor field. If v is a vector field, ∇ v is a second-order tensor field. The rate of change of v in
any direction n is given by

∂v
=∇ v ⋅ n
∂n

References

1. R. Aris, Vectors, Tensors, and the Basic Equations of Fluid Mechanics, Dover Publications,
1989 (original by Prentice-Hall, 1962).

2. R. B. Bird, W. E. Stewart, and E. N. Lightfoot, Transport Phenomena, John Wiley, 2007.

3. J. Slattery, Momentum, Energy, and Mass Transfer in Continua, McGraw-Hill, 1972.

16
Momentum transport

Momentum transport deals with the transport of momentum which is responsible for flow in
fluids. Momentum transport describes the science of fluid flow also called fluid dynamics. A few
basic assumptions are involved in fluid flow and these are discussed below.

No slip boundary condition

This is the first basic assumption used in momentum transport. It deals with the fluid flowing
over a solid surface, and states that whenever a fluid comes in contact with any solid boundary,
the adjacent layer of the fluid in contact with the solid surface has the same velocity as the solid
surface. Hence, we assumed that there is no slip between the solid surface and the fluid or the
relative velocity is zero at the fluid–solid interface. For example, consider a fluid flowing inside
a stationary tube of radius R as shown in Fig 7.1. Since the wall of the tube at r=R is stationary,
according to the no-slip condition implies that the fluid velocity at r=R is also zero.

Fig 1.1 Fluid flow in a circular tube of radius R

In the second example as shown in Fig. 7.2, there are two plates which are separated by a
distance h, and some fluid is present between these plates. If the lower plate is forced to move
with a velocity V in x direction and the upper plate is held stationary, no-slip boundary
conditions may be written as follows

Fig 1.2 Two parallel plates at stationary condition


Thus, every layer of fluid is moving at a different velocity. This leads to shear forces which are
described in the next section.

Newton’s Law of Viscosity

Newton’s law of viscosity may be used for solving problem for Newtonian fluids. For many
fluids in chemical engineering the assumption of Newtonian fluid is reasonably acceptable. To
understand Newtonian fluid, let us consider a hypothetical experiment, in which there are two
infinitely large plates situated parallel to each other, separated by a distance h. A fluid is present
between these two plates and the contact area between the fluid and the plates is A.
A constant force F1 is now applied on the lower plate while the upper plate is held stationary.
After steady state has reached, the velocity achieved by the lower plate is measured as V1. The
force is then changed, and the new velocity of the plate associated with this force is measured.
The experiment is then repeated to take sufficiently large readings as shown in the following
table.
If the F/A is plotted against V/h, we may observe that they lie on a straight line passing through
the origin.

Fig 1.3 Shear stress vs. shear stain

Thus, it may be said that F/A is proportional to v/h for a Newtonian fluid.

It may be noted that it is the velocity gradient which leads to the development of shear forces.
The above equation may be re-written as

In the limiting case, as h → 0, we have

where, µ is a constant of proportionality, and is called as the viscosity of the fluid. The quantity

F/A represents the shear forces/stress. It may be represented as , where the subscript x
indicates the direction of force and subscript y indicates the direction of outward normal of the
surface on which this force is acting. The quantity or the velocity gradient is also called
the shear rate. µ is a property of the fluid and is measured the resistance offered by the fluid to
flow. Viscosity may be constant for many Newtonian fluids and may change only with
temperature.

Thus, the Newton’s law of viscosity, in its most basic form is given as

Here, both ‘+’ or ‘–’ sign are valid. The positive sign is used in many fluid mechanics books
whereas the negative sign may be found in transport phenomena books. If the positive sign is

used then may be called the shear force while if the negative sign is used may be
referred to as the momentum flux which flows from a higher value to a lower value.

The reason for having a negative sign for momentum flux in the transport phenomena is to have
similarities with Fourier's law of heat conduction in heat transport and Ficks law of diffusion in
mass transport. For example, in heat transport, heat flows from higher temperature to lower
temperature indicating that heat flux is positive when the temperature gradient is negative. Thus,

a minus sign is required in the Fourier's law of heat conduction. The interpretation of as
the momentum flux is that x directed momentum flows from higher value to lower value in y
direction.

The dimensions of viscosity are as follows:


The SI unit of viscosity is kg/m.s or Pa.s. In CGS unit is g/cm.s and is commonly known as poise
(P). where 1 P = 0.1 kg/m.s. The unit poise is also used with the prefix centi-, which refers to
one-hundredth of a poise, i.e. 1 cP = 0.01 P. The viscosity of air at 25oC is 0.018 cP, water at
25oC is 1 cP and for many polymer melts it may range from 1000 to 100,000 cP, thus showing a
long range of viscosity.

Laminar and turbulent flow

Fluid flow can broadly be categorized into two kinds: laminar and turbulent. In laminar flow, the
fluid layers do not inter-mix, and flow separately. This is the flow encountered when a tap is just
opened and water is allowed to flow very slowly. As the flow increases, it becomes much more
irregular and the different fluid layers start mixing with each other leading to turbulent flow.
Osborne Reynolds tried to distinguish between the two kinds of flow using an ingenious
experiment and known as the Reynolds’s experiment. The basic idea behind this experiment is
described below.

Reynolds’s experiment
Fig 1.4 Reynolds’s experiments

The experiment setup used for performing the Reynolds's experiment is shown in Fig. 7.5. The
average velocity of fluid flow through the pipe diameter can be varied. Also, there is an
arrangement to inject a colored dye at the center of the pipe. The profile of the dye is observed
along the length of the pipe for different velocities for different fluids. If this experiment is
performed, it may be seen that for certain cases the dye shows a regular thread type profile,
which is seen at low fluid velocity and flow is called laminar flow. when the fluid velocity is
increased the dye starts to mixed with the fluid and for larger velocities simply disappears. At
this point fluid flow becomes turbulent.
For the variables average velocity of fluid vz avg, pipe diameter D, fluid density ρ, and the fluid
viscosity µ, Reynolds found a dimensionless group which could be used to characterize the type
of fluid flow in the tube. This dimensionless quantity is known as the Reynolds number. From
the experiment, It was observed that if Re >2100, the dye simply disappeared and the flow has
changed to laminar to turbulent flow.

8
Thus, for Re <2100, we have laminar flow, i.e., no mixing in the radial direction leading to a
thread like flow and for Re >2100, we have the turbulent flow, i.e., mixing in the radial direction
between layers of fluid.

In laminar flow, the fluid flows as a stream line flow with no mixing between layers. In turbulent
flow, the fluid is unstable and mixes rapidly due to fluctuations and disturbances in the flow. The
disturbance might be present due to pumps, friction of the solid surface or any type of noise
present in the system. This makes solving fluid flow problem much more difficult. To understand
the difference in the velocity profile in two kinds of fluid flows, we consider a fluid flowing to a
horizontal tube in z direction under steady state condition. Then, we can intuitively see the
velocity profile may be shown below

For laminar flow, it is observed that fluid flows as smooth stream line and all other components
of velocity are zero. Thus

For turbulent flow, if we observe the fluid flows at a local point. It is observed that fluid flows in
very random manner in all directions where these local velocities may be the function of any
dimensions.

Thus, we see that for laminar flow there is only one component of velocity present and it
depends only on one coordinate whereas the solution of turbulent flow may be vary complex.
For turbulent flow, one can ask the question that if the fluid is flowing in the z direction then why
are the velocity components in r and θ direction non-zero? The mathematical answer for this
question can be deciphered from the equation of motion. The equation of motion is a non-linear
partial differential equation. This non-linear nature of the equation causes instability in the
system which produces flow in other directions. The instability in the system may occur due to
any disturbances or noise present in the environment. On the other hand, if the velocity of fluid is
very low the deviation due to disturbances may decay with time, and becomes negligible after
that. Thus the flow remains in laminar region. Consider a practical example in which some cars

9
are moving on the highway in the same direction but in the different lanes at different speeds. If
suddenly, some obstacle comes on the road, then if the car's speed is sufficiently low, it can move
on to other lane smoothly and come back to its original lane after the obstacle is crossed. This is
the regular laminar case. On the other hand, if the car is moving at a high speed and suddenly
encounters an obstacle, then the driver may lose control, and this car may move haphazardly and
hit other cars and after that traffic may never return to normal traffic conditions. This is the
turbulent case.

Internal and external flows

Depending on how the fluid and the solid boundaries contact each other, the flow may be
classified as internal flow or external flow. In internal flows, the fluid moves between solid
boundaries. As is the case when fluid flows in a pipe or a duct. In external flows, however, the
fluid is flowing over an external solid surface, the example may be sited is the flow of fluid over
a sphere as shown in Fig. 8.1.

Fig 1.5 External flow around a sphere

Boundary layers and fully developed regions

Let us now consider the example of fluid flowing over a horizontal flat plate as shown in Fig.

The velocity of the fluid is before it encounters the plate. As the fluid touches the
plate, the velocity of the fluid layer just adjacent to the plate surface becomes zero due to the
no slip boundary condition. This layer of fluid tries to drag the next fluid layer above it and
reduces its velocity. As the fluid proceeds along the length of the plate (in x-direction), each
layer starts to drag adjacent fluid layer but the effect of drag reduces as we go further away
from the plate in y-direction. Finally, at some distance from the plate this drag effect disappears
or becomes insignificant. This region where the velocity is changing or where the velocity
gradients exists, is called the boundary layer region. The region beyond boundary layer where
the velocity gradients are insignificant is called the potential flow region.
Fig 1.6 External flow over a flat plate

As depicted in Fig. 8.2, the boundary layer keeps growing along the x-direction, and may be
referred to as the developing flow region. In internal flows (e.g. fluid flow through a pipe), the
boundary layers finally merge after flow over a distance as shown in Fig. 8.3 below.

Fig 8.3 Developing flow and fully developed flow region

The region after the point at which the layers merge is called the fully developed flow region and
before this it is called the developing flow region. In fact, fully developed flow is another
important assumption which is taken for finding solution for varity of fluid flow problem. In the
fully developed flow region (as shown in Figure 8.3), the velocity vz is a function of r direction
only. However, the developing flow region, velocity vz is also changing in the z direction.
Non-Newtonian fluids
Non-Newtonian fluids are the fluids which do not obey Newton’s law of viscosity. For
describing Non-Newtonian fluids, let’s recall the Newton's law of viscosity experiment. There
are two long parallel plate situated at distance h to each other. Top plate is stationary and bottom
plate is moving with velocity as shown in Fig. (20.1).

Fig 1.30 Non-Newtonian flow between two parallel plates

If a force, F, is applied to move plate, then ( )

and under steady state conditions when h is small and when

Now, we calculate by repeating experiments for different applied forces and velocity
achieved by the bottom plate and plotting a graph as shown in Fig. (20.2). Depending on the
nature of fluid, different types of curves may be obtained.

84
Fig 1.31 Shear stress vs. shear strain diagram for Newtonian and non-Newtonian fluids
Rheological behaviour of fluids

If fluid shows the behaviour like curve (1) then it is a Newtonian fluid. Other fluids are non-
Newtonian fluids. Curve (2) represents a Pseudo-plastic fluid, curve (3) represents a Dilatant
fluid, and curve (4) represents a Bingham plastic fluid. There are several Theoretical and
empirical models available to describe the rheological behaviour of non-Newtonian fluids. Here,
we discuss some of them, which come under the group of generalized Newtonian models. Basic
equation for a generalized non-Newtonian fluid is given below

Here, is the apparent viscosity, which is clearly a function of shear rate as may be seen from
Fig. (20.2). Therefore,

If the apparent viscosity increases with increase in shear rate, , then the fluid is called

Dilatant fluid and if it decreases with increase in shear rate, then fluid is called Pseudo-
plastic fluid. Some fluids require a critical shear stress to initiate the flow. These fluids are called
Bingham fluids. Some important rheological models for non-Newtonian fluids are given below.

85
Rheological models

1 Power Law or Ostwald De Waele model

Power law or Ostwald De Waele model is the most generalized model for non-Newtonian fluids.
The expression of this model is given in Equation (20.3)

Here, apparent viscosity is defined as,

This is a two-parameter model where m and n are the two parameters.


If n = l then =m
where m is similar to the viscosity of the fluid and model shows the Newtonian behaviour .
If n>1, then increases with increasing shear rate and the model shows the Dilatant behaviour.

If n<1, then decreases with increasing shear rate and the model shows the Pseudo-plastic
behaviour.

Modulus sign

In power law model, modulus sign can be removed according to the value of shear rate.

1. If is positive, then

2. If is negative, then

86
Several fluids do not show single type of rheological behaviour. They show Newtonian
behaviour for a range of shear stress and Non-Newtonian behaviour for some other ranges of
shear stresses. Several models have been suggested for these types of fluids. Some popular
models like Eyring model, Ellis model, Reiner Philipp off model and Bingham Fluid model are
discussed here.

2. Eyring model

Eyring model is a two-parameter model. The equation of Eyring model is as follow

where A, B are the two parameters.

In Eyring model, if → 0 which means very low shear forces, we have

Therefore, as → 0, the model shows Newtonian behaviour

Here, viscosity =

If is very large, the model shows Non-Newtonian behaviour as shown Fig. 20.3

87
Fig 1.32 Shear stress vs. shear strain diagram for Eyring model

Therefore, Eyring model may be used for a fluid which shows Newtonian behaviour at low shear
rates and non- Newtonian behaviour at high shear rates.

3. Ellis model

Ellis model is a three-parameter model. The equation of this model is as follows

Here, , and are the three parameters .

Here, we consider some special cases,

1. If then Equation (20.11) reduce to

or

which is same as Newton’s law of viscosity with as the viscosity of the fluid.

88
2. If , then

which is similar to a Power law model

3. If >1 and is small then the second term is approximately zero and equation reduces to

which is similar to Newton’s law of viscosity.

4. If <1 and is very large, then again, second term is negligible and we have

Which again shows Newtonian behaviour. Therefore, Ellis model may be used for fluids which
show Newtonian behaviour at very low and very high shear stresses, but non-Newtonian
behaviour at intermediate value of shear stresses.

89
Fig 1.33 Shear stress vs. shear strain diagram for Ellis model

This type of behaviour may be shown by some polymer melts

4. Reiner Philipp off model

This is also a three-parameter model. The equation of Reiner Philipp off model is as follows,

where, , and are the three parameters.

In Reiner Philipp off model, if is very large, the equation reduces to,

or

90
which is same as the Newton’s law of viscosity,

If is very small then equation reduces to

or

which is also same as the Newton’s law of viscosity. Therefore, Reiner Philipp off model may be
used for a fluid which shows Newtonian behaviour at very low and very high shear stresses but
non-Newtonian behaviour for intermediate values of shear stress. Here, and represent
the viscosity of fluid at very low and very high shear stress conditions respectively.

5. Bingham Fluid model

Bingham fluid is special type of fluid which require a critical shear stress to start the flow.
The equation of Bingham fluid model are given below

if

91
if or

A typical shear stress vs. shear rate diagram for a Binghum model is shown below

Fig 20.5 Shear stress vs. shear strain diagram for Bingham model
Momentum transport problems involving Power law and Bingham fluids:

You might also like