CVX5531 Navier Stokes Equations Note

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Session 01 Navier-Stokes Equations and

Approximations

AIM:
To study the fundamentals of viscous fluid flow.

OBJECTIVES:
 To derive the Navier-Stokes equation for an
incompressible Newtonian fluid.
 To write down the equations of motion for any
incompressible Newtonian fluid flow.
 To explain the usefulness of Hele-Shaw model.
 To solve some simple Newtonian fluid flow problems
using appropriate co-ordinate systems.

1.1 Introduction
Fluids with zero viscosity do not exist in reality. However, there
are many problems where an assumption that the fluid viscosity is
negligible will simplify the analysis and at the same time lead to
meaningful results. All fluids possess a viscosity, which is the
shearing motion of a fluid. The fluid in which the shearing stress
is linearly proportional to the rate of strain is termed a Newtonian
fluid. A non-Newtonian fluid is one for which the dependence of
shearing stress on rate of strain is nonlinear.

In solid mechanics, a material is considered perfectly elastic if it


undergoes a finite deformation when a load is applied. The
deformation vanishes as the load is removed. Since the
deformation is finite and constant for a given load, it cannot have
any rate of deformation as long as the applied stress does not
exceed the yield stress of the material. The proportionality of the
stress and the strain is constant, and the material is within elastic
limits. When the applied load exceeds the yield stress of the
material, the strain is no longer constant, and a continuous
deformation associated with a certain rate appears. If the ratio of
the stress and the rate of strain is linear, the material is ideally
plastic.

1
Fig. 1.1 Rheological Classification of Fluids

The differences in shearing stress and rate of strain relationship


for solids and fluids are shown in Fig. 1.1. The behavior of non-
viscous (ideal) fluid for which there is no shearing stress
represents a perfectly elastic material. The curves of the
Newtonian fluid and the ideal plastic solid have constant slope.
However, the latter behaves like perfectly elastic representing the
non-Newtonian fluid display the variable coefficient of viscosity.

In this lesson, we shall derive the general equations of motion


namely Navier-Stokes equations, which may be used to describe
the flow of a Newtonian fluid. It will be seen that the complete
equations can only be solved for a few and relatively simple cases.
The various approximations for solving the equations for a wider
range of problems will be discussed with some simple examples.

1.2 Continuity Equation


In a real fluid, mass must be conserved; it cannot be created or
destroyed. To develop a mathematical equation to express this
concept, consider a very small cube located with its center at x,y,z
in a Cartesian coordinate system as shown in Fig. 1.2. For the
cube with sides x, y and z , the rate at which fluid mass flows
into the cube across the various faces must equal the sum of the
rate of mass accumulation in the cube and the mass fluxes out of
the faces.

2
Fig. 1.2 Reference cube in a fluid.

Taking first the x-face at x  x / 2 , the rate at which the fluid


mass flows in is equal to the velocity component in the x-direction
times the area through which it is crossing, all multiplied times
the density of the fluid,  . Therefore, the mass inflow rate at
x  x / 2 , or side ACEG, is
 x  x / 2, y, z ux  x / 2, y, z yz
where the terms in parentheses denote the coordinate location.

This mass flow rate can be related to that at the center of the cube
by the truncated Taylor series, keeping in mind the smallness of
the cube,
Taylor series,

f x   f a   f (a)
x  a   f (a) x  a 2  .......... .....
1! 2!

x
x x
Replacing, 2 ax

x
xa 
i.e 2
we obtain,

 x   x 
 x  , y, z u x  , y, z yz 
 2   2 
 u  x  2 u   x  1 
2

  x, y, z u x, y, z       ..... yz


 x 2 x  2  2
2


3
The mass flow rate out of the other x-face, at x  x / 2 , or side
BDFH, can also be represented by the Taylor series,
 x   x 
 x  , y, z u x  , y, z yz 
 2   2 
 u  x  2 u   x  1 
2

   x , y , z u  x , y , z       ..... yz
 x 2 x  2  2
2

Disregarding higher-order terms O x   and by subtracting the
4

mass rate out from the mass rate in, the net flux of mass into the
cube in the x-direction is obtained, that is, the rate of mass
accumulation in the x-direction:
  u 

x
xyz  O x 4  
Now considering all three directions, the net rate of mass
accumulation inside the control volume due to flux across all six
faces is,
u  v  w
 xyz  xyz  xyz  Ox 4 
x y z

The mass of volume at time, t =  (t ) xyz

The mass of volume at time, t  t =  (t  t )xyz

Recalling Taylor series,


Replacing, x  t  t a t
i.e x  a  t
we obtain,
 t  2  t 
2
 t  t    t     .......... .....
t 1! t 2 2!
The increase in mass is given by,
 (t  t )   t xyz
From Taylor series expansion we can obtain,

 (t  t )   t xyz    t  Ot 2 xyz


 t 
where, Ot  represents the higher order terms in the Taylor
2

series. Since mass must be conserved, this increase in mass must


be due to the net inflow rate occurring over a time increment t ,
that is,

4
  2
 t t  Ot   xyz 
  u   v   w
 x  y  z  xyzt  O x t
4
 
 

This results,
 u  v  w
   0 E (1-1)
t x y z

By expanding the product terms, a different form of the continuity


equation can be derived.
 u  v  w 
 u   v  w 0
t x x y y z z
       u v w 
 t  u x  v y  w z     x  y  z   0
   

By chain rule, we obtain,


d  x  y  z 
   
dt t t x t y t z
d    
 u v w
dt t x y z
Therefore,
d   u   v   w 
   0 E (1-2)
dt  x y z 

If the fluid is incompressible,


d  u v w 
0 i.e.  x  y  z   0 E (1-3)
dt  

This equation is also referred to as “Continuity Equation”.

5
Fig. 1.3 Cylindrical Coordinates
The continuity equation for incompressible fluid flow in terms of
cylindrical coordinates x, r, θ becomes,

𝜕𝑢 𝜕𝑣𝑟 𝑣𝑟 1 𝜕𝑣𝜃
𝜕𝑥
+ 𝜕𝑟
+ 𝑟
+𝑟 𝜕𝜃
=0 E (1-4)

where u, 𝑣𝑟 and 𝑣𝜃 are the components of velocity in the axial (x),


radial (r) and circumferential (θ) directions respectively.

1.3 The Equations of Motion


The equations of motion are obtained by applying Newton‟s
second law, i.e. the rate of change of momentum is equal to the
sum of the external forces. We may consider the forces
experienced by the fluid to be of two types.
(i) Body forces such as gravitational and magnetic forces and
(ii) Boundary forces such as pressure forces and viscous
forces. If the body force per unit volume has components
x, y, z and the boundary force on unit area has components
Fx , Fy , Fz , the equations of motion may be written for
flow in the directions of the x, y and z axes.

By chain rule, we obtain total acceleration du / dt  as,


du u u u u
 u v w
dt t x y z E (1-5)

where, u / t  is the local acceleration and uu x  vu y  wu z is the


advective acceleration.

6
Fig 1.4 Shear and Normal Stresses on a Fluid Cube.

There are nine stresses that are exerted on the cube faces. Three of
these stresses include the pressure, as the normal stresses are
written as,
 xx   p   xx
 yy   p   yy
 zz   p   zz
  xx   yy   zz 
Where, p   
 [hydrostatic pressure]
 3 
[shear stresses  xx , yy , zz are not well-defined]
The equation of motion in the x-direction can be formulated as:
du
F x m
dt
From Figure 2, the surface forces can be obtained on the six faces
via the truncated Taylor series,
  xx x    xx x    yx y 
  xx  yz    xx  yz   yx  xz

 x 2   x 2   y 2 
  yx y    z 
  yx  xz   zx  zx
 xy
 y 2   z 2 
  z  du
  zx  zx xy  xyzX  xyz
 z 2  dt

xyzX is to represent any body forces in x-direction, where X


is the body force per unit mass.

Combining terms and dividing by the volume of the cube yields,

7
du 1   xx  yx  zx 
     X
dt   x y z 

 xx   p   xx
We know,  yy   p   yy
 zz   p   zz

du 1 p 1   xx  yx  zx 
      X
dt  x   x y z 

By similar developments, the equations of motion are obtained for


y- and z-directions as:
dv 1 p 1   xy  yy  zy 
     Y
dt  y   x y z 

dw 1 p 1   xz  yz  zz 
     Z
dt  z   x y z 

Y , Z are the body forces per unit mass in y- and z-directions


respectively. [In z-direction Z   g ]. The above equations are
called Cauchy‟s equations.

In many real flow cases, the flow is turbulent and shear stresses
are influenced by the turbulence. The Newton‟s law of viscosity
gives,
u   u v   u w 
 xx  2   divV ,  xy      ,  xz     
x  y x   z x 
 u v  v   v w 
 yx      ,  yy  2  divV ,  yz     
 y x  y  z y 
 u w   v w  w 
 zx      ,  zy      ,  zz  2  divV
 z x   z y  y


For incompressible flow, divV  0 . Then the Cauchy‟s equations
can be re-written as;
x-direction
du 1 p    2 u  2 u  2 v  2 u  2 w 
  2      X
dt  x   x 2 y 2 xy z 2 xz 
du 1 p    2 u  2 u  2 u     u v w 
          X
dt  x   x 2 y 2 z 2   x  x y z 

8
u v w
We know,    0 , hence we can write the following
x y z
equations for all three directions.
du 1 p    2 u  2 u  2 u 
      X
dt  x   x 2 y 2 z 2 
dv 1 p    2 v  2 v  2 v 
     Y
dt  y   x 2 y 2 z 2 
dw 1 p    2 w  2 w  2 w 
     g
dt  z   x 2 y 2 z 2 
E (1-6)

These are the equation of motion for the laminar flow of a viscous
compressible fluid and they are known as Nervier-Stokes
equations.

The Navier-Stokes equation in cylindrical coordinates are given


below, where u, 𝑣𝑟 , v  are the components of velocity in the
directions of x, r and  coordinates respectively. The body forces
are omitted.

du 1 p    2 u  2 u 1 u 1  2 u 
      
dt  x   x 2 r 2 r r r 2  2 
dvr 1 p    2 v r  2 v r 1 v r v r 1  2 vr 2 v 
 
  2      
dt  r   x r 2 r r r 2 r 2  2 r 2  
dv 1 p    2 v  2 v 1 v v 1  2 v 2 v r 
 
  2      
dt r    x r 2 r r r 2 r 2  2 r 2  
E (1-7)

where total derivative is given by,


du u u u v u
 u  vr 
dt t x r r 

dvr v r v v v v v2
  u r  vr r   r  
dt t x r r  r

dv v v v v v vv
  u   vr      r 
dt t x r r  r E (1-8)

9
1.3.1 Approximations for Cases of Small and Large Viscosity
The equations of motion may be simplified to suit more restricted
classes of flow. For example, if we consider the case of flows with
very small viscosity, we obtain from Eq. (1.6) the following
equation of motion for an incompressible inviscid fluid. This
condition which is approach as the Reynolds number becomes
very large.
du 1 p

dt  x
dv 1 p

dt  y
dw 1 p
 g
dt  z E (1-9)

The other Limiting case is that for flow with low Reynolds
number, in which the viscous forces are large in comparison with
the inertia forces. In this case, the inertia terms in the equation of
motion may be neglected, and we obtain  q  grad p or
2

p   2u  2u  2u 
   2  2  2   X
x  x y z 

p   2v  2v  2v 
   2  2  2   Y
y  x y z 

p  2w 2w 2w 


   2  2  2   g E (1-10)
z  x y z 

Flows of this type in which forces predominate, are often referred


to as creeping motions. Hydrodynamic lubrication and flow
through porous media are some of the examples of creeping flow.

1.3.2 Hele-Shaw Model


Flow between two parallel plates kept at very small distance „2h‟
apart could be analogous to potential flow. This flow‟s usefulness
lies in the fact that its stream lines, as viewed from above the
plates, are those corresponding to a potential flow and for that
reason can be used in a study of two-dimensional irrotational flow
past obstacles of arbitrary shapes. The stream lines in such an
apparatus, in which the top plate is made out of glass, are made
visible by injection of a dye into the fluid between the plates.

10
Steady creeping motion takes place in which the Reynolds
number is very small (Re << 1).

Fig. 1.5 Hele-Shaw Model

Continuity equation of 2D incompressible flow


u v
 0
x y

Navier-Stokes equation in x-direction


p *   2u  2u  2u 
   2  2  2 
x  x y z 

N-S Equation in y-direction


p *   2v  2v  2v 
   2  2  2 
y  x y z 

Thus taking only the terms whose magnitudes are not small we
obtain,
p *  2u p *  2v
 2  2
x z y z

Integrating twice w.r.t. z and applying boundary conditions z = +


h; u = v = 0 we obtain the following expressions velocity
components

Fig. 1.6 Shaw Model Past an Aerofoil

11
h 2 p *  z 2  p *
u 1  2   C ( z )
2 x  h  x
h 2 p *  z 2  p *
v 1  2   C ( z )
2 y  h  y

Considering the flow in x, y plane


(a) Continuity equation
u v  2 p*  2 p*
 0    2 p*
y y x 2 y 2
(b) Vorticity
u v  2 p*  2 p*
  C ( z)  C ( z) 0
y x xy yx

Thus observing the flow in x-y plane, it is a potential flow. The


equations for the velocity components u, and v are derivable from
a potential
h2  z2  
  x, y    1  2  p
2  h 
and therefore streamline of the Hele-Shaw flow corresponds to
those of a potential flow.

1.4 Application of Navier-Stokes Equation


Although the assumption of incompressible flow leads to marked
simplifications of the equations of motion, a general analytical
solution of the Navier – Stokes equations for Laminar flow has
never been obtained for particular systems which allow the
equations of motion to be further simplified.

SAQ. 01
Determine the distribution of velocity of the laminar flow of an
incompressible fluid through the annular space between two
coaxial cylinders of radius and b. Fluid flows steadily parallel to
the axis of the cylinder determine the shearing stress at the all of
the inner cylinder.

12
Solution
The only non-zero velocity component is the axial component
v  v r  0 . Thus the continuity equation (Eq. 1.4) gives,
u
0
x

Neglecting the inertial terms in the equations of motion (Eq. 1.7),


the flow in the radial and circumferential directions vanish giving
p p
  0
r 

While the equation of flow in the axial direction becomes,


d 2 u 1 du 1 dp
 
dr 2 r dr  dx

d  du  r dp
or r  
dr  dr   dx

Integration of this expression gives the velocity gradient,


du r du C1
 
dr 2 dr r

Integrating the velocity gradient gives,


r 2  dp 
u    C1 ln r  C 2
4  dx 

The boundary conditions are that there is no slip adjacent to both


walls

Fig. 1.7 Flow through Gap between Cylinders

i.e. r = a or b u=0

13
u
1  dp  2

  r  C1 ln r  C 2
4  dx 

a 2  C1 ln a  C 2  0
b 2  C1 ln b  C 2  0

 
C1   b 2  a 2 / lnb a 
C 2  a 2  C1 ln a
So,

u
1 dp 2
4  dx

r  a 2  C1 lnr / a  
du 1 dp  1
 2 r  C 
dr 4 dx  r 
1

The shearing stress in the laminar boundary is given by,


du
 r a  
dr r  a

 r a  .
1 dp 


b2  a2 1  
 2 a 
4 dx  lnb / a  a 

 r a 
1 dp 

b 2  a 2  1 
 2 a 
4 dx  lnb / a  a 

SAQ. 02
Determine the distribution of velocity in the annular gap between
two concentric cylinders, the inner cylinder being stationary and
the outer cylinder rotating with constant angular velocity.

Solution

Fig 1.8 Rotating Cylinder

14
Here u  vr  o and the continuity equation, Eq. (1.4) gives
v
O


It means that the velocity varies with equation radius only


considering the equations of motion in cylindrical coordinates [Eq.
1.7],
(i) The equation for axial flow vanishes
p
0
x
(ii) The equation for radial flow becomes
dp 2 v dp
2
v
   
r dr r  dr
(iii) The equation for circumferential flow gives [since
p   0 ],
d 2 v 1 dv v
  0
dr 2 r dr r 2

d 2 v d v 
or    0
dr 2
dr  r 

The boundary conditions are, r  a v  0;


r  b  v  b

The first integration of the equation for circumferential flow gives,

dv v
  2C , or
1 d
rv   2C1
dr r r dr

Second integration of the same equation gives,


C2
v  C1r 
r

Considering the boundary conditions,


(i) r  a; v  0
C2
So C1 a   0;
a
C 2  C1 a 2

15
(ii) r  b, v  b
C2  a2 
b  C1b   C 
1 1  b
b2  b2 

C1 
 a2 
1  2 
 b 
So velocity distribution,
  a2 
v   r  
 a 2   r 
1  2 
 b 

16

You might also like