Download as doc, pdf, or txt
Download as doc, pdf, or txt
You are on page 1of 20

TRANSPORT PHENOMENA

Unit-II

Theory of viscosity of gases at low density


Consider a pure gas composed of rigid non-attracting spherical molecules of diameter
‘d’, mass ‘m’, present in a concentration of ‘n’ number of molecules per unit volume.
Assume that ‘n’ to be very small enough so that the distance between molecules is many
times their diameter ‘d’.
In such a gas mixture the average molecular speed U in m/s according to kinetic
SKT
theory of gases given by U   (1)
m

Where U = avg molecular speed in m/s.

kgm 2
K = Bolt’s mann constant in 2
sk
T = absolute temperature in K.
m = mass of the gas molecule in kg.
The average molecular collision on one side of any stationary surface per unit area in
1
given by Z  nU  (2).
4
Where Z = wall collision frequency. 1
m2 sec

n = number of molecules per unit volume 1


m3

U = average mole. Speed m/s


The average distance traveled by a gas molecule between successive collision is given
1 1
by    (3) is also called mean free path.
2 d 2 n
Where,
 = mean free path ‘m’
d = diameter of gas molecule ‘m’
n = number of molecular per unit volume 1/m
Figure

Distance between the planes assumed here is given as a, given by

2
a   (4)
3
Where
a  distance between planes in m
  mean free path in m.
To determine the viscosity of such gases in term of molecular property, consider the
dVx
behaviour of gas when it flows in the direction parallel to x axis with a velocity gradient
dy
. The ‘x’ component momentum flux at any given constant ‘y’ can be written as
yx = Rate of momentum in – Rate of momentum out from plane y.
 Z m Vx|ya  (5)

= Z m Vx|ya

kg kg
2
 2 .m / s
ms m sec

The expressions for v x / y a and v x/y+a by various other assumptions can be written as
follows
dVx 
Vx / y a  v x / y  32 
dy 
  (6)
dVx 
Vx / y a  vx/ y  3 
2
dy 
Substitute equation (6) & equation (2) in equation (5)

 yx  z m v x / y a  Z m Vx / y a
= 1
4nUm U x/y  23  dVx
1
dy   4 nUm Vx/y  3  dy 
2 dVx

 2 dV 2 dV 
= 41 nUm  Vx / y   x  Vx / y   x 
 3 dy 3 dy 
  2 dV  
= nUm  2   x  
1
4
  3 dy  
1 dV
 yx   nUm x  (7)
3 dy

We know that by Newton’s law of viscosity


dVx
 yx    (8)
dy

Compare equation (7) and (8)

dVx 1 dVx
   nUm
dy 3 dy
1
= nUm  (9)
3
1
 = nU  (10) n= 1
m3
m = kg.
3

The expression for viscosity in some other form can be obtained by substituting
equation (1) and equation (3) in equation (9).

1 8KT 1 1
 n m
3 m 2 d 2 n
1 KTm 2 1
= 2 2
3 m2 2 d2

2 KTm

3d 2 3

Equation represents viscosity of gas composed hard spheres at low density.


The prediction of viscosity in independent of pressure agrees well with experimental
data upto 10 atm pressure and predicted temperature depends in less satisfactory. Therefore
the co- efficient of viscosity at absolute temperature to of a pure t of a pure mono atomic gas
of molecular weight M may be written in terms of parameters  (Characteristic diameter or
collision diameter)  (characteristic energy of interaction between the molecules is

MT
  2.669  10 5  (12)
 2 

Where,  = collision integral (dimensionless)


Although this formula was derived for monoatomic gases it has been found to be the
seasonably good for polyatomic gases. It should be noted that viscosity of gases at low
density increases with temperature roughly as the 0.6 – 1.0 power of the temperature and this
also noted that there is no dependence on pressure in the low density rate.

Molecular theory of viscosity of liquids.


A rigorous kinetic theory of transport properties of monatomic liquids was developed
by Kirkwood and coworkers. However this theory does not lead to easy to use results. An
older theory, developed by Erying and coworkers, although less well grounded theoretically,
does give a qualitative picture of the mechanism of momentum transport in liquids and
permits rough estimation of the viscosity from other physics properties. In a pure liquid at
rest the individual molecules are constantly in motion. However, because of the close
packing, the motion in largely confined to a vibration of each molecule within a “cage”
formed by its nearest neighbours. This cage is represented by an energy barrier of height
G 0 / N, in which G0 is the molar free energy of activatic for escape from the “cage” into
an adjoin “hole”, and that the molecules thus more in each of the coordinate directions in
jumps of length a at a frequency v per molecule. The frequency is given by the rate equation,

KT 
v exp(  Go / RT)  (1)
h
In which k and h are the Boltzmann and Planck constants, N
 is the Avogadro number
and R = N k in the gas constant.

dVx
In a fluid that is flowing in the x direction with a velocity gradient , the frequency of
dy
molecular rearrangements is increased. The effect can be explained by considering the
potential energy barrier as distorted under the applied stress yx, so that

a  yx V 
 G    G o       (2)
 s   2 
Where V is the volume of a mole of liquid and (a / s)(  yx V / 2) in an approximation
to the work done on the molecules as they more at the top of the energy barrier, moving with
the applied shear stress (plus sign) or a against the applied shear stress (minus sign). We now
define V+ as the frequency of forward jumps and v - as the frequency of backward jumps.
Then from equations we find that
kT
v  exp(  G o / RT)exp( a yx V / 2RT)  (3)
h
The net velocity with which molecules in layer A slip ahead of those in layer B in just
distance traveled per jump (a) times the net frequency of forward jumps (v+ - v-); this gives
v xA  v xB  a(v   v  )  (4)

The velocity profile can be considered to be linear over the very small distance 
between the layers A and B, so that
dVx  a 
   (v   v 1 )  (5)
dy   

By combining equation, we obtain finally

 ayx v 
dVx  a  KT
  
dy    h

exp  G o / RT    exp  2RT    exp(a yx v / 2R)
  
 ayx v 
=   exp  G o / RT   2 sin h
 
a KT
   (6)
   h  2RT 

This predicts a nonlinear relation between the shear stress (momentum flux) and the
velocity gradient – that in, non-Newtonian flow such nonlinear behaviour is discussed further
in.
The usual situation, however, is that ayx v / 2RT  1 . Then we can use the Tayler
series sin hx = x  (1/ 30 x 3  ( 1 50 x 5  .........) and retain only one term. Equation 6 is then of
the form of which the viscosity beings given by

 Nh
a

    exp(  G o / RT)  (7)


a v

The factora can be taken to be unity, this simplification involves no loss of

accuracy. Since Go is usually determined empirically to make the equation agree with

explanation viscosity data.

It has been found that free energies of activation,  Go determined by fitting equation
7 to experimental data on viscosity for a given fluid and are simply related to the internal
energy of vaporization at the normal boiling point, as follows

 G o  0.408 U vap  (8)



By using this empiricism and setting  1 equation 7 becomes
a

Nh
 exp 0.408 U vap / RT  (9)
v
The energy of vaporization at the normal boiling point can be estimated roughly from
Trouton’s rule

U vap  H

vap  RTb  9.4RTb  (10)

With this further approximation equation (9) becomes

Nh
 exp(3.8 Tb / T)  (11)
v
Equation (9) and (11) are in agreement with the long-used and apparently successful
empiricism    exp(B / T) . The theory, although only approximation in nature, does give
the observed decrease of viscosity with temperature, but errors of as much as 30% are
common when equation (9) and (11) are used. They should not be used for very long slender
molecules, such as n  C 20 H 42 .

Escape process in the flow of a liquid (Molecule 1 must pass through a “bottleneck” to reach
the vacant site).

General procedure for setting up and solving momentum balance boundary conditions
for solving momentum transfer
Velocity distribution in laminar flow
The velocity distributions under laminar condition for different types of flow
condition are derived by writing the momentum balance.
Shell momentum balance:
The shell momentum balance under steady state condition can be written as
Rate of momentum in - Rate of momentum out + sum of forces acting on the system = 0 
(1)
Momentum transfer in due to two factors:
 Due to fluid property – bulk viscosity of fluid
 Due to bulk motion of fluid.
The momentum transfer due to 1st factor occurs in the direction perpendicular to the
fluid flow direction, and then the second factor it occurs in the direction of fluid flow.
The possible forces that may act on any system:
 Pressure force
 Viscous force (or) shear force
 Gravitational force
General procedure for setting up and solving momentum balance:
1. Write the momentum balance in the form of equation (1) for a given shell of finite
thickness.
2. Let this thickness approaches zero and make use of mathematical definations of 1 st
order derivations to obtain corresponding differential equations describing momentum
flux distribution.
3. Insert appropriate Newtonian or non-Newtonian expressions for momentum flux to
obtain differential equation for velocity distribution.
4. The integration of these two differential equations (equation from step (2) & (3)),
yields momentum flux and velocity distribution for the given system respectively,
which is further useful to calculate various other quantities such as average velocity,
maximum velocity, volumetric flow, pressure drop and forces on the boundaries.
5. In the two integrations mentioned above various constants of integrations appears
which are then evaluated by the use of boundary conditions.

Expression for Max velocity & Average velocity in a flow of falling film.
1. At fluid solid interfaces the velocity of fluid in equals to the velocity with which the
solid surface itself moving.
2. At liquid-gas interfaces the momentum flux (and hence the velocity gradient) in the
liquid phase is very nearly zero and can be assumed to be zero in most of the cases.
3. At liquid-liquid interfaces the momentum flux perpendicular to the interface and the
velocity are continuous across the interface.

Consider the steady flow of Newtonian fluid with constant density and viscosity over a flat
inclined surface from a reservoir as show in figure, our attention in focused over a region of
length L where entrance disturbance or exist disturbance are not included.
Let  - density of fluid kg/m3
 - viscousity of liquid kg/ms
L – region of length L in which the velocity profile in fully developed(or) length
of the liquid film which in under study in ‘m’.
s – Thickness of liquid film in ‘m’
W - Width of liquid film ‘m’ in ‘y’ direction.
 - angle of inclination of flat surface from the reservoir vertical wall.
xz – the ‘Z’ momentum flux at any constant position ‘x’ kg/ms2
vz – Velocity in Z direction m/s
x – small thickness of liquid film over which the momentum balance is made, ‘m’.
The steady state momentum balance over a shell of thickness x length ‘L’ with ‘W’
can be written as, Rate of momentum in – Rate of momentum out + sum of forces acting
on the system = l
Rate of momentum in across the = Momentum flux  area
xz X
surface due to bulk viscosity of fluid at x = x = LW
x=x

Rate of momentum out across the


xz X
surface due to bulk viscosity of liquid = LW
x = x

at x = x + x
Rate of momentum in across the
momentum
surface due to bulk motion of liquid =
time
mass  velocity
at Z = 0 =
time
den  vol  velocity
=
time
length  V
= den  Area 
time

= den  Area  Velocity2


= AV2

= (Wx)v z Z  0
2

Rate of momentum of

= (Wx)v z Z  L
2
across the surface due to

bulk motion of liquid Z = L

The gravitational force acting on the system = mass  gZ


= density  volume  gZ
= (LWx)g Z

= (LWx)g cos 
Substitute all the above expression in equation (1):
0
LWxZ x  x  LWxz x  x x  Wx v z2 Z 0

0
Wx v 2z Z L
 (LWx)g cos  = 0

The third and fourth term get cancelled because Vz is same at Z = 0 and Z = L
LWxZ x x  LWxz x x x  (LWx)g cos  = 0

 the above equation by LW x


xz xz
xx  g cos   0
x x x  x x

xz
xx  xz xxx
 g cos   0
x
xz
xx  xz xxx
 g cos   0
x
Taking the limit x  0
xz xx xz
Lt x  x x
 g cos 
x 0 x
The left hand side of the above equation in the definition for 1st order differential equation,
d( xz )
 g cos 
dx
d( xz )
 g cos 
dx
Integrating

 d( xz )  g cos  dx
xz  g cos  x  C1  (2)

Boundary condition:
(i) At x = 0 xz = 0
Apply B.C at equation (2)
0  g cos (0)  C1
C1  0
xz  g cos x  (3)

Equation (3) in the expression for momentum flux distribution at any given point ‘x’ since
the fluid under study in Newtonian, Newton’s law of viscosity is applicable is given by
dv z
xz    (4)
dx
Compare equation (3) & (4)
dv z
  g cos  x
dx
dv z g cos x

dx 
Integrating
dv z g cos  x
 dx

  x dx
g cos  x 2
vz    C 2  (5)
 2
0
B.C (ii) @ x = s , vz = 0

Apply B.C in equation (5)


g cos s 2
0  C2
 2
g cos  s 2
C2   (6)
2
sub (6) in (5)
g cos x 2 g cos  s 2
vz   
 2 2
g cos 2
vz  (s  x 2 )
2

g cos  2
s (1   x s   (7)
2
vz 
2

Equation (7) in the expression for velocity distribution at any given point ‘x’.
Maximum velocity:
B.C (iii) x = 0, vz = vx max
Sub B.C. in equation (7)

g cos  s 2
v z max   (8)
2

Average Velocity:
Average velocity is obtained by summing up all the velocities at a particular c.s
divided by the c.s. area.
w s

  v dx dy
z

 v z  0 o
w s

  dx dy
0 o
Ww
g cos 2
= 1 Ws   s (1  x 2 s 2 )dx dy
o o
2
g cos s 2 w
[(x  s12   ] dy
2ws 0
 x3
3
s
0

g coss w s3
2w o
 s  dy
3 s2
g coss 2s w
 (y)o
2w 3
g coss 2
 w
w 3

g cos  s 2
 v z   (9)
3

Volumetric flow rate:


Q = c.s. Area  average. Velocity
m3 s  m2  m s
g cos s 2
= ws 
3
g cos  s 3 w
Q  (10)
3

Expression for film thickness s:


Rearrange equation (9) & (10) for s

 v z  3
s
g cos
or
3
s=
g cos w

Forces on the wall surface:


Fs

A
Fs  xz x s x S.A xz equation (3)
= (g cosx) x=s x(LW)
Fs  shear force, at watt so Fw
Fw  g cos s L w

All the above expression are valid only if the flow in laminar.

Flow through a circular tube


Consider a steady flow of a Newtonian fluid with constant viscosity under steady state
condition throw a circular tube.
Assumption made:
1. The flow in laminar (i.e., NRe  2100)
2. Density of fluid is constant [incompressible flow]
3. The flow is independent of time. [steady state]

 dv 
4. The fluid is Newtonian  rz   z 
 dr 
5. End effects are neglected [if the selection of pipe of interest must be free from
entrances and exist disturbances.]
6. The fluid behaves as a continuum [fully developed flow].
7. There is no slip at the wall. [Velocity of fluid at wall surfaces in zero].
Let ‘R’ is the radius of the circular tube ‘m’.
‘L’ is the length of tube in ‘m’.
‘r’ in the thickness of the shell over which the momentum balance is made.
‘r’ radius of the shell which is ranging from o  R.
‘Po’ the pressure intensity at the z = 0 N/m2.
‘PL’ in the pressure intensity at Z = L.
rz in the momentum flux at any given constant radial distance ‘r’. kg/ms2.
vz Z – component velocity m/s
Q volumetric flow rate m3/s
The steady state momentum balance over the cylindrical shell of thickness ‘r’ can be
written as

Rate of momentum in – Rate of momentum out + sum of forces acting on the system
= 0 -----1)
Rate of momentum in across the cylindrical surface due to viscous transport at r = r
= mom. flux  Area

= rz rr (2 rL)

= (rrz r  rr 2 L)

Rate of momentum out across the cylindrical surface due to viscous transport at r = r + r
 (rrz ) rrr 2L

Rate of momentum in across the annular surface due to bulk motion at z = 0.

= (2 rr) v z z0


2

Rate of momentum out across the annular surface due to bulk motion at z =L.
 (2 rr) v 2z zL

The pressure force acting on the annular surface at z = 0,


FP
P
A
FP  P  A
= Po (2 rr)

The pressure force acting on the annular surface at z =L


FP  PL (2 rr) (down stream pressure)

-ve sign because pressure is acting opposite to direction of lfuid flow.


The gravitational force acting on the cylindrical volume
Fg = mg
= den  vol  g
= (2 rLr)g
Substitute all above equations in equations (1).
(rrz ) r r (2L)  [(r  r) rz ] r  rr 2 L 
0 0
(2 r r) U 2z z0  PL (2 r r )v 2z zL 
Po (2 rLr)  PL (2rr)+P(2rLr)g=0

Since vz = 0 and z = L, so 3rd & 4th term get cancelled.


(r  r)rz (2 L)  (rrz) rr (2 L) 
r r

2rr[Po  PL  PLg]

 by 2Lr
(r  r)r2 rr  rrz rr r
 (Po  PL )  rg
r L
The left hand of the above equation can be written as on differentiating

P P
(rrz )  r  o L  g 
d
dr  L 
P  P  gL 
 rrz   r  o L
d
dr  L 
d  Po  gz z 0  [PL  gz z 1 
[rr z ]  r  
dr  L 
P P
[rrz ]  r  o L   (2)
d
dr  L 
Where,
Po  Po  gz z 0
PL  PL  gz z L

PO and PL represents the combined effect of pressure and gravitational force.


Integrating,
Po  PL
 d(r rz )
L 
r dr

P  P r2
rrz  o L  c1
L 2
Po  PL
rz 

Here c, must be 0, since rz is cannot be intimate, so C1 = 0

P0  PL r
rz 
L 2
Equation (3) Z – momentum flux distribution at give radial distance ‘r’.
Since the fluid is Newtonian, Newton’s law of viscosity is applicable,
dv z
rz    (4)
dr
Compare equation (3) and (4)
dv z Po  PL r
 
dr L 2
dv z (PL  Po )r

dr 2L
Integrating
(PL  Po )
 dv z 
2 
rd

(PL  Po ) r 2
vz   c2
2L 2
r  R, v z  0
(PL  Po ) R 2
0  c2
2L 2
 (P  P ) 
c2    L o R 2 
 4 
(P  P ) r (PL  Po )rR 2
2
vz  L o 
2L 2 4
(P  P )
v z  L o [r 2  R 2 ]
4L
(PL  Po )r.R 2  r  2 
vz     1
4L  R  
(Po  PL ) 2 2
vz  [R  r ]
4L

Po  PL 2 
 
R 1  r R   (5)
2
vz 
4L  

Equation (5) in the expression for velocity distribution, and thin results tells that, for a
steady laminar incompressible Newtonian fluid in a circular tube follows the parabolic path.

Maximum Velocity:
r = 0; vz = vz max
Then equation (5)
Po  PL
v z max   (6)
4L

Average velocity:
<vz> is obtained by summing up all the velocity divided by the c.s. area
Sub equation (5)
2 R

 v z r dr d
 v z  0 0
2 R

  r dr d
0 0

 
2 R
 Po  PL 2 2 
0 0  4L R 1  r
R  r dr d

 
 v z 
 
2 R

 r R 0 d
2

 
2 R
 Po  PL  2
 R   r  R 2 dr d
r2
 4L
=   0 0
2

R
2
2 d
0

 
2 R
 Po  PL  2 r 2
 4L  R 
1 r4
2  R 2 4 d
  0
 0
R  0
2 2 
2 0

 Po  PL  2  R 2 R 2  2 
 4L  R  2  4   0  0
   
R2
 2
2
 Po  PL  2  2 R 2 
 4L  R  4  2 
   2 

R 2

(Po  PL )R 2
 v z   (7)
8L

Ratio of average velocity to maximum velocity


Divide equation (7) by (6)
 vz 

v z max
(Po  PL )R 2 4L

8L(Po  PL )R 2

 vz  1
 
v z max 2

Volumetric flow rate:


Q = Average velocity  c.s. Area
Q = <vz>  R2

(Po  PL )R 4
= Q  (9)
8L

Expression for pressure drop per unit length:


In terms of
(i) Average velocity:
(Po  PL ) 2
 v z  R  equ (7)
8L
From (7)
Po  PL 8  v z 

L R2
Po  PL 8  v z 

 
2
L P
2
P Po  PL 32  v z 
   (10)
L L D2

Equation (10) is Hagen’s poisuillie equation viscosity can be determine using this equation
(significance)
(ii) In terms of volumetric flow rate:
From (9)
(Po  PL )R 4
Q
8L
Po  PL 8Q

L R 4
8Q
=
 
4
 D2
Po  PL 128Q

L .D 4
P Po  PL 128Q
   (11)
L L D 4

Equation (10) & (11) are known as Hagen – Poisuillie equation.


The z-component forces on wetted wall surface:
Fz  rz (surface area)
r R

From equation (3) rz

 r(P  P ) 
Fz   o L  ( 2 RL )
 2 L  r R
Fz  (Po  PL )R 2  (12)

You might also like