Anda di halaman 1dari 20

MECH 430 FLUIDS 2

Flow of a compressible fluid in a constant area duct with friction

For an incompressible fluid where the density remains constant, friction results in
a pressure drop in the flow direction. For a compressible flow, viscous dissipation heats
up the gas resulting in a density change. Thus, the flow velocity and pressure can
increase or decrease depending on the rate of expansion due to density decrease and the
convective mass flux through the cross section of the pipe. As in the case of area
change, friction has opposite effect if the flow is subsonic or supersonic.
Let us first review the steady flow of an incompressible fluid in constant area pipe
with friction (which results in a wall shear stress W w ). Referring to the sketch below:
L

p + dp

Ww
The conservation of mass for steady flow gives U u=constant and since U constant, the
flow velocity is constant throughout the length of the duct. The momentum equation
gives:
pA  p  dp A  W w Aw

where A is the area of the duct ( A


( Aw

SD2

for a circular duct) and Aw is the wetted area


4
SDL for a circular pipe). Rearranging the momentum equation gives:

 dp W w

Aw
A

Ww L
D

where we have assumed a circular pipe for convenience. The wall shear stress is given
by:

wu

wr r

wu
where P is the coefficient of viscosity and
wr r

Ww

is the velocity gradient at the wall (if

we assume a laminar flow). The velocity profile for steady laminar incompressible flow
in a circular pipe is given by:

u
u

r 2
2 1 
R

where u is the mean flow velocity (i.e. the classical Hagen-Poiseuille Flow.) This
parabolic velocity profile gives the wall shear stress as:

Ww

wu

wr r


R

4 Pu
R

 8Pu
D

The negative sign denotes the direction of W w that gives the parabolic profile and we have
already considered the appropriate sign (i.e. direction) of W w when we write the
momentum equation. Thus, we write:
 dp
L

where Re D

p1  p2
L

4W w
D

32 Pu
D2

64 Uu 2 1

Re D 2 D

UuD
is the Reynolds number based on the pipe diameter. In general, we
P

define a friction factor f as:


f

4W w
u2
U
2

where f Re D is determined from experiments. (Moody diagram where f is given for a


wide range of Reynolds number and wall roughness). For the present case of laminar
Poiseuille Flow:

64
Re D

f
We may write the pressure drop as:
p1  p2
L

64
Re D

Uu 2 1

2 D

f Uu 2
D 2

For incompressible flow, we need not worry about the energy equation since U constant
and the problem is a dynamic problem with pressure forces balancing the friction forces.

Compressible Flow (Fanno Flow)


For steady compressible flow in a constant area, adiabatic duct with friction as illustrated
below, i.e.:

P+dp
U  dU
T+dT
u+du

u
p
T
Ww
dx

The conservation of mass gives:


m
dU

UuA constant


du
u

since dA=0 for a constant area duct.


The conservation of momentum can be written as:

1.

dpA  W w Dsx

mdu

2.

where S and A are the perimeter and the area of the cross-section of the tube, respectively.
Defining a friction factor:

Ww

u2
2

and a hydraulic diameter DH by:


4 * area
wetted perimeter

DH

4A
S

Equation 2 becomes:

Uudu  dp 

Uu 2 4 fdx
DH

3.

If the flow is adiabatic, the energy equation is given by:


u2
h
2

ho

constant

Note that there is no work done by the viscous stress W w since the velocity at the wall
vanishes. The no slip (i.e. u=0) condition at the wall generates a velocity profile resulting
in viscous dissipation. The viscous heating is at the expense of the kinetic energy
decrease in the flow but since there is no heat transfer, the viscous heating remains in the
flow. In the present assumption of a quasi one dimensional flow, the flow velocity is
uniform across the cross section of the duct. We should consider the flow velocity as an
averaged value to permit the presence of a velocity gradient for the viscous stress to
occur. In an inviscid flow, there will be no shear stress at the wall. For adiabatic (not
isentropic) flow where
h

u2
2

H0

cons tan t

we get
dh+udu=0

Since h=cpT, and c p

JR
J 1

JRT , the energy equation can be written as:

and c 2

dT
du
 J  1 M 2
T
u
The equation of state for a perfect gas is:

4.

URT

p
and hence
dU

dp
p

dT
T

5.

Rewriting Equation 3 as:

Uudu

 dp Uu 2 4 fdx

p
2 p DH

p
and noting that c 2

dp
Jp
, we obtain an expression for
as
p
U
dp
p

du JM 2
 JM

2
u
2

4 fdx

DH

From Equations 1, 4, 5 and 6, we can get an expression for

du
u

JM 2

4 fdx
2 1  M DH

6.

du
as
u
7.

From Equation 1, we see that:

dU

du
u

 JM 2 4 fdx
2 1  M 2 DH

From Equation 4, we get

dT
T

 J  1 M 2

du
u

8.

and substituting Equation 7 into the above yields

-J J -1 M 4 4 fdx
2 1- M 2 DH

dT
T

9.

Substituting Equations 8 and 9 into Equation 5, gives:


dp
p

 JM 2 1  J  1 M 2 4 fdx
DH
2 1  M 2

u
, thus
c

To get an expression for the Mach number, we note M

dM
M
and since c 2

JRT ,

dc
c

10.

du dc

u
c

1 dT
2 T

Hence,

dM
M

du 1 dT

u 2 T

And substituting Equations 7 and 9 into the above, gives:

dM
M

J 1

M 2
4 fdx
2
DH
2 1  M 2

JM 2 1 

11.

For the variation of the stagnation pressure along the pipe, we note that the definition of
the stagnation pressure is the pressure one would obtain if we decelerate a flow
isentropically to zero velocity. Thus,
J

po
p

J  1 2 J 1

M
1

and differentiating, we get:


dpo
po

dp

p

1

JM 2
J 1
2

M2

dM
M

The variation of p and M along the pipe is given by Equations 10 and 11, thus, we get:
dpo
po

JM 2 4 fdx
2

12.

Dh

The variation of entropy along the pipe can be found from:

Tds = dh-vdp

with h

cpT

JRT
, pv=RT, the above becomes:
J 1
dT J  1 dp

T
J p

ds
cp

dT
dp
and
, thus we obtain for the change in
T
p
entropy along the pipe the following expression:

Equations 9 and 10 give the variation of

ds
cp

ds
R

or

since

c p J  1

J  1 JM 2 4 fdx

J 2 DH

JM 2 4 f
2

DH

dx

 dpo
po

13.

R . As in the case of the normal shock, the entropy increase can be

correlated to the stagnation pressure loss due to viscous dissipation.


Equations 7-13 give the variation of u, U , T, P, M, po and s along the pipe
dpo
ds
and
, all the equations contain the term (1-M2) in the
respectively. Except for
R
po
denominator. Thus, depending on whether the flow is subsonic M<1, and (1-M2)>0 or
supersonic, M>1 and (1-M2) <0, the various thermodynamic states can either increase or
7

decrease due to friction. The table below summarizes the effect of friction on the fluid
and thermodynamic states. The sign + or denotes whether the variable increases or
decreases along the pipe (i.e. increased value of x or positive dx).

Variable
Velocity
Mach Number
Pressure
Temperature
Density
Stagnation pressure
Entropy

Subsonic

Supersonic

+
+
+

+
+
+
+

u
M
p
T

po
s

The difference between subsonic and supersonic flow is a result of the


competition between the rate of expansion due to viscous heating of the flow and the rate
of mass convected through the cross section of the pipe. Thus, for subsonic flow, the rate
of expansion (characterized by the sound speed) dominates, and the density decreases
resulting in an increase in the flow velocity, a decrease in temperature and pressure due to
the expansion. The reverse happens for supersonic flow when the rate of convection
(characterized by the flow velocity) dominates over the rate of expansion. However, for
the stagnation pressure and entropy, both subsonic and supersonic flows indicate the
same behavior of loss in stagnation pressure and increase in entropy from irreversible
viscous dissipation.

Integration of the equations


To get the variation of the various state variables along the duct, the differential
equations have to be integrated. The important one to start is the variation of the Mach
number with distance given by Equation 11. Rewriting it in the following form:
4 fdx
DH

1  M 2 dM 2

JM 4 1 

J 1
2

M 2

we can integrate the above using the method of partial fraction. The above can be
expanded as:

J 1
dM 2
dM 2 J  1 dM 2 J  1
2


JM 4
2J M 2
2J J  1 2
1 
M

4 fdx
DH

Since friction always drives the flow towards M=1, we integrate the above from
0<x<L* and 0<M2<1 (i.e. we assume the Mach number at x=0 is M and at x=L*, M=1. In
general, the friction factor is a function of the Reynolds number and varies along the duct
as the fluid properties change. However, we shall assume an average value for the
friction factor from 0<x<L*. We then obtain:
J 1

4 fL*
DH

J  1 2 2J
2
M

1 M
2


ln
JM 2
1  J  1 M 2


2

14.

The above equation says that if we have a duct where at x=0, the Mach number is
M, then at x=L*, M=1 and the flow is choked. Equation 14 is computed for a range of
Mach numbers and tabulated (Fanno flow tables). So, if we have a pipe of length L and
inlet Mach number M1, and if we want to know the Mach number at L, we do the
following:

M2

M1

L*M2

L
LM1

Referring to the sketch above:


L

L* M1  L* M 2

or

4 fL* M1 4 fL* M 2

DH
DH

4 fL
DH

4 fL* M1
We know L and M1, we can look up the table for L such that
corresponds
DH
to M 1 . Then, we can find the value for:
*
1

4 fL* M 2
DH

4 fL* M 2 4 fL


DH
DH

We can then find the value of M2 corresponding to the value of

4 fL* M 2
.
DH

Equation 14 gives M as a function of L*. If we want to know the other flow variables
dp dT
like p, T, U , etc. we can integrate the other equations for
,
, etc. However, it is
p T
more convenient to express them in terms of the Mach number so that once M at a
section is known, we can find the other variables. From Equations 10 and 11, we write:

 1  J  1 M 2 JM 2 4 fdx

DH
2 1  M 2

dp
p
dM
M

JM 2

J  1 2
M
1
4 fdx

2
2

DH
1  M 2

and simplifying the above yields:

dp
p

 1  J  1 M 2 dM 2
J  1 2 M 2

M
2 1 

Using partial fraction, the above can be written as:

dp
p

J 1
dM 2
dM 2
2


2M 2
J 1 2
M
2 1 
2

10

which integrates to yield:


p
p*

1
M

2 1 

J 1
J 1
2

15.

The above equation gives the pressure p at M and p = p* when M=1. Similarly, we
can find expressions for the other variables in terms of the Mach number as follows:

T
T*

u
u*

c
*
c

2 1 

J 1
J 1
2

J  1

17.

J  1 2

2 1 
M

J  1 2

2 1 
M

2
J 1

U
U*

u
u

po
po *

J  1 2 2 J 1

2 1 
M

1
2

M
J  1

16.

1
M

18.

J 1

s  s*
cp

J  1
2

ln M
J  1 2

2
2 M 1  2 M

11

19.

J 1
2J

20.

Equations 15 to 20 are also computed for various Mach numbers and tabulated together
4 fL*
in the Fanno tables for convenience in solving problems of compressible flow
with
DH
in ducts with friction.

The Fanno Line


The adiabatic flow of a compressible fluid in a constant area duct with friction
satisfies the conservation of mass and energy i.e.

Uu constant
dU

U
h

u2
2

du
u

ho

constant

dh  udu

From these two equations, we can obtain an expression for the variation of the enthalpy
(or equivalently the temperature) and the entropy along the duct. From the Tds equation,
i.e.:

dh  vdp

Tds
we get:

dh J  1 dp

h
J p

ds
cp

and from the equation of state p


dp
p

URT , we get:
dU

dT
T

Hence,

12

dU

dh
h

dh J  1 dU dh



J U
h
h

ds
cp

1 dh J  1 dU

J h
J U

From the conservation of mass, we write:


dU

du
u

du 2
2u 2

From the energy equation, we note that:


u2
2
and

du 2
2

 dh

d ho  h since ho

h o h

constant.

Thus, we may write:


dU

d ho  h
2 ho  h

and the equation for the entropy becomes:


1 dh J  1 d ho  h

J h
2J ho  h

ds
cp

We can integrate the above as following:

ds
s* c
p

dh J  1 ho  h d ho  h

h* h
2J ho  h* ho  h

and obtain:
s  s*
cp

Since ho

J 1

ln

h J  1 ho  h

ln
*
2J
h*
ho  H

h* , we can eliminate ho in the above equation and obtain the equation for

2
the Fanno Line as:

13

s  s*
cp

J 1
J 1
1

J 2 2J J  1
2J
h
h


ln *
 *
h J  1 2
h

21.

The Fanno line gives the variation of the enthalpy as a function of the entropy for
a given mass flow rate in and stagnation enthalpy ho (or equivalently h*). The lower
branch is for supersonic flow and the upper branch is for subsonic flow. For any given
value of M, friction will drive the flow towards sonic conditions M=1 where the entropy
is a maximum. To prove that entropy is a maximum when the flow becomes sonic, one
ds
could differentiate Equation 20 and equate
0 and solve for M. Alternately, we do
dM
the following: from the Tds equation, we write:

dh 

Tds
and from the energy equation dh

dp

udu, we get:
udu 

Tds
From the conservation of mass Uu

dp

constant , we get :

Udu  udU

If we express U s, p , we can write:

wU
wU
dU s, p ds  dp
ws p
wp s

1
wU
ds  2 dp
c
ws p

since the sound speed c is defined by:

c2

wp

wU s

Replacing dU in the continuity equation by the above expression yields :

Udu  udU

Udu 

u
wU
dp  u ds
2
c
ws p

14

22.

From the Tds equation, we get:

 Uudu  UTds

dp

and replacing dp in Equation 22 by the above and rearranging gives:

UT

wU

ws p
c

Udu 1  M 2  uds

23.

wU
is always positive and is always negative (since density decreases
c
ws p
when heat is added at constant pressure and heat added means entropy increases, i.e.
wU  0 as ds>0 at constant p). The bracketed term is always positive, i.e.:
Since UT

pT wU
 ! 0
c 2 ws p
Thus, as M o 1 , ds must vanish, thus s corresponds to an extremum (maximum) value
when M=1.
Another alternate expression for the Fanno Line that involves the Mach number can be
obtained as follows: From the Tds equation, we write:

ds

dh Rdp

T
p

and rearranging, we write:

dh
ds

1
1 R dp

T p dh

From the energy equation, we get:


dh  du

And the conservation of mass gives:

dU

du
u

15

1 du 2
2
u2

we can express dh as:

 du

dh
And using the equation of state p

u2

dU

URT , we can express:


dU

dp dT

p
T

Hence,

dh u 2

And solving for the slope

p
1  J  1 M 2
2
u

dh
as:
ds
dh
ds

or

dp dT

u 2 
T
p

dp
, we get:
dh
dp
dh

We can now write

dU

dh
ds

1
p
1 R p
 2  J  1 M 2 2
u
T pu

JM 2T
M 2 1

The above is an alternate form of the equation for the Fanno Line and it is clear that for
dh
subsonic flow where M<1,
 0 which corresponds to the upper branch of the Fanno
ds
dh
! 0 , the locus of states is represented by the lower branch of the
Line. For M>1,
ds
dh
Fanno Line. When M=1,
o f and s o max .
ds
To plot a Fanno Line, we first note that the Fanno Line involves the conservation of mass
and energy, i.e.

16

Uu

m

u2
h
2

constant

ho

constant

Thus, a Fanno Line corresponds to a value for m and ho . Combining conservation of


energy and mass, we get:
h

Where v

m 2v 2
2

ho

constant

24.

is the specific value. So, given m and ho , we pick different values for U

and compute h from Equation 24. And Equation 21 gives s as a function of h. Note
2
ho , thus the value of h* can be found for a given stagnation enthalpy ho.
that h*
J 1

Effect of duct length and back pressure

The flow inside a duct with friction is influenced by the duct length and the back
pressure of the exit of the duct. Consider first the case of subsonic flow in the duct. We
assume the inlet to the duct is from a converging nozzle connected to a large plenum
chamber where the stagnation pressure and temperature is po and To. Consider first
just the nozzle. For a subsonic jet, the exit pressure must be the same as the back
pressure. Thus, for a given exit pressure, the Mach number at the jet exit is given by the
isentropic relationship
J

po
p

J  1 2 J 1
M
1 
2

If we now connect a pipe to the exit of the converging nozzle, then the exit of the nozzle
corresponds to the entrance to the duct. Let the end of the duct be connected to a large
plenum chamber where the pressure can be controlled. The system is shown
schematically in the sketch below.

17

stagnation
conditions

x Pe

plenum
chamber
PB

Po, To

Pe
a

x Pe = PB

x Pe = PB

x Pe = P* = PB
x Pe = P*
PB < P*
ho

Me

x Me
Me = 1 x

a
b

c

decreasing m

For curve a, the flow is subsonic throughout the duct, the exit pressure pe=pB . The
4 fL
4 fL*
corresponding to the inlet Mach number M1.
value of
is less than the value
DH
DH

When we lower the back pressure pB, the flow is similar but the inlet Mach number to the
duct (hence the mass flow rate) is increased. When the back pressure is reduced to a
value where the flow is choked at the exit, then pe=p*=pB and Me=1. For this case, the
18

duct length corresponds to the critical value L* for the inlet Mach number (curve c). If
the back pressure is further decreased, i.e. pB<p*, no change occurs within the duct and
further expansion occurs downstream of the duct exit to permit the jet pressure to adjust
to the ambient pressure. The three cases of a b and c are also illustrated by the
4 fL 4 fL*
) if
Fanno Lines on the h-s diagram. After choking occurs, (i.e. curve c,
DH
Dh
the length of the pipe is increased beyond this critical length, then the inlet Mach number
will have to decrease to permit choking to occur in the longer length of pipe. Thus, for
subsonic flow, the inlet conditions to the pipe can adjust to the different values of the
back pressure (or exit pressure) and the length of the duct.
For supersonic flow, the inlet of the duct must be connected to the stagnation tank
A
via a converging diverging nozzle. Once the area ratio 1* of the nozzle is fixed, then
A
the exit Mach number, pressure, flow rate, etc. are all determined from the isentropic
flow relationships (look up the tables). Mach number at the duct inlet is supersonic and
conditions downstream in the duct cannot influence the inlet condition unless a shock
wave is driven back into the nozzle. Consider first a duct length less the critical value for
the given inlet Mach number M1. For a sub critical duct length, the exit Mach number is
still supersonic and the flow in the duct is supersonic throughout. This is the case
provided the exit pressure of the duct is greater than the back pressure pB in the plenum
*
chamber, i.e. pe<p
since the flow is not choked at the exit and pe<p
>
> B . The under
expanded jet will adjust to ambient pressure via expansion waves. When the back
pressure is increased to the exit pressure, i.e. pe=pB, then the jet exit pressure matches the
ambient pressure and the jet is parallel. When pB>pe, but not significantly greater, then
conditions inside the duct remains unchanged and adjustment to ambient pressure occurs
downstream of the jet exit via oblique shock waves. When pB increased further, the
oblique shock becomes stronger by increasing its angle to the flow. The limiting
condition will be when a normal shock occurs just at the duct exit so that the pressure
downstream of the normal shock matches the back pressure pB. When pB increases
beyond this limiting value, the shock moves into the duct and becomes stronger since the
Mach number is higher. Downstream of the shock is subsonic, thus the exit pressure will
have to be equaled to the ambient pressure.
The location of the shock is such that this condition is satisfied. The above
discussion is for the case when the duct length is less than the critical value
corresponding to the inlet Mach number.
If the duct length is the critical value, then pe=p* and Me=1. Again, when
*
pe=p >pB, the under expanded sonic jet will undergo further expansion downstream of the
duct exit to adjust to the ambient pressure. If pB=p*, the same kind of events as described
earlier for the case of L<L* occurs. For high pB, a normal shock will eventually be
formed inside the duct. Once a normal shock occurs inside the duct, the flow is subsonic
downstream of the duct and for a subsonic flow; the exit pressure must match the ambient
pressure. Similarly, if the duct length L>L* for the inlet Mach number, a normal shock
will be formed inside the duct. The position of the shock is such that the exit pressure
matches the ambient pressure. If the duct is very long, then the maximum Mach number
the subsonic flow downstream of the shock inside the duct can accelerate to is M=1.
Thus, the flow is chocked at the exit, pe=p*. If pB=p*, further adjustment occurs
19

downstream of the duct exit. For pe=p*, there is only one location where the shock inside
the duct is located.
For supersonic inlet flow and a very high pB so that a shock is driven inside the
duct, the process is illustrated in the h-s diagram below.

ho

M< 1

isentropic flow
in nozzle

20

x Me , Pe = PB

normal shock
transition
M> 1

Anda mungkin juga menyukai