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

Duct Flow with Heat Transfer

In case of one dimensional flow three factors that can change flow
properties continuously are (i) changes in area (ii) wall friction (iii) heat
exchange.

The effects of area and friction have been discussed previously.

The effect of heat transfer in which stagnation enthalpy or temperature


changes is now studied.

Compressible flow problems often encountered in practice involve


chemical reactions such as combustion, nuclear reactions as well as heat
gain or heat loss through the duct wall.

Many of these problems are difficult to analyze exactly since they may
involve significant changes in chemical composition during flow, and the
conversion of chemical or nuclear energies to thermal energy.

The essential features of such complex flows can still be captured by


a simple analysis by modeling the generation or absorption of thermal
energy as heat transfer through the duct wall.
The heat transfer can be assumed at the same rate and any change in
chemical composition is neglected.

This simplified problem is still too complicated for an elementary


treatment of the topic since the flow may involve friction, variations in duct
area, and multidimensional effects.

However, when heating or cooling is achieved by external heat exchange


the temperature difference between duct wall and fluid can be very large
and frictional effects can be neglected.

In such cases flow can be modeled as simple one-dimensional flow in a


duct of constant cross-sectional area with negligible frictional effects.

Steady, one-dimensional flow of an ideal gas with constant specific heats


through a constant-area duct with heat transfer, but with negligible friction
is referred to as Rayleigh flow.

The conservation of mass, momentum, and energy equations for the


control volume shown in Figure are written as follows:

𝑚̇ = 𝜌1 𝐴1 𝑉1 = 𝜌2 𝐴2 𝑉2

The momentum equation is:


𝑝1 𝐴1 − 𝑝2 𝐴2 = 𝜌2 𝐴2 𝑉22 − 𝜌1 𝐴1 𝑉12

For constant area

𝜌1 𝑉1 = 𝜌2 𝑉2 𝑝1 + 𝜌𝑉12 = 𝑝2 + 𝜌𝑉22

The energy equation is:

𝑉22 𝑉12
𝑄̇ = 𝑚̇(ℎ02 − ℎ01 ) = 𝑚̇ (ℎ2 + − ℎ1 − )
2 2

The ideal gas, Tds and entropy relations are given as

𝑑𝑝 𝑇2 𝑝2
𝑝 = 𝜌𝑅𝑇 𝑇𝑑𝑠 = 𝑑ℎ − 𝑠2 − 𝑠1 = 𝐶𝑝 𝑙𝑛 ( ) − 𝑅𝑙𝑛 ( )
𝜌 𝑇1 𝑝1

Consider a gas with known properties R, γ, and Cp. For a specified inlet
state 1, the inlet properties P1, T1 and V1 etc. are known.

The exit properties can be determined from three governing equations for
any specified value of heat transfer Q.
There are, however, infinite number of possible downstream states 2
corresponding to a given upstream state 1.

Plotting the results on a T-s or h-s diagram gives a curve passing through
the specified inlet state.

The plot of Rayleigh flow on a T-s diagram is called the Rayleigh line.

In other words, it is a locus of points on T-s / h-s diagram passing through


the given initial point. The locus is obtained from continuity and
momentum equations (along with equation of state and entropy relations).

For a given initial state, the fluid cannot exist at any downstream state
outside the Rayleigh line on a T-s diagram since it is the locus of all
physically attainable downstream states corresponding to an initial state.

On the line impulse function is same but (usually) with different value of
stagnation enthalpy.

In Rayleigh line points A and B are on upper and lower branches and
have same entropy.

The kinetic energy of B is greater (since static temperature / enthalpy is


less) than that at state A. Therefore upper branch represents subsonic
flow and lower branch is supersonic flow.

Extrema of Rayleigh line

The differential forms of equations are:

𝑑𝜌 𝑑𝑉
𝜌𝑉 = constant → =− (1)
𝜌 𝑉

𝑑𝑝
𝑃 + (𝜌𝑉 )𝑉 = constant → = −𝜌𝑉 (2)
𝑑𝑉
𝑑𝑝 𝑑𝜌 𝑑𝑇
= + (3)
𝑝 𝜌 𝑇

𝑑𝑇 𝑑𝑝
𝑑𝑠 = 𝐶𝑝 −𝑅 (4)
𝑇 𝑝

From (3) and (4)

𝑑𝑇 𝑑𝜌 𝑑𝑇
𝑑𝑠 = 𝐶𝑝 −𝑅( + )
𝑇 𝜌 𝑇

𝑅 𝑑𝑇 𝑑𝜌
𝑑𝑠 = −𝑅( ) (5)
𝛾−1 𝑇 𝜌

Dividing both sides of (5) by dT and combining with (1)

𝑑𝑠 𝑅 𝑅 𝑑𝑉
= − ( ) (6)
𝑑𝑇 𝑇(𝛾 − 1) 𝑉 𝑑𝑇

Dividing (3) by dV and combining it with (1) and (2) after rearranging

𝑑𝑉 𝑇 𝑉
= − (7)
𝑑𝑇 𝑉 𝑅
Substituting Equation 7 into Equation 6 and rearranging

𝑑𝑠 𝑅 (𝛾𝑅𝑇 − 𝑉 2 )
=
𝑑𝑇 𝑇(𝛾 − 1)(𝑅𝑇 − 𝑉 2 )

Setting ds/dT = 0 for maximum entropy gives

𝑉 = √𝛾𝑅𝑇 or 𝑀 = 1 (8)

Setting dT /ds = (ds /dT)-1 = 0 for maximum entropy

1
𝑉 = √𝑅𝑇 or 𝑀 = (9)
√𝛾

This can be proved using the differential forms of the energy equations

𝑉2
𝛿𝑄 = 𝑑ℎ0 = 𝑑 (ℎ + ) = 𝐶𝑝 𝑑𝑇 + 𝑉𝑑𝑉 (10)
2

𝛿𝑄 𝑑𝑇 𝑉𝑑𝑉 𝑑𝑉 𝑉 𝑑𝑇 (𝛾 − 1)𝑉 2
= + = ( + ) (11)
𝐶𝑝 𝑇 𝑇 𝐶𝑝 𝑇 𝑉 𝑑𝑉 𝑇 𝛾𝑅𝑇

Solving (7) and (11)

𝛿𝑄 𝑑𝑉 𝑉 𝑇 𝑉 2
𝑑𝑉 𝑉2
= ( ( − ) + (𝛾 − 1)𝑀 ) = (1 − + 𝛾𝑀2 − 𝑀2 ) (12)
𝐶𝑝 𝑇 𝑉 𝑇 𝑉 𝑅 𝑉 𝑇𝑅

𝑑𝑉 𝛿𝑄 1
= (13)
𝑉 𝐶𝑝 𝑇 (1 − 𝑀2 )

Several important observations can be made from Rayleigh plot and the
above results:

Entropy increases with heat gain

Thus we proceed to the right on the Rayleigh line as heat is transferred to


the fluid.
Similarly, a process proceeds to the left with heat rejection regardless of
the initial value of the Mach number.

At M =1 entropy is maximum

M = 1 at point ‘a’ (Eq. 8), which is the point of maximum entropy.

Upper branch is subsonic flow since M =1 at point of maximum entropy


and static temperature is higher on the upper branch.

Similarly flow is supersonic on the lower branch.

Heating increases the stagnation temperature

It is clear from the energy balance Q = Cp(T02 - T01) that T0 increases for
both subsonic and supersonic flows, and cooling decreases it.

Heating also increases the static temperature T except for the narrow
1
Mach number range of < M < 1 in subsonic flow (Eq. 9).
√𝛾

Both temperature and Mach number increase with heating in subsonic


1
flow, but T reaches a maximum Tmax at 𝑀 = (which is 0.845 for air)
√𝛾

and then decreases.

It may seem peculiar that the temperature of a fluid drops as heat is


transferred to it.

Heating initially increases static temperature (internal energy) and kinetic


energy.

It can be assumed that, however, after particular amount of addition of


external heat, it becomes impossible to increase both internal (hence
temperature) and kinetic energy (hence velocity).
As a result beyond a certain limit temperature decreases but velocity
(hence KE) continues to increase.

1
The cooling effect in < M < 1 is thus due to the large increase in the
√𝛾

fluid velocity and the accompanying drop in temperature in accordance


with the relation T0 = T + V2/2Cp.

Flow velocity increases with heat addition in subsonic flow and that the
opposite occurs in supersonic flow (Eq. 13).

In subsonic flow, i.e. M<1, heat transfer and velocity change have the
same sign. As a result, heating the fluid (dQ > 0) increases the flow
velocity while cooling decreases it.

In supersonic flow, however M>1 and heat transfer and velocity change
have opposite signs.
In supersonic flow the fluid has sufficient kinetic energy and significantly
less static temperature. The addition of heat results in rapid increase of
temperature but decrease in velocity and Mach number.

The momentum equation P + KV = constant, where K = ρV = constant


(from the continuity equation) reveals that velocity and static pressure
have opposite trends.

The relations for variation of properties like pressure, temperature etc.


can be written in terms of Mach numbers.

From momentum equation:

𝑝1 + 𝜌𝑉12 = 𝑝2 + 𝜌𝑉22

Replacing velocity with Mach no and using ideal gas relation we get:

𝑉 = 𝑀√𝛾𝑅𝑇 𝑝 = 𝜌𝑅𝑇

𝑝2 1 + 𝛾𝑀12
= (A)
𝑝1 1 + 𝛾𝑀22

From continuity and ideal gas equations

𝜌2 𝑉1 𝑀1 √𝛾𝑅𝑇1 𝑀1 √𝑇1
= = = (B)
𝜌1 𝑉2 𝑀2 √𝛾𝑅𝑇2 𝑀2 √𝑇2

𝑇2 𝑝2 𝜌1
= (C)
𝑇1 𝑝1 𝜌2

Substituting pressure ratio from equation (A) and density ratio from (B) in
(C)

𝑇2 1 + 𝛾𝑀12 𝑀1 √𝑇1
=( )( )
𝑇1 1 + 𝛾𝑀22 𝑀2 √𝑇2
2
𝑇2 𝑀2 (1 + 𝛾𝑀12 )
=( )
𝑇1 𝑀1 (1 + 𝛾𝑀22 )

𝜌2 𝑉1 𝑀12 (1 + 𝛾𝑀22 )
= =
𝜌1 𝑉2 𝑀22 (1 + 𝛾𝑀12 )

Flow properties at sonic conditions are usually easy to determine, and


thus the critical state corresponding to M = 1 serves as a convenient
reference point in compressible flow.

Taking state 2 to be the sonic state

𝑝 1+𝛾
=
𝑝∗ 1 + 𝛾𝑀2

2
𝑇 𝑀(1 + 𝛾 )
= ( )
𝑇∗ 1 + 𝛾𝑀2

𝑉 𝜌 (1 + 𝛾 )𝑀2
= =
𝑉 ∗ 𝜌∗ 1 + 𝛾𝑀2

2
𝑇0 𝑇0 𝑇 𝑇 ∗ 𝛾 − 1 2 𝑀(1 + 𝛾 ) 𝛾 − 1 −1
= = (1 + 𝑀 )( ) (1 + )
𝑇0∗ 𝑇 𝑇 ∗ 𝑇0∗ 2 1 + 𝛾𝑀2 2

𝑇0 (𝛾 + 1)𝑀2 [2 + (𝛾 − 1)𝑀2 ]
=
𝑇0∗ (1 + 𝛾𝑀2 )2

Similarly

𝛾⁄𝛾−1
𝑃0 (𝛾 + 1) 2 + (𝛾 − 1)𝑀2
= ( )
𝑃0∗ (1 + 𝛾𝑀2 ) 𝛾+1
Effect of heating / cooling on stagnation enthalpy

Figure shows the Rayleigh line (which represents the locus of static
states) together with the corresponding stagnation reference lines.

Remember that for a perfect gas this h–s diagram is equivalent to a T–s
diagram.

There are two stagnation curves, one for subsonic flow and the other for
supersonic flow. It can be shown that the supersonic stagnation curve is
the top one.

The differential form of the energy equation is:

From definition of entropy

From above two equations


or

Note that the equation gives the slope of the stagnation curve in terms of
the static temperature.

Now draw a constant-entropy line on Figure. This line will cross the
subsonic branch of the (static) Rayleigh line at a higher temperature than
where it crosses the supersonic branch.

Consequently, the slope of the subsonic stagnation reference curve


will be greater than that of the supersonic stagnation curve.

Since both stagnation curves must come together at the point of


maximum entropy, this means that the supersonic stagnation curve is a
separate curve lying above the subsonic one.

Correlation with Shocks

Similarities exist between Rayleigh flow and normal shocks.

The end points before and after a normal shock represent states with the
same mass flow per unit area, the same impulse function, and the same
stagnation enthalpy.
A Rayleigh line represents states with the same mass flow per unit area
and the same impulse function. All points on a Rayleigh line however do
not have the same stagnation enthalpy.

From figure it can be noticed that for every point on the supersonic
branch of the Rayleigh line, there is a corresponding point on the
subsonic branch with the same stagnation enthalpy.
Thus these two points satisfy all three conditions for the end
points of a normal shock and could be connected by such a shock

The shock merely jumps the flow from the supersonic branch to the
subsonic branch of the same Rayleigh line.

This also provides another reason why the supersonic stagnation curve
must lie above the subsonic stagnation curve.

If this were not so, a shock would exhibit a decrease in entropy, which is
not correct.

Thermal Choking

In Fanno flow, recall that once sufficient duct was added, or the receiver
pressure was lowered far enough, we reached a Mach number of unity at
the end of the duct.

Further reduction of the receiver pressure could not affect conditions in


the flow system.
The addition of any more duct caused the flow to move along a new
Fanno line at a reduced flow rate.

Subsonic Rayleigh flow is quite similar. Figure shows a given duct fed by
a large tank and converging nozzle. Once sufficient heat has been added,
M = 1 at the end of the duct.

The T–s diagram for this is shown as path 1–2–3. This is called thermal
choking.

It is assumed that the receiver pressure is at p3 or below. Reduction of


the receiver pressure below p3 would not affect the flow conditions
inside the system.
However, the addition of more heat will change these conditions.
Now suppose that we add more heat to the system.

This would probably be done by increasing the heat transfer rate through
the walls of the original duct.

However, it is more convenient to indicate the additional heat transfer at


the original rate in an extra piece of duct, as shown in Figure.

The only way that the system can reflect the required additional entropy
change is to move to a new Rayleigh line at a decreased flow rate (due to
high pressure ratio).

This is shown as path 1–2’–3’– 4 on the T–s diagram.

Whether or not the exit velocity remains sonic depends on how much
extra heat is added and on the receiver pressure imposed on the system.

The important thing to remember is that once a subsonic flow is thermally


choked, the addition of more heat causes the flow rate to decrease.

Just how much it decreases and whether or not the exit remains sonic
depends on the pressure that exists after the exit.

The parallel between choked Rayleigh and Fanno flow does not quite
extend into the supersonic regime. In Fanno flow extension in length of
duct in case of supersonic flow results in shock.

This does not happen in supersonic flows.

Figure shows a M = 3.53 flow that hasTt/Tt* = 0.6139.

For a given total temperature at this section, the value of Tt/Tt* is a direct
indication of the amount of heat that can be added to the choke point.
If a normal shock were to occur at this point, the Mach number after the
shock would be 0.450, which also has Tt/Tt * = 0.6139.

Thus the heat added after the shock is exactly the same as it would be
without the shock.

The situation above is not surprising since heat transfer is a function of


stagnation temperature, and this does not change across a shock.

The shock may occur at some location preceding the Rayleigh flow such
as in a converging–diverging nozzle which produces the supersonic flow

You might also like