Vous êtes sur la page 1sur 20

Chapter 2

Energy and the Laws of Thermodynamics


THE FIRST LAW OF THERMODYNAMICS
This law, also known as the law of conservation of energy, is one of the most fundamental laws in
nature and states that “energy can neither be created or destroyed during a process; it can only be
converted from one form to another”.

Consider 1 kg of working fluid which may be a liquid, gas or vapour, or any combination, and which is
flowing steadily through a control volume (Figure 1). The fluid may possess energy in a number of
forms between which conversions are possible:
1. Flow work or pressure work, given by pv
2. Kinetic energy C2/2 due to the movement of the fluid as a whole
3. Internal (thermal ) energy u, due to the energy of the fluid molecules
4. Potential or gravimetric energy, due to the height Z above some datum line and given by Zg
5. Chemical, electrical or magnetic energies may also be added, but these are not involved in the
overwhelming cases encountered in thermal power cycles
6. Heat Q may enter or leave the control volume
7. Mechanical energy (W) may be added or removed with some of the added energy being used to
pump the fluid into the control volume or drive it out again.
8. Accumulated (stored) energy in the control volume as a whole ecv

Control Volume

Figure 1 Steady-flow control volume

From the law of conservation of energy, the net energy input should be equal to the net energy output
plus the energy that accumulates in the control volume. For properties per unit mass (specific properties
denoted by lower case letters) and consistent energy units (J) for all terms we can write
 2   2 
 p1v1 + C1 + u1 + gZ 1  + q1 + w1 = ecv +  p2 v2 + C 2 + u 2 + gZ 2  + q2 + w2 (1)
 2   2 
   

1 JGhojel 2008
Energy is not usually allowed to accumulate in the control volume of practical thermal power plants
operating on thermodynamic cycles, and therefore the term ecv will be henceforth ignored and
equation 1 is reduced to
 2   2 
 p1v1 + C1 + u1 + gZ 1  + q1 + w1 =  p 2 v 2 + C 2 + u 2 + gZ 2  + q2 + w2 (2)
 2   2 
   

Since specific enthalpy h = u + pv , equation (2) can be written as

 2   2 
 h1 + C1 + gZ 1  + q1 + w1 =  h2 + C 2 + gZ 2  + q 2 + w2 (3)
 2   2 
   

For a fluid flowing steadily at the rate of m


& kg/s, the energy equation becomes
 C12   C 22 
m&  p1v1 + + u1 + gZ 1  + Q1 + W1 = m&  p 2 v2 + + u 2 + gZ 2  + Q& 2 + W& 2 (4a)
 2   2 
 C 2   C 2 
m&  h1 + 1 + gZ 1  + Q& 1 + W&1 = m&  h2 + 2 + gZ 2  + Q& 2 + W& 2 (4b)
 2   2 

with the terms acquiring the units of power (Nm/s, J/s or W).

For a control volume with multiple flows into and out of the system, the general steady-flow energy
equation can be written as
 C12   C2 
&  h1 +
∑m + gZ 1  + Q& 1 + W&1 = ∑ m&  h2 + 2 + gZ 2  + Q& 2 + W& 2 (5)
 2 2
   

For each inflow For each outflow

Certain specific cases can be obtained from equation (3):

Boiler feed-pump: The working fluid is liquid and the difference in the potential energy terms (gZ2 -
gZ1) may be large compared with the other terms (Figure 3). The work done in raising boiler feed-water
from the level of Z1 of the exit from the condenser to the height Z2 of the point of entry to the boiler
may be an appreciable part of the load on the boiler feed-pump.

Figure 3 Control volume of a pump without heat transfer to or from the liquid

2 JGhojel 2008
W&1 = m& (h2 − h1 ) + m& ( gZ 2 − gZ 1 ) (6)
If (gZ2 - gZ1) is negligible, then
W&1 = m& ( h2 − h1 ) (7)
For an isentropic process, ( h2 − h1 ) = v( p 2 − p1 ) , and
W&1 = m& ( h2 − h1 ) = m& v( p 2 − p1 )
Or, since h = u + pv ,
W&1 = m& ( p 2 v2 − p1v1 ) + m& ( u 2 − u1 )

If heat is added to the system during the pumping process, the energy equation is then written as
Q& 1 + W&1 = m& ( h2 − h1 )
or
Q& 1 + W&1 = m& ( p 2 v2 − p1v1 ) + m& ( u 2 − u1 )

If the working fluid is gas the Z term or the difference (gZ2 - gZ1) is often negligible as in the case of
the gas turbine, although it is of importance when considering chimney draught.

Throttling valve: throttling valves are flow restricting devices that cause a significant drop of pressure
in the fluid without producing work. These could be in the form of a restricting valve , a porous plug or
a capillary tube (Figure 4)

Figure 4. Types of throttling valves

The kinetic and potential energy changes and heat transfer to and from the valve (control volume) are
negligible reducing equation (4) to
h1 = h2
h = const
u + p v =u + p v (8)
1 1 1 2 2 2
u + pv = const

For a perfect gas h = cpT so that T1 = T2. In the case of an imperfect gas, the flow energy pv may
increase during the throttling process at the expense of the internal energy (u decreases) causing a drop
in temperature (Joule-Thompson cooling effect).

3 JGhojel 2008
Simple nozzle or diffuser: A nozzle is a device that increases the velocity of a gas or vapour at the
expense of pressure. A diffuser is a device that increases the pressure of a gas or vapour at the expense
of velocity.

Figure 5 Schematic diagrams of a nozzle and diffuser

The input kinetic energy, potential energy change and heat and work transfer changes are usually
negligible and the energy equation becomes
C 22
h1 = h2 + (9)
2

If the input kinetic energy is of considerable value (the gas is approaching the nozzle at a high velocity)
the energy equation for the nozzle is
C12 C2
h1 + = h2 + 2 (10)
2 2

Reciprocating internal combustion engine: The gases flow into and out of the engine slowly and the
change in kinetic energy can be neglected. Also, the potential energy change (gZ2 - gZ1) is very small
and can be ignored. The energy equation now becomes
Q& 1 = m& (h2 − h1 ) + Q& 2 + W& 2 − W&1

Figure 6 is an illustration of the representation of a control volume for a car engine. In this example, no
work is crossing the boundary into the control volume or heat crossing the boundary into or out of the
control volume.

Figure 6 Control-volume representation of a car engine

4 JGhojel 2008
The energy equation is often used for the combustion and expansion (power) stroke in order to assess
the process of heat release by the burning fuel. As both the inlet and exhaust valves are closed during
this process, the process is a closed system with no added mechanical work and negligible potential
energy, flow work and kinetic energy. The energy equation in (2) is now reduced to

u1 + q1 = u 2 + q 2 + w2 , or
q1 = (u 2 − u1 ) + q 2 + w2

Both the volume and pressure change as the gases expand in the cylinder producing work w2 = ∫ pdv .
The energy equation then becomes

q1 = du + ∫ pdv + q 2 (11)

This convenient form of the energy equation equates the energy input as heat from the combustion of
the fuel to the sum of the change of internal energy of the gases as their temperature change, work done
by the gases as they expand in the cylinder during the power stroke and the heat losses to the
surroundings.

Gas turbine: The process is steady-state, steady-flow where the potential and kinetic energy changes
are negligible (Figure 7).

Figure 7 Schematic diagram of a turbine

Equation (4b) is then reduced to

m& h1 + Q& 1 = m& h2 + Q& 2 + W& 2 (12a)

If the expansion process in the turbine is adiabatic,

W& 2 = m& ( h1 − h2 ) (12b)

5 JGhojel 2008
Refrigeration system: A typical vapour compression refrigeration system is shown in Figure 8.

Figure 8 Schematic diagram of a vapour compression refrigeration system

The potential and kinetic energy changes as well as the output mechanical work are assumed
negligible. The energy equation in (3) is reduced to
h1 − h2 + q1 + w1 = q 2 (13)

h1 - h2 represents the change of state of the refrigerant, q1 is the heat pumped out of the cold chamber,
w1 is the work input (usually work to run a compressor) and q2 is the heat absorbed by the cooler or
condenser. If the working fluid does not undergo a state change within the cycle,
q1 + w1 = q 2 (14)

High-speed compressor: The potential energy change, input heat, inlet velocity and output mechanical
work are negligible in the case of the compressor shown in Figure 9.

Figure 9 Schematic diagram of a air compressor

Equation (4b) is now reduced to


2
C
m& (h1 − h2 ) + W&1 = m& 2 + Q& 2 (15)
2
or
enthalpy change + work input = kinetic energy of discharge + cooling losses
6 JGhojel 2008
If the velocity of the gas is reduced in a volute chamber at the exit so that the kinetic energy of
discharge is negligible and there is no appreciable cooling, then

W&1 = m& (h2 − h1 ) (16)

For a perfect gas

W&1 = m& c p (T2 − T1 ) (17)

Heating gas at constant pressure or constant volume: Equation (2) is reduced to

q1 = ( h2 − h1 ) = c p (T2 − T1 ) (18)

When the gas is heated at constant volume, equation (2) is reduced to

q1 = (u 2 − u1 ) = cv (T2 − T1 ) (19)

Liquid flowing at constant temperature: There is no change in internal energy, and there is no heat
or work exchange with the surroundings, and since the fluid is liquid (assumed incompressible), the
specific volume remains constant and equation (2) is reduced to

 2   2 
 p1v + C1 + gZ 1  =  p 2 v + C 2 + gZ 2 
 2   2 
   

If we replace the specific volume v by 1 ρ , where ρ is the density, we obtain Bernoulli’s equation

 p1 C12   p 2 C 22 
 + + gZ = + + gZ  (20)
 ρ 2
1   ρ 2
2 
   

THE SECOND LAW OF THERMODYNAMICS


This law states that “it is impossible to construct a system which will operate as a cycle, extract heat
from a reservoir and do an equivalent amount of work on the surroundings.” It follows that part of the
extracted heat must be rejected to another reservoir at a lower temperature. Two cases can be
identified:

• Heat transfer will occur down a temperature gradient as a natural phenomena and part of this
heat is converted to work with the balance rejected to a low temperature reservoir (Figure 10a).
• Heat can be transferred from a low-temperature reservoir up a temperature gradient to a high-
temperature reservoir with the assistance of external work (Figure 10b).

7 JGhojel 2008
(a) (b)
Figure 10 Schematic arrangements of a heat engine (a) and heat pump (b)

The second law of thermodynamics is also stated as the Law of Degradation of Energy whereby the
quantity of energy is conserved, but its quality (the potential to produce useful work) is not. Every time
energy changes from or is transferred from one system to another, its potential to produce useful work
is reduced irreversibility for ever. It is then said that energy has degraded.

This law is the reason we may face an energy crisis. All the energy that we use ultimately ends up as
waste heat transferred to the earth’s atmosphere and then to space.

Entropy is the measure of irreversibility or degradation and is defined as


dQ dq
dS = kJ/K, or ds = kJ/kg K
T T
dS – total entropy change
ds – specific entropy change
dQ – heat transferred reversibly
T – absolute temperature at which heat is transferred.

• If heat is added to a system dQ will be positive and ds will be positive (entropy increases).
• If heat is removed from a system dQ will be negative and ds will also be negative (entropy
decreases)
• If ds = 0 during a process the process is isentropic. The frictionless adiabatic process is an
isentropic process.

A reversible process occurs when both the system and the surroundings are returned to their original
conditions after the process and reverse process have been carried out. Processes in nature are however
irreversible because reversal always causes some change to occur in the system and/or surroundings.
Factors causing irreversibility include:
• Friction
• Unrestricted expansion
• Heat transfer through a finite temperature difference
• Mixing of two different gases
• Chemical reactions.

8 JGhojel 2008
THE ZEROTH LAW OF THERMODYNAMICS
If two systems are in thermal equilibrium with a third system, then they are in thermal equilibrium with
each other and the three systems are said to be at the same temperature.

This law was added to the laws of thermodynamics early in the 20th century because it was realized that
the concept of equal-in-temperature is a prerequisite to a logical development of those laws. And to be
logical, it was named the Zeroth Law of thermodynamics.

The Thermodynamic Scale of Temperature


Temperature is a unique property of matter in that it is a fundamental concept, not expressible in terms
of other units. It is unsatisfactory, from a thermodynamic point of view, to express temperature in terms
of the length of a column of mercury or alcohol in a liquid-in-glass thermometer, or as the
electromotive force generated in a thermocouple. If temperature is to be used in thermodynamic
equations, and as an indicator of heat engine efficiency, then it must somehow be expressible in terms
of some thermodynamic function.

We should know first just what it is that temperature measures. If we look in the dictionary, we read
that temperature measures the relative "hotness" of things. But this definition in reality tells us nothing,
for when we look up hotness, the dictionary blandly tells us that it is something characterized by high
temperature. Here the kinetic theory helps us out of our dilemma. According to this theory, the pressure
of a gas varies directly with the kinetic energy of its molecules. But according to the gas laws, we can
see that the pressure of a gas also varies directly with the absolute temperature. Therefore, we may
conclude that the kinetic energy of molecules also varies directly as the absolute temperature, and is in
fact what temperature actually measures. The higher the molecular agitation the higher the temperature,
and the slower the molecular movement the lower the temperature becomes, until there would be no
temperature or pressure at all if we could somehow cool the gas to where the molecules were frozen
motionless.

This atomistic or microscopic explanation of temperature may be satisfying, in that it gives us a


physical picture that we can comprehend, but it does not tell us how we can measure temperature in
some fundamental way.

Thomson (Lord Kelvin) was very much aware of this perplexing problem. The answer came to him in
1848 during his pondering on the significance and implications of Carnot's reversible cycle. He had the
acuity to see that here was a means of establishing a universal and absolute temperature scale,
independent of any thermometric substance and based only on energy conversion.

To show how it is possible to arrive at an absolute scale, we shall first perform a hypothetical
experiment showing that an absolute zero of temperature must exist, and then extend this line of
reasoning to develop an "energy" or "thermodynamic" temperature scale.

Let us set up a small Carnot reversible engine to operate between a high-temperature region and a low-
temperature region as shown in Figure 11.

9 JGhojel 2008
Figure 11 The Carnot reversible engine

Since no engine can have a higher efficiency than the Carnot engine, the ratio of heat received to heat
rejected has a unique value when operating between temperatures (1) and (2).

Q1 T1 Q2 T2
= or =
Q2 T2 Q1 T1

where T represents a property called temperature on some thermodynamic scale which is independent
of any thermometric substance.

We have defined thermal efficiency of a heat engine as the work done divided by the heat received
W
thermal efficiency =
Q1

But according to the First Law of Thermodynamics heat energy cannot be destroyed, and therefore the
difference between the heat received and heat rejected by the cycle must be the work done, i.e.
Ql - Q2 = W. Making this substitution, then, we have,

Q1 − Q2 Q
thermal efficiency = or 1 − 2
Q1 Q1

Now substituting the temperature ratio for the heat ratio as shown in the first equation, we have,
T
thermal efficiency = 1 − 2 (21)
T1
From this expression of thermal efficiency we must conclude that T2, the temperature of the receiver,
can never be zero or less than zero. For if T2 were equal to zero, we see that the thermal efficiency of
the heat engine would be 1, or 100 per cent. This is inconceivable, since it would result in a complete
conversion of heat to work in a heat engine and would constitute a violation of the Second Law of
Thermodynamics. Also, if T2 were less than zero (a negative number), the thermal efficiency would be

10 JGhojel 2008
greater than 100 per cent which would mean a reversal of the direction of Q2, and thus heat would be
drawn directly from the low-temperature reservoir. This likewise would be a violation of the Second
Law. So, in theory, we can approach infinitely close to an absolute zero of temperature but never quite
attain it.

Next let us continue our hypothetical experiment by operating our little Carnot engine between a high-
temperature reservoir at the temperature of boiling water, T1, and a low-temperature reservoir at the
temperature of melting ice, T2. Suppose we had some metering device that could measure the amount
of heat received, Qsteam and the amount of heat rejected, Qice. If we actually were able to make these
Qsteam
measurements, we would find that = 1.3662 . But from our previous reasoning we note that
Qice
this must also be the ratio of the thermodynamic temperature of these two points. Thus,
Qsteam Tsteam
= = 1.3662 (22a)
Qice Tice

We might decide that we wanted, let us say, 100 divisions or degrees to separate the temperature of
boiling water and the temperature of melting ice on a scale called “Centigrade scale or C for short”.
Then we have
Tsteam – Tice =100oC (22b)

Here we have two equations with two unknowns, and by solving them simultaneously we can calculate
the temperature of boiling water and melting ice on a thermodynamic scale:

Tsteam = 373.07
degrees Kevin, K
Tice = 273.07

If we decide to have another scale where 180 divisions or degrees separate the temperature of boiling
water and the temperature of melting ice (Fahrenheit scale, F), then
Tsteam – Tice =180o which leads to

Tsteam = 672
degrees Rankine, R
Tice = 492

THE THIRD LAW OF THERMODYNAMICS


The absolute entropy of a pure crystalline substance in complete internal equilibrium is zero at zero
degree absolute.

The Third Law allows the determination of absolute entropies from thermal data. Knowing the absolute
o
entropy at a reference state ( s ), the absolute entropy of a substance at any other state is given by
T ,p
o
sT , p = s + ∫ ds (23)
To , p o
The reference state is usually taken at 0.1 MPa and 25oC.

11 JGhojel 2008
dT dp
ds = Cp −R (24)
T p

T p
o dT dp
sT , p = s + ∫ Cp −R ∫
To T po p

o T dT p
sT , p = s + ∫ C p − R ln (25)
To T po

so

The absolute entropies at 298 K and 0.1 MPa are known for most substances and the entropy change
ds for an ideal gas at any temperature other than the reference temperature can be determined if C p is
known at that temperature. Thermodynamic tables such as the JANAF tables list the absolute entropy
o
values for substances at a constant pressure of 0.1 MPa as a function of temperature; i.e. the value s
o T dT
so = s + ∫ C p (26)
To T
Equation (25) is used to calculate the absolute entropy values at pressures other than the reference
pressure of 0.1 MPa. When using equation (25) for chemical reactions, the pressure ratio in the last
term is replaced by the mole concentration of each substance.

ISENTROPIC EFFICIENCIES OF STEADY STATE FLOW SYSTEMS


Irreversibilities (process degradation) accompany all actual processes and entropy is the measure of this
degradation. Another way of measuring irreversibilities is to define process efficiencies. Process
efficiencies are defined in such a way that they are always positive numbers and less than unity.

Isentropic Efficiency of Turbines


The isentropic efficiency of a turbine when the gases expand from pressure p1 to p2 (Figure 12) is
defined as:
wa
ηt =
ws
where wa is the actual work and ws is the isentropic work done by the turbine during expansion.

12 JGhojel 2008
Figure 12 h-s diagram of the expansion process in an adiabatic turbine

From equation 12b the efficiency can be rewritten as


h1 − h2 a
ηt = (27)
h1 − h2 s

For an ideal gas


T1 − T2 a
ηt = (28)
T1 − T2 s

Isentropic Efficiency of Compressors


The isentropic efficiency of a compressor when the gases are compressed from pressure p1 to p2 (Figure
13) is defined as:
ws
ηc =
wa
where wa is the actual work and ws is the isentropic work done to compress the gas the compressor. The
work is provided from an external source.

13 JGhojel 2008
Figure 13 h-s diagram of the compression process in an adiabatic compressor

From equation 16 the efficiency can be rewritten as


h2 s − h1
ηc = (29)
h2 a − h1

For an ideal gas


T2 s − T1
ηc = (30)
T2 a − T1

Isentropic Efficiency of Nozzles


The isentropic efficiency of a nozzle in which the fluid is expanding from pressure p1 to p2 (Figure 14)
is defined as:
( KE ) a C 22a
ηn = =
( KE ) s C22s
Where ( KE ) a is the actual kinetic energy of the fluid at the nozzle exit and ( KE ) s is the isentropic
kinetic energy at the exit of an isentropic nozzle for the same inlet state and exit pressure. C2a and C2s.
are the corresponding exit velocities.

Figure 14 h-s diagram of the expansion process in an adiabatic nozzle

From equation (9)


C 22a C 22s
= h1 − h2 a and = h1 − h2 s and the efficiency is
2 2
h1 − h2 a
ηn = (31)
h1 − h2 s

For an ideal gas

14 JGhojel 2008
T1 − T2 s
ηn = (32)
T1 − T2 a

Isentropic Efficiency of Diffusers


A diffuser increases the pressure of the fluid by decreasing its kinetic energy. Diffusers are used as
intake ducts in turbojet and ramjet engines. The expression for the isentropic (intake) efficiency of a
diffuser is similar to the expression for the isentropic efficiency of a compressor. Referring to Figure
13,
h2 s − h1
ηd = (33)
h2 a − h1
or
T2 s − T1
ηd = (34)
T2 a − T1
HEAT ENGINE THEORY
One of Carnot’s great contributions to the science of thermodynamics was the recognition that heat
engines must operate with cyclic processes. A cycle occurs when a thermodynamic system, having
undergone a series of processes∗, arrives at a final state which is exactly the same as its initial state. In
other words, the working fluid must periodically return to the same state. Therefore, for any working
fluid, the work done and heat rejected during a cycle must all come from the original input heat. In
Carnot’s own words: “The thermal agency by which mechanical effect may be obtained is the
transference of heat from one body to another at a lower temperature.” Carnot also considered the
problem of determining the maximum work that can be obtained from the transfer of heat from high to
low temperature. He deduced a definition of a perfect thermodynamic engine thus; “Whatever amount
of mechanical effect it can derive from a certain thermal agency, if an equal amount be spent in
working it backwards, an equal reverse thermal effect will be produced.” Such an engine has come to
be known as the reversible engine and the quotation as the Carnot Principle. Furthermore, Carnot stated
that the maximum limits of temperature between which any actual heat engine can work are the
temperature of combustion of the fuel and the temperature of the coldest body we can easily find and
use in nature, usually the water of the locality. He firmly recognized that temperature and not pressure
was the fundamental criterion in determining heat engine efficiency. Figure 15 shows the Carnot
engine and cycle.


A process occurs when a system changes from one state to another. Thus, when air in a cylinder at an initial state with
properties V1 and T1 is heated, the volume may increase and the temperature may rise at a final state having properties V2
and T2
15 JGhojel 2008
Figure 15 Ideal Carnot Engine and Cycle

Process 1-2 Isothermal expansion (PV = const.) with heat addition


Process 2-3 Reversible adiabatic expansion (PVk = Const.)
Process 3-4 Isothermal compression (PV = const.) with heat rejection
Process 4-1 Reversible adiabatic compression (PVk = Const.)

Thermal efficiency of a this heat engine is


QH − QL Q T
η= =1− L =1− L (35)
QH QH TH

16 JGhojel 2008
APPENDIX B

POLYTROPIC (SMALL-STAGE) EFFICIENCY


The isentropic compressor and turbine efficiencies defined earlier vary with variation in the number of
stages in the compressor (variation in pressure ratio). This is due to the fact that the vertical distance
between any two constant pressure lines increase with entropy increase. Therefore it is incorrect to
assume constant isentropic efficiencies for the compressor and turbine over the wide range of pressure
ratios encountered in some applications. The concept of polytropic or small-stage efficiency of
compression and expansion is useful, since it can be assumed to be constant and independent of the
pressure ratio.

Consider an axial compressor comprising several infinitely small stages. The small stage efficiency ηp
is defined as the constant isentropic efficiency for each small stage (Figure1).

Figure1 Infinitely small multistage compression

For the small stage isentropic compression of a perfect gas


dTs
η pc = = const (1)
dTa
Tγ T
but = const , or = const
p ( γ −1 ) p ( γ −1 ) γ

Figure 2 Variation of compressor and turbine isentropic efficiencies with


17 JGhojel 2008
pressure ratio at constant polytropic efficiency

In differential form

dTs γ − 1 dp dT γ − 1 dp
= , or η pc =
T γ p T γ p

Integrating between points 1 and 2, keeping in mind that ηpc is by definition constant, we obtain

ln( p2 p1 )( γ −1 ) γ
η pc = (2)
ln( T2 a T1 )

This equation enables to calculate ηpc from the known values of p and T at the inlet and outlet of
compressor. Rearranging equation (2) yields

T2 a p ( γ −1 ) γη pc
=( 2 ) (3)
T1 p1
ln(T1 T2a )
η pt = γ −1
ln( p1 p2 ) γ

The overall isentropic efficiency from Figure 1 is

T2 s p
− 1 ( 2 )( γ −1 ) γ − 1
∆T T − T T p1
ηc = s = 2 s 1 = 1 = (4)
∆T T2 a − T1 T2 a − 1 ( p2 )( γ −1 ) γη pc − 1
T1 p1
This equation enables to determine the variation of ηc with the compressor pressure ratio for a given
polytropic efficiency ηpc. As Figure 2 shows, the compressor isentropic efficiency decreases with
increasing pressure ratio at constant polytropic efficiency (0.9). For an irreversible adiabatic
(polytropic) process

dT n − 1 dp dT γ − 1
= and =
T n p T γη pc
γ −1 n−1
=
γη pc n
dT γ − 1 dp
and since = ,
T γη pc p

18 JGhojel 2008
γ −1 n −1
∴ = (5)
γη pc n
hence the origin of the term polytropic efficiency.

A similar expression can be obtained for the expansion process in a multistage turbine by defining the
efficiency of an infinitely small expansion stage (Figure 3)

Figure 3 Infinitely small multistage expansion

By definition

dTa
η pt = = const
dTs
(6)
T1 p η ( γ −1 ) γ
= ( 1 ) pt
T2 a p2

The overall isentropic efficiency of the expansion process is

T2 a p η ( γ −1 ) / γ
1− 1 − ( 2 ) pt
T −T T1 p1
ηt = 1 2 a = = (7)
T1 − T2 s T p
1 − 2s 1 − ( 2 )( γ −1 ) γ
T1 p1

The variation of the isentropic efficiency of the turbine as function of pressure ratio at constant
polytropic efficiency of 0.9 is shown in Figure 2 . The temperature equivalents (in terms of stagnation
temperatures) of the work transfers for a given pressure ratio for both the compressor and turbine
respectively are

19 JGhojel 2008
 (γ − 1) γ η
pc 
 p  
= T   − 1
2
T −T (8)
2t 1t 1t
 1 
p

 

 (γ − 1 )η
pt
γ 
 p  
= T 1 −   
2
T −T (9)
1t 2t 1t
  p
1 
 

20 JGhojel 2008

Vous aimerez peut-être aussi