You are on page 1of 35

Thermodynamics and Propulsion

Next: 3.2 Generalized Representation of Up: 3. The First Law Previous: 3. The
First Law Contents Index

3.1 Some Properties of Engineering


Cycles; Work and Efficiency
As preparation for our discussion of cycles (and as a foreshadowing of the
second law), we examine two types of processes that concern interactions
between heat and work. The first of these represents the conversion of work
into heat. The second, which is much more useful, concerns the conversion of
heat into work. The question we will pose is how efficient can this conversion be
in the two cases.

Figure 3.1: Examples of the conversion of work into heat

Three examples of the first process are given in Figure 3.1. The first is the
pulling of a block on a rough horizontal surface by a force which moves through
some distance. Friction resists the pulling. After the force has moved through
the distance, it is removed. The block then has no kinetic energy and the same
potential energy it had when the force started to act. If we measured the
temperature of the block and the surface we would find that it was higher than
when we started. (High temperatures can be reached if the velocities of pulling
are high; this is the basis of inertia welding.) The work done to move the block
has been converted totally to heat.

The second example concerns the stirring of a viscous liquid. There is work
associated with the torque exerted on the shaft turning through an angle. When
the stirring stops, the fluid comes to rest and there is (again) no change in
kinetic or potential energy from the initial state. The fluid and the paddle wheels
will be found to be hotter than when we started, however.

The final example is the passage of a current through a resistance. This is a


case of electrical work being converted to heat, indeed it models operation of an
electrical heater.

All the examples in Figure 3.1 have 100% conversion of work into heat. This
100% conversion could go on without limit as long as work were supplied. Is
this true for the conversion of heat into work?

To answer the last question, we need to have some basis for judging whether
work is done in a given process. One way to do this is to ask whether we can
construct a way that the process could result in the raising of a weight in a
gravitational field. If so, we can say ``Work has been done.'' It may sometimes
be difficult to make the link between a complicated thermodynamic process and
the simple raising of a weight, but this is a rigorous test for the existence of
work.

One example of a process in which heat is converted to work is the isothermal


(constant temperature) expansion of an ideal gas, as sketched in Figure 3.2.
The system is the gas inside the chamber. As the gas expands, the piston does
work on some external device. For an ideal gas, the internal energy is a
function of temperature only, so that if the temperature is constant for some
process the internal energy change is zero. To keep the temperature constant

during the expansion, heat must be supplied. Because the first law

takes the form . This is a process that has 100% conversion of heat
into work.

Figure 3.2: Isothermal expansion

The work exerted by the system is given by

where 1 and 2 denote the two states at the beginning and end of the process.
The equation of state for an ideal gas is
with the number of moles of the gas contained in the chamber. Using the
equation of state, the expression for work can be written as

(3..1
)

For an isothermal process, , so that . The


work can be written in terms of the pressures at the beginning and end as

(3..2)

The lowest pressure to which we can expand and still receive work from the
system is atmospheric pressure. Below this, we would have to do work on the
system to pull the piston out further. There is thus a bound on the amount of
work that can be obtained in the isothermal expansion; we cannot continue
indefinitely. For a power or propulsion system, however, we would like a source
of continuous power, in other words a device that would give power or
propulsion as long as fuel was added to it. To do this, we need a series of
processes where the system does not progress through a one-way transition
from an initial state to a different final state, but rather cycles back to the initial
state. What is looked for is in fact a thermodynamic cycle for the system.

We define several quantities for a cycle:

 is the heat absorbed by the system.

 is the heat rejected by the system.


 is the net work done by the system.

The cycle returns to its initial state, so the overall energy change, , is zero.
The net work done by the system is related to the magnitudes of the heat
absorbed and the heat rejected by
The thermal efficiency of the cycle is the ratio of the work done to the heat
absorbed. (Efficiencies are often usefully portrayed as ``What you get'' versus
``What you pay for.'' Here what we get is work and what we pay for is heat, or
rather the fuel that generates the heat.) In terms of the heat absorbed and
rejected, the thermal efficiency is

(3..3)

The thermal efficiency can only be 100% (complete conversion of heat into

work) if ; a basic question is what is the maximum thermal efficiency


for any arbitrary cycle? We examine this for several cases, including the Carnot
cycle and the Brayton (or Joule) cycle, which is a model for the power cycle in a
jet engine.

Next: 3.2 Generalized Representation of Up: 3. The First Law Previous: 3. The
First Law Contents Index
UnifiedTP
Thermodynamics and Propulsion

Next: 3.4 Refrigerators and Heat Up: 3. The First Law Previous: 3.2
Generalized Representation of Contents Index

3.3 The Carnot Cycle


A Carnot cycle is shown in Figure 3.4. It has four processes. There are two
adiabatic reversible legs and two isothermal reversible legs. We can construct a
Carnot cycle with many different systems, but the concepts can be shown using
a familiar working fluid, the ideal gas. The system can be regarded as a
chamber enclosed by a piston and filled with this ideal gas.

Figure 3.4: Carnot cycle -- thermodynamic diagram on left and schematic of the
different stages in the cycle for a system composed of an ideal gas on the right

The four processes in the Carnot cycle are:

1. The system is at temperature at state . It is brought in contact with


a heat reservoir, which is just a liquid or solid mass of large enough
extent such that its temperature does not change appreciably when
some amount of heat is transferred to the system. In other words, the
heat reservoir is a constant temperature source (or receiver) of heat. The
system then undergoes an isothermal expansion from to , with heat

absorbed .
2. At state , the system is thermally insulated (removed from contact with
the heat reservoir) and then let expand to . During this expansion the
temperature decreases to . The heat exchanged during this part of

the cycle, )
3. At state the system is brought in contact with a heat reservoir at

temperature . It is then compressed to state , rejecting heat in


the process.
4. Finally, the system is compressed adiabatically back to the initial

state . The heat exchange .

The thermal efficiency of the cycle is given by the definition

(3..4)

In this equation, there is a sign convention implied. The quantities , as


defined are the magnitudes of the heat absorbed and rejected. The

quantities , on the other hand are defined with reference to heat


received by the system. In this example, the former is negative and the latter is
positive. The heat absorbed and rejected by the system takes place during
isothermal processes and we already know what their values are from Eq. (3.1):

The efficiency can now be written in terms of the volumes at the different states
as

(3..5)

The path from states to and from to are both adiabatic and reversible.
For a reversible adiabatic process we know that . Using the
ideal gas equation of state, we have . Along curve - ,
therefore, . Along the curve - , .
Thus,

Comparing the expression for thermal efficiency Eq. (3.4) with Eq. (3.5) shows
two consequences. First, the heats received and rejected are related to the
temperatures of the isothermal parts of the cycle by

(3..6)

Second, the efficiency of a Carnot cycle is given compactly by

(3..7)

The efficiency can be 100% only if the temperature at which the heat is rejected
is zero. The heat and work transfers to and from the system are shown
schematically in Figure 3.5.

Figure 3.5: Work and heat transfers in a Carnot cycle between two heat reservoirs
Muddy Points

Since , looking at the - graph, does that mean the farther

apart the , isotherms are, the greater efficiency? And that if they were
very close, it would be very inefficient? (MP 3.2)

In the Carnot cycle, why are we only dealing with volume changes and not
pressure changes on the adiabats and isotherms? (MP 3.3)

Is there a physical application for the Carnot cycle? Can we design a Carnot
engine for a propulsion device? (MP 3.4)

How do we know which cycles to use as models for real processes? (MP 3.5)

Next: 3.4 Refrigerators and Heat Up: 3. The First Law Previous: 3.2
Generalized Representation of Contents Index
UnifiedTP
Thermodynamics and Propulsion

Next: 3.5 The Internal combustion Up: 3. The First Law Previous: 3.3 The Carnot
Cycle Contents Index

Subsections


o 3.4.0.1 Refrigerator Hardware

3.4 Refrigerators and Heat Pumps


The Carnot cycle has been used for power, but we can also run it in reverse. If
so, there is now net work into the system and net heat out of the system. There

will be a quantity of heat rejected at the higher temperature and a quantity of

heat absorbed at the lower temperature. The former of these is negative


according to our convention and the latter is positive. The result is that work is
done on the system, heat is extracted from a low temperature source and
rejected to a high temperature source. The words ``low'' and ``high'' are relative
and the low temperature source might be a crowded classroom on a hot day,
with the heat extraction being used to cool the room. The cycle and the heat
and work transfers are indicated in Figure 3.6. In this mode of operation the
cycle works as a refrigerator or heat pump. ``What we pay for'' is the work, and
``what we get'' is the amount of heat extracted. A metric for devices of this type
is the coefficient of performance, defined as
Figure 3.6: Operation of a Carnot refrigerator

For a Carnot cycle we know the ratios of heat in to heat out when the cycle is
run forward and, since the cycle is reversible, these ratios are the same when
the cycle is run in reverse. The coefficient of performance is thus given in terms
of the absolute temperatures as

This can be much larger than unity.

The Carnot cycles that have been drawn are based on ideal gas behavior. For
different working media, however, they will look different. We will see an
example when we discuss two-phase situations. What is the same whatever the
medium is the efficiency for all Carnot cycles operating between the same two
temperatures.

3.4.0.1 Refrigerator Hardware

Typically the thermodynamic system in a refrigerator analysis will be a working


fluid, a refrigerant, that circulates around a loop, as shown in Figure 3.7. The
internal energy (and temperature) of the refrigerant is alternately raised and
lowered by the devices in the loop. The working fluid is colder than the
refrigerator air at one point and hotter than the surroundings at another point.
Thus heat will flow in the appropriate direction, as shown by the two arrows in
the heat exchangers.
Figure 3.7: Schematic of a domestic refrigerator

Starting in the upper right hand corner of the diagram, we describe the process
in more detail. First the refrigerant passes through a small turbine or through an
expansion valve. In these devices, work is done by the refrigerant so its internal
energy is lowered to a point where the temperature of the refrigerant is lower
than that of the air in the refrigerator. A heat exchanger is used to transfer
energy from the inside of the refrigerator to the cold refrigerant. This lowers the
internal energy of the inside and raises the internal energy of the refrigerant.
Then a pump or compressor is used to do work on the refrigerant, adding
additional energy to it and thus further raising its internal energy. Electrical
energy is used to drive the pump or compressor. The internal energy of the
refrigerant is raised to a point where its temperature is hotter than the
temperature of the surroundings. The refrigerant is then passed through a heat
exchanger (often coils at the back of the refrigerator) so that energy is
transferred from the refrigerant to the surroundings. As a result, the internal
energy of the refrigerant is reduced and the internal energy of the surroundings
is increased. It is at this point where the internal energy of the contents of the
refrigerator and the energy used to drive the compressor or pump are
transferred to the surroundings. The refrigerant then continues on to the turbine
or expansion valve, repeating the cycle.

Next: 3.5 The Internal combustion Up: 3. The First Law Previous: 3.3 The Carnot
Cycle Contents Index

UnifiedTP
Thermodynamics and Propulsion

Next: 3.6 Diesel Cycle Up: 3. The First Law Previous: 3.4 Refrigerators and
Heat Contents Index

Subsections

 3.5.1 Efficiency of an ideal Otto cycle


 3.5.2 Engine work, rate of work per unit enthalpy flux

3.5 The Internal combustion engine


(Otto Cycle)
[VW, S & B: 9.13]

The Otto cycle is a set of processes used by spark ignition internal combustion
engines (2-stroke or 4-stroke cycles). These engines a) ingest a mixture of fuel
and air, b) compress it, c) cause it to react, thus effectively adding heat through
converting chemical energy into thermal energy, d) expand the combustion
products, and then e) eject the combustion products and replace them with a
new charge of fuel and air. The different processes are shown in Figure 3.8:

1. Intake stroke, gasoline vapor and air drawn into engine ( ).

2. Compression stroke, , increase ( ).


3. Combustion (spark), short time, essentially constant volume ( ). Model:

heat absorbed from a series of reservoirs at temperatures to .


4. Power stroke: expansion ( ).
5. Valve exhaust: valve opens, gas escapes.
6. ( ) Model: rejection of heat to series of reservoirs at

temperatures to .
7. Exhaust stroke, piston pushes remaining combustion products out of chamber
( ).

We model the processes as all acting on a fixed mass of air contained in a


piston-cylinder arrangement, as shown in Figure 3.10.

Figure 3.8: The ideal Otto cycle

Figure 3.9: Sketch of an actual Otto cycle


Figure 3.10: Piston and valves in a four-stroke internal combustion engine

The actual cycle does not have the sharp transitions between the different
processes that the ideal cycle has, and might be as sketched in Figure 3.9.

3.5.1 Efficiency of an ideal Otto cycle


The starting point is the general expression for the thermal efficiency of a cycle:

The convention, as previously, is that heat exchange is positive if heat is flowing into

the system or engine, so is negative. The heat absorbed occurs during


combustion when the spark occurs, roughly at constant volume. The heat absorbed
can be related to the temperature change from state 2 to state 3 as:
The heat rejected is given by (for a perfect gas with constant specific heats)

Substituting the expressions for the heat absorbed and rejected in the expression for
thermal efficiency yields

We can simplify the above expression using the fact that the processes from 1 to 2 and
from 3 to 4 are isentropic:

The quantity is called the compression ratio. In terms of compression


ratio, the efficiency of an ideal Otto cycle is:

Figure 3.11: Ideal Otto cycle thermal efficiency

The ideal Otto cycle efficiency is shown as a function of the compression ratio

in Figure 3.11. As the compression ratio, , increases, increases, but so


does . If is too high, the mixture will ignite without a spark (at the wrong
location in the cycle).

3.5.2 Engine work, rate of work per unit enthalpy flux


The non-dimensional ratio of work done (the power) to the enthalpy flux through the
engine is given by

There is often a desire to increase this quantity, because it means a smaller engine for
the same power. The heat input is given by

where

 is the heat of reaction, i.e. the chemical energy liberated per unit mass
of fuel,

 is the fuel mass flow rate.

The non-dimensional power is

The quantities in this equation, evaluated at stoichiometric conditions are:

so
Muddy Points

How is calculated? (MP 3.6)

What are ``stoichiometric conditions?'' (MP 3.7)

Next: 3.6 Diesel Cycle Up: 3. The First Law Previous: 3.4 Refrigerators and
Heat Contents Index

UnifiedTP
Thermodynamics and Propulsion

Next: 3.7 Brayton Cycle Up: 3. The First Law Previous: 3.5 The Internal
combustion Contents Index

3.6 Diesel Cycle


The Diesel cycle is a compression ignition (rather than spark ignition) engine. Fuel is

sprayed into the cylinder at (high pressure) when the compression is complete,
and there is ignition without a spark. An idealized Diesel engine cycle is shown in
Figure 3.12.

Figure 3.12: The ideal Diesel cycle

The thermal efficiency is given by:


This cycle can operate with a higher compression ratio than the Otto cycle because
only air is compressed and there is no risk of auto-ignition of the fuel. Although for a
given compression ratio the Otto cycle has higher efficiency, because the Diesel
engine can be operated to higher compression ratio, the engine can actually have
higher efficiency than an Otto cycle when both are operated at compression ratios that
might be achieved in practice.

Muddy Points

When and where do we use and ? Some definitions use . Is

it ever ? (MP 3.8)

Explanation of the above comparison between Diesel and Otto. (MP 3.9)

Next: 3.7 Brayton Cycle Up: 3. The First Law Previous: 3.5 The Internal
combustion Contents Index

UnifiedTP
Thermodynamics and Propulsion

Next: 3.8 Muddiest points on Up: 3. The First Law Previous: 3.6 Diesel
Cycle Contents Index

Subsections

 3.7.1 Work and Efficiency


 3.7.2 Gas Turbine Technology and Thermodynamics
 3.7.3 Brayton Cycle for Jet Propulsion: the Ideal Ramjet
 3.7.4 MIT Cogenerator

3.7 Brayton Cycle


[VW, S & B: 9.8-9.9, 9.12]

The Brayton cycle (or Joule cycle) represents the operation of a gas turbine
engine. The cycle consists of four processes, as shown in
Figure 3.13 alongside a sketch of an engine:

 a - b Adiabatic, quasi-static (or reversible) compression in the inlet and


compressor;
 b - c Constant pressure fuel combustion (idealized as constant pressure heat
addition);
 c - d Adiabatic, quasi-static (or reversible) expansion in the turbine and exhaust
nozzle, with which we
1. take some work out of the air and use it to drive the compressor, and
2. take the remaining work out and use it to accelerate fluid for jet
propulsion, or to turn a generator for electrical power generation;
 d - a Cool the air at constant pressure back to its initial condition.
Figure 3.13: Sketch of the jet engine components and corresponding thermodynamic states

The components of a Brayton cycle device for jet propulsion are shown in
Figure 3.14. We will typically represent these components schematically, as in
Figure 3.15. In practice, real Brayton cycles take one of two forms.
Figure 3.16(a) shows an ``open'' cycle, where the working fluid enters and then
exits the device. This is the way a jet propulsion cycle works.
Figure 3.16(b)shows the alternative, a closed cycle, which recirculates the
working fluid. Closed cycles are used, for example, in space power generation.

Figure 3.14: Schematics of typical military gas turbine engines. Top: turbojet with
afterburning, bottom: GE F404 low bypass ratio turbofan with afterburning (Hill and Peterson,
1992).
Figure 3.15: Thermodynamic model of gas turbine engine cycle for power generation

[Open cycle operation] [Closed cycle

operation]

Figure 3.16: Options for operating Brayton cycle gas turbine engines

Muddy Points
Would it be practical to run a Brayton cycle in reverse and use it as refrigerator?
(MP 3.10)

3.7.1 Work and Efficiency

The objective now is to find the work done, the heat absorbed, and the thermal
efficiency of the cycle. Tracing the path shown around the cycle from - - -
and back to , the first law gives (writing the equation in terms of a unit
mass),

Here is zero because is a function of state, and any cycle returns the system to
its starting state3.2. The net work done is therefore

where , are defined as heat received by the system ( is negative). We thus


need to evaluate the heat transferred in processes - and - .

For a constant pressure, quasi-static process the heat exchange per unit mass
is

We can see this by writing the first law in terms of enthalpy (see Section 2.3.4)

or by remembering the definition of .

The heat exchange can be expressed in terms of enthalpy differences between


the relevant states. Treating the working fluid as a perfect gas with constant
specific heats, for the heat addition from the combustor,

The heat rejected is, similarly,

The net work per unit mass is given by


The thermal efficiency of the Brayton cycle can now be expressed in terms of
the temperatures:

(3..8
)

To proceed further, we need to examine the relationships between the different


temperatures. We know that points and are on a constant pressure

process as are points and , and ; . The other two legs


of the cycle are adiabatic and reversible, so

Therefore , or, finally, . Using this relation in


the expression for thermal efficiency, Eq. (3.8) yields an expression for the
thermal efficiency of a Brayton cycle:

(3..9)

The temperature ratio across the compressor, . In terms of


compressor temperature ratio, and using the relation for an adiabatic reversible
process we can write the efficiency in terms of the compressor (and cycle)
pressure ratio, which is the parameter commonly used:
(3..10)

Figure 3.17: Gas turbine engine pressures and temperatures

Figure 3.17 shows pressures and temperatures through a gas turbine engine
(the PW4000, which powers the 747 and the 767).
Figure 3.18: Gas turbine engine pressure ratio trends (Jane’s Aeroengines, 1998)

Figure 3.19: Trend of Brayton cycle thermal efficiency with compressor pressure ratio

Equation (3.10) says that for a high cycle efficiency, the pressure ratio of the
cycle should be increased. This trend is plotted in Figure 3.19.
Figure 3.18 shows the history of aircraft engine pressure ratio versus entry into
service, and it can be seen that there has been a large increase in cycle
pressure ratio. The thermodynamic concepts apply to the behavior of real
aerospace devices!

Muddy Points

When flow is accelerated in a nozzle, doesn't that reduce the internal energy of
the flow and therefore the enthalpy? (MP 3.11)

Why do we say the combustion in a gas turbine engine is constant pressure?


(MP 3.12)

Why is the Brayton cycle less efficient than the Carnot cycle? (MP 3.13)

If the gas undergoes constant pressure cooling in the exhaust outside the
engine, is that still within the system boundary? (MP 3.14)

Does it matter what labels we put on the corners of the cycle or not? (MP 3.15)

Is the work done in the compressor always equal to the work done in the turbine
plus work out (for a Brayton cyle)? (MP 3.16)

3.7.2 Gas Turbine Technology and Thermodynamics

The turbine entry temperature, , is fixed by materials technology and cost. (If the
temperature is too high, the blades fail.) Figures 3.20 and 3.21 show the progression of
the turbine entry temperatures in aeroengines. Figure 3.20 is from Rolls Royce and
Figure 3.21 is from Pratt & Whitney. Note the relation between the gas temperature
coming into the turbine blades and the blade melting temperature.
Figure 3.20: Rolls-Royce high temperature technology

Figure 3.21: Turbine blade cooling technology [Pratt & Whitney]

For a given level of turbine technology (in other words given maximum
temperature) a design question is what should the compressor be? What
criterion should be used to decide this? Maximum thermal efficiency? Maximum
work? We examine this issue below.
Figure 3.22: Efficiency and work of two Brayton cycle engines

The problem is posed in Figure 3.22, which shows two Brayton cycles. For
maximum efficiency we would like as high as possible. This means that the
compressor exit temperature approaches the turbine entry temperature. The net

work will be less than the heat received; as the heat received
approaches zero and so does the net work.

The net work in the cycle can also be expressed as , evaluated in


traversing the cycle. This is the area enclosed by the curves, which is seen to

approach zero as .

The conclusion from either of these arguments is that a cycle designed for
maximum thermal efficiency is not very useful in that the work (power) we get
out of it is zero.

A more useful criterion is that of maximum work per unit mass (maximum power
per unit mass flow). This leads to compact propulsion devices. The work per
unit mass is given by:

where is the maximum turbine inlet temperature (a design constraint) and is


atmospheric temperature. The design variable is the compressor exit
temperature, , and to find the maximum as this is varied, we differentiate the

expression for work with respect to :

The first and the fourth terms on the right hand side of the above equation are
both zero (the turbine entry temperature is fixed, as is the atmospheric
temperature). The maximum work occurs where the derivative of work with

respect to is zero:

(3..11)

To use Eq. (3.11), we need to relate and . We know that

Hence,

Plugging this expression for the derivative into Eq. (3.11) gives the compressor

exit temperature for maximum work as . In terms of temperature


ratio,
The condition for maximum work in a Brayton cycle is different than that for
maximum efficiency. The role of the temperature ratio can be seen if we
examine the work per unit mass which is delivered at this condition:

Ratioing all temperatures to the engine inlet temperature,

To find the power the engine can produce, we need to multiply the work per unit
mass by the mass flow rate:

(3..12
)

The trend of work output vs. compressor pressure ratio, for different

temperature ratios , is shown in Figure 3.23.


Figure 3.23: Trend of cycle work with compressor pressure ratio, for different temperature

ratios
[Gas turbine engine core] [Core power vs. turbine

entry temperature]

Figure 3.24: Aeroengine core power [Koff/Meese, 1995]

Figure 3.24 shows the expression for power of an ideal cycle compared with
data from actual jet engines. Figure 3.24(a) shows the gas turbine engine layout
including the core (compressor, burner, and turbine). Figure 3.24(b) shows the
core power for a number of different engines as a function of the turbine rotor
entry temperature. The equation in the figure for horsepower (HP) is the same
as that which we just derived, except for the conversion factors. The analysis
not only shows the qualitative trend very well but captures much of the
quantitative behavior too.

A final comment (for this section) on Brayton cycles concerns the value of the
thermal efficiency. The Brayton cycle thermal efficiency contains the ratio of the
compressor exit temperature to atmospheric temperature, so that the ratio is not
based on the highest temperature in the cycle, as the Carnot efficiency is. For a
given maximum cycle temperature, the Brayton cycle is therefore less efficient
than a Carnot cycle.

Muddy Points

What are the units of in ? (MP 3.17)


Question about the assumptions made in the Brayton cycle for maximum
efficiency and maximum work (MP 3.18)

You said that for a gas turbine engine modeled as a Brayton cycle the work

done is , where is the heat added and is the heat rejected.


Does this suggest that the work that you get out of the engine doesn't depend
on how good your compressor and turbine are? since the compression and
expansion were modeled as adiabatic. (MP 3.19)

3.7.3 Brayton Cycle for Jet Propulsion: the Ideal Ramjet


A schematic of a ramjet is given in Figure 3.25.

Figure 3.25: Ideal ramjet [J. L. Kerrebrock, Aircraft Engines and Gas Turbines]

In the ramjet there are ``no moving parts.'' The processes that occur in this
propulsion device are:

 : Isentropic diffusion (slowing down) and compression, with a decrease

in Mach number, .
 : Constant pressure combustion.
 : Isentropic expansion through the nozzle.

The ramjet thermodynamic cycle efficiency can be written in terms of flight

Mach number, , as follows:

and
so

See also Section 11.6.3 for other figures of merit.

Muddy Points

Why don't we like the numbers 1 and 2 for the stations? Why do we go 0-3?
(MP 3.20)

For the Brayton cycle efficiency, why does ? (MP 3.21)

3.7.4 MIT Cogenerator

MIT operates a Brayton cycle power generator on campus. For more


information, see the website at https://cogen.mit.edu/ctg.cfm .

Next: 3.8 Muddiest points on Up: 3. The First Law Previous: 3.6 Diesel
Cycle Contents Index

UnifiedTP

You might also like