Sunteți pe pagina 1din 260

1

CHAPTER ONE

FUNDAMENTALS OF
ENERGY CONVERSION

1.1 Introduction

Energy conversion engineering (or heat-power engineering, as it was called prior to the
Second World War), has been one of the central themes in the development of the
engineering profession. It is concerned with the transformation of energy from sources
such as fossil and nuclear fuels and the sun into conveniently used forms such as
electrical energy, rotational and propulsive energy, and heating and cooling.
A multitude of choices and challenges face the modern energy conversion
engineer. A few years ago major segments of the energy conversion industry were
settled into a pattern of slow innovation. Most automobile manufacturers were satisfied
to manufacture engines that had evolved from those produced twenty years earlier,
some of which boasted 400 horsepower and consumed a gallon of leaded gasolene
every eight or nine miles. Many electric power utilities were content with state-of-the-
art, reliable, fossil-fuel-consuming steam power plants, except for a few
forward-looking, and in several cases unfortunate, exceptions that risked the nuclear
alternative.
Then came the oil embargo of the 1970s, high fuel prices, and threatened
shortages. Also, the public and legislatures began to recognize that air pollution
produced by factories, power plants, and automobiles and other forms of environmental
pollution were harmful. International competitors, producing quality automobiles with
smaller, lower-pollution engines, exceptional gas mileage, and lower prices shook the
automobile industry. The limitations of the Earth's resources and environment started
to come into clearer focus. These and other influences have been helping to create a
more favorable climate for consideration, if not total acceptance, of energy conversion
alternatives and new concepts.
There are opposing factors, however. Among them are limited research and
development funding due to budgetary constraints, emphasis on short-term rather than
long-term goals because of entrepreneurial insistence on rapid payback on investment,
and managerial obsession with the bottom line. But more open attitudes have become
established. New as well as previously shelved ideas are now being considered or
reconsidered, tested, and sometimes implemented. A few examples are combined
steam and gas turbine cycles, rotary combustion engines, solar and windmill power
farms, stationary and vehicular gas turbine power plants, cogeneration, photovoltaic
solar power, refuse-derived fuel, stratified charge engines, turbocharged engines,
fluidized-bed combustors, and coal-gasification power plants. We are living in a
2

rapidly changing world that requires continuing adaptation of old technologies and the
development of new ones. Energy conversion engineering is a more stimulating,
complex, and viable field today because of this altered climate.

A Look Backward

How did we get where we are today? A good answer requires a study of the history of
science and engineering worthy of many volumes. Table 1.1 identifies a few pivotal
ideas and inventions, some of them landmarks to energy conversion engineers, and the
names of the thinkers and movers associated with them. Of course, the table cannot
present the entire history of energy conversion engineering. Omitted are the
contributions of Newton and Euler, the first rocket engine, the V-8 engine, the ramjet
and fanjet. The reader could easily come up with many other glaring omissions and
extend the table indefinitely.
While the names of one or two persons are associated with each landmark
achievement, most of these landmarks were the products of teams of unheralded
individuals whose talents were crucial to success. Moreover, the successes did not
occur in a vacuum, but benefited and followed from the advances and failures of others.
Unknown or renowned, each engineer and his or her associates can make a contribution
to the progress of mankind.
Table 1.1 can only hint at how the persons, ideas, and events listed there relied on
their predecessors and on a host of less well-known scientific and technological
advances. A brief bibliography of historical sources is given at the end of the chapter.
These works chronicle the efforts of famous and unsung heroes, and a few villains, of
energy conversion and their struggles with ideas and limiting tools and resources to
produce machines for man and industry.
The historical progress of industry and technology was slow until the fundamentals
of thermodynamics and electromagnetism were established in the ninteenth century.
The blossoming of energy technology and its central role in the industrial revolution is
well known to all students of history. It is also abundantly clear that the development of
nuclear power in the second half of the twentieth century grew from theoretical and
experimental scientific advances of the first half century. After a little reflection on
Table 1.1, there should be no further need to justify a fundamental scientific and
mathematical approach to energy conversion engineering.

TABLE 1-1 Some Significant Events in the History of Energy Conversion


___________________________________________________________________________
Giovanni Branca Impulse steam turbine proposal 1629

Thomas Newcomen Atmospheric engine using steam (first widely used 1700
Heat engine)

James Watt Separate steam condenser idea; 1765


and first Boulton and Watt condensing steam engine 1775
3

Table 1.1 (continued)


___________________________________________________________________________
John Barber Gas turbine ideas and patent 1791

Benjamin Thompson Observed conversion of mechanical energy 1798


(Count Rumford) to heat while boring cannon

Robert Fulton First commercial steamboat 1807

Robert Stirling Stirling engine 1816

N. L. Sadi Carnot Principles for an ideal heat engine 1824


(foundations of thermodynamics)

Michael Faraday First electric current generator 1831

Robert Mayer Equivalence of heat and work 1842

James Joule Basic ideas of the First Law of Thermodynamics; 1847


and measured the mechanical equivalent of heat 1849

Rudolph Clausius Second Law of Thermodynamics 1850

William Thompson Alternate form of the Second Law of 1851


(Lord Kelvin) Thermodynamics

Etienne Lenoir Internal combustion engine without


without mechanical compression 1860

A. Beau de Rochas Four-stroke cycle internal combustion engine concept 1862

James C. Maxwell Mathematical principles of electromagnetics 1865

Niklaus Otto Four-stroke cycle internal combustion engine 1876

Charles Parsons Multistage, axial-flow reaction steam turbine 1884

Thomas Edison Pearl Street steam-engine-driven electrical power plant 1884

C.G.P. de Laval Impulse steam turbine with convergent-divergent nozzle 1889

Rudolph Diesel Compression ignition engine 1892

___ First hydroelectric power at Niagara Falls 1895

Albert Einstein Mass-energy equivalence 1905

Ernst Schrodinger Quantum wave mechanics 1926

Frank Whittle Turbojet engine patent application; 1930


and first jet engine static test 1937
4

Table 1.1 (concluded)


___________________________________________________________________________
Otto Hahn Discovery of nuclear fission 1938

Hans von Ohain First turbojet engine flight 1939

J. Ackeret, C. Keller Closed-cycle gas turbine electric power generation 1939

Enrico Fermi Nuclear fission demonstration at the University of Chicago 1942

Felix Wankel Rotary internal combustion engine 1954

Production of electricity via nuclear fission by a utility 1957


at Shippingport, Pennsylvania .

NASA Rocket-powered landing of man on the moon 1969

Electricité de France Superphénix 1200-MW fast breeder reactor – first 1986


grid power
________________________________________________________________

Since energy conversion engineering is deeply rooted in thermodynamics, fluid


mechanics, and heat transfer, this chapter briefly reviews those aspects of these
disciplines that are necessary for understanding, analysis, and design in the field of
energy conversion.

1.2 Fundamentals of Thermodynamics

The subject of thermodynamics stems from the notions of temperature, heat, and work.
Although, the following discussion makes occasional reference to molecules and
particles, useful in clarifying and motivating concepts in thermodynamics,
thermodynamics as a science deals with matter as continuous rather than as discrete or
granular.

System, Surroundings, and Universe

We define a pure substance as a homogeneous collection of matter. Consider a fixed


mass of a pure substance bounded by a closed, impenetrable, flexible surface. Such a
mass, called a system, is depicted schematically in Figure 1.1(a). For example, the
system could be a collection of molecules of water, air, refrigerant, or combustion gas
confined in a closed container such as the boundary formed by a cylinder and a fitted
piston, Figure 1.1 (b). A system should always be defined carefully, to ensure that the
same particles are in the system at all times. All other matter which can interact with the
system is called the surroundings. The combination of the system and the surroundings
is termed the universe, used here not in a cosmological sense, but to include only the
5

system and all matter which could interact with the system. Thermodynamics and
energy conversion are concerned with changes in the system and in its interactions with
the surroundings.

State

The mass contained within a system can exist in a variety of conditions call states.
Qualitatively, the concept of state is familiar. For example, the system state of a
gasmight be described qualitatively by saying that the system is at a high temperature
and a low pressure. Values of temperature and pressure are characteristics that identify
a particular condition of the system. Thus a unique condition of the system is called a
state.

Thermodynamic Equilibrium

A system is said to be in thermodynamic equilibrium if, over a long period of time, no


change in the character or state of the system is observed.
6

Thermodynamic Properties

It is a fundamental assumption of thermodynamics that a state of thermodynamic


equilibrium of a given system may be described by a few observable characteristics
called thermodynamic properties, such as pressure, temperature, and volume.
Obviously, this approach excludes the possibility of description of the condition of the
molecules of the system, a concern that is left to the fields of statistical and quantum
mechanics and kinetic theory. Nevertheless, it is frequently useful to think of
thermodynamic phenomena in molecular terms.

The Temperature Property. Temperature is a measure of the vigor of the molecular


activity of a system. How can it be observed? A thermometer measures a system
property called temperature when it is in intimate and prolonged contact
(thermodynamic equilibrium) with the system. A mercury-in-glass thermometer, for
instance, functions by thermal expansion or contraction of mercury within a glass bulb.
The bulb must be in intimate thermal contact with the observed system so that the
temperatures of the bulb and the system are the same. As a result of the equilibrium,
elongation or contraction of a narrow column of mercury connected to the bulb
indicates the temperature change of the system with which it is in contact.

The Pressure Property. Another way to observe changes in the state of a liquid or
gaseous system is to connect a manometer to the system and observe the level of the
free surface of the manometer fluid . The manometer free surface rises or drops as the
force per unit area or pressure acting on the manometer-system interface changes.

Defining a State

It has been empirically observed that an equilibrium state of a system containing a


single phase of a pure substance is defined by two thermodynamic properties. Thus, if
we observe the temperature and pressure of such a system, we can identify when the
system is in a particular thermodynamic state.

Extensive and Intensive properties

Properties that are dependent on mass are known as extensive properties. For these
properties that indicate quantity, a given property is the sum of the the corresponding
properties of the subsystems comprising the system. Examples are internal energy and
volume. Thus, adding the internal energies and volumes of subsystems yields the
internal energy and the volume of the system, respectively.
In contrast, properties that may vary from point to point and that do not change
with the mass of the system are called intensive properties. Temperature and pressure
are well-known examples. For instance, thermometers at different locations in a system
may indicate differing temperatures. But if a system is in equilibrium, the temperatures
7

of all its subsystems must be identical and equal to the temperature of the system. Thus,
a system has a single, unique temperature only when it is at equilibrium.

Work

From basic mechanics, work, W, is defined as the energy provided by an entity that
exerts a force, F, in moving one or more particles through a distance, x. Thus work
must be done by an external agent to decrease the volume, V, of a system of molecules.
In the familiar piston-cylinder arrangement shown in Figure 1.1(b), an infinitesimal
volume change of the system due to the motion of the piston is related to the
differential work through the force-distance product:

dW = Fdx = pAdx = pdV [ft-lbf | n-m] (1.1a)

or

dw = pdv [Btu/lbm | kJ/kg] (1.1b)

where p is the system pressure, and A is the piston cross-sectional area.


Note that in Equation (1.1b), the lower case letters w and v denote work and
volume on a unit mass basis. All extensive properties, i.e., those properties of state that
are proportional to mass, are denoted by lowercase characters when on a unit mass
basis. These are called specific properties. Thus, if V represents volume, then v denotes
specific volume. Although work is not a property of state, it is dealt with in the same
way.
Also note that the English units of energy in Equation (1.1a) are given in
mechanical units. Alternately, the British Thermal unit [Btu] may be used, as in
Equation (1.1b). The two sets of units are related by the famous conversion factor
known as the mechanical equivalent of heat, 778 ft-lbf/Btu. The student should pay
close attention to the consistency of units in all calculations. Conversion factors are
frequently required and are not explicitly included in many equations. For the
convenience of the reader, Appendix A lists physical constants and conversion factors.
When work decreases the volume of a system, the molecules of the system move
closer together. The moving molecules then collide more frequently with each other
and with the walls of their container. As a result, the average forces (and hence
pressures) on the system boundaries increase. Thus the state of the system may be
changed by work done on the system.

Heat

Given a system immersed in a container of hot fluid, by virtue of a difference in


temperature between the system and the surrounding fluid, energy passes from the fluid
to the system. We say that heat, Q [Btu | kJ], is transferred to the system. The system is
8

observed to increase in temperature or to change phase or both. Thus heat transfer to


or from the system, like work, can also change the state of the matter within the
system.
When the system and the surrounding fluid are at the same temperature, no heat
is transferred. In this case the system and surroundings are said to be in thermal
equilibrium. The term adiabatic is used to designate a system in which no heat crosses
the system boundaries. A system is often approximated as an adiabatic system if it is
well insulated.

Heat and Work Are Not Properties

Mechanics teaches that work can change the kinetic energy of mass and can change the
elevation or potential energy of mass in a gravitational field. Thus work performed by
an outside agent on the system boundary can change the energy associated with the
particles that make up the system. Likewise, heat is energy crossing the boundary of a
system, increasing or decreasing the energy of the molecules within. Thus heat and
work are not properties of state but forms of energy that are transported across system
boundaries to or from the environment. They are sometimes referred to as energy in
transit. Energy conversion engineering is vitally concerned with devices that use and
create energy in transit.

Internal Energy and The First Law of Thermodynamics

A property of a system that reflects the energy of the molecules of the system is called
the internal energy, U. The Law of Conservation of Energy states that energy can be
neither created nor destroyed. Thus the internal energy of a system can change only
when energy crosses a boundary of the system, i.e., when heat and/or work interact
with the system. This is expressed in an equation known as the First Law of
Thermodynamics. In differential form the First Law is:

du = dq – dw [Btu/lbm | kJ/kg] (1.2)

Here, u is the internal energy per unit mass, a property of state, and q and w are,
respectively, heat and work per unit mass. The differentials indicate infinitesimal
changes in quantity of each energy form. Here, we adopt the common sign convention
of thermodynamics that both the heat entering the system and work done by the system
are positive. This convention will be maintained throughout the text. Thus Equation
(1.2) shows that heat into the system (positive) and work done on the system (negative)
both increase the system’s internal energy.

Cyclic Process

A special and important form of the First Law of Thermodynamics is obtained by


9

integration of Equation (1.2) for a cyclic process. If a system, after undergoing


arbitrary change due to heat and work, returns to its initial state, it is said to have
participated in a cyclic process. The key points are: (1) the integral of any state
property differential is the difference of its limits, and (2) the final state is the same as
the initial state (hence there is no change in internal energy of the system)

Šdu = uf – ui = 0

where the special integral sign indicates integration over a single cycle and subscripts i
and f designate, respectively, initial and final states. As a consequence, the integration
of Equation (1.2) for a cycle yields:

Šdq = Šdw [Btu/lbm | kJ/kg] (1.3)

This states that the integral of all transfers of heat into the system, taking into account
the sign convention, is the integral of all work done by the system. The latter is the net
work of the system. The integrals in Equation (1.3) may be replaced by summations for
a cyclic process that involves a finite number of heat and work terms. Because many
heat engines operate in cyclic processes, it is sometimes convenient to evaluate the net
work of a cycle using Equation (1.3) with heat additions and losses rather than using
work directly.

Arbitrary Process of a System

Another important form of the First Law of Thermodynamics is the integral of


Equation (1.2) for an arbitrary process involving a system:

q = uf – ui + w [Btu/lbm | kJ/kg] (1.4)

where q and w are, respectively, the net heat transferred and net work for the process,
and uf and ui are the final and initial values of the internal energy. Equation (1.4), like
Equation (1.2), shows that a system that is rigid (w = 0) and adiabatic (q = 0) has an
unchanging internal energy. It also shows, like Equation (1.3), that for a cyclic process
the heat transferred must equal the work done.

Reversibility and Irreversibility

If a system undergoes a process in which temperature and pressure gradients are always
small, the process may be thought of as a sequence of near-equilibrium states. If each of
the states can be restored in reverse sequence, the process is said to be internally
reversible. If the environmental changes accompanying the process can also be reversed
in sequence, the process is called externally reversible. Thus, a reversible process is
10

one that is both internally and externally reversible. The reversible process becomes
both a standard by which we measure the success of real processes in avoiding losses
and a tool that we can use to derive thermodynamic relations that approximate reality.
All real processes fail to satisfy the requirements for reversibility and are therefore
irreversible. Irreversibility occurs due to temperature, pressure, composition, and
velocity gradients caused by heat transfer, solid and fluid friction, chemical reaction,
and high rates of work applied to the system. An engineer’s job frequently entails
efforts to reduce irreversibility in machines and processes.

Entropy and Enthalpy

Entropy and enthalpy are thermodynamic properties that, like internal energy, usually
appear in the form of differences between initial and final values. The entropy change of
a system, ªs [Btu/lbm-R | kJ/kg-K], is defined as the integral of the ratio of the system
differential heat transfer to the absolute temperature for a reversible thermodynamic
path, that is, a path consisting of a sequence of well-defined thermodynamic states. In
differential form this is equivalent to:

ds = dqrev /T [Btu/lbm-R | kJ/kg-K] (1.5)

where the subscript rev denotes that the heat transfer must be evaluated along a
reversible path made up of a sequence of neighboring thermodynamic states. It is
implied that, for such a path, the system may be returned to its condition before the
process took place by traversing the states in the reverse order.

An important example of the use of Equation (1.5) considers a thermodynamic cycle


composed of reversible processes. The cyclic integral, Equation (1.3), may then be used
to show that the net work of the cycle is:

wn = Šdq = ŠTds [Btu/lbm | kJ/kg]

This shows that the area enclosed by a plot of a reversible cyclic process on a
temperature-entropy diagram is the net work of the cycle.

The enthalpy, h, is a property of state defined in terms of other properties:

h = u + pv [Btu/lbm | kJ/kg] (1.6)

where h, u and v are, respectively, the system specific enthalpy, specific internal
energy, and specific volume, and p is the pressure.
11

Two other important forms of the First Law make use of these properties.
Substitution of Equations (1.1) and (1.5) in Equation (1.2) yields, for a reversible
process

Tds = du + pdv [Btu/lbm | kJ/kg] (1.7)

and differentiation of Equation (1.6), combined with elimination of du in Equation


(1.7), gives

Tds = dh - vdp [Btu/lbm | kJ/kg] (1.8)

Equations (1.7) and (1.8) may be regarded as relating changes in entropy for reversible
processes to changes in internal energy and volume in the former and to changes in
enthalpy and pressure in the latter. The fact that all quantities in these equations are
properties of state implies that entropy must also be a thermodynamic property.
Because entropy is a state property, the entropy change between two equilibrium
states of a system is the same for all processes connecting them, reversible or
irreversible. Figure 1.2 depicts several such processes 1-a-b-c-2, 1-d-2, and a sequence
of nonequilibrium states not describable in thermodynamic terms indicated by the
dashed line (an irreversible path). To use Equation (1.5) directly or as in Equations
(1.7) and (1.8), a reversible path must be employed. Because of the path independence
of state property changes, any reversible path will do. Thus the entropy change, s2 – s1,
12

may be evaluated by application of Equations (1.5), (1.7), or (1.8) to either of the


reversible paths shown in Figure 1.2 or to any other reversible path connecting states 1
and 2.

The Second Law of Thermodynamics

While Equation (1.5) may be used to determine the entropy change of a system, the
Second Law of Thermodynamics, is concerned with the entropy change of the universe,
i.e., of both the system and the surroundings. Because entropy is an extensive property,
the entropy of a system is the sum of the entropy of its parts. Applying this to the
universe, the entropy of the universe is the sum of the entropy of the system and its
surroundings. The Second Law may be stated as "The entropy change of the universe is
non-negative":

ÎSuniv $ 0 [Btu/R | kJ/K] (1.9)

Note that the entropy change of a system may be negative (entropy decrease) if the
entropy change of its environment is positive (entropy increase) and sufficiently large
that inequality (1.9) is satisfied.
As an example: if the system is cooled, heat is transferred from the system. The
heat flow is therefore negative, according to sign convention. Then, according to
Equation (1.5), the system entropy change will also be negative; that is, the system
entropy will decrease. The associated heat flow, however, is into the environment,
hence positive with respect to the environment (considered as a system). Then Equation
(1.5) requires that the environmental entropy change must be positive. The Second Law
implies that, for the combined process to be possible, the environmental entropy change
must exceed the magnitude of the system entropy change.
The First Law of Thermodynamics deals with how the transfer of heat influences
the system internal energy but says nothing about the nature of the heat transfer, i.e.,
whether the heat is transferred from hotter or colder surroundings. Experience tells us
that the environment must be hotter to transfer heat to a cooler object, but the First
Law is indifferent to the condition of the heat source. However, calculation of the
entropy change for heat transfer from a cold body to a hot body yields a negative
universe entropy change, violates the Second Law, and is therefore impossible. Thus
the Second Law provides a way to distinguish between real and impossible processes.
This is demonstrated in the following example:

EXAMPLE 1.1

(a) Calculate the entropy change of an infinite sink at 27°C temperature due to heat
transfer into the sink of 1000 kJ.
(b) Calculate the entropy change of an infinite source at 127°C losing the same amount
of heat.
13

(c) What is the entropy change of the universe if the aforementioned source supplies
1000 kJ to the sink with no other exchanges?
(d) What are the entropy changes if the direction of heat flow is reversed and the
source becomes the sink?

Solution
(a) Because the sink temperature is constant, Equation (1.5) shows that the entropy
change of the sink is the heat transferred reversibly divided by the absolute temperature
of the sink. This reversible process may be visualized as one in which heat is transferred
from a source which is infinitesimally hotter than the system:

ªSsink = 1000/(273 + 27) = + 3.333 kJ/K.

(b) Treating the source in the same way:

ªSsource = – 1000/(273 + 127) = – 2.5 kJ/K.

(c) Because the entropy change of the universe is the sum of the entropy changes of
source and sink, the two acting together to transfer 1000kJ irreversibly give:

ªSuniverse = 3.333 – 2.5 = +0.833 kJ/K > 0

which satisfies the Second Law inequality (1.9).

(d) A similar approach with the direction of heat flow reversed, taking care to observe
the sign convention, gives

ÎSsink = (– 1000 )/(273 + 27) = – 3.333 kJ/K

ÎSsource = (1000)/(273 + 127) = + 2.555 kJ

ÎSuniv = – 3.333 + 2.5 = – 0.833 kJ/K.

Thus we see that heat flow from a low to a high temperature reduces the entropy of the
universe, violates the Second Law, and therefore is not possible.
____________________________________________________________________
Parts a, b, and c of Example 1.1 show that the entropy change of the universe depends
on the temperature difference driving the heat transfer process:

ÎSuniv = Q(1/Tsink – 1/ Tsource) = Q( Tsource – Tsink) / Tsource Tsink

Note that if the temperature difference is zero, the universe entropy change is also zero
and the heat transfer is reversible. For finite positive temperature differences, ÎSuniv
14

exceeds zero and the process is ireversible. As the temperature difference increases,
ÎSuniv increases. This exemplifies the fact that the entropy change of the universe
produced by a process is a measure of the irreversibility of the process.
For an isolated system, there is no change in the entropy of the surroundings.
Hence the system entropy change is the entropy change of the universe and therefore
must be non-negative. In other words, the entropy of an isolated system can only
increase or at best stay constant.

1.3 Control Volumes and Steady Flows

In many engineering problems it is preferrable to deal with a flow of fluid particles as


they pass through a given region of space rather than following the flow of a fixed
collection of particles. Thus, putting aside the system concept (fixed collection) for the
moment, consider a volume with well-defined spatial boundaries as shown in Figure
1-3. This is called a control volume. Mass at state 1 enters at a rate m1 and leaves at
state 2 with mass flow m2. If one mass flow rate exceeds the other, mass either
accumulates in the volume or is depleted. The important special case of steady flow, in
which no accumulation or depletion of mass occurs in the control volume, is considered
here. In steady flow, the conservation of mass requires equal mass flows in and out, i.e.,
m1 = m2, [lbm /s, | kg /s].
If Q-dot is the rate of heat flow into the control volume and W-dot is the rate at
which shaft work is delivered from the control volume to the surroundings,
conservation of energy requires that the excess of inflowing heat over outgoing work
equal the net excess of the energy (enthalpy) flowing out of the ports, i.e.,
[Btu/s | kJ/s] (1.10)
15

where summations apply to inflows i and outflows o, and where other types of energy
terms, such as kinetic and potential energy flows, are assumed negligible. For clarity,
the figure shows only one port in and one port out. Kinetic and potential energy terms
may be added analogous to the enthalpy termsat each port, if needed.
Equation (1.10) may be the most important and frequently used equation in this
book. Mastery of its use is therefore essential. It is known as the steady flow form of
the First Law of Thermodynamics. It may be thought of as a bookkeeping relation for
keeping track of energy crossing the boundaries of the control volume.
The Second Law of Thermodynamics applied to steady flow through an adiabatic
control volume requires that m2s2 $ m1s1, or by virtue of mass conservation:

s2 $ s1 [Btu/lbm-R | kJ/kg-K] (1.11)

That is, because entropy cannot accumulate within the control volume in a steady flow,
the exit entropy must equal or exceed the inlet entropy. In steady flows, heat transfer
can increase or decrease the entropy of the flow, depending on the direction of heat
transfer, as long as the entropy change of the surroundings is such that the net effect is
to increase the entropy of the universe.
We will often be concerned with adiabatic flows. In the presence of fluid friction
and other irreversibilities, the exit entropy of an adiabatic flow exceeds its inlet entropy.
Adiabatic flows that have no irreversibilities also have no entropy change and therefore
are called isentropic flows.

1.4 Properties of Vapors: Mollier and T-s Diagrams

When heated, liquids are transformed into vapors. The much different physical
character of liquids and vapors makes engines in which phase change takes place
possible. The Newcomen atmospheric engine, for instance condensed steam to liquid
water in a piston-cylinder enclosure to create a partial vacuum. The excess of
atmospheric pressure over the low pressure of the condensed steam, acting on the
opposite face of the piston, provided the actuating force that drove the first successful
engines in the early eighteenth century. In the latter half of the eighteenth century,
engines in which work was done by steam pressure on the piston rather than by the
atmosphere, replaced Newcomen-type engines. Steam under pressure in reciprocating
engines was a driving force for the industrial revolution for about two centuries. By the
middle of the twentieth century, steam turbines and diesel engines had largely replaced
the steam engine in electric power generation, marine propulsion, and railroad
locomotives. .
Figure 1.4 shows typical saturation curves for a pure substance plotted in
temperature and entropy coordinates. A line of constant pressure (an isobar) is shown
in which the subcooled liquid at state 1 is heated, producing increases in entropy,
temperature, and enthalpy, until the liquid is saturated at state 2. Isobars in the
16

subcooled region of the diagram lie very close to the saturated liquid curve. The
separation of the two is exaggerated for clarity.
Once the substance has reached state 2, further transfer of heat fails to increase the
system temperature but is reflected in increased enthalpy and entropy in a vaporization
or boiling process. During this process the substance is converted from a saturated
liquid at state 2 to a mixture of liquid and vapor, and finally to a saturated vapor at
state 3. The enthalpy difference between the saturation values, h3 – h2, is called the
enthalpy of vaporization or heat of vaporization.
Continued addition of heat to the system, starting at state 3, superheats the steam
to state 4, again increasing temperature, enthalpy, and entropy.
Several observations about the isobaric process may be made here. Equation (1.5)
and Figure 1.4 show that the effect of adding heat is to always increase system entropy
and that of cooling to always decrease it. A similar conclusion can be drawn from
Equation (1.10) regarding heat additions acting to increase enthalpy flow through a
control volume in the absence of shaft work.
A measure of the proximity of a superheated state (state 4 in the figure) to the
saturated vapor line is the degree of superheat. This is the difference between the
temperature T4 and the saturated vapor temperature T3, at the same pressure. Thus the
degree of superheat of superheated state 4 is T 4 - T 3.
In the phase change from state 2 to state 3, the temperature and pressure give no
indication of the relative quantities of liquid and vapor in the system. The quality x is
defined as the ratio of the mass of vapor to the mass of the mixture of liquid and vapor
at any point between the saturation curves at a given pressure. By virtue of this
definition, the quality varies from 0 for a saturated liquid to 1 for a saturated vapor.
17

Because extensive properties are proportional to mass, they vary directly with the
vapor quality in the mixed region. The entropy, for example, varies from the entropy of
the saturated liquid sl at state 2 to the saturated vapor entropy sv at state 3 in
accordance with the following quality equation:

s = sl + x(sv – sl) [Btu/lbm-R | kJ/kg-K] (1.12)

where s is the entropy per unit mass. Other extensive properties such as enthalpy and
volume vary with quality in the same way.
A variable closely related to the quality is moisture fraction (both quality and
moisture fraction can be expressed as percentages). Moisture fraction, M, is defined as
the ratio of the mass of liquid to the total mass of liquid and vapor. It can be easily
shown that the sum of the quality and the moisture fraction of a mixture is one.
A Mollier chart, a diagram with enthalpy as ordinate and entropy as abscissa, is
much like the temperature-entropy diagram. A Mollier diagram for steam is included in
Appendix B. An isobar on a Mollier chart, unlike that on a T-s diagram, has a
continuous slope. It shows both enthalpy and entropy increasing monotonically with
heat addition. Such a diagram is frequently used in energy conversion and other areas
because of the importance of enthalpy in applying the steady-flow First Law.

1.5 Ideal Gas Basics

Under normal ambient conditions, the average distance between molecules in gases is
large, resulting in negligible influences of intermolecular forces. In this case, molecular
behavior and, therefore, system thermodynamics are governed primarily by molecular
translational and rotational kinetic energy. Kinetic theory or statistical thermodynamics
may be used to derive the ideal gas or perfect gas law:

pv = RT [ft-lbf /lbm | kJ/kg] (1.13)

where p [lbf /ft2 | kN/m2], v [ft3/lbm | m3/kg] and T [°R | °K] are pressure, specific
volume, and temperature respectively and R [ft-lbf /lbm-°R | kJ/kg-°K] is the ideal gas
constant. The gas constant R for a specific gas is the universal gas constant R divided
by the molecular weight of the gas.
Thus, the gas constant for air is (1545 ft-lbf /lb-mole-°R) / (29 lbm/lb-mole) =
53.3 ft-lbf /lbm-°R in the English system and (8.31 kJ/kg-mole-°K) / (29 kg/kg-mole)
= 0.287 kJ/kg-°K in SI units.
The specific heats or heat capacities at constant volume and at constant pressure,
respectively, are:

cv = (Mu / MT)v [Btu/lbm-°R | kJ/kg-°K] (1.14)

and
18

cp = (Mh /MT)p [Btu/lbm-°R | kJ/kg-°K] (1.15)

As thermodynamic properties, the heat capacities are, in general, functions of two


other thermodynamic properties. For solids and liquids, pressure change has little
influence on volume and internal energy, so that to a very good approximation: cv = cp.
A gas is said to be thermally perfect if it obeys Equation (1.13) and its internal
energy, enthalpy, and heat capacities are functions of temperature only. Then

du = cv(T) dT [Btu/lbm | kJ/kg] (1.16)


and
dh = cp(T) dT [Btu/lbm | kJ/kg] (1.17)

A gas is said to be calorically perfect if in addition to being thermally perfect it


also has constant heat capacities. This is reasonably accurate at low and moderate
pressures and at temperatures high enough that intermolecular forces are negligible but
low enough that molecular vibrations are not excited and dissociation does not occur.
For air, vibrational modes are not significantly excited below about 600K, and
dissociation of oxygen does not occur until the temperature is above about 1500K.
Nitrogen does not dissociate until still higher temperatures. Excitation of molecular
vibrations causes specific heat to increase with temperature increase. Dissociation
creates further increases in heat capacities, causing them to become functions of
pressure.
It can be shown (see Exercise 1.4) that for a thermally perfect gas the heat
capacities are related by the following equation:

cp = cv + R [Btu/lbm-R | kJ/kg-K] (1.18)

This relation does not apply for a dissociating gas, because the molecular weight of the
gas changes as molecular bonds are broken. Note the importance of assuring that R and
the heat capacities are in consistent units in this equation.
Another important gas property is the ratio of heat capacities defined by k = cp /cv.
It is constant for gases at room temperatures but decreases as vibrational modes
become excited. The importance of k will be seen in the following example.

EXAMPLE 1.2

(a) Derive an expression for the entropy change of a system in terms of pressure and
temperature for a calorically perfect gas.
(b) Derive a relation between p and T for an isentropic process in a calorically perfect
gas.

Solution
(a) For a reversible process, Equation (1.8) gives Tds = dh - vdp. Dividing by T and
applying the perfect gas law gives ds = cp dT/T - Rdp/p.
19

Then integration between states 1 and 2 yields

s2 - s1 = cp ln(T2 /T1) - R ln( p2 /p1)

(b) For an isentropic process, s2 = s1. Then the above equation gives

T2 /T1 = (p2/p1)(R/cp)

But R/cp = (cp - cv )/cp = (k - 1)/k. Hence T2 /T1 = (p2 /p1)(k - 1)/k.
____________________________________________________________________

This and other important relations for an isentropic process in a calorically perfect gas
are summarized as follows

T2 /T1 = (p2 /p1)(k - 1)/k [dl] (1.19)

T2 /T1 = (v2 /v1)(k - 1) [dl] (1.20)

p2 /p1 = (v1 /v2)k [dl] (1.21)

These relations show that the ratio of heat capacities governs the variation of
thermodynamic properties in an isentropic process. For this reason the ratio of heat
capacities is sometimes called the isentropic exponent.

1.6 Fundamentals of Fluid Flow

Almost all energy conversion devices involve the flow of some form of fluid. Air, liquid
water, steam, and combustion gases are commonly found in some of these devices.
Here we review a few of the frequently used elementary principles of fluid flow.
The volume flow rate, Q [ft3/s | m3/s] at which a fluid flows across a surface is the
product of the area, A [ft2 | m2], of the surface and the component of velocity normal to
the area, V [ft/s | m/s]. The corresponding mass flow rate is the ratio of the volume rate
and the specific volume, v [ft3/lbm | m3/kg]:

m = AV/v = Q/v [lbm /s | kg /s] (1.22)

Alternatively the flow rate can be expressed in terms of the reciprocal of the specific
volume, the density, , [lbm /ft3 | kg /m3]:

m = AV, = Q, [lbm /s | kg /s] (1.23)


20

The first important principle of fluid mechanics is the conservation of mass, a


principle that we have already used in Section 1.3. For a steady flow, the net inflow to a
control volume must equal the net outflow. Any imbalance between the inflow and
outflow implies an accumulation or a reduction of mass within the control volume, i.e.,
an unsteady flow. Given a control volume with n ports, the conservation of mass
provides an equation that may be used to solve for the nth port flow rate, given the
other n-1 flow rates. These flows may be (1) given, (2) calculated from data at the
ports using Equation (1.22) or (1.23), (3) obtained by solving n-1 other equations, or
(4) a combination of the preceding three.
For isentropic flow of an incompressible (constant density, , ) fluid, the Bernoulli
equation applies:

p1 /, + V12/2 = p2 /, + V22/2 [ft-lbf/lbm | kJ/kg] (1.24)

This is an invariant form, i.e. an equation with the same terms on both sides, p/, +
V2/2. The subscripts identify the locations in the flow where the invariants are
evaluated. The first term of the invariant is sometimes called the pressure head, and the
second the velocity head. The equation applies only in regions where there are no
irreversibilities such as viscous losses or heat transfer.
The invariant sum of the two terms on either side of Equation (1.24) may be called
the total head or stagnation head. It is the head that would be observed at a point
where the velocity approaches zero. The pressure associated with the total head is
therefore called the total pressure or stagnation pressure, po = p + ,V2/2. Each point
in the flow may be thought of as having its own stagnation pressure resulting from an
imaginary isentropic deceleration.
In the event of significant irreversibilities, there is a loss in total head and the
Bernoulli equation may be generalized to:

p1 /, + V12/2 = p2/, + V22/2 + loss [ft-lbf/lbm | kJ/kg] (1.25a)


or
po1 /, = po2 /, + loss [ft-lbf/lbm | kJ/kg] (1.25b)

Stagnation pressure or head losses in ducts, such as due to flow turning or sudden
area change, are tabulated in reference books as fractions of the upstream velocity head
for a variety of geometries. Another example is the famous Darcy-Weisbach equation
which gives the head loss resulting from fluid friction in a pipe of constant cross-
section.

1.7 Compressible Flow

While many engineering analyses may reasonably employ incompressible flow


principles, there are cases where the compressibility of gases and vapors must be
considered. These are situations where the magnitude of the kinetic energy of the flow
21

is comparable to its enthalpy such as in supersonic nozzles and diffusers, in turbines and
compressors, and in supersonic flight. In these cases the steady-flow First Law must be
generalized to include kinetic energy per unit mass terms. For two ports:

[Btu /s | kJ/s] (1.26a)

Care should be taken to assure consistency of units, because enthalpy is usually stated
in thermal units [Btu /lbm | kJ/kg] and velocity in mechanical units [ft /s | m /s].
Another invariant of significance appears in Equation (1.26a). The form

ho = h + V2/2 [Btu/lbm | kJ/kg] (1.27)

is seen to be invariant in applications where heat transfer and shaft work are
insignificant. The invariant, ho, is usually given the name stagnation enthalpy because it
is the enthalpy at a point in the flow (real or imagined) where velocity approaches zero.
In terms of stagnation enthalpy, Equation (1.26a) may be rewritten as

[Btu/s | kJ/s] (1.26b)

where conservation of mass with steady flow through two ports has been assumed.
Writing dho = cp dTo with cp constant, we get

ho2 - ho1 = cp(To2 - To1)

Combining this with Equation (1.27), we are led to define another invariant, the
stagnation temperature for a calorically perefect gas:

To = T + V2/2cp [ R | K] (1.28)

The stagnation temperature may be regarded as the temperture at a real or imaginary


point where the gas velocity has been brought to zero adiabatically. For this special
case of a constant heat capacity, Equation (1.26b) may be written as

[Btu /s | kJ/s] (1.26c)

In both incompressible and compressible flows, the mass flow rates at all stations
in a streamtube are the same. Because the specific volume and density are constant in
incompressible flow, Equation (1.22) shows that the volume flow rates are the same at
all stations also. However for compressible flow, Equation (1.23) shows that density
change along a streamtube implies volume flow rate variation. Thus, while it is
frequently convenient to think and talk in terms of volume flow rate when dealing with
incompressible flows, mass flow rate is more meaningful in compressible flows and in
general.
22

A measure of the compressiblity of a flow is often indicated by a Mach number, defined


as the dimensionless ratio of a flow velocity to the local speed of sound in the fluid. For
ideal gases the speed of sound is given by

a = (kp/,)½ = (kRT)½ [ft /s | m /s] (1.29)

Compressible flows are frequently classified according to their Mach number:

M=0 Incompressible
0<M<1 Subsonic
M=1 Sonic
M>1 Supersonic

Studies of compressible flows show that supersonic flows have a significantly different
physical character than subsonic and incompressible flows. For example, the velocity
fields in subsonic flows are continuous, whereas discontinuities known as shock waves
are common in supersonic flows. Thus the student should not be surprised to find that
different relations hold in supersonic flows than in subsonic flows.

1.8 Energy Clasification

Energy exists in a variety of forms. All human activities involve conversion of energy
from one form to another. Indeed, life itself depends on energy conversion processes.
The human body, through complex processes, transforms the chemical energy stored in
food into external motion and work produced by muscles as well as electrical impulses
that control and activate internal functions.
It is instructive to examine some of the processes for transformation between
technically important forms of energy. Table 1.2 shows a matrix of energy forms and
the names of some associated energy converters.

Table 1.2 Energy Transformation Matrix


From: To: Thermal Energy Mechanical Energy Electrical Energy

Chemical Energy Furnace Diesel engine Fuel cell

Thermal Energy Heat exchanger Steam turbine Thermocouple

Mechanical Energy Refrigerator, heat Gearbox Electrical generator


pump

Nuclear Energy Fission reactor Nuclear steam turbine Nuclear power plant

The table is far from complete, and other energy forms and energy converters
could readily be added. However, it does include the major energy converters of
interest to mechanical engineers. It is a goal of this book to present important aspects
of the design, analysis, performance, and operation of most of these devices.
One of the major criteria guiding the design of energy conversion systems is
23

efficiency. Each of the conversions in Table 1.2 is executed by a device that operates
with one or more relevant efficiencies. The following section explores some of the
variety of definitions of efficiency used in design and performance studies of energy
conversion devices.

1.9 Efficiencies

Efficiency is a measure of the quality of an operation or of a characteristic of a device.


Several types of efficiencies are widely used. It is important to clearly distinguish
among them. Note that the terms work and power are equally applicable here.
The efficiency of a machine that transmits mechanical power is measured by its
mechanical efficiency, the fraction of the power supplied to the transmission device
that is delivered to another machine attached to its output, Figure 1.5(a). Thus a gear-
box for converting rotational motion from a power source to a device driven at another
speed dissipates some mechanical energy by fluid and/or dry friction, with a consequent
loss in power transmitted to the second machine. The efficiency of the gearbox is the
ratio of its power output to the power input, a value less than one. For example, a
turboprop engine with a gearbox efficiency of 0.95 will transmit only 95% of its power
output to its propeller.
24

Another type of efficiency that measures internal losses of power is used to


indicate the quality of performance of turbomachines such as pumps, compressors and
turbines. These devices convert flow energy to work (power), or vice versa. Here the
efficiency compares the output with a theoretical ideal in a ratio [Figure 1.5(b)]. The
resulting efficiency ranges from 0 to 1 as a measure of how closely the process
approaches a relevant isentropic process. A turbine with an efficiency of 0.9, will, for
example, deliver 90% of the power of a perfect (isentropic) turbine operating under the
same conditions. This efficiency, sometimes referred to as isentropic efficiency or
turbine efficiency, will be considered in more detail in the next chapter.
Another form of isentropic efficiency, sometimes called compressor efficiency (or
pump efficiency), is defined for compressors (or pumps). It is the ratio of the isentropic
work to drive the compressor (or pump) to the actual work required. Because the
actual work required exceeds the isentropic work, this efficiency is also less than to 1.
A third type of efficiency compares the magnitude of a useful effect to the cost of
producing the effect, measured in comparable units. An example of this type of
efficiency compares the net work output, wn, of a heat engine to the heat supplied, qa,
to operate the engine. This is called the to (th = wn /qa) [Figure 1.5(c)]. For example,
the flow of natural gas to an electrical power plant provides a chemical energy flow
rate or heat flow rate to the plant that leads to useful electric power output. It is known
from basic thermodynamics that this efficiency is limited by the Carnot efficiency, as
will be discussed in the next section.
Another example of this type of efficiency as applied to refrigerators and heat
pumps [Figure 1.5(d)] is called the coefficient of performance, COP. In this case the
useful effect is the rate of cooling or heating, and the cost to produce the effect is the
power supplied to the device. The term "coefficient of performance" is used instead of
efficiency for this measure of quality because the useful effect usually exceeds the cost
in comparable units of measure. Hence, unlike other efficiencies, the COP can exceed
unity. As seen in Figure 1.5(d), there are two definitions for COP, one for a refrigerator
and another for a heat pump. It may be shown using the First Law of Thermodynamics,
that a simple relationship exists between the two definitions: COPhp = COPrefr + 1.

1.10 The Carnot Engine

On beginning the study of the energy conversion ideas and devices that will serve us in
the twenty-first century, it is appropriate to review the theoretical cycle that stands as
the ideal for a heat engine. The ideas put forth by Sadi Carnot in 1824 in his
“Reflections on the Motive Power of Heat” (see Historical Bibliography) expressed the
content of the Second Law of Thermodynamics relevant to heat engines, which, in
modern form (attributed to Kelvin and Planck) is: “It is impossible for a device which
operates in a cycle to receive heat from a single source and convert the heat completely
to work.” Carnot’s great work also described the cycle that today bears his name and
provides the theoretical limit for efficiency of heat engine cycles that operate between
two given temperature levels: the Carnot cycle.
The Carnot cycle consists of two reversible, isothermal processes separated by two
25

reversible adiabatic or isentropic processes, as shown in Figure 1.6. All of the heat
transferred to the working fluid is supplied isothermally at the high temperature
TH = T3, and all heat rejected is transferred from the working medium at the low
temperature TL = T1. No heat transfer takes place, of course, in the isentropic
processes. It is evident from Equation (1.5) and the T-s diagram that the heat added is
T3(s3 - s2 ), the heat rejected is T1 (s1 - s4 ), and, by the cyclic integral relation, the net
work is T3(s3 - s2 ) + T1 (s1 - s4 ). The thermal efficiency of the Carnot cycle, like that of
other cycles, is given by wn / qa and can be expressed in terms of the high and low cycle
temperatures as :
26

because both isothermal processes operate between the same entropy limits.
The Carnot efficiency equation shows that efficiency rises as TL drops and as
TH increases. The message is clear: a heat engine should operate between the widest
possible temperature limits. Thus the efficiency of a heat engine will be limited by the
maximum attainable energy-source temperature and the lowest available heat-sink
temperature.
Students are sometimes troubled by the idea of isothermal heat transfer processes
because they associate heat transfer with temperature rise. A moments reflection,
however, on the existence of latent heats–e.g., the teakettle steaming on the stove at
constant temperature-makes it clear that one should not always associate heat transfer
with temperature change.
It is important here, as we start to consider energy conversion devices, to recall the
famous Carnot Theorem, the proof of which is given the most thermodynamics texts.
It states that it is impossible for any engine operating in a cycle between two reservoirs
at different temperatures to have an efficiency that exceeds the Carnot efficiency
corresponding to those temperatures. It can also be shown that all reversible engines
operating between two given reservoirs have the same efficiency and that all irreversible
engines must have lower efficiencies. Thus the Carnot efficiency sets an upper limit on
the performance of heat engines and therefore serves as a criterion by which other
engines may be judged.

1.11 Additional Second-Law Considerations

The qualitative relationship between the irreversibility of a process and the entropy
increase of the universe associated with it was considered in section 1.2. Let us now
consider a quantitative approach to irreversibility and apply it to a model of a power
plant.

Reversible Work

Instead of comparing the work output of the power plant with the energy supplied from
fuel to run the plant, it is instructive to compare it with the maximum work achievable
by a reversible heat engine operating between the appropriate temperature limits, the
reversible work. It has been established that any reversible engine would have the same
efficiency as a Carnot engine. The Carnot engine provides a device for determining the
reversible work associated with a given source temperature, TH, and a lower sink
temperature,TL.
27

Irreversibility

The irreversibility, I, of a process is defined as the difference between the reversible


work and the actual work of a process:
I = Wrev - Wact [Btu | kJ]

It is seen that the irreversibility of a process vanishes when the actual work is the same

as that produced by an appropriate Carnot engine. Moreover, the irreversibility of a


non-work-producing engine is equal to the reversible work. It is clear that I $ 0,
because no real engine can produce more work than a Carnot engine operating between
the same limiting temperatures.

Second-Law Efficiency

We can also define a “second-law efficiency,” II, as the ratio of the actual work of a
process to the reversible work:

[dl]

This is an efficiency that is limited to 100%, as opposed to the thermal efficiency of a


heat engine, sometimes referred to as a “first-law efficiency,” which may not exceed
that of the appropriate Carnot engine. Note that an engine which has no irreversibility
is a reversible engine and has a second-law efficiency of 100%.

A Power Plant Model

Let us consider a model of a power plant in which a fuel is burned at a high


temperature, TH, in order to transfer heat to a working fluid at an intermediate
temperature, TINT. The working fluid, in turn, is used in an engine to produce work and
reject heat to a sink at the low temperature, TL. Figure 1.7 presents a diagram of the
model that shows explicitly the combustion temperature drop from the source
temperature, TH, to the intermediate temperature, TINT, the actual work-producing
engine, and a Carnot engine used to determine the reversible work for the situation.
The Carnot engine has an efficiency of C = 1 - TL /TH and develops work in the
following amount:

Wrev = WC = QIN - QC = CQIN [Btu|kJ]

Suppose that the engine we are considering is a Carnot engine that operates from the
intermediate source at TINT. Its efficiency and work output are, respectively,
I = 1 - TL /TINT and Wact = QIN – QI = I QIN . The irreversibility, I, of the power plant
is then
28

I = Wrev - Wact = (C - I )QIN

= [1 - TL /TH - ( 1 - TL /TINT)]QIN

= TLQIN ( TH - TINT ) /TINTTH [Btu|kJ]

and the second-law efiiciency is:

II = Wact /Wrev = (1 - TL/TINT ) / (1 - TL/TH) [dl]

Note that when TINT = TH, the irreveribility vanishes and the second-law efficiency
becomes 100%. Also, when TINT = TL, the irreversibility is equal to the reversible work
of the Carnot engine and the second- law efficiency is zero. The latter condition
indicates that a pure heat transfer process or any process that produces no useful work
causes a loss in the ability to do work in the amount of Wrev . Thus the reversible work
associated with the extremes of a given process is a measure of how much capability to
do work can be lost, and the irreversibility is a measure of how much of that work-
producing potential is actually lost. The following example illustrates these ideas.

EXAMPLE 1.3

Through combustion of a fossil fuel at 3500°R, an engine receives energy at a rate of


3000 Btu/s to heat steam to 1500°R. There is no energy loss in the combustion
process. The steam, in turn, produces 1000 Btu/s of work and rejects the remaining
energy to the surroundings at 500°R.

(a) What is the thermal efficiency of the plant?


(b) What are the reversible work and the Carnot efficiency corresponding to the source
29

and sink temperatures?


(c) What is the irreversibility?
(d) What is the second-law efficiency?
(e) What would the irreversibility and the second-law efficiency be if the working fluid
were processed by a Carnot engine rather than by the real engine?

Solution
(a) The thermal efficiency, or first-law efficiency, of the plant is I = Wact/QIN =
1000/3000 = 0.333 or 33.3%.
(b) The relevant Carnot efficiency is 1 - 500/3500 = 0.857, or 85.7%. The engine’s
reversible work is then CQIN = 0.857(3000) = 2571 Btu/s.

(c) The plant irreversibility is 2571 - 1000 = 1571 Btu/s.

(d) The second-law efficiency is then


I /C = 0.333/0.857 = 0.389, or 38.9%
or
Wact /Wrev = 1000/2571 = 0.389, or 38.9%.

(e) The Carnot efficiency corresponding to the maximum temperature of the working
fluid is 1 - 500/1500 = 0.667 or 66.7%. The second-law efficiency for this system is
then 0.667/0.857 = 0.778, compared with 0.389 for the actual engine. The actual work
produced by the irreversibly heated Carnot engine is 0.667(3000) = 2001 Btu/s, and its
irreversibility is then I = 2571 -2001 = 570 Btu/s.
___________________________________________________________________

In summary, the thermal efficiency, or first-law efficiency, of an engine is a


measure of how well the engine converts the energy in its fuel to useful work. It says
nothing about energy loss, because energy is conserved and cannot be lost: it can only
be transformed. The second-law efficiency, on the other hand, recognizes that some of
the energy of a fuel is not available for conversion to work in a heat engine and
therefore assesses the ability of the engine to convert only the available work into useful
work. This is a reason why some regard the second-law efficiency as more significant
than the more commonly used first-law efficiency.

Bibliography and References

1. Van Wylen, Gordon J., and Sonntag, Richard E., Fundamentals of Classical
Thermodynamics, 3rd ed. New York: Wiley, 1986.

2. Balmer, Robert, Thermodynamics. Minneapolis: West, 1990.

3. Cengel, Yunus A., and Boles, Michael A., Thermodynamics. New York: McGraw-
Hill, 1989.
30

4. Faires, Virgil Moring, Thermodynamics, 5th ed. New York: Macmillan, 1970.

5. Silver, Howard F., and Nydahl, John E., Engineering Thermodynamics.


Minneapolis: West, 1977.

6. Bathie, William W., Fundamentals of Gas Turbines. New York: Wiley, 1984

7. Wilson, David Gordon, The Design of High Efficiency Turbomachinery and Gas
Turbines. Boston: MIT Press, 1984.

8. Anderson, John D., Modern Compressible Flow. New York: McGraw-Hill, 1982.

9. Anderson, John D., Introduction to Flight. New York: McGraw-Hill, 1978.

10. Chapman, Alan J. and Walker, William F., Introductory Gas Dynamics. New York:
Holt, Rinehart, and Winston, 1971.

Historical Bibliography

1. Barnard, William N., Ellenwood, Frank E., and Hirshfeld, Clarence F., Heat-Power
Engineering. Wiley, 1926.

2. Bent, Henry, The Second Law.NewYork: Oxford University Press, 1965

3. Cummins, C. Lyle, Jr., Internal Fire, rev. ed. Warrendale, Penna.: Society of
Automotive Engineers, 1989.

4. Tann, Jennifer, The Selected Papers of Boulton and Watt. Boston: MIT Press, 1981.

5. Carnot, Sadi, Reflexions Sur la Puissance Motrice de Feu. Paris: Bachelor,1824.

6. Potter, J. H., “The Gas Turbine Cycle.” ASME Paper presented at the Gas Turbine
Forum Dinner, ASME Annual Meeting, New York, Nov. 27, 1972.

7. Grosser, Morton, Diesel: The Man and the Machine. New York: Atheneum, 1978.

8. Nitske, W. Robert, and Wilson, Charles Morrow, Rudolph Diesel: Pioneer of the
Age of Power. Norman, Okla: University of Oklahoma Press, 1965.

9. Rolt, L.T. C., and Allen, J. S., The Steam Engine of Thomas Newcomen. New York:
Moorland Publishing Co., 1977.

10. Briggs, Asa, The Power of Steam. Chicago: University of Chicago Press, 1982.
31

11. Thurston, Robert H., A History of the Growth of the Steam Engine. 1878; rpr.
Ithaca, N.Y.: Cornell University Press, 1939.

EXERCISES

1.1 Determine the entropy of steam at 1000 psia and a quality of 50%.

1.2 Show that the moisture fraction for a liquid water-steam mixture, defined as the
ratio of liquid mass to mixture mass, can be written as 1 - x, where x is steam quality.

1.3 Write expressions for the specific entropy, specific enthalpy, and specific volume as
functions of the moisture fraction. Determine the values of these properties for steam at
500°F and a moisture fraction of 0.4.

1.4 Show that for a thermally perfect gas, cp - cv = R.

1.5 During a cyclic process, 75kJ of heat flow into a system and 25kJ are rejected from
the system later in the cycle. What is the net work of the cycle.

1.6 Seventy-five kJ of heat flow into a rigid system and 25 kJ are rejected later. What
are the magnitude and sign of the change in internal energy? What does the sign
indicate?

1.7 The mass contained between an insulated piston and an insulated cylinder
decreases in internal energy by 50 Btu. How much work is involved, and what is the
sign of the work term? What does the sign indicate?

1.8 Derive Equation (1.8) from Equation (1.7).

1.9 Use Equation (1.8) to derive an expression for the finite enthalpy change of an
incompressible fluid in an isentropic process. If the process is the pressurization of
saturated water initially at 250 psia, what is the enthalpy rise, in Btu /lbm and in
ft-lbf / lbm, if the final pressure is 4000 psia? What is the enthalpy rise if the initial
pressure is 100 kPa and the final pressure is 950 kPa?

1.10 Sixty kg /s of brine flows into a device with an enthalpy of 200 kJ/kg. Brine flows
out of the other port at a flow rate of 20 kg /s. What is the net inflow? Is the system in
steady flow? Explain.

1.11 Use the steam tables in Appendices B and C to compare the heats of vaporization
at 0.01, 10, and 1000 psia. Compare the saturated liquid specific volumes at these
pressures. What do you conclude about the influence of pressure on these properties?
32

1.12 Using the heat capacity equation for nitrogen:

cp = 9.47 - 3.47 x 103/T + 1.16 x 106/T2

where cp is in Btu / lb-mole and T is in degrees Rankine (From Gordon J. Van Wylen
and Richard E. Sonntag, Fundamentals of Classical Thermodynamics, 3rd ed. New
York: Wiley, 1986). Compare the enthalpy change per mole of nitrogen between
540°R and 2000°R for nitrogen as a thermally perfect gas and as a calorically perfect
gas.

1.13 Use Equation (1.7) to derive Equation (1.20) for a calorically perfect gas.

1.14 Use Equation (1.7) to derive a relation for the entropy change as a function of
temperature ratio for a constant-volume process in a calorically perfect gas.

1.15 Use Equation (1.8) to derive a relation for the entropy change as a function of
temperature for an isobaric process in a calorically perfect gas.

1.16 A convergent nozzle is a flow passage in which area decreases in the streamwise
direction (the direction of the flow). It is used for accelerating the flow from a low
velocity to a higher velocity. Use the generalized form of the steady-flow First Law
given in Equation (1.26a) to derive an equation for the exit velocity for an adiabatic
nozzle.

1.17 Derive an equation for the pressure drop for a loss-free incompressible flow in a
varying-area duct as a function of area ratio.

1.18 Two units of work are required to transfer 10 units of heat from a refrigerator to
the environment. What is the COP of the refrigerator? Suppose that the same amount
of heat transfer instead is by a heat pump into a house. What is the heat pump COP?

1.19 A power plant delivers 100 units of work at 30% thermal efficiency. How many
heat units are supplied to operate the plant? How many units of heat are rejected to the
surroundings?

1.20 A steam turbine has an efficiency of 90% and a theoretical isentropic power of'
100 kW. What is the actual power output?

1.21 Thomas Newcomen used the fact that the specific volume of saturated liquid is
much smaller than the specific volume of saturated steam at the same pressure in his
famous "atmospheric engine." Calculate the work done on the piston by the atmosphere
if steam is condensed at an average pressure of 6 psia by cooling in a tightly fitted
piston-cylinder enclosure if the piston area is 1 ft2 and the piston stroke is 1 ft. If the
process takes place 10 times a minute, what is the power delivered? Discuss what can
33

be done to increase the power of the engine. Describe the characteristics of a


Newcomen engine that would theoretically deliver 20 horsepower.

1.22 Expand Table 1.2 to include solar and geothermal sources.

1.23 Twenty pounds of compressed air is stored in a tank at 200 psia and 80/F. The
tank is heated to bring the temperature to 155/F. What is the final tank pressure, and
how much heat was added?

1.24 Ten kilograms of compressed air is stored in a tank at 250 kPa and 50/C. The
tank is heated to bring the air temperature to 200 °C. What is the final tank pressure,
and how much heat was added?
3
1.25 An 85-ft tank contains air at 30 psia and 100/ F. What mass of air must be added
to bring the pressure to 50 psia and the temperature to 150/F?

1.26 A 20-L tank contains air at 2 bar and 300K. What mass of air must be added to
bring the pressure to 2 bar and the temperature to 375K?

1.27 Air enters a wind tunnel nozzle at 160/F, 10 atm, and a velocity of 50 ft/s. The
entrance area is 5ft2. If the heat loss per unit mass is 10 Btu/lbm and the exit pressure
and velocity are, respectively, 1.5 atm and 675 ft/s, what are the exit temperature and
area?

1.28 Air enters a wind tunnel nozzle at 90°C, 250 kPa, and a velocity of 40 m/s. The
entrance area is 3 m2. If the heat loss per unit mass is 7 kJ/kg and the exit pressure and
velocity are, respectively, 105 kPa and 250 m/s, what are the exit temperature and
area?

1.29 Air enters a wind tunnel nozzle at 160/F, 10 atm, and a velocity of 50 ft/s. The
entrance area is 5ft2. If the heat loss per unit mass is 8 Btu/lbm and the exit pressure and
temperature are, respectively, 1.25 atm and 120/F, what are the exit velocity and area?

1.30 Air enters a wind tunnel nozzle at 90/C, 250 kPa, and a velocity of 40 m/s. The
entrance area is 3 m2. If the heat loss per unit mass is 5 kJ/kg and the exit pressure and
temperature are, respectively, 120 kPa and 43°C, what are the exit velocity and area?

1.31 Sketch a Mollier diagram showing the character of three isotherms and three
isobars for a calorically perfect gas. Label each curve with a value in SI units to show
the directions of increasing temperature and pressure. Explain how the diagram would
differ if the gas were not calorically perfect.
35

CHAPTER TWO

Fundamentals of
Steam Power

2.1 Introduction

Much of the electricity used in the United States is produced in steam power plants.
Despite efforts to develop alternative energy converters, electricity from steam will
continue, for many years, to provide the power that energizes the United States and
world economies. We therefore begin the study of energy conversion systems with this
important element of industrial society.
Steam cycles used in electrical power plants and in the production of shaft power
in industry are based on the familiar Rankine cycle, studied briefly in most courses in
thermodynamics. In this chapter we review the basic Rankine cycle and examine
modifications of the cycle that make modern power plants efficient and reliable.

2.2 A Simple Rankine-Cycle Power Plant

The most prominent physical feature of a modern steam power plant (other than its
smokestack) is the steam generator, or boiler, as seen in Figure 2.1. There the
combustion, in air, of a fossil fuel such as oil, natural gas, or coal produces hot
combustion gases that transfer heat to water passing through tubes in the steam
generator. The heat transfer to the incoming water (feedwater) first increases its
temperature until it becomes a saturated liquid, then evaporates it to form saturated
vapor, and usually then further raises its temperature to create superheated steam.
Steam power plants such as that shown in Figure 2.1, operate on sophisticated
variants of the Rankine cycle. These are considered later. First, let’s examine the
simple Rankine cycle shown in Figure 2.2, from which the cycles of large steam power
plants are derived.
In the simple Rankine cycle, steam flows to a turbine, where part of its energy is
converted to mechanical energy that is transmitted by rotating shaft to drive an
electrical generator. The reduced-energy steam flowing out of the turbine condenses to
liuid water in the condenser. A feedwater pump returns the condensed liquid
(condensate) to the steam generator. The heat rejected from the steam entering the
condenser is transferred to a separate cooling water loop that in turn delivers the
rejected energy to a neighboring lake or river or to the atmosphere.
36

As a result of the conversion of much of its thermal energy into mechanical energy,
or work, steam leaves the turbine at a pressure and temperature well below the turbine
entrance (throttle) values. At this point the steam could be released into the
atmosphere. But since water resources are seldom adequate to allow the luxury of one-
time use, and because water purification of a continuous supply of fresh feedwater is
costly, steam power plants normally utilize the same pure water over and over again.
We usually say that the working fluid (water) in the plant operates in a cycle or
undergoes of cyclic process, as indicated in Figure 2.2. In order to return the steam to
the high-pressure of the steam generator to continue the cycle, the low- pressure steam
leaving the turbine at state 2 is first condensed to a liquid at state 3 and then
pressurized in a pump to state 4. The high pressure liquid water is then ready for its
next pass through the steam generator to state 1 and around the Rankine cycle again.
The steam generator and condenser both may be thought of as types of heat
exchangers, the former with hot combustion gases flowing on the outside of water-
37

filled tubes, and the latter with external cooling water passing through tubes on which
the low- pressure turbine exhaust steam condenses. In a well-designed heat exchanger,
both fluids pass through with little pressure loss. Therefore, as an ideal, it is common
to think of steam generators and condensers as operating with their fluids at
unchanging pressures.
It is useful to think of the Rankine cycle as operating between two fixed pressure
levels, the pressure in the steam generator and pressure in the condenser. A pump
provides the pressure increase, and a turbine provides the controlled pressure drop
between these levels.
Looking at the overall Rankine cycle as a system (Figure 2.2), we see that work is
delivered to the surroundings (the electrical generator and distribution system) by the
turbine and extracted from the surroundings by a pump (driven by an electric motor or
a small steam turbine). Similarly, heat is received from the surroundings (combustion
gas) in the steam generator and rejected to cooling water in the condenser.
38

At the start of the twentieth century reciprocating steam engines extracted thermal
energy from steam and converted linear reciprocating motion to rotary motion, to
provide shaft power for industry. Today, highly efficient steam turbines, such as shown
in Figure 2.3, convert thermal energy of steam directly to rotary motion. Eliminating
the intermediate step of conversion of thermal energy into the linear motion of a piston
was an important factor in the success of the steam turbine in electric power
generation. The resulting high rotational speed, reliability, and power output of the
turbine and the development of electrical distribution systems allowed the centralization
of power production in a few large plants capable of serving many industrial and
residential customers over a wide geographic area.
The final link in the conversion of chemical energy to thermal energy to mechanical
energy to electricity is the electrical generator. The rotating shaft of the electrical
generator usually is directly coupled to the turbine drive shaft. Electrical windings
attached to the rotating shaft of the generator cut the lines of force of the stator
windings, inducing a flow of alternating electrical current in accordance with Faraday's
Law. In the United States, electrical generators turn at a multiple of the generation
frequency of 60 cycles per second, usually 1800 or 3600 rpm. Elsewhere, where 50
cycles per second is the standard frequency, the speed of 3000 rpm is common.
Through transformers at the power plant, the voltage is increased to several hundred
thousand volts for transmission to distant distribution centers. At the distribution
centers as well as neighborhood electrical transformers, the electrical potential is
reduced, ultimately to the 110- and 220-volt levels used in homes and industry.
39

Since at present there is no economical way to store the large quantities of


electricity produced by a power plant, the generating system must adapt, from moment
to moment, to the varying demands for electricity from its customers. It is therefore
important that a power company have both sufficient generation capacity to reliably
satisfy the maximum demand and generation equipment capable of adapting to varying
load.

2.3 Rankine-Cycle Analysis

In analyses of heat engine cycles it is usually assumed that the components of the
engine are joined by conduits that allow transport of the working fluid from the exit of
one component to the entrance of the next, with no intervening state change. It will be
seen later that this simplification can be removed when necessary.
It is also assumed that all flows of mass and energy are steady, so that the steady
state conservation equations are applicable. This is appropriate to most situations
because power plants usually operate at steady conditions for significant lengths of
time. Thus, transients at startup and shutdown are special cases that will not be
considered here.
Consider again the Rankine cycle shown in Figure 2.2. Control of the flow can be
exercised by a throttle valve placed at the entrance to the turbine (state 1). Partial valve
closure would reduce both the steam flow to the turbine and the resulting power
output. We usually refer to the temperature and pressure at the entrance to the turbine
as throttle conditions. In the ideal Rankine cycle shown, steam expands adiabatically
and reversibly, or isentropically, through the turbine to a lower temperature and
pressure at the condenser entrance. Applying the steady-flow form of the First Law of
Thermodynamics [Equation (1.10)] for an isentropic turbine we obtain:

q = 0 = h2 – h1 + wt [Btu/lbm | kJ/kg]

where we neglect the usually small kinetic and potential energy differences between the
inlet and outlet. This equation shows that the turbine work per unit mass passing
through the turbine is simply the difference between the entrance enthalpy and the
lower exit enthalpy:

wt = h1 – h2 [Btu/lbm | kJ/kg] (2.1)

The power delivered by the turbine to an external load, such as an electrical generator,
is given by the following:

Turbine Power = mswt = ms(h1 – h2) [Btu/hr | kW]

where ms [lbm /hr | kg/s] is the mass flow of steam though the power plant.
40

Applying the steady-flow First Law of Thermodynamics to the steam generator,


we see that shaft work is zero and thus that the steam generator heat transfer is

qa = h1 – h4 [Btu/lbm | kJ/kg] (2.2)

The condenser usually is a large shell-and-tube heat exchanger positioned below or


adjacent to the turbine in order to directly receive the large flow rate of low-pressure
turbine exit steam and convert it to liquid water. External cooling water is pumped
through thousands of tubes in the condenser to transport the heat of condensation of
the steam away from the plant. On leaving the condenser, the condensed liquid (called
condensate) is at a low temperature and pressure compared with throttle conditions.
Continued removal of low-specific-volume liquid formed by condensation of the high-
specific-volume steam may be thought of as creating and maintaining the low pressure
in the condenser. The phase change in turn depends on the transfer of heat released to
the external cooling water. Thus the rejection of heat to the surroundings by the
cooling water is essential to maintaining the low pressure in the condenser. Applying
the steady-flow First Law of Thermodynamics to the condensing steam yields:

qc = h3 – h2 [Btu/lbm | kJ/kg] (2.3)

The condenser heat transfer qc is negative because h2 > h3. Thus, consistent with sign
convention, qc represents an outflow of heat from the condensing steam. This heat is
absorbed by the cooling water passing through the condenser tubes. The condenser-
cooling-water temperature rise and mass-flow rate mc are related to the rejected heat
by:
ms|qc| = mc cwater(Tout - Tin) [Btu/hr | kW]

where cwater is the heat capacity of the cooling water [Btu/lbm-R | kJ/kg-K]. The
condenser cooling water may be drawn from a river or a lake at the temperature Tin and
returned downstream at Tout, or it may be circulated through cooling towers where heat
is rejected from the cooling water to the atmosphere.
We can express the condenser heat transfer in terms of an overall heat transfer
coefficient, U, the mean cooling water temperature, Tm = (Tout + Tin)/2, and the
condensing temperature T3:

ms|qc| = UA(T3 - Tm) [Btu/hr | kJ/s]

It is seen for given heat rejection rate, the condenser size represented by the tube
surface area A depends inversely on (a) the temperature difference between the
condensing steam and the cooling water, and (b) the overall heat-transfer coefficient.
For a fixed average temperature difference between the two fluids on opposite
sides of the condenser tube walls, the temperature of the available cooling water
controls the condensing temperature and hence the pressure of the condensing steam.
41

Therefore, the colder the cooling water, the lower the minimum temperature and
pressure of the cycle and the higher the thermal efficiency of the cycle.
A pump is a device that moves a liquid from a region of low pressure to one of
high pressure. In the Rankine cycle the condenser condensate is raised to the pressure
of the steam generator by boiler feed pumps, BFP. The high-pressure liquid water
entering the steam generator is called feedwater. From the steady-flow First Law of
Thermodynamics, the work and power required to drive the pump are:

wp = h3 – h4 [Btu/lbm | kJ/kg] (2.4)

and

Pump Power = mswp = ms(h3 – h4) [Btu/hr | kW]

where the negative values resulting from the fact that h4 > h3 are in accordance with the
thermodynamic sign convention, which indicates that work and power must be supplied
to operate the pump.
The net power delivered by the Rankine cycle is the difference between the turbine
power and the magnitude of the pump power. One of the significant advantages of the
Rankine cycle is that the pump power is usually quite small compared with the turbine
power. This is indicated by the work ratio, wt / wp, which is large compared with one
for Rankine cycle. As a result, the pumping power is sometimes neglected in
approximating the Rankine cycle net power output.
It is normally assumed that the liquid at a pump entrance is saturated liquid. This
is usually the case for power-plant feedwater pumps, because on the one hand
subcooling would increase the heat edition required in the steam generator, and on the
other the introduction of steam into the pump would cause poor performance and
destructive, unsteady operation. The properties of the pump inlet or condenser exit
(state 3 in Figure 2.2) therefore may be obtained directly from the saturated-liquid
curve at the (usually) known condenser pressure.
The properties for an isentropic pump discharge at state 4 could be obtained from
a subcooled-water property table at the known inlet entropy and the throttle pressure.
However, such tables are not widely available and usually are not needed. The
enthalpy of a subcooled state is commonly approximated by the enthalpy of the
saturated-liquid evaluated at the temperature of the subcooled liquid. This is usually
quite accurate because the enthalpy of a liquid is almost independent of pressure. An
accurate method for estimating the pump enthalpy rise and the pump work is given later
(in Example 2.3).
A measure of the effectiveness of an energy conversion device is its thermal
efficiency, which is defined as the ratio of the cycle net work to the heat supplied from
external sources. Thus, by using Equations (2.1), (2.2), and (2.4) we can express the
ideal Rankine-cycle thermal efficiency in terms of cycle enthalpies as:
42

th = (h1 – h2 + h3 – h4)/(h1 – h4) [dl] (2.5)

In accordance with the Second Law of Thermodynamics, the Rankine cycle


efficiency must be less than the efficiency of a Carnot engine operating between the
same temperature extremes. As with the Carnot-cycle efficiency, Rankine-cycle
efficiency improves when the average heat-addition temperature increases and the heat-
rejection temperature decreases. Thus cycle efficiency may be improved by increasing
turbine inlet temperature and decreasing the condenser pressure (and thus the
condenser temperature).
Another measure of efficiency commonly employed by power plant engineers is the
heat rate, that is, the ratio of the rate of heat addition in conventional heat units to the
net power output in conventional power units. Because the rate of heat addition is
proportional to the fuel consumption rate, the heat rate is a measure of fuel utilization
rate per unit of power output. In the United States, the rate of heat addition is usually
stated in Btu/hr, and electrical power output in kilowatts, resulting in heat rates being
expressed in Btu/kW-hr. The reader should verify that the heat rate in English units is
given by the conversion factor, 3413 Btu/kW-hr, divided by the cycle thermal efficiency
as a decimal fraction, and that its value has a magnitude of the order of 10,000
Btu/kW-hr. In the SI system of units, the heat rate is usually expressed in kJ/kW-hr, is
given by 3600 divided by the cycle efficiency as a decimal fraction, and is of the same
order of magnitude as in the English system. It is evident that a low value of heat rate
represents high thermal efficiency and is therefore desirable.

EXAMPLE 2.1

An ideal Rankine cycle (see Figure 2.2) has a throttle state of 2000 psia/1000°F and
condenser pressure of 1 psia. Determine the temperatures, pressures, entropies, and
enthalpies at the inlets of all components, and compare the thermal efficiency of the
cycle with the relevant Carnot efficiency. Neglect pump work. What is the quality of
the steam at the turbine exit?

Solution
The states at the inlets and exits of the components, following the notation of
Figure 2.2, are listed in the following table. The enthalpy and entropy of state 1 may be
obtained directly from tables or charts for superheated steam (such as those in
Appendices B and C) at the throttle conditions. A Mollier chart is usually more
convenient than tables in dealing with turbine inlet and exit conditions.
For an ideal isentropic turbine, the entropy is the same at state 2 as at state 1. Thus
state 2 may be obtained from the throttle entropy (s2 = s1 = 1.5603 Btu/lbm-R) and the
condenser pressure (1 psia). In general, this state may be in either the superheated-
steam region or the mixed-steam-and-liquid region of the Mollier and T-s diagrams. In
the present case it is well into the mixed region, with a temperature of 101.74°F and an
enthalpy of 871 Btu/lbm.
43

The enthalpy, h3 = 69.73 Btu/lbm, and other properties at the pump inlet are obtained
from saturated-liquid tables, at the condenser pressure. The steady-flow First Law of
Thermodynamics, in the form of Equation (2.4), indicates that neglecting isentropic
pump work is equivalent to neglecting the pump enthalpy rise. Thus in this case
Equation (2.4) implies that h3 and h4 shown in Figure (2.2) are almost equal. Thus we
take h4 = h3 as a convenient approximation.

State Temperature Pressure Entropy Enthalpy


(°F) (psia) (Btu/lbm-°R) (Btu/lbm)

1 1000.0 2000 1.5603 1474.1

2 101.74 1 1.5603 871.0

3 101.74 1 0.1326 69.73

4 101.74 2000 0.1326 69.73

The turbine work is

h1 – h2 = 1474.1 – 871 = 603.1 Btu/lbm.

The heat added in the steam generator is

h1 – h4 = 1474.1 – 69.73 = 1404.37 Btu/lbm.

The thermal efficiency is the net work per heat added = 603.1/1404.37 = 0.4294
(42.94%). This corresponds to a heat rate of 3413/0.4294 = 7946 Btu/kW-hr. As
expected, the efficiency is significantly below the value of the Carnot efficiency
of 1 – (460 + 101.74)/(460 + 1000) = 0.6152 (61.52%), based on a source
temperature of T1 and a sink temperature of T3.
The quality of the steam at the turbine exit is

(s2 – sl)/(sv – sl) = (1.5603 – 0.1326)/(1.9781 – 0.1326) = 0.7736

Here v and l indicate saturated vapor and liquid states, respectively, at pressure p2.
Note that the quality could also have been obtained from the Mollier chart for steam as
1 - M, where M is the steam moisture fraction at entropy s2 and pressure p2.
__________________________________________________________________

Example 2-2
If the throttle mass-flow is 2,000,000 lbm/hr and the cooling water enters the condenser
at 60°F, what is the power plant output in Example 2.1? Estimate the cooling-water
mass-flow rate.
44

Solution: The power output is the product of the throttle mass-flow rate and the power
plant net work. Thus

Power = (2 × 106)(603.1) = 1.206 × 109 Btu/hr

or

Power = 1.206 × 109 / 3413 = 353,413 kW.

The condenser heat-transfer rate is

msqc = ms ( h3 – h2 ) = 2,000,000 × (69.73 – 871) = – 1.603×109 Btu/hr

The condensing temperature, T3 = 101.74 °F, is the upper bound on the cooling
water exit temperature. Assuming that the cooling water enters at 60°F and leaves at
95°F, the cooling-water flow rate is given by

mc = ms|qc| / [ cwater(Tout – Tin)] = 1.603×109 /[(1)(95 - 60)] = 45.68×106 lbm/hr

A higher mass-flow rate of cooling water would allow a smaller condenser cooling-
water temperature rise and reduce the required condenser-heat-transfer area at the
expense of increased pumping power.
____________________________________________________________________

2.4 Deviations from the Ideal – Component Efficiencies

In a power plant analysis it is sometimes necessary to account for non-ideal effects such
as fluid friction, turbulence, and flow separation in components otherwise assumed to
be reversible. Decisions regarding the necessity of accounting for these effects are
largely a matter of experience built on familiarity with the magnitudes of the effects,
engineering practices, and the uses of the calculated results.

Turbine

In the case of an adiabatic turbine with flow irreversibilities, the steady-flow First Law
of Thermodynamics gives the same symbolic result as for the isentropic turbine in
Equation (2.1), i.e.,

wt = h1 – h2 [Btu/lb | kJ/kg] (2.6)

except that here h2 represents the actual exit enthalpy and wt is the actual work of an
adiabatic turbine where real effects such as flow separation, turbulence, irreversible
internal heat transfers, and fluid friction exist.
45

An efficiency for a real turbine, known as the isentropic efficiency, is defined as


the ratio of the actual shaft work to the shaft work for an isentropic expansion between
the same inlet state and exit pressure level. Based on the notation of Figure 2.4, we see
that the turbine efficiency is:

turb = (h1 – h2 )/(h1 – h2s ) [dl] (2.7)

where h2s, the isentropic turbine-exit enthalpy, is the enthalpy evaluated at the turbine
inlet entropy and the exit pressure. For the special case of an isentropic turbine,
h2 = h2s and the efficiency becomes 1. Note how state 2 and the turbine work change
in Figure 2.4 as the efficiency increases toward 1. The diagram shows that the
difference between the isentropic and actual work, h2 – h2s, represents work lost due to
irreversibility. Turbine isentropic efficiencies in the low 90% range are currently
achievable in well-designed machines.
Normally in solving problems involving turbines, the turbine efficiency is known
from manufacturers’ tests, and the inlet state and the exhaust pressure are specified.
State 1 and p2 determine the isentropic discharge state 2s using the steam tables. The
actual turbine-exit enthalpy can then be calculated from Equation (2.7). Knowing both
p2 and h2, we can then fully identify state 2 and account for real turbine behavior in any
cycle analysis.

Pump

Work must be supplied to a pump to move liquid from a low pressure to a high
pressure. Some of the work supplied is lost due to irreversibilities. Ideally the remaining
effective work to raise the pressure is necessarily less than that supplied. In order for
46

the efficiency of a pump to be less than or equal to 1, it is defined in inverse fashion to


turbine efficiency. That is, pump efficiency is the ratio of the isentropic work to the
actual work input when operating between two given pressures. Applying Equation
(2.4) and the notation of Figure (2.5), the isentropic pump work, wps = h3 – h4s, and the
pump isentropic efficiency is

pump = wps /wp = (h4s – h3)/(h4 – h3) [dl] (2.8)

Note the progression of exit states that would occur in Figure 2.5 as pump efficiency
increases for a fixed inlet state and exit pressure. It is seen that the pump lost work,
given by h4 – h4s decreases and that the actual discharge state approaches the isentropic
discharge state.
States 4 and 4s are usually subcooled liquid states. As a first approximation their
enthalpies may be taken to be the saturated liquid enthalpy at T3. More accurate
approximations for these enthalpies may be obtained by applying the First Law for a
closed system undergoing a reversible process, Equation (1.8): Tds = dh - vdp. For an
isentropic process it follows that dh = vdp. Because a liquid is almost incompressible,
its specific volume, v, is almost independent of pressure. Thus, using the notation of
Figure 2.5, integration with constant specific volume yields

h4s = h3 + v3 ( p4 – p3 ) [Btu/lbm | kJ/kg]

where a knowledge of state 3 and p4 determines h4s.


47

Using Equation (2.8), and without consulting tables for subcooled water, we can
then calculate the pump work from

wp = v3(p3 – p4)/p [ft-lbf/lbm | kN-m/kg] (2.9)

Note that the appropriate conversion factors must be applied for dimensional
consistency in Equation (2.9).

EXAMPLE 2.3

Calculate the actual work and the isentropic and actual discharge enthalpies for an 80%
efficient pump with an 80°F saturated-liquid inlet and an exit pressure of 3000 psia.

Solution
From the saturated-liquid tables, for 80°F, the pump inlet conditions are 0.5068 psia,
48.037 Btu/lbm, and 0.016072 ft3/lbm.
Using Equation (2.9), we find that the pump work is

wp = [0.016072(0.5068 – 3000)(144)]/0.8 = – 8677 ft-lbf / lbm


or
wp = – 8677/778 = – 11.15 Btu/lbm.

Note the importance of checking units here.


The actual discharge enthalpy is

h4 = h3 – wp = 48.037 – (–11.15) = 59.19 Btu/lbm.

and the isentropic discharge enthalpy is

h4s = h3 – p wp = 48.037 – (0.8)(– 11.15) = 56.96 Btu/lbm.


____________________________________________________________________

EXAMPLE 2.4

What is the turbine work, the net work, the work ratio, and the cycle thermal efficiency
for the conditions of Example 2.1 if the turbine efficiency is 90% and the pump
efficiency is 85%? What is the turbine exit quality?

Solution
By the definition of isentropic efficiency, the turbine work is 90% of the isentropic
turbine work = (0.9)(603.1) = 542.8 Btu/lbm.
By using Equation (2.9), the isentropic pump work is
[(0.01614)(1 – 2000)(144)] / 778 = – 5.97 Btu/lbm.
48

The actual pump work is then – 5.97/.85 = – 7.03 Btu/lbm and the work ratio is
542.8/| – 7.03| = 77.2
The cycle net work is wt + wp = 542.8 – 7.03 = 535.8 Btu/lbm..

Applying the steady-flow First Law of Thermodynamics to the pump, we get the
enthalpy entering the steam generator to be

h4 = h3 – wp = 69.73 – (– 7.03) = 76.76 Btu/lbm.

The steam-generator heat addition is then reduced to 1474.1 – 76.76 = 1397.3


Btu/lbm. and the cycle efficiency is 535.8/1397.3 = 0.383. Study of these examples
shows that the sizable reduction in cycle efficiency from that in Example 2.1 is largely
due to the turbine inefficiency, not to the neglect of pump work.
From Equation (2.6), the true turbine exit enthalpy is the difference between the
throttle enthalpy and actual turbine work = 1474.1 - 542.8 = 931.3 Btu/lbm.
The quality is then x = (h2 – hl)/(hv – hl) = (931.3 – 69.73)/(1105.8 – 69.73) =
0.832.
Thus the turbine inefficiency increases the turbine exhaust quality over the
isentropic turbine value of 0.774.
____________________________________________________________________

2.5 Reheat and Reheat Cycles

A common modification of the Rankine cycle in large power plants involves


interrupting the steam expansion in the turbine to add more heat to the steam before
completing the turbine expansion, a process known as reheat. As shown in Figure 2.6,
steam from the high-pressure (HP) turbine is returned to the reheat section of the steam
generator through the "cold reheat" line. There the steam passes through heated tubes
which restore it to a temperature comparable to the throttle temperature of the high
pressure turbine. The reenergized steam then is routed through the "hot reheat" line to
a low-pressure turbine for completion of the expansion to the condenser pressure.
Examination of the T-s diagram shows that reheat increases the area enclosed by
the cycle and thus increases the net work of the cycle by virtue of the cyclic integral,
Equation (1.3). This is significant, because for a given design power output higher net
work implies lower steam flow rate. This, in turn, implies that smaller plant components
may be used, which tends to reduce the initial plant cost and to compensate for added
costs due to the increased complexity of the cycle.
Observe from Figure 2.6 that the use of reheat also tends to increase the average
temperature at which heat is added. If the low-pressure turbine exhaust state is
superheated, the use of reheat may also increase the average temperature at which heat
is rejected. The thermal efficiency may therefore increase or decrease, depending on
specific cycle conditions. Thus the major benefits of reheat are increased net work,
49

drying of the turbine exhaust (discussed further later), and the possibility of improved
cycle efficiency.
Note that the net work of the reheat cycle is the algebraic sum of the work of the
two turbines and the pump work. Note also that the total heat addition is the sum of the
heat added in the feedwater and reheat passes through the steam generator. Thus the
50

thermal efficiency of the reheat cycle is:

(h1 – h2) + (h3 – h4) + (h5 – h6)


th = ----------------------------------- [dl] (2.10)
(h1 – h6) + (h3 – h2)

Relations such as this illustrate the wisdom of learning to analyze cycles using
definitions and applying fundamentals to components rather than memorizing equations
for special cases such as Equation (2.5) for the efficiency of the simple Rankine cycle.
Note that the inclusion of reheat introduces a third pressure level to the Rankine
cycle. Determination of a suitable reheat pressure level is a significant design problem
that entails a number of considerations. The cycle efficiency, the net work, and other
parameters will vary with reheat pressure level for given throttle and condenser
conditions. One of these may be numerically optimized by varying reheat pressure level
while holding all other design conditions constant.
Reheat offers the ability to limit or eliminate moisture at the turbine exit. The
presence of more than about 10% moisture in the turbine exhaust can cause erosion of
blades near the turbine exit and reduce energy conversion efficiency. Study of Figure
2.6 shows that reheat shifts the turbine expansion process away from the two-phase
region and toward the superheat region of the T-s diagram, thus drying the turbine
exhaust.

EXAMPLE 2.5

Reanalyze the cycle of Example 2.1 (2000 psia/1000°F/1 psia) with reheat at 200 psia
included. Determine the quality or degree of superheat at the exits of both turbines.
Assume that reheat is to the HP turbine throttle temperature.

Solution
Referring to Figure 2.6, we see that the properties of significant states are the
following:

State Temperature Pressure Entropy Enthalpy


(°F) (psia) (Btu/lbm-°R) (Btu/lbm)

1 1000.0 2000 1.5603 1474.1

2 400.0 200 1.5603 1210.0

3 1000.0 200 1.84 1527.0

4 101.74 1 1.84 1028.0

5 101.74 1 0.1326 69.73

6 101.74 2000 0.1326 69.73


51

Properties here are obtained from the steam tables and the Mollier chart as follows:

1. The enthalpy and entropy at state 1 are read from the superheated-steam tables
at the given throttle temperature and pressure.

2. State 2 is evaluated from the Mollier diagram at the given reheat pressure and
the same entropy as in state 1 for the isentropic turbine expansion.

3. Reheat at constant pressure p3 = p2 to the assumed throttle temperature T3 = T1


gives s3 and h3. Normally, T3 is assumed equal to T1 unless otherwise specified.

4. The second turbine flow is also specified as isentropic with expansion at s4 = s3


to the known condenser pressure p4.

5. The condenser exit (pump entrance) state is assumed to be a saturated liquid at


the known condenser pressure.

6. Pump work is neglected here. The steady-flow First Law then implies that
h6 = h5, which in turn implies the T6 = T5.

The turbine work is the sum of the work of both turbines:


(1474.1 - 1210) + (1527 - 1028) = 763.1 Btu/lbm.

The heat added in the steam generator feedwater and reheat passes is
(1474.1 - 69.73) + (1527 - 1210) = 1721.4 Btu/lbm.

The thermal efficiency then is 763.1/1721.4 = 0.443, or 44.3%.


Both the net work and the cycle efficiency are higher than in the simple Rankine
cycle case of Example 2.1. From the Mollier chart in Appendix B it is readily seen that
state 2 is superheated, with 400 - 381.8 = 18.2 Fahrenheit degrees of superheat; and
state 4 is wet steam, with 7.4% moisture, or 0.926 (92.6%) quality. Thus the first
turbine has no moisture and the second is substantially drier than 0.774 quality value in
Example 2.1.
____________________________________________________________________

Reheat is an important feature of all large, modern fossil-fueled steam power plants.
We now consider another key feature of these plants, but temporarily omit reheat, for
the purpose of clarity.

2.6 Regeneration and Feedwater Heaters

The significant efficiency advantage of the Carnot cycle over the Rankine cycle is due
to the fact that in the Carnot cycle all external heat addition is at a single high
52

temperature and all external heat rejection at a single low temperature. Examination of
Figures 2.2 and 2.6 shows that heat addition in the steam generator takes place over a
wide range of water temperature in both the simple and reheat Rankine cycles.
Presumably, the Rankine-cycle thermal efficiency could be improved by increasing the
average water temperature at which heat is received. This could be accomplished by an
internal transfer of heat from higher-temperature steam to low-temperature feedwater.
An internal transfer of heat that reduces or eliminates low-temperature additions of
external heat to the working fluid is known as regeneration.

Open Feedwater Heaters

Regeneration is accomplished in all large-scale, modern power plants through the use
of feedwater heaters. A feedwater heater (FWH) is a heat exchanger in which the latent
heat (and sometimes superheat) of small amounts of steam is used to increase the
temperature of liquid water (feedwater) flowing to the steam generator. This provides
the internal transfer of heat mentioned above.
An open feedwater heater is a FWH in which a small amount of steam mixes
directly with the feedwater to raise its temperature. Steam drawn from a turbine for
feedwater heating or other purposes is called extraction steam. Feedwater heaters in
which extraction steam heats feedwater without fluid contact will be discussed later.
Consider the regenerative Rankine-cycle presented in Figure 2.7. The steam
leaving the high-pressure (HP) turbine is split with a small part of the mass flow
extracted to an open FWH and the major part of the flow passing to a low pressure
(LP) turbine. The T-s diagram shows that steam entering the FWH at state 2 is at a
higher temperature than the subcooled feedwater leaving the pump at state 5. When the
two fluids mix in the FWH, the superheat and the heat of vaporization of the extraction
steam are transferred to the feedwater, which emerges with the condensed extraction
steam at a higher temperature, T6.. It is assumed that all streams entering and leaving
the FWH are the same pressure so that the mixing process occurs at constant pressure.
The T-s and flow diagrams show that heat from combustion gases in the steam
generator need only raise the water temperature from T7 to T1 rather than from T5 when
extraction steam is used to heat the feedwater. The average temperature for external
heat addition must therefore increase. Despite the reduced flow rate through the low-
pressure turbine, we will see by example that the thermal efficiency of the steam cycle is
improved by the transfer of energy from the turbine extraction flow to the feedwater.
The analysis of cycles with feedwater heaters involves branching of steam flows.
In Figure 2.7, for example, conservation of mass must be satisfied at the flow junction
downstream of the high-pressure-turbine exit. Thus, assuming a mass flow of 1 at the
HP turbine throttle and a steam mass-flow fraction, m1, through the feedwater heater,
the low-pressure-turbine mass-fraction must be 1 - m1. Note that the latter flow passes
through the condenser and pump and is reunited with the extraction flow, m1, in the
FWH at state 6, where the exit-flow-rate fraction is again unity.
53

It will be seen later that it is common for more than one FWH to be used in a
single power plant. When more than one FWH is present, mass flows m1, m2...mn are
defined for each of the n FWHs. Conservation of mass is used to relate these flows to
54

condenser flow rate and the reference throttle flow rate. This is accomplished by taking
a mass flow of 1 at the high-pressure-turbine throttle as a reference, as in the case of a
single FWH discussed above. After solving for each of the thermodynamic states and
FWH mass fractions, actual mass flow rates are obtained as the products of the known
(or assumed) throttle flow rate and FWH mass-flow fractions.
The function of feedwater heaters is to use the energy of extraction steam to
reduce the addition of low-temperature external heat by raising the temperature of the
feedwater before it arrives at the steam generator. Feedwater heaters are therefore
insulated to avoid heat loss to the surroundings. Because the resulting heat loss is
negligible compared with the energy throughflow, feedwater heaters are usually treated
as adiabatic devices.
In order to avoid irreversibility associated with unrestrained expansion, constant
pressure mixing of the streams entering the FWH is necessary. Returning to Figure 2.7,
this implies that the pressures of the feedwater at state 5 and at the FWH exit state 6
are chosen to be the same as that of the extraction steam at state 2.
Note that, as with reheat, the inclusion of a FWH also introduces an additional
pressure level into the Rankine cycle as seen in the T-s diagram. In the figure, the
extraction pressure level, p2, is another parameter under the control of the designer.
The extraction mass flow rate, m1, is in turn controlled by the designer’s choice of p2.
The mass-flow rate is determined by the physical requirement that the feedwater
entering the FWH at state 5 increase in temperature to T6 through absorption of the
heat released by the condensing extraction steam. This is accomplished by applying the
steady-flow First Law of Thermodynamics, using appropriate mass fractions, to the
insulated open FWH:

q = 0 = (1)h6 – m1h2 – (1 – m1 )h5 + 0 [Btu/lbm | kJ/kg]

Every term in this equation has dimensions of energy per unit throttle mass, thus
referring all energy terms to the mass-flow rate at the throttle of the high-pressure
turbine. For example, the second term on the right is of the form:

FWH Extraction mass Enthalpy at state 2 Enthalpy at state 2


-------------------------- × -------------------------- = ----------------------
Throttle mass FWH Extraction mass Throttle mass

Similarly, the structure of the third term on the right has the significance of

Pump mass Enthalpy at state 5 Enthalpy at state 5


---------------- × --------------------- = ----------------------
Throttle mass Pump mass Throttle mass
55

Solving for the extraction mass fraction, we obtain

m1 = (h6 – h5) / (h2 – h5) [dl] (2.11)

For low extraction pressures, the numerator is usually small relative to the
denominator, indicating a small extraction flow. The T-s diagram of Figure 2.7 shows
that increasing the extraction pressure level increases both h6 and h2. Thus, because the
small numerator increases faster than the large denominator, we may reason, from
Equation (2.11), that the extraction mass-flow fraction must increase as the extraction
pressure level increases. This conforms to the physical notion that suggests the need
for more and hotter steam to increase the feedwater temperature rise. While such
intuitions are valuable, care should be exercised in accepting them without proof.
The total turbine work per unit throttle mass flow rate is the sum of the work of
the turbines referenced to the throttle mass-flow rate. Remembering that 1 - m1 is the
ratio of the low-pressure turbine mass flow to the throttle mass flow, we obtain:

wt = (h1 – h2 ) + (1 – m1 )(h2 – h3 ) [Btu/lbm | kJ/kg] (2.12)

The reader should examine the structure of each term of Equation (2.12) in the light of
the previous discussion. Note that it is not important to remember these specific
equations, but it is important to understand, and be able to apply, the reasoning by
which they are obtained.
For a given throttle mass flow rate, mthr [lbm/s | kg/s], the total turbine power
output is given by mthrwt [Btu/s | kW].
We see in Figure 2.7 that the heat addition in the steam generator is reduced, due
to extraction at pressure p6 = p2, by about h7 – h5 to

qa = h1 – h7 [Btu/lbm | kJ/kg] (2.13)

At the same time, the net work also decreases, but more slowly, so that the net effect is
that the cycle efficiency increases with increased extraction.

EXAMPLE 2.6

Solve Example 2.1 (2000 psia /1000°F/1 psia) operating with an open feedwater heater
at 200 psia.

Solution
Referring to Figure 2.7, we find that the properties of significant states are:
56

State Temperature Pressure Entropy Enthalpy


(°F) (psia) (Btu/lbm-°R) (Btu/lbm)

1 1000.0 2000 1.5603 1474.1

2 400.0 200 1.5603 1210.0

3 101.74 1 1.5603 871.0

4 101.74 1 0.1326 69.73

5 101.74 200 0.1326 69.73

6 381.8 200 0.5438 355.5

7 381.8 2000 0.5438 355.5

States 1 through 4 are obtained in the same way as in earlier examples. Constant
pressure mixing requires that p5 = p6 = p2, the extraction pressure level. State 6, a
pump entrance state, is assumed to be a saturated-liquid state as usual. Subcooled-
liquid states are approximated, as before, consistent with the neglect of pump work.
The extraction mass fraction obtained by applying the steady-flow First Law of
Thermodynamics to the FWH, Equation (2.11), is

m1 = (355.5 – 69.73)/(1210 – 69.73) = 0.251.

The net work (neglecting pump work) by Equation (2.12), is then

wn = (1474.1 – 1210) + (1 – 0.251)(1210 – 871) = 518.1 Btu/lbm

This may be compared with the simple-cycle net work of 603.1 Btu/lbm.
The heat added in the steam generator by Equation (2.13) is

qa = h1 – h7 = 1474.1 – 355.5 = 1118.6 Btu/lbm.

The resulting cycle efficiency is th = 518.1/1118.6 = 0.463, or 46.3%, a


significantly higher value than the 42.94% for the corresponding simple Rankine cycle.
Note, however, that the LP-turbine exhaust quality is the same as for the simple
Rankine cycle, an unacceptable 77.4%. This suggests that a combination of reheat and
regeneration through feedwater heating may be desirable. We will investigate this
possibility later after looking at closed feedwater heaters.
_____________________________________________________________________

Closed Feedwater Heaters

We have seen that feedwater heating in open feedwater heaters occurs by mixing of
extraction steam and feedwater. Feedwater heating also is accomplished in shell-and-
57

tube-type heat exchangers, where extraction steam does not mix with the feedwater.
Normally, feedwater passes through banks of tubes whereas steam condenses on the
outside of the tube surfaces in these heaters. Such heat exchangers are called closed
feedwater heaters.

Pumped Condensate. Closed feedwater heaters normally are employed in two


configurations in power plants. In the configuration shown in figure 2.8, condensate is
pumped from the condenser through the FWH and the steam generator directly to the
turbine along the path 4-5-8-9-1. Ideally, p5 = p1 assuming no pressure drop in the
FWH and steam generator.
Note that if m1 mass units of steam are extracted from the turbine for use in the
FWH, only 1 - m1 units of feedwater pass throught the condenser, pump, and the tubes
of the FWH. The condensed extraction steam (condensate) emerging from the FWH at
state 6 is pumped separately from p6 = p2 to throttle pressure p7 = p1, where it becomes
part of the steam generator feedwater. The pumped condensate at state 7 thus mixes
with the heated feedwater at state 8 to form the total feedwater flow at state 9.
Constant pressure mixing ( p7 = p8 = p9) is required at this junction to avoid losses
associated with uncontrolled flow expansion.
The enthalpy of the feedwater entering the steam generator can be determined by
applying the steady-flow First Law of Thermodynamics to the junction of the feedwater
and FWH streams:

h9 = (1 – m1 )h8 + m1h7 [Btu/lbm | kJ/kg]

As in the open FWH analysis, the extraction mass fraction depends on the choice
of intermediate pressure p2 and is obtained by applying the steady-flow First Law of
Thermodynamics to the feedwater heater.

Throttled Condensate. The second closed FWH configuration is shown in Figure


2.9 where the FWH condensate drops in pressure from p6 = p2 through a trap into the
condenser at pressure p7 = p3 = p4. The trap allows liquid only to pass from the FWH at
state 6 in a throttling process to state 7. As usual, it is assumed that the throttling
process is adiabatic. The T-s diagram shows that the saturated liquid at state 6 flashes
into a mixture of liquid and vapor in the condenser with no change in enthalpy, h7 = h6.
For this configuration, the closed FWH condensate mass-flow rate is equal to the
extraction mass-flow rate. As a result, conservation of mass applied to the condenser
shows that the mass-flow rate leaving the condenser and passing through the pump and
FWH tubes is the same as the throttle mass-flow rate. The throttled-condensate, closed
feedwater heater is the preferred configuration in power plants, because it is
unnecessary for each FWH to have a condensate pump.
58
59
60

EXAMPLE 2.7

Rework Example 2.1 (2000 psia/1000°F/1 psia) with reheat and a closed feedwater
heater with extraction from the cold reheat line and FWH condensate throttled to the
condenser. Both reheat and extraction are at 200 psia. Assume that the feedwater
leaving the FWH is at the temperature of the condensing extraction stream. Draw
appropriate T-s and flow diagrams.

Solution
Referring to the notation of Figure 2.10, verify that the significant the
thermodynamic state properties are:

State Temperature Pressure Entropy Enthalpy


(°F) (psia) (Btu/lbm-°R) (Btu/lbm)

1 1000.0 2000 1.5603 1474.1

2 400.0 200 1.5603 1210.0

3 1000.0 200 1.84 1527.0

4 101.74 1 1.84 1028.0

5 101.74 1 0.1326 69.73

6 101.74 2000 0.1326 69.73

7 381.8 200 0.5438 355.5

8 101.74 1 __ 355.5

9 381.8 2000 __ 355.5

Applying the steady-flow First Law of Thermodynamics to the FWH, we obtain:

0 = h9 + m1h7 – m1h2 – h6 + 0

which, solved for m1, yields:

m1 = ( h9 – h6 )/( h2 – h7 ) = (355.5 – 69.73)/(1210 – 355.5) = 0.3344

The total net work per unit of mass flow at the throttle of the HP turbine is the sum of
the specific work of each of the turbines adjusted for the HP turbine throttle mass flow:

wn = h1 – h2 + (1 – m1)( h3 – h4 )

= 1474.1 – 1210 + (1 – 0.3344)(1527 – 1028) = 596.2 Btu/lbm


61
62

As in the earlier examples in this series, pump work has been neglected.
The heat addition per unit HP-turbine-throttle mass is the sum of the heat addition
in the main pass and reheat pass through the steam generator, the latter as adjusted for
the reduced mass flow. Thus the steady-flow First Law of Thermodynamics yields

qa = h1 – h9 + (1 – m1 )( h3 – h2 )

= 1474.1 – 355.5 + (1 – 0.3344)(1527 – 1210) = 1329.6 Btu/lbm

The thermal efficiency of the cycle is wn / qa = 596.2 / 1329.6 = 0.448, or 44.8%.


The Mollier chart shows that the discharge of the first turbine (state 2) has 20
degrees of superheat and the second turbine (state 4) 7.4% moisture, or a quality of
0.926.
___________________________________________________________________

In the above calculation it was assumed that the feedwater temperature leaving the
FWH had risen to the temperature of the condensing extraction steam. Since the FWH
is a heat exchanger of finite area, the feedwater temperature T9 usually differs from the
condensing temperature of the extraction steam T7. If the surface area of the FWH is
small, the feedwater will emerge at a temperature well below the extraction-steam
condensing temperature. If the area were increased, the feedwater temperature would
approach the condensing temperature. This aspect of FWH design is reflected in the
parameter known as the terminal temperature difference, TTD, defined as

TTD = Tsat - Tfw [R | K]

where Tfw is the temperature of the feedwater leaving the tubes and Tsat is the
condensing temperature of the extraction steam in the closed FWH. In Figure 2.10, for
instance, Tfw = T9 and Tsat = T7. Thus, if the TTD and the extraction pressure are
known, the true FWH exit temperature may be determined. An application of the TTD
will be considered in a later example.
Table 2.1 summarizes, for comparison, the results of the calculations for the
several plant configurations that we have considered. The reader is cautioned that since
these calculations have not accounted for turbine inefficiency, the thermal efficiencies
are unusually high. While the efficiency differences with respect to the simple cycle may
seem insignificant, they are of great economic importance. It must be realized that
hundreds of millions of dollars may be spent on fuel each year in a power plant and that
capital costs are equally impressive. As a result, the choice of cycle and design
characteristics are of great significance. Some further improvement in net work and
efficiency could be shown by selecting extraction and reheat pressure levels to
maximize these parameters.
63

Table 2.1 Comparison of Rankine Cycle Modifications


Net Work Efficiency Heat Rate Turbine Exit
(Btu/lbm) % (Btu/kW-hr) Quality

Simple cycle 603.1 42.9 7956 0.774

Reheat cycle 763.1 44.3 7704 0.926

One open FWH 518.1 46.3 7371 0.774

One closed FWH 596.2 44.8 7618 0.926


and reheat

Multistage Extraction

It has been shown that increases in cycle efficiency may be accomplished in a steam
power plant through regeneration via the feedwater heater. Large steam power plants
typically employ large numbers of feedwater heaters for this purpose. Multistage
extraction refers to the use of multiple extractions to supply steam to these feedwater
heaters. Earlier discussions of examples involved extractions taken only from the flows
between turbines. However, the number of extractions is not limited by the number of
turbines. In fact, large turbines are designed with several extraction points through
which steam may be withdrawn for feedwater heating and other purposes.

Assigning Extraction-Pressure Levels. Given n feedwater heaters, it is necessary to


assign values to the n associated extraction pressures. For preliminary design purposes,
the extraction-pressure levels assigned may be those that give equal feedwater
temperature rises through each heater and through the steam generator to the boiling
point. Thus, for n heaters the appropriate temperature rise is given by

Topt = ( Tsl – Tcond )/( n + 1) [R | K] (2.14)

where Tsl is the temperature the saturated liquid at the throttle pressure and Tcond is the
temperature the feedwater leaving the condenser. The corresponding steam condensing
temperature in the ith heater is then

Ti = Tcond + ( i )Topt

= Tcond + i ( Tsl – Tcond )/( n + 1) [R | K] (2.15)

where i = 1, 2..., n. Steam tables may then be used to evaluate the corresponding
extraction-pressure levels. It is, of course, possible and sometimes necessary to assign
extraction-pressure levels in other ways.
64

EXAMPLE 2.8

Evaluate the recommended extraction-pressure levels for single heater for the
1000° F/2000 psia throttle and one psia condenser that have been used throughout this
chapter.

Solution
The feedwater temperature rise to establish an appropriate extraction-pressure
level for a single heater for a plant such as that shown in Figures 2.7 through 2.9 is
(Tsl – T4 )/2 = (635.8 – 101.74)/2 = 267.05°F where Tsl was evaluated at p1 = 2000
psia. This would make T6 = 101.74 + 267.05 = 368.79°F and the corresponding
extraction pressure level p6 = p2 = 171 psia, using the saturated-steam tables.
____________________________________________________________________

At this point we have the tools necessary to evaluate the performance and penalties
associated with a given configuration. The following example examines the gains that
follow from the use a single feedwater heater and the sensitivity of the thermal
efficiency to the assigned feedwater temperature rise.

EXAMPLE 2.9

Consider a single open feedwater heater operating in a Rankine cycle with a 2000 psia
saturated-vapor throttle and a 1 psia condenser. Evaluate the thermal efficiency as a
function of feedwater temperature rise. Compare the temperature rise that maximizes
the thermal efficiency with the results of Equation (2.14).

Solution

Utilizing the notation of Figure 2.7 and taking the throttle state as a saturated vapor,
we get the results that are summarized in spreadsheet format in Table 2.2. (This table is
a direct reproduction of a Quattro Pro spreadsheet used in the analysis. Care should be
taken if this spreadsheet is used for "what if" studies, because it is dependent on manual
entry of thermodynamic properties. To explore other cases, appropriate properties must
be obtained from steam tables or charts and inserted in the spreadsheet. Despite this
drawback, the spreadsheet provides a convenient means of organizing, performing, and
displaying calculations.) Details of the methodology are given in the right-most column.
It is seen that the net work drops, as expected, as more extraction steam is used to heat
the feedwater. Figure 2.11 shows the percentage increase in thermal efficiency as a
function of the feedwater temperature rise for this case. Over a 9% increase in thermal
efficiency is achieved with feedwater temperature rises between 200/F and 300°F.
Thus the prediction of Topt = 267°F using Equation (2.14) in Example 2.8 is clearly in
this range.
_____________________________________________________________________
65

Example 2.9 shows that improved thermal efficiency is achieved over a broad range of
feedwater temperature rise and therefore extraction pressure. This gives the designer
freedom to assign extraction-pressure levels so as to make use of existing designs for
feedwater heaters and turbines without severely compromising the efficiency of the
plant design.

Calculation Methodology. Once the extraction- and reheat-pressure levels are


established for a cycle with multistage extraction, and once throttle and condenser
conditions, turbomachine efficiencies, and FWH terminal temperature differences are
known, significant state properties should be determined. Symbols for extraction
mass-fraction variables should be assigned for each heater and related to other
unknown flows using mass conservation assuming unit mass flow at the high-
pressure-turbine throttle. The steady-flow First Law of Thermodynamics should then
be applied to each of the FWHs, starting with the highest extraction pressure and
progressing to the lowest-pressure FWH. Analyzing the heaters in this order allows
each equation to be solved immediately for a mass fraction rather than solving all of
the equations simultaneously. Important performance parameters such as thermal
efficiency, net work, and work ratio may then be evaluated taking care to account
properly for component mass flows. The following example illustrates this
methodology.
66
67

EXAMPLE 2.10

Consider a power plant with 1000/F/2000-psia throttle, reheat at 200 psia back to
1000/F, and 1-psia condenser pressure. The plant has two closed feedwater heaters,
both with terminal temperature differences of 8/F. The high-pressure (HP) heater
condensate is throttled into the low-pressure (LP) heater, which in turn drains into the
condenser. Turbomachine efficiencies are 0.88, 0.9, and 0.8 for the HP turbine, the LP
turbine, and the boiler feed pump, respectively. Draw relevant T-s and flow diagrams
and evaluate FWH mass fractions, thermal efficiency, net work, and work ratio.

Solution
The notation used to study this plant is shown in Figure 2.12. The pertinent
thermodynamic properties and part of the analysis are presented in the spreadsheet
given in Table 2.3. The earlier-stated caution (Example 2.9) about using spreadsheets
that incorporate external data applies here as well, because changing parameters may
require changes in steam-table lookup values.
To start the analysis we first determine the extraction-pressure levels. The ideal
FWH temperature rise is given by

( Tsl – T7 )/3 = ( 635.8 – 101.74)/3 = 178.02°F

where the saturation temperature is evaluated at the HP-turbine throttle pressure of


2000 psia. The corresponding extraction condensing temperatures and extraction-
pressure levels are

101.74 + 178.02 = 280/F $ p9 = p5 = 49 psia

and

101.74 + (2)(178.02) = 457.8°F $ p12 = p2 = 456 psia

where the extraction pressures have been evaluated using the saturated-steam tables.
After the entropy and enthalpy at state 1 are evaluated, the enthalpy h3s at the HP-
turbine isentropic discharge state 3s is determined from s1 and p3. The HP-turbine
efficiency then yields h3 and the steam tables give s3. The entropy and enthalpy at the
HP-turbine extraction state 2 may be approximated by drawing a straight line on the
steam Mollier diagram connecting states 1 and 3 and finding the intersection with the
HP-extraction pressure P2. This technique may be used for any number of extraction
points in a turbine.
Once the hot reheat properties at state 4 are determined from the steam tables, the
LP-turbine exit and extraction states at 6 and 5 may be obtained by the same method
used for the HP turbine.
68
69

The determination of the FWH condensate temperatures and pressures at states 9


and 12 have already been discussed. The temperatures of the heated feedwater leaving
the FWHs may be determined from the terminal temperature differences:

T11 = T9 - TTD = 281 - 8 = 273°F

T14 = T12 - TTD = 457.5 - 8 = 449.5°F

Recalling that the enthalpy of a subcooled liquid is almost independent of pressure, we


note that the enthalpies h11 and h14 may be found in the saturated-liquid tables at T11 and
T14, respectively.
70

The pump discharge state 8 is a subcooled-liquid state, which may be


approximated in the same way as in Examples 2.3 and 2.4. Thus

h8s = h7 + ( p8 – p7 )v7

= 69.7 + (2000 – 1)(144)(0.016136)/778 = 75.7 Btu/lbm

and

h8 = h7 + ( h8s – h7 )/p = 69.7 + (75.7 – 69.7)/0.8 = 77.2 Btu/lbm

The pump work is then

wp = h7 – h8 = 69.7 – 77.2 = – 7.5 Btu/lbm

The extraction mass-flow fractions designated m1 and m2 relate other flows to the
unit mass flow at the high-pressure-turbine throttle. For example, the condensate flow
rate from the LP heater at state 10 is given by m1 + m2.
The steady-flow First Law of Thermodynamics may now be applied to the heaters.
For the HP FWH:

0 = m1h12 + (1)h14 – m1h2 – (1)hll

may be rewritten as

m1 = ( h14 – hll )/( h2 – h12 ) [dl]

This and the T-s diagram show that the HP extraction-flow enthalpy drop from
state 2 to state 12 provides the heat to raise the enthalpy in the feedwater from state 11
to state 14. Also, for the LP FWH:

0 = (1)h ll + (m2 + m1 )h9 – (1)h8 – m2h5 – mlhl3

becomes

m2 = [ m1( h9 - h13 ) + h11 – h8 ]/( h5 – h9 ) [dl]

This and the T-s diagram show that the discharge from the HP FWH at state 13 aids
the mass flow m2 in heating the LP FWH flow from state 8 to state 11. The values of
m1 and m2 are evaluated at the bottom of spreadsheet in Table 2.3.
71

With all states and flows known, we may now determine some plant performance
parameters. The turbine work referenced to the throttle mass-flow rate is easily
obtained by summing the flow contributions through each section of the turbines:

wt = h1 – h2 + (1 – ml )( h2 – h3) + (1 – ml )( h4 – h5)
+ (1 – ml – m2 )( h5 – h6) [Btu/lbm | kJ /kg]

The net work is then wt + wp, and the heat added in the steam generator is the sum of
heat additions in the feedwater pass and the reheat pass:

qa = h1 - h14 + (1 - ml )( h4 - h3) [Btu/lbm | kJ/kg]

These parameters and the work ratio are evaluated in Table 2.3.
____________________________________________________________________

Example 2.10 shows that a good thermal efficiency and net work output are
possible with the use of two feedwater heaters despite taking into account realistic
turbomachine inefficiencies. The high work ratio clearly demonstrates the low-
compression work requirements of Rankine cycles.

2.7 A Study of a Modern Steam Power Plant

Modern steam power plants incorporate both reheat and feedwater heating. A
flowsheet for the Public Service Company of Oklahoma (PSO) Riverside Station Unit
#1, south of Tulsa, is shown in Figure 2.13. This natural-gas-burning plant was sized
for two nominal 500-megawatt units. Several other plants in the PSO system have
similar unit flowsheets, including a coal-burning plant. Note the flowsheet coding W, H,
F, and A for flow rate in lbm/hr, enthalpy in Btu/lbm, temperature in °F, and pressure in
psia, respectively.
The steam generator, not shown on the flowsheet, interacts through the feedwater
and steam lines on the right-hand side of the diagram. The high pressure turbine throttle
is at 1000°F and 3349 psia and has a mass-flow rate of 2,922,139 lbm/hr. This type of
unit is called supercritical, because the pressure in the main steam line to the HP-
turbine throttle exceeds the 3208.2-psia critical pressure of steam. Note that a large
fraction of the HP-turbine mass-flow rate enters the cold reheat line at 630 psia and is
reheated to the intermediate-pressure (IP) turbine throttle conditions of 1000°F and
567 psia.
Most of the steam flow through the IP turbine passes through the crossover at 186
psia to the double-flow low-pressure (DFLP) turbine. The term double-flow refers to
the fact that the incoming flow enters at the middle, splits, and flows axially in opposite
72
73

directions through the turbine. This causes the large axial force components on the
blades and shaft to oppose each other so that the resultant axial thrust is small and does
not necessitate heavy thrust bearings. The combined HP and IP turbines are similarly
configured.
The plant is equipped with six closed FWHs and one open FWH (the deaerator).
Note that the condensate of each of the closed feedwater heaters is throttled to the next
lowest pressure FWH or, in the case of the lowest-pressure heater, to the condenser.
The extraction steam for the four lowest-pressure FWHs flows from the DFLP turbine.
Extraction steam for the highest pressure FWH is provided by the HP turbine, and the
IP turbine supplies heater HTR1-6 and the open feedwater heater identified as the
deaerator. The deaerator is specially designed to remove non-condensable gases from
the system, in addition to performing its feedwater heating duties.
The feedwater starts at the "hot well" of the condenser on the left of the diagram,
enters the condensate pump at 101.1°F and 2"Hg abs., and starts its passage through
the FWHs. Note that the feedwater increases in temperature from 102.1° to 180°,
227.2°, 282.7°, and 314.4° in passing through the 4 lowest pressure FWHs. The
feedwater from the deaerator is pumped to 405 psia by the booster pump and
subsequently to 3933 psia by the boiler feed pump (BFP). The BFP exit pressure
exceeds the HP-turbine throttle pressure of 3349 psia in order to overcome flow losses
in the high pressure heater, the boiler feed line, the steam generator main steam pass,
and the main steam line, all of which operate at supercritical pressure.
The boiler feed pump turbine (BFPT) shown in the upper left of the diagram
supplies the shaft power to drive the BFP at the lower right. The BFPT receives steam
from an extraction line of the DFLP turbine and exhausts directly to the condenser.
The reader should study Figure 2.13 thoroughly in the light of the preceding
discussions of reheat and feedwater heating. It is particularly useful to consider the flow
rates with respect to mass and energy conservation. Mastery of this flow sheet will
make it possible to quickly understand flowsheets of other major power plants.

Example 2.11

Verify that the steam generator feedwater flow rate satisfies the conservation of mass
into all the feedwater heaters shown for the Riverside Unit #1 in Figure 2.13. You may
neglect all flows of less than 2000 lbm/hr.

Solution
The shell side of the low pressure heater, labeled HTR1-1, receives condensate
from heaters 2, 3 and 4 as well as steam entering from the LP turbine. The total
condensate from the low-pressure heaters into the condenser are:
74

Source Flow rate, lbm/hr


_______________________________________________
Condensate from HTR1-4 75,005

Extraction steam into HTR1-3 125,412

Extraction steam into HTR1-2 102,897


----------
Total condensate into HTR1-1 303,314

Extraction steam into HTR1-1 157,111


---------
Total condensate leaving HTR1-1 460,425

The feedwater flow rate through the four low-pressure heaters (the condenser
condensate pump flow rate) is the sum of the flows into the condenser:

460,425 + 162,701 + 1,812,971 = 2,436,097 lbm/hr.

An easier approach to evaluating this flow rate is by imagining a control volume


around the entire left side of the diagram that cuts it in two parts between the deaerator
and HTR1-4 and through the crossover steam line. Because these are the only points
where the control volume is penetrated by large mass flows, the two flows must be
equal. Consequently the crossover mass-flow rate of 2,434,357 lbm/hr agrees very well
with our above calculation of the feedwater flow rate into the deaerator.
Now, observing that the boiler feedwater all flows from the deaerator through the
booster pump, we sum all of the flows into the deaerator:

Feedwater into deaerator 2,434,357

Steam to deaerator 148,321

Steam to HTR1-6 107,661

Steam to HTR1-7 222,876


-----------
Total feedwater into HTR1-7 2,913,215 lbm/hr

This compares well with the tabulated value of 2,922,139 lbm /hr to the steam
generator. Accounting for the small flows should improve the agreement.
______________________________________________________________
75

2.8 Deviations from the Ideal - Pressure Losses

It is evident from study of Figure 2.13 that there are significant pressure drops in the
flows through the steam generator between the HP FWH and the HP-turbine throttle
and in the reheat line between the HP and IP turbines. While we have neglected such
losses in our calculations, final design analysis requires their consideration. A first
attempt at this may be made by applying a fractional pressure drop based on
experience. Two per cent pressure drops through the main steam and feedwater lines
and a 3.7% loss through the steam generator would, for instance, account for the
indicated 14.8% loss from the boiler feed pump to the HP turbine.
In the final analysis, of course, when realistic values are available for flow rates and
properties, known fluid mechanic relations for pressure drop may be employed to
account for these losses.

Bibliography and References

1. Anon., Steam, Its Generation and Use. New York: Babcock and Wilcox, 1978.

2. Singer, J. G., (Ed.), Combustion/Fossil Power Systems. Windsor, Conn.:


Combustion Engineering, 1981.

3. Wood, Bernard, Applications of Thermodynamics, 2nd ed. Reading, Mass.: Addison-


Wesley, 1981.

4. Li, Kam W., and Priddy, A. Paul, Powerplant System Design. New York: Wiley,
1985.

5. El-Wakil, M. M., Power Plant Technology. New York: McGraw-Hill, 1984.

6. Skrotzi, B. G. A. and Vopat, W. A., Power Station Engineering and Economy.


New York: McGraw-Hill, 1960.

EXERCISES

2.1 An ideal Rankine-cycle steam power plant has 800-psia saturated steam at the
turbine throttle and 5-psia condenser pressure. What are the turbine work, pump work,
net work, steam generator heat addition, thermal efficiency, maximum cycle
temperature, and turbine exit quality? What is the Camot efficiency corresponding to
the temperature extremes for this cycle?

2.2 A Rankine-cycle steam power plant has an 800-psia/900/F throttle and 5-psia
condenser pressure. What are the net work, turbine work, pump work, steam generator
76

heat addition, thermal efficiency, and turbine exit quality? What is the Carnot efficiency
corresponding to the temperature extremes for this cycle?

2.3 Solve Exercise 2.2 for the cases of (a) an 85% efficient turbine, (b) an 85%
efficient pump, and (c) both together. Tabulate and discuss your results together with
those of Exercise 2.2.

2.4 Solve Exercise 2.2 for the case of (a) 1000/F throttle, (b) 2000-psia throttle,
(c) 2-psia condenser, and (d) all three changes simultaneously. Make a table comparing
net work, quality, and thermal efficiency, including the results of Exercise 2.2. What
conclusions can you draw from these calculations?

2.5 Sketch coordinated, labeled flow and T-s diagrams for the ideal Rankine cycle.
Tabulate the temperatures, entropies, pressures, enthalpies, and quality or degree of
superheat for each significant state shown on the diagram for a throttle at 1000 psia and
1000/F and a condenser at 5 psia. Determine the net work, heat added, thermal
efficiency, heat rate, and heat rejected in the condenser. If the power plant output is
100 megawatts and the condenser cooling-water temperature rise is 15 Rankine
degrees, what is the steam flow rate and cooling-water flow rate? Neglect pump work.

2.6 Consider a simple Rankine cycle with a 2000-psia/1100/F throttle and 1-psia
condenser. Compare the thermal efficiencies and net work for cycles with a perfect
turbine and one having 86% turbine isentropic efficiency. Assume isentropic pumping.

2.7 A boiling-water reactor operates with saturated vapor at 7500 kPa at the throttle
of the high-pressure turbine. What is the lowest turbine exit pressure that ensures that
the turbine exit moisture does not exceed 12% if the turbine is isentropic? What would
the lowest pressure be if the turbine isentropic efficiency were 85%?

2.8 Consider a steam plant with a single reheat and a single open feedwater heater that
takes extraction from the cold reheat line. Sketch carefully coordinated and labeled T-s
and flow diagrams. If the throttle is at 1000/F and 3000psia, the condenser is at 1 psia,
and reheat is to 1000/F at 400 psia, what is the extraction mass fraction, the heat rate,
and the thermal efficiency? The turbine efficiency is 89%. Neglect pump work.

2.9 A Rankine-cycle power plant condenses steam at 2 psia and has 1000/F and 500
psia at the turbine throttle. Assume an isentropic turbine.
(a) Tabulate the temperature, pressure, entropy, and enthalpy of all states. Determine
the quality and moisture fraction for all mixed states.
(b) Calculate the heat transferred in the condenser and the steam generator and the
turbine work, all per unit mass. What is the thermal efficiency?
(c) Calculate the pump work. What is the ratio of turbine to pump work?
77

(d) What is the turbine work and thermal efficiency if the turbine efficiency is 85%?
Include pump work.

2.10 For throttle conditions of 1000/F and 1000 psia and a condenser pressure of 2
psia, compare the net work, thermal efficiency, and turbine discharge quality or degree
of superheat for a simple cycle and two reheat cycles with reheat to 1000/F at 50 and
200 psia. Tabulate your results. Sketch a single large, labeled T-s diagram comparing
the cycles. Turbine isentropic efficiencies are 85%.

2.11 Consider a regenerative Rankine cycle with a 1000/F and 500-psia throttle, 2-psia
condenser, and an open feedwater heater operating between two turbines at 50 psia.
Turbine efficiencies are 85%. Neglect pump work.
(a) Draw labeled, coordinated T-s and flow diagrams.
(b) Determine the fraction of the throttle mass flow that passes through the extraction
line.
(c) Calculate the turbine work per unit mass at the throttle.
(d) Calculate the cycle efficiency, and compare it to the simple-cycle efficiency.

2.12 Consider a 1120°F, 2000-psia, 10-psia steam cycle with reheat at 200 psia to
1000/F and a closed feedwater heater taking extraction from a line between two
turbines at 100psia. The FWH condensate is throttled to the condenser, and the
feedwater in the FWH is raised to the condensing temperature of the extraction steam.
(a) Draw labeled T-s and flow diagrams for this plant.
(b) Tabulate the enthalpies for each significant state point.
(c) What is the extraction fraction to the FWH?
(d) What are the net work and work ratio?
(e) What are the thermal efficiency and the heat rate?

2.13 A turbine operates with a 860/F, 900-psia throttle. Calorimetric measurements


indicate that the discharge enthalpy is 1250 Btu/lbm at 100 psia. What is the isentropic
efficiency?

2.14 An ideal Rankine cycle has 1000-psia saturated steam at the turbine throttle. The
condenser pressure is 10psia. What are the turbine work, steam generator heat addition,
maximum cycle temperature, turbine exit quality, and Carnot efficiency corresponding
to the temperature extremes of the cycle? Neglect pump work.

2.15 Assume that the extraction mass-flow rate to FWH #7 in Figure 2.13 is not
known. Calculate the FWH extraction mass fraction (relative to the HP-turbine throttle
flow) and the extraction mass-flow rate. Compare the extraction-steam energy loss rate
with the feedwater energy gain rate.
78

2.16 Compare the inflow and outflow of steam of the DFLP turbine in Figure 2.13,
and calculate the percentage difference. Calculate the power output of the DFLP
turbine in Btu/hr and in kW.

2.17 Calculate the power delivered by the PSO Riverside Unit #1 boiler feed pump
turbine, BFPT. Based on the feedwater enthalpy rise across the BFP, determine its
power requirements, in kilowatts. What fraction of the plant gross output is used by the
BFPT?

2.18 Without performing a detailed analysis of the FWHs, determine the PSO
Riverside Unit #1 feedwater flow rate from heater number 4 to the deaerator. Explain
your methodology.

2.19 Total and compare the inflows and outflows of mass and energy to the PSO
Riverside Unit #1 deaerator.

2.20 Rework Example 2.4 neglecting pump work. Repeat your calculations for an 80%
efficient pump. Compare and comment on the significance of accounting for pump
work and turbomachine efficiency.

2.21 For a 1080/F, 2000-psia, 5-psia Rankine cycle with 85% turbine efficiency and
60% pump efficiency:
(a) Compare the actual net work and the isentropic turbine work and the isentropic net
work.
(b) Calculate the actual heat transfer and work for each component, and evaluate the
cyclic integrals of Q and W.
(c) Compare the real cycle efficiency with that for the ideal Rankine cycle.

2.22 For a 1080/F, 2000-psia, 5-psia Rankine cycle with 85% turbine efficiency and
60% pump efficiency, evaluate the effect of a single reheat to 1080/F at 500 psia on:
(a) Heat addition in the steam generator.
(b) Work of each turbine, total turbine work, and net work. Compare the net work
with the cyclic integral of the external transfers of heat.
(c) Cycle efficiency and heat rate.
(d) Quality or degree of superheat at the exit of the turbines.
Draw labeled flow and T-s diagrams.

2.23 Consider a 1080/F, 2000-psia, 5-psia Rankine cycle with 85% turbine efficiency
and 60% pump efficiency. Compare the simple cycle with the same cycle operating with
a single reheat to 1080/F at 1000 psia with respect to:
(a) Heat addition in the steam generator.
(b) Work of each turbine, total turbine work and net work, condenser heat rejection,
and cyclic integral of heat added.
79

(c) Cycle efficiency.


(d) Quality or degree of superheat at the exit of the turbines.
Draw labeled flow and T-s diagrams.

2.24 For a 1080/F, 2000-psia, 5-psia Rankine cycle with 85% turbine efficiencies and
60% pump efficiencies and using a single open feedwater heater operating at 500 psia:
(a) Draw labeled and coordinated flow and T-s diagrams.
(b) Evaluate the feedwater heater mass fraction.
(c) Evaluate heat addition in the steam generator, work of each turbine, total turbine
work, and net work, all per pound of steam at the HP-turbine throttle.
(d) Evaluate condenser heat transfer per unit mass at the HP-turbine throttle.
(e) Evaluate cycle efficiency and heat rate. Compare with simple-cycle efficiency.
(f) Evaluate the cyclic integral of the differential heat addition, and compare it with
the net work.

2.25 Consider a 1080/F, 2000-psia, 5-psia Rankine reheat-regenerative cycle with


perfect turbomachinery and a closed feedwater heater taking extraction from the cold
reheat line at 500 psia. FWH condensate is pumped into the feedwater line downstream
of the feedwater heater. Assume that the enthalpy of the feedwater entering the steam
generator is that of the saturated liquid leaving the FWH.
(a) Draw coordinated and labeled flow and T-s diagrams.
(b) Determine the extraction mass fraction, the net work, and the total heat addition.
(c) Determine the thermal efficiency and heat rate.
(d) Determine the superheat or quality at the turbine exhausts:

2.26 Taking the reheat-pressure level as a variable, plot net work, thermal efficiency,
and turbine exhaust superheat and/or moisture against reheat pressure for the
conditions of Example 2.5. Select a suitable design value based on your analysis.

2.27 Solve Example 2.6 for 1200/F throttle substituting a closed FWH for the open
heater. Consider two cases in which the FWH condensate is (a) throttled to the
condenser, and (b) pumped to throttle pressure.

2.28 Solve Example 2.6 using the method for assigning extraction-pressure levels
given in the subsection of Section 2.6 on multistage extraction systems.

2.29 Solve Example 2.7 using the method for assigning extraction-pressure levels
given in the section on multistage extraction systems, and determine by trial and error
the reheat-pressure level that maximizes the thermal efficiency.

2.30 Solve Example 2.7 with the extraction condensate from the closed FWH pumped
ahead to the feedwater-pressure level.
80

2.31 Solve Example 2.6 for 900/F throttle temperature with the open FWH replaced
by a closed FWH where the feedwater is (a) throttled to the condenser, and (b) pumped
into the feedwater line downstream of the FWH.

2.32 Compare the work and exhaust quality of 90% efficient turbines with 2500-psia
throttle pressure and 1000/F and 1200/F throttle temperatures exiting to a 2-psia
condenser.

2.33 Draw a large T-s diagram showing the states associated with the important flows
of the PSO Riverside Unit #1 (Figure 2.13).

2.34 A Rankine-cycle steam power plant has 5-MPa saturated steam at the turbine
throttle and 25-kPa condenser pressure. What are the net work, steam generator heat
addition, thermal efficiency, heat rate, maximum cycle temperature, and turbine exit
quality? What is the Carnot efficiency corresponding to the temperature extremes for
this cycle?

2.35 A Rankine-cycle steam power plant has a 5-MPa/450/C throttle and 10-kPa
condenser pressure. What are the net work, steam generator heat addition, thermal
efficiency, heat rate, and turbine exit quality? What is the Carnot efficiency
corresponding to the temperature extremes for this cycle?

2.36 Solve Exercise 2.35 for the cases of (a) an 85% efficient turbine, (b) an 85%
efficient pump, and (c) both together. What conclusions may be inferred from your
results?

2.37 Solve Exercise 2.35 for the case of (a) a 550/C throttle, (b) a 15-MPa throttle,
(c) a 5-kPa condenser, and (d) all three changes simultaneously. What conclusions can
you draw from these calculations?

2.38 Sketch coordinated, labeled flow and T-s diagrams for the following Rankine
cycle. Tabulate the temperatures, entropies, pressures, enthalpies, and quality or degree
of superheat for each significant state shown on the diagram for a throttle at 10 MPa
and 550/C and condenser at 5 kPa. Determine the net work, heat added, thermal
efficiency, and heat rejected in the condenser. If the power plant output is 100
megawatts and the condenser cooling water temperature rise is 15/Rankine, what is the
steam flow rate and cooling-water flow rate? Neglect pump work.

2.39 Consider a Rankine cycle with a 20MPa/600/C throttle and 3-kPa condenser.
Compare the thermal efficiencies and net work for cycles with a perfect turbine and one
having 86% turbine isentropic efficiency.
81

2.40 Consider a steam plant, with a single reheat and a single open feedwater heater,
that takes extraction from the cold reheat line. Sketch carefully coordinated and labeled
T-s and flow diagrams. If the throttle is at 550/C and 15 MPa, the condenser is at 5
kPa, and reheat is to 3 MPa and 550° C, what are the extraction mass fraction, work
ratio, and thermal efficiency? The pump and turbine efficiencies are 82% and 89%,
respectively.

2.41 A Rankine-cycle power plant condenses steam at 10 kPa and has 550/C and 5
MPa at the turbine throttle. Assume an isentropic turbine.
(a) Tabulate the temperature, pressure, entropy, and enthalpy of all states. Determine
the quality and moisture fraction for all mixed states.
(b) Calculate the heat transferred in the condenser and steam generator and the turbine
work, all per unit mass. What is the thermal efficiency?
(c) Calculate the pump work. What is the ratio of turbine to pump work?
(d) What is the turbine work and thermal efficiency if the turbine efficiency is 85%?
Include pump work.

2.42 For throttle conditions of 550°C and 5 MPa and a condenser pressure of 10 kPa,
compare the net work, thermal efficiency, and turbine discharge quality or degree of
superheat for a simple cycle and two reheat cycles. Consider reheat to 500/C at (a)
4MPa and (b) 1 MPa. Tabulate and compare your results. Sketch a large, labeled T-s
diagram for a reheat cycle. Turbine efficiencies are 85%.

2.43 Consider a regenerative Rankine cycle with a 600/C and 4-MPa throttle, a 5-kPa
condenser, and an open feedwater heater at 500 kPa. Turbine efficiencies are 85%.
Neglect pump work.
(a) Draw labeled, coordinated T-s and flow diagrams.
(b) Determine the fraction of the throttle mass flow that passes through the extraction
line.
(c) Calculate the turbine work per unit mass at the throttle.
(d) Calculate the cycle efficiency, and compare it with the simple-cycle efficiency.
(e) Calculate the heat rate.

2.44 Consider a 600/C, 15-MPa steam cycle with reheat at 2 MPa to 600/C and
extraction to a closed feedwater heater at 600 kPa. The FWH condensate is throttled to
the condenser at 5 kPa, and the feedwater in the FWH is raised to the condensing
temperature of the extraction steam. Neglecting pump work:
(a) Draw labeled T-s and flow diagrams for this plant.
(b) Tabulate the enthalpies for each significant state point.
(c) What is the extraction fraction to the FWH?
(d) What is the net work?
(e) What is the thermal efficiency?
82

(f) What is the heat rate?

2.45 A turbine operates with a 600°C, 7-MPa throttle. Calorimetric measurements


indicate that the discharge enthalpy is 3050 kJ/kg at 0.8 MPa. What is the turbine
isentropic efficiency?

2.46 A pressurized water-reactor nuclear power plant steam generator has separate
turbine and reactor water loops. The steam generator receives high-pressure hot water
from the reactor vessel to heat the turbine feedwater. Steam is generated from the
feedwater in the turbine loop. The water pressure in the reactor is 15 MPa, and the
water temperature in and out of the reactor is 289°C and 325°C, respectively. The plant
has one turbine with a single extraction to a closed FWH with condensate throttled to
the condenser. Throttle conditions are 300/C and 8 MPa. The extraction and condenser
pressures are 100 kPa and 5 kPa, respectively. The reactor-coolant flow rate is 14,000
kg/s. Assume no heat losses in heat exchangers and isentropic turbomachines. Neglect
pump work.
(a) What is the rate of heat transfer from the reactor in MWt?
(b) Draw coordinated flow and T-s diagrams that show both loops.
(c) Determine the extraction mass fraction of the throttle flow rate.
(d) Determine the cycle net work, heat rate, and thermal efficiency.
(e) Calculate the steam flow rate.
(f) Assuming the electrical generator has 97% efficiency, calculate the power output,
in MWe (electric).

2.47 Perform an optimization of the extraction pressure of a Rankine cycle with a


2000-psia saturated-vapor throttle, a 1-psia condenser with a single closed feedwater
heater, as in Example 2.9. Compare the optimum extraction temperature given by
Equation (2.14) with your results.

2.48 Prepare an optimization study of thermal efficiency with a table and plot of net
work and thermal efficiency as a function of reheat pressure level for Example 2.5.
Discuss the selection of reheat pressure for this case. How does the reheat pressure
used in Example 2.5 compare with your results?

2.49 Solve Exercise 2.25 for reheat and extraction at 200 psia. Compare the extraction
mass fraction, net work, thermal efficiency, heat rate, and turbine exit conditions with
those of Exercise 2.25.

2.50 Rework Exercise 2.25, accounting for 90% turbine efficiencies and a 10/F
terminal temperature difference.

2.51 A 1000/F/2000-psia-throttle high-pressure turbine discharges into a cold reheat


line at 200 psia. Reheat is to 1000/F. The low-pressure turbine discharges into the
83

condenser at 0.5 inches of mercury absolute. Both turbines are 90% efficient. Design
the cycle for the use of three feedwater heaters. Draw coordinated T-s and flow
diagrams. State and discuss your decisions on the handling of the feedwater heater
design.

2.52 A steam turbine receives steam at 1050/F and 3000 psia and condenses at 5 psia.
Two feedwater heaters are supplied by extraction from the turbine at pressures of 1000
psia and 200 psia. The low-pressure heater is an open FWH, and the other is closed
with its condensate throttled to the open heater. Assuming isentropic flow in the turbine
and negligible pump work:
(a) Sketch accurately labeled and coordinated T-s and flow diagrams for the system,
and create a table of temperature, pressure, and enthalpy values for each state.
(b) What are the extraction flows to each feedwater heater if the throttle mass flow
rate is 250,000 pounds per hour?
(c) How much power, in kW, is produced by the turbine?
(d) Compare the thermal efficiency of the system with the efficiency if valves of both
extraction lines are closed.
(e) What is the heat rate of the system with both feedwater heaters operative?

2.53 Apply the steady-flow First Law of Thermodynamics to a single control volume
enclosing the two turbines in Example 2.7. Show that the same equation is obtained for
the turbine work as when the work of individual turbines is summed.

2.54 Apply the steady-flow First Law of Thermodynamics to a single control volume
enclosing the two turbines in Example 2.10. Show that the same equation is obtained
for the turbine work as when the work of individual turbines is summed.

2.55 Resolve Example 2.7 with 4% pressure drops in the main steam pass and reheat
pass through the steam generator. Make a table comparing your results with those of
the example to show the influence of the losses on plant performance. Calculate and
display the percentage differences for each parameter. Assume turbine throttle
conditions are unchanged.

2.56 Draw labeled and coordinated T-s and flow diagrams for a steam power plant
with 1000°F / 3000-psia / 2" Hg absolute conditions, assuming isentropic
turbomachinery. The plant has reheat at 500 psia to ll00/F. The plant has the following
feedwater heaters:
1. A closed FWH with extraction at 1000 psia and pumped condensate.
2. A closed FWH at 400 psia with condensate throttled into the next-lowest-
pressure FWH.
3. An open FWH at 20 psia.
Define mass fraction variables. Show mass-flow variable expressions, with arrows
indicating mass fractions along the various process paths on the T-s and flow diagrams.
84

Write equations for conservation of energy for the FWHs that allow you to solve easily
for the mass fractions in terms of known state enthalpies and other mass fractions.
Indicate a solution method for the mass fractions that involves simple substitution only.

2.57 A pressurized-water nuclear-reactor steam generator has separate turbine and


reactor loops. The steam generator linking the two loops cools high-pressure hot water
from the reactor vessel and transfers the heat to the turbine feedwater producing steam.
The water pressure in the reactor is 2250 psia, and the water temperatures in and out of
the reactor are 559°F and 623°F, respectively. The plant has one turbine with a single
extraction to an open FWH. Throttle conditions are 555/F and 1100 psia. The
extraction and condenser pressures are 100 psia and 1 psia, respectively. The reactor-
coolant flow rate is 147,000,000 lbm/hr. Assume no heat losses in heat exchangers and
isentropic turbomachines.
(a) What is the rate of heat transfer from the reactor, in Btu/hr and in MWt?
(b) Draw coordinated flow and T-s diagrams that show both loops with states in their
proper relations with respect to each other.
(c) Determine the extraction mass fraction relative to the throttle flow rate.
(d) Determine the cycle net work.
(e) What are the cycle thermal efficiency and heat rate?
(f) Calculate the turbine-steam flow rate.
(g) Assuming the electrical generator has 100% efficiency, calculate the turbine power,
in Btu/hr and in MWe.

2.58 Determine the efficiencies of the boiler feed pump and boiler feed pump turbine of
the PSO Riverside Station Unit 4/1 (Figure 2.13).

2.59 A Rankine cycle with a single open feedwater heater has a 1040/F and 550-psia
throttle. Extraction from the exit of the first turbine (assumed isentropic) is at 40/F of
superheat. The second turbine has an efficiency of 85% and expands into the condenser
at 5 psia.
(a) Draw matched, labeled T-s and flow diagrams.
(b) Accurately calculate and tabulate the enthalpies of all significant states. Neglect
pump work.
(c) What is the feedwater-heater mass fraction relative to the mass flow at the first
throttle?
(d) What is the quality or degree of superheat at the condenser inlet?
(e) What are the net work, thermal efficiency, and heat rate?
(f) Estimate the feedwater-heater condensate pump work and its percentage of turbine
work.
85

CHAPTER 3

FUELS AND COMBUSTION

3.1 Introduction to Combustion

Combustion Basics

The last chapter set forth the basics of the Rankine cycle and the principles of operation
of steam cycles of modern steam power plants. An important aspect of power
generation involves the supply of heat to the working fluid, which in the case of steam
power usually means turning liquid water into superheated steam. This heat comes from
an energy source. With the exception of nuclear and solar power and a few other exotic
sources, most power plants are driven by a chemical reaction called combustion, which
usually involves sources that are compounds of hydrogen and carbon. Process
industries, businesses, homes, and transportation systems have vast heat requirements
that are also satisfied by combustion reactions. The subject matter of this chapter
therefore has wide applicability to a variety of heating processes.

Combustion is the conversion of a substance called a fuel into chemical compounds


known as products of combustion by combination with an oxidizer. The combustion
process is an exothermic chemical reaction, i.e., a reaction that releases energy as it
occurs. Thus combustion may be represented symbolically by:

Fuel + Oxidizer Y Products of combustion + Energy

Here the fuel and the oxidizer are reactants, i.e., the substances present before the
reaction takes place. This relation indicates that the reactants produce combustion
products and energy. Either the chemical energy released is transferred to the
surroundings as it is produced, or it remains in the combustion products in the form of
elevated internal energy (temperature), or some combination thereof.
Fuels are evaluated, in part, based on the amount of energy or heat that they
release per unit mass or per mole during combustion of the fuel. Such a quantity is
known as the fuel's heat of reaction or heating value.
Heats of reaction may be measured in a calorimeter, a device in which chemical
energy release is determined by transferring the released heat to a surrounding fluid.
The amount of heat transferred to the fluid in returning the products of combustion to
their initial temperature yields the heat of reaction.
86

In combustion processes the oxidizer is usually air but could be pure oxygen, an
oxygen mixture, or a substance involving some other oxidizing element such as
fluorine. Here we will limit our attention to combustion of a fuel with air or pure
oxygen.
Chemical fuels exist in gaseous, liquid, or solid form. Natural gas, gasoline, and
coal, perhaps the most widely used examples of these three forms, are each a complex
mixture of reacting and inert compounds. We will consider each more closely later in
the chapter. First let's review some important fundamentals of mixtures of gases, such
as those involved in combustion reactions.

Mass and Mole Fractions

The amount of a substance present in a sample may be indicated by its mass or by the
number of moles of the substance. A mole is defined as the mass of a substance equal to
its molecular mass or molecular weight. A few molecular weights commonly used in
combustion analysis are tabulated below. For most combustion calculations, it is
sufficiently accurate to use integer molecular weights. The error incurred may easily be
evaluated for a given reaction and should usually not be of concern. Thus a gram-mole
of water is 18 grams, a kg-mole of nitrogen is 28 kg, and a pound-mole of sulfur is 32
lbm.
_____________________________________________________________________
Molecule Molecular Weight
-------------------------------------------
C 12
N2 28
O2 32
S 32
H2 2
_____________________________________________________________________

The composition of a mixture may be given as a list of the fractions of each of the
substances present. Thus we define the mass fraction, of a component i, mfi, as the
ratio of the mass of the component, mi, to the mass of the mixture, m:

mfi = mi /m

It is evident that the sum of the mass fractions of all the components must be 1. Thus

mf1 + mf2 + ... = 1

Analogous to the mass fraction, we define the mole fraction of component i, xi, as
the ratio of the number of moles of i, ni, to the total number of moles in the mixture, n:

xi = ni /n
87

The total number of moles, n, is the sum of the number of moles of all the components
of the mixture:

n = n1 + n2 + ...

It follows that the sum of all the mole fractions of the mixture must also equal 1.

x1 + x2 + ... = 1

The mass of component i in a mixture is the product of the number of moles of i and its
molecular weight, Mi. The mass of the mixture is therefore the sum, m = n1M1 + n2M2 +
..., over all components of the mixture. Substituting xin for ni, the total mass becomes

m = (x1M1 + x2M2 + ...)n

But the average molecular weight of the mixture is the ratio of the total mass to the
total number of moles. Thus the average molecular weight is

M = m /n = x1M1 + x2M2 + ...

EXAMPLE 3.1

Express the mass fraction of component 1 of a mixture in terms of: (a) the number of
moles of the three components of the mixture, n1, n2, and n3, and (b) the mole fractions
of the three components. (c) If the mole fractions of carbon dioxide and nitrogen in a
three component gas containing water vapor are 0.07 and 0.38, respectively, what are
the mass fractions of the three components?

Solution
(a) Because the mass of i can be written as mi = niMi , the mass fraction of component
i can be written as:

mfi = niMi /(n1M1 + n2M2 + ..) [dl]

For the first of the three components, i = 1, this becomes:

mf1 = n1M1/(n1M1 + n2M2 + n3M3)

Similarly, for i = 2 and i = 3:

mf2 = n2M2/(n1M1 + n2M2 + n3M3)

mf3 = n3M3/(n1M1 + n2M2 + n3M3)


88

(b) Substituting n1 = x1 n, n2 = x2 n, etc. in the earlier equations and simplifying, we


obtain for the mass fractions:

mf1 = x1M1/(x1M1 + x2M2 + x3M3)

mf2 = x2M2/(x1M1 + x2M2 + x3M3)

mf3 = x3M3 /(x1M1 + x2M2 + x3M3)

(c) Identifying the subscripts 1, 2, and 3 with carbon dioxide, nitrogen, and water
vapor, respectively, we have x1 = 0.07, x2 = 0.38 and x3 = 1 – 0.07 – 0.038 = 0.55.
Then:

mf1 = (0.07)(44)/[(0.07)(44) + (0.38)(28) + (0.55)(18)]

= (0.07)(44)/(23.62) = 0.1304

mf2 = (0.38)(28)/(23.62) = 0.4505

mf3 = (0.55)(18)/(23.62) = 0.4191

As a check we sum the mass fractions: 0.1304 + 0.4505 + 0.4191 = 1.0000.


________________________________________________________________

For a mixture of gases at a given temperature and pressure, the ideal gas law
shows that pVi = niúT holds for any component, and pV = núT for the mixture as a
whole. Forming the ratio of the two equations we observe that the mole fractions have
the same values as the volume fraction:

xi = Vi /V = ni /n [dl]

Similarly, for a given volume of a mixture of gases at a given temperature, piV = niúT
for each component and pV = núT for the mixture. The ratio of the two equations
shows that the partial pressure of any component i is the product of the mole fraction
of i and the pressure of the mixture:

pi = pni /n = pxi

EXAMPLE 3.2

What is the partial pressure of water vapor in Example 3.1 if the mixture pressure is
two atmospheres?
89

Solution
The mole fraction of water vapor in the mixture of Example 3.1 is 0.55. The partial
pressure of the water vapor is therefore (0.55)(2) = 1.1 atm.
_____________________________________________________________________

Characterizing Air for Combustion Calculations

Air is a mixture of about 21% oxygen, 78% nitrogen, and 1% other constituents by
volume. For combustion calculations it is usually satisfactory to represent air as a 21%
oxygen, 79% nitrogen mixture, by volume. Thus for every 21 moles of oxygen that
react when air oxidizes a fuel, there are also 79 moles of nitrogen involved. Therefore,
79/21 = 3.76 moles of nitrogen are present for every mole of oxygen in the air.
At room temperature both oxygen and nitrogen exist as diatomic molecules, O2
and N2, respectively. It is usually assumed that the nitrogen in the air is nonreacting at
combustion temperatures; that is, there are as many moles of pure nitrogen in the
products as there were in the reactants. At very high temperatures small amounts of
nitrogen react with oxygen to form oxides of nitrogen, usually termed NOx. These small
quantities are important in pollution analysis because of the major role of even small
traces of NOx in the formation of smog. However, since these NOx levels are
insignificant in energy analysis applications, nitrogen is treated as inert here.
The molecular weight of a compound or mixture is the mass of 1 mole of the
substance. The average molecular weight, M, of a mixture, as seen earlier, is the linear
combination of the products of the mole fractions of the components and their
respective molecular weights. Thus the molecular weight for air, Mair, is given by the
sum of the products of the molecular weights of oxygen and nitrogen and their
respective mole fractions in air. Expressed in words:

Mair = Mass of air/Mole of air = (Moles of N2 /Mole of air)(Mass of N2 /Mole of N2)

+ (Moles of O2/Mole of air)(Mass of O2 /Mole of O2)


or

Mair = 0.79 Mnitrogen + 0.21 Moxygen

= 0.79(28) + 0.21(32) = 28.84

The mass fractions of oxygen and nitrogen in air are then

mfoxygen = (0.21)(32)/28.84 = 0.233, or 23.3%

and

mfnitrogen = (0.79)(28)/28.84 = 0.767, or 76.7%


90

3.2 Combustion Chemistry of a Simple Fuel

Methane, CH4, is a common fuel that is a major constituent of most natural gases.
Consider the complete combustion of methane in pure oxygen. The chemical reaction
equation for the complete combustion of methane in oxygen may be written as:

CH4 + 2O2 Y CO2 + 2H2O (3.1)

Because atoms are neither created nor destroyed, Equation (3.1) states that methane
(consisting of one atom of carbon and four atoms of hydrogen) reacts with four atoms
of oxygen to yield carbon dioxide and water products with the same number of atoms
of each element as in the reactants. This is the basic principle involved in balancing all
chemical reaction equations.
Carbon dioxide is the product formed by complete combustion of carbon through
the reaction C + O2 Y CO2. Carbon dioxide has only one carbon atom per molecule.
Since in Equation (3.1) there is only one carbon atom on the left side of the equation,
there can be only one carbon atom and therefore one CO2 molecule on the right.
Similarly, water is the product of the complete combustion of hydrogen. It has two
atoms of hydrogen per molecule. Because there are four hydrogen atoms in the
reactants of Equation (3.1), there must be four in the products, implying that two
molecules of water formed. These observations require four atoms of oxygen on the
right, which implies the presence of two molecules (four atoms) of oxygen on the left.
The coefficients in chemical equations such as Equation (3.1) may be interpreted as
the number of moles of the substance required for the reaction to occur as written.
Thus another way of interpreting Equation (3.1) is that one mole of methane reacts
with two moles of oxygen to form one mole of carbon dioxide and two moles of water.
While not evident in this case, it is not necessary that there be the same number of
moles of products as reactants. It will be seen in numerous other cases that a different
number of moles of products is produced from a given number of moles of reactants.
Thus although the numbers of atoms of each element must be conserved during a
reaction, the total number of moles need not. Because the number of atoms of each
element cannot change, it follows that the mass of each element and the total mass must
be conserved during the reaction. Thus, using the atomic weights (masses) of each
element, the sums of the masses of the reactants and products in Equation (3.1) are
both 80:

CH4 + 2O2 Y CO2 + 2H2O


[12 + 4(1)] + 4(16) Y [12 + 2(16)] + 2[2(1) + 16] = 80

Other observations may be made with respect to Equation (3.1). There are 2 moles of
water in the 3 moles of combustion products, and therefore a mole fraction of water in
the combustion products of xwater = 2/3 = 0.667. Similarly, xCarbon dioxide = 1/3 = 0.333
moles of CO2 in the products.
There are 44 mass units of CO2 in the 80 mass units of products for a mass
91

fraction of CO2 in the products,

mfcarbon dioxide = 44/80 = 0.55

Likewise, the mass fraction of water in the products is 2(18)/80 = 0.45.


We also observe that there are 12 mass units of carbon in the products and
therefore a carbon mass fraction of 12/80 = 0.15. Note that because the mass of any
element and the total mass are conserved in a chemical reaction, the mass fraction of
any element is also conserved in the reaction. Thus the mass fraction of carbon in the
reactants is 0.15, as in the products.

Combustion in Air

Let us now consider the complete combustion of methane in air. The same combustion
products are expected as with combustion in oxygen; the only additional reactant
present is nitrogen, and it is considered inert. Moreover, because we know that in air
every mole of oxygen is accompanied by 3.76 moles of nitrogen, the reaction equation
can be written as

CH4 + 2O2 + 2(3.76)N2 Y CO2 + 2H2O + 2(3.76)N2 (3.2)

It is seen that the reaction equation for combustion in air may be obtained from the
combustion equation for the reaction in oxygen by adding the appropriate number of
moles of nitrogen to both sides of the equation.
Note that both Equations (3.1) and (3.2) describe reactions of one mole of
methane fuel. Because the same amount of fuel is present in both cases, both reactions
release the same amount of energy. We can therefore compare combustion reactions in
air and in oxygen. It will be seen that the presence of nitrogen acts to dilute the
reaction, both chemically and thermally. With air as oxidizer, there are 2 moles of water
vapor per 10.52 moles of combustion products, compared with 2 moles of water per 3
moles of products for combustion in oxygen. Similarly, with air, there is a mass fraction
of CO2 of 0.1514 and a carbon mass fraction of 0.0413 in the combustion products,
compared with 0.55 and 0.15, respectively, for combustion in oxygen.
The diluting energetic effect of nitrogen when combustion is in air may be reasoned
as follows: The same amount of energy is released in both reactions, because the same
amount of fuel is completely consumed. However, the nonreacting nitrogen molecules
in the air have heat capacity. This added heat capacity of the additional nitrogen
molecules absorbs much of the energy released, resulting in a lower internal energy per
unit mass of products and hence a lower temperature of the products. Thus the energy
released by the reaction is shared by a greater mass of combustion products when the
combustion is in air.
Often, products of combustion are released to the atmosphere through a chimney,
stack, or flue. These are therefore sometimes referred to as flue gases. The flue gas
composition may be stated in terms of wet flue gas (wfg) or dry flue gas (dfg), because
92

under some circumstances the water vapor in the gas condenses and then escapes as a
liquid rather than remaining as a gaseous component of the flue gas. When liquid water
is present in combustion products, the combustion product gaseous mass fractions may
be taken with respect to the mass of flue gas products, with the product water present
or omitted. Thus, for Equation (3.2), the mass of dry combustion products is 254.56.
Hence the mass fraction of carbon dioxide is 44/254.56 = 0.1728 with respect to dry
flue gas, and 44/290.56 = 0.1514 with respect to wet flue gas.
In combustion discussions reference is frequently made to higher and lower heating
values. The term higher heating value, HHV, refers to a heating value measurement in
which the product water vapor is allowed to condense. As a consequence, the heat of
vaporization of the water is released and becomes part of the heating value. The lower
heating value, LHV, corresponds to a heating value in which the water remains a
vapor and does not yield its heat of vaporization. Thus the energy difference between
the two values is due to the heat of vaporization of water, and

HHV = LHV + (mwater /mfuel)hfg [Btu/lbm | kJ/kg]

where mwater is the mass of liquid water in the combustion products, and hfg is the latent
heat of vaporization of water.

Air-Fuel Ratio

It is important to know how much oxygen or air must be supplied for complete
combustion of a given quantity of fuel. This information is required in sizing fans and
ducts that supply oxidizer to combustion chambers or burners and for numerous other
design purposes. The mass air-fuel ratio, A/F, or oxygen-fuel ratio, O/F, for complete
combustion may be determined by calculating the masses of oxidizer and fuel from the
appropriate reaction equation. Let’s return to Equation (3.2):

CH4 + 2O2 + 2(3.76)N2 Y CO2 + 2H2O + 2(3.76)N2 (3.2)

The A/F for methane is [(2)(32) + (2)(3.76)(28)]/(12 + 4) = 17.16 and the O/F is
2(32)/(12 + 4) = 4. Thus 4 kg of O2 or 17.16 kg of air must be supplied for each
kilogram of methane completely consumed.
Of course it is possible, within limits, to supply an arbitrary amount of air to a
burner to burn the fuel. The terms stoichiometric or theoretical are applied to the
situation just described, in which just enough oxidizer is supplied to completely convert
the fuel to CO2 and H2O. Thus the stoichiometric O/F and A/F ratios for methane are
4.0 and 17.16, respectively. If less than the theoretical amount of air is supplied, the
products will contain unburned fuel. Regardless of the magnitude of A/F, when
unburned fuel remains in the products (including carbon, carbon monoxide, or
hydrogen), combustion is said to be incomplete. Because air is virtually free and fuel is
expensive, it is usually important to burn all of the fuel by using more air than the
theoretical air-fuel ratio indicates is needed. Thus most burners operate with excess air.
93

The actual air-fuel ratio used in a combustor is frequently stated as a percentage of


the theoretical air-fuel ratio

% theoretical air = 100(A/F)actual /(A/F)theor (3.3)

Thus, for methane, 120% of theoretical air implies an actual mass air-fuel ratio of
(120/100)(17.16) = 20.59.
Excess air is defined as the difference between the actual and the theoretical air
supplied. Accordingly, the percentage of excess air is

% excess air = 100[(A/F)actual – (A/F)theor ]/(A/F)theor (3.4)

Thus, for methane, 120% of theoretical air implies

% excess air = (100)(20.59 – 17.16)/17.16 = 20%.

Note also that combining Equations (3.4) and (3.3) yields the following general result:

% excess air = % theoretical air – 100% (3.5)

Again, the excess air percentage is 120% – 100% = 20%. Table 3.1 shows examples of
ranges of excess air used with certain fuels and combustion systems.
The air/fuel parameters just discussed emphasize the amount of air supplied to burn
a given amount of fuel relative to the theoretical requirement. An alternate approach
considers a given amount of air and indicates the mass of fuel supplied , the fuel-air
ratio, F/A, which is the inverse of the air-fuel ratio. A measure of how much fuel is
actually supplied, called the equivalence ratio, is the ratio of the actual fuel-air ratio to
the theoretical fuel-air ratio:

M = (F/A)actual / (F/A)theor = (A/F)theor / (A/F)actual

= 100/( % theoretical air)

Thus 100% theoretical air corresponds to an equivalence ratio of 1, and 20% excess air
to M = 100/120 = 0.833. When the equivalence ratio is less than 1, the mixture is called
lean; when greater than 1, it is called rich.
This section has dealt with the application of combustion chemistry or
stoichiometry applied to methane gas. Other fuels for which a reaction equation such as
Equation (3.1) or (3.2) is available may be treated in a similar way. Before considering
more complex combustion problems, it is appropriate to investigate the nature and
description of the various types of fossil fuels.
94

3.3 Fossil Fuel Characteristics

Most chemical fuels are found in nature in the form of crude oil, natural gas, and coal.
These fuels are called fossil fuels because they are believed to have been formed by the
decay of vegetable and animal matter over many thousands of years under conditions of
high pressure and temperature and with a deficiency or absence of oxygen. Other fuels
such as gasoline, syngas (synthetic gas), and coke may be derived from fossil fuels by
some form of industrial or chemical processing. These derived fuels are also called
fossil fuels.

Coal

Coal is an abundant solid fuel found in many locations around the world in a variety of
forms. The American Society for Testing Materials, ASTM, has established a ranking
system (ref. 3) that classifies coals as anthracite (I), bituminous (II), subbituminous
(III), and lignite (IV), according to their physical characteristics. Table 3.2 lists
95

seventeen of the many United States coals according to this class ranking.
Coal is formed over long periods of time, in a progression shown from left to right
in Figure 3.1. The bars on the ordinate show the division of the combustibles between
fixed carbon and volatile matter in the fuels. “Fixed carbon” and “volatile matter”
indicate roughly how much of the fuel burns as a solid and as a thermally generated gas,
respectively. It is seen that the volatile matter and oxygen contained in the fuels
decrease with increasing age.
Peat is a moist fuel, at the geologically young end of the scale, that has a relatively
low heating value. It is not considered a coal but, nevertheless, follows the patterns of
characteristics shown in the figure. Peat is regarded as an early stage or precursor of
coal. At the other extreme, anthracite is a geologically old, very hard, shiny coal with
high carbon content and high heating value. Bituminous is much more abundant than
anthracite, has a slightly lower carbon content, but also has a high heating value.
Subbituminous coal, lignite, and peat have successively poorer heating values and
higher volatile matter than bituminous.
Coal is a highly inhomogeneous material, of widely varying composition, found in
seams (layers) of varying thickness at varying depths below the earth's surface. The
wide geographic distribution of coal in the United States is shown in Figure 3.2.
96

According to reference 1, the average seam in the United States is about 5.5 ft. thick.
The largest known seam is 425 ft. thick and is found in Manchuria.

Coal Analyses

It is often difficult to obtain representative samples of coal because of composition


variations from location to location even within a given seam. As a result there are
limits on the accuracy and adequacy of coal analyses in assessing coal behavior in a
given application. Before discussing the nature of these analyses, it is important to
establish the basis on which they are conducted.
Coal contains varying amounts of loosely held moisture and noncombustible
materials or mineral matter (ash), which are of little or no use. The basis of an analysis
helps to specify the conditions under which the coal is tested. The coal sample may be
freshly taken from the mine, the as-mined basis. It may have resided in a coal pile for
months, and be analyzed just before burning, the as-fired basis. It may be examined
immediately after transport from the mine, the as-received basis. Exposure to rain or
dry periods, weathering, and separation and loss of noncombustible mineral matter
through abrasion and the shifting of loads during transport and storage may cause the
same load of coal to have changing mineral matter and moisture content over time. It is
therefore important to specify the basis for any test that is conducted. Published
tabulations of coal properties are frequently presented on a dry, ash-free, or dry and
ash-free basis, that is, in the absence of water and/or noncombustible mineral matter.
Coal ranking and analysis of combustion processes rely on two types of analysis of
coal composition: the proximate analysis and the ultimate analysis. The proximate
analysis starts with a representative sample of coal. The sample is first weighed, then
raised to a temperature high enough to drive off water, and then reweighed. The weight
97

loss divided by the initial weight gives the coal moisture content, M. The remaining
material is then heated at a much higher temperature, in the absence of oxygen, for a
time long enough to drive off gases. The resulting weight-loss fraction gives the
volatile matter content, VM, of the coal. The remainder of the sample is then burned in
air until only noncombustibles remain. The weight loss gives the fixed carbon, FC, and
the remaining material is identified as non-combustible mineral matter or ash, A.
The proximate analysis may be reported as percentages (or fractions) of the four
quantities moisture, ash, volatile matter, and fixed carbon, as in Table 3.2, or without
ash and moisture and with the FC and VM normalized to 100%. Sulfur, as a fraction of
the coal mass, is sometimes reported with the proximate analysis. The proximate
analysis, while providing very limited information, can be performed with limited
laboratory resources.
A more sophisticated and useful analysis is the ultimate analysis, a chemical
analysis that provides the elemental mass fractions of carbon, hydrogen, nitrogen,
oxygen, and sulfur, usually on a dry, ash-free basis. The ash content of the coal and
heating value are sometimes provided also.
Data from a dry, ash-free analysis can be converted to another basis by using the
basis adjustment factor, 1 - A - M, as follows. The mass of coal is the mass of ultimate
or proximate analysis components plus the masses of water (moisture) and ash:

m = mcomp + mash + mmoist [lbm | kg]


98

Dividing through by the total mass m and rearranging, we get the following as the ratio
of the mass of components to the total mass:

mcomp / m = 1 – A – M [dl]

where A is the ash fraction and M is the moisture fraction of the total coal mass. A
component of a coal analysis may be converted from the dry, ash-free basis to some
other basis by forming the product of the component fraction and the basis adjustment
factor. Thus an equation for the wet and ashy volatile matter fraction in the proximate
analysis may be determined from the dry, ash-free proximate analysis by using

VMas-fired = (Mass of combustibles/Total mass)VMdry,ashfree

= ( 1 - A - M ) VMdry,ash-free (3.6)

where A and M are, respectively, the ash and moisture fractions for the as-fired coal.
Here the as-fired (wet, ashy) mass fraction of volatile matter is the product of the dry,
ash-free mass fraction and the basis adjustment factor. Fixed carbon, heating values,
and components of the ultimate analysis may be dealt with in a similar way.
Table 3.3 gives proximate and ultimate analyses for a number of United States
coals on a dry basis. Another extensive tabulation of the characteristics of American
and world coals is given in Appendix E.

EXAMPLE 3.3

If the as-fired moisture fraction for Schuylkill, Pa. anthracite culm characterized in
Table 3.3 is 4.5%, determine the as-fired proximate and ultimate analysis and heating
value of the coal. (Culm is the fine coal refuse remaining from anthracite mining.)

Solution
The FC, VM, and ash contents are given in Table 3.3. Because ash is already present in
the analysis, the appropriate adjustment factor is 1 - A - M = 1 – 0.0 – 0.045 = 0.955.
Using Equation (3.6) and the data from Table 3.3, we get

VMas-fired = (0.955)(8.3) = 7.927


FCas-fired = (0.955)(32.6) = 31.133
Aas-fired = (0.955)(59.1) = 56.411
Mas-fired = 4.500
Check Sum = 99.971
Heating valueas-fired = (0.955)(4918) = 4697 Btu/lbm

Similarly, the as-fired ultimate analysis is 32% C, 1.15% H2, 4.87% O2, 0.57% N2,
0.48% S, 56.44% ash, and 4.5% moisture, with a checksum of 100.01.
_____________________________________________________________________
99
100
101
102

As a solid fuel, coal may be burned in a number of ways. Starting with the smallest
of installations, coal may be burned in a furnace, in chunk form on a stationary or
moving grate. Air is usually supplied from below with combustion gases passing
upward and ash falling through a stationary grate or dropping off the end of a moving
grate into an ash pit. A wide variety of solid fuels can be burned in this way.
Though all furnaces were onced fired manually, today many are fired by or with the
assistance of mechanical devices called stokers. Figure 3.3 shows a spreader stoker,
which scatters coal in a uniform pattern in the furnace, the finer particles burning in
suspension in the rising combustion air stream while the heavier particles drop to the
grate as they burn. The particles that reach the grate burn rapidly in a thin layer, and the
remaining ash drops off the end into the ash pit. This type of combustion system has
been in use for over fifty years for hot water heating and steam generation.
In large installations, coal is crushed to a particular size, and sometimes pulverized
to powder immediately before firing, to provide greater surface exposure to the
103

oxidizer and to ensure rapid removal of combustion gases. Because of the wide
variation in the characteristics of coals, specialized types of combustion systems
tailored to a specific coal or range of coal characteristics are used.

Natural Gas

Natural gas is a mixture of hydrocarbons and nitrogen, with other gases appearing in
small quantities. Table 3.4 shows the composition of samples of natural gases found in
several regions of the United States. For these samples, it is seen that the gases contain
83-94% methane (CH4), 0-16% ethane (C2H6), 0.5-8.4% nitrogen and small quantities
of other components, by volume. The ultimate analysis shows that the gases contain
about 65-75% carbon, 20-24% hydrogen, 0.75-13% nitrogen, and small amounts of
oxygen and sulfur in some cases. The higher heating values are in the neighborhood of
1000 Btu/ft3 on a volume basis and 22,000 Btu/lbm on a mass basis. In regions where it
is abundant, natural gas is frequently the fuel of choice because of its low sulfur and ash
content and ease of use.
104

EXAMPLE 3.4

Determine the molecular weight and stoichiometric mole and mass air-fuel ratios for the
Oklahoma gas mole composition given in Table 3.4.

Solution
Equation (3.2),

CH4 + 2O2 + 2(3.76)N2 Y CO2 + 2H2O + 2(3.76)N2 (3.2)

shows that there are 2 + 2(3.76) = 9.52 moles of air required for complete combustion
of each mole of methane. Similarly for ethane, the stoichiometric reaction equation is:

C2H6 + 3.5O2 + (3.5)(3.76)N2 Y 2CO2 + 3H2O + 13.16N2

where 2 carbon and 6 hydrogen atoms in ethane require 2 CO2 molecules and 3 H2O
molecules, respectively, in the products. There are then 7 oxygen atoms in the
products, which implies 3.5 oxygen molecules in the reactants. This in turn dictates the
presence of (3.5)(3.76) = 13.16 nitrogen molecules in both the reactants and products.
The reaction equation then indicates that 3.5(1 + 3.76) = 16.66 moles of air are
required for complete combustion of one mole of ethane.
In Table 3.5, the molecular weight of the gas mixture, 18.169, is found in the
fourth column by summing the products of the mole fractions of the fuel components
and the component molecular weights. This is analogous to the earlier determination of
the average air molecular weight from the nitrogen and oxygen mixture mole fractions.
The products of the mole fractions of fuel components and the moles of air
required per mole of fuel component (as determined earlier and tabulated in the fifth
column of Table 3.5) then yield the moles of air required for each combustible per mole
of fuel (in the sixth column). Summing these, the number of moles of air required per
mole of fuel yields the stoichiometric mole air-fuel ratio, 9.114.
The stoichiometric mass A/F is then given by the mole A/F times the ratio of air
molecular weight to fuel molecular weight: (9.114)(28.9)/18.169 = 14.5.

Table 3.5 Calculations for Example 3.4


i Mi xi xiMi Moles air per Moles air per mole fuel
mole i

Methane 16 0.841 13.456 9.52 (0.841)(9.52) = 7.998

Ethane 30 0.067 2.010 16.66 (0.067)(16.66) = 1.116

CO2 44 0.008 0.351 0.0

Nitrogen 28 0.084 2.352 0.0

Totals 1.000 18.169 moles air /mole fuel = 9.114


105

Liquid Fuels

Liquid fuels are primarily derived from crude oil through cracking and fractional
distillation. Cracking is a process by which long-chain hydrocarbons are broken up into
smaller molecules. Fractional distillation separates high-boiling-point hydrocarbons
from those with lower boiling points. Liquid fuels satisfy a wide range of combustion
requirements and are particularly attractive for transportation applications because of
their compactness and fluidity. Table 3.6 gives representative analyses of some of these
liquid fuels. Compositions of liquid and solid fuels, unlike gaseous fuels, are usually
stated as mass fractions.

3.4 Combustion Reactions and Analysis

Mechanism of Combustion

Details of the mechanics of combustion depend to a great extent on the fuel and the
nature of the combustion system. They are sometimes not well understood and are
largely beyond the scope of this book. There are, however, certain fundamentals that
are useful in dealing with combustion systems.
The chemical reaction equations presented here do not portray the actual
mechanism of combustion; they merely indicate the initial and final chemical
compositions of a reaction. In most cases the reactions involve a sequence of steps,
leading from the reactants to the products, the nature of which depends on the
temperature, pressure, and other conditions of combustion. Fuel molecules, for
instance, may undergo thermal cracking, producing more numerous and smaller fuel
molecules and perhaps breaking the molecules down completely into carbon and
hydrogen atoms before oxidation is completed.
In the case of solid fuels, combustion may be governed by the rate at which
oxidizer diffuses from the surrounding gases to the surface and by the release of
combustible gases near the surface. Combustion of solids may be enhanced by
increasing the fuel surface area exposed to the oxidizer by reducing fuel particle size.
106

The following simple model illustrates the effect.

Example 3.5 is, of course, an idealized example. In reality, the reacting surface area of
solid fuels is usually much larger than the spherical surface area implied by their size.
We have seen that, for combustion to occur, molecules of oxidizer must affiliate
with fuel molecules, an action enhanced by the three T’s of combustion: turbulence,
time, and temperature. Chemical reactions take place more rapidly at high temperatures
but nevertheless require finite time for completion. It is therefore important that burners
be long enough to retain the fuel-air mixture for a sufficiently long time so that
combustion is completed before the mixture leaves. Turbulence, or mixing, enhances
the opportunities for contact of oxidizer and fuel molecules and removal of products of
combustion.
A flame propagates at a given speed through a flammable mixture. It will
propagate upstream in a flow of a combustible mixture if its flame speed exceeds the
flow velocity. If a fixed flame front is to exist at a fixed location in a duct flow in which
the velocity of the combustion gas stream exceeds the propagation speed, some form of
flame stabilization is required. Otherwise the flame front is swept downstream and
flameout occurs. Stabilization may be achieved by using fixed flameholders (partial
107

flow obstructions that create local regions of separated flow in their bases where the
flame speed is greater than the local flow velocity) or by directing a portion of the flow
upstream to provide a low-speed region where stable combustion may occur.
Each combination of oxidizer and fuel has been seen to have a particular
stoichiometric oxidizer-fuel ratio for which the fuel is completely burned with a
minimum of oxidizer. It has also been pointed out that it is usually desirable to operate
burners at greater than the theoretical air-fuel ratio to assure complete combustion of
the fuel and that this is sometimes referred to as a lean mixture. Occasionally it may be
desirable to have incomplete combustion, perhaps to produce a stream of products in
which carbon monoxide exists or to assure that all the oxidizer in the mixture is
consumed. In that case a burner is operated at less than the stoichiometric air-fuel ratio
with what is called a rich mixture.
There are limits to the range of air-fuel ratios for which combustion will occur
called limits of flammability. Here the density of the mixture is important. The limits of
flammability around the stoichiometric A/F are reduced at low densities. If combustion
is to occur reliably in mixtures at low densities, it is necessary to closely control the
air-fuel ratio.

Combustion Analysis of Solid Fuels

In the determination of the air-fuel ratio and flue gas composition for the combustion of
solid fuels, it is important to account for the ash and moisture in the fuel in the as-fired
condition. In the following analyses, all of the elements of the reactants in the fuel and
oxidizer are assumed to be present in the flue gas products except for the ash, which is
assumed to fall as a solid or flow as molten slag to the furnace bottom. Nitrogen and
oxygen are present in many solid fuels and should be accounted for in predicting the
flue gas composition. While both carbon monoxide and oxygen may be present in
combustion products at the same time because of imperfect mixing of combustibles and
oxygen in some instances, we will assume for prediction of the flue gas composition
that perfect mixing occurs such that no carbon monoxide is present when excess air is
supplied.

EXAMPLE 3.6

A coal with a dry, ash-free composition of 0.87 C, 0.09 H2, 0.02 S, and 0.02 O2 is
burned with 25% excess air. The as-fired ash and moisture contents are 6% and 4%,
respectively.
(a) What are the stoichiometric and actual air-fuel ratios?
(b) What is the flue gas composition?

Solution
(a) Before performing combustion calculations, it is necessary to convert coal
composition data to an as-fired basis. The ratio of as-fired to dry, ash-free
108
109

mfj = (kg j / kg fg) = (kg j / kg coal) / (kg fg / kg coal)


110
111
112
113
114
115

(Note: The reference conditions of the 1985 JANAF Thermochemical tables used differ
slightly from those of preceding editions.) Heats of formation are usually determined
based on statistical thermodynamics and spectroscopic measurements.
By definition, heats of formation are zero for all elements in the standard state.
Hence, from the steady-flow First Law of Thermodynamics, the heat transferred in a
formation reaction of a compound created from elements in the standard state is the
heat of formation for the compound, as in the hydrogen-to-water example just
mentioned.

Heat Transfer in a Chemically Reacting Flow

Consider now the combustion problem in which fuel and oxidizer flow into a control
volume and combustion products flow out. The steady-flow First Law of Thermo-
dynamics applied to the control volume may be written as

Q = Hp – Hr + Ws [Btu | kJ] (3.7)

where Q is heat flow into the control volume, Ws is the shaft work delivered by the
control volume, and the enthalpies, H, include chemical as well as thermal energy. The
subscripts r and p refer to the reactants entering and products leaving the control
volume, respectively. The enthalpy Hp is the sum of the enthalpies of all product
streams leaving the control volume. A similar statement applies to Hr for the entering
reactant streams.
116

The individual enthalpies may each be written as the product of the number of
moles of the component in the reaction equation and its respective enthalpy per mole of
the component. For example, for k products:

Hp = n1h1 + n2h2 +...+ nkhk [Btu | kJ] (3.8)

where the n’s are the stoichiometric coefficients of the chemical equation for the
combustion reaction, and the enthalpies are on a per-mole bases.

EXAMPLE 3.9

Write an equation for the enthalpy of the products of complete combustion of octane in
oxygen.

Solution The balanced equation for the complete combustion of octane is

C8H18 + 12.5O2 Y 8CO2 + 9H2O

The mole coefficients, 8 and 9, of the products are stoichiometric coefficients that yield

Hp = 8h(CO2) + 9h(H2O) [Btu | kJ]

per mole of octane consumed.


___________________________________________________________________

The enthalpy of any component of the reactants or products may be written as the
sum of (1) its enthalpy of formation at the standard temperature, To, and standard
pressure, and (2) its enthalpy difference between the actual state and the standard state
of the components. Thus, for each component:

h(T) = hf (To) + [h(T) – h(To)] [Btu /mole | kJ /mole] (3.9)

where it is assumed that the sensible gas enthalpy difference is independent of pressure.
Sensible enthalpies (those that depend on temperature but do not involve reactions or
phase change) relative to the standard reference state are given in Appendix D.
Thus, returning to the formation reaction for the combustion of hydrogen in
oxygen at the standard state to produce water, as discussed in the preceding section,
we see that the steady-flow First Law of Thermodynamics becomes

Q = (1)hf, H2O – (1)hf, H2 – (0.5)hf, O2 = –103,996 – 0 – 0 = – 103,996 Btu/lb-mole

with water in the vapor phase as the product of combustion of one mole of H2. Here
the sensible enthalpy differences are zero, because both the products and the reactants
are at the standard state. Note that because the stoichiometric coefficients of both
117

hydrogen and water are the same in this reaction, the resulting heat transfer may be
interpreted as either per mole of water formed or per mole of hydrogen consumed.
If, instead, liquid water is the combustion product, the heat transfer is given by

Q = – 122,885 – 0 – 0 = – 122,885 Btu/lb-mole

of H2O. The difference between the two cases, 18,919 Btu/lb-mole H2O, is equivalent
to 18,935/18 = 1,051.9 Btu/lbm of water, the enthalpy of vaporization of water. This
result compares with the enthalpy or latent heat of vaporization of water at 77°F,
1050.1 Btu/lbm, given in the steam tables.
With either liquid or gaseous water as the product, the heat transfer term for the
control volume is negative, indicating, in accordance with the usual sign convention,
that the heat flows from the control volume to the surroundings. The two calculations
above illustrate the fact that the heat transfer in a formation reaction at the standard
state is the heat of formation of the compound created.

EXAMPLE 3.10

What is the enthalpy of water vapor at 1800°R and 1 bar? What is the heat transfer in
the formation reaction of water from hydrogen and oxygen if the products are at
1800°R and the reactants are (a) at the standard state, and (b) at 900°R?

Solution
The heat of formation of water vapor at the standard state of 298.15K. (536.7°R)
and one bar is – 103,966 Btu/lb-mole. The enthalpy of water vapor at 1800°R is the
sum of the heat of formation at the standard state and the sensible enthalpy difference
of water vapor between 536.7°R and 1800°R. Thus:

Hp = – 103,966 + 11,185 = – 92,781 Btu/lb-mole of water.

(a) In this case, the reactants, oxygen and hydrogen, have zero enthalpies because they
are in the standard state and, as elements, their heats of formation are zero. Thus the
heat transferred is – 92,781 Btu/lb-mole, or 5154.5 Btu/lbm of water.

(b) For reactants at 900°R , Appendix D gives hH2(900) – hH2(536.7) = 2530 Btu/mole
of H2 and hO2(900) – hO2(536.7) = 2617 Btu/mole of O2. The enthalpy of the reactants
is then Hr = (1.0)(2530) + (0.5)(2617) = 3838.5 Btu/lb-mole H2O. The heat transferred
is then:

Q = Hp – Hr = – 92,781 – 3,838.5 = – 96,619.5 Btu/lb-mole of water,


or
Q = – 96,619.5 / 18 = – 5,367.8 Btu/lbm of water.
118

Thus more heat must be transferred from the control volume to form water vapor at
1800°R if the reactants are at 900°R than if they are in the 536.7°R standard state.

Combustion Flame Temperature

In many combustion problems, the reactants enter the combustor near room tempera-
ture and products emerge at an elevated temperature. The temperature of the products
flowing from the control volume depends on the energy released in the combustion
reaction and heat gain or loss through the boundary of the control volume. The
resulting combustion product temperature is sometimes called the flame temperature.

EXAMPLE 3.11

Methane and stoichiometric air enter a combustor at the standard state. Using a spread-
sheet, calculate the heat transfer as a function of the exit (flame) temperature of the
products of complete combustion. Assume the water in the products remains a vapor.

Solution
The reaction equation for complete combustion of methane in air is:

CH4 + 2O2 + (2)3.76N2 Y CO2 + 2H2O + 7.52N2

The enthalpy of the products at temperature T and of the reactants at the standard state
is

Hp = (1)hf, CO2 + (1)[hCO2 (T) – hCO2 (537)] + (2)hf, H2O


+ (2)[hH2O (T) – hH2O (537)] + (7.52)[hN2 (T) – hN2 (537)]

Hr = hf, CH4 = – 32,189.6 Btu/lb-mole

of methane, where the heats of formation of elemental nitrogen and oxygen are zero
and the heat of formation of water is for the vapor phase. Writing the enthalpy
differences as )h’s and applying the steady-flow First Law of Thermodynamics, we get

Q = Hp – Hr = – 169,184 + )hCO2(T) + (2)(– 103,966)


+ (2) )hH2O(T) + 7.52 )hN2(T) – (– 32,189.6)

= – 344,926.4 + )hCO2(T) + (2))hH2O(T) + 7.52 )hN2 Btu/lb-mole

of methane. This function is tabulated in the spreadsheet in Table 3.10 and plotted in
Figure 3.4 using values of enthalpies at the temperature T from the the JANAF tables.
Negative values of Q indicate that heat must be rejected from the control volume to
maintain product effluent temperature below about 4200°R. Beyond 4200°R, the CO2,
119

N2, and H2O outflow carries more energy than is released in the control volume by
chemical reaction; hence, heat must flow into the control volume to achieve the
resulting high exit temperatures. Thus the final flame temperature clearly depends on
the chemical composition of the flow and on the consequent control volume heat
transfer.
____________________________________________________________________
120

Heat of Combustion and Heating Value

The heat of combustion, or enthalpy of combustion, of a fuel is defined as the energy


transferred during a steady-flow process in which the fuel is completely burned and
where the products are returned to the temperature and pressure of the reactants. It will
be seen that the enthalpy of combustion evaluated at the standard state may be deter-
mined from the heats of formation. The heat of combustion of hydrogen has, in fact,
been determined in a preceding section that examined the formation reaction for water.
The negative of the enthalpy of combustion of a fuel burned in air is usually
referred to as the heating value of the fuel. When water in the combustion products is
condensed, the heat of vaporization of the water adds to the chemical energy released,
and the resulting heating value is called the higher heating value, HHV. Recall also
that the heating value obtained when the product water stays a vapor is called the lower
heating value, LHV. The difference between HHV and LHV has been illustrated in the
previous section for the formation reaction of water resulting from the combustion of
hydrogen.
For methane, note also that the heat of combustion, and thus the magnitude of the
lower heating value, appears in the value of Q in the top row of table 3.10, since there
the combustion products are at the reference temperature.

EXAMPLE 3.12

Illinois no. 6 raw coal has the following dry mass composition: 61.6% C, 4.2% H2,
9.7% O2, 1.3% N2, 4.6% S, and 18.5% ash. Using heats of formation, determine the
higher and lower heating values, in kJ / kg, of the as-fired coal with 10% moisture, and
compare them with the heating value in Table 3.3.

Solution
To adjust the composition for 10% moisture, the factor 1 – A – M becomes 1 – 0
– 0.1 = 0.9. The resulting moist coal composition is given in the following table.
It was seen earlier that the heat of reaction of hydrogen in its standard state, and
thus its heat of combustion, is the heat of formation of its product of combustion. The
reaction equation

H2 + 0.5O2 Y H2O

shows that one mole of hydrogen produces one mole of water. Thus, from Table 3.9,
the heat of formation of steam, – 241,826 kJ per kg-mole of water formed or per kg-
mole of hydrogen burned, is also the heat of combustion of hydrogen in the standard
state. Thus the hydrogen contributes 241,826/2 kJ per kg of hydrogen in the coal. The
total energy released by the hydrogen in the coal is then 241,826/2 times the mass
fraction of hydrogen in the coal, as shown in the following table. Similar arguments
may be made for hydrogen with product water in the liquid phase and the carbon and
sulfur components of the coal.
121

Element i Dry mfi Wet mfi Heat of Combustion kJ/kg coal

H2:

For LHV 0.042 0.0378 (241,826)(0.0378)/2 = 4,570.5 (v)

For HHV (285,830)(0.0378)/2 5402.2 (l)

C 0.616 0.5544 (393,522)(0.5544)/12 = 18,180.7

O2 0.097 0.0873

N2 0.013 0.0117

S 0.046 0.0414 (296,842)(0.0414)/32 = 384.0

Ash 0.185 0.1665

H2O _____ 0.1000

0.999 0.9991 23,135.2 (v)

23,966.9 (l)

Thus the lower and higher heating values of the coal are 23,135.2 and 23,966.9 kJ/kg,
respectively. Table 3.3 lists a heating value of dry Illinois no. 6 raw coal of 11,345
Btu/lbm. The corresponding heating value for the wet coal is (0.9)(11,345) = 10,210.5
Btu/lbm. This corresponds to (10,210.5)/(0.43) = 23,745 kJ/kg.
____________________________________________________________________

Adiabatic Flame Temperature

The results of Example 3.11, tabulated in Figure 3.10, show that for a given air-fuel
mixture there is a unique product temperature for which the control volume is
adiabatic. This temperature is known as the adiabatic flame temperature. It can be
determined as in Example 3.11, or it may be calculated from the steady-flow First Law
of Thermodynamics by setting Q = 0. The resulting First Law equation for the adiabatic
flame temperature, designated T*, becomes:

Hp (T*) = Hr (T r) [Btu | kJ] (3.10)

where the reactants are at the temperature Tr . The enthalpy terms depend on the
individual enthalpies of the components as functions of temperature. Thus a trial-and-
error solution is required using data on heat of formation from Table 3.9 and the
enthalpy tables in Appendix D. Given a known Tr , the adiabatic flame temperature may
also be obtained as the intercept (Q = 0) on a graph of Q versus temperature, T, such as
Figure 3.4.
Adiabatic Flame Temperature for Solid Fuels
122

As a final example, we will determine the combustion products, heat of combustion,


and adiabatic flame temperature for a solid fuel specified by its ultimate analysis. The
solution is presented in a spreadsheet in which enthalpies are tabulated as a function of
temperature for the relevant chemical components as given in the JANAF tables.
123
124

The number of moles of excess O2 in the flue gas per pound of coal may also be
obtained from the excess air-fuel ratio:

[(A/F)actual – (A/F)theor ](0.233)/32 = (8.288 – 6.375)(0.233)/32 = 0.0139

The number of moles of flue gas nitrogen is also given by

[(A/F)actual ](1 – 0.233)/28 + d = 8.288(0.767)/28 + 0.000357 = 0.227

With the actual balanced reaction equation known, the mole coefficients may then be
used to write an equation for the enthalpy of the products per unit mass of coal.
Because all the reactants are assumed to be elements at the reference temperature, the
enthalpy of the reactants is zero. By setting Q = Hp = 0, we can solve this First Law
equation for the adiabatic flame temperature by trial and error. However, with a
spreadsheet, it is convenient to calculate the heat transfer at each temperature-enthalpy
data point and determine the adiabatic flame temperature by inspection and, if
additional precision is required, by explicit interpolation. The heat transfer equation is
shown on the spreadsheet, and the values of the flue gas mole coefficients and heats of
formation are shown above the appropriate JANAF enthalpies to which they relate.

It is seen that the adiabatic flame temperature for combustion with 30% excess air
is about 2110K, or 3800°R. Note also that the heat of combustion of the coal, 8506.7
Btu/lbm of coal, may be read from the spreadsheet at the JANAF table reference
temperature. This is possible because the heat of reaction is independent of the amount
of excess air employed in the reaction.
_____________________________________________________________________

The spreadsheet for Example 3.13 is easily modified to compute and plot the
composition of the flue gas as a function of the percentage of excess air. Figure 3.5
shows that the mole fractions of excess oxygen and of nitrogen increase while the
fractions of other products decrease. Excess oxygen measured from a flue gas sample
is commonly used as a measure of excess air in adjusting the air-fuel ratio of
combustion systems.

3.6 Molecular Vibration, Dissociation, and Ionization

The temperature of a gas is a measure of the random translational kinetic energy of


molecules. The simple kinetic theory of the heat capacity of a gas predicts heat
capacities that are independent of temperature and determined by the number of
degrees of freedom of the molecules. The kinetic theory is usually regarded as
applicable at low pressures and moderate and high temperatures, conditions at which
collisions between molecules are rare.
125

For monatomic gases, kinetic theory predicts an internal energy per atom of

u = 3kT/2

and an enthalpy of

h = u + pv = 3kT/2 + kT = 5kT/2

Here k, the Boltzmann constant, is the ideal-gas constant per molecule, which can be
calculated from the universal gas constant, ú, and Avogadro’s number of molecules
per mole. Thus

k = ú/No = 8.32/(6.025 × 1023) = 1.38 × 10 -23 J/K

From their definitions, Equations (1.14) and (1.15), this leads to cv = 3k/2 and cp = 5k/2
per atom and to a heat capacity ratio of k = cp /cv = 5/3 = 1.667.
The concept of equipartition of energy assumes that the energy of a particle is
equally divided among its various degrees of freedom. Each mode of energy storage of
a molecule is assigned an energy kT/2. The theory then represents the internal energy
as nkT/2, the enthalpy as (n + 2)kT/2, and the heat capacity ratio as k = (n + 2)/n,
where n is the number of modes of energy storage, or degrees of freedom, of the
molecule. For atoms with three translational and no rotational degrees of freedom,
n = 3 (an atom is presumed to be a point and to have no rotational kinetic energy
126

because it has zero radius); and these relations reduce to the findings described in the
preceding paragraph.
The simple kinetic theory suggests that a diatomic molecule has five degrees of
freedom: three degrees of translational freedom and two modes having significant
rotational kinetic energy. For this case we obtain u = 5kT/2, h = 7kT/2, and
k = 7/5 = 1.4, in agreement with experiment for oxygen and nitrogen at moderate
temperatures and densities.
At higher temperatures, diatomic molecules start to vibrate, adding additional
degrees of freedom that reduce the heat capacity ratio k below 1.4. As the temperature
increases and the collisions between molecules become more vigorous, molecules not
only vibrate but they start to be torn apart, each forming separate atoms, the process
known as dissociation. The energy required to break the bonds between atoms in
molecules is called the dissociation energy. At low temperatures, few molecules have
sufficient kinetic energy to provide the energy needed to cause dissociation by collision.
At higher temperatures, when more molecules have energies exceeding the dissociation
energy, the chemical equilibrium shifts to a composition in which there are more atoms
and fewer molecules in the gas. This trend continues as temperature increases.
At still higher temperatures, when particle kinetic energies exceed the ionization
energy of the gas, outer electrons are separated from atoms, forming positively charged
ions, in a process known as the ionization. Particles lose kinetic energy when causing
dissociation and ionization. This lost energy is then not reflected in the temperature.
Thus, when vibration, dissociation, and ionization occur, the internal energy and
enthalpy increase more rapidly than the temperature. Then the simple linear relations
just given for u and T are no longer correct. Stated another way, the temperature of a
gas rises less rapidly with heat addition when it is at temperature levels where
significant vibration, dissociation, or ionization take place. The phenomenon may be
thought of as analogous to phase change, in which enthalpy increases with heat addition
while temperature does not.
As a result of these phenomena, high flame temperatures determined with ideal gas
enthalpies may be overestimated. At temperatures exceeding 2000K (3600°R), flame
temperature calculations based on the JANAF gas enthalpies may start to become
inaccurate. At these temperature levels, dissociation (and at still higher temperatures,
ionization) starts to influence the composition of the gases and hence their
thermodynamic properties. The progression of dissociation and ionization with
temperature is shown for air at sea level density in Figure 3.6. It is seen that, at this
density, little dissociation of nitrogen occurs below about 6000 K. but that oxygen
starts to dissociate significantly above 3000 K.. At lower densities, the onset of
dissociation occurs at progressively lower temperatures. In Figure 3.6, the number of
particles per initial atom of air may be obtained by multiplying the ordinates by 1.993.
Usually, dissociation does not seriously influence combustion calculations when the
oxidizer is air, but the high combustion temperatures resulting from use of pure oxygen
may be significantly influenced by dissociation. The reader is referred to advanced
thermodynamics, physical chemistry, and advanced engineering texts for methods of
predicting the effects of dissociation and ionization.
127

Bibliography and References

1. Anon, Steam, Its Generation and Use, 39th ed. New York,: Babcock and Wilcox,
1978.

2. Singer, Joseph G. (Ed.), Combustion / Fossil Power Systems. Windsor, Conn.:


Combustion Engineering Inc., 1981.

3. Anon., Classification of Coals by Rank, Standard D-388. Philadelphia: American


Society for Testing Materials, Section 5, Volume 05.05, 1983.

4. Chase, M. W. Jr., et.al., JANAF Thermochemical Tables, 3rd ed., J. Phys. Chem.
Ref. Data 14, Supplement No. 1, 1985.

5. Van Wylen, Gordon J., and Sonntag, Richard E., Fundamentals of Classical
Thermodynamics. New York: Wiley, 1986.
128

6. El-Wakil, M. M., Powerplant Technology. New York: McGraw-Hill, 1984.

7. Campbell, Ashley S., Thermodynamic Analysis of Combustion Engines. New York:


Wiley, 1979.

8. Culp, Archie W., Principles of Energy Conversion. New York: McGraw-Hill, 1979.

9. Wood, Bernard D., Applications of Thermodynamics. Reading, Mass.: Addison-


Wesley, 1982.

10. Lefebvre, Arthur H., Gas Turbine Combustion, New York: McGraw-Hill, 1983.

11. Baumeister, Theodore, and Marks, Lionel S. (Eds.), Standard Handbook for
Mechanical Engineers, 7th ed. New York: McGraw-Hill, 1967.

12. Sorenson, Harry A., Energy Conversion Systems. New York: Wiley, 1983.

13. Hill, Phillip G., Power Generation. Cambridge, Mass.: MIT Press, 1977.

14. Anon., Coal Quality Information Book, CS-5421, Interim Report. Palo Alto:
Electric Power Research Institute, December 1987.

15. Moeckel, W. E., and Weston, Kenneth C., “Composition and Thermodynamic
Properties of Air in Chemical Equilibrium,” NACA TN 4265, April 1958.

EXERCISES

3.1 Determine the mass fractions of a mixture of six grams of carbon, three grams of
sulfur, and one gram of sodium chloride.

3.2 Determine the mole fractions of a gas consisting of a mole of oxygen, eight moles
of nitrogen, a mole of CO, and two moles of CO2. Determine also the mass fractions.
What is the average molecular weight of the gas?

3.3 Write the balanced reaction equation for the complete combustion of sulfur in
oxygen. What are the mass and mole fractions of oxygen in the reactants?

3.4 Write the balanced reaction equation for the complete combustion of carbon in
oxygen. What are the mass and mole fractions of oxygen in the reactants?

3.5 Write the balanced reaction equation for the complete combustion of carbon in air.
What are the mass and mole fractions of fuel, air, oxygen, and nitrogen in the reactants?
129

3.6 Write the balanced reaction equation for the complete combustion of ethane, C2H6,
in air. What are the mass and mole fractions of fuel, air, oxygen, and nitrogen in the
reactants? What are the mass and mole fractions of carbon dioxide and water vapor in
the combustion products?

3.7 Write the balanced reaction equation for the complete combustion of propane,
C3H8,, in air. What are the mass and mole fractions of fuel, air, oxygen, and nitrogen in
the reactants? What are the mass and mole fractions of carbon dioxide and water vapor
in the combustion products. What are the mass and mole air-fuel ratios?

3.8 Write the balanced reaction equation for the complete combustion of C8H18 in air.
What are the mass and mole fractions of fuel, air, oxygen, and nitrogen in the reactants?
What are the mass and mole fractions of carbon dioxide and water vapor in the
combustion products? What are the mass and mole air-fuel ratios?

3.9 Gasoline, sometimes represented as C8H18, is burned in 25% excess air mass. What
are the mass and mole stoichiometric and actual air-fuel ratios? Determine the mass
and mole fractions of the combustion products.

3.10 Determine the lower and higher heating values of methane using the JANAF table
of heats the formation.

3.11 Determine the as-fired stoichiometric and actual air-fuel ratios for Greene,
Pennsylvania raw coal (Table 3.3) with 5% moisture and the mass and mole flue gas
compositions for combustion with 20% excess air.

3.12 Compare the stoichiometric and actual air-fuel ratios and the mole flue gas
composition for combustion with 20% excess air for the following raw and clean
(process #1) coals (Table 3.3): (a) Freestone, Texas, big brown lignite; (b) Indiana,
Pennsylvania, Freeport (upper); (c) British Columbia, Hat Creek (A zone); (d) Perry,
Illinois no. 6; (e) Muhlenberg, Kentucky no. 9; (f) Nicholas, West Virginia, Kittanning;
(g) Belmont, Ohio, Pittsburgh; (h) Big Horn, Montana, Robinson; (i) Greene,
Pennsylvania, Sewickley; (j) Kanawha, West Virginia, Stockton-Lewiston; (k) Belmont,
Ohio, Waynesburg.

3.13 If Union, Kentucky no. 11 raw coal has 10% moisture, as mined, determine the as-
mined proximate and ultimate analyses for this coal.

3.14 If Big Horn, Montana, Robinson raw coal has 15% moisture, as-mined, what are
its as-mined proximate and ultimate analyses?

3.15 Determine the ultimate analyses of the raw coals listed in Exercise 3.12 (a-e)
assuming 10% as-mined moisture.
130

3.16 Write the balanced chemical equation for the combustion of methane in
stoichiometric air. Use the table of heats of formation to determine the heats of
reaction of methane (in Btu/lbm and Btu/lb-mole), with products and reactants all at the
standard state and product water as liquid. What are the values if the product water is
vapor? Would the heat of reaction be different if the combustion were in pure oxygen?
Compare your results with tabulated heating values for methane.

3.17 Solve Exercise 3.16 in SI units: kJ/gm-mole and kJ/kg.

3.18 Calculate the heating values, in Btu/lbm and Btu/lb-mole, for the complete
combustion of hydrogen, with product water in liquid and in vapor phases. Compare
with tabulated heating values. Repeat the calculations in SI units.

3.19 Determine the heat transferred when ethane is burned (a) in stoichiometric air, and
(b) in 100% excess air. In both cases the reactants are in the standard state and
products at 1000K. Use a heat of formation of –36,420 Btu/lb-mole.

3.20 What is the adiabatic flame temperature for the combustion of ethane in air,
ignoring dissociation? Use a heat of formation of –36,420 Btu/lb-mole of ethane.

3.21 Compare the adiabatic flame temperatures for the stoichiometric combustion of
hydrogen in air and in pure oxygen, ignoring dissociation.

3.22 Determine the adiabatic flame temperature for the stoichiometric combustion in
air of Illinois no. 6 coal after the clean #2 process. Determine also its heat of
combustion, and compare with the tabulated value.

3.23* Develop a spreadsheet that determines the air-fuel ratio for coal characterized by
a dry ultimate analysis, such as given in Table 3.3. Apply it to several of the coals in
the table, as assigned by your instructor.

3.24* Develop a spreadsheet that determines the air-fuel ratio for a coal characterized
by a dry ultimate analysis, such as given in Table 3.3, and a given moisture content.
Apply it to a coal in the table and several different moisture contents, as assigned by
your instructor.

3.25* Develop a spreadsheet that determines the mass and mole, wet and dry flue gas
compositions for a coal characterized by a dry ultimate analysis, such as given in Table
3.3, and a given percentage of excess air. Apply it to several coals in the table for 20%
and 40% excess air.

______________________
*Exercise numbers with an asterisk involve computer usage.
131

3.26* Develop a spreadsheet that determines the average molecular weight, and the
mass and mole, wet and dry flue gas compositions for a coal characterized by a dry
ultimate analysis, such as given in Table 3.3, a given percentage of excess air, and a
given moisture content. Apply it to a coal in the table for 10% moisture and 10% and
20% excess air.

3.27* Develop a spreadsheet that determines the theoretical air-fuel ratio for a gas
characterized by any combination of the components of Table 3.4. Apply it to the
gases in the table, as assigned by your instructor.

3.28* Develop a spreadsheet that determines the average molecular weight, and the
mass and mole, wet and dry flue gas compositions for a gas characterized by any
combination of the components of Table 3.4 and a given percentage of excess air.
Apply it to several the gases in the table for 20% excess air.

3.29 What are the theoretical and actual air-fuel ratios and the wet and dry mole flue
gas compositions for 20% excess air for the Oklahoma natural gas in Table 3.4?

3.30 What are the theoretical and actual air-fuel ratios and the wet and dry mole flue
gas compositions for 15% excess air for the Ohio natural gas in Table 3.4?

3.31 An adiabatic gas turbine combustor burns methane at 77°F with air at 400°F. The
combustion products emerge from the combustion chamber at 3200°F. What is the air-
fuel ratio? What is the equivalent external heat transferred per lbm of air to produce this
temperature rise, assuming a mean heat capacity of 0.24 Btu/lbm-R?

3.32 An adiabatic gas turbine combustor burns methane at 25°C with air at 250°C. The
combustion products emerge from the combustion chamber at 2000K. What is the air-
fuel ratio? What is the equivalent external heat transferred per kilogram of air to
produce this temperature rise, assuming a mean heat capacity of 1.005 kJ/kg-K?

3.33 An adiabatic combustor burns methane with 400% excess air. Both air and
methane are initially at 298K. What is the exit temperature?

3.34 An adiabatic combustor burns methane with 500% theoretical air. Both air and
methane are initially at 77°F. What is the flame temperature?

3.35 An adiabatic combustor burns methane in 100% excess oxygen. Both fuel and
oxidizer enter at the JANAF tables reference temperature. What is the flame
temperature?

3.36 Calculate and tabulate the higher and lower heating values of methane, in kJ/kg-
mole and in kJ/kg.
132

3.37 Determine the theoretical air-fuel ratio and the air-fuel ratio for 20% excess air
for Muhlenberg, Kentucky, raw coal with 10% moisture.

3.38 Determine the air-fuel ratio and wet and dry flue gas mole and mass fractions for
dry Muhlenberg, Kentucky no. 9 raw coal with 20% excess air.

3.39 Solve Exercise 3.38 for the coal having 10% moisture in the as-fired condition.

3.40 Solve Exercise 3.38 for 40% excess air.

3.41* Set up a spreadsheet to solve Exercises 3.38–3.40 where it is necessary to


change only one parameter for each of the latter cases.

3.42 Using the compound composition data of Table 3.4, calculate the ultimate
(elemental) analysis, and compare with the tabular results for the following natural
gases: (a) Pennsylvania, (b) Southern California, (c) Ohio, (d) Louisiana, (e) Oklahoma.

3.43* Develop an interactive computer program to solicit and receive ultimate analysis
data for an arbitrary coal, arbitrary as-fired ash and moisture fractions, an excess air
percentage, and output the appropriate air-fuel ratio and mass and mole, wet and dry
flue gas compositions.

3.44* Develop an interactive computer program to solicit and receive ultimate analysis
data for an arbitrary coal, arbitrary as-fired ash and moisture fractions, and Orsat CO,
CO2, and O2 data. The program should determine the actual operating air-fuel ratio.

3.45* Apply the spreadsheet of Example 3.13 to determine the adiabatic flame
temperature, heat of combustion, and wet flue gas composition for stoichiometric
combustion of the liquid fuels in Table 3.6.

3.46* Develop a well-organized spreadsheet in which the user may enter a coal
ultimate analysis for C, H, O, N, and S; dry flue gas mole compositions for CO, CO2,
and 02; and as-fired moisture and ash mass fractions to determine theoretical and actual
air-fuel ratios and percentage of excess air.
133

CHAPTER FOUR

ASPECTS of STEAM
POWER PLANT DESIGN

4.1 Introduction

After studying the fundamental thermodynamic cycles of steam power plants and con-
sidering the characteristics and thermochemistry of fuels, it is appropriate to consider
the design of the systems and flow processes that are operative in steam plants and
other large-scale power production facilities. This chapter will focus first on the
processing of several fundamental streams that play a major role in power plant
operation. Up to this point, a great deal of attention has been focused on the water path
from the point of view of the thermodynamics of the steam cycle. Additional aspects of
the water path related to plant design are considered here.
Another fundamental flow in the power plant, the gas stream, includes the intake
of combustion air, the introduction of fuel to the air stream, the combustion process,
combustion gas cooling in the furnace heat exchange sections, and processing and
delivery of the gas stream to the atmosphere through a chimney or stack. A third
important stream involves the transportation and preparation of fuel up to the point that
it becomes part of the combustion gas.
A major non-physical aspect of power production is the economics of power plant
design and operation. This is considered in conjunction with some preliminary design
analyses of a prototype plant. Environmental considerations also play an important part
in planning and design. The chapter concludes with back-of-the-envelope type
calculations that define the magnitudes of the flows in a large plant and identify major
design aspects of steam power plants.

4.2 The Water Path

The Liquid-Water-to-Steam Path

Several pumps are employed in the feedwater path of a steam power plant to push the
working fluid through its cycle by progressively elevating the pressure of the water
from the condenser to above the turbine throttle pressure. These pumps are usually
driven by electric motors powered by electricity generated in the plant or by steam
turbines powered by steam extracted from the main power cycle.
134

The power requirement of a pump is proportional to the liquid mass-flow rate and the
pump work, as given by Equation 2.9, and inversely proportional to the pump
efficiency:

Power = mvsat p/pump [ft-lbf /s | kW]

The pumps are required to overcome frictional pressure losses in water-flow and
steam-flow passages, to provide for the pressure differences across turbines, and to
elevate the liquid to its highest point in the steam generator. The pump power
requirements are typically a small percentage of the gross power output of the plant.
Thus condensate leaving the condenser passes through one or more pumps and
feedwater heaters on its way to the steam generator. A typical shell-and-tube closed
feedwater heater is shown in Figure 4.1. Normally, feedwater passes through the tubes
while extracton steam enters at the top and condenses as it flows over the tubes to the
bottom exit.
135

After passing through the chain of feedwater heaters and pumps, the feedwater
enters the steam generator through the economizer. An economizer is a combustion-
gas-to-feedwater tubular heat exchanger that shares the gas path in a steam generator,
as seen in Figure 4.2. The economizer heats the feedwater by transferring to it some of
the remaining energy from the cooled exhaust gas before the gas passes to the air
heater, the pollution control equipment, and the stack.
136

Steam Generators

Figures 4.3 and 4.4 show two-drum steam generators in which the design
vaporization pressure is below the critical pressure of water. Hot, subcooled liquid
feedwater passes from the economizer and through the boiler tube walls to the drum
loop located near the top of the steam generator. Liquid water circulates by free
convection through the many boiler tubes between the drums until it is vaporized by the
hot gas stream flowing over the tubes. A fixed liquid level is maintained in the upper
steam drum, where the steam separates from the liquid and passes to the superheater.
Solids settle in the bottom of the so-called “mud drum” below it.
A steam drum mounted on a railroad flat car en route to a construction site is
shown in Figure 4.5. The many stub tubes around the bottom and on top are to be
137

connected to steam-generating loops and to steam-superheating pipes, respectively, as


seen at the top of Figure 4.4. The large pipes on the bottom and the ends are for
connection to downcomers, which supply recirculated liquid water to various heating
circuits in the steam generator.
Steam produced in the steam drum, at a boiling temperature corresponding to the
vapor pressure in the drum, passes to superheater tube or plate heat-exchanger banks.
The superheater tube banks are located in the gas path upstream of the drum loop, as
seen in Figures 4.3 and 4.4, taking advantage of the highest gas temperatures there to
superheat the steam to throttle temperature. The hottest gases are used to heat the
138

hottest water-tube banks, to minimize the irreversibility associated with the heat
transfer through the large temperature differences between the combustion gas and the
steam or liquid water. The dry steam from the superheater then passes from the steam
generator through the main steam line to the HP turbine. The progression of tube
banks, with decreasing water temperatures exposed to successively cooler gas temp-
eratures from the secondary superheater to the economizer to the air heater are also
seen in the universal-pressure (supercritical pressure) steam generator in Figure 4.2.
As the design throttle steam pressure increases toward the critical pressure of
water, the density difference between liquid water and vapor decreases, finally
vanishing at the critical point (3208.2 psia, 705.47°F.). As a consequence, in steam-
generator boiling loops, natural convection water circulation—which is driven by the
density difference between liquid and steam—becomes impractical at pressures above
about 2500 psia. Thus modern high-throttle-pressure power plants use circulating
pumps to provide forced to circulation to augment or replace natural circulation of
water in the steam generator.
In single-drum steam generators, water flows downward from the steam drum
through large pipes called downcomers located outside the furnace wall, then through
circulating pumps to headers at the bottom of the steam generator. From the headers,
water flows upward in vertical tubes forming the inside of the furnace walls. The water
is heated by the furnace gases as it rises, and eventually boils and forms a two-phase
flow that returns to the steam drum. There, vapor separates and passes to the
superheater.
Steam generators may utilize natural convection flow through downcomers and
139

vapor-laden upward flow through the tube walls alone or may combine natural
convection with the use of booster pumps to provide adequate circulation for a wider
range of operating loads. It is important to recognize that at the same time steam is
being generated in the boiler, the tube walls are being cooled by the water. Adequate
water circulation must be ensured to provide waterside heat transfer rates high enough
to maintain tube wall temperatures below their limiting design values and thereby to
avoid tube failure.
A once-through supercritical steam generator, operates at a throttle pressure
above the critical pressure of water as in the Riverside station discussed in Chapter 2.
There are no drums and no water recirculation in a once-through steam generator.
Water from the economizer passes to the bottom of the furnace, where it starts its
upward flow through the furnace tube walls. Steam formed in the tubes flows upward
to be collected in headers and mixed to provide a unifrom feed to the superheater.
The feedwater passes directly from the liquid to the vapor phase as it is heated at a
pressure above the saturation pressure. It may be compared to a flow of water pumped
through a highly heated tube with a downstream valve. The state of the steam emerging
from the tube depends on the valve setting, the heat addition rate, and the feedwater
flow rate. In the same way, the steam conditions at the turbine throttle may be adjusted
by changing the turbine throttle valve setting, the fuel firing rate, and the feedwater
flow rate. If the flow rate is decreased by closing the throttle valve, it is necessary to
decrease the fuel firing rate to maintain the same thermodynamic conditions at the
throttle. On the other hand, if the rate of heat transfer is increased without changing the
flow rate, the steam discharge temperature will increase. Other adjustments, such as
increasing condenser cooling-water flow rate, may then be appropriate to avoid an
increase in condenser temperature and pressure. Similarly, an increase in fuel flow rate
must be accompanied by an increase in air flow rate to maintain a constant air-fuel
ratio.
In cycles with reheat, the reduced-pressure steam from the HP turbines passes
through the cold reheat line to the reheater section in the steam generator, where the
steam temperature is returned to approximately the original throttle temperature. The
steam then returns to the next turbine through the hot reheat steam line, as Figure 2.13
indicates.
After leaving the LP turbine, low-pressure steam then passes over the
water-cooled tubes in the condenser and returns to the feedwater heating system as
saturated liquid condensate. A condensate pump then raises the pressure of the liquid
and transports it to the first low-pressure feedwater heater, where it begins another trip
through the cycle.
In order to avoid corrosion, scaling and the deposits of solids along the water path
can result in losses of efficiency and unscheduled shutdowns, water of extreme purity is
required in the steam cycle. Chemical and filtration processes are employed to ensure
that high water quality is maintained, to avoid deterioration or clogging of water path
components. An example of the potential deposits when proper water treatment is
neglected is seen in Figure 4.6. The deaerator, an open feedwater heater mentioned in
140

Chapter 2, provides for the removal of noncondensable gases, particularly oxygen,


from the working fluid. The deaerator allows noncondensable gases to escape to the
atmosphere through a vent condenser, while accompanying steam is retained by
condensing it on cool surfaces and returning it to the feedwater heater stream by
gravity flow.
The turbine room at the Bull Run coal-burning power plant of the Tennessee
Valley Authority (TVA) is shown in Figure 2.3. Electrical generators are seen in the
left and right foreground. Behind them, high-pressure turbines on the left are seen
joined to low-pressure turbines on the right by two large, vee-shaped crossover steam
lines. The side-by-side condensers are seen on either side of the low pressure turbines,
a departure from the usual practice of locating the condenser below the low-pressure
turbines. Figure 4.7 shows the turbogenerator room at TVA’s Brown’s Ferry nuclear
power plant with a turbine in the foreground.

The Condenser Cooling-Water Loop

The cooling loop, in which water passes through tubes in the condenser removing heat
from the condensing steam, is an important water path in large steam plants. This
cooling water, clearly separate from the working fluid, is usually discharged into a
nearby body of water (a river, or a natural or man-made lake) or into the atmosphere.
Figure 4.8 shows a typical wood-framed, induced-draft cooling tower used to
dissipate heat from the condenser cooling water into the atmosphere. The tower is
usually located a few hundred yards from a plant. Typically, the cooling water entering
the tower is exposed to a flow of air created by upward-blowing fans at the bases of the
funnels at the top of the towers. A fraction of the condenser cooling water, which
passes over extensive aerating surfaces in the tower, evaporates and exits to the
141

atmosphere, cooling the rest of the water. The remaining chilled water is then returned
to the condenser by a cooling-water circulating pump. A continuing supply of liquid
makeup water is required for these towers to compensate for vapor loss to the
atmosphere.
In areas where large structures associated with power plants are acceptable, large
natural-draft cooling towers may be used. Figure 4.9 shows two large natural-draft
hyperbolic cooling towers serving a large power plant. The height of these towers,
which may reach over 500 feet, creates an upward draft due to the difference in density
between the warm air in the tower and the cooler ambient air. Heat from the condenser
cooling water warms the air, inducing an upward air flow through the heat transfer
surfaces at the base. These towers offer long-term fan-power savings over mechanical
draft towers. Under some conditions, these power savings may offset high construction
costs of hyperbolic towers.
142
143

4.3 The Fuel Path for a Coal-Burning Plant

The supply and handling of fuel for a modern coal-burning power plant is a complex
and expensive undertaking. In contrast to the relatively simple steady flow of fluid fuels
in power plants that consume natural gas and fuel oil, solid-fuel-burning plants offer
major and continuing challenges to engineers. The discussion here focuses on these
operations and their challenges.

Getting the Coal to the Plant

The source of coal for a plant may be a surface mine or a deep underground mine.
Power plants are sometimes located adjacent to mines, where conveyors may provide
the only transportation required. This significantly reduces coal transportation costs
which otherwise can be higher than the cost of the coal alone. Such plants are called
mine-mouth plants. A mine-mouth plant may be an attractive option if its selection
does not result in significant transmission costs to bring the electrical power to distant
load centers where the utilities’ customers are located.
Today the power plant and coal mines are likely to be a considerable distance
apart, perhaps a thousand miles or more. The most widely used modern transportation
link between mine and plant is the unit train, a railroad train of about a hundred cars
dedicated to transporting a bulk product such as coal. Although slurry pipelines (a
slurry is a fluid mixture of solid lumps and liquid, usually water, which can be pumped
through pipes by continuous motion) sometimes offer attractive technical solutions to
coal transport problems, economic and political forces frequently dictate against their
use. Dedicated truck transport is an occasional short-haul solution, and barges are
sometimes used for water transport. Here, we will focus on unit trains.
Several unit trains may operate continuously to supply a single plant. Trains
carrying low-sulfur coal from Wyoming, Montana, and the Dakotas supply coal to
plants as distant as Michigan, Illinois, and Oklahoma. This strange situation, in which
utilities located in states with large quantities of coal purchase coal from distant states,
was a response to pollution control requirements. It was preferred to purchase low-
sulfur coal from a distant state rather than pay a high price for sulfur removal
equipment (perhaps 10% of the cost of the plant construction), some of which has a
reputation for unreliability. In response to such choices, the Oklahoma legislature
passed a law requiring utilities to burn at least 10% Oklahoma coal in their
coal-burning steam generators. Such mixing of small quantities of high-sulfur coal with
low-sulfur coal is an expedient to protect local businesses and to spread out resource
utilization geographically.
New plants, however, no longer have the option to choose low-sulfur coal
or sulfur removal equipment. “Best available control technology,” BACT, has become
the rule. The Environmental Protection Agency now requires new plants to have
scrubbers (sulfur removal equipment) even if the plants use low-sulfur coal, and they
are required to employ the currently most effective pollution control technology.
144

Coal Unloading and Storage

On arrival at the plant, the unit train passes through an unloading station. Some coal
cars have doors on the bottom that open and dump their load to a conveyor below.
Others have couplings between cars that allow the rotation of individual cars about
their coupling-to-coupling axis, by a dumping machine, without detachment from the
train, as seen in Figure 4.10. The figure shows a breaker in the dumping facility that
reduces large coal chunks to a smaller, more uniform size for transport on a belt
conveyor. The under-track conveyor at the unloading station then carries the newly
arrived coal up and out to a bunker or to a stacker-reclaimer in the coal yard as seen in
Figure 4.11. The stacker-reclaimer either feeds the coal through a crusher to the plant
or adds it to the live storage pile.
A permanent coal storage pile sufficient to supply the plant for several months is
usually maintained. While the first-in-first-out approach common in handling perishable
goods seems logical, a first-in-last-out storage system is usually used. A primary
reason for this approach is the hazard and expense of coal pile fires, which can occur
due to spontaneous combustion. Once a stable storage pile is achieved by packing and
other treatment to restrict air acccess, it is usually not disturbed unless coal must be
withdrawn by the stacker-reclaimer to satisfy unusual demands caused by labor strikes,
extreme weather, rail accidents, or the like.
145
146

A conveyor transports coal from the reclaimer to a crusher house, where hammer
mills, ball crushers, or roller crushers break up large chunks to a more manageable
size. Another conveyor may then carry the crushed coal to one of several bunkers or
silos for temporary storage prior to firing. Some of these features may be seen in the
photograph of the PSO Northeastern Station in Figure 2.1.
The rate of feeding coal from the silos is controlled to maintain the desired steam
generator energy-release rate. In a pulverized-coal plant, the coal is fed from the silos
to pulverizers, where it is further reduced in size to a powdery form. Warm air drawn
through an air preheater in the steam generator by the primary air fan flows through
the pulverizer, where it picks up the fine coal particles and transports them pneumatic-
ally through piping to the steam generator burners. Several arrangements of silos,
feeders, pulverizers, and pneumatic transport systems are seen in Figures 4.2 to 4.4.

4.4 The Gas Path

Fans

While natural or free convection may be used to provide combustion air to small boilers
and heaters, modern power plants employ large fans or blowers to circulate air to the
burners and to assist flue gas in escaping from the furnace. These fans are called
forced-draft fans and induced-draft fans, respectively. A common arrangement of
these fans is shown in Figure 4.4. Atmospheric air drawn into the steam generator by
one or more forced-draft fans is heated as it passes through the cold gas side of an air
heater on its way to the furnace. At the same time, combustion gases that have passed
through the furnace heat transfer sections are cooled as they passed through the hot
side of the air heater on their way to induced-draft fans and thence to the stack. In the
case of pulverized-coal-burning plants, primary air fans, as seen in Figures 4.3 and 4.4,
supply enough pre-heated air to pulverizers to transport coal pneumatically to the
burners. Primary air usually pre-heated to 300-600°F to dry the coal as it passes
through the pulverizer.
With a forced-draft fan alone, the furnace pressure is above atmospheric pressure,
causing large outward forces on the furnace walls and a tendency for leakage of
combustion gas from the furnace. On the other hand, the use of an induced-draft fan
alone would cause the furnace pressure to be below atmospheric pressure, producing
large inward forces on the walls and possible air leakage into the furnace. The forces
on the walls, which can be significant, can be minimized by keeping the furnace
pressure near atmospheric by using balanced draft, that is, the use of both forced- and
induced-draft fans, which produce a gas path pressure distribution such as shown by
the heavy line in Figure 4.12. In this design the forced-draft fan raises the pressure to
15 in. of water gauge entering the steam generator, and the induced-draft fan depresses
its inlet pressure about 21 in. of water below atmospheric. As a result, the furnace
inside-wall pressure is less than an inch of water below atmospheric. This substantially
reduces both the potential for furnace leakage and the forces on the furnace walls.
147

The power requirements for fans may be determined in much the same way that
pump power requirements are determined. Fans are primarily gas-moving devices that
produce small pressure rises. Pressure and density changes across fans are usually
small fractions of the fan inlet values. This justifies the approximation that the fan
process is incompressible. Fan power requirements then closely follow the pump
power prediction method discussed earlier. Thus, for a forced-draft fan, power may be
estimated by using

PowerFD = Qair pair/FDfan [ft-lbf/s | kW]

For the induced-draft fan

PowerID = ,airQair pgas(1 + F/A)/,gasFDfan [ft-lbf/s | kW]

where Qair is the volume flow rate of air entering the forced-draft fan. The second
equation accounts for the additional fuel mass handled by the induced-draft fan and the
148

density of the gas leaving the furnace, assuming no leakage or diversion of air from that
leaving the forced-draft fan.
The drawing of a centrifugal forced-draft fan is shown in Figure 4.13; a
photograph of a rotor and open housing is presented in Figure 4.14. In a centrifugal
fan, air is spun by the rotor blades, producing tangential motion and pressure rise and
leaving behind a vacuum for air to flow in along the axis of the fan. It the fan entrance
is open to the atmosphere, its exhaust is pressurized; if its exhaust is atmospheric, its
entrance pressure is below atmospheric. Fans typically do not produce large pressure
rises but do produce large flows of gases.
A diagram of an axial-flow fan designed for induced-draft use is shown in Figure
4.15. Figure 4.16 presents a photo of the same type of fan. Induced-draft fans
must be able to withstand high-temperature service and erosion due to airborne
particulates. Large electrostatic precipitators located upstream of the induced-draft
fans remove most of the flyash by inducing a static charge on the flowing particles
and collecting them on plates of opposite charge. The plates are periodically rapped
mechanically to free ash deposits that drop to the bottom of the precipitator and are
collected and removed. Figure 2.1 shows electrostatic precipitators, to the right of the
steam generator. The large structure behind the fans, air heater, and ducting and below
the stack in Figure 4.4 is an electrostatic precipitator.

Air Preheaters

The air leaving the forced-draft fan usually flows through an air preheater to a windbox
around the furnace and then to the burners. A Ljungstrom rotary air preheater, used
149
150

in many large plants, is shown in Figures 4.17 and 4.18. The rotary air heater is a
slowly rotating wheel with many axial-flow passages, having large surface area and
heat capacity, through which air and flue gas pass in counterflow parallel to the wheel
axis. When the wheel surfaces heated by the flue gas rotate to the air side, they are
cooled by the air from the forced-draft fan. As result, the air temperature rises several
hundred degrees before passing to the furnace windbox.
Although the Ljungstrom rotary air heater is widely used in utility and industrial
power plants, heat-pipe air and process heaters are now being considered and applied
for use in power plants and industry. A heat pipe, shown in Figure 4.19, is a sealed
tube in which energy is transported from one end to the other by a thermally driven
vapor. The heat-pipe working fluid absorbs heat and vaporizes at the lower, hot end.
After rising to the higher, cold end, the vapor condenses, releasing its heat of
vaporization, which is carried away by conduction and convection through external fins
to the combustion air. The liquid then returns to the hot end by gravity and/or by
capillary action through wicking, to complete the cycle. The wicking may be spiral
grooves around the inside of the tube that ensure that the entire inside surface is wetted
for maximum heat transfer. The wicking in the cold section is particularly important,
because it provides increased surface area that increases inside-gas heat transfer rates.
The outside the tube is usually finned to provide adequate external heat transfer rates,
both from the flue gas and to the incoming air.
151
152

As a heat transfer device, a well-designed heat pipe has an effective thermal


conductance many times that of a copper rod. Note that the energy transfer inside the
heat pipe is essentially isothermal, since the liquid and the vapor are in near
equilibrium. Although heat pipes will operate in a horizontal orientation, their
effectiveness is augmented by gravity by inclining them about 5°– 10° to assist in liquid
return to the hot end.
In power plant air heater applications (see Figures 4.19 and 4.20), finned heat
pipes supported by a central partition between the incoming air stream and the flue gas
153

stream are free to expand outward. This reduces thermal expansion problems and
virtually eliminates the possibility of leakage between the flows. Such heaters often
have been installed in process plants and have been retrofitted in power plants originally
built without air heaters, because of their ease of installation and compact size
compared with stationary tubular air heaters.

EXAMPLE 4.1

The heat-pipe air heater in Figure 4.20 has an air flow of 360,800 lbm/hr and a flue gas
flow rate of 319,000 lbm/hr. The flue gas enters at 705°F and leaves at 241°F; the
combustion air enters at 84°F. What is the rate of energy recovery from the flue gas,
and what is the air temperature entering the windbox? Assume a flue gas heat capacity
of 0.265 Btu/lbm-R.

Solution
The rate of heat recovery from the flue gas is

mcp (Tout – Tin) = 319,000(0.265)(705 – 241) = 39,224,240 Btu/hr

The air temperature is then

84 + 39,224,240/[(0.24)(360,800)] = 537°F
_____________________________________________________________________

An analysis of the gas flow through a steam generator must take into account the
streamwise pressure rise through the fans and pressure losses due to friction and losses
through flow restrictions and turns. These include losses due to flow through furnace
tube and plate heat exchanger banks and other passages, such as in both the air and the
gas passes through the air heater.

Power Plant Burners

Burner design depends on the choice of fuel and the steam generator design. Figure
4.21 shows a burner designed for forced-draft applications burning natural gas and oil.
Oil and steam under pressure are mixed in the central feed rod to atomize the oil to a
fine mist coming out of the oil tip. The cone at the oil tip stabilizes the flame in the
surrounding air flow. The gas pilot next to the oil feed rod provides a continuous
ignition source. A separate duct for the natural gas supply feeds gas to the two types
of gas tip. Separate air registers control the flow of air to the gas and oil tips. Figure
4.22 shows an oil burner, with a water-cooled throat, installed in a furnace wall.
Registers that control the flow of air from the windbox are also visible.
154

In the case of the pulverized-coal plant, primary air flows through the pulverizer and
carries the fuel directly to the burners (Figure 4.23). Secondary (and sometimes
tertiary) air helps to control the temperature of the control nozzle and of the furnace
wall, and mixes with the combustion gases to provide for essentially complete
combustion of the fuel. Features of a burner designed for pulverized-coal firing in
planar furnace walls are shown in Figure 4.24. Note that the secondary air flow
through the windbox registers helps to cool the nozzle through which the coal and the
primary air flow.
Another approach to burning pulverized coal uses corner burners in tangentially
fired steam generators. A plan view of such a furnace is shown in Figure 4.25. The
burners induce a circular motion, in the horizontal plane, on the upward-rising
combustion gases, promoting vigorous mixing, which hastens completion of
combustion in the furnace. For control purposes, the corner burners can be tilted in the
vertical direction to adjust the furnace heat transfer distribution.
A cyclone furnace type of burner installation, used for burning slagging coals
(those that form liquid ash, or slag, at moderate temperatures) in steam generators, is
shown in Figure 4.2. The cyclone furnace is a cylindrical furnace with very large
155
156

volumetric heat release rates that lies adjacent to and opens onto the main furnace.
Details of a cyclone furnace are shown in Figures 4.26 and 4.27.
The coal supplied to cyclone furnaces, which is crushed but not pulverized, is fed
to the cyclone by a mechanical feeder. The coal and primary air entering the cyclone
157

move tangentially to the inside of the furnace cylinder. There the momentum of the
coal carried by the swirl flows forces the coal pieces toward the cylindrical burner wall.
The very high temperature and vigorous mixing produce a high rate of burning. As a
result, combustion is virtually complete by the time the combustion gas flow enters the
158

main furnace. The cooled walls stimulate formation of a protective slag layer on the
cylinder walls. Because the main furnace is only required for steam generation and to
cool the combustion gases, and not to provide time for completion of combustion, the
cyclone furnace steam generator can be significantly smaller than the pulverized-coal
steam generator. A steady flow of slag drains from the cyclone furnace into a slag tank
at the bottom of the main furnace.
In both cyclone and pulverized-coal steam generators the combustion gases flow
upward from the burners, transferring heat to the tube walls by radiation and
convection. The cooling gases then flow through superheater, reheat, and boiler tube or
plate sections. The combustion gas temperature drops as it passes through these steam
generator sections in essentially a counterflow arrangement with the water flow. The
combustion gases undergo their final cooling as they pass through the economizer and
then the air preheater. From there they pass through an electrostatic precipitator for
removal of airborne particles and through scrubbers for control of oxides of nitrogen
and sulfur (NOx and SOx) and through an induced-draft fan before entering the stack.
The serious degradation of the environment caused by oxides of sulfur and
nitrogen in the flue gas of power plants and from other sources has led to widespread
chemical processing of flue gases. Figure 4.28 shows a schematic diagram of the gas
flow path for removal of NOx and SOx after particulate removal in the precipitator and
passage through an induced-draft fan. In this scheme, gas-to-gas heat exchangers
(GGH) provide the proper temperatures for the flue gas desulfurization (FGD) unit and
the DENOX catalyst unit. This additional equipment increases the pressure drop
through the system, sometimes necessitating an additional fan.
With high smokestacks, the stack effect also influences the gas path and must be
taken into account.The stack effect is the upward movement of exhaust gas produced
by the density difference between the hot gases inside the stack and the surrounding
cooler atmospheric air. Because of hot gas buoyancy, a smaller pressure gradient along
the stack length is required to expel the combustion gases from the stack. This effect is
opposed by the usual viscous friction pressure losses. The diameter and height of the
stack control the relative influence of frictional forces in opposing the stack effect.
Other considerations, such as cost and a possible need to disperse the stack gas above a
particular height, also have a significant influence on these dimensions.
It is obvious that the air heater of the steam generator should extract as much
energy from the combustion gas as possible to maximize its regenerative heat transfer
effect. This implies cooling the gas to a low tempeature. However, practical limits exist
on the minimum combustion gas temperature, to avoid the condensation of water vapor
in the presence of sulfur and nitrogen compounds in the gas and to meet the temp-
erature requirements of the pollution control equipment. Condensation of water vapor
in the presence of gaseous oxides of sulfur and nitrogen leads to the formation of acids
that erode the materials on which the liquid condenses. The temperature at which the
vapor condenses is called the acid dew point. Typical acid dew points for coal range to
about 320°F. As a result, stack gas design temperatures may exceed that value,
depending on the coal and flue gas treatment.
159

4.5 Introduction to Engineering Economics

The success of any engineering undertaking depends on adequate financial planning


to ensure that the proceeds of the activity will exceed the costs. The construction of a
new power plant or the upgrading of an old one involves a major financial investment
for any energy company. Financial planning therefore starts long before ground is
broken, detailed design is begun, and orders are placed for equipment. Cost analysis
and fiscal control activities continue throughout the construction project and the
operating life of the plant. This section briefly introduces fundamentals of engineering
economics, with a slant toward power plant cost analysis as well as issues of
maintenance and equipment replacement.
160

The cost to construct a power plant, waterworks, dam, bridge, factory, or other
major engineering work is called its capital cost. It is common to discuss the capital
cost of building a power plant in terms of dollars per kilowatt of plant power output. A
plant may cost $1100 per kilowatt of installed power generation capacity, for instance.
In addition to the cost of building the plant, there are many additional expenditures
required to sustain its operation. These are called operating costs. They may be
occasional, or they may occur regularly and continue throughout the life of the plant.
Often these costs are periodic, or are taken to be periodic for convenience of analysis.
There are, for instance, annual fuel costs, salary expenses, and administrative and
maintenance costs that are not associated with the initial cost of the plant but are the
continuing costs of generating and selling power. Operating costs are sometimes
related to the amount of electrical energy sold. Usually they are expressed in cents per
kilowatt-hour of energy distributed to customers.
Thus the expenses associated with power generation and other business endeavors
may be thought of as two types: (1) initial costs usually associated with the purchase of
land, building site preparation, construction, and the purchase of plant equipment; and
(2) recurring operating costs of a periodic or cyclic nature.
It is frequently desirable to express all costs on a common basis. The company and
its investors may wish to know what annual sum of money is equivalent to both the
capital and operating costs. The company may, for example, borrow money to finance
the capital cost of the plant and then pay the resulting debt over the expected useful life
of the plant, say, 30 or 40 years. On the other hand, they may wish to know what
present sum would be required to ensure the payment of all future expenses of the
enterprise.
It is clear that $100 in hand today is not the same as $100 in hand ten years from
now. One difference is that money can earn interest. One hundred dollars invested
today at 8% annual compound interest will become $215.89 in ten years. Clearly, an
important aspect of engineering economics is the time value of money.

Compound Interest

If Alice lends Betty $500, who agrees to pay $50 each year for five years for the use of
the money, together with the original $500, then at the end of the fifth year, Alice will
have earned $250 in simple interest and receive a total of $750 in return. The annual
interest rate is

i = Annual interest / Capital = 50 / 500 = 0.1

or (0.1)(100) = 10% rate of return.


If, however, Betty keeps the interest instead of paying it to Alice annually, and
eventually pays 10% on both the retained interest and the capital, the deal involves
compound interest. The total sum to be returned to Alice after 5 years is computed as
follows: At the end of the first year Alice has earned $50 in interest. The interest for the
161

next year should be paid on the original sum and on the $50 interest earned in the first
year, or $550. The interest on this sum for the second year is 0.1 × $550 = $55. The
following table shows the calculation of the annual debt for the five-year loan of $500
at 10% interest:
At the End The Accumulated Debt is: This Sum
of:

First year $500 + 0.1 x $500 = $550.00

Second year $550 + 0.1 x $550 = $605.00

Third year $605 + 0.1 x $605 = $665.50

Fourth year $665.50 + 0.1 x $665.50 = $732.05

Fifth year $732.05 + 0.1 x $732.05 = $805.26

It is evident that the interest earned on the preceding interest accumulation causes
the annual indebtedness to grow at a increasing rate. It can be shown that the future
sum, S, is given by

S = P (1 + i)n
where P is the principal, the initial sum invested; i is the interest rate, and n is the
number of investment periods, in this case the number of years. Here the factor
multiplying the principal,

S / P = (1 + i)n

is called the compound amount factor, CAF. The difference between simple and
compound interest may not be spectacular for short investment periods but it is very
impressive for long periods of time such as the operating life of a power plant. For our
example, the CAF is (1 + 0.1)5 = 1.6105, and S = 500(1.6105) = $805.26. Now,
consider the following closely related problem.

EXAMPLE 4.2

What sum is required now, at 8% interest compunded annually, to produce one million
dollars in 25 years?

Solution
The future sum is

S = P (1 + i)n = 1,000,000 = P (1 + 0.08)25

Solving for P, the present sum is 1,000,000/(1.08)25 = $146,017.90. Thus, compound


interest brings a return of almost over seven times the original investment here. The
162

same present sum invested at 8% simple interest for twenty-five years would produce a
future sum of less than half a million dollars.
_____________________________________________________________________

In the example, the inverse of the CAF was used to determine the present worth of
a future sum. The inverse of the CAF is called the present-worth factor, ( PWF):

PWF = P/ S = 1 / (1 + I)n

Thus we see that the time value of money is related to the compound interest that
can be earned, and that taking compounding into account can be important. To
recklessly adapt an old adage, “A dollar in the hand is worth two (or more) in the
future (if invested wisely).”

Capital Recovery

Another important aspect of compound interest is the relationship between a


present sum of money and a regular series of uniform payments. Consider a series of
five annual payments of R dollars each, when the interest rate is i. What is the present
dollar equivalent of these payments? Applying the CAF as in the preceding example,
with R as the future sum, the present sum associated with the first payment is R/(1 + i).
The present sum associated with the second payment is R/(1 + i)2. Thus the present
worth of the five payments is

P = R [ (1 + i) – 1 + (1 + i) – 2 + (1 + i) – 3 + (1 + i) – 4 + (1 + i) – 5 ]

It may be shown that this expression can be written as

P = R [(1 + i)5 – 1]/[i(1 + i)5].

The factor multiplying the annual sum R is called the series present-worth factor,
SPWF, which for n years is:

SPWF = P/ R = [(1 + i)n – 1]/[i(1 + i)n] (4.1)

Solving for R, we obtain an expression for the regular annual payment for n years
needed to fund a present expenditure of P dollars at an interest rate i. The resulting
factor is called the capital recovery factor, CRF, which is the reciprocal of the series
present worth factor:

CRF = R / P = i(1 + i) n / [(1 + i) n – 1] (4.2)


163

EXAMPLE 4.3

What uniform annual payments are required for forty years at 12% interest to retire the
debt associated with the purchase of a $500,000,000 power plant?

Solution
Using equation (4.2), we get

R = Pi (1 + i)/[(1 + i) n – 1] = 5×108(0.12)(1 + 0.12) 40/(1.1240 – 1)


= $60,651,813

This sum may be regarded as part of the annual operating expense of the plant. It must
be recovered annually by the returns from the sale of power.
_____________________________________________________________________

4.6 A Preliminary Design Analysis of a 500-MW Plant

Consider the design of a 500-megawatt steam power plant with a heat rate of
10,000 Btu/kW-hr and a water-cooled condenser with a 20°F cooling-water
temperature rise produced by heat transfer from the condensing steam. The plant uses
coal with a heating value of 10,000 Btu/lbm. Let us estimate the magnitude of some of
the parameters that characterize the design of the plant. The reader should verify
carefully each of the following calculations.
A 500-megawatt plant operating at full load produces 500,000 kW and an annual
electrical energy generation of

500,000 ( 365 ( 24 = 4.38 × 109 kW-hr

With a heat rate of 10,000 Btu/kW-hr, this requires a heat addition rate of

500,000 ( 10,000 = 5 × 109 Btu/hr

Coal with an assumed heating value of 10,000 Btu/lbm must therefore be supplied at a
rate of 5 × 109 / 104 = 500,000 lbm/hr or 500,000 / 2000 = 250 tons/hr. A dedicated
coal car carries about 100 tons. Hence the plant requires 250 /100 = 2.5 cars per hour
of continuous operation. A coal unit train typically has about 100 cars. Then the plant
needs 2.5 ( 24 / 100 = 0.6 unit trains per day, or a unit train roughly every two days.
If coal costs $30 per ton, the annual cost of fuel will be

30 ( 250 ( 24 ( 365 = $65,700,000

The cost of fuel alone per kW-hr, based on 100% annual plant capacity, will be

65,700,000/(500,000 ( 365 ( 24) = $0.015/kW-hr ; 1.5 cents/kW-hr


164

The annual plant factor, or annual capacity factor, expressed as a decimal fraction , is
the ratio of the actual annual generation to the annual generation at 100 % capacity.
If the coal has 10% ash, the plant will produce 250 ( 0.1 = 25 tons of ash per hour.
Under some circumstances the ash may be used in the production of cement or other
paving materials. If it is not marketable, it is stabilized and stored in nearby ash ponds
until it can be moved to a permanent disposal site.
Similarly, if 2% of the coal is sulfur and half of it is removed from the combustion
products, 2.5 tons per hour is produced for disposal. If the sulfur is of sufficient purity,
it may be sold as an industrial chemical.
With an air-fuel ratio of 14, an air flow rate of 14 ( 500,000 = 7,000,000 lbm/hr is
required for combustion. This information is important in determining the size of the
induced- and forced-draft fans, that of their driving motors or turbines, and of the
plant’s gas path flow passages.
The heat rate of 10,000 Btu/kW-hr corresponds to a thermal efficiency of
3413/10,000 = 0.3413 or 34.13%. If we approximate the heat of vaporization of water
as 1000 Btu/lbm, the throttle steam flow rate, with no superheat, would be about

10,000 ( 500,000 / 1000 = 5,000,000 lbm/hr

This determines the required capacity of the feedwater pumps and is important in sizing
the passages for the water path. The above thermal efficiency implies that about 65% of
the energy of the fuel is rejected into the environment, mostly through the condenser
and the exiting stack-gas energy. As an upper limit, assume that all of the heat is
rejected in the condenser. Thus

(1 – 0.3413)(5 × 109) = 3.29 × 109 Btu/hr

must be rejected to condenser cooling water. With 20° water temperature rise in the
condenser, this rate of cooling requires a cooling-water flow rate to the condenser of

3.29 × 109/(1.0 × 20) = 1.65 × 108 lbm / hr

assuming a water heat capacity of 1.0 Btu/lbm-R. This gives information relevant to the
design sizing of cooling-water lines, cooling towers, and water pump capacities.
These back-of-the-envelope calculations should not be regarded as precise, but
they are reasonable estimates of the magnitudes of important power plant parameters.
Such estimates are useful in establishing a conceptual framework of the relationships
among design factors and of the magnitude of the design problem.

EXAMPLE 4.4

Relating to the above rough design of a 500-MW plant, and assuming the capital cost
information of Example 4.3, determine the capital cost per kW of generation capacity
and estimate the minimum cost of generation for the plant if it is predicted to have an
165

annual plant factor of 80% and maintenance and administrative costs of $0.007 /kW-hr.

Solution
The unit cost of the power plant is

$500,000,000/(500 (1000) = $1000 per kW-hr

of capacity. The capital cost part of the annual cost of power generation is

(60,651,813 (100)/(365 ( 24 ( 0.8 (500,000) = 1.73 cents per kW-hr

The cost of coal was determned to be 1.5 cents/kW-hr. The minimum cost of
producing electricity is then

1.73 + 1.5 + 0.7 = 3.93 cents per kW-hr


_____________________________________________________________________

Bibliography and References

1. Singer, Joseph G., (Ed.), Combustion /Fossil Power Systems. Windsor, Conn.:
Combustion Engineering, Inc.,, 1981.

2. Anon., Steam, It’s Generation and Use. New York: Babcock and Wilcox, 1978.

3. Hensley, John C., Cooling Tower Fundamentals. Mission, Kan.: Marley Cooling
Tower Co., 1985.

4. Li, Kam W., and Priddy, A. Paul, Power Plant Systems Design. New York: Wiley,
1985.

EXERCISES

4.1 Derive an equation for the sum, S, resulting from P dollars invested at simple
interest rate i for a period of n years.

4.2 For the power plant design discussed in Section 4.6, estimate the horsepower of a
motor required to drive the fans used to overcome a steam generator gas-path pressure
drop of 1 psia. Assume a fan efficiency of 80%. What is the fractional and percentage
reduction in power plant output due to the fans?

4.3 Estimate the horsepower required by the feedwater pumps in the Section 4.6
design if the HP-turbine throttle pressure is 3200 psia. Assume a pump efficiency of
70%. What fractional and percentage reduction of the power plant output does this
represent?
166

4.4 What are the annual savings in fuel costs in the Section 4.6 plant design if the plant
heat rate can be reduced to 8500 Btu/kW-hr?

4.5 If the total capital cost of the Section 4.6 plant design is $600,000,000 and the
annual administrative and maintenance costs are one cent per kW-hr, what is the
minimum cost of electricity per kW-hr, assuming an annual interest rate of 9% and an
expected plant lifetime of thirty-five years?

4.6 What is the present worth of a sequence of five annual payments of $4500, $6500,
$3500, $7000, and $10,000 at an annual interest rate of 8%?

4.7 You have collected the following data on 1.5-MW steam turbines, as alternatives to
the purchase of utility power, for a new process plant to operate at 60% plant factor:

Turbine Number
1 2 3
Heat rate, Btu/kW-hr 12,000 10,600 9,500
Installed cost $124,000 $190,000 $245,000
Estimated annual $2,000 $1,800 $2,550
maintenance cost

Coal (14,000 Btu/lbm) is the fuel to be used, at a cost of $26 per ton. Assuming an
annual interest rate of 8%, compare the annual cost of of the turbines for thirty-year
turbine lifetimes. Which turbine would you select? What other factors would you
consider before making a decision?

4.8 Using the data of Exercise 4.7, compare the turbines on the basis of present worth
of all costs.

4.9 For the power plant design discussed in Section 4.6, estimate the motor horse-
power required to drive a 75% efficient fan that is used to overcome a steam generator
gas-path pressure drop of 50 kPa.

4.10 Estimate the total power required by 65% efficient feedwater pumps operating in
parallel in the Section 4.6 design for a throttle pressure of 20 MPa.

4.11 Work out a back-of-the-envelope analysis similar to that of Section 4.6 in SI units.

4.12* Develop an interactive computer program that implements a steam power plant
system analysis of the type presented in Section 4.6 at one of the following levels, to be
167

assigned by your instructor.


Level 1: User supplies parameters in response to screen prompts, and one or more
output screens display the resulting input and output parameters.
Level 2: Same as Level 1, but also provide a capability for the user to change the
design by varying one input while holding all others constant.
Level 3: Allow the user to select a dependent variable from a list of outputs, and a
parameter to be varied and its range from another list. Display a graph of the variation
of user-selected outputs over the range of the parameter.

4.13* Construct a spreadsheet that systematizes the computations for a steam power
plant along the lines presented in this chapter. Set up a version of the spreadsheet that
allows easy variation of input parameters. Use the spreadsheet to develop graphs that
show the influence of plant heat rate on fuel costs and sulfur byproduct production.

4.14 An electric utility, expecting to increase its system capacity by 400 megawatts,
must choose between a high-technology combined-cycle plant, at a cost of $1800 per
kW of installed capacity, and an oil-burning steam plant, at $1150 per kW. The
combined-cycle plant has a variable cost of 18 mills per kW-hr, while the oil-burning
plant variable cost is 39 mills per kW-hr. For an annual plant factor of 0.6 and fixed
charges of 15% of the capital cost, determine (a) the total annual cost of operation of
each plant, and (b) the cost of electricity, in cents per kW-hr, for each plant.

4.15 Estimate the mass-flow rate of makeup water required by an evaporative cooling
tower satisfying the cooling requirements of the example power plant of Section 4.6.

4.16 An 800-MW steam power plant operates at a heat rate of 8700 Btu/kW-hr. It has
a 16°F rise in condenser cooling-water temperature. Neglecting energy losses, estimate
the condenser cooling-water flow rate and the flow-rate of cooling-tower makeup
water. Estimate the amount of pump power required to circulate the cooling water to
the cooling tower.

4.17 Plot a curve of condenser cooling-water flow rate and makeup-water flow rate as
a function of condenser cooling-water temperature rise for Riverside Station Unit #1.

4.18 The average temperature in an 800 ft. high power plant exhaust gas stack is 350°F
and the ambient temperature is 60°F. Neglecting fluid friction and exhaust gas kinetic
energy, estimate the pressure inside the base of the stack.

4.19 The average temperature in a 200 meter power plant exhaust gas stack is 150°C,
and the ambient temperature is 20°C. Neglecting fluid friction and exhaust gas
momentum, estimate the pressure, in kPa, at the inside of the base of the stack.
168

4.20 Estimate the heat transfer rates in the Riverside Station Unit #1 in the air pre-
heater, an economizer that heats liquid water to saturation, the boiling surfaces, the re-
heater, and the superheater. Estimate the temperature drops in the combustion gas
across each of these, assuming that they are arranged in the same order as just listed.

4.21 After completing Exercise 4.20, estimate the flow area of combustion gas through
a crossflow economizer in the Riverside Station Unit #1, and define a suitable design.

4.22 After completing Exercise 4.20, estimate the flow area of combustion gas through
a pendant superheater consisting of parallel U-tubes in cross flow, and define a suitable
design.

4.23 The following is a list of ten air and gas path components of a steam power plant.
Number the components so as to put them in order, starting with the air into the plant
as number 1 and concluding with the flue gas out as number 10.
_________ superheater __________ boiler tubewall
_________ economizer __________ windbox
_________ induced-draft fan __________ air heater, gas side
_________ burner __________ electrostatic precipitator
_________ forced-draft fan __________ air heater, air side

4.24 Upon hiring on with Hot Stuff Engineering Company after graduation, you
purchase a $30,000 automobile to establish an image as a prosperous engineer. You
pay no money down, but 1% interest per month, compounded monthly, for four years.
What are your monthly payments? What will your payments be if you are paying
simple interest?

4.25 A forced-draft fan with an efficiency of 70% supplies 1,000,000 ft3 per minute of
air to a furnace that produces a pressure drop of 0.7 psia. What is the fan power
requirement, in horsepower and in kilowatts?
169

CHAPTER FIVE

GAS TURBINES AND JET ENGINES

5.1 Introduction

History records over a century and a half of interest in and work on the gas turbine.
However, the history of the gas turbine as a viable energy conversion device began
with Frank Whittle's patent award on the jet engine in 1930 and his static test of a jet
engine in 1937. Shortly thereafter, in 1939, Hans von Ohain led a German
demonstration of jet-engine-powered flight, and the Brown Boveri company intro-
duced a 4-MW gas-turbine-driven electrical power system in Neuchatel, Switzerland.
The success of the gas turbine in replacing the reciprocating engine as a power plant for
high-speed aircraft is well known. The development of the gas turbine was less rapid as
a stationary power plant in competition with steam for the generation of electricity and
with the spark-ignition and diesel engines in transportation and stationary applications.
Nevertheless, applications of gas turbines are now growing at a rapid pace as research
and development produces performance and reliability increases and economic benefits.

5.2 An Ideal Simple-Cycle Gas Turbine

The fundamental thermodynamic cycle on which gas turbine engines are based is called
the Brayton Cycle or Joule cycle. A temperature-entropy diagram for this ideal cycle
and its implementation as a closed-cycle gas turbine is shown in Figure 5.1. The cycle
consists of an isentropic compression of the gas from state 1 to state 2; a constant
pressure heat addition to state 3; an isentropic expansion to state 4, in which work is
done; and an isobaric closure of the cycle back to state 1.
As Figure 5.1(a) shows, a compressor is connected to a turbine by a rotating shaft.
The shaft transmits the power necessary to drive the compressor and delivers the
balance to a power-utilizing load, such as an electrical generator. The turbine is similar
in concept and in many features to the steam turbines discusssed earlier, except that it is
designed to extract power from a flowing hot gas rather than from water vapor. It is
important to recognize at the outset that the term "gas turbine" has a dual usage: It
designates both the entire engine and the device that drives the compressor and the
load. It should be clear from the context which meaning is intended. The equivalent
term “combustion turbine” is also used occasionally, with the same ambiguity.
170

Like the simple Rankine-cycle power plant, the gas turbine may be thought of as a
device that operates between two pressure levels, as shown in Figure 5.1(b). The
compressor raises the pressure and temperature of the incoming gas to the levels of p2
and T2. Expansion of the gas through the turbine back to the lower pressure at this
point would be useless, because all the work produced in the expansion would be
required to drive the compressor.
171

Instead, it is necessary to add heat and thus raise the temperature of the gas before
expanding it in the turbine. This is achieved in the heater by heat transfer from an
external source that raises the gas temperature to T3, the turbine inlet temperature.
Expansion of the hot gas through the turbine then delivers work in excess of that
needed to drive the compressor. The turbine work exceeds the compressor requirement
because the enthalpy differences, and hence the temperature differences, along
isentropes connecting lines of constant pressure increase with increasing entropy (and
temperature), as the figure suggests.
The difference between the turbine work, Wt, and the magnitude of the compressor
work, |Wc|, is the net work of the cycle. The net work delivered at the output shaft may
be used to drive an electric generator, to power a process compressor, turn an airplane
propeller, or to provide mechanical power for some other useful activity.
In the closed-cycle gas turbine, the heater is a furnace in which combustion gases
or a nuclear source transfer heat to the working fluid through thermally conducting
tubes. It is sometimes useful to distinguish between internal and external combustion
engines by whether combustion occurs in the working fluid or in an area separate from
the working fluid, but in thermal contact with it. The combustion-heated, closed-cycle
gas turbine is an example, like the steam power plant, of an external combustion
engine. This is in contrast to internal combustion engines, such as automotive engines,
in which combustion takes place in the working fluid confined between a cylinder and a
piston, and in open-cycle gas turbines.

5.3 Analysis of the Ideal Cycle

The Air Standard cycle analysis is used here to review analytical techniques and to
provide quantitative insights into the performance of an ideal-cycle engine. Air
Standard cycle analysis treats the working fluid as a calorically perfect gas, that is, a
perfect gas with constant specific heats evaluated at room temperature. In Air Standard
cycle analysis the heat capacities used are those for air.
A gas turbine cycle is usually defined in terms of the compressor inlet pressure and
temperature, p1 and T1, the compressor pressure ratio, r = p2/p1, and the turbine inlet
temperature, T3, where the subscripts correspond to states identified in Figure 5.1.
Starting with the compressor, its exit pressure is determined as the product of p1 and
the compressor pressure ratio. The compressor exit temperature may then be deter-
mined by the familiar relation for an isentropic process in an ideal gas, Equation (1.19):

T2 = T1( p2 /p1)(k–1)/k [R | K] (5.1)

For the two isobaric processes, p2 = p3 and p4 = p1. Thus the turbine pressure ratio,
p3/p4, is equal to the compressor pressure ratio, r = p2 /p1. With the turbine inlet
temperature T3 known, the turbine discharge temperature can be determined from

T4 = T3/( p2/p1)(k–1)/k [R | K] (5.2)


172

and the temperatures and pressures are then known at all the significant states.
Next, taking a control volume around the compressor, we determine the shaft
work required by the compressor, wc, assuming negligible heat losses, by applying the
steady-flow energy equation:

0 = h2 – h1 + wc

or

wc = h1 – h2 = cp( T1 – T2) [Btu/lbm | kJ/kg] (5.3)

Similarly, for the turbine, the turbine work produced is

wt = h3 - h4 = cp ( T3 – T4) [Btu/lbm | kJ/kg] (5.4)

The net work is then

wn = wt + wc = cp ( T3 – T4 + T1 – T2) [Btu/lbm | kJ/kg] (5.5)

Now taking the control volume about the heater, we find that the heat addition per
unit mass is

qa = h3 – h2 = cp ( T3 – T2) [Btu/lbm | kJ/kg] (5.6)

The cycle thermal efficiency is the ratio of the net work to the heat supplied to the
heater:

th = wn /qa [dl] (5.7)

which by substitution of Equations (5.1), (5.2), (5.5), and (5.6) may be simplified to

th = 1 – ( p2/p1) – (k–1)/k [dl] (5.8)

It is evident from Equation (5.8) that increasing the compressor pressure ratio increases
thermal efficiency.
Another parameter of great importance to the gas turbine is the work ratio, wt /|wc|.
This parameter should be as large as possible, because a large amount of the power
delivered by the turbine is required to drive the compressor, and because the engine net
work depends on the excess of the turbine work over the compressor work. A little
algebra will show that the work ratio wt /|wc| can be written as:

wt /|wc| = (T3 /T1) / ( p2 /p1) (k–1)/k [dl] (5.9)


173

Note that, for the ideal cycle, the thermal efficiency and the work ratio depend on only
two independent parameters, the compressor pressure ratio and the ratio of the turbine
and compressor inlet temperatures. It will be seen that these two design parameters are
of utmost importance for all gas turbine engines.
Equation (5.9) shows that the work ratio increases in direct proportion to the ratio
T3 /T1 and inversely with a power of the pressure ratio. On the other hand, Equation
(5.8) shows that thermal efficiency increases with increased pressure ratio. Thus, the
desirability of high turbine inlet temperature and the necessity of a tradeoff involving
pressure ratio is clear. Equation (5.9) also suggests that increases in the ratio T3 /T1
allow the compressor pressure ratio to be increased without reducing the work ratio.
This is indicative of the historic trend by which advances in materials allow higher
turbine inlet temperatures and therefore higher compressor pressure ratios.
It was shown in Chapter 1 that the area of a reversible cycle plotted on a T–s
diagram gives the net work of the cycle. With this in mind, it is interesting to consider a
family of cycles in which the compressor inlet state, a, and turbine inlet temperatures
are fixed, as shown in Figure 5.2. As the compressor pressure ratio pb/pa approaches 1,
the cycle area and hence the net work approach 0, as suggested by the shaded cycle
labeled with single primes. At the other extreme, as the compressor pressure ratio
approaches its maximum value, the net work also approaches 0, as in the cycle denoted
by double primes. For intermediate pressure ratios, the net work is large and positive,
indicating that there is a unique value of compressor pressure ratio that maximizes the
net work. Such information is of great significance in gas turbine design,
174

because it indicates the pressure ratio that yields the highest power output for given
turbomachine inlet temperatures and mass flow rate. This is an important approach to
the pressure ratio tradeoff mentioned earlier. It will be considered from an analytic
viewpoint for a more realistic gas turbine model in a later section.
Up to this point the discussion has focused on the closed-cycle gas turbine, an
external combustion or nuclear-heated machine that operates with a circulating fixed
mass of working fluid in a true cyclic process. In fact, the same Air Standard cycle
analysis may be applied to the open-cycle gas turbine. The open cycle operates with
atmospheric air that is pressurized by the compressor and then flows into a combustion
chamber, where it oxidizes a hydrocarbon fuel to produce a hot gas that drives the
turbine. The turbine delivers work as in the closed cycle, but the exiting combustion
gases pass into the atmosphere, as they must in all combustion processes.
A diagram of the cycle implementation is shown in Figure 5.3. Clearly, the open-
cycle gas turbine is an internal combustion engine, like the automotive engine. Note
that the diagram is consistent with Figure 5.1 and all the preceding equations in this
chapter. This is true because (1) the atmosphere serves as an almost infinite source and
sink that may be thought of as closing the cycle, and (2) the energy released by
combustion has the same effect as the addition of external heat in raising the
temperature of the gas to the turbine inlet temperature. A cutaway of an open-cycle
utility gas turbine is presented in Figure 5.4.
175

5.4 Realistic Simple-Cycle Gas Turbine Analysis

The preceding analysis of the Air Standard cycle assumes perfect turbomachinery, an
unachievable but meaningful ideal, and room-temperature heat capacities. Realistic
quantitative performance information can be obtained by taking into account
efficiencies of the compressor and the turbine, significant pressure losses, and more
realistic thermal properties.

Properties for Gas Turbine Analysis

It is pointed out in reference 1 that accurate gas turbine analyses may be performed
using constant heat capacities for both air and combustion gases. This appears to be a
specialization of a method devised by Whittle (ref. 4). The following properties are
therefore adopted for all gas turbine analyses in this book:
Air:
cp = 0.24 Btu/lbm-R or 1.004 kJ /kg-K

k = 1.4 implies k /(k – 1) = 3.5

Combustion gas:
cp,g = 0.2744 Btu/lbm-R or 1.148 kJ /kg-K

kg = 1.333 implies kg/(kg – 1) = 4.0

The properties labeled as combustion gas above are actually high-temperature-air


properties. Because of the high air-fuel ratio required by gas turbines, the
176

thermodynamic properties of gas turbine combustion gases usually differ little from
those of high-temperature air. Thus the results given below apply equally well to
closed-cycle machines using air as the working fluid and to open-cycle engines.

Analysis of the Open Simple-Cycle Gas Turbine

A simple-cycle gas turbine has one turbine driving one compressor and a
power-consuming load. More complex configurations are discussed later. It is assumed
that the compressor inlet state, the compressor pressure ratio, and the turbine inlet
temperature are known, as before. The turbine inlet temperature is usually determined
by the limitations of the high-temperature turbine blade material. Special metals or
ceramics are usually selected for their ability to withstand both high stress at elevated
temperature and erosion and corrosion caused by undesirable components of the fuel.
As shown in Figure 5.3, air enters the compressor at a state defined by T1 and p1.
The compressor exit pressure, p2, is given by

p2 = rp1 [lbf /ft2 | kPa] (5.10)

where r is the compressor pressure ratio. The ideal compressor discharge temperature,
T2s is given by the isentropic relation

T2s = T1 r (k–1)/k [R | K] (5.11)

The compressor isentropic efficiency, defined as the ratio of the compressor isentropic
work to the actual compressor work with both starting at the same initial state and
ending at the same pressure level, may be written as

c = ( h1 – h2s )/( h1 – h2 ) = ( T1 – T2s )/( T1 – T2 ) [dl] (5.12)

Here the steady-flow energy equation has been applied to obtain expressions for the
work for an irreversible adiabatic compressor in the denominator and for an isentropic
compressor in the numerator. Solving Equation (5.12) for T2, we get as the actual
compressor discharge temperature:

T2 = T1 + ( T2s – T1 ) /  c [R | K] (5.13)

Equation (5.3) then gives the work needed by the compressor, wc:

wc = cp ( T1 – T2 ) = cp ( T1 – T2s )/ c [Btu /lbm | kJ/kg] (5.14)

Note that the compressor work is negative, as required by the sign convention that
defines work as positive if it is produced by the control volume. The compressor power
requirement is, of course, then given by mawc [Btu/hr | kW], where ma is the
177

compressor mass flow rate [lbm / hr | kg / s].


After leaving the compressor at an elevated pressure and temperature, the air then
enters the combustion chamber, where it completely oxidizes a liquid or a gaseous fuel
injected under pressure. The combustion process raises the combustion gas temperature
to the turbine inlet temperature T3. One of the goals of combustion chamber design is
to minimize the pressure loss from the compressor to the turbine. Ideally, then, p3 = p2,
as assumed by the Air Standard analysis. More realistically, a fixed value of the
combustor fractional pressure loss, fpl, (perhaps about 0.05 or 5%) may be used to
account for burner losses:

fpl = (p2 – p3)/p2 [dl] (5.15)

Then the turbine inlet pressure may be determined from

p3 = (1 – fpl) p2 [lbf /ft2 | kPa] (5.16)

Rather than deal with its complexities, we may view the combustion process simply as
one in which heat released by exothermic chemical reaction raises the temperature of
combustion gas (with hot-air properties) to the turbine inlet temperature. The rate of
heat released by the combustion process may then be expressed as:

Qa = ma(1 + f )cp,g(T3 – T2) [Btu/hr | kW] (5.17)

where f is the mass fuel-air ratio. The term ma(1 + f) is seen to be the sum of the air and
fuel mass flow rates, which also equals the mass flow rate of combustion gas. For gas
turbines it will be seen later that f is usually much less than the stoichiometric fuel-air
ratio and is often neglected with respect to 1 in preliminary analyses.
The turbine in the open-cycle engine operates between the pressure p3 and
atmospheric pressure, p4 = p1, with an inlet temperature of T3. If the turbine were
isentropic, the discharge temperature would be

T4s = T3( p4 /p3 ) (kg– 1) / kg [R | K] (5.18)

From the steady-flow energy equation, the turbine work can be written as

wt = cp,g (T3 – T4 ) = t cp,g( T3 – T4s) [Btu/lbm | kJ/kg] (5.19)

referenced to unit mass of combustion gas, and where t is the turbine isentropic
efficiency. The turbine power output is then ma(1 + f)wt, where, as seen earlier, ma(1 +
f) is the mass flow rate of combustion gas flowing through the turbine. The net work
based on the mass of air processed and the net power output of the gas turbine, Pn, are
then given by
178

wn = (1 + f )wt + wc [Btu/lbm air| kJ/kg air] (5.20)

and

Pn = ma [(1 + f )wt + wc ] [Btu/hr | kW] (5.21)

and the thermal efficiency of the engine is

th = Pn /Qa [dl] (5.22)

EXAMPLE 5.1

A simple-cycle gas turbine has 86% and 89% compressor and turbine efficiencies,
respectively, a compressor pressure ratio of 6, a 4% fractional pressure drop in the
combustor, and a turbine inlet temperature of 1400°F. Ambient conditions are 60°F and
one atmosphere. Determine the net work, thermal efficiency, and work ratio for the
engine. Assume that the fuel-mass flow rate is negligible compared with the air flow
rate.

Solution
The notation for the solution is that of Figure 5.3. The solution details are given in
Table 5.1 in a step-by-step spreadsheet format. Each line presents the parameter name,
symbol, and units of measure; its value; and the right-hand side of its specific
determining equation.
179

When an entire cycle is to be analyzed, it is best to start at the compressor with the
inlet conditions and proceed to calculate successive data in the clockwise direction on
the T-s diagram. The compressor isentropic and actual discharge temperatures and
work are determined first using Equations (5.11), (5.13), and (5.14). The turbine
pressure ratio is determined next, accounting for the combustor pressure loss, using
Equation (5.16). The isentropic relation, Equation (5.18), gives the isentropic turbine
exit temperature, and the turbine efficiency and Equation (5.19) yields the true turbine
exit temperature and work. Once all the turbomachine inlet and exit temperatures are
known, other cycle parameters are easily determined, such as the combustor heat
transfer, net work, thermal efficiency, and work ratio.
_____________________________________________________________________

An important observation may be made on the basis of this analysis regarding the
magnitude of the compressor work with respect to the turbine work. Much of the
turbine work is required to drive the compressor. Compare the work ratio of 1.66, for
example, with the much higher values for the steam cycles of Chapter 2 (the Rankine-
cycle pumps have the same function there as the compressor here). Example 2.4 for the
Rankine cycle with a 90% turbine efficiency has a work ratio of 77.2. Thus the gas
turbine’s pressurization handicap relative to the Rankine cycle is substantial.
The unimpressive value of the thermal efficiency of the example gas turbine, 25%
(not typical of the current state of the art) compares with a Carnot efficiency for the
same cycle temperature extremes of 72% The large amount of compressor work requi-
red clearly contributes to this weak performance. Nevertheless, current gas turbines are
competitive with many other engines on an efficiency basis, and have advantages such
as compactness and quick-start capability relative to Rankine cycle power plants. One
approach to the improvement of thermal efficiency of the gas turbine will be addressed
later in Section 5.5. First let’s look at what can be done about gas turbine work.

Maximizing the Net Work of the Cycle

Using Equations (5.14) and (5.19), we can rewrite the cycle net work as

wn = t cp,g ( T3 – T4s) – cp( T2s – T1 )/c [Btu/lbm | kJ/kg] (5.23)

where the fuel-air ratio has been neglected with respect to 1. In the following, the
combustor pressure losses and the distinction between hot-gas and air heat capacities
will be neglected but the very important turbomachine efficiencies are retained.
Nondimensionalizing the net work with the constant cpT1 we get:

wn/cpT1 = t (T3 /T1)(cp,g/cp)(1 – r – (k–1)/k) + (1 – r (k–1)/k)/c [dl] (5.24)

By differentiating wn with respect to the compressor pressure ratio r and setting the
180

result equal to 0, we obtain an equation for r*, the value of r that maximizes the net
work with fixed turbomachine efficiencies and with a constant ratio of the temperatures
of the turbomachine inlets, T3 /T1. For constant gas properties throughout, the result is

r* = (c t T3/T1)k/2(k–1) [dl] (5.25)

This relation gives a specific value for the compressor pressure ratio that defines an
optimum cycle, in the sense of the discussion of Figure 5.2. There it was established
qualitatively that a cycle with maximum net work exists for a given value of T3/T1.
Equation (5.25) defines the condition for this maximum and generalizes it to include
turbomachine inefficiency.
The pressure ratio r* given by Equation (5.25) increases with increasing
turbomachine efficiencies and with T3/T1. This is a clear indicator that increasing
turbine inlet temperature favors designs with higher compressor pressure ratios. This
information is important to the gas turbine designer but does not tell the whole design
story. There are other important considerations; for example, (1) compressors and
turbines become more expensive with increasing pressure ratios, and (2) the pressure
ratio that maximizes thermal efficiency is different from that given by Equation (5.25).
Figure 5.5 shows the influence of compressor pressure ratio on both efficiency and net
work and the position of the value given by Equation (5.25). Thus, when all factors are
taken into account, the final design pressure ratio is likely to be in the vicinity of, but
not necessarily identical to, r*.
181

5.5 Regenerative Gas Turbines

It was shown in Chapter 2 that the efficiency of the Rankine cycle could be improved
by an internal transfer of heat that reduces the magnitude of external heat addition, a
feature known as regeneration. It was also seen in Chapter 2 that this is accomplished
conveniently in a steam power plant by using a heat exchanger known as a feedwater
heater.
Examination of Example 5.1 shows that a similar opportunity exists for the gas
turbine cycle. The results show that the combustion process heats the incoming air from
924°R to 1860°R and that the gas turbine exhausts to the atmosphere at 1273°R. Thus
a maximum temperature potential of 1273 – 924 = 349°F exists for heat transfer. As in
the Rankine cycle, this potential for regeneration can be exploited by incorporation of a
heat exchanger. Figure 5.6 shows a gas turbine with a counterflow heat exchanger that
extracts heat from the turbine exhaust gas to preheat the compressor discharge air to Tc
ahead of the combustor. As a result, the temperature rise in the combustor is reduced to
T3 – Tc, a reduction reflected in a direct decrease in fuel consumed.
Note that the compressor and turbine inlet and exit states can be the same as for a
simple cycle. In this case the compressor, turbine, and net work as well as the work
ratio are unchanged by incorporating a heat exchanger.
The effectiveness of the heat exchanger, or regenerator, is a measure of how well
it uses the available temperature potential to raise the temperature of the compressor
discharge air. Specifically, it is the actual rate of heat transferred to the air divided by
the maximum possible heat transfer rate that would exist if the heat exchanger had
infinite heat transfer surface area. The actual heat transfer rate to the air is mcp(Tc – T2),
and the maximum possible rate is mcp(T4 – T2). Thus the regenerator effectiveness can
be written as

reg = ( Tc – T2 )/( T4 – T2 ) [dl] (5.26)

and the combustor inlet temperature can be written as

Tc = T2 + reg( T4 – T2 ) [R | K] (5.27)

It is seen that the combustor inlet temperature varies from T2 to T4 as the


regenerator effectiveness varies from 0 to 1. The regenerator effectiveness increases as
its heat transfer area increases. Increased heat transfer area allows the cold fluid to
absorb more heat from the hot fluid and therefore leave the exchanger with a higher Tc.
On the other hand, increased heat transfer area implies increased pressure losses
on both air and gas sides of the heat exchanger, which in turn reduces the turbine
pressure ratio and therefore the turbine work. Thus, increased regenerator effectiveness
implies a tradeoff, not only with pressure losses but with increased heat exchanger size
and complexity and, therefore, increased cost.
182

The exhaust gas temperature at the exit of the heat exchanger may be determined
by applying the steady-flow energy equation to the regenerator. Assuming that the heat
exchanger is adiabatic and that the mass flow of fuel is negligible compared with the air
flow, and noting that no shaft work is involved, we may write the steady-flow energy
equation for two inlets and two exits as

q = 0 = he + hc – h2 – h4 + w = cp,gTe + cpTc – cpT2 – cp,gT4 + 0

Thus the regenerator combustion-gas-side exit temperature is:

Te = T4 – (cp/cp,g)( Tc – T2 ) [R | K] (5.28)

While the regenerator effectiveness does not appear explicitly in Equation (5.28),
the engine exhaust temperature is reduced in proportion to the air temperature rise in
the regenerator, which is in turn proportional to the effectiveness. The dependence of
the exhaust temperature on reg may be seen directly by eliminating Tc from Equation
(5.28), using Equation (5.27) to obtain

T4 – Te = reg (cp/cp,g)(T4 – T2) [R | K] (5.29)


183

The regenerator exhaust gas temperature reduction, T4 – Te, is seen to be jointly


proportional to the effectiveness and to the maximum temperature potential, T4 – T2.
The regenerator, like other heat exchangers, is designed to have minimal pressure
losses on both air and gas sides. These may be taken into account by the fractional
pressure drop approach discussed in connection with the combustor.

EXAMPLE 5.2

Let’s say we are adding a heat exchanger with an effectiveness of 75% to the engine
studied in Example 5.1. Assume that the same frictional pressure loss factor applies to
both the heat exchanger air-side and combustor as a unit, and that gas-side pressure
loss in the heat exchanger is negligible. Evaluate the performance of the modified
engine.

Solution
The solution in spreadsheet format, expressed in terms of the notation of Figures
5.6 and 5.7, is shown in Table 5.2. Examination of the spreadsheet and of the T-s
diagram in Figure 5.7 shows that the entry and exit states of the turbomachines are not
influenced by the addition of the heat exchanger, as expected. (There would have been
a slight influence if a different pressure loss model had been assumed.)
With the heat exchanger, it is seen that the combustor inlet temperature has
increased about 262° and the exhaust temperature reduced 229°. The net work and
184

work ratio are clearly unchanged. Most importantly, however, the thermal efficiency
has increased 10 percentage counts over the simple cycle case in Example 5.1. Such a
gain must be traded off against the added volume, weight, and expense of the
regenerator. The efficiency gain and the associated penalties may be acceptable in
stationary power and ground and marine transportation applications, but are seldom
feasible in aerospace applications. Each case, of course, must be judged on its own
merits.
_____________________________________________________________________

Figure 5.8 shows the influence of regenerator effectiveness and turbine inlet
temperature on the performance of the gas turbine, all other conditions being the same
as in the example. The values for reg = 0 correspond to a gas turbine without
regenerator. The abscissa is arbitrarily truncated at reg = 0.8 because gas turbine heat
exchanger effectivenesses usually do not exceed that value. The impressive influence of
both design parameters is a strong motivator for research in heat exchangers and high-
temperature materials. The use of regeneration in automotive gas turbines is virtually
mandated because good fuel economy is so important.
185

Figure 5.9 shows the layout of a regenerative gas turbine serving a pipeline
compressor station. Gas drawn from the pipeline may be used to provide the fuel for
remotely located gas-turbine-powered compressor stations. (A later figure, Figure 5.12,
shows details of the turbomachinery of this gas turbine.)
186

5.6 Two-Shaft Gas Turbines

Problems in the design of turbomachinery for gas turbines and in poor part-load or
off-design performance are sometimes avoided by employing a two-shaft gas turbine, in
which the compressor is driven by one turbine and the load by a second turbine. Both
shafts may be contained in a single structure, or the turbines may be separately
packaged. Figure 5.10 shows the flow and T-s diagrams for such a configuration. The
turbine that drives the compressor is called the compressor turbine. The compressor,
combustor, compressor-turbine combination is called the gas generator, or gasifier,
because its function is to provide hot, high-pressure gas to drive the second turbine, the
power turbine. The compressor-turbine is sometimes also referred to as the gasifier
turbine or gas-generator turbine.
The analysis of the two-shaft gas turbine is similar to that of the single shaft
machine, except in the determination of the turbine pressure ratios. The pressure rise
produced by the compressor must be shared between the two turbines. The manner in
which it is shared is determined by a power, or work, condition. The work condition
expresses mathematically the fact that the work produced by the gasifier turbine is used
to drive the compressor alone. As a result, the gas generator turbine pressure ratio,
p3/p4, is just high enough to satisfy the compressor work requirement.
Thus the compressor power (work) input is the same as the delivered gas-
generator turbine power(work) output:

|wc| = mech (1 + f )wt [Btu /lbm | kJ/kg] (5.30)

where f is the fuel–air ratio and mech is the mechanical efficiency of transmission of
power from the turbine to the compressor. The mechanical efficiency is usually close to
unity in a well-designed gas turbine. For this reason, it was not included in earlier
analyses.
The gasifier turbine work may be written in terms of the turbine pressure ratio:

wt = t cp,gT3( 1 – T4s /T3 )

= t cp,gT3[1 – 1/( p3/p4)(kg-1)/kg] [Btu /lbm | kJ/kg] (5.31)

With the compressor work determined, as before, by the compressor pressure ratio and
the isentropic efficiency, the compressor-turbine pressure ratio, p3/p4, is obtained by
combining Equations (5.30) and (5.31):

p3/p4 = [ 1 – |wc| /mecht cp,g (1 + f )T3]–kg/(kg-1)

= ( 1 – wf )–kg/(kg-1) [dl] (5.32)

where wf is the positive, dimensionless work factor, |wc| /mecht cp,g (1 + f )T3, used as
187

a convenient intermediate variable. The power turbine pressure ratio may then be
determined from the identity p4/p5 = p4/p1 = ( p4/p3)( p3/p2)( p2/p1). This shows that the
power turbine pressure ratio is the compressor pressure ratio divided by the gasifier
turbine pressure ratio when there is no combustion chamber pressure loss ( p3 = p2).
With the pressure ratios known, all the significant temperatures and performance
parameters may be determined.

EXAMPLE 5.3

Let’s consider a two-shaft gas turbine with a regenerative air heater. The compressor
pressure ratio is 6, and the compressor and gas generator turbine inlet temperatures are
520°R and 1860°R, respectively. The compressor, gasifier turbine, and power turbine
isentropic efficiencies are 0.86, 0.89, and 0.89, respectively. The regenerator
effectiveness is 75%, and a 4% pressure loss is shared by the high-pressure air side of
the regenerator and the combustor. Determine the pressure ratios of the two turbines,
and the net work, thermal efficiency, and work ratio of the engine.
188

Solution
The solution in spreadsheet form shown in Table 5.3 follows the notation of Figure
5.10. The solution proceeds as before, until the calculation of the turbine pressure
ratios. The available pressure ratio shared by the two turbines is p3/p5 = p3/p1 =
(p2/p1)(p3/p2) = r ( 1 – fpl) = 5.76. The gasifier turbine pressure ratio is determined by
the work-matching requirement of the compressor and its driving turbine, as expressed
in Equation (5.32), using the dimensionless compressor work factor, wf. The resulting
gas generator and power turbine pressure ratios are 2.61 and 2.2, respectively.
Comparison shows that the design point performance of the two-shaft gas turbine
studied here is not significantly different from that of the single-shaft machine
considered an Example 5.2. While the performance of the two machines is found to be
essentially the same, the single-shaft machine is sometimes preferred in applications
189

with fixed operating conditions where good part-load performance over a range of
speeds is not important. On the other hand, the independence of the speeds of the gas
generator and power turbine in the two-shaft engine allows acceptable performance
over a wider range of operating conditions.
_____________________________________________________________________
Let us examine further the characteristics of regenerative two-shaft gas turbines,
starting with the spreadsheet reproduced in Table 5.3. By copying the value column of
that spreadsheet to several columns to the right, a family of calculations with identical
methodologies may be performed.. The spreadsheet /EDIT-FILL command may then
be used to vary a parameter in a given row by creating a sequence of numbers with a
specified starting value and interval. Such a parametric study of the influence of
compressor pressure ratio on two-shaft regenerative gas turbine performance is shown
in Table 5.4, where the pressure ratio is varied from 2 to 7. The fifth numeric column
contains the values from Table 5.3. The data of Table 5.4 are included in the Example
190

5-4.wk3 spreadsheet that accompanies this text.


Table 5.4 shows that, for the given turbine inlet temperature, the thermal efficiency
maximum is at a pressure ratio between 4 and 5, while the net work maximum is at a
pressure ratio of about 7. The work ratio is continually declining because the
magnitude of the compressor work requirement grows faster with compressor pressure
ratio than the turbine work does.
Figure 5.11, plotted using the spreadsheet, compares the performance of the
regenerative two-shaft gas turbine with a nonregenerative two-shaft engine (reg = 0 ).
Net work for both machines has the same variation with pressure ratio. But notice the
high efficiency attained with a low-compressor pressure ratio, a significant advantage
attributable to regeneration.
A cutaway view of a two-shaft regenerative gas turbine of the type used in pipeline
compressor stations such as that shown in Figure 5.9 is seen in Figure 5.12. The figure
shows that the compressor blade heights decrease in the direction of flow as the gas is
compressed. The exhaust from the last of the sixteen compressor stages is reduced in
velocity by a diffusing passage and then exits through the right window-like flange,
which connects to a duct (not shown) leading to the regenerator. The heated air from
the external regenerator reenters the machine combustor casing, where it flows around
and into the combustor cans, cooling them. The air entering near the combustor fuel
nozzles mixes with the fuel and burns locally in a near-stoichiometric mixture. As the
mixture flows downstream, additional secondary air entering the combustor through
slots in its sides mixes with, and reduces the temperature of, the combustion gas before
it arrives at the turbine inlet.
A cutaway view of an industrial two-shaft gas turbine with dual regenerators is
presented in Figure 5.13. From the left, the air inlet and radial compressor and axial
flow gasifier turbine and power turbine are seen on the axis of the machine, with the
combustion chamber above and one of the rotary regenerators at the right. Due to the
relatively low pressure ratios required by regenerative cycles, centrifugal compressors
are normally used in regenerative machines because of their simplicity, good efficiency,
compactness, and ruggedness.
Performance data for the turbine of Figure 5.13 is graphed in Figure 5.14. The GT
404 gas turbine delivers about 360 brake horsepower at 2880-rpm output shaft speed.
The torque-speed curve of Figure 5.14 shows an important characteristic of two-shaft
gas turbines with respect to off-design point operation. Whereas the compressor
pressure ratio and output torque of a single-shaft gas turbine drop as the shaft speed
drops, the compressor speed and pressure ratio in a two-shaft machine is independent
of the output speed. Thus, as the output shaft speed changes, the compressor may
maintain its design speed and continue to develop high pressure and mass flow. Thus
the torque at full stall of the output shaft of the GT404 is more than twice the full-load
design torque. This high stall torque is superior to that of reciprocating engines and is
important in starting and accelerating rotating equipment that has high initial turning
resistance. This kind of engine may be used in truck, bus, and marine applications as
well as in an industrial setting.
191
192
193

A unique patented feature of some of the Allison gas turbines, called “power
transfer,” is the ability to link the duel shafts. A hydraulic clutch mechanism between
the two turbine shafts acts to equalize their speeds. This tends to improve part-load
fuel economy, and provides engine braking and overspeed protection for the power
turbine. When the clutch mechanism is fully engaged, the shafts rotate together as a
single-shaft machine.

5.7 Intercooling and Reheat

Intercooling

It has been pointed out that the work of compression extracts a high toll on the output
of the gas turbine. The convergence of lines of constant pressure on a T-s diagram
indicates that compression at low temperatures reduces compression work. The ideal
compression process would occur isothermally at the lowest available temperature.
Isothermal compression is difficult to execute in practice. The use of multistage
compression with intercooling is a move in that direction.
Consider replacing the isentropic single-stage compression from p1 to p2 = p2* in
Figure 5.15 with two isentropic stages from p1 to pis and pi* to p2s. Separation of the
compression processes with a heat exchanger that cools the air at Tis to a lower
temperature Ti* acts to move the final compression process to the left on the T-s
194

diagram and reduces the discharge temperature following compression to T2s. A heat
exchanger used to cool compressed gas between stages of compression is called an
intercooler.
The work required to compress from p1 to p2s = p2* = p2 in two stages is

wc = cp [(T1 – Tis) + ( T1* – T2s)] [Btu/lbm | kJ/kg]

Note that intercooling increases the net work of the reversible cycle by the area
is–i*–2s–2*–is. The reduction in the work due to two-stage intercooled compression
is also given by this area. Thus intercooling may be used to reduce the work of
compression between two given pressures in any application. However, the favorable
effect on compressor work reduction due to intercooling in the gas turbine application
may be offset by the obvious increase in combustor heat addition, cp (T2* – T2s), and by
increased cost of compression system. The next example considers the selection of the
optimum pressure level for intercooling, pi = pis = pi*.

EXAMPLE 5.4

Express the compressor work, for two-stage compression with intercooling back to the
original inlet temperature, in terms of compressor efficiencies and pressure ratios.
Develop relations for the compressor pressure ratios that minimize the total work of
compression in terms of the overall pressure ratio.

Solution

Taking Ti* = T1 as directed in the problem statement, and letting r = p2/p1, r1 = pis/p1 and
r2 = p2s/pi* = r/r1 as in Figure 5.15, we get for the compression work,

wc = cp [(T1 – Tis)/c1 + ( T1 – T2s)/c2 ]

= cp T1[(1 – Tis /T1)/c1 + ( 1 – T2s / T1)/c2 ]

= cp T1[(1 – r1(k – 1 )/k )/c1 + ( 1 – r2(k – 1 )/k )/c2 ] [Btu /lbm | kJ/kg]

Eliminating r2, using r = r1r2, yields

wc = cp T1[(1 – r1(k – 1 )/k )/c1 + ( 1 – (r/r1) (k – 1 )/k )/c2 ]

Differentiating with respect to r1 for a fixed r and setting the result equal to zero, we
obtain

– r1– 1/k /c1 + (r (k – 1 )/k /c2) r1 – (2k – 1 )/k = 0

which simplifies to
195

r1opt = (c1/c2)k/(2k – 1) r1/2


Using this result we find also that

r2opt = (c2/c1)k/(2k – 1) r1/2

Examination of these equations shows that, for compressors with equal efficiencies,
both compressor stages have the same pressure ratio, which is given by the square root
of the overall pressure ratio. For unequal compressor efficiencies, the compressor with
the higher efficiency should have the higher pressure ratio.
_____________________________________________________________________

Reheat

Let us now consider an improvement at the high-temperature end of the cycle. Figure
5.16 shows the replacement of a single turbine by two turbines in series, each with
appropriately lower pressure ratios, and separated by a reheater. The reheater may be a
combustion chamber in which the excess oxygen in the combustion gas leaving the first
turbine burns additional fuel, or it may be a heater in which external combustion
provides the heat necessary to raise the temperature of the working fluid to Tm*. The
high temperature at the low-pressure turbine inlet has the effect of increasing the area
of the cycle by m–m*–4–4*m and hence of increasing the net work.
196

Like intercooling, the increase in net work is made possible by the spreading of the
constant pressure lines on the T-s diagram as entropy increases. Thus the increase in
turbine work is

wt = cp,g [(Tm* – T4) – ( Tm – T4*) ] [Btu/lbm | kJ/kg] (5.33)

Also as with intercooling, the favorable effect in increasing net work is offset by the
reduction of cycle efficiency resulting from increased addition of external heat from the
reheater:

qrh = cp,g (Tm* – Tm) [Btu/lbm | kJ/kg] (5.34)

As with intercooling, the question arises as to how the intermediate pressure for
reheat will be selected. An analysis similar to that of Example 5.4 shows the unsur-
prising result that the reheat pressure level should be selected so that both turbines have
the same expansion ratio if they have the same efficiencies and the same inlet
temperatures.

Combining Intercooling, Reheat, and Regeneration

Because of their unfavorable effects on thermal efficiency, intercooling and reheat alone
or in combination are unlikely to be found in a gas turbine without another feature that
has already been shown to have a favorable influence on gas turbine fuel economy: a re-
generator. The recuperator or regenerator turns disadvantage into advantage in a cycle
involving intercooling and/or reheat. Consider the cycle of Figure 5.17, which
incorporates all three features.
The increased turbine discharge temperature T4 produced by reheat and the
decreased compressor exit temperature T2 due to intercooling both provide an enlarged
temperature potential for regenerative heat transfer. Thus the heat transfer cp (Tc – T2)
is accomplished by an internal transfer of heat from low pressure turbine exhaust gas.
This also has the favorable effect of reducing the temperature of the gas discharged to
the atmosphere. The requisite external heat addition for this engine is then

qa = cp [(T3 – Tc) + ( Tm* – Tm)] [Btu/lbm | kJ/kg] (5.35)

Thus the combination of intercooling, reheat, and regeneration has the net effect of
raising the average temperature of heat addition and lowering the average temperature
of heat rejection, as prescribed by Carnot for an efficient heat engine.
197

The Ericsson cycle

Increasing the number of intercoolers and reheaters without changing the overall pres-
sure ratio may be seen to cause both the overall compression and the overall expansion
to approach isothermal processes. The resulting reversible limiting cycle, consisting of
two isotherms and two isobars, is called the Ericsson cycle. With perfect internal heat
198

transfer between isobaric processes, all external heat addition would be at the maximum
temperature of the cycle and all heat rejected at the lowest temperature. Analysis of the
limiting reversible cycle reveals, as one might expect, that its efficiency is that of the
Carnot cycle. Plants with multistage compression, reheat, and regeneration can have
high efficiencies; but complexity and high capital costs have resulted in few plants that
actually incorporate all these features.

5.8 Gas Turbines in Aircraft –Jet engines

Gas turbines are used in aircraft to produce shaft power and hot, high-pressure gas for
jet propulsion. Turbine shaft power is used in turboprop aircraft and helicopters to
drive propellers and rotors. A modern turboprop engine and an aircraft that uses it are
shown in Figures 5.18 through 5.20. While its jet exhaust provides some thrust, the
bulk of the propulsive thrust of the turboprop is provided by its propeller. The rear-
199

ward acceleration of a large air mass by the propeller is responsible for the good fuel
economy of turboprop aircraft. Thus the turboprop is popular as a power plant for
small business aircraft. At higher subsonic flight speeds, the conventional propeller
loses efficiency and the turbojet becomes superior.
Auxiliary power units, APUs, are compact gas turbines that provide mechanical
power to generate electricity in transport aircraft while on the ground. The thermo-
dynamic fundamentals of these shaft-power devices are the same those of stationary gas
turbines, discussed earlier. Their design, however, places a premium on low weight
and volume and conformance to other constraints associated with airborne equipment.
Thus their configuration and appearance may differ substantially from those of other
stationary gas turbines.
The jet engine consists of a gas turbine that produces hot, high-pressure gas but
has zero net shaft output. It is a gasifier. A nozzle converts the thermal energy of the
hot, high-pressure gas produced by the turbine into a high-kinetic-energy exhaust
stream. The high momentum and high exit pressure of the exhaust stream result in a
forward thrust on the engine.
Although the analysis of the jet engine is similar to that of the gas turbine, the
configuration and design of jet engines differ significantly from those of most
stationary gas turbines. The criteria of light weight and small volume, mentioned
earlier, apply here as well. To this we can add the necessity of small frontal area to
minimize the aerodynamic drag of the engine, the importance of admitting air into the
200

engine as efficiently (with as little stagnation pressure loss) as possible, and the efficient
conversion of high-temperature turbine exit gas to a high-velocity nozzle exhaust. The
resulting configuration is shown schematically in Figure 5.21.
Up to now we have not been concerned with kinetic energy in the flows in gas
turbines, because the flows at the stations of interest are usually designed to have low
velocities. In the jet engine, however, high kinetic energy is present in the free stream
ahead of the engine and in the nozzle exit flow. The analysis here will therefore be
presented in terms of stagnation, or total, temperatures and pressures, where kinetic
201

energy is taken into account implicitly, as discussed in Section 1.7. The preceding
analyses may be readily adapted to deal with the stagnation properties associated with
compressible flow. In the following discussion, engine processes are first described and
then analyzed.
It should be recalled that if there are no losses, as in an isentropic flow, the
stagnation pressure of a flow remains constant. All loss mechanisms, such as fluid
friction, turbulence, and flow separation, decrease stagnation pressure. Only by doing
work on the flow (with a compressor, for example) is it possible to increase stagnation
pressure.
In Figure 5.21, free-stream ambient air, denoted by subscript a, enters an engine
inlet that is carefully designed to efficiently decelerate the air captured by its frontal
area to a speed low enough to enter the compressor, at station 1, with minimal
aerodynamic loss. There is stagnation pressure loss in the inlet, but efficient
deceleration of the flow produces static and stagnation pressures at the compressor
entrance well above the ambient free-stream static pressure. This conversion of relative
kinetic energy of ambient air to increased pressure and temperature in the engine inlet is
sometimes called ram effect.
The compressor raises the stagnation pressure of the air further to its maximum
value at station 2, using power delivered by the turbine. Fuel enters the combustion
chamber and is burned with much excess air to produce the high turbine inlet
temperature at station 3. We adopt here, for simplicity, the familiar idealization that no
pressure losses occur in the combustion chamber. The hot gases then expand through
the turbine and deliver just enough power to drive the compressor (the work condition
again). The gases leave the turbine exit at station 4, still hot and at a stagnation pres-
sure well above the ambient. These gases then expand through a nozzle that converts
the excess pressure and thermal energy into a high-kinetic-energy jet at station 5. The
forward thrust on the engine, according to Newton’s Second Law, is produced by the
reaction to the internal forces that accelerate the internal flow rearward to a high jet
velocity and the excess of the nozzle exit plane pressure over the upstream ambient
pressure.

Inlet Analysis

Given the flight speed, Va , and the free-stream static temperature and pressure, Ta and
pa, at a given altitude, the free-stream stagnation temperature and pressure are

Toa = Ta + Va2 / 2cp [R | K] (5.36)

and

poa = pa( Toa/Ta ) k/(k – 1) [lbf /ft2 | kPa] (5.37)

Applying the steady-flow energy equation to the streamtube entering the inlet, we find
202

that the stagnation enthalpy hoa = ho1 for adiabatic flow. For subsonic flight and
supersonic flight at Mach numbers near one, the heat capacity of the air is essentially
constant. Thus constancy of the stagnation enthalpy implies constancy of the stagnation
temperature. Hence, using Equation (5.36),

To1 = Toa = Ta + Va2 / 2cp [R | K] (5.38)

The effects of friction, turbulence, and other irreversibilities in the inlet flow are
represented by the inlet pressure recovery, PR, defined as

PR = po1 / poa [dl] (5.39)

where an isentropic flow through the inlet has a pressure recovery of 1.0. Lower values
indicate reduced inlet efficiency and greater losses. For subsonic flow, values on the
order of 0.9 to 0.98 are typical. At supersonic speeds the pressure recovery decreases
with increasing Mach number.

Compressor Analysis

With the stagnation conditions known at station 1 in Figure 5.21, the compressor
pressure ratio, r = po2 /po1, now yields po2; and the isentropic relation, Equation (1.19),
gives the isentropic temperaure, To2s:

To2s = To1r(k – 1)/k [R | K] (5.40)

The actual compressor discharge stagnation temperature is then obtained from the
definition of the compressor efficiency in terms of stagnation temperatures:

comp = ( To1 – To2s ) / ( To1 – To2 ) [dl] (5.41)

Combustor and Turbine Analysis

The turbine inlet temperature, T03, is usually assigned based on turbine blade material
considerations. For preliminary analysis it may be assumed that there are negligible
pressure losses in the combustion chamber, so that p03 = p02. As with the two-shaft gas
turbine, the condition that the power absorbed by the compressor equal the power
delivered by the turbine determines the turbine exit temperature, T04:

cp(To2 – To1 ) = (1 + f )cp,gTo3 ( 1 – To4 / To3 ) [Btu | kJ] (5.42)

where f is the engine fuel-air ratio, which often may be neglected with respect to 1 (as
in our earlier studies) when high precision is not required.
203

The turbine efficiency equation then yields the isentropic discharge temperature
T04s, and Equation (1.19) yields the turbine pressure ratio:

To4s = To3 – ( To3 – To4) /turb [R | K] (5.43)

po3 /po4 = ( To3/To4s)kg/(kg – 1) [dl] (5.44)

Thus the stagnation pressure and temperature at station 4 are known. Note that the
turbine pressure ratio is usually significantly lower than the compressor pressure ratio.

Nozzle Analysis

The flow is then accelerated to the jet velocity at station 5 by a convergent nozzle that
contracts the flow area. A well-designed nozzle operating at its design condition has
only small stagnation pressure losses. Hence the nozzle here is assumed to be loss-free
and therefore isentropic.
Under most flight conditions the exhaust nozzle is choked; that is, it is passing the
maximum flow possible for its upstream conditions. A choked nozzle has the local
flow velocity at its minimum area, or throat, equal to the local speed of sound. As a
result, simple relations exist between the upstream stagnation conditions at station 4
and the choked conditions at the throat. Thus, for a choked isentropic nozzle,

To4 = To5 = T5 + V5 2 / 2cp,g

= T5 + kgRT5/2cp,g = T5 ( 1 + kgR/2cp,g)

= T5( kg + 1 )/2 [R | K] (5.45)

where kgR/cp,g = kg – 1. With kg = 1.333 for the combustion gas, this determines the exit
temperature T5. Combining the isentropic relation with Equation (5.45) then gives the
nozzle exit static pressure p5:

p5 /po4 = (T5 /T04)kg/(kg – 1)

= [2/(kg + 1)]kg/(kg – 1) [dl] (5.46)

EXAMPLE 5.5

The stagnation temperature and pressure leaving a turbine and entering a convergent
nozzle are 970.2K. and 2.226 bar, respectively. What is the static pressure and
temperature downstream if the nozzle is choked? If the free-stream ambient pressure is
0.54 bar, is the nozzle flow choked? Compare the existing nozzle pressure ratio with
the critical pressure ratio.
204

Solution
If the nozzle is choked, then from Equations (5.45) and (5.46),

T5 = 2To4/( kg + 1) = 2(970.2)/2.333 = 831.6 K

and the static pressure at the nozzle throat is

p5 = po4 [2/(kg + 1)]kg/(kg – 1) = 2.226( 2 / 2.333 )4 = 1.202 bar

The fact that p5 > pa = 0.54 indicates that the nozzle is choked.
The critical pressure ratio of the nozzle is

po4 /p5 = [(kg + 1)/2]kg/(kg – 1) = (2.333/2)4 = 1.852

and the applied pressure ratio is 2.226/0.54 = 4.12. Thus the applied pressure ratio
exceeds the critical pressure ratio.This also indicates that the nozzle is choked.
_____________________________________________________________________

For the isentropic nozzle, the steady-flow energy equation gives

0 = h5 + V52/2 – ho4

or, with cp,g constant,

V5 = [2cp,g(To4 – T5)]½ [ft/s | m/s] (5.47)

Thus the jet velocity is determined from Equation (5.47), where T5 is obtained from
Equation (5.45).
The thrust of the engine is obtained by applying Newton’s Second Law to a
control volume, as shown in Figure 5.22. If the mass flow rate through the engine is m,
the rates of momentum flow into and out of the control volume are mVa and mV5,
respectively. The net force exerted by the exit pressure is (p5 – pa)A5, where A5 is the
nozzle exit area. Thus, applying Newton’s Second Law to the control volume, we can
relate the force exerted by the engine on the gases flowing through, F, and the net exit
pressure force to the rate of increase of flow momentum produced by the engine:

m(V5 – Va) = F – (p5 – pa)A5 [lbf | kN]

or

F = m(V5 – Va) + ( p5 – pa)A5 [lbf | kN] (5.48)

Here, F is the engine force acting on the gas throughflow. The reaction to this force is
the thrust on the engine acting in the direction of flight. Thus the magnitude of the
205

thrust is given by Equation (5.48). It is the sum of all the pressure force components
acting on the inside the engine in the direction of flight.
The exit area, A5, is related to the mass flow rate by

m = A5,5V5 [lbm /s | kg /s] (5.49)

where the density at station 5 is obtained from the perfect gas law using p5 and T5 from
Equations (5.45) and (5.46). If A5 is known, the mass flow rate through the engine may
be determined from Equation (5.49) and the thrust from Equation (5.48).
Another type of nozzle used in high-performance engines and in rocket nozzles is a
convergent-divergent nozzle, one in which the flow area first contracts and then
increases. It differs from the convergent nozzle in that it can have supersonic flow at
the exit. For such a fully expanded, convergent-divergent nozzle operating at its design
condition, p5 = pa, and the engine thrust from Equation (5.48) reduces to m(V5 – Va).

Jet Engine Performance

It is seen that engine thrust is proportional to the mass flow rate through the engine and
to the excess of the jet velocity over the flight velocity. The specific thrust of an engine
is defined as the ratio of the engine thrust to its mass flow rate. From Equation (5.48)
206

the specific thrust is

F/m = (V5 – Va) + ( p5 – pa)A5/m [lbf -s/lbm | kN-s/kg] (5.50)

Because the engine mass flow rate is proportional to its exit area, as seen in Equation
(5.49), A5/m depends only on design nozzle exit conditions. As a consequence, F/m is
independent of mass flow rate and depends only on flight velocity and altitude.
Assigning an engine design thrust then determines the required engine-mass flow rate
and nozzle exit area and thus the engine diameter. Thus the specific thrust, F/m, is an
important engine design parameter for scaling engine size with required thrust at given
flight conditions.
Another important engine design parameter is the thrust specific fuel consumption,
TSFC, the ratio of the mass rate of fuel consumption to the engine thrust

TSFC = mf /F [lbm / lbf-s | kg / kN-s ] (5.51)

Low values of TSFC, of course, are favorable. The distance an aircraft can fly without
refueling, called its range, is inversely proportional to the TSFC of its engines. The
following example demonstrates the evaluation of these parameters.

EXAMPLE 5.6

An aircraft flies at a speed of 250 m/s at an altitude of 5000 m. The engines operate at
a compressor pressure ratio of 8, with a turbine inlet temperature of 1200K. The
compressor and turbine efficiencies are 0.9 and 0.87, respectively, and there is a 4%
pressure loss in the combustion chamber. The inlet total pressure recovery is 0.97, and
the engine-mass flow rate is 100 kg/s. Use an engine mechanical efficiency of 0.99 and
a fuel heating value of 43,000 kJ/kg. Assume that the engine has a convergent,
isentropic, nozzle flow. Determine the nozzle exit area, the engine thrust, specific
thrust, fuel flow rate, and thrust specific fuel consumption.

Solution
The solution details are presented in Table 5.5 in spreadsheet form. At 5000 m
altitude, the ambient static temperature and pressure are determined from standard-
atmosphere tables such as those given in Appendix H. The ambient stagnation pressure
and temperature are then determined for the given flight speed. The stagnation
temperature at the compressor entrance is the same as the free-stream value for an
adiabatic inlet. The inlet pressure recovery determines the total pressure at the com-
pressor entrance. Using the notation of Figures 5.21 and 5.22, and given the com-
pressor pressure ratio and turbine inlet temperature, the stagnation conditions at the
compressor, combustor, and turbine exits may be determined in the same way as the
static conditions were determined earlier for stationary two-shaft gas turbines.
207

The calculated available nozzle pressure ratio po4/pa = 4.118 is then compared with
the critical pressure ratio p04/pc = 1.852, which indicates that the convergent nozzle is
choked; i.e., sonic velocity exists at the throat. The nozzle exit plane pressure must
then be given by p04 divided by the critical pressure ratio. Equation (5.45) for a sonic
condition then determines the nozzle exit plane temperature, and the exit plane density
follows from the ideal gas law. Because the exit is choked, the exit plane temperature
determines the exit velocity through the sonic velocity relation.
The nozzle exit area may then be determined by using the given mass flow rate and
208

the exit velocity and density from Equation (5.49). The fuel-air ratio for the combustor
is estimated from a simple application of the steady-flow energy equation to the
combustor, which neglects the fuel sensible heat with respect to its heating value and
assumes hot-air properties for the combustion products. The thrust, specific thrust, and
TSFC are then found from Equations (5.48), (5.50), and (5.51).
_____________________________________________________________________

The spreadsheet in Table 5.5 (available from Spreadsheet Examples as Example


5.6) is set up with conditional statements that treat the convergent nozzle for both the
choked and subsonic exit conditions. The calculations of T5, V5, and p5 depend on
whether the throat is choked or not. The format for the spreadsheet conditional
statements is:

Conditional test, Result if true, Result if not true

The low value of fuel-air ratio, f = 0.0174, obtained in this example is typical of most
gas turbines and jet engines. In comparison with the stoichiometric value of 0.068, it
corresponds to an equivalence ratio of 0.256.

Modern Jet Engines

Full and cutaway views of a small modern jet engine used in business aircraft are shown
in Figure 5.23. The engine is a turbofan engine, a type of gas turbine engine that is
used in all large commercial aircraft and is gaining popularity in the business jet market.
The fan referred to in the turbofan name is seen at the left in Figure 5.23(b). The
incoming air splits after passing through the fan, with the flow through the outer
annulus passing to a nozzle that expels it without heating. The inner-core flow leaving
the fan passes through an axial flow compressor stage and a centrifugal compressor
before entering the combustor, turbine stages, and exit nozzle. Thus the core flow
provides the turbine power to drive the fan and the compressors. The thrust specific
fuel consumption and the thrust/weight ratio of turbofans is superior to those of
conventional jet engines, because a larger mass flow rate of air is processed and exits at
high velocity. The lower average exit velocity of the turbofan engine (compared to
turbojets) is secondary in importance to the increased total mass flow rate through the
engine.
A large turbofan engine designed to power Boeing 747 and 767, Airbus A310 and
A300, and McDonnell-Douglas MD-11 aircraft is shown in Figure 5.24. In large
turbofans and most large jet engines, axial compressors and turbines are used rather
than centrifugal compressors. The axial compressors are capable of much higher
pressure ratios, allow compression without turning the flow through a large angle, and
have somewhat higher efficiencies than centrifugals. Large axial compressors have
many axial stages and are capable of overall pressure ratios in excess of 30. Turbofans
are studied in more detail in Chapter 9.
209

Afterburning

The exhaust of a jet engine contains a large amount of unused oxygen because of the
high air-fuel ratios necessary to limit the gas stagnation temperature to which the
210

turbine blading is exposed. This excess oxygen at the turbine exit makes it possible to
burn additional fuel downstream and thereby to increase the nozzle exit temperature
and jet velocity. By extending the interface between the turbine and the nozzle in a jet
engine and by adding fuel spray bars to create a large combustion chamber called an
afterburner, it is possible to dramatically increase the thrust of a jet engine. Much
higher afterburner stagnation temperatures are allowed than those leaving the
combustor, because (a) there is no highly stressed rotating machinery downstream of
the afterburner, and (b) afterburner operating periods are usually limited to durations of
a few minutes. Afterburners are used for thrust augmentation of jet aircraft to assist in
takeoff and climb and to provide a brief high-speed-dash capability and increased
maneuver thrust in military aircraft. However, the substantial fuel consumption penalty
of afterburning restricts its use to brief periods of time when it is badly needed.
The T-s diagram for a jet engine with afterburner seen in Figure 5.25 shows that
afterburning is analogous to reheating in an open-cycle stationary gas turbine, as shown
in Figure 5.16. The energy release in the afterburner at approximately constant
stagnation pressure shifts the nozzle expansion process to the right on the T-s diagram.
As a result, the nozzle enthalpy and temperature differences between T-s diagram
211

constant-pressure lines increase as the nozzle inlet stagnation temperature increases.


This produces higher jet velocities. Using Equation (5.47) and the notation
of Figure 5.16, we get for the ratio of fully expanded nozzle isentropic jet velocity with
afterburning to that without:

V/Vwo = [(Tom* – T4)/(Tom – T4*)]½

= (Tom* /Tom)½ [(1 – T4 /Tom*)/(1 – T4*/Tom)]½

= (Tom* /Tom)½
where the static-to-stagnation temperature ratios are eliminated using the
corresponding equal isentropic pressure ratios. Thus, for example, the jet velocity ratio,
and therefore the thrust ratio, for a reheat temperature ratio of 4 is about 2. The
analysis of a jet engine with afterburner is illustrated in the following example.

EXAMPLE 5.7

Consider the performance of the engine analyzed in Example 5.6 when an afterburner is
added. Assume that heat addition in the afterburner raises the nozzle entrance
stagnation temperature to 2000 K, with a 5% stagnation pressure loss in the
afterburner. What is the increase in nozzle exit temperature, jet velocity, and thrust?
212

Solution
Table 5.6 is from an adaptation of the spreadsheet (Table 5.5) used for Example
5.6. The calculations are identical up to the turbine exit at station 4. Heat addition in
the afterburner raises the stagnation temperature at the nozzle entrance, Tob, to 2000K,
while the 5% pressure loss drops the stagnation pressure to 2.115 bar. Comparing the
applied nozzle pressure ratio (pob/pa) with the critical pressure ratio (pob/pc) shows that
213

the nozzle remains choked. The remainder of the calculation follows Example 5.6
except that the fuel consumption associated with the afterburner temperature rise is
taking into account in the fuel-air ratio, fuel flow rate, and specific fuel consumption.
_____________________________________________________________________

Table 5.7 compares some of the performance parameters calculated in Examples 5.6
and 5.7. The table shows clearly the striking gain in thrust provided by the high nozzle
exit temperature produced by afterburning and the accompanying high penalty in fuel
consumption. Note, however, that even with afterburning the overall equivalence ratio
0.0448/0.068 = 0.659 is well below stoichiometric.

Table 5.7 Examples 5.6 and 5.7 Compared


Parameter Without Afterburner With Afterburner

Temperature, K

Nozzle entrance 970.2 2000.0

Nozzle exit 831.6 1714.3

Nozzle exit velocity, m/s 564.1 809.9

Thrust, kN 54.7 87.99

Fuel-air Ratio 0.0174 0.0448

TSFC, kg/kN-s 0.032 0.051

Bibliography and References

1. Cohen, H., Rogers, G. F. C., and Saravanamuttoo, H. I. H., Gas Turbine Theory,
3rd Edition. New York: Longman Scientific and Technical, 1987.

2. Potter, J. H., “The Gas Turbine Cycle." ASME paper presented at the Gas Turbine
Forum Dinner, ASME Annual Meeting, New York, N.Y., November 27, 1972.

3. Bathie, William W., Fundamentals of Gas Turbines. New York: Wiley, 1984.

4. Whittle, Sir Frank, Gas Turbine Aerothermodynamics. New York: Pergamon, 1981.

5. Wilson, David Gordon, The Design of High-Efficiency Turbomachinery and Gas


Turbines. Cambridge, Mass.: MIT Press, 1984.

6. Chapman, Alan J., and Walker, William F., Introductory Gas Dynamics. New York:
Holt, Rinehart and Winston, 1971.
214

7. Oates, Gordon C., Aerothermodynamics of Gas Turbine and Rocket Propulsion.


Washington, D. C.: American Institute of Aerodynamics and Astronautics, 1988.

8. Bammert, K., and Deuster, G., “Layout and Present Status of the Closed-Cycle
Helium Turbine Plant Oberhausen.” ASME paper 74-GT-132, 1974.

9. Bammert, Karl, “The Oberhausen Heat and Power Station with Helium Turbine.”
Address on the Inauguration of the Helium Turbine Power Plant of EVO at
Oberhausen-Sterkrade, Westdeutschen Allgemeinen Zeitung, WAZ, December 20,
1974.

10. Zenker, P., “The Oberhausen 50-MW Helium Turbine Plant,” Combustion, April
1976, pp. 21-25.

11. Weston, Kenneth C., “Gas Turbine Analysis and Design Using Interactive
Computer Graphics.” Proceedings of Symposium on Applications of Computer
Methods in Engineering, University of Southern California, August 23-26, 1977.

12. United States Standard Atmosphere, 1976, NOAA, NASA, and USAF, October
1976.

13. Weston, Kenneth C., “Turbofan Engine Analysis and Optimization Using
Spreadsheets.” ASME Computers in Engineering Conference, Anaheim, California,
July 30-August 3, 1989.

14. Kerrebrock, Jack K., Aircraft Engines and Gas Turbines. Cambridge, Mass.: MIT
Press, 1983.

15. Dixon, S. L., Fluid Mechanics, Thermodynamics of Turbomachinery. New York:


Pergamon Press, 1978.

16. Bammert, K., Rurik, J., and Griepentrog, H., “Highlights and Future Development
of Closed-Cycle Gas Turbines.” ASME paper 74-GT-7, 1974.

EXERCISES

5.1 For the Air Standard Brayton cycle, express the net work in terms of the
compressor pressure ratio, r, and the turbine-to-compressor inlet temperature ratio,
T3/T1. Nondimensionalize the net work with cpT1, and derive an expression for the
pressure ratio that maximizes the net work for a given value of T3/T1.

5.2 For a Brayton Air Standard cycle, work out an expression for the maximum
possible compressor pressure ratio for a given turbine-to-compressor inlet temperature
215

ratio. Draw and label the cycle on a T-s diagram. What is the magnitude of the net
work for this cycle? Explain.

5.3 For a calorically perfect gas, write an expression for the temperature difference,
T2 – T1, on an isentrope between two lines of constant pressure in terms of the initial
temperature T1 and the pressure ratio p2/p1. Sketch a T-s diagram showing two different
isentropes between the two pressure levels. Explain how your expression demonstrates
that the work of an isentropic turbomachine operating between given pressure levels
increases with temperature.

5.4 Derive an expression for the enthalpy difference, h2 – h1, along a calorically perfect
gas isentrope spanning two fixed pressure levels, p2 and p1, in terms of the discharge
temperature T2. Note that as T2 increases, the enthalpy difference also increases.

5.5 Derive Equation (5.9).

5.6 Derive Equations (5.24) and (5.25).

5.7 A simple-cycle stationary gas turbine has compressor and turbine efficiencies of
0.85 and 0.9, respectively, and a compressor pressure ratio of 20. Determine the work
of the compressor and the turbine, the net work, the turbine exit temperature, and the
thermal efficiency for 80°F ambient and 1900°F turbine inlet temperatures.

5.8 A simple-cycle stationary gas turbine has compressor and turbine efficiencies of
0.85 and 0.9, respectively, and a compressor pressure ratio of 20. Determine the work
of the compressor and the turbine, the net work, the turbine exit temperature, and the
thermal efficiency for 20°C ambient and 1200°C turbine inlet temperatures.

5.9 A regenerative-cycle stationary gas turbine has compressor and turbine isentropic
efficiencies of 0.85 and 0.9, respectively, a regenerator effectiveness of 0.8, and a
compressor pressure ratio of 5. Determine the work of the compressor and the turbine,
the net work, the turbine and regenerator exit temperatures, and the thermal efficiency
for 80°F ambient and 1900°F turbine inlet temperatures. Compare the efficiency of the
cycle with the corresponding simple-cycle efficiency.

5.10 A regenerative-cycle stationary gas turbine has compressor and turbine isentropic
efficiencies of 0.85 and 0.9, respectively, a regenerator effectiveness of 0.8, and a
compressor pressure ratio of 5. Determine the work of the compressor and the turbine,
the net work, the turbine and regenerator exit temperatures, and the thermal efficiency
for 20°C ambient and 1200°C turbine inlet temperatures. Compare the efficiency of the
cycle with the corresponding simple-cycle efficiency.

5.11 A two-shaft stationary gas turbine has isentropic efficiencies of 0.85, 0.88, and 0.9
216

for the compressor, gas generator turbine, and power turbine, respectively, and a
compressor pressure ratio of 20.
(a) Determine the compressor work and net work, the gas generator turbine exit
temperature, and the thermal efficiency for 80°F ambient and 1900°F compressor-
turbine inlet temperatures.
(b) Calculate and discuss the effects of adding reheat to 1900°F ahead of the
power turbine.

5.12 A two-shaft stationary gas turbine has isentropic efficiencies of 0.85, 0.88, and
0.9 for the compressor, gas generator turbine, and power turbine, respectively, and a
compressor pressure ratio of 20.
(a) Determine the compressor work and net work, the gas generator turbine exit
temperature, and the thermal efficiency for 20°C ambient and 1200°C compressor-
turbine inlet temperatures.
(b) Calculate and discuss the effects of adding reheat to 1200°C ahead of the
power turbine.

5.13 A two-shaft stationary gas turbine with an intercooler and reheater has
efficiencies of 0.85, 0.88, and 0.9 for the compressor, gas-generator turbine, and power
turbine, respectively, and a compressor pressure ratio of 5.
(a) Determine the compressor work and net work, the gas generator turbine exit
temperature, and the thermal efficiency for 80°F ambient and 1900°F turbine inlet
temperatures.
(b) Calculate and discuss the effect on thermal efficiency, exhaust temperature, and
net work of adding a regenerator with an effectiveness of 75%.

5.14 A two-shaft stationary gas turbine with an intercooler and reheater has
efficiencies of 0.85, 0.88, and 0.9 for the compressor, gas-generator turbine, and power
turbine, respectively, and a compressor pressure ratio of 5.
(a) Determine the compressor work and net work, the gas generator turbine exit
temperature, and the thermal efficiency for 20°C ambient and 1200°C turbine inlet
temperatures.
(b) Calculate and discuss the effect on thermal efficiency, exhaust temperature, and
net work of adding a regenerator with an effectiveness of 75%.

5.15 Consider a pulverized-coal-burning, single-shaft gas turbine in which the


combustion chamber is downstream of the turbine to avoid turbine blade erosion and
corrosion. The combustion gases leaving the burner heat the compressor discharge air
through the intervening walls of a high temperature ceramic heat exchanger.
(a) Sketch the flow and T-s diagrams for this gas turbine, showing the influence of
pressure drops through the combustor and the heat exchanger. The ambient, turbine
inlet, and combustor exhaust temperatures are 80°F, 1900°F, and 3000°F, respectively.
The compressor pressure ratio is 5. Assume perfect turbomachinery.
217

(b) For zero pressure drops, determine the net work, the thermal efficiency, and
the heat exchanger exhaust temperature.
(c) If the coal has a heating value of 14,000 Btu/lbm, what is the coal consumption
rate, in tons per hour, for a 50-MW plant?

5.16 Consider a pulverized-coal-burning, single-shaft gas turbine in which the


combustion chamber is downstream of the turbine to avoid turbine blade erosion and
corrosion. The combustion gases leaving the burner heat the compressor discharge air
through the intervening walls of a high-temperature ceramic heat exchanger.
(a) Sketch the flow and T-s diagrams for this gas turbine, showing the influence of
pressure drops through the combustor and the heat exchanger. The ambient, turbine
inlet, and combustor exhaust temperatures are 20°C, 1200°C, and 2000°C,
respectively. The compressor pressure ratio is 5. Assume perfect turbomachinery.
(b) For zero pressure drops, determine the net work, the thermal efficiency, and
the heat exchanger exhaust temperature.
(c) If the coal has a heating value of 25,000 kJ/kg, what is the coal consumption
rate, in tons per hour, for a 50-MW plant?

5.17 A stationary gas turbine used to supply compressed air to a factory operates with
zero external shaft load. Derive an equation for the fraction of the compressor inlet air
that can be extracted ahead of the combustion chamber for process use in terms of the
compressor pressure ratio, the ratio of turbine-to-compressor inlet temperatures, and
the turbomachinery efficiencies. Plot the compressor mass extraction ratio as a
function of compressor pressure ratio for temperature ratios of 3 and 5, perfect
turbomachinery, and identical high- and low-temperature heat capacities.

5.18 A stationary gas turbine used to supply compressed air to a factory operates with
zero external shaft load. Derive an equation for the fraction of the inlet air that can be
extracted ahead of the combustion chamber for process use in terms of the compressor
pressure ratio, the ratio of turbine-to-compressor inlet temperatures, and the turbo-
machinery efficiencies. What is the extraction mass flow for a machine that has a
compressor pressure ratio of 10, turbomachine inlet temperatures of 1800°F and 80°F,
turbomachine efficiencies of 90%, and a compressor inlet mass flow rate of 15 lbm/s?

5.19 A stationary gas turbine used to supply compressed air to a factory operates with
zero external shaft load. Derive an equation for the fraction of the inlet air that can be
extracted ahead of the combustion chamber for process use in terms of the compressor
pressure ratio, the ratio of turbine-to-compressor inlet temperatures, and the turbo-
machinery efficiencies. What is the extraction mass flow for a machine that has a
compressor pressure ratio of 8, turbomachine inlet temperatures of 1000°C and 25°C,
turbomachine efficiencies of 90%, and a compressor inlet mass flow rate of 10 kg /s?

5.20 Do Exercise 5.18, but account for combustion chamber fractional pressure drops
218

of 3% and 5% of the burner inlet pressure. How does increased pressure loss influence
process mass flow?

5.21 Do Exercise 5.19, but account for combustion chamber fractional pressure drops
of 3% and 5% of the burner inlet pressure. How does increased pressure loss influence
process mass flow?

5.22 A gas-turbine-driven car requires a maximum of 240 shaft horsepower. The


engine is a two-shaft regenerative gas turbine with compressor, gas generator turbine,
and power turbine efficiencies of 0.86, 0.9, and 0.87, respectively, and a regenerator
effectiveness of 0.72. The compressor pressure ratio is 3.7, and the turbine and
compressor inlet temperatures are 1800°F and 90°F, respectively. What air flow rate
does the engine require? What is the automobile exhaust temperature? What are the
engine fuel-air ratio and specific fuel consumption if the engine burns gaseous methane?

5.23 A gas-turbine-driven car requires a maximum of 150kW of shaft power. The


engine is a two-shaft regenerative gas turbine with compressor, gas generator turbine,
and power turbine efficiencies of 0.84, 0.87 and 0.9, respectively, and a regenerator
effectiveness of 0.75. The compressor pressure ratio is 4.3, and the turbine and
compressor inlet temperatures are 1250°C and 20°C, respectively. What air flow rate
does the engine require? What is the automobile exhaust temperature? What are the
engine fuel-air ratio and specific fuel consumption if the engine burns gaseous
hydrogen?

5.24 A simple-cycle gas turbine is designed for a turbine inlet temperature of 1450°F,
a compressor pressure ratio of 12, and compressor and turbine efficiencies of 84% and
88%, respectively. Ambient conditions are 85°F and 14.5 psia.
(a) Draw and label a T-s diagram for this engine.
(b) Determine the compressor, turbine, and net work for this cycle.
(c) Determine the engine thermal efficiency.
(d) Your supervisor has requested that you study the influence of replacing the
compressor with dual compressors and an intercooler. Assume that the new
compressors are identical to each other and have the same efficiencies and combined
overall pressure ratio as the original compressor. Assume intercooling to 85°F with no
intercooler pressure losses. Show clearly the T-s diagram for the modified system
superimposed on your original diagram. Calculate the revised system net work and
thermal efficiency.

5.25 Determine the air and kerosene flow rates for a 100-MW regenerative gas turbine
with 1800K turbine inlet temperature, compressor pressure ratio of 5, and 1 atm. and
300K ambient conditions. The compressor and turbine efficiencies are 81% and 88%,
respectively, and the heat exchanger effectiveness is 75%. Use a heating value for
kerosene of 45,840 kJ/kg. What is the engine specific fuel consumption?
219

5.26 A regenerative gas turbine has compressor and turbine discharge temperatures of
350K and 700K, respectively. Draw and label a T-s diagram showing the relevant
states. If the regenerator has an effectiveness of 70%, what are the combustor inlet
temperature and engine exhaust temperature?

5.27 A gas turbine has a turbine inlet temperature of 1100K, a turbine pressure ratio of
6, and a turbine efficiency of 90%. What are the turbine exit temperature and the
turbine work?

5.28 A two-shaft gas turbine with reheat has turbine inlet temperatures of 1500°F, a
compressor pressure ratio of 16, and turbomachine efficiencies of 88% each. The
compressor inlet conditions are 80°F and 1 atm. Assume that all heat capacities are
0.24 Btu/lbm-R and k = 1.4.
(a) Draw T-s and flow diagrams.
(b) Make a table of temperatures and pressures for all real states, in °R and atm.
(c) What are the compressor work and power turbine work?
(d) What is the power-turbine-to-compressor work ratio?
(e) What is the cycle thermal efficiency?
(f) Evaluate the recommendation to add a regenerator to the system. If a 4-count
(0.04) increase in thermal efficiency can be achieved, the addition of the regenerator is
considered economically feasible. Give your recommendation, supporting arguments,
and substantiating quantitative data.

5.29 Consider a gas turbine with compressor and turbine inlet temperatures of 80°F
and 1200°F, respectively. The turbine efficiency is 85%, and compressor pressure
ratio is 8.
(a) Draw coordinated T-s and plant diagrams.
(b) What is the turbine work?
(c) What is the minimum compressor efficiency required for the gas turbine to
produce a net power output?
(d) What is the thermal efficiency if the compressor efficiency is raised to 85%?

5.30 The first closed-cycle gas turbine power plant in the world using helium as a
working fluid is a 50-MW plant located in Oberhausen, Germany (ref. 8). It was
designed as an operating power plant and as a research facility to study aspects of
component design and performance with helium as a working fluid. It has two
compressors, with intercooling, connected directly to a high-pressure turbine. The high-
pressure turbine is in turn connected through a gearbox to a low-pressure turbine with
no reheat. Helium is heated first by regenerator, followed by a specially designed
heater that burns coke-oven gas. A water-cooled pre-cooler returns the helium to the
low-pressure compressor inlet conditions. The high-pressure turbine mass flow rate is
84.4 kg/s, and the heater efficiency is 92.2%. The following design data are given in
the reference:
220

Temperature, °C Pressure, Bar

1. Low-pressure compressor inlet 25 10.5

2. Intercooler inlet 83 15.5

3. High-pressure compressor inlet 25 15.4

4. Regenerator inlet, high-pressure side 125 28.7

5. Heater inlet 417 28.2

6. High-pressure turbine inlet 750 27.0

7. Low-pressure turbine inlet 580 16.5

8. Regenerator inlet, low-pressure side 460 10.8

9. Precooler inlet 169 10.6

In the following, assume k = 1.67 and cp = 5.197 kJ/kg-K for helium.


(a) Sketch and label flow and T-s diagrams for the plant.
(b) What is the overall engine pressure ratio of the gas turbine?
(c) Estimate the mechanical power output and plant thermal efficiency. Reference
8 gives 50-MW and 31.3% as net electrical output and efficiency, respectively.
Evaluate your calculations with these data.

5.31 Determine the thrust for a turbojet engine flying at 200 m/s with a compressor
inlet temperature of 27°C, a compressor pressure ratio of 11, a turbine inlet
temperature of 1400K, and compressor and turbine efficiencies of 0.85 and 0.9,
respectively. The engine mass flow rate is 20 kg/s. You may use k = 1.4 and cp =
1.005 kJ/kg-K throughout and neglect the differences between static and stagnation
properties in the turbomachinery. Assume an ambient pressure of 0.2615 atmospheres.

5.32 The gas generator of a two-shaft gas turbine has a compressor pressure ratio of 5
and compressor and turbine inlet temperatures of 80°F and 2000°F, respectively, at sea
level. All turbomachines have efficiencies of 90%, and the inlet air flow is 50 lbm/s.
(a) What are the net work, pressure ratio, and horsepower of the power turbine
and the cycle efficiency?
(b) Suppose the power turbine is removed and the gas generator exhaust gas flows
isentropically through a convergent-divergent propulsion nozzle that is fully expanded
(exit pressure is ambient). What are the nozzle exhaust velocity and the static thrust?
(c) Repeat part (b) for a choked conversion nozzle.

5.33 A gas turbine with reheat has two turbines with efficiencies e1 and e2. Derive
relations for the turbine pressure ratios r1 and r2 that maximize the total turbine work
for a given overall turbine pressure ratio, r, if both turbines have the same inlet
temperature. How do the pressure ratios compare if the turbine efficiencies are equal?
221

If the efficiency of one turbine is 50% higher than the other, what is the optimum
pressure ratio?

5.34* For a simple-cycle gas turbine, develop a multicolumn spreadsheet that


tabulates and plots (a) the net work, nondimensionalized by using the product of the
compressor constant-pressure heat capacity and the compressor inlet temperature, and
(b) the thermal efficiency, both as a function of compressor pressure ratio (only one
graph with two curves). Use 80°F and 2000°F as compressor and turbine inlet
temperatures, respectively, and 0.85 and 0.90 as compressor and turbine efficiencies,
respectively. Use appropriate constant heat capacities. Each of the input values should
be entered in separate cells in each column so that the spreadsheet may be used for
studies with other parametric values. Use your plot and table to determine the pressure
ratio that yields the maximum net work. Compare with your theoretical expectation.

5.35* Solve Exercise 5.34 for a regenerative cycle. Use a nominal value of 0.8 for
regenerator effectiveness.

5.36* Solve Exercise 5.34 for a two-shaft regenerative cycle. Use a nominal value of
0.8 for regenerator effectiveness and 0.88 for the power turbine efficiency. Account
also for 3% pressure losses in both sides of the regenerator.

5.37 For the conditions of Exercise 5.31, but using more realistic properties in the
engine hot sections, plot curves of thrust and specific fuel consumption as a function of
gaseous hydrogen air-fuel ratio. Determine the maximum thrust corresponding to the
stoichiometric limit for gaseous hydrogen fuel.

5.38* Use the spreadsheet corresponding to Table 5.4 to plot a graph showing the
influence of turbine inlet temperature on net work and thermal efficiency for two-shaft
gas turbines, with and without regeneration, for a compressor pressure ratio of 4.

5.39 Extend Example 5.6 by using hand calculations to evaluate the thrust and the
thrust specific fuel consumption for the engine if it were fitted with a fully expanded
(exit pressure equal to ambient pressure) convergent-divergent isentropic nozzle.

5.40* Modify the spreadsheet corresponding to Table 5.6 to evaluate the thrust and the
thrust specific fuel consumption for the engine if it were fitted with a fully expanded
(exit pressure equal to ambient pressure) convergent-divergent isentropic nozzle.

5.41 A gas turbine engine is being designed to provide work and a hot, high-velocity
exhaust flow. The compressor will have a pressure ratio of 4 and an isentropic
efficiency of 90% at the design point. The compressor and load are driven by separate
___________________
* Exercise numbers with asterisks involve computer usage.
222

turbines, but the overall expansion pressure ratio across the turbines will be 3 to 1, and
the efficiency of each turbine will be 90%. The exhaust of the low-pressure turbine is
expanded through a convergent nozzle to provide the high-velocity exhaust. At the
design point the turbine inflow air temperature will be held to 1140°F, and the air flow
rate will be 300,000 pounds mass per hour. Ambient conditions are 60°F and 14.7 psia.
Calculate the brake power of the engine, in kW, and the temperatures at the entrance
and the exit of nozzle.

5.42 A ramjet is a jet engine that flies at speeds high enough that the pressure rise
produced by ram effect in the inlet makes a compressor and turbine unnecessary. At
50,000 feet and Mach 3, the inlet has a stagnation pressure recovery of 85%.
Combustion raises the air temperature to 2500K. What is the thrust per unit mass flow
rate of air and the exit velocity of the engine if the nozzle is (a) convergent, and (b)
fully expanded (exit pressure equal to ambient pressure) convergent-divergent?

5.43 Using the First Law of Thermodynamics, derive an equation for the work of
compression in a reversible steady flow in terms of volume and pressure. Use the
equation to derive an expression for the reversible isothermal work of compression of a
calorically perfect gas, with compressor pressure ratio as an independent variable.

5.44 Sketch a pressure-volume diagram comparing isothermal and isentropic


compressions starting at the same state and having the same pressure ratio. Show for a
thermally perfect gas that, at a given state, the isentrope has a steeper (negative) slope
than an isotherm. Use your diagram to prove that the isothermal work of compression
is less than the isentropic work.

5.45 A supersonic aircraft flies at Mach 2 at an altitude of 40,000 feet. Its engines
have a compressor pressure ratio of 20 and a turbine inlet temperature of 2000°F. The
inlets have total pressure recoveries of 89%, the compressors and turbines all have
efficiencies of 90%, and the nozzles are convergent and isentropic. Determine the
nozzle exit velocity and thrust, and estimate the thrust specific fuel consumption of
each engine. The mass flow rate of air of each engine is 750 lbm/s. Use a fuel heating
value of 18,533 Btu/lbm.

5.46 A supersonic aircraft flies at Mach 2 at an altitude of 13,000 m. Its engines have
a compressor pressure ratio of 20 and a turbine inlet temperature of 1500K. The inlets
have total pressure recoveries of 89%, the compressors and turbines all have
efficiencies of 90%, and the nozzles are convergent and isentropic. Determine the
nozzle exit velocity and thrust, and estimate the thrust specific fuel consumption of an
engine for an engine air mass flow rate of 100 kg /s. Use a fuel heating value of 43,100
kJ/kg.

5.47 A supersonic aircraft flies at Mach 2 at an altitude of 40,000 feet. Its engines
223

have a compressor pressure ratio of 20 and a turbine inlet temperature of 2000°F. The
inlets have total pressure recoveries of 89%, and the compressors and turbines all have
efficiencies of 90%. The engines have afterburners that raise the temperature of the gas
entering the nozzles to 3000°F. The nozzles are convergent and isentropic. Compare
the design thrust and thrust specific fuel consumption with the afterburner on and off
for an engine air mass flow rate of 100 lbm/s.

5.48* For a simple-cycle gas turbine, develop a multicolumn spreadsheet that tabulates
and plots (a) the net work, nondimensionalized by using the compressor constant-
pressure heat capacity and the compressor inlet temperature, and (b) the thermal
efficiency, as a function of compressor pressure ratio (only one graph with two curves).
Use 30° C and 1500°C as compressor and turbine inlet temperatures, respectively, and
0.85 and 0.90 as compressor and turbine efficiencies, respectively. Assume appropriate
constant heat capacities. Each of the input values should be entered in separate cells in
each column, so that the spreadsheet may be used for studies with other parametric
values. Use your plot and table to determine the pressure ratio that yields the
maximum net work. Compare with your theoretical expectation.

5.49* Solve Exercise 5.48 for a regenerative cycle. Use a nominal value of 0.85 for
regenerator effectiveness.

5.50 Methane is burned in an adiabatic gas turbine combustor. The fuel enters the
combustor at the reference temperature for the JANAF tables and mixes with air
compressed from 80°F through a pressure ratio of 16 with a compressor efficiency of
90.5%. Determine the equivalence ratio that limits the turbine inlet temperature to
2060°F by:
(a) Using the JANAF tables.
(b) Using an energy balance on the combustion chamber and the lower heating
value for methane.
224

CHAPTER 6

RECIPROCATING INTERNAL
COMBUSTION ENGINES

6.1 Introduction

Perhaps the best-known engine in the world is the reciprocating internal combustion
(IC) engine. Virtually every person who has driven an automobile or pushed a power
lawnmower has used one. By far the most widely used IC engine is the spark-ignition
gasoline engine, which takes us to school and work and on pleasure jaunts. Although
others had made significant contributions, Niklaus Otto is generally credited with the
invention of the engine and with the statement of its theoretical cycle.
Another important engine is the reciprocating engine that made the name of Rudolf
Diesel famous. The Diesel engine, the workhorse of the heavy truck industry, is widely
used in industrial power and marine applications. It replaced the reciprocating steam
engine in railroad locomotives about fifty years ago and remains dominant in that role
today.
The piston, cylinder, crank, and connecting rod provide the geometric basis of the
reciprocating engine. While two-stroke-cycle engines are in use and of continuing
interest, the discussion here will emphasize the more widely applied four-stroke-cycle
engine. In this engine the piston undergoes two mechanical cycles for each
thermodynamic cycle. The intake and compression processes occur in the first two
strokes, and the power and exhaust processes in the last two. These processes are made
possible by the crank-slider mechanism, discussed next.

6.2 The Crank-Slider Mechanism

Common to most reciprocating engines is a linkage known as a crank-slider mechan-


ism. Diagramed in Figure 6.1, this mechanism is one of several capable of producing
the straight-line, backward-and-forward motion known as reciprocating. Fundament-
ally, the crank-slider converts rotational motion into linear motion, or vice-versa. With
a piston as the slider moving inside a fixed cylinder, the mechanism provides the vital
capability of a gas engine: the ability to compress and expand a gas. Before delving into
this aspect of the engine, however, let us examine the crank-slider mechanism more
closely.
225

It is evident from Figure 6.2 that, while the crank arm rotates through 180°, the
piston moves from the position known as top-center (TC) to the other extreme, called
bottom-center (BC). During this period the piston travels a distance, S, called the
stroke, that is twice the length of the crank.
For an angular velocity of the crank, T, the crank pin A has a tangential velocity
component TS/2. It is evident that, at TC and at BC, the crank pin velocity component
in the piston direction, and hence the piston velocity, is zero. At these points,
corresponding to crank angle 2 = 0° and 180°, the piston reverses direction. Thus as 2
varies from 0° to 180°, the piston velocity accelerates from 0 to a maximum and then
returns to 0. A similar behavior exists between 180° and 360°.
The connecting rod is a two-force member; hence it is evident that there are both
axial and lateral forces on the piston at crank angles other than 0° and 180°. These
lateral forces are, of course, opposed by the cylinder walls. The resulting lateral force
component normal to the cylinder wall gives rise to frictional forces between the piston
rings and cylinder. It is evident that the normal force, and thus the frictional force,
alternates from one side of the piston to the other during each cycle. Thus the piston
motion presents a challenging lubrication problem for the control and reduction of both
wear and energy loss.
The position of the piston with respect to the crank centerline is given by

x = (S/2)cos2 + LcosN [ft | m] (6.1)

where yA = (S/2)sin2 = LsinN can be used to eliminate N to obtain

x/L = (S/2L)cos2 + [1– (S/2L)2 sin2 2 ]½ [dl] (6.2)

Thus, while the axial component of the motion of the crank pin is simple harmonic,
xA = (S/2)cos2, the motion of the piston and piston pin is more complex. It may be
226

seen from Equation (6.2), however, that as S/L becomes small, the piston motion
approaches simple harmonic. This becomes physically evident when it is recognized
that, in this limit, the connecting rod angle, N , approaches 0 and the piston motion
approaches the axial motion of the crank pin. Equations (6.1) and (6.2) may be used to
predict component velocities, accelerations, and forces in the engine.
The volume swept by the piston as it passes from TC to BC is called the piston
displacement, disp. Engine displacement, DISP, is then the product of the piston
displacement and the number of cylinders, DISP = (n)(disp). The piston displacement is
the product of the piston cross-sectional area and the stroke. The cylinder inside
diameter (and, approximately, also the piston diameter) is called its bore. Cylinder bore,
stroke, and number of cylinders are usually quoted in engine specifications along with
or instead of engine displacement. It will be seen later that the power output of a
reciprocating engine is proportional to its displacement. An engine of historical interest
that also used the crank-slider mechanism is discussed in the next section.

6.3 The Lenoir Cycle

An early form of the reciprocating internal combustion engine is credited to Etienne


Lenoir. His engine, introduced in 1860, used a crank-slider-piston-cylinder arrangement
227

in which a combustible mixture confined between the piston and cylinder is ignited after
TC. The resulting combustion gas pressure forces acting on the piston deliver work by
way of the connecting rod to the rotating crank. When the piston is at BC, combustion
gases are allowed to escape. The rotational momentum of the crank system drives the
piston toward TC, expelling additional gases as it goes. A fresh combustible mixture is
again admitted to the combustion chamber (cylinder) and the cycle is repeated.
The theoretical Lenoir cycle, shown in Figure 6.3 on a pressure-volume diagram,
consists of the intake of the working fluid (a combustible mixture) from state 0 to state
1, a constant-volume temperature and pressure rise from state 1 to state 2, approxim-
ating the combustion process, an isentropic expansion of the combustion gases to state
3, and a constant-pressure expulsion of residual gases back to state 0. Note that a
portion of the piston displacement, from state 0 to state 1, is used to take in the
combustible mixture and does not participate in the power stroke from state 2 to state
3. The engine has been called an explosion engine because the power delivered is due
only to the extremely rapid combustion pressure rise or explosion of the mixture in the
confined space of the cylinder.
Hundreds of Lenoir engines were used in the nineteenth century, but the engine is
quite inefficient by todays standards. In 1862, Beau de Rochas pointed out that the
228

efficiency of internal combustion could be markedly improved in reciprocating engines


by compression of the air-fuel mixture prior to combustion. In 1876 Niklaus Otto (who
is thought to have been unaware of Rochas’ suggestion) demonstrated an engine that
incorporated this important feature, as described next.

6.4 The Otto Cycle

The Otto cycle is the theoretical cycle commonly used to represent the processes in
the spark ignition (SI) internal combustion engine. It is assumed that a fixed mass of
working fluid is confined in the cylinder by a piston that moves from BC to TC and
back, as shown in Figure 6.4. The cycle consists of isentropic compression of an
air-fuel mixture from state 1 to state 2, constant-volume combustion to state 3,
isentropic expansion of the combustion gases to state 4, and a constant-volume heat
rejection back to state 1. The constant-volume heat rejection is a simple expedient to
close the cycle. It obviates the need to represent the complex expansion and outflow of
229

combustion gases from the cylinder at the end of the cycle. Note that the Otto cycle is
not concerned with the induction of the air-fuel mixture or with the expulsion of
residual combustion gases. Thus only two mechanical strokes of the crank-slider are
needed in the Otto cycle, even when it is used to represent an ideal four-stroke-cycle
Otto engine. In this case the remaining strokes are used to execute the necessary intake
and exhaust functions. Because it involves only two strokes, the Otto cycle may also
represent a two-stroke-cycle engine. The two-stroke-cycle engine is in principle
capable of as much work in one rotation of the crank as the four-stroke engine is in
two. However, it is difficult to implement because of the necessity of making the
intake and exhaust functions a part of those two strokes. It is therefore not as highly
developed or widely used as the four-stroke-cycle engine. We will focus on the four-
stroke-cycle here.
The simplest analysis of the Otto cycle assumes calorically perfect air as the work-
ing fluid in what is called the Air Standard cycle analysis. Following the notation of
Figure 6.4, the compression process can be represented by the isentropic relation for a
calorically perfect gas, Equation (1.21), as

p2/p1 = (V1/V2)k [dl] (6.3)

where the compression ratio, CR = V1/V2, is a fundamental parameter of all recipro-


cating engines. The diagram shows that the expansion ratio for the engine, V4 /V3, has
the same value, V1/V2. The clearance volume, V2, is the volume enclosed between the
cylinder head and the piston at TC. Thus the compression ratio may be expressed as the
ratio of the sum of the clearance and displacement volumes to the clearance volume:

CR = [V2 + (V1 – V2)]/V2


Thus, for a given displacement, the compression ratio may be increased by reducing the
clearance volume.
The efficiency of the cycle can be most easily determined by considering constant-
volume-process heat transfers and the First Law cyclic integral relation, Equation (1.3).
The heat transferred in the processes 2$3 and 4$1 are

q2$3 = cv (T3 – T2) [Btu/lbm | kj/kg] (6.4)

and

q4$1 = cv (T1 – T4) [Btu/lbm | kJ/kg] (6.5)

Both the expansion process, 3$4, and the compression process, 1$2, are assumed to
be isentropic. Thus, by definition, they are both adiabatic. From the cyclic integral, the
net work per unit mass is then:

w = q2$3 + q4$1 = cv (T3 – T2 + T1 – T4) [Btu/lbm | kJ/kg] (6.6)


230

As before, the cycle thermal efficiency is the ratio of the net work to the external heat
supplied:

0Otto = w/q2$3 = cv (T3 – T2 + T1 – T4) / [cv (T3 – T2)]

= 1 + (T1 – T4) / (T3 – T2)

= 1 – T1/T2 = 1 – 1 / CR k-1 [dl] (6.7)

where Equation (1.20) has been used to eliminate the temperatures. Equation (6.7)
shows that increasing compression ratio increases the cycle thermal efficiency. This is
true for real engines as well as for the idealized Otto engine. The ways in which real
spark ignition engine cycles deviate from the theoretical Otto cycle are discussed later.

EXAMPLE 6.1

An Otto engine takes in an air-fuel mixture at 80°F and standard atmosphere presssure.
It has a compression ratio of 8. Using Air Standard cycle analysis, a heating value of
20,425 Btu/lbm, and A/F = 15, determine:
(a) The temperature and pressure at the end of compression, after combustion, and
at the end of the power stroke.
(b) The net work per pound of working fluid.
(c) The thermal efficiency.

Solution
We use the notation of Figure 6.4:

(a) p2 = p1(V1/V2)k = 1(8)1.4 = 18.38 atm

T2 = T1(V1/V2)k – 1 = (540)(8)0.4 = 1240.6°R

T3 = T2 + qa /cv = T2 + (F/A)(HV)k/cp = 1240.6 + 1.4(20,425/15(0.24 = 9184°R

p3 = p2T3 /T2 = 18.38(9184/1240.6) = 136.1 atm

T4 = T3 /CRk–1 = 9184/ 80.4 = 3997.2°R

p4 = p3 /CRk = 136.1/81.4 = 7.4 atm

(b) The constant-volume heat addition is governed by the fuel-air ratio and the fuel
heating value:

qa = HV(F/A) = 20,425/15 = 1361.7 Btu/lbm of air


231

qr = cv (T1 – T4) = (0.24/1.4)( 540 – 3997.4) = – 592.7 Btu/lbm

w = qa + qr = 1361.7 + ( – 592.7) = 769 Btu/lbm

(c) The cycle termal efficiency may then be determined from the definition of the
heat engine thermal efficiency or Equation (6.7):

0th = w/qa = 769/1361.7 = 0.565

0th = 1 – 1/80.4 = 0.565


_____________________________________________________________________

In view of the discussion of gas properties and dissociation in Chapter 3, the values
of T3 and T4 in Example 6.1 are unrealistically high. Much of the energy released by the
fuel would go into vibration and dissociation of the gas molecules rather than into the
translational and rotational degrees of freedom represented by the temperature. As a
result, significantly lower temperatures would be obtained. Thus, while the analysis is
formally correct, the use of constant-low-temperature heat capacities in the Air
Standard cycle makes it a poor model for predicting temperature extremes when high
energy releases occur. Some improvement is achieved by using constant-high-
temperature heat capacities, but the best results would be achieved by the use of real
gas properties, as discussed in several of the references.

6.5 Combustion in a Reciprocating Engine

The constant-volume heat transfer process at TC in the Otto cycle is an artifice to


avoid the difficulties of modeling the complex processes that take place in the
combustion chamber of the SI engine. These processes, in reality, take place over a
crank angle span of 30° or more around TC. Let us consider aspects of these processes
and their implementation in more detail.
Normally, the mixture in the combustion chamber must have an air-fuel ratio in the
neighborhood of the stoichiometric value for satisfactory combustion. A more or less
homogeneous mixture may be produced outside the cylinder in a carburetor, by
injection into the intake manifold, or by throttle-body injection into a header serving
several intake manifolds. In the case of the carburetor, fuel is drawn into the engine
from the carburetor by the low pressure created in a venturi through which the
combustion air flows. As a result, increased air flow causes lower venturi pressure and
hence increased fuel flow. The fuel system thus serves to provide an air-fuel mixture
that remains close to the stoichiometric ratio for a range of air flow rates. Various
devices designed into the carburetor further adjust the fuel flow for the special
operating conditions encountered, such as idling and rapid acceleration.
Maximum fuel economy is usually attained with excess air to ensure that all of the
fuel is burned. A mixture with excess air is called a lean mixture. The carburetor
232

usually produces this condition in automobiles during normal constant-speed driving.


On the other hand, maximum power is achieved with excess fuel to assure that all
of the oxygen in the air in the combustion chamber is reacted. It is a matter of exploit-
ing the full power-producing capability of the displacement volume. A mixture with
excess fuel is called a rich mixture. The automotive carburetor produces a rich mixture
during acceleration by supplying extra fuel to the air entering the intake manifold.
The equivalence ratio is sometimes used to characterize the mixture ratio, whether
rich or lean. The equivalence ratio, M, is defined as the ratio of the actual fuel-air ratio
to the stoichiometric fuel-air ratio. Thus M > 1 represents a rich mixture and M < 1
represents a lean mixture. In terms of air-fuel ratio, M = (A/F)stoich /(A/F).
Homogeneous air-fuel mixtures close to stoichiometric may ignite spontaneously
(that is, without a spark or other local energy source) if the mixture temperature
exceeds a temperature called the autoignition temperature. If the mixture is brought to
and held at a temperature higher than the autoignition temperature, there is a period of
delay before spontaneous ignition or autoignition This time interval is called the
ignition delay, or ignition lag. The ignition delay depends on the characteristics of the
fuel and the equivalence ratio and usually decreases with increasing temperature.
In spark-ignition engines, compression ratios and therefore the temperatures at the
end of compression are low enough that the air-fuel mixture is ignited by the spark plug
before spontaneous ignition can occur. SI engines are designed so that a flame front
will propagate smoothly from the spark plug into the unburned mixture until all of the
mixture has been ignitied. However, as the flame front progresses, the temperature and
pressure of the combustion gases behind it rise due to the release of the chemical
energy of the fuel. As the front propagates, it compresses and heats the unburned
mixture, sometimes termed the end-gas. Combustion is completed as planned when the
front smoothly passes completely through the end-gas without autoignition. However,
if the end-gas autoignites, a pinging or low-pitched sound called knock is heard.
The avoidance of knock due to autoignition of the end-gas is a major constraint on
the design compression ratio of an SI engine. If hot spots or thermally induced
compression of the end-gas ignite it before the flame front does, there is a more rapid
release of chemical energy from the end-gas than during normal combustion. Knock is
sometimes thought of as an explosion of the end gas that creates an abrupt pulse and
pressure waves that race back and forth across the cylinder at high speed, producing
the familiar pinging or low-pitched sound associated with knock. Knock not only
reduces engine performance but produces rapid wear and objectionable noise in the
engine. Thus it is important for a SI engine fuel to have a high autoignition
temperature. It is therefore important for SI engine fuel to have a high autoignition
temperature. Thus the knock characteristics of commercially available fuels limit the
maximum allowable design compression ratio for SI engines and hence limit their best
efficiency.
The octane number is a measure of a gasoline's ability to avoid knock. Additives
such as tetraethyl lead have been used in the past to suppress engine knock. However,
the accumulation of lead in the environment and its penetration into the food cycle has
233

resulted in the phaseout of lead additives. Instead refineries now use appropriate blends
of hydrocarbons as a substitute for lead additives in unleaded fuels.
The octane number of a fuel is measured in a special variable-compression-ratio
engine called a CFR (Cooperative Fuels Research) engine. The octane rating of a fuel is
determined by comparison of its knocking characteristics with those of different
mixtures of isooctane, C8H18, and n-heptane, C7H16. One hundred percent isooctane is
defined as having an octane number of 100 because it had the highest resistance to
knock at the time the rating system was devised. On the other hand, n-heptane is
assigned a value of 0 on the octane number scale because of its very poor knock
resistance. If a gasoline tested in the CFR engine has the same knock threshold as a
blend of 90% isooctane and 10% n-heptane, the fuel is assigned an octane rating of 90.
In combustion chamber design, the designer attempts to balance many factors to
achieve good performance. Design considerations include locating intake valves away
from and exhaust valves near spark plugs, to keep end-gas in a relatively cool area of
the combustion chamber and thereby suppress hot-surface-induced autoignition
tendencies. Valves are, of course, designed as large as possible to reduce induction and
exhaust flow restrictions. More than one intake and one exhaust valve per cylinder are
now used in some engines to improve “engine breathing.” In some engines, four valves
in a single cylinder are employed for this purpose. The valves are also designed to
induce swirl and turbulence to promote mixing of fuel and air and to improve
combustion stability and burning rate.
Pollution and fuel economy considerations have in recent years profoundly
influenced overall engine and combustion chamber design. Stratified-charge engines,
for example, attempt to provide a locally rich combustion region to control peak
temperatures and thus suppress NOx formation. The resulting combustion gases
containing unburned fuel then mix with surrounding lean mixture to complete the
combustion process, thus eliminating CO and unburned hydrocarbons from the exhaust.
These processes occur at lower temperatures than in conventional combustion chamber
designs and therefore prevent significant nitrogen reactions.

6.6 Representing Reciprocating Engine Perfomance

In an earlier section, the theoretical work per unit mass of working fluid of the Otto
engine was evaluated for a single cycle of the engine, using the cyclic integral of the
First Law of Thermodynamics. The work done by pressure forces acting on a piston
can also be evaluated as the integral of pdV. It is evident therefore that the work done
during a single engine cycle is the area enclosed by the cycle process curves on the
pressure-volume diagram. Thus, instead of using the cyclic integral or evaluating pdV
for each process of the cycle, the work of a reciprocating engine can be found by
drawing the theoretical process curves on the p–V diagram and graphically integrating
them. Such a plot of pressure versus volume for any reciprocating engine, real or
theoretical, is called an indicator diagram.
234

In the nineteenth and early twentieth centuries a mechanical device known as an


engine indicator was used to produce indicator cards or diagrams to determine the
work per cycle for slow-running steam and gas reciprocating engines. The indicator
card was attached to a cylinder that rotated back and forth on its axis as the piston
oscillated, thus generating a piston position (volume) coordinate. At the same time a
pen driven by a pressure signal from the engine cylinder moved parallel to the cylinder
axis, scribing the p-V diagram over and over on the card. The work of high speed
engines is still evaluated from traces of pressure obtained with electronic sensors and
displayed on electronic monitors and through digital techniques.
The work done per cycle (from an indicator card, for instance) can be represented
as an average pressure times a volume. Because the displacement volumes of engines
are usually known, an engine performance parameter known as the mean effective
pressure, MEP, is defined in terms of the piston displacement. The mean effective
pressure is defined as the value of the pressure obtained by dividing the net work per
cylinder per cycle at a given operating condition by the piston displacement volume:

MEP = W/disp [lbf/ft2 | kPa] (6.8)

Thus the MEP is a measure of the effectiveness of a given displacement volume in


producing net work.
The power output of an engine with identical cylinders may be represented as the
product of the work per cycle and the number of cycles executed per unit time by the
engine. Thus if the engine has n cylinders, each executing N identical thermodynamic
cycles per unit time, and delivering W work units per cylinder, with a piston
displacement, disp, the power output is given by

P = n(N(W = n(N (MEP ( disp [ft-lbf /min | kW] (6.9)

Expressed for the entire engine, the engine displacement is DISP = n(disp and the
engine work is MEP (DISP. Hence the engine power is:

P = N (MEP(DISP [ft-lbf /min | kW] (6.10)

where N, the number of thermodynamic cycles of a cylinder per unit time, is the number
of crank-shaft revolutions per unit time for a two-stroke-cycle engine and one-half of
the revolutions per unit time for a four-stroke-cycle engine. The factor of ½ for the
four-stroke-cycle engine arises because one thermodynamic cycle is executed each time
the crank rotates through two revolutions.

EXAMPLE 6.2

What is the displacement of an engine that develops 60 horsepower at 2500 rpm in a


four-stroke-cycle engine having an MEP of 120 psi?
235

Solution
From Equation (6.10), the displacement of the engine is

DISP = P/(N (MEP) = (60)(33,000)(12)/[(2500/2)(120)] = 158.4 in3

Checking units: (HP)(ft-lbf/HP-min)(in/ft)/[(cycles/min)(lbf/in2)] = in3


_____________________________________________________________________

If the work is evaluated from an indicator diagram the work is called indicated
work; the MEP is called the indicated mean effective pressure, IMEP; and the power is
indicated power, IP. Note that the indicated work and power, being associated with the
work done by the combustion chamber gases on the piston, do not account for
frictional or mechanical losses in the engine, such as piston-cylinder friction or the drag
of moving parts (like connecting rods) as they move through air or lubricating oil.

Brake Performance Parameters

Another way of evaluating engine performance is to attach the engine output shaft
to a device known as a dynamometer, or brake. The dynamometer measures the
torque, T, applied by the engine at a given rotational speed. The power is then
calculated from the relation

P = 2B(rpm (T [ft-lbf /min | N-m/min] (6.11)

A simple device called a prony brake, which was used in the past, demonstrates the
concept for the measurement of the shaft torque of engines. Figure 6.5 shows the prony
brake configuration in which a stationary metal band wrapped around the rotating
flywheel of the engine resists the torque transmitted to it by friction. The product of the
force measured by a spring scale, w, and the moment arm, d , gives the resisting torque.
The power dissipated is then given by 2B(rpm)w (d.
Modern devices such as water brakes and electrical dynamometers long ago
replaced the prony brake. The water brake is like a centrifugal water pump with no
outflow, mounted on low-friction bearings, and driven by the test engine. As with the
prony brake, the force required to resist turning of the brake (pump) housing provides
the torque data. This, together with speed measurement, yields the power output from
Equation (6.11). The power dissipated appears as increased temperature of the water
in the brake and heat transfer from the brake. Cool water is circulated slowly through
the brake to maintain a steady operating condition. The torque measured in this way is
called the brake torque, BT, and the resulting power is called the brake power, BP. To
summarize: while indicated parameters relate to gas forces in the cylinder, brake
parameters deal with output shaft forces.
Thus the brake power differs from the indicated power in that it accounts for the
effect of all of the energy losses in the engine. The difference between the two is
referred to as the friction power, FP. Thus FP = IP – BP.
236

Friction power varies with engine speed and is difficult to measure directly. An
engine is sometimes driven without fuel by a motor-dynamometer to evaluate friction
power. An alternative to using friction power to relate brake and indicated power is
through the engine mechanical efficiency, 0m:

0m = BP/IP [dl] (6.12)

Because of friction, the brake power of an engine is always less than the indicated
power; hence the engine mechanical efficiency must be less than 1. Clearly, mechanical
efficiencies as close to 1 as possible are desired.
The engine indicated power can also be expressed in terms of torque, through
Equation (6.11). Thus an indicated torque, IT, can be defined. Similarly, a brake mean
effective pressure, BMEP, may be defined that, when multiplied by the engine displace-
ment and speed, yields the brake power, analogous to Equation (6.10). Table 6.1
summarizes these and other performance parameters and relations.
The thermal efficiency, as for other engines, is a measure of the fuel economy of a
reciprocating engine. It tells the amount of power output that can be achieved for a
given rate of heat release from the fuel. The rate of energy release is, in turn, the
product of the rate of fuel flow and the fuel heating value. Thus, for a given thermal
efficiency, power output can be increased by employing a high fuel flow rate and/or
selecting a fuel with a high heat of combustion.
If the thermal efficiency is evaluated using the brake power, it is called the brake
thermal efficiency, BTE. If the evaluation uses the indicated power, it is called the
indicated thermal efficiency, ITE.
237

It is common practice in the reciprocating engine field to report engine fuel


economy in terms of a parameter called the specific fuel consumption, SFC, analogous
to the thrust specific fuel consumption used to describe jet engine performance. The
specific fuel consumption is defined as the ratio of the fuel-mass flow rate to the power
output. Typical units are pounds per horsepower-hour or kilograms per kilowatt-hour.
Obviously, good fuel economy is indicated by low values of SFC. The SFC is called
brake specific fuel consumption, BSFC, if it is defined using brake power or indicated
specific fuel consumption, ISFC, when based on indicated power. The SFC for a
reciprocating engine is analogous to the heat rate for a steam power plant in that both
are measures of the rate of energy supplied per unit of power output, and in that low
values of both are desirable.

Volumetric Efficiency

The theoretical energy released during the combustion process is the product of the
mass of fuel contained in the combustion chamber and its heating value if the fuel is
completely reacted. The more air that can be packed into the combustion chamber, the
Table 6.1 Engine Performance Parameters
Indicated Brake Friction

Mean effective pressure IMEP BMEP FMEP = IMEP – BMEP


m = BMEP / IMEP
Power IP BP FP = IP – BP
m = BHP / IHP
Torque IT BT FT = IT – BT
m = BT / IT
Thermal efficiency ITE BTE m = BTE / ITE
Specific fuel consumption ISFC BSFC m = ISFC / BSFC

more fuel that can be burned with it. Thus a measure of the efficiency of the induction
system is of great importance. The volumetric efficiency, 0v, is the ratio of the actual
mass of mixture in the combustion chamber to the mass of mixture that the displace-
ment volume could hold if the mixture were at ambient (free-air) density. Thus the
average mass-flow rate of air through a cylinder is 0v (disp) DaN. Pressure losses
across intake and exhaust valves, combustion-chamber clearance volume, the influence
of hot cylinder walls on mixture density, valve timing, and gas inertia effects all
influence the volumetric efficiency.

EXAMPLE 6.3

A six-cylinder, four-stroke-cycle SI engine operates at 3000 rpm with an indicated


mean effective pressure of five atmospheres using octane fuel with an equivalence ratio
238

of 0.9. The brake torque at this condition is 250 lbf–ft., and the volumetric efficiency is
85%. Each cylinder has a five inch bore and 6 inch stroke. Ambient conditions are
14.7 psia and 40°F. What is the indicated horsepower, brake horsepower, and friction
horsepower; the mechanical efficiency; the fuel flow rate; and the BSFC?

Solution
The six cylinders have a total displacement of

DISP = 6B×52×6/4 = 706.86 in3

Then the indicated horsepower is

IP = MEP×DISP×N /[12×33,000] [lbf /in2][in3][cycles/min]/[in/ft][ft-lbf /HP-min]

= (5)(14.7)(706.86)(3000/2)/[12×33,000] = 196.8 horsepower

The brake horsepower, from Equation (6.11), is:

BP = 2B × 3000 × 250 / 33,000 = 142.8 horsepower

Then the friction power is the difference between the indicated and brake power:

FP = 196.8 – 142.8 = 54 horsepower

and the mechanical efficiency is

0m = 142.8/196.8 = 0.726

The ambient density is

Da = 14.7 × 144/ [53.3 × 500] = 0.0794 lbm /ft3

and the mass flow rate of air to the engine is

ma = 0.85×0.0794×706.86×(3000/2)/1728 = 41.4 lbm /min

For octane the stoichiometric reaction equation is

C8H18 + 12.5O2 + (12.5×3.76)N2 $ 8CO2 + 9H2O + (12.5×3.76)N2

The fuel-air ratio is then

F/A = 0.9×[(8×12) + (18×1)]/[12.5(32 + 3.76×28)] = 0.0598 lbm-fuel /lbm-air


239

The fuel flow rate is

mf = ma (F/A) = 41.4 × 0.0598 = 2.474 lbm /min

The brake specific fuel consumption is

BSFC = 60 mf /BHP = 60×2.474/142.8 = 1.04 lbm /BHP-hr


____________________________________________________________________

6.7 Spark-Ignition Engine Performance

A typical indicator diagram showing intake and exhaust processes, valve actuation,
and spark timing for a four-stroke-cycle SI engine is shown in Figure 6.6. It is assumed
that an appropriate air-fuel mixture is supplied from a carburetor through an intake
manifold to an intake valve, IV, and that the combustion gas is discharged through an
exhaust valve, EV, into an exhaust manifold.
The induction of the air-fuel mixture starts with the opening of the intake valve at
point A just before TC. As the piston sweeps to the right, the mixture is drawn into the
cylinder through the IV. The pressure in the cylinder is somewhat below that in the
intake manifold due to the pressure losses across the intake valve. In order to use the
momentum of the mixture inflow through the valve at the end of the intake stroke to
improve the volumetric efficiency, intake valve closure is delayed to shortly after BC at
point B. Power supplied from inertia of a flywheel (and the other rotating masses in the
engine) drives the piston to the left, compressing and raising the temperature of the
trapped mixture.
The combustion process in a properly operating SI engine is progressive in that the
reaction starts at the spark plug and progresses into the unburned mixture at a finite
speed. Thus the combustion process takes time and cannot be executed instantaneously
as implied by the theoretical cycle. In order for the process to take place as near to TC
as possible, the spark plug is fired at point S. The number of degrees of crank rotation
before TC at which the spark occurs is called the ignition advance. Advances of 10° to
30° are common, depending on speed and load. The spark advance may be controlled
by devices that sense engine speed and intake manifold pressure. Microprocessors are
now used to control spark advance and other functions, based on almost instantaneous
engine performance measurements.
Recalling the slider-crank analysis, we observ that the piston velocity at top center
is momentarily zero as the piston changes direction. Therefore no work can be done at
this point, regardless of the magnitude of the pressure force. Thus, to maximize the
work output, it is desired to have the maximum cylinder pressure occur at about 20°
after TC. Adjustment of the spark advance (in degrees before TC) allows some control
of the combustion process and the timing of peak pressure. For a fixed combustion
duration, the combustion crank-angle interval must increase with engine speed. As a
consequence, the ignition advance must increase with increasing engine speed to
240

maintain optimum timing of the peak pressure.


Following combustion, the piston continues toward bottom center as the high
pressure gases expand and do work on the piston during the power stroke. As the
piston approaches BC, the gases do little work on the piston as its velocity again
approaches zero. As a result, not much work is lost by early opening of the exhaust
valve before BC (at point E) to start the blowdown portion of the exhaust process. It is
expedient to sacrifice a little work during the end of the power stroke in order to
reduce the work needed to overcome an otherwise-high exhaust stroke cylinder
pressure. Inertia of the gas in the cylinder and resistance to flow through the exhaust
valve opening slow the drop of gas pressure in the cylinder after the valve opens.
Thus the gases at point E are at a pressure above the exhaust manifold pressure and,
during blowdown, rush out through the EV at high speed. Following blowdown, gases
remaining in the cylinder are then expelled as the piston returns to TC. They remain
above exhaust manifold pressure until reaching TC because of the flow resistance of the
exhaust valve. The EV closes shortly after TC at point C, terminating the exhaust
process. The period of overlap at TC between the intake valve opening at point A and
exhaust valve closing at point C in Figure 6.6 allows more time for the intake and
exhaust processes at high engine speeds, when about 10 milliseconds may be available
for these processes. At low engine speed and at idling there may be some mixture loss
through the exhaust valve and discharge into the intake manifold during this valve
overlap period.
The combined exhaust and induction processes are seen to form a “pumping loop”
that traverses the p-V diagram in a counterclockwise direction and therefore
241

represents work input rather than work production. The higher the exhaust stroke
pressure and the lower the intake stroke pressure, the greater the area of the pumping
loop and hence the greater the work that must be supplied by the power loop
(clockwise) to compensate. Great attention is therefore paid to valve design and other
engine characteristics that influence the exhaust and induction processes. Volumetric
efficiency is a major parameter that indicates the degree of success of these efforts.

Performance Characteristics

A given ideal Otto-cycle engine produces a certain amount of work per cycle. For such
a cycle, MEP = W/disp is a constant. Equating the power equations (6.9) and (6.11)
shows that the average torque is proportional to MEP and independent of engine
engine speed. Therefore power output for the ideal engine is directly proportional to
the number of cycles executed per unit time, or to engine speed. Thus an Otto engine
has ideal torque and power characteristics, as shown by the solid lines in Figure 6.7.
The characteristics of real engines (represented by the dashed lines) tend to be
similar in nature to the ideal characteristics but suffer from speed-sensitive effects,
particularly at low or high speeds. Torque and power characteristics for a 3.1 liter V6
engine (ref. 9) are shown by the solid lines in Figure 6.8. Note the flatness of the
torque-speed curve and the expected peaking of the power curve at higher speed than
the torque curve. Rather than present graphical characteristics such as this in their
242

brochures, automobile manufacturers usually present only values for the maximum
power and torque and the speeds at which they occur. Engine characteristics such as
those shown in the figure are invaluable to application engineers seeking a suitable
engine for use in a product.

6.8 The Compression-Ignition or Diesel Cycle

The ideal Diesel cycle differs from the Otto cycle in that combustion is at constant
pressure rather than constant volume. The ideal cycle, shown in Figure 6.9, is
commonly implemented in a reciprocating engine in which air is compressed without
fuel from state 1 to state 2. With a typically high compression ratio, state 2 is at a
temperature high enough that fuel will ignite spontaneously when sprayed directly into
the air in the combustion chamber from a high-pressure fuel injection system.
By controlling the fuel injection rate and thus the rate of chemical energy release in
relation to the rate of expansion of the combustion gases after state 2, a constant-
243

pressure process or other energy release pattern may be achieved as in Figure 6.9. For
example, if the energy release rate is high, then pressure may rise, as from 2 to 3', and
if low may fall to 3''. Thus constant-pressure combustion made possible by controlling
the rate of fuel injection into the cyclinder implies the use of a precision fuel injection
system.
Instead of injecting fuel into the high-temperature compressed air, the cycle might
be executed by compression of an air-fuel mixture, with ignition occurring either
spontaneously or at a hot spot in the cylinder near the end of the compression process.
Inconsistency and unpredictability of the start of combustion in this approach, due to
variations in fuel and operating conditions, and to lack of control of the rate of heat
release with the possibility of severe knock, makes the operation of such an engine
unreliable, at the least, and also limits the maximum compression ratio. The Diesel
engine therefore usually employs fuel injection into compressed air rather than
carbureted mixture formation.
In the Air Standard cycle analysis of the Diesel cycle, the heat addition process is at
constant pressure:

q2$3 = cp(T3 – T2) [Btu/lbm | kJ/kg] (6.13)

and, as with the Otto cycle, the closing process is at constant volume:

q4$1 = cv(T1 – T4) [Btu/lbm | kJ/kg] (6.14)


244

The net work and thermal efficiency are then:

w = q2$3 + q4$1

= cp(T3 – T2) + cv(T1 – T4)

= cvT1[k(T3/T1 – T2/T1) + 1 – T4/T1] [Btu/lbm | kJ/kg] (6.15)

0Diesel = w/q2$3 = 1 + q4-$1/q2$3 = 1 + (cv/cp)(T1 – T4)/(T3 – T2)

= 1 – (1/k)(T1/T2)(T4/T1 – 1)/(T3/T2 – 1) [dl] (6.16)

The expressions for the net work and cycle efficiency may be expressed in terms
two parameters, the compression ratio, CR = V1/V2 (as defined earlier in treating the
Otto cycle) and the cutoff ratio, COR = V3/V2. The temperature ratios in Equations
(6.15) and (6.16) may be replaced by these parameters using, for the constant-pressure
process,

COR = V3/V2 = T3/T2

and by expanding the following identity:

T4 /T1 = (T4/T3)(T3/T2)(T2 /T1)

= (V3 /V4)k-1(V3/V2)(V1/V2)k-1

= [(V3/V4)(V1/V2)]k-1COR = (COR)k-1COR

= CORk

where the product of the volume ratios was simplified by recognizing that V4 = V1.
Thus the nondimensionalized net work and Diesel-cycle thermal efficiency are given by

w /cvT1 = kCRk-1(COR – 1) + (1 – CORk) [dl] (6.17)

and

0Diesel = 1 – (1/k)[(CORk – 1)/(COR – 1)]/CRk-1 [dl] (6.18)

where the cutoff ratio, COR, is the ratio of the volume at the end of combustion, V3, to
that at the start of combustion, V2. Thus the cutoff ratio may be thought of as a
measure of the duration of fuel injection, with higher cutoff ratios corresponding to
longer combustion durations.
245

Diesel-cycle net work increases with both compression ratio and cutoff ratio. This is
readily seen graphically from Figure 6.9 in terms of p-V diagram area. As with the
Otto cycle, increasing compression ratio increases the Diesel-cycle thermal efficiency.
Increasing cutoff ratio, however, decreases thermal efficiency. This may be rationalized
by observing from the p-V diagram that much of the additional heat supplied when
injection is continued is rejected at increasingly higher temperatures. Another view is
that heat added late in the expansion process can produce work only over the remaining
part of the stroke and thus adds less to net work than to heat rejection.

EXAMPLE 6.4

A Diesel engine has a compression ratio of 20 and a peak temperature of 3000K. Using
an Air Standard cycle analysis, estimate the work per unit mass of air, the thermal
efficiency, the combustion pressure, and the cutoff ratio.

Solution
Assuming an ambient temperature and pressure of 300K and 1 atmosphere, the
temperature at the end of the compression stroke is

T2 = (300)(20)1.4 – 1 = 994.3K

and the combustion pressure is

p2 = (1)(20)1.4 = 66.3 atm

Then the cutoff ratio is

V3/V2 = T3/T2 = 3000/994.3 = 3.02

The expansion ratio is calculated as follows:

V4 /V3 = (V1/V2)/(V3 /V2) = 20/3.02 = 6.62

T4 = T3 (V3 /V4)1.4 – 1 = 3000/6.620.4 = 1409K

w = 1.005(3000 – 994.3) + (1.005/1.4)(300 – 1409) = 1219.6 kJ/kg

qa = 1.005(3000 – 994.3) = 2015.7 kJ/kg

0th = w/qa = 1219.6/2015.6 = 0.605, or 60.5%


_____________________________________________________________________
246

6.9 Comparing Otto-Cycle and Diesel-Cycle Efficiencies

A reasonable question at this point is: Which cycle is more efficient, the Otto cycle or
the Diesel cycle? Figure 6.10 assists in examining this question. In general notation,
the cycle efficiency may be written as

0th = wnet /qin = wnet /(wnet + |qout|)

= 1 /(1 + |qout| /wnet) [dl] (6.19)

Comparing the Otto cycle 1–2–3–4 and the Diesel cycle with the same compression
ratio 1–2–3'–4, we see that both have the same heat rejection but that the Otto cycle
has the higher net work. Equation (6.19) then shows that, for the same compression
ratio, the Otto cycle has the higher efficiency.
It has been observed that Diesel-cycle efficiency decreases with increasing cutoff
ratio for a given compression ratio. Let us examine the limit of the Diesel-cycle
efficiency for constant CR as COR approaches its minimum value, 1. We may write
Equation (6.18) as

0Diesel = 1 – 1 /(kCRk-1) f (COR)

where f(COR) = (CORk – 1)/(COR – 1). Applying L'Hospital's rule, with primes
247

designating differentiation with respect to COR, to the limit of f(COR) as COR $1,
yields

lim f(COR) = lim (CORk – 1)'/ Lim (COR– 1)' = lim kCORk – 1 = k
COR$1 COR$1 COR$1

and

lim0
COR$1 Diesel
= 1 – 1 /CRk – 1 = 0Otto

Thus the limit of the Diesel-cycle efficiency as COR approaches 1 is the Otto cycle
efficiency. Hence Equation (6.18) shows that the efficiency of the Diesel cycle must be
less than or equal to the Otto-cycle efficiency if both engines have the same
compression ratio, the same conclusion we reached by examination of the p-V diagram.
Suppose, however, that the compression ratios are not the same. Compare the Otto
cycle 1–2'–3'–4 with the Diesel cycle 1–2–3'–4 having the same maximum temperature
in Figure 6.10. The Otto cycle has a smaller area, and therefore less work, than the
Diesel cycle, but the same heat rejection. Equation (6.19) demonstrates that the Otto
cycle has a lower thermal efficiency than the Diesel cycle with the same maximum
temperature.
The conclusion that must be drawn from the above comparisons is quite clear. As in
most comparative engineering studies, the result depends on the ground rules which
were adopted at the start of the study. The Otto cycle is more efficient if the
compression ratio is the same or greater than that of the competing Diesel cycle. But
knock in spark-ignition (Otto) engines limits their compression ratios to about 12,
while Diesel-engine compression ratios may exceed 20. Thus, with these higher
compression ratios, the Air Standard Diesel-cycle efficiency can exceed that of the Otto
cycle. In practice, Diesel engines tend to have higher efficiencies than SI engines
because of higher compression ratios.

6.10 Diesel-Engine Performance

In 1897, five years after Rudolph Diesel's first patents and twenty-one years after
Otto's introduction of the spark-ignition engine, Diesel's compression-ignition engine
was proven to develop 13.1 kilowatts of power with an unprecedented brake thermal
efficiency of 26.2% (ref. 7). At that time, most steam engines operated at thermal
efficiencies below 10 %; and the best gas engines did not perform much better than the
steam machines.
Diesel claimed (and was widely believed) to have developed his engine from the
principles expounded by Carnot. He had developed "the rational engine." Whether his
claims were exaggerated or not, Diesel's acclaim was well deserved. He had developed
an engine that operated at unprecedented temperatures and pressures, had proven his
concept of ignition of fuel by injection into the compressed high-temperature air, and
had overcome the formidable problems of injecting a variety of fuels in appropriate
248

amounts with the precise timing required for satisfactory combustion. His is a
fascinating story of a brilliant and dedicated engineer (refs. 7, 8).
In the Diesel engine, the high air temperatures and pressures prior to combustion
are attributable to the compression of air alone rather than an air-fuel mixture.
Compression of air alone eliminates the possibility of autiognition during compression
and makes high compression ratios possible. However, because of the high pressures
and temperatures, Diesel engines must be designed to be structurally more rugged.
Therefore, they tend to be heavier than SI engines with the same brake power.
The energy release process in the Diesel engine is controlled by the rate of injection
of fuel. After a brief ignition lag, the first fuel injected into the combustion chamber
autoignites and the resulting high gas temperature sustains the combustion of the
remainder of the fuel stream as it enters the combustion chamber. Thus it is evident that
the favorable fuel characteristic of high autoignition temperature for an SI engine is an
unfavorable characteristic for a Diesel engine.
In the Diesel engine, a low autoignition temperature and a short ignition delay are
desirable. Knock is possible in the Diesel engine, but it is due to an entirely different
cause than knock in a spark-ignition engine. If fuel is ignited and burns as rapidly as it
is injected, then smooth, knock-free combustion occurs. If, on the other hand, fuel
accumulates in the cylinder before ignition due to a long ignition lag, an explosion or
detonation occurs, producing a loud Diesel knock. The cetane number is the parameter
that identifies the ignition lag characteristic of a fuel.
The cetane number, like the octane number, is determined by testing in a CFR
engine. The ignition lag of the test fuel is compared with that of a mixture of n-cetane,
C16H34, and heptamethylnonane, HMN (ref. 10). Cetane, which has good ignition
qualities, is assigned a value of 100; and HMN, which has poor knock behavior, a value
of 15. The cetane number is then given by the sum of the percentage of n-cetane and
0.15 times the percentage of HMN in the knock-comparison mixture. A cetane number
of 40 is the minimum allowed for a Diesel fuel.

6.11 Superchargers and Turbochargers

The importance of the volumetric efficiency, representing the efficiency of induction


of the air-fuel mixture into the reciprocating-engine cylinders, was discussed earlier.
Clearly, the more mixture mass in the displacement volume, the more chemical energy
can be released and the more power will be delivered from that volume.
During the Second World War, the mechanical supercharger was sometimes used
with SI aircraft engines to increase the power and operational ceiling of American
airplanes. Today supercharging is used with both Diesel engines and SI engines. The
supercharger is a compressor that supplies air to the cylinder at high pressure so that
the gas density in the cylinder at the start of compression is well above the free-air
density. The piston exhaust gases are allowed to expand freely to the atmosphere
through the exhaust manifold and tailpipe. The supercharger is usually driven by a belt
or gear train from the engine crank shaft.
249

Figure 6.11 shows a modification of the theoretical Otto cycle to accommodate


mechanical supercharging. The supercharger supplies air to the engine cyclinders at
pressure p7 in the intake process 7 $ 1. The processes 4 $ 5 $ 6 purge most of the
combustion gas from the cylinder. The most striking change in the cycle is that the
induction-exhaust loop is now traversed counterclockwise, indicating that the cylinder
is delivering net work during these processes as well as during the compression-
expansion loop. It should be remembered, however, that part of the cycle indicated
power must be used to drive the external supercharger.
The turbosupercharger or turbocharger, for short, is a supercharger driven by a
turbine using the exhaust gas of the reciprocating engine, as shown schematically in
Figure 6.12. A cutaway view of a turbocharger is shown in Figure 6.13(a). Figure
6.13(b) presents a diagram for the turbocharger. Compact turbochargers commonly
increase the brake power of an engine by 30% or more, as shown in Figure 6.8, where
the performance of an engine with and without turbocharging is compared. There, a
substantial increase in peak torque and flattening of the torque-speed curve due to
turbocharging is evident.
For a supercharged engine, the brake power, BP, is the indicated power (as in
Figure 6.11) less the engine friction power and the supercharger shaft power:

BP = DISP ( IMEP ( N – Pm – FP [ft-lbf /min | kJ/s] (6.15)


250

where Pm is the supercharger-shaft mechanical power supplied by the engine (0 for a


turbocharger). The IMEP includes the positive work contribution of the exhaust loop.
The exhaust back pressure of the reciprocating engine is higher with a turbocharger
than for a naturally aspirated or mechanically supercharged engine because of the drop
in exhaust gas pressure through the turbine. The engine brake power increases
primarily because of a higher IMEP due to the added mass of fuel and air in the
cylinder during combustion. Intercooling between the compressor and the intake
manifold may be used to further increase the cylinder charge density. Turbocharging
may increase engine efficiency, but its primary benefit is a substantial increase in brake
power.
In a turbocharged engine, a wastegate may be required to bypass engine exhaust
gas around the turbine at high engine speeds. This becomes necessary when the
compressor raises the intake manifold pressure to excessively high levels, causing
engine knock or threatening component damage. Thirty to forty percent of the exhaust
flow may be bypassed around the turbine at maximum speed and load (ref. 1).
251
252

6.12 The Automobile Engine and Air Pollution

Since the Second World War, concern for environmental pollution has grown from
acceptance of the status quo to recognition and militance of national and international
scope. Among other sources, causes of the well-known Los Angeles smog problem
were identified as hydrocarbons (HC) and oxides of nitrogen (NOx) in exhaust
emissions from motor vehicle reciprocating engines. As a result, national and
California automobile air pollution limits for automobiles have been established and
toughened. Prior to the Clean Air Act of 1990, the U.S. federal exhaust-gas emissions
standards limited unburned hydrocarbons, carbon monoxide, and oxides of nitrogen to
0.41, 3.4, and 1.0 g/mile, respectively. According to reference 12, today it takes 25
autos to emit as much CO and unburned hydrocarbons and 4 to emit as much NOx as a
single car in 1960. The reference anticipated that, led by existing California law and
other factors, future engine designs should be targeted toward satisfying a tailpipe
standard of 0.25, 3.4, 0.4 g/mile. Indeed, the 1990 Clean Air Act (refs. 15,16)
specified these limits for the first 50,000 miles or five years of operation for all
passenger cars manufactured after 1995. In addition to the regulations on gaseous
emissions, the Clean Air Act of 1990 adopted the California standard for particulate
matter of 0.08 g/mile for passenger cars. The standards on particulates are particularly
difficult for the Diesel engine, because of its of soot-producing tendency.
The automobile air pollution problem arises in part because the reactions in the
exhaust system are not in chemical equilibrium as the gas temperature drops. Oxides of
nitrogen, once formed in the cylinder at high temperature, do not return to equilibrium
concentrations of nitrogen and oxygen in the cooling exhaust products. Likewise, CO
formed with rich mixtures or by dissociation of CO2 in the cylinder at high temperature
does not respond rapidly to an infusion of air as its temperature drops in the exhaust
system. Their concentrations may be thought of as constant or frozen. Unburned
hydrocarbons are produced not only by rich combustion but also by unburned mixture
lurking in crevices (such as between piston and cylinder above the top piston ring), by
lubricating oil on cylinder walls and the cylinder head that absorbs and desorbs
hydrocarbons before and after combustion, and by transient operating conditions.
Starting in 1963, positive crankcase ventilation was used in all new cars to duct
fuel-rich crankcase gas previously vented to the atmosphere back into the engine intake
system. Later in the ‘60s, various fixes were adopted to comply with regulation of
tailpipe unburned hydrocarbons and CO, including lowering compression ratios.
In 1973, NOx became federally regulated, and exhaust gas recirculation (EGR) was
employed to reduce NOx formation through reduced combustion temperatures. At the
same time, HC and CO standards were reduced further, leading to the use of the
oxidizing catalytic converter. Introduction of air pumped into the tailpipe provided
additional oxygen to assist in completion of the oxidation reactions. In 1981, a
reducing catalytic converter came into use to reduce NOx further. This device does not
perform well in an oxidizing atmosphere. As a result, two-stage catalytic converters
were applied, with the first stage reducing NOx in a near-stoichiometric mixture and the
253

second oxidizing the combustibles remaining in the exhaust with the help of air
introduced between the stages. This fresh air does not the increase NOx significantly,
because of the relatively low temperature of the exhaust. The three-way catalytic
converter using several exotic metal catalysts to reduce all three of the gaseous
pollutants was also introduced.
The use of catalytic converters to deal with all three pollutants brought about
significant simultaneous reductions in the three major gaseous pollutants from
automobiles. This allowed fuel-economy-reducing modifications that had been
introduced earlier to satisfy emission reduction demands to be eliminated or relaxed,
leading to further improvements in fuel economy.
Catalytic converters, however, require precise control of exhaust gas oxygen to
near-stoichiometric mixtures. The on-board computer has made possible control of
mixture ratio and spark timing in response to censor outputs of intake manifold
pressure, exhaust gas oxygen, engine speed, air flow, and incipient knock. The oxygen,
or lambda, censor located in the exhaust pipe upstream of the three-way converter
or between the two-stage converters is very sensitive to transition from rich to lean
exhaust and allows close computer control of the mixture ratio to ensure proper
operation of the catalytic converter. Computer control of carburetors or fuel injection
as well as other engine functions has allowed simultaneous improvement in fuel
economy and emissions in recent years. Thus, while emissions have been drastically
reduced since 1974, according to reference 11 the EPA composite fuel economy of the
average U.S. passenger car has nearly doubled; although this improvement has not
come from the engine alone. Despite the hard-won gains in emissions control and fuel
economy, further progress may be expected.

EXAMPLE 6.5

The 1990 NOx emissions standard is 0.4 grams per mile. For an automobile burning
stoichiometric octane with a fuel mileage of 30 mpg, what is the maximum tailpipe
concentration of NOx in parts per million? Assume that NOx is represented by NO2 and
that the fuel density is 692 kilograms per cubic meter.

Solution
For the stoichiometric combustion of octane, C8H18, the air-fuel ratio is 15.05 and
the molecular weight of combustion products is 28.6. The consumption of octane is

mf = (692)(1000)(3.79×10-3)/ 30 = 87.4 g/mile

[Note: (kg/m3)(g/kg)(m3/gal)/(mile/gal) = g/mile.] The concentration of NOx is the ratio


of the number of moles of NOx to moles of combustion gas products:

mole Nox /mole cg = (mNOx /mf)(mf / mcg)(Mcg /MNOx)

= (0.4/87.4)(28.6/46)/ (15.05 + 1) = 0.0001773


254

or 177.3 parts per million (ppm).


_____________________________________________________________________

Bibliography and References

1. Heywood, John B., Internal Combustion Engine Fundamentals. New York:


McGraw-Hill, 1988.

2. Ferguson, Colin R., Internal Combustion Engines. New York: Wiley, 1986.

3. Adler, U., et al., Automotive Handbook, 2nd ed. Warrendale, Pa.: Society of
Automotive Engineers., 1986.

4. Lichty, Lester C., Internal Combustion Engines. New York: McGraw Hill, 1951.

5. Crouse, William H., Automotive Engine Design. New York: McGraw-Hill, 1970.

6. Obert, Edward, Internal Combustion Engines, Analysis and Practice. Scranton,


Pa.: International Textbook Co., 1944.

7. Grosser, Morton, Diesel: The Man and the Engine. New York: Atheneum, 1978.

8. Nitske, W. Robert, and Wilson, Charles Morrow, Rudolph Diesel: Pioneer of the
Age of Power. Norman, Okla.: University of Oklahoma Press, 1965.

9. Demmler, Albert W. Jr., et al., “1989 Technical Highlights of Big-three U.S.


Manufacturers,” Automotive Engineering. Vol. 96, No. 10, October 1988, p. 81.

10. Anon., “Ignition Quality of Diesel Fuels by the Cetane Method,” ASTM D 613-84,
1985 Annual Book of ASTM Standards, Section 5.

11. Amann, Charles A., “The Automotive Spark Ignition Engine-A Historical
Perspective,” American Society of Mechanical Engineers, ICE-Vol. 8, Book No.
100294, 1989.

12. Amann, Charles A., “The Automotive Spark-Ignition Engine-A Future


Perspective,” Society of Automotive Engineers Paper 891666, 1989.

13. Amann, Charles A., “The Passenger Car and the Greenhouse Effect,” Society of
Automotive Engineers Paper, 1990.

14. Taylor, Charles Fayette, The Internal Combustion Engine in Theory and Practice,
2nd ed., revised. Cambridge, Mass.: MIT Press, 1985.
255

15. Public Law 101-549, “An Act to Amend the Clean Air Act to Provide for
Attainment and Maintenance of Health, Protection, National Air Quality Standards,
and Other Purposes,” November 15, 1990.

16. Anon., “Provisions–Clean Air Amendments,” Congressional Quarterly, November


24, 1990.

EXERCISES

6.1 Plot dimensionless piston position against crank angle for S/2L = 0.5, 0.4, 0.3,
and 0.2.

6.2* Obtain expressions for the piston velocity and acceleration as a function of the
crank angle, constant angular velocity, and S/2L ratio. Use a spreadsheet to
calculate and plot velocity and acceleration against crank angle for S/2L = 0.5,
0.4, 0.3, and 0.2.

6.3 Determine the equation for the piston motion for a scotch yoke mechanism in
terms of crank angle. Obtain an equation for the piston velocity for a crank that
turns with a given angular velocity, T.

6.4 Derive an equation for the Otto-engine net work by integration of pdV for the Air
Standard cycle. Compare with Equation (6.6).

6.5* Use a spreadsheet to calculate and plot cycle efficiency as a function of


compression ratio for the Diesel cycle for cutoff ratios of 1, 2, and 3. Indentify
the Otto-cycle efficiency on the plot. Explain and show graphically from the plot
how a Diesel engine can be more efficient than an Otto engine.

6.6 A single-cylinder Air Standard Otto engine has a compression ratio of 8.5 and a
peak temperature of 3500°F at ambient conditions of 80°F and one atmosphere.
Determine the cycle efficiency, maximum cylinder pressure, and mean effective
pressure.

6.7 A six-cylinder engine with a compression ratio of 11 runs at 2800 rpm at 80°F
and 14.7 psia. Each cylinder has a bore and stroke of three inches and a
volumetric efficiency of 0.82. Assume an Air Standard, four-stroke Otto cycle

_______________________
* Exercise numbers with an asterisk indicate that they involve computer usage.
256

with stoichiometric octane as fuel. Assume that the energy release from the
fuel is equally divided between internal energy increase in cylinder gases and
cylinder wall heat loss. What are the cylinder mean effective pressures and the
engine horsepower and specific fuel consumption?

6.8 A single-cylinder four-stroke-cycle spark-ignition engine has a BSFC of 0.4


kg/kW-hr and a volumetric efficiency of 78% at a speed of 45 rps. The bore is 6
cm and the stroke is 8.5 cm. What is the fuel flow rate, fuel-air ratio, and brake
torque if the brake power output is 6 kW with ambient conditions of 100 kPa and
22°C?

6.9 A single-cylinder four-stroke-cycle spark-ignition engine operating at 3500 rpm


has a brake mean effective pressure of 1800 kPa and a displacement of 400 cm3.
Atmospheric conditions are 101kPa and 27°C.

(a) If the stroke is 6 cm, what is the bore?


(b) What is the brake power?
(c) If the mass air-fuel ratio is 16 and the fuel flow rate is 0.00065 kg/s, what
is the volumetric efficiency?
(d) Compare your results with the performance of a two-cylinder engine with
the same overall geometric characteristics.

6.10 A four-cylinder four-stroke-cycle spark-ignition engine operating at 3500 rpm has


a brake mean effective pressure of 80 psi and a displacement of 400 cm3.
Atmospheric conditions are one bar and 80°F.

(a) If the stroke is 3 inches, what is the bore?


(b) What is the brake power?
(c) If the mass air-fuel ratio is 16 and the fuel flow rate is 0.00065 kg/s, what
is the volumetric efficiency?

6.11 A four-cylinder four-stroke-cycle spark-ignition engine with 200 cm3


displacement and operating in air at 27°C and 110 kPa has a friction power of 27
kW and a brake power output of 136 kW at 3600 rpm.

(a) What is the mechanical efficiency?


(b) If it has a volumetric efficiency of 74% and burns liquid methanol with
15% excess air, what is the brake specific fuel consumption?

6.12 An eight-cylinder four-stroke-cycle engine has a bore of three inches and a stroke
of 4 inches. At a shaft speed of 3000 rpm, the brake horsepower is 325 and the
mechanical efficiency is 88%. Fuel with a heating value of 19,000 Btu/lbm is
supplied at a rate of 80 lbm/hr. What are the engine displacement, BMEP, brake
257

torque, and indicated specific fuel consumption in lbm/HP-hr?

6.13 An eight-cylinder four-stroke-cycle engine has a bore of 10 cm and a stroke of 12


cm. At a shaft speed of 53 rps, the brake power is 300kW and the mechanical
efficiency is 85%. Fuel with a heating value of 40,000 kJ/kg is supplied at a rate
of 40 kg/hr. What are the engine displacement, BMEP, brake torque, and
indicated specific fuel consumption in kg/kW-hr?

6.14 Consider a naturally aspirated eight-cylinder four-stroke-cycle Diesel engine with


a compression ratio of 20 and cutoff ratio of 2.5. Air is inducted into the cylinder
from the atmosphere at 14.5 psia and 80°F. Assume an Air Standard cycle.

(a) Determine the temperatures and pressures immediately before and after
combustion.
(b) What is the heat added in the combustion process, in Btu/lbm?
(c) What is the net work, in Btu/lbm, and the thermal efficiency?
(d) If the volumetric efficiency is 85%, the engine displacement is 300 in3, and the
engine speed is 2000 rpm, what is the mass flow rate of air through the engine in
lbm/min?
(e) What is the engine horsepower?
(f) Assuming that losses through the valves cause a 20-psi pressure differential
between the pressures during the exhaust and intake strokes, estimate the
actual and fractional losses, in horsepower, due to these processes. Sketch
the appropriate p-V diagram.

6.15 A two-cylinder four-stroke-cycle engine produces 30 brake horsepower at a brake


thermal efficiency of 20% at 2600 rpm. The fuel is methane burning in air with an
equivalence ratio of 0.8 and a heating value of 21,560 Btu/lbm. Ambient
conditions are 520°R and 14.7 psia. The engine mechanical efficiency is 88%,
and the volumetric efficiency is 92%. What are the fuel flow rate, the
displacement volume per cylinder, and the brake specific fuel consumption? What
is the bore if the bore and stroke or equal?

6.16 Sketch carefully a single p–V diagram showing Otto and Diesel cycles having the
same minimum and maximum temperatures. Shade the area representing the
difference in net work between the cycles. Repeat for cycles having the same
compression ratio. Discuss the implications of these sketches.

6.17 An eight-cylinder reciprocating engine has a 3-in. bore and a 4-in. stroke and
runs at 1000 cycles per minute. If the brake horsepower is 120 and the mechanical
efficiency is 80%, estimate the indicated mean effective pressure.
258

6.18 Consider a naturally aspirated eight-cylinder two-stroke-cycle Diesel engine with


a compression ratio of 20 and a cutoff ratio of 2.5. Air is inducted into the
cylinder at 1 atm and 23°C. Assume an Air Standard cycle.

(a) Determine the temperatures and pressures immediately before and after
combustion.
(b) What is the heat added in the combustion process, in kJ/kg?
(c) What is the net work, in kJ/kg, and the thermal efficiency?
(d) If the volumetric efficiency is 85%, the engine displacement is 2500 cc, and
the engine speed is 2200 rpm, what is the mass flow rate of air through the
engine, in lbm per minute?
(e) What is the engine horsepower?
(f) If losses through the valves cause a 120-kPa pressure differential between the
pressures during the exhaust and intake strokes, estimate the actual and
fractional losses, in horsepower, due to these processes. Sketch the
appropriate p–V diagram.

6.19 A twelve-cylinder four-stroke-cycle Diesel engine has a 4-in. bore, a 4.5-in.


stroke, and a compression ratio of 20. The mechanical efficiency is 89%, the
cutoff ratios is 2, and the engine speed is 1200 rpm. The air entering the cylinder
is at 14.5 psia and 60° F. Assuming Air Standard cycle performance, determine
the cycle temperatures, indicated power, IMEP, and engine brake horsepower.

6.20 A hypothetical engine cycle consists of an isentropic compression, a constant-


pressure heat addition, and a constant-volume blowdown, consecutively.

(a) Draw and label a p–V diagram for the cycle.


(b) Use the cyclic integral of the First Law to derive an equation for the cycle net
work in terms of the temperature.
(c) Use the definition of mechanical work to derive an equation for the cycle net
work also. Show that your equation is equivalent to the result obtained in
part (b) using the cyclic integral.
(d) Express an equation for the cycle thermal efficiency in terms of cycle
temperature ratios and k.
(e) If T1 = 60°F and the volume ratio is 10, determine the other cycle
temperatures; and compare the cycle efficiency with the efficiency the Otto
cycle having the same compression ratio.

6.21 A slightly more complex model of a reciprocating engine cycle than those
discussed combines constant-pressure and constant-volume heat additions in a
single Air Standard cycle.

(a) Sketch and label a p–V diagram for this cycle that consists of the following
259

consecutive processes: isentropic compression, constant-volume heat addition,


constant-pressure heat addition, isentropic expansion, and constant-volume
blowdown.
(b)The engine may be characterized by three parameters: the compression ratio;
the Diesel engine cutoff ratio; and a third parameter, the ratio of the pressure
after to that before the constant-volume combustion. Define these parameters in
terms of the symbols in your sketch and derive an equation for the thermal
efficiency of the cycle.
(c) Show how varying the parameters appropriately reduces your efficiency
equation to the equations for the Otto and Diesel cycles.

6.22 As a plant engineer you must recommend whether electric power for a plant
expansion (2.5 MW continuous generation requirement) will be purchased from a
public utility or generated using a fully attended Diesel-engine-driven generator or
an automatic remotely controlled gas turbine generator set. The price of
electricity is 4.8 cents per kW-hr, and the price of natural gas is 60 cents per
thousand cubic feet. Both Diesel engine and gas turbine are to be natural-gas
fired. The gas turbine has a heat rate of 11,500 Btu/kW-hr, and the Diesel engine
13,200 Btu/kW-hr. The initial costs of Diesel engine and gas turbine are
$750,000 and $850,000, respectively. Control equipment for the gas turbine
costs an additional $150,000. The engines and control equipment are estimated
to have a useful life of 20 years. The annual wages and benefits for a Diesel
engine operating engineer working eight-hour daily shifts is $36,000. Assume a
10% per annum interest-rate. Evaluate the alternatives for a natural gas heating
value of 1000 Btu/ft3, and present a table of their costs, in cents per kW-hr.
Discuss your recommendation.

6.23 Evaluate the alternatives in Exercise 6.22 based on the present-worth method.

6.24 In terms of the notation of Figure 6.3, what are the piston displacement,
compression ratio, and expansion ratio for the Lenoir cycle?

6.25 What are the fuel and air flow rates and brake specific fuel consumption for an
eight-cylinder engine with a 3.75-in. bore and 3.5-in. stroke delivering 212
horsepower at 3600 rpm with a brake thermal efficiency of 25%? The fuel is
C8H18, and the equivalence ratio is 1.2. What is the power per cubic inch of
engine displacement?

6.26* Construct a spreadsheet to perform an Air Standard cycle analysis for a Diesel
engine with a compression ratio of 20 and a range of peak temperatures from
1000K to 3000K, in 500° increments. Use it to tabulate and plot both the net work
per unit mass of air and the thermal efficiency against the cutoff ratio.
260

6.27 Determine the maximum tailpipe concentrations of the three federally regulated
gaseous pollutants based on the existing standards for an automobile that achieves
28 mile/gal of iso-octane. Assume that the engine mixture equivalence ratio is
0.9, that NOx is represented by NO2 and unburned hydrocarbons by monatomic
carbon, and that the fuel density is 700 kg/m3.

6.28 A single-cylinder Air Standard Otto engine has a compression ratio of 9.0 and a
peak temperature of 3000°F at 80°F and one atmosphere ambient conditions.
Determine the net work, cycle efficiency, maximum cylinder pressure, and mean
effective pressure.

6.29 A six-cylinder engine with a compression ratio of 11 runs at 3200 rpm and 80°F
and 14.7 psia. Each cylinder has a bore of 3 inches, a stroke of 3.25 inches, and a
volumetric efficiency of 0.85. Assume an Air Standard four-stroke Otto cycle
with stoichiometric octane as fuel. Assume that the energy release from the fuel
is equally divided between internal energy increase in cylinder gases and cylinder
wall heat loss. What are the cylinder mean effective pressures and the engine
horsepower and specific fuel consumption? Assume a heating value of 20,600
Btu/lbm.

6.30 An eight-cylinder four-stroke-cycle compression-ignition engine operates with a


fuel-air ratio of 0.03 at 2400 rpm. It has a turbocharger and intercooler, as
diagrammed nearby, with compressor pressure ratio of 1.7 and intercooler exit
temperature of 320K. The engine bore and stroke are 10 cm and 12 cm,
respectively. The compressor efficiency, turbine efficiency, and volumetric
efficiency are 70%, 75%, and 87%, respectively. The entrance temperature of the
turbine gases is 1000K. What are the compressor power, the turbine pressure
ratio, and the engine power, in kW and in horsepower? Assume that the engine is
constructed of ceramic components that minimize engine heat losses so that they
may be neglected—and ideal “adiabatic” engine.
261

S-ar putea să vă placă și