Fundamentals of Energy Conversion
Fundamentals of Energy Conversion
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 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 called states.
Qualitatively, the concept of state is familiar. For example, the system state of a gas might
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. At a given
7
state, all the properties of a system have fixed values. If the value of even one property
changes, the state will change to a different one.
Thermodynamic Equilibrium
Thermodynamic Properties
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.
Properties that are dependent on mass (or size) are known as extensive properties. For
these properties that indicate quantity, a given property is the sum of 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 of all its
9
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.
Energy Transfer
Energy can cross the boundary of a closed system in two distinct forms: heat and work (Fig.
2.4). It is important to distinguish between these two forms of energy. Therefore, they will
be discussed first, to form a sound basis for the development of the laws of
thermodynamics.
10
Figure 2. 4: Energy can cross the boundaries of a closed system in the form of heat and
work.
Heat
Heat is defined as the form of energy that is transferred between two systems (or a system
and its surroundings) by virtue of a temperature difference. Then it follows that there
cannot be any heat transfer between two systems that are at the same temperature. 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 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 (Fig. 2.4). A system is often approximated as an adiabatic system if it is well
insulated.
There are two ways a process can be adiabatic: Either the system is well insulated so that
only a negligible amount of heat can pass through the boundary, or both the system and the
surroundings are at the same temperature and therefore there is no driving force
(temperature difference) for heat transfer.
11
Figure 2. 5: During an adiabatic process, a system exchanges no heat with its surroundings.
WORK
Work, like heat, is an energy interaction between a system and its surroundings. As
mentioned earlier, energy can cross the boundary of a closed system in the form of heat or
work. Therefore, if the energy crossing the boundary of a closed system is not heat, it must
be work. Heat is easy to recognize: Its driving force is a temperature difference between the
system and its surroundings. Then we can simply say that an energy interaction that is not
caused by a temperature difference between a system and its surroundings is work.
More specifically, work is the energy transfer associated with a force acting through a
distance.
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 2.1(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:
12
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.
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 and work are directional quantities, and thus the complete description of a heat or
work interaction requires the specification of both the magnitude and direction. One way of
doing that is to adopt a sign convention. The generally accepted formal sign convention
for heat and work interactions is as follows: heat transfer to a system and work done by a
system are positive; heat transfer from a system and work done on a system are negative.
Another way is to use the subscripts in and out to indicate direction (Fig. 2.6).
13
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.
It is a common experience that a cup of hot coffee left on the table eventually cools off and
a cold drink eventually warms up. That is, when a body is brought into contact with another
body that is at a different temperature, heat is transferred from the body at higher
temperature to the one at lower temperature until both bodies attain the same temperature
(Fig. 2.7). At that point, the heat transfer stops, and the two bodies are said to have
reached thermal equilibrium. The equality of temperature is the only requirement for
thermal equilibrium.
Figure 2.7: Two bodies reaching thermal equilibrium after being brought into
contact in an isolated enclosure.
The zeroth law of thermodynamics states that if two bodies are in thermal equilibrium
with a third body, they are also in thermal equilibrium with each other. It may seem silly
that such an obvious fact is called one of the basic laws of thermodynamics. However, it
cannot be concluded from the other laws of thermodynamics, and it serves as a basis for
the validity of temperature measurement. By replacing the third body with a thermometer,
the zeroth law can be restated as two bodies are in thermal equilibrium if both have the
same temperature reading even if they are not in contact.
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:
15
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 systems internal energy.
Cyclic Process
A special and important form of the First Law of Thermodynamics is obtained by 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)
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
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
16
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.
Another important form of the First Law of Thermodynamics is the integral of Equation (1.2)
for an arbitrary process involving a system:
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.
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 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 engineers job frequently entails efforts to reduce
irreversibility in machines and processes.
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
17
system, s [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:
where the subscript rev denotes that the heat transfer must be evaluated along a
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.
where h , u and v are, respectively, the system specific enthalpy, specific internal
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
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, 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.
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. The second law states that heat flows naturally from
regions of higher temperature to regions of lower temperature, but that it will not flow
naturally the other way.
EXAMPLE 2.1
20
a) Calculate the entropy change of an infinite sink at 27C temperature due to heat
transfer into the sink of 1000 kJ.
b) Calculate the entropy change of an infinite source at 127C losing the same amount
of heat.
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:
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:
d) A similar approach with the direction of heat flow reversed, taking care to observe
the sign convention, gives
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:
21
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, exceeds
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.
A large number of engineering problems involve mass flow in and out of a system and,
therefore, are modeled as control volumes. A water heater, a car radiator, a turbine, and a
compressor all involve mass flow and should be analyzed as control volumes (open systems)
instead of as control masses (closed systems).
23
24
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, [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.,
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 terms at each port, if needed.
Equation (1.10) may be the most important and frequently used equation in this chapter.
Mastery of its use is therefore essential. It is known as the steady flow form of the First Law
25
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
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.
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
26
the liquid is saturated at state 2. Isobars in the 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, h3h2 , 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)
27
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 T 4 and
the saturated vapor temperature T3, at the same pressure. Thus the degree of superheat of
superheated state 4 is T4 - T3.
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.
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
where s is the entropy per unit mass. Other extensive properties such as enthalpy and
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.
where p [kN/m2], v [m3/kg] and T [K] are pressure, specific volume, and
temperature respectively and 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 (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:
And
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
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
29
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:
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.
30
This and other important relations for an isentropic process in a calorically perfect gas are
summarized as follows
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.
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 [m3/s] at which a fluid flows across a surface is the product of the
area, A [m2], of the surface and the component of velocity normal to the area, V [m/s]. The
corresponding mass flow rate is the ratio of the volume rate and the specific volume, v
[m3/kg]:
31
Alternatively the flow rate can be expressed in terms of the reciprocal of the specific
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.
equation applies:
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
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:
32
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.
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 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:
Care should be taken to assure consistency of units, because enthalpy is usually stated in
thermal units [kJ/kg] and velocity in mechanical units [m /s].
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.
where conservation of mass with steady flow through two ports has been assumed.
Combining this with Equation (1.27), we are led to define another invariant, the stagnation
temperature for a calorically perfect gas:
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.
1.9 Efficiencies
The key word in the above definition is useful. The First Law of Thermodynamics tells us
that energy is conserved in all its transformations. So the ratio of energy output to energy
input is always unity, or 100%.
The meaning of the word useful depends on the purpose of the device. For example, if the
device is an electric heater, the useful energy output is heat, and the energy input is
electricity. Electricity is converted to heat. Heat is also obtained from electricity in a light
35
bulb, as we well know. But this is not the useful energy obtained from a light bulb; the
purpose of a light bulb is to convert electricity into light. Table 4-1 summarizes the useful
energy output and energy input for some common energy conversion devices. Figures 4-2
and 4-3 are illustrations of how to use the information provided in Table 4-1 for the case of
two ubiquitous devices, an electric motor and a furnace. We may know, or may be
interested in knowing, how they work, but this is not necessary for our purposes. For
becoming an energy-informed and (perhaps more importantly) energy-conscious member of
society, all one needs is the information provided in Table 4-1.
36
Second Law of Thermodynamics states that 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. A real
device cannot be 100% efficient.
A cycle during which a net amount of work is produced is called a power cycle. A power
cycle receives heat at a high temperature, converts some of this energy into mechanical
work, and rejects reminder at a lower temperature. By virtue of second law of
thermodynamics, no power cycle can convert more heat into work than the Carnot cycle.
37
The theoretical maximum efficiency of any heat engine is defined by the Carnot Cycle. The
Carnot cycle is a hypothetical engine involving four processes: an adiabatic reversible
(isentropic) compression and expansion and a constant temperature (isothermal) heat
addition and rejection. The Carnot cycle consists of two reversible, isothermal processes
separated by two 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
38
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:
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., a tea-kettle steaming on the stove at constant temperature-
makes it clear that one should not always associate heat transfer with temperature change.
Carnot Theorem 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.
39
Latent energy is the internal energy associated with the phase of a system.
Latent heat is the amount of energy absorbed or released during a phase-change process.
Latent heat
It is the change in heat content of a substance, when its physical state is changed without a
change in temperature.