Review Thermodynamics: 2.1 The First Law of Thermodynamics
Review Thermodynamics: 2.1 The First Law of Thermodynamics
Review of thermodynamics
Turbomachinery and gas turbines may be fully analyzed, to the degree reached in
this text, by rather simple thermodynamic relations. All are based on applications of
the first and second laws of thermodynamics to so-called "flow" systems. From these,
powerful and widely applicable tools such as the energy-enthalpy relations for various
components and the one-dimensional compressible-flow functions can be quickly and
easily derived.
The very simplicity of these tools can lead to their misuse. Serious errors can result
when a relation derived for incompressible flow is used for a high-Mach-number gas
stream, or when an isentropic-flow function is employed to relate upstream to downstream
conditions in frictional flow. Rigor is vital. We shall derive the significant relations from
first principles so that the conditions under which they may individually be used may be
fully understood. This review will also serve to establish a language and a system of
symbols that will be used in the remainder of the book.
4S
46 Ch. 2. Review of thermodynamics
8q = dE + 8w
where
8q(= Oqill) is the heat transferred in the process (positive for heat transfer
into the material);
d EA is the increase of energy level of the material inside the system A ,
termed the "internal" energy; and
ow(= oW ex ) is the work transferred out of the material to the surroundings
(positive for work done by the material on the surroundings).
The change in internal energy, dE A , includes all forms of energy storage, such as
kinetic, gravitational, elastic, electric, magnetic, and thermal. In the following discussion
we will give thermal-energy storage the special symbol du o
Of these quantities only E , the internal energy, is a property of the material. That
is, its value depends only on the state of the material and not on how that state was
reached. The energy transfers Oqill and oWex are not properties, and their values depend
absolutely on the processes used to attain the final state. That is why different symbols
are used for the differentials (0 and d). Later the apparently similar Gibbs ' equation will
be compared with this statement of the first law.
In many engineering examples there may be several identifiably different heat flows
and work transfers, and the material may be composed of parts having different internal
energies. The first law can be applied to the whole mass of material with the various
quantities summed algebraically, or to individual components.
A2
Pst, 2
Ust ,2
v st , 2
®
Control
volume "A"
dV •
C1 ~ =\{=o
dt
Zl
Datum
Figure 2.2. Energy flows for a steady stream of fluid passing through a control
volume
only two ports is not a necessary assumption, and using it results in no loss of generality.
Likewise, without loss of generality we shall consider only one of each heat transfer
rate entering (8qin/dt) and exiting (8qex/dt) the control volume, and one of each power
transfer entering (8Win /dt) and exiting (8w ex /dt) the control volume.
For flow-process control volumes we have to take into account not only heat and
work transfers through the surfaces of the control volume, but also the various categories
of energy that the material takes with it as it passes through the inlets and outlets of
the control volume. One of these categories of energy is kinetic (the effect of high
flow velocity as material passes through the ports). Another category accounts for the
displacement of mass through the port at pressure p . Another category accounts for
potential-energy changes between ports due to changes of elevation z with respect to
an arbitrary elevation level. There are of course other forms of energy, such as surface
tension, electromagnetic forces, etc., which should be included in general formulations,
but are seldom of significant influence in turbomachinery problems.
When applying energy balances to turbomachinery problems it is sometimes conve-
nient to add to the thermal-energy term the kinetic-energy term; and at other times it is
more convenient to add to the thermal-energy term the kinetic-energy and the potential-
energy term, evaluated from an arbitrary elevation level. This leads to the definition
of what are known as "static", "stagnation", and "total" properties, as explained further
below.
We shall henceforth use the subscript st ("static") to denote property values that
do not include the effects of kinetic-energy or potential-energy terms, the subscript 0
("stagnation") to denote property values that include the effects of kinetic energy but
do not include the effects of potential energy, and the subscript T ("total") to denote
property values that include the effects of kinetic energy as well as potential energy. The
48 Ch. 2. Review of thermodynamics
following examples illustrate when it is convenient to use each one of these types of
properties.
At any flow station with appreciable area-averaged flow velocity C perpendicular to
the flow area A, incompressible liquids have (constant) density P and temperature T
(corresponding to static conditions), the mass-flow rate is given by m = pAC, and the
fluid has different values of static, stagnation, and total pressure (Pst, Po and PT) . Perfect
or semi-perfect gases (defined below) have different static and stagnation temperatures,
pressures, and densities (T.,t. To, Pst. Po, Pst and Po), and the mass-flow rate is given by
m= Pst AC.
The energy categories that usually need to be accounted for at the inlets and outlets
of control volumes of turbomachinery problems are:
where
8m is the mass of material passing into or out of the control volume in time
81
C is the velocity of the material
ge is a unit-matching constant in Newton's law (gc - 1 for SI and other
"consistent" systems of units I)
Z is height above a datum
Vst static specific volume (i.e., measured at the speed of the flowing materi-
als), where Vst is the reciprocal of static density Pst
Pst is static pressure
C2
8E=8m ( Ust+PstVs t+ - + -
gz) (2.2)
2gc ge
Accounting for the heat and work flows, and the inlet and outlet flows of system A,
shown in figure 2.2, with energy EA and volume VA in a region where the surrounding
(atmospheric) environmental pressure is Pen at any time, the first law (energy balance)
I In Newton's law. force = (mass x acceleration)lgc , gc == I for SI and other "consistent" systems of units.
gc is useful for unit systems having the same units for mass and weight (at sea level), e.g., pounds-mass and
pounds-force (for which gc=32.IS [Ibm · ft]/[lbf· s2]) and kilograms-mass and kilograms-force (for which
g<.=9.S07 [kgm . m]/[kgf . s2]).
Sec. 2.1. The first law of thermodynamics 49
+-d
dml (
Ust , J
ci +-
+ P st, I Vst , J+ -2 gZI)
t gc gc
-Tt
dm z (
U st , 2
C~ gZZ)
+ P st,Z Vs1, 2 + 2gc + & (2.3)
Equation 2,3 should be used whenever transients are of interest (such as transient
flow through a compressor or turbine, hot-air balloon races, charging and discharging
rigid or non-rigid vessels, etc,). For steady-flow processes m= dm/dt, Q = oq/dt, and
W = ow /dt , In practice turbomachinery components do not change volume with time
(VA = 0), Also, the characteristic response time of the flow in turbomachinery is usually
orders of magnitude shorter than the characteristic time of the overall transient problem
(Korakianitis et al., 1994), so that most transients encountered can be defined as "slow",
That is, most cases may be analyzed by examining a series of steady-state conditions, in
which there is no storage of mass or energy inside the control volume (dmA/dt = 0, or
m] = mz).
Therefore the energy flows into and out of the control volume in steady flow can be
equated, and the steady-flow form of equation 2.3 becomes:
.
Wex
,
+ Qex C~
+ m 2 ( Ust , 2 + P st,ZVst,2 + -2 gZ2)
+-
gc gc
=
. + Qin
"'in
. +m]. (Ust , ] + P st , ]Vst , J + -2
Ci +-
gZ]) (2.4)
gc gc
This is known as the "steady-flow energy equation", (SFEE). If there is only one inlet
and one outlet flow port, as in figure 2,2, then the mass flow rate is m= mJ = mz. The
SFEE has very wide usefulness for the analysis of turbomachinery and associated steady-
flow processes. Some examples of its use are given below. It is convenient to simplify
equation 2.4 by defining the property enthalpy, mostly used for gas flows (compressors,
gas turbines), and stagnation pressure, mostly used for liquid flows (hydraulic machinery).
Enthalpy
The static properties in (ust + P stVsr) are properties of the material at a given point.
Properties, by definition, are functions only of the thermodynamic state of the material.
This uniqueness means that new, unique, properties can be formed from any combination
of other properties. It is obviously convenient to define U st + P s t Vst as one property. It
is given the name "static enthalpy" and the symbol h st :
While pressure and volume can be defined in terms of measurable length and force
quantities, the internal thermal energy, and therefore the enthalpy, can be measured only
50 Ch. 2. Review of thermodynamics
from a defined datum state. Different tables often use different datum states. When using
two sets of data to cover widely differing states, one must take care that they have the
same datum or base state.
The steady-flow energy equation incorporating this "new" property, enthalpy, is
written
gz)
. . . . 2
Qin + Win ~ Qex - Wex = fli (h st + £+ (2.6)
m 2gc gc
where the fl operator implies that the difference of the quantity in parentheses should be
taken from outlet station 2 to inlet station 1.
The enthalpy has the subscript sf because, as mentioned previously, it is the value
for so-called "static" or "stream" conditions. These are the conditions measured with
instruments that are "static" or stationary with respect to the fluid . It is diffic ult to
measure static properties (pressure and temperature) in high-speed flow s. The stagnation
properties are more easily measured by means of upstream-facing "Pi tot" or "stagnation"
probes, shielded to reduce heat transfer by radiation and conduction (figure 2.3). The
steady-flow energy equation can be written for conditions (1) upstream and (2) within
a stagnation probe, assuming that the velocity of flow within the probe is vanishingly
small, that is C2 ~ 0:
(2.9)
We shall henceforth distinguish between stagnation and static conditions, using appro-
priate subscripts, except in those circumstances where all properties are of the same type.
Where motion is slow, for instance, the static properties equal the stagnation properties.
And in dealing with the relationships of a thermodynamic cycle, stagnation conditions
should be used throughout.
Using this definition, we can write the steady-flow energy equation as:
(2.10)
In air and gas turbomachinery, changes of height have a negligible effect on the
enthalpy change, and this general form of the steady-flow energy equation can be ap-
proximated by:
The steady-flow energy equation (SFEE) is usually simplified for most of the com-
ponents of plants incorporating turbomachinery. This is because some are virtually ther-
mally isolated, so that the heat transfer is negligible and the process undergone is adi-
abatic (Qill = Qex = Q = 0), for instance, compressors, pumps, and turbines. In
other turbomachinery components there is considerable heat transfer but no work transfer
(W = 0), for instance, boilers, combustion chambers, and heat exchangers. In yet other
turbomachinery components there is neither heat transfer nor work transfer, as in nozzles,
throttles, diffusers, and ducts. Table 2.1 summarizes how the SFEE, equation 2.4, is used
in components with gas, vapor, or liquid, and how the restricted SFEE, equation 2.11, is
used in components with gas, vapor, or liquid in horizontal flow.
Use of the SFEE is especially convenient for steady-flow devices that have gases
as their working fluids, because over a range of commonly encountered pressures, the
enthalpy of gases at ranges well above their critical temperatures is a function of tem-
perature only. Tabulated data, such as those in Gas Tables by Keenan and Kaye (1983),
W Q ~Tho
Nozzles 0 0 0
Valves 0 0 0
Throttles 0 0 0
Ducts, pipes, diffusers 0 0 0
Boilers 0 Qifl/ m
Heat exchangers 0 Q/m
Combustion chambers 0 Qill/ m
Compressors, pumps, etc. O(?*) Win/fir
Turbines O(?*) Wex /m
• The queries are entered because in special cases there is substantial heat transfer in compressors and turbines.
Compressors may incorporate intercoolers. Turbine blades, disks and duct walls may incorporate cooling
circuits. Reheat combustors can be used between turbine stages or groups of stages.
52 Ch . 2. Review of thermodynamics
give enthalpies for air and other common gases as functions only of temperature for "low
pressures". Comparison with the only high-pressure data available when the tables were
originally published (1948) confirms that the tabulated values are fully valid for pressures
up to 1.4 MPa (200 psia) and that "the characteristics of the isentropic as given by ta-
ble 1 [of the Gas Tables] are suitable for precise work even to pressures in the hundreds
of pounds per square inch" (quoted from the appendix of the 1948 edition). For the
examples and problems in this book we use the simple table of temperature-dependent
properties of air (table A.1) and the correlations of the properties of air and of combustion
products given in appendix A.
(2.14)
For a given left-hand side the losses in a pump result in unwanted internal-energy in-
creases (~U s t) due to friction that reduce the total-pressure increase on the right-hand
side. Similarly for a given total-pressure drop the losses in a water turbine result in
unwanted internal-energy increases (~Ust) due to friction that reduce the work output of
the turbine. Because of the large specific heat capacity of most liquids, these internal-
energy increases result in small but finite temperature increases in both pumps and liquid
turbines.
Sec. 2.1. The first law of thermodynamics 53
Wex =
-.- [(Ho, I + Zl) - (Ho,2 + Z2) ] -g - (U s /,2 - Ust , I) (2.16)
m gc
UPPER RESERVOIR
, , RESERVOIR
,' LOWER , , , ,
Figure 2.4. Hydraulic-turbine heads
54 Ch. 2. Review of thermodynamics
or, in words: [actual specific work] = [ideal specific work] - [losses]. The units of
specific work are in SI units (and British-U.S. units) Joules per kilogram (and British
thermal units per pound-mass) or kilowatts per kilogram per second (and horsepower per
pound-mass per second): kJ/kg (and Btullbm); or kW/(kg/s) (and hp/(lbrn/s)).
Gas-turbine expander
A gas-turbine expander is measured on test to be passing 84 kg/s of combustion
products of known mean specific heat capacity at constant pressure 1,130 J/(kg . K).
What is the shaft power output if the mean inlet temperature is 1,250 °C and the mean
exhaust temperature is 550 °C, both measured with stagnation probes?
For gases, the changes in enthalpy between two temperatures can be calculated from
the change in temperature multiplied by the mean specific heat at constant pressure
between those two temperatures (see equation 2.44 below). Therefore
mt.iho = mCpt.iTo
= 84 kg x 1,130_J _ x 0 , 250 - 550)K
s kg· K
66.44 X 106 J/ s = 66.44 MW
also correct if the power is that at the rotor blading and if the friction losses are ducted
away in oil drains and ventilation ducts, external to the working fluid. When none of
these situations applies, a correction must be made for the additional energy transfer
across the control-volume boundaries.
Intercooler
Find the energy loss in the air passing into an air-compressor intercooler at 50 psia,
650 OR, and leaving at 550 OR. There is a stagnation-pressure loss in the air of 10 psi.
The air mass flow is 10 Ibm/s.
Application of the SFEE, equation 2.11, gives:
states and then release them to interact. Rapidly the assembly would progress from a state
of complete order (uniformity) to a state of maximum disorder. The second law states
simply that this process from order to disorder, from disequilibrium to equilibrium, oc-
curs. (As a simple analogy, consider an assembly of dice, each with faces of six different
colors. At the start, all show the same color uppermost. Then the dice are thrown, one
at a time, in a random sequence. The state of the assembly-the "macro-state"-would
proceed from order to disorder, from maximum disequilibrium to maximum equilibrium.)
We could conceptually quantify the approach to equilibrium by calculating the number
of ways the particles could be arranged among the "microstates" to give the "macrostate"
of the assembly. Let us call this number Q. When the particles are all at the identical
(mean-energy) microstate, there is only one way the particles can be arranged, since they
are all identical, and that number is 1. As equilibrium is approached, the number becomes
extremely large, even for a very small collection of particles (figure 2.5).
If two collections or assemblies of particles, A and B, are allowed to interact, the
new number of ways the combined macrostate C could be arranged is the product of the
number of ways A and B could individually be arranged:
(2.17)
It is more appropriate that this measure of disorder of the combined system should
be the sum, rather than the product, of the "disorders" of the constituents, and that the
numbers be smaller than they would otherwise be for realistic collections of molecules.
We therefore define an "extensive" property (one that is proportional to the mass of the
system) S, which we call entropy, for reasons which will be more obvious later, and
which is proportional to the logarithm of Q:
S == K InQ (2.18)
100 u -_ _ _ _ _ _ _ _ _- - - '
TIME - - -
Therefore
(2.19)
and
(2.20)
It is more rigorous to define entropy as the time average of K lnf:?. This overcomes
to some extent the major objection to this model, which is that it could allow the entropy
of an isolated assembly to decrease occasionally even though the general direction of the
approach to equilibrium is for the entropy to increase. But to decrease through even one
interaction would contravene the second law, which can be restated: The entropy of an
isolated assembly must increase or in the limit remain constant.
Also
Sc = SA + SB (2.23)
dSc = dS A + dS B (2.24)
(2.25)
dUA = -dUB = 8q
Equation 2.25 then simplifies to
(2.26)
where we have used lower-case s for "specific entropy", or entropy per unit mass.
This, then, is the condition for thermal equilibrium. We can define this condition as
one in which the temperatures of the two assemblies are equal. We could in fact define the
quantities in equation 2.26 as thermodynamic temperature. The result, however, would
be the inverse of what we commonly call temperature. Therefore we make the following
definition:
We now incorporate this definition into equation 2.26 and apply it to a second in-
teraction between the assemblies A and B in which we allow the common boundary to
move by an increase in the volume of, say, A and an equal decrease in the volume of B,
while maintaining thermal equilibrium, until pressure equilibrium has been attained:
(2.28)
TB
= -dUB
= -dVB
dS c 0
(2.29)
This condition applies for equality of pressure, and of temperature, for the two as-
semblies. The quantities in equation 2.29 have the dimensions of pressure divided by
temperature and are given that definition:
(2.30)
Sec. 2.3. The second law of thermodynamics 59
Incorporating this definition, equation 2.28 becomes the Gibbs' equation for a simple
substance in the absence of energy storage due to motion, gravity, electricity, magnetism,
and capillarity:
Tds=du+pdv (2.31)
Tds = du + pdv
First law:
oq = du +ow
Gibbs' equation applies to any change, ideal or not: its only restriction is that it must
apply to a simple substance. It is composed entirely of properties or changes in property
values.
The first-law equation applies to a defined collection of material but is otherwise
generally applicable. It includes only one property, duo The quantities oq and ow are
dependent on the type of process undergone.
If one considers a thermodynamically reversible process in simple circumstances, one
can write for ow the work of a reversible process, equal to pdv. Then, by comparison
with Gibbs' equation, oq for a reversible process is Tds, and the equations are identical.
This discussion underscores the essential difference in application of the first and second
laws. The first law makes no condition about the type of process undergone, only about
the end states. The second law is concerned about the process, and allows the end state
to be predicted, given the initial state, or vice versa.
We examine process descriptions in the following derivations from Gibbs' equation.
2Some thennodynamics texts define a perfect gas as one with constant specific-heat capacities, and an ideal
gas as one with heat capacities as functions of temperature. Other thennodynamics texts define perfect and
ideal gases exactly inversely. Both expressions, "perfect" and "ideal", imply perfection in some way. In this
text we define as perfect gas one with constant specific-heat capacities, and as semi-perfect gas (a "less than
perfect" gas) one with specific-heat capacities that are functions of temperature. Equation Cp - Cv = Rand
definition y == CplCv apply to both perfect and semi-perfect gases.
60 Ch. 2. Review of thermodynamics
state:
m
pV = nfiT = - f i T = mRT
Mw
pv RT (2.32)
p = pRT
where n is the number of kmoles of the substance, fi is the "universal gas constant",
Mw is the molecular weight of the gas, and R = fi/ Mw is the "gas constant".
The universal gas constant is:
8, 3l3.219 Jjkmole/K (2.33)
1.98587 Btu/lbmole;oR
= 1,545.32 ft.lbf/lbmolet R
The molecular weight for dry air is Mw = 28.970 kg/kmole. Therefore the gas constant
for dry air is
R 286.96 J/kg/K (2.34)
= 0.068549 Btu/ lbmtR
= 53.32 ft.lbf/lbmt R
The above equation of state dictates that the specific internal energy and specific
enthalpy of perfect and semi-perfect gases also obey the following relations:
(2.35)
(2.36)
Perfect and semi-perfect gases have a value of R == fi/M w . For perfect gases Rand
one other value of either Cp, Cv, or y specify the remaining two using Cv = Cp - R
and y == Cp/Cv. For semi-perfect gases Cp(T) is evaluated from temperature, and R
and Cp(T) are used to evaluate Cv(T) = Cp(T) - Rand y == Cp(T)/Cv(T).
Equations 2.31 and 2.33, and the the relation for the velocity of propagation of a
small pressure wave (a sound wave), a, through a fluid, are used (Keenan, 1970, pp.
96-105) to derive:
(2.39)
Sec. 2.3. The second law of thermodynamics 61
v (~~)s = -p ~ -y RvT
(2.40)
where we have noted that the appropriate temperature is 'Fs/ for the actual speed of sound
a s/·
Models for semi-perfect gases usually give mathematical expressions for Cp as a
function of temperature (Cp(T)). For example such expressions for air and products of
combustion modeled as semi-perfect gases are given in Appendix A. All of the above
expressions are valid for both perfect and semi-perfect gases. For perfect gases one can
evaluate changes between two states as:
(2.41)
(2.42)
For semi-perfect gases the expressions are modified to:
U2 - UI = r
iT,
Tz
CvdT ~ C v(T2 - TI ) (2.43)
dh-vdp = Tds=du+pdv
Tds = CvdT + pdv (2.45)
Therefore the left- and right-hand sides of the above equation respectively lead to:
(2.46)
(2.47)
Equation 2.46 states that the slope of any constant-pressure line (isobar) is a function
only of temperature. (Cp for a perfect gas is constant. For a real gas it is a function
only of temperature for all normal pressures, table A.I.) The constant-pressure lines on
a T - s diagram are therefore identical, have an increasing slope with temperature, and
are displaced horizontally from one another (figure 2.7). They do not "diverge", as is
often stated; the T - s diagram of figure 2.8 (used in some textbooks) is wrong and
misleading. Similarly (equation 2.47) the slope of the constant-volume lines (isochors)
is a function only of temperature, but it is higher than that of the constant pressure lines
(because Cp > C v ).
The apparent virtue of assuming diverging rather than parallel constant-pressure lines
is that the temperature drop in a high-temperature expansion process between two pres-
sures can be larger than the temperature (and therefore the enthalpy) rise in the compres-
sion process between the same pressures. This inexactitude is not necessary, as will be
seen.
cpaT)
(- - -RT (2.49)
ap/p s
Sec. 2.3. The second law of thermodynamics 63
-
IJJ
a:
:::>
~
<t
a:
IJJ
c..
~
IJJ
~
ENTROPY, S
w
a::
:::>
~
<t
a::
w
c..
:E
w
~
ENTROPY,S
" ,
The work required for an isentropic process between two pressure levels is therefore
proportional to the absolute temperature. An isentropic turbine expanding from 1, 500 K
(2, 700 oR, 2,240 OF) would produce five times the work required by an isentropic com-
pressor compressing from 300 K (540 oR, 80 OF) between the same pressure levels.
where ~n and Tex are the local temperatures at which heat transfers dqill and dqex occur,
and dSir is the entropy generated by irreversibility inside the control volume due to
losses.
For steady-flow and quasi-steady flow the corresponding "steady-flow entropy equa-
. +f - f
tion" (SFSE) is:
mS2 . +
Qex = ms) -Qin + S·ir (2.51)
T Tex T Tin
which, in words, states that the entropy transferred out of the control volume via bulk
flow through the ports or heat transfers is equal to the entropy transferred into the control
volume via bulk flow through the ports or heat transfers plus the entropy generated within
the control volume due to irreversibilities (losses).
The relationship between stagnation and static condition at any flow station is by
definition isentropic. Therefore changes in entropy th(S2 - s)) may be evaluated: either
by using stagnation properties at both 1 and 2; or by using static properties at both 1
and 2. For simple substances S2 - s) may be evaluated from tables. The entropy change
for an incompressible substance with constant density p and specific heat capacity C is
given by:
T2
= Cln-
S2 - S, (2.52)
T)
Gibbs' equation 2.31 and the equation of state for perfect and semi-perfect gases
lead to:
dh - vdp dT dp
ds C --R-
T P T P
du + pdv dT dv
ds C v T
- + R v- (2.53)
T
dh - vdp dv dp
ds = = C -+Cv -
T P v P
Sec. 2.4. Examples of the use of the SFEE and the SFSE 65
Therefore for a perfect gas the entropy change between any two states 1 and 2 is given
by anyone of:
T2 P2
S2 - sl C In- - Rln-
p TI PI
S2 - Sl = Cvln-
T2
+ Rln-
V2
(2.54)
TJ VI
S2 - s] = Cpln-
V2
+ Cv ln-
P2
VI PI
For a semi-perfect gas one may use equations 2.53 and integrate, or approximate the
entropy change with equations 2.54 with a suitably averaged value of Cp as described
above. (Cv is given by C v = Cp - R.)
If two states 1 and 2 are connected via an isentropic process, or if the flow between
two flow stations 1 and 2 is isentropic, then S2 - Sl = 0 and equations 2.54 lead to:
R I I
T2 (;~) c = (;~) [( Cpj R) -
p = (~~) [(CpjR) -I]
= l] (2.55)
T\
R R R
Tz
TJ
( ;~ ) C
p
= (~) C
v
= (~~) C
V
T2
TJ
(;~) 7 (;~r-J (~~r-I
= =
Since the relationship between stagnation and static conditions at any flow station is by
definition isentropic, the above equations also relate stagnation to static properties at any
location for perfect gases. For semi-perfect gases the above equation can be used as an
approximation, with average values of Cp and C v evaluated as described above. Therefore
for semi-perfect gases y is also a function of temperature (y = CplCv = Cpl(Cp - R)).
We as" ume that the compressor is adiabatic and receives work Wi". Application of the
SFEE between compressor inlet and outlet, flow station land 2, give:
The maximum stagnation pressure at compre sor outlet is attained by an isentropic com-
pressor, and it is found by applying equation 2.55 between stagnation conditions at compressor
inlet and outlet:
~
PO.2.mx = (TO'l ) R
Po,I To,I
=} Po.2.nu = 449.33 kPa [447.83 kPa]
then the entropy generated by irreversibility in the compressor is given by the top expression
of equation 2.54 applied between stagnation conditions at compressor inlet and outlet:
S·iT / me
. SO. 2 - SO• • = Cp l n-r;-
To. 2
- R l n -PO.l
0. 1 PO. I
=> Sir = 162.5 WfK [162.51 W/K]
The stagnation conditions at the inlet are environmental, bur tbe corresponding static tem-
perature and static pressure are lower than the environmental levels. The tagnation temperature
at the outlet of an adiabatic compressor is determined by the power input into the compressor
and the mass-flow rate, and it is not affected by compressor inefficiencies. Given the amount
of power input into the compressor, the maximum tagnation pressure at the outlet is achieved
by an isentropic compressor. Compressor inefficiencies reduce the maximum attainable pres-
sure at compressor outlet, and generate entropy by irreversibility inside the compressor control
volume. This increases the entropy of the outgoing stream from SO.l to $0.2 (tbe entropy gen-
erated by irreversibil.iry inside the compressor leaves the control volume with the outgoing air
stream).
ho - h st
we obtain
To
(2.57)
Tot
Po
(2.58)
Pst
Po
(2.59)
Pst
Equation 2.57 has a significant difference from equations 2.58 and 2.59. Equation 2.57
is based on the first law only, and on adiabatic (no-heat-transfer), but not necessarily
isentropic (not loss-less) flow. However, equations 2.58 and 2.59 are valid only for
isentropic flows.
m Pst M gcCpTst
A RTot
(1- 1)
m
. JRTsrlgc
Apst
M[I-~J-2 (2.60)
We usually know the stagnation, rather than the static conditions, and we therefore convert
this relation into a more useful form using equations 2.57 and 2.58:
= (2.61)
Sec. 2.5. Compressible-flow functions for a perfect gas 71
y -
M.JY [ 1+ ( - 2 -
1) M
2J-2rY~\)
The gas constant R can be found from the universal gas constant m and the molecular
weight Mw (values given by equations 2.33 and 2.34). A chart of equation 2.61 is shown
in figure 2.9.
Substituting equation 2.57 into equation 2.61 gives:
c (2.62)
= (2.63)
0.7r-----+------r-----+------~----+_----~----~
0 .6 r-------j-----_+_
0 . 51-------t------+~r#_-t------l_----_+_----+----____j
0'3t------t-H---+------j------t-----+------+-------l
0.21-----litj----+---t----I----+---+------1
0.1 r--I---+------+------~----+_----_+_----_I_----___l
t2f----+----+----I----f----+--~~--__l
LOf----+----+--
~ Qef------\---+---=
"-
U
0.6·t----+---+-~H---+_--+--_+--___l
0 .4t----+-~4--_I---+__--+--_+--_I
0.21--~~--_+--_+---l_--+_--_I_--~
Figure 2.10. Velocity function versus Mach number for perfect gases
Equation 2.62 indicates that the left-hand side, a velocity function, is a function of M
and Cp / R. A chart of equation 2.62 is shown in figure 2.10.
The use of the above equations and of figures 2.9 and 2.10 is illustrated in the
following examples.
Sec. 2.5. Compressible-flow functions for a perfect gas 73
gc R To----Ll
y-
ast = (2.64)
~2 + ~-1
Sec. 2.6. Turbomachine-efficiency definitions 75
4. The actual work transfer may be just that in the machine blading (sometimes called
the "internal" or "blading" work, and hence efficiency), or it may also include disk
and seal friction and bearing losses; which of these is to be used should be specified.
Even in scientific papers it is common to identify only the ideal process. Let us
discuss the implications of these aspects.
Ideal processes
The two principal ideal processes used are the isentropic for adiabatic processes
and the isothermal for gas compression when intercooling is (or can be) employed. In
76 Ch . 2 . Review of thermodynamics
addition, a variation of the isentropic efficiency is to take the limiting value at unity
pressure ratio. This is termed the "small-stage" or "polytropic" efficiency, for reasons
that will become obvious later. In machines using liquids, such as pumps and hydraulic
turbines, the ideal isentropic process is also isothermal (equation 2.52). The ideal work
transfer in these various cases is found as follows.
Isentropic process
Consider the SFEE for gas turbomachinery (equation 2.11) for an adiabatic process
(Qin = Qex = 0). If the machine is a compressor, then Wex = 0 and Will > O. If the
machine is a turbine, then W;n =
0 and Wex > O. In either case, the actual and isentropic
processes result in:
(2.66)
(2.67)
where h ~ex , s is defined as the enthalpy after an isentropic process from hO,in and PO,in
to P~ ex ' the defined outlet pressure (to be discussed below). These are the general
expressions.
In the special case of a perfect gas,
®
PO,ex
r
PO, in
where r is the pressure ratio, representing the compressor pressure ratio rc > 1, or the
turbine pressure ratio re < 1. Equation 2.68 illustrates why the authors are keen on the
"(CpjR)" rather than the "y" form of the compressible flow and isentropic equations.
This equation, and others like it, are usually given as
leading users to find a value of Cp appropriate for temperature but to take y as constant
with temperature. This approach leads to errors and inconsistencies in precise calcula-
tions.
Substituting equations 2.66, 2.67, and 2.68 in equations 2.65, the "internal" isentropic
efficiencies for a compressor and turbine, Tis ,c and Tis,e, respectively, are:
The above equations in terms of temperature differences and r are exact for perfect
gases, and approximations (with appropriate values of Cp = Cp ) for semi-perfect gases,
In the special case of an adiabatic liquid machine with equal inlet and outlet elevations
the SFEE (equation 2.4) reduces to:
Isothermal process
For incompressible liquids, equation 2.71 gives the isentropic and isothermal ideal
power required to attain P~ex'
For any simple substance, let T = TO ,in = TO ,ex be the temperature of isothermal
compression from inlet pressure PO,in to outlet pressure P~ex ' Assuming inlet and outlet
are at the same elevation, and starting from the SFEE (equation 2.4),
=
= ~f~ho - T ~f~so == (ho, ex - hO,in) - T(so .ex - SO.in) (2.72)
where we used Qinlrh = T t.s st = T t.so for the isothermal process. Equation 2.72 and
property tables can be used to evaluate the entropy change from inlet to outlet for most
substances.
78 Ch. 2. Review of thermodynamics
The isothermal power required for perfect and semi-perfect gases is obtained by
substituting in the general equation 2.72 ho,ex - hO,in = Cp(To,ex - TO,in) = 0 and ~so
from the first of equations 2.54, where T = TO.in = To. ex :
.. 0
( Win ~ Wex ). = T R In PO,ex (2.73)
m II PO, in
Isentropic,
(H6:ex - HO,in)(glgc)
incompr.
liquid, I1s Will 1m
Isothermal,
general gas,
h~ex - hO,in t:.~:ho - T t:.~~So - W8 31m
Winlm t:.~~ ho - T t:.~~ So
l1it
Isothermal,
perfect gas, Not used for expanders
l1it
Notes: The exponent n' is defined in section 2.6. W g3 indicates group 3 losses as defined in chapters 7, 8, and
9, the E[extemallosses]. 6 7: ho = hO,ex - hO.in and 67:so = SO,ex - SO,in '
t:.ho, 1 = t:.hO,2
>-
Do-
--1
«
I
t-
Z
w
ENTROPY,s
Therefore Ilho.1,s = Ilh o,2,s' But Ilho,2,s > Ilh~.2, s (see equations 2.46 and 2.68). For
the two stages
Ilho, l,s + Ilh~, 2,s
TIs ,c = Ilh 0 , I + Ilh 0,2
Therefore the isentropic efficiency of the two-stage compressor is less than that of the
single-stage compressor.
The losses are indicated by the increase of entropy in the working fluid per unit change
of enthalpy: in other words, by the slope of the process line on the h - s chart. This
slope is a function of the exponent n in pv n = constant, which can be used to represent
the process. (In fact this relation need connect only the end points of the process. The
intermediate points will presumably be close to those given by the relation, but the process
does not have to follow the "polytropic" equation for the following to be true.)
To avoid the influence of pressure ratio on the isentropic efficiency, the limiting value
of the isentropic efficiency for a given polytropic process can be used as the pressure ratio
approaches unity. In steam turbines this efficiency is usually known as the "small-stage"
efficiency. In gas-turbine expanders and compressors, it is known as the "polytropic"
efficiency.
The combination of the polytropic relation, pv n = constant, and the equation of state
for a perfect gas (equation 2.33) leads to
r
= PO.ex
® _ ( PO, ex )n
PO,in PO,in
for a polytropic, and
PO,ex To,ex
r
PO ,in TO, in
Sec. 2.6. Turbomachine-efficiency definitions 81
TO. ex )
( - - -r
(n ;;- 1) - rn' where
, n- 1
n ;;:--.
TO,in - -, n
The actual enthalpy rise (work input) of an adiabatic compressor is
To ,ex )
t!..ho = Cp(To. ex - TO,in) = CpTO,in ( y;-:- - n'
1 = Cp TO ,in(r - 1) (2.75)
O,In
t!..hO ,s = Cp T.o , In
. (r R / Cp - 1) (2.76)
n
(1 + R/ Cp - I 1 + (R/Cp)~ + .. . + ... - 1
T)s, c = (1 + nn' - 1 = 1 + n'~ + .. , + ... - 1
(2.78)
For perfect gases the polytropic efficiencies defined by equations 2.78 lead to the
following useful relations:
TO ,ex
- - = rc
[( ~) ry:J (2.79)
TO ,in
for compressors, and
TO ,ex
- - = re
[( ~) T)p. eJ (2.80)
TO ,in
for expanders (turbines).
(2.81)
82 Ch. 2. Review of thermodynamics
to be obtained directly from the pressure ratio. But which value of P~ex should be used?
It is the value for which I]p is defined. A stagnation-to-static efficiency at the outlet
flange, for instance, defines the useful stagnation pressure at this plane as the actual static
pressure there.
Figure 2.12 shows four equivalent ideal polytropic adiabatic compression processes
for different definitions of useful outlet pressure. All processes have the same inlet To, 1
and outlet TO,2 and therefore the same actual work. Their differing values of n' depend
only on which of the four possibilities is used for p~ 2:
In a compressor the diffuser outlet and the casing flange may provide other locations
for the outlet pressure to be defined. When P~ 2 is defined as a static pressure, the
efficiency obtained is termed a "stagnation-to-static" efficiency, I]os. When P~ 2 is a
stagnation pressure, the efficiency obtained is "stagnation to stagnation", 1]00.
It is probable that turbomachine efficiencies are poorly defined and hence misused
more because of confusion over the definition of P~ 2 than through any other factor. Here
are some comments intended to clear up some of the confusion.
l. There is no "right" definition of efficiency. One may choose to use any definition
one wants. It is of course essential that whichever is chosen is exactly identified.
b::
...J
<[
:r:
I-
z
w
hC\1 t---r'-~C-f--r-----
(TO))
ENTROPY, S
saves some effort in making calculations and in accounting for energy flows and
loss locations. For instance, the appropriate definition of P~ 2 for the efficiency
of a compressor in a turbojet engine is the static pressure at the end of the com-
pressor diffuser, the boundary between the compressor and the combustor, because
the dynamic pressure at the point cannot be used by the combustor. There would
be nothing thermodynamically wrong in choosing to use the stagnation pressure,
but doing so would have undesirable consequences. The combustor would seem
to have a higher pressure drop than under the other definition, because the inlet
pressure would now be listed as the stagnation, instead of the static, pressure. The
compressor designer would then have an incentive to cut back the diffuser, or
maybe to eliminate it altogether, which would further increase the numerical value
of the compressor efficiency and further increase the allocated combustor pressure
drop. At this point we would have gone beyond merely energy accounting. The
actual losses would be increased if the diffuser were eliminated, so that defining the
compressor efficiency in terms of a stagnation outlet pressure could have wholly
undesirable consequences.
Therefore the choice of the "appropriate" value of P~2 is one that must take in-
centives into account.
In the case of the turbine of a turbojet engine, the appropriate P~ 2 to use in the
turbine-efficiency definition is the stagnation pressure upstream of the propulsion
nozzle that produces the jet. Obviously in this case the dynamic pressure in the
stream is useful, and there is no need for an incentive for the turbine designer to
diffuse the leaving flow.
3. Only the ideal work is affected by the choice of definition of useful outlet pressure,
P~2' The actual work is entirely unaffected. One has only one decision to make
with regard to the actual work: whether or not to include the "external" friction
losses (bearings, windage, and seals).
The equation for a polytropic process is pV" = constant. Differentiating and dividing
give
~ dp + dv =0
n p v
Differentiating the equation of state for a perfect gas pv = RT and dividing gives
dp dv dT
-p + -
v
= T-
Gibbs' equation (equation 2.31) with equation 2.35 is
Tds = CpdT - vdp
ds
dT
= Cp
T
(1- R/Cp
n'
) (2.82)
Using the definitions of compressor and expander polytropic efficiencies, equations 2.78,
we can now derive the slope of the polytropic processes in the following simple forms,
for compression and expansion respectively:
(2.83)
These differ from the slope of the constant-pressure lines, given in equation 2.46, by a
constant: the efficiency function, (aT/as)p = (T/Cp)'
The slope of the compression line is positive, and that of the expansion line is negative.
For a polytropic efficiency of 100 percent, the slope becomes infinite: the line is vertical
and is an isentrope (figure 2.13). At a temperature of absolute zero, both lines have zero
slope.
The variation of curvature of the polytropic lines is of little significance for machines
of small pressure ratio. However, for large pressure ratios a misrepresentation of the
shape of these process lines can lead to the drawing of incorrect conclusions.
Entropy, S
Therefore
( ~)
op/p s
= pv = Zc RT (2.84)
The general equation of state, pv = Zc RT, applies to non-perfect gases, with the so-
called compressibility factor, Zc, being a slowly varying function of (reduced) pressure
and temperature. R is the gas constant. Equation 2.84 shows that the isentropic enthalpy
drop between lines of constant pressure increases with absolute temperature as with
perfect gases, for which Zc = 1, and for which the appropriate relation was derived in
equation 2.49.
Therefore, when a non-isentropic adiabatic expansion, such as that from 01 to 02
in figure 2.14, is divided into a number of actual or hypothetical small expansions, the
isentropic enthalpy drop of each (for example, ai, b') will be larger than the corresponding
section (a, b) of the overall isentropic enthalpy drop flho ,s'
That is a' > a, b' > b, and "£flho,s,se > flho ,s , where
r.
~
Q...
-.J l:;h O•s
<!
:r:
I-
z
W
ENTROPY, S
1'}s
l:;h o = ( ~l:;ho,se ) . (~f...ho,s,se)
f...ho,s ~ f...ho,s,se f...ho,s
1'}s 1'}p R h (2.85)
where
1'}p == f...ho,se/ f...ho,s,se, the "small-stage" isentropic efficiency, and
Rh == ~l:;ho,s,se/l:;ho,s, the "reheat factor".
The reheat factor is obviously a function of pressure ratio, being unity for very low
pressure ratios, and of polytropic efficiency, being unity for all pressure ratios when the
polytropic efficiency is unity.
the polytropic efficiency is used, both the single-stage and the multistage machines will
have the same polytropic efficiency of about 91 percent.
The polytropic efficiency has a double advantage in analysis of Brayton power cycles,
particularly when the design pressure ratio is being varied. First, as mentioned, a single
value of polytropic efficiency can be entered for each compression and expansion process,
instead of isentropic efficiencies that vary with pressure ratio. Second, the equations using
polytropic efficiencies are often simpler than those for isentropic efficiencies, leading to
a saving in computation time.
In practice, however, the elimination of the reheat factor by the use of polytropic effi-
ciencies does not eliminate all influence of pressure ratio from efficiency. A compressor
or turbine of high pressure ratio will have higher Mach numbers, in general, and blading
that is relatively longer in the low-pressure region and shorter in the high-pressure area,
all factors leading to increased losses. The variation in maximum efficiency with pressure
ratio is discussed in chapter 3.
In[I-17s,e(l- r R / Cp )]
T/p .e (2,89)
InrR/Cp
88 Ch. 2. Review of thermodynamics
As before, the outlet stagnation pressure, P~ex' must be chosen to be consistent with
the particular definition of efficiency being used.
References
Gyftopoulos, Elias P., and Beretta, Gian Paolo (1991). Thermodynamics: Foundations and Appli-
cations. Macmillan, NY, NY.
Keenan, Joseph H. (1970). Thermodynamics. The MIT Press, Cambridge, MA.
Keenan, Joseph H., Jing Chao, and Joseph Kaye (1983). Gas Tables. Wiley, NY.
Korakianitis, T., Vlachopoulos, N.E., and Zou, D. (1994). Models for the prediction of transients
in regenerative gas turbines with centrifugal impellers. ASME paper 94-GT-361, Transactions
of the ASME, Journal of Engineering for Gas Turbines and Power (in print).
Reynolds, William C. (1968). Thermodynamics. McGraw-Hill, NY, NY.
Problems
1. Does the stagnation temperature of the working fluid rise or fall in passing through a gas-
turbine expander? Why?
2. Does the temperature of the water rise or fall in passing through a water turbine? Why?
3. By how much, and in which direction, does the temperature of water change when it falls
over a waterfall 50 m high? What is the mechanism for this change?
4. Sketch in your qualitative estimates of the variation of stagnation and static enthalpy and
pressure through the intercooled compressor shown in figure P2.4. Some end points are
shown. (Four lines are required.)
5. For the subsonic axial-flow air expander specified, calculate the stagnation and static pressures
and temperatures and the Mach number at the rotor-inlet plane (figure P2.5). Also find the
rotor rotational speed and the nozzle-inlet blade height for constant-outer-diameter blading.
Sketch the form of the complete turbine expansion on an enthalpy-entropy chart. The axial
velocity will be constant at this design point.
I TWO-STAGE
DIFFUSER
~ATER- COOLED I
~ALJis GUIDE
TWO-STAGE
RECEIVER
DIFFUSER
OUTLET
.---'---'li'-
AXIAL - FLOW AXIAL-FLOW
~~~~
ho,ex
PO,ex
PO,in
Pst, in
Figure P2.4
® CD ® @ @
Figure P2.S
90 Ch . 2. Review of thermodynamics
6. Complete the table of efficiencies, table P2.6, for the turbine expander of problem 2.5, for
inlet conditions 1, 500 K and 8 x 105 N/m2. The turbine stagnation enthalpy drop is twice
the mean blade speed squared (f:..ho = 2u~) and the mean blade speed is 500 m/s. The flow
velocity at rotor exit is axial and is 0.728 of the mean blade speed, and it is reduced by 50
percent in the diffuser. The rise in static pressure in the diffuser is 75 percent of the value it
would have if the diffuser flow were isentropic. Also calculate the power, in kW, delivered
by the blading. (Do not count the "external" losses-bearing friction and so forth-here or
in the efficiencies.) The inlet conditions to be used for the efficiencies are the stagnation
conditions at plane N .
Table P2.6
Outlet plane
Efficiency 2 3
'1,,11 0.95
'1"ts
'1p .tt
'1p.ts
7. Given that an axial-compressor stage is a row of rotor blades which act as diffusers, followed
by a row of stator blades which also act as diffusers, explain (from first-law and second-law
considerations) why a high Mach number is necessary if the compressor is to give a high
pressure ratio.
9. Using figure P2.9, calculate the rotor rotational speed, the static temperature and pressure at
nozzle exit, and the (absolute) Mach number at that point, of a radial-inward-flow turbine
expander of nozzle-outlet diameter (surface 1) of 250 mm. As an approximation, take this
as the rotor-inlet diameter also. The nozzle-exit direction of the flow is 75 degrees from the
radial direction, and the axial height of the blade passage is 10 percent of the rotor diameter.
The turbine nozzles are supplied with 2 kg/s of air at 2 bars stagnation pressure and 125 °C
stagnation temperature. The rotor peripheral speed is 90 percent of the tangential velocity of
the air at nozzle exit. The stagnation-pressure losses of the flow through the nozzles are small
enough to be neglected. Sketch the nozzle expansion on a temperature-entropy diagram.
10. Draw lines approximating the changes of stagnation and static enthalpy and pressure through
the compressor test rig shown diagrammatically below (figure P2.IO). The centrifugal com-
pressor is driven by a motor and by the energy-recovery turbine, which expands the flow
Problems 91
Figure P2.9
Turbine
Partly closed
Diffuser
Compressor Diffuser
PO. 1 r----r----~--+_----~r_-~--------r-+_--+_--_4.----_4-----------
Pst.1
Figure P2.10. Stagnation and static pressure and enthalpy through a compressor test rig
back to atmospheric pressure (as static pressure). A throttle valve is used to produce different
back pressures on the compressor.
11. Figure P2.tt shows a block diagram of the compressor test rig of problem 2.10. Your task is
to draw control volumes around each of the three components, and then one around the three
as an assembly, excluding the motor. Calculate and discuss the external energy flows and
changes of enthalpy using the steady-flow energy equation. Do this for two different cases.
In the first case, the throttle valve is wide open, so that there is no loss of stagnation pressure
between the outlet of the compressor diffuser and the inlet of the turbine bell mouth. In the
second case, the throttle reduces the turbine expansion ratio from 10 to 5 to 1.
In both cases the compressor takes in 25 kg/s of air at 288 K and compresses it through
a stagnation-to-static pressure ratio (to diffuser exit) of 10 : 1 and a polytropic efficiency
(same conditions) of 0.80. The turbine takes the flow from the throttle valve and expands
92 Ch. 2. Review of thermodynamics
f---'-'- '---'-'
._____ .L._ ._._CV1 CV3~
C E
-.-.~~--.-.-:
Valve /
.L .
__ . _ . _ _ _ . _ . _ . J
CV2
-~~~~-.-.-.-.-.-
13. Calculate the power output in kW of a radial-inflow turbine for a Diesel-engine turbocharger.
The exhaust gas has a molecular weight of 28 .5 and a mass flow of 0.25 kg/so The gas enters
the turbine at a stagnation temperature of 675 K and a stagnation enthalpy of 695 kJ/kg.
The stagnation pressure at inlet is 280,000 Pa. The rotor is 150 mm diameter and runs at
50, 000 rpm. The gas enters the rotor with a velocity of 400 mls at 70° to the radial direction,
and leaves without swirl at a Mach number of 0.22. The static pressure at diffuser outlet is
100,000 Pa.
14. Suppose that you are a manufacturer of high-speed boats, wanting to specify and purchase the
most effective axial-flow water-pump jet-propulsion units to be connected to the boat's engine
(see figure P2.14). Write down the four factors needed to define turbomachine efficiency.
Then choose your specifications for each factor to enable you to choose the best pump for
your application.
15. Figure P2.15 is a diagram of a test setup of a device with which an inventor wants to
heat buildings under construction. Plot the qualitative variations of four air properties: the
stagnation and static pressures and temperatures. Assume that the motor waste heat (from
inefficiencies) is conducted into the airflow.
16. At what flow angle will the flow become just sonic downstream of the variable-setting-angle
nozzles of a high-pressure turbine? The annulus into which the nozzles discharge has a shroud
diameter of 283 mm, and a hub/shroud diameter ratio of 0.85. The mass flow is 8 kg/s; the
Problems 93
Propulsion
unit
Electric fan
Kerosine
Venturi for air-flow Diffuser combustor
measurement ~~
I
----..
Cold air in
Electric power
I Vanes '
I straightening
I flow
Figure P2.1S. Stagnation and static pressure and temperatures in building-heating device
stagnation temperature and pressure at nozzle exit are 1550 K and 15 bar; R is 286.96 J/
(kg · K); and Cp / R is 4.1. The flow enters the nozzles without swirl, and an impulse turbine
is downstream of the nozzles.
17. Does the steady-flow energy equation apply precisely to a turbine in which the blades are
cooled? Explain.