Chapter 2
Chapter 2
When you can measure what you are speaking about, and express it in numbers, you
know something about it. When you cannot measure it, your knowledge is meager
and unsatisfactory.
Lord Kelvin
The energy balance is based on the postulate of conservation of energy in the universe. This postu-
late is known as the first law of thermodynamics. It is a “law” in the same sense as Newton’s The energy
laws. It is not refuted by experimental observations within a broadly defined range of conditions, balance is also
known as the first
but there is no mathematical proof of its validity. Derived from experimental observation, it quanti- law of
tatively accounts for energy transformations (heat, work, kinetic, potential). We take the first law as thermodynamics.
a starting point, a postulate at the macroscopic level, although the conservation of energy in elastic
collisions does suggest this inference in the absence of radiation. Facility with computation of
energy transformations is a necessary step in developing an understanding of elementary thermody-
namics. The first law relates work, heat, and flow to the internal energy, kinetic energy, and poten-
tial energy of the system. Therefore, we precede the introduction of the first law with discussion of
work and heat.
39
40 Unit I First and Second Laws
6. Calculate ideal gas or liquid properties relative to an ideal gas or liquid reference state,
using the ideal gas law for the vapor phase properties and heats of vaporization.
dW Fapplied dx Fsystem dx
where we have used Newton’s principle of equal and opposite forces acting on a boundary to relate
the applied and system forces. Since it is more convenient to use the system force in calculations, we
use the latter form, and drop the subscript with the understanding that we are calculating the work
done on the system and basing the calculation on the system force. For a constant force, we may
write
W F x
ò
W = – F dx 2.1
For a fluid acting on a surface of constant area A, the system force and pressure are related,
P F/A Þ F P·A
W
EC ò
= – PA dx = – P dV
ò 2.2
In evaluating this expression, a nagging question of perspective comes up. It would be a trivial
question except that it causes major headaches when we later try to keep track of positive and neg-
ative signs. The question is essentially this: In the discussion above, is positive work being done on
the system, or is negative work being done by the system? When we add energy to the system, we
consider it a positive input into the system; therefore, putting work into the system should also be
considered as a positive input. On the other hand, when a system does work, the energy should go
down, and it might be convenient to express work done by the system as positive. The problem is
that both perspectives are equally valid—therefore, the choice is arbitrary. Since various textbooks
choose differently, there is always confusion about sign conventions. The best we can hope for is to
be consistent during our own discussions. We hereby consider work to be positive when per-
formed on the system. Thus, energy put into the system is positive. Because volume decreases
when performing work of compression, the sign on the integral for work is negative,
Section 2.3 Work Associated with Flow 41
W
EC ò
= – P dV (reversible change of system size) 2.3 Expansion/
Contraction work is
associated with a
where P and V are of the system. Clarification of “reversible” is given in Section 2.4 on page 42. By change in system
size.
comparing Eqn. 2.3 with the definitions of work given by Eqns. 2.1 and 2.2, it should be obvious that
the dV term results from expansion/contraction of the boundary of the system. The P results from
the force of the system acting at the boundary. Therefore, to use Eqn. 2.3, the pressure in the inte-
gral is the pressure of the system at the boundary, and the boundary must move. A system which
does not have an expanding/contracting boundary does not have expansion/contraction work.1
out
WS =
ò V dP reversible shaft work, flowing system 2.4
in
Note that Eqns. 2.3 and 2.4 are distinct and should not be interchanged. Eqn. 2.4 is restricted to
shaft work in an open system and Eqn. 2.3 is for expansion/contraction work in a closed system.
We later show how selection of the system boundary in a flow system relates the two types of terms
on page 54.
1. Some texts refer to expansion/contraction work as PV work. This leads to confusion since Section 2.3 shows that work
associated with flow is PV, and the types of work are distinctly different. We have chosen to use the term “expansion/con-
traction” for work involved in moving boundaries to help avoid this ambiguity.
42 Unit I First and Second Laws
Fluid
System
A
Surroundings
Figure 2.1 Schematic illustration of
flow work.
Let us first consider a fluid entering a system as shown in Fig. 2.1. We have dW Fdx, and the
work interaction of the system is positive since we are pushing fluid into the system. The rate of
· ·
work is W = Fx· , but x· is velocity, and F P A. Further rearranging, recognizing Ax· = V , and
that the volumetric flow rate may be related to the mass specific volume and the mass flow rate,
· in
V = Vm· ,
· in · in in
Work associ- W flow = PAx· in = PV = PV in m· 2.5
ated with fluid
in
flowing in/out of
boundaries is called
where PV are the properties of the fluid at the point where it crosses the boundary, and m· is the
flow work. absolute value of the mass flow rate across the boundary. When fluid flows out of the system, work
is done on the surroundings and the work interaction of the system is
· in in out out
W flow = PV m· – PV m· 2.7
in out
where m· and m· are absolute values of the mass flow rates. For more streams, we simply follow
the conventions established, and add inlet streams and subtract outlet streams.
cookie batter. Does the cookie batter become unmixed if you stir in the reverse direction? Of course
not. The shaft work of stirring has been degraded to effect the randomness of the ingredients. It is
impossible to completely recover the work lost in the randomness of this irreversible process. Any
real process involves some degree of stirring or mixing, so lost work cannot be eliminated, but we
can hope to minimize unnecessary losses if we understand the issue properly.
Consider a process involving gas enclosed in a piston and cylinder. Let the piston be oriented
upward so that an expansion of the gas causes the piston to move upward. Suppose that the pressure
in the piston is great enough to cause the piston to move upward when the latch is released. How
can the process be carried out so that the expansion process yields the maximum work? First, we
know that we must eliminate friction to obtain the maximum movement of the piston.
Friction decreases the work available from a process. Frequently we neglect fric- Friction results
tion to perform a calculation of maximum work. in “lost work.”
If we neglect friction, what will happen when we release the latch? The forces are not bal-
anced. Let us take z as our coordinate in the vertical direction, with increasing values in the upward
direction. The forces downward on the piston are the force of atmospheric pressure ( Patm A,
where A is the cross-sectional area of the piston) and the force of gravity ( m g). These forces will
be constant throughout movement of the piston. The upward force is the force exerted by the gas (P
A). Since the forces are not balanced, the piston will begin to accelerate upward (F ma). It will
continue to accelerate until the forces become balanced.2 However, when the forces are balanced,
the piston will have a non-zero velocity. As it continues to move up, the pressure inside the piston
continues to fall, making the upward force due to the inside pressure smaller than the downward
force. This causes the piston to decelerate until it eventually stops. However, when it stops at the
top of the travel, it is still not in equilibrium because the forces are again not balanced. It begins to
move downward. In fact, in the absence of dissipative mechanisms we have set up a perpetual
motion.3 A reversible piston would oscillate continuously between the initial state and the state at
the top of travel. This would not happen in a real system. One phenomenon which we have failed to
consider is viscous dissipation (the effect of viscosity).
Let us consider how velocity gradients dissipate linear motion. Consider two diatomic mole-
cules touching one another which both have exactly the same velocity and are traveling in exactly
the same direction. Suppose that neither is rotating. They will continue to travel in this direction at
the same velocity until they interact with an external body. Now consider the same two molecules
in contact, again moving in exactly the same direction, but one moving slightly faster. Now there is
a velocity gradient. Since they are touching one another, the fact that one is moving a little faster
than the other causes one to begin to rotate clockwise and the other counter-clockwise because of
friction as one tries to move faster than the other. Naturally, the kinetic energy of the molecules will
stay constant, but the directional velocities are being converted to rotational (directionless) ener-
gies. This is an example of viscous dissipation in a shear situation. In the case of the oscillating pis-
ton, the viscous dissipation prevents complete transfer of the internal energy of the gas to the piston
during expansion, resulting in a stroke that is shorter than a reversible stroke. During compression,
viscous dissipation results in a fixed internal energy rise for a shorter stroke than a reversible pro-
cess. In both expansion and compression, the temperature of the gas at the end of each stroke is
higher than it would be for a reversible stroke, and each stroke becomes successively shorter.
2. Two other possibilities exist: 1) The piston may hit a stop before it has finished moving upward, a case that will be con-
sidered below, or; 2) The piston may fly out of the cylinder if the cylinder is too short, and there is no stop.
3. However, this is not a useful perpetual motion machine because the net effect on the surroundings and the piston is zero
at the end of each cycle. If we tried to utilize the motion, we would damp it out.
44 Unit I First and Second Laws
We can see that friction and viscosity play an important role in the loss of capability to perform
useful work in real systems. In our example, these forces cause the oscillations to decrease some-
what with each cycle until the piston comes to rest. Another possibility of motion that might occur
with a piston is interaction with a stop, which limits the travel of the piston. As the piston travels
upward, if it hits the stop, it will have kinetic energy which must be absorbed. In a real system, this
kinetic energy is converted to internal energy of the piston, cylinder, and gas.
Kinetic energy is dissipated to internal energy when objects collide inelastically,
Inelastic colli- such as when a moving piston strikes a stop. Frequently we imagine systems
sions result in lost where the cylinder and piston can neither absorb nor transmit heat; therefore, the
work. lost kinetic energy is returned to the gas as internal energy.
So far, we have identified three dissipative mechanisms. Additional mechanisms are diffusion
along a concentration gradient and heat conduction along a temperature gradient, which will be
discussed in Chapter 4. Velocity, temperature, and concentration gradients are always associated
with losses of work. If we could eliminate them, we could perform maximum work (but it would
require infinite time).
Approaching Reversibility
We can approach reversibility by eliminating gradients within our system. To do this, we can per-
form motion by differential changes in forces, concentrations, temperatures, and so on. Let us con-
sider a piston with a weight on top, at equilibrium. If we slide the weight off to the side, the change
in potential energy of the weight is zero, and the piston rises, so its potential energy increases. If the
piston hits a stop, kinetic energy is dissipated. Now let us subdivide the weight into two portions. If
we move off one-half of the weight, the piston strikes the stop with less kinetic energy than before,
and in addition, we have now raised half of the weight. If we repeat the subdivision again we would
find that we could move increasing amounts of weight by decreasing the weight we initially move
off the piston. In the limit, our weight would become like a pile of sand, and we would remove one
grain at a time. Since the changes in the system are so small, only infinitesimal gradients would
ever develop, and we would approach reversibility. The piston would never develop kinetic energy
which would need to be dissipated.
it must be at equilibrium. Because it is static, we could just as easily go one way as another Þ
“reversible.” Thus, reversible processes are the result of infinitesimal driving forces.
W EC = – P dV
ò
V2 V2 V2
RT RT dV
------ = – RT ln æ ------ö
P = ------- Þ W EC = –
V V1 V
ò
------- dV = – RT
V1 V
ò è V 1ø
(ig) 2.8
V2 RT P P 1-
------ = ---------------------2- = -----1- = ----- (ig) 2.9
V1 RT P 1 P2 10
Qblock 1 = – Q block 2
Heat is transferred at a boundary between the blocks. Therefore, heat is not a property of the system.
It is a form of interaction at the boundary which transfers internal energy. If heat is added to a system
for a finite period of time, then the energy of the system increases because the kinetic energy of the
molecules is increased. When an object feels hot to our touch, it is because the kinetic energy of
molecules is readily transferred to our hand.
Since the rate of heating may vary with time, we must recognize that the total heat flows must
be summed (or integrated) over time. In general, we can represent a differential contribution by
·
dQ = Q dt
We can also relate the internal energy change and heat transfer for either block in a differential
form:
dU ·
dU = dQ or ------- = Q 2.10
dt
An idealized system boundary that has no resistance to heat transfer but is impervious to mass
Diathermal. is called a diathermal wall.
The heat and work transfer necessary for a change in state are dependent on the pathway taken
The work and between the initial and final states. A state property is one that is independent of the pathway taken.
heat transfer neces- For example, when the pressure and temperature of a gas are changed and the gas is returned to its
sary for a change in initial state, the net change in temperature, pressure, and internal energy is zero, and these proper-
state are depen-
dent on the pathway ties are therefore state properties. However, the net work and net heat transfer will not necessarily
taken between the be zero; their values will depend on the path taken. Also, it is helpful to recall that heat and work
initial and final state. are not properties of the system; therefore, they are not state properties.
1
P
2
3
V
Figure 2.2 Schematic for Example 2.2.
3
1.2 moles 8.314cm MPa 298 K - = 14 865 cm 3
V = nR T P = --------------------------------------------------------------------------------------- (ig)
mole K 0.2 MPa
ò
Þ W EC = – P dV = – nRT 1 ln V 2 V 1 (ig) 2.11
3
1.2 moles 8.314 cm MPa 298 K
= ---------------------------------------------------------------------------------------------- ln 2 = – 2060 J
mole K
2. Isobarically cool down to V1:
V1
3
ò
WEC = – P 2 dV = – P 2 V 1 – V 2 = –0.1MPa –14 865 cm = 1487 J 2.12
V2
W = W1 2
+ W2 3 + W3 1 = – nRT1 ln V 2 V 1 + nP 2 V 2 – V 1 = – 573 J (ig) 2.13
Exercise: If we reverse the path, the work will be different; in fact, it will be positive instead of
negative ( 573.6 J). If we change the path to isobarically expand the gas to double the volume
48 Unit I First and Second Laws
(W 2973 J), cool to T1 at constant volume (W 0 J), then isothermally compress to the origi-
nal volume (W 2060 J), the work will be 913 J.
Note: Heat was added and removed during the process of Example 2.2 which has not been
Work and heat accounted for above. The above process transforms work into heat, and all we have done is com-
are path properties. puted the amount of work. The amount of heat is obviously equal in magnitude and opposite in
sign, in accordance with the first law. The important thing to remember is that work is a path func-
tion, not a state function.
Experimentally, scientists discovered that if heat and work are measured for a cyclical process
which returns to the initial state, the heat and work interactions together always sum to zero. This is
an important result! This means that, in non-cyclical processes where the sum of heat and work is
non-zero, the system has stored or released energy, depending on whether the sum is positive or
negative. In fact, by performing enough experiments, scientists decided that the sum of heat and
work interactions in a closed system is the change in energy of the system! To develop the closed-
system energy balance, let us first express the balance in terms of words.
System 1 (Fluid)
System 2 System 3
(Container) (Surroundings)
Piston
·
W EC
4. Other possibilities include electric or magnetic fields, or mechanical springs, etc., which we do not address in this text.
Section 2.7 The Closed-System Energy Balance 49
energy
heat flow + work done
accumulation within = 2.14
into system on system
system boundaries
Energy within the system is composed of the internal energy (e.g., U), and the kinetic (mu2/2gc) and
potential energy (mgz/gc) of the center of mass. For closed systems, the “check list” equation is:
2
v gz
m d U + -------- + ----- = dQ + dW S + dW EC 2.15
2g c g c
The left-hand side summarizes changes occurring within the system boundaries and the right-
hand side summarizes changes due to interactions at the boundaries. It is a recommended prac-
tice to always write the balance in this convention when starting a problem. We will follow this
convention throughout example problems in Chapters 2–4 and relax the practice subsequently.
The kinetic and potential energy of interest in Eqn. 2.15 is for the center of mass, not the random
kinetic and potential energy of molecules about the center of mass. The balance could also be
expressed in terms of molar quantities, but if we do so, we need to introduce molecular weight in
the potential and kinetic energy terms. Since the mass is constant in a closed system, we may divide
the above equation by m,
2
v gz
d U + -------- + ----- = dQ + dW S + dW EC 2.16
2g c gc Closed-system
balance. The left-
hand side summa-
rizes changes inside
where heat and work interactions are summed for multiple interactions at the boundaries. We can the boundaries, and
integrate Eqn. 2.16 to obtain the right-hand side
summarizes interac-
tions at the bound-
æ 2
v gzö aries.
ç U + -------- + -----÷ = Q + W S + W EC 2.17
è 2gc gcø
2 · · ·
d v gz
----- U + -------
- + ----- = Q + W S + W EC 2.18
dt 2gc gc
· · ·
where Q = dQ dt , W S = dW S dt , and W EC = dW EC dt . Frequently, the kinetic and poten-
tial energy changes are small (as we will show in Example 2.9), in a closed system shaft work is not
common, and the balance simplifies to
dU · ·
------- = Q + W EC or U = Q + W EC 2.19
dt
.
50 Unit I First and Second Laws
Solution: First choose a system boundary. Let us initially place system boundaries around each
of the blocks. Let the warm block be block1 and the cold block be block2. Next, eliminate terms
which are zero or are not important. The problem statement says nothing about changes in posi-
tion or velocity of the blocks, so these terms can be eliminated from the balance. There is no
shaft involved, so shaft work can be eliminated. The problem statement doesn’t specify the pres-
sure, so it is common to assume that the process is at a constant atmospheric pressure of 0.101
MPa. The cold block does expand slightly when it is warmed, and the warm block will contract;
however, since we are dealing with solids, the work interaction is so small that it can be
neglected. For example, the blocks together would have to change 10 cm3 at 0.101 MPa to equal
1 J out of the 2000 J that are transferred.
2
v gz
d U + -------- + ----- = dQ + dW S + dW EC
2g c g c
The magnitude of the heat transfer between the blocks is the same since no heat is transferred to
the surroundings, but how about the signs? Let’s explore that further. Now, placing the system
boundary around both blocks, the energy balance becomes:
2
v gz
d U + -------- + ----- = dQ + dW S + dW EC
2g c g c
Note that the composite system is an isolated system since all heat and work interactions across
the boundary are negligible. Therefore, U = 0 or by dividing in subsystems,
U block1 + Ublock2 = 0 which becomes Ublock1 = – U block2 . Notice that the signs are
important in keeping track of which system is giving up heat and which system is gaining heat.
In this example, it would be easy to keep track, but other problems will be more complicated,
and it is best to develop a good bookkeeping practice of watching the signs. In this example the
heat transfer for the initially hot system will be negative, and the heat transfer for the other sys-
tem will be positive. Therefore, the internal energy changes are U block1 = – 2000 J and
U block2 = 2000 J.
Section 2.8 The Open-System, Steady-State Balance 51
Although very simple, this example has illustrated several important points.
1. Before simplifying the energy balance, the boundary should be clearly described by a state-
ment and/or a sketch.
2. A system can be subdivided into subsystems. The composite system above is isolated, but
the subsystems are not. Many times, problems are more easily solved, or insight is gained
by looking at the overall system. If the subsystem balances look difficult to solve, try an
overall balance. Four important
3. Positive and negative energy signs are important to use carefully. points about solv-
ing problems.
4. Simplifications can be made when some terms are small relative to other terms. Calcula-
tion of the expansion contraction work for the solids is certainly possible above, but it has a
negligible contribution. However, if the two subsystems had included gases, then this sim-
plification would have not been reasonable.
1. All state properties throughout the system are invariant with respect to time. The properties
may vary with respect to position within the system.
2. The system has constant mass, that is, the total inlet mass flow rate equals the total outlet
mass flow rates, and all flow rates are invariant with respect to time.
3. The center of mass for the system is fixed in space. (This restriction is not strictly required,
but will be used throughout this text.)
To begin, we write the balance in words, by adding flow to our previous closed-system bal-
ance. There are only three ways the surroundings can interact with the system: flow, heat, and
work. A schematic of an open steady-state system is shown in Fig. 2.4. In consideration of the types Steady-state
of work encountered in steady-state flow, recognize that expansion/contraction work is rarely flow systems are
usually fixed size,
involved, so this term is omitted at this preliminary stage. This is because we typically apply the
so WEC 0.
steady-state balance to systems of rigid mechanical equipment, and there is no change in the size of
the system. Therefore, the expansion/contraction work term is set to 0.
Adiabatic wall ·
System boundary Q
Heat-conducting wall
·
WS
System 1 (Fluid)
m· in = m· out m· out = m· in
System 2 System 3
(Container) (Surroundings)
The balance in words becomes time-dependent since we work with flow rates:
Again, we follow the convention that the left-hand side quantifies changes inside our system.
Consider the change of energy inside the system boundary given by the left-hand side of the equa-
tion. Due to the restrictions placed on the system by steady-state, there is no accumulation of energy
within the system boundaries, so the left-hand side of Eqn. 2.20 becomes 0.
As a result,
2 in 2 out
v - + gz in v - + gz out · · ·
m· – m·
0 =
å U + -------
2g c g c
-----
å U + -------
2g c g c
----- + Q + W S + W flow 2.21
inlets outlets
where heat and work interactions are summed over all boundaries. The flow work from Eqn. 2.7
may be inserted and summed over all inlets and outlets,
2 in 2 out
v gz in v gz out · ·
0 = å U + -------- + ----- m· – å U + -------- + ----- m· + Q + WS 2.22
2g c g c 2g c g c
inlets outlets
in in out out
m· – m·
+
å PV
å PV
inlets outlets
2 in 2 out
v gz in v gz out · ·
m· – m·
0 =
å U + PV + -------- + -----
2gc g c å U + PV + -------- + -----
2g c g c
+ Q + WS 2.23
inlets outlets
Enthalpy
Enthalpy is a Note that the quantity (U PV) arises quite naturally in the analysis of flow systems. Flow systems
mathematical prop- are very common, so it makes sense to define a single symbol that denotes this quantity:
erty defined for con-
venience in problem
solving.
H U PV
Thus, we can tabulate precalculated values of H and save steps in calculations for flow systems. We
call H the enthalpy.
Section 2.8 The Open-System, Steady-State Balance 53
2 in 2 out
v gz in v gz out · ·
m· – m·
0 =
å H + -------- + -----
2gc g c å H + -------- + -----
2g c g c
+ Q + WS 2.24 Open-system,
steady-state
inlets outlets
balance.
where the heat and work interactions are summations of the individual heat and work interactions
over all boundaries.
Note: Q is positive when the system gains heat energy; W is positive when the system
in out in out
gains work energy; m· and m· are always positive; and m· system = m· – m·
is positive when the systems gains mass and zero for steady-state flow. Mass may be
replaced with moles in a non-reactive system with appropriate care for unit conversion.
Note that the relevant potential and kinetic energies are for the fluid entering and leaving the
boundaries, not for the fluid which is inside the system boundaries. When only one inlet and one
outlet stream are involved, the steady-state flow rates must be equal, and
2 in 2 out
v gz in v gz out · ·
0 = H + -------- + ----- m· – H + -------- + ----- m· + Q + WS (one inlet/outlet) 2.25
2g c g c 2g c g c Several com-
mon ways the
steady-state bal-
ance can be written.
When kinetic and potential energy changes are negligible, we may write
· ·
0 = – Hm· + Q + W S (one inlet/outlet) 2.26
out in
where H = H – H . We could use molar flow rates for Eqns. 2.24 through 2.26 with the
usual care for unit conversions of kinetic and potential energy. For an open steady-state system
meeting the restrictions of Eqn. 2.26, we may divide through by the mass flow rate to find
In common usage, it is traditional to relax the convention of keeping only system properties on
the left side of the equation. More simply we often write:
Compare Eqns. 2.19 and 2.28. Energy and enthalpy do not come from different energy bal-
ances, where the “closed system” balance uses U and WEC and the “open system” balance uses H
and Ws. Rather, the terms result from logically simplifying the generalized energy balance shown in
the next section.
Comment on Notation
In a closed system we use the symbol to denote the change of a property from initial state to final
state. In an open, steady-state system, the left-hand side of the energy balance is zero. Therefore,
we frequently write as a shorthand notation to combine the first two flow terms on the right-hand Explanation of
side of the balance, with the symbol meaning “outlet relative to inlet” as shown above. You need to the use of .
learn to recognize which terms of the energy balance are zero or insignificant for a particular prob-
54 Unit I First and Second Laws
lem, whether a solution is for a closed or open system, and whether the symbol denotes “outlet
relative to inlet” or “final relative to initial.”
In the conception of a closed-system fluid packet, no mass moves across the system boundary.
The system, as we have chosen it, does not include a shaft even though it will move past the shaft.
If you have trouble seeing this, remember that the system boundaries are defined by the conceived
packet of mass. Since the system boundary does not contain the shaft before the packet enters, or
after the packet exits, it cannot contain the shaft as it moves through the turbine. The system simply
deforms to envelope the shaft. Therefore, all work for this closed system is technically expansion/
contraction work; the closed-system expansion/contraction work is composed of the flow work and
shaft work that we have seen from the open-system perspective. It is difficult to describe exactly
what happens to the system at every point, but we can say something about how it begins and how
it ends. This observation leads to what is called an integral method of analysis.
System: closed, adiabatic; Basis: packet of mass m. The kinetic and potential energy changes
are negligible:
2
v g
d U + -------- + ----- z = dQ + dW S + dW EC
Note that this
2g c g c
derivation neglects
kinetic and poten-
tial energy changes. Integrating from the inlet (initial) state to the outlet (final) state:
out in
U –U = W EC
We may change the form of the integral representing work via integration by parts:
out out
out
W EC = –
òin P dV = – PV in +
òin V dP
We recognize the term PV as representing the work done by the flowing fluid entering and leaving
the system; it does not contribute to the work of the device. Therefore, the work interaction with the
out
turbine is the remaining integral, W S = òin V dP . Substitution gives,
out
out in
Þ U + PV – U + PV =
òin V dP 2.29
Section 2.8 The Open-System, Steady-State Balance 55
out in
H –H = Q + WS 2.30
Recalling that H U PV and comparing the last two equations means Ws òVdP is the work done
using the pump, compressor, or turbine as the system. Furthermore, the appearance of the PV con-
tribution in combination with U occurs naturally as part of the integration by parts. Physically,
work is always “force times distance.” Though this derivation has been restricted to an adiabatic
device, the result is general to devices including heat transfer as we show later in Section 5.7.
out
H = WS =
òin V dP reversible shaft work 2.31 Shaft work for a
pump or turbine
where kinetic and
Note: The shaft work given by dWS VdP is distinct from expansion/contraction potential energy
work, dWEC PdV. Moreover, both are distinct from flow work, dWflow PVdm. changes are small.
Several practical issues may be considered in light of Eqn. 2.31. First, the work done on the
system is negative when the pressure change is negative, as in proceeding through a turbine or
expander. This is consistent with our sign convention. Second, when considering gas flow, the inte-
gration may seem daunting if an ideal gas is not involved because of the complicated manner that V
changes with T and P. Rather, for gases, we can frequently work with the enthalpy for a given state
change. The enthalpy values for a state change read from a table or chart lead to Ws directly using
Eqn. 2.30. For liquids, however, the integral can be evaluated quickly. Volume can often be approx- Shaft work for a
imated as constant, especially when Tr < 0.75. In that case, we obtain by integration an equation for liquid pump or
turbine where
estimating pump work: kinetic and
potential energy
L out in P changes are small
H = WS V P –P = ------- liquid 2.32 and Tr < 0.75 so
that the fluid is
incompressible.
A convenient way of converting units for these calculations is to multiply and divide by the gas
constant, noting its different units. This shortcut is especially convenient in this case, e.g.,
The saturation enthalpy is read from the saturation tables as 83.95 kJ/kg. The values given in the
compressed liquid table (at the end of the steam tables) are 88.6 kJ/kg at 5 MPa and 130 kJ/kg at 50
MPa, corresponding to estimated work values of 4.65 and 46.1 kJ/kg. The estimation error in the
computed work is about 7 to 9%, and smaller for lower pressures. This degree of precision is gen-
erally satisfactory because the pump work itself is usually small relative to other work and terms
(like the work produced by a turbine in a power cycle).
2 2 in 2 out
d æ v gzö v gz in v gz out
Complete ----- m ç U + -------
- + -----÷ = å H + -------- + ----- m· – å H + -------- + ----- m· 2.33
energy balance. dt è 2g c g cø 2g c g c 2g c g c
inlets outlets
· · ·
+ Q + W EC + W S
Adiabatic wall ·
Q
System boundary ·
Heat-conducting wall WS
System 1 (Fluid)
m· in m· out
System 2 System 3
(Container) (Surroundings)
Piston
·
W EC
where the heat and work interactions are summations of the individual heat and work interactions
over all boundaries. We also may write this with the time dependence implied:
2 2 in 2 out
æ v gzö v gz in v gz out
d m ç U + -------- + -----÷
è 2g c g cø
= å H + -------- + -----
2g c g c
dm – å H + -------- + -----
2gc g c
dm 2.34
inlets outlets
+ dQ + dW EC + dW S
Note: The signs and conventions are the same as presented following Eqn. 2.24.
Usually, the closed-system or the steady-state equations are sufficient by themselves. But for
unsteady-state open systems, the entire equation must be considered. Fortunately, even when the
entire energy balance is applied, some of the terms are usually not necessary for a given problem,
so fewer terms are usually needed than shown in Eqn. 2.33. An objective of this text is to build your
ability to recognize which terms apply to a given problem.
U-ö
æ ------
CV 2.35
è Tø V Definition
of CV.
Since temperature changes are easily measured, internal energy changes can be calculated once CV
is known. CV is not commonly tabulated, but, as shown below, it can be easily determined from the
constant pressure heat capacity, which is commonly available.
Hö
æ ------
Definition CP - 2.36
of CP.
è Tø P
The use of two heat capacities, CV and CP, forces us to think of constant volume or constant
pressure as the important distinction between these two quantities. The important quantities are
really internal energy versus enthalpy. You simply must convince yourself to remember that CV
refers to changes in U at constant volume, and CP refers to changes in H at constant pressure.
For liquids or solids, we typically calculate H and correct the calculation if necessary as
explained below. For liquids, it has been experimentally determined that internal energy is only
very weakly dependent on pressure below Tr 0.75. In addition, the molar volume is insensitive to
pressure below Tr 0.75. We demonstrated in Example 2.4 that,
T2
Liquid below Tr 0.75 or solid: reason-
H
òT 1
C P T dT + V P able approximation. 2.42
Note: These formulas do not account for phase changes which may occur.
Note that the heat capacity of a monatomic ideal gas can be obtained by differentiating the
internal energy as given in Chapter 1, resulting in CV = 3/2 R and CP 5/2 R. Heat capacities for
diatomics are larger, CP 7/2 R, and CV 5/2 R near room temperature, and polyatomics are larger
still. According to classical theory, each degree of freedom5 contributes 1/2R to CV. Kinetic and
potential energy each contribute a degree of freedom in each dimension. A monatomic ideal gas has
only three kinetic energy degrees of freedom, thus CV = 3/2 R. Diatomic molecules are linear so
they have two additional degrees of freedom for the linear (one-dimensional) bond that has kinetic
and potential energy both. In complicated molecules, the vibrations are characterized by modes.
See the end flap to make a quick comparison. Monatomic solids have three degrees of freedom
each for kinetic and vibrational energy, one for each principle direction, thus the law of Dulong and
Petit, CVS = 3R is a first approximation. Low-temperature heat capacities of monatomic solids are
explored more in Example 6.8. If you become curious about the manner in which the heat capaci-
ties of polyatomic species differ from those of the spherical molecules discussed in Chapter 1, you
will find introductions to statistical thermodynamics explain the contributions of translation, rota-
tion, and vibration. For polyatomic molecules, the heat capacity increases with molecular weight Whenever we
due to the increased number of degrees of freedom for each bond. In this text, ideal gas heat capac- assume heat capac-
ity to be tempera-
ity values at 298 K are summarized inside the back cover of the book, and may be assumed to be ture independent in
independent of temperature over small temperature ranges near room temperature. The increase in this text, we mark
heat capacity with temperature for diatomics and polyatomics is dominated by the vibrational con- the equation with a
(*) symbol near the
tribution. The treatment of heat capacity by statistical thermodynamics is particularly interesting right margin.
because it is a theory6 that often gives more accurate results than experimental calorimetric mea-
surements. Commonly, engineers correlate ideal gas heat capacities with expressions like polyno-
mials.
5. The degrees of freedom discussed here are different from those discussed for the Gibbs phase rule.
6. This theory also requires experimental spectroscopic measurements, but those are quite different from the calorimetric
measurement of enthalpy changes with respect to temperature.
60 Unit I First and Second Laws
We will frequently ignore the heat capacity dependence on T to make an approximate calcula-
tion. Whenever we assume heat capacity to be temperature independent in this text, we mark the
equation with a (*) symbol near the right margin. Heat capacities represented as polynomials of
temperature are available in Appendix E. The heat capacity depends on the state of aggregation.
The heat For example, water has a different heat capacity when solid (ice), liquid, or vapor (steam). The con-
capacity of a sub- tribution of the heat capacity integral to the energy balance is frequently termed the sensible heat
stance depends
on the state of
to communicate its contribution relative to latent heat (due to phase changes) or heat of reaction
aggregation. to be discussed later. Note that these are called “heats” even though they are enthalpy changes.
Solution: The ideal gas change is calculated via Eqn. 2.41 and is independent of pressure. The
heat capacity constants are obtained from Appendix E.
T2 T2
2 3
H2ig H1ig òT 1
C P dT = òT 1
A + BT + CT + DT dT
B 2 2 C 3 3 D 4 4
A(T2 T1) --- T 2 – T 1 + ---- T 2 – T 1 + ---- T 2 – T 1
2 3 4
0.3063
---------------- (463.152 378.152)
4.224(463.15 378.15)
2
–4 –8
– 1.586 10 - 3.215 10 -
------------------------------ (463.153 378.153) --------------------------- (463.154 378.154) 8405 J/mol
3 4
This is a com- Solution: This is a common calculation needed for working with power plant condensate
mon calculation streams at high pressure. The relevant equation is Eqn. 2.42, but we can eliminate the tempera-
needed for working ture integral by selecting saturated water at the same temperature and then applying the pressure
with power plant
condensate streams correction, i.e., applying Eqn. 2.39, H V P relative to the saturation condition, giving H =
at high pressure. Hsat + V P. The numerical calculations have already been done in Example 2.4 on page 55. Both
calculations use the same approximation, even though the paths are slightly different. A more
rigorous analysis is shown later in Example 6.1.
Section 2.10 Internal Energy, Enthalpy, and Heat Capacities 61
Solution: System is the gas. Closed system, system size changes, adiabatic.
2
v g
d U + -------- + ----- z = dQ + dW S + dW EC
2g c g c
ò dU = ò dWEC = WEC
dU = C V dT Þ W EC =
ò CV dT = CV T = CP – R T (*ig)
5
= æ ---ö 8.314 200 = 4157 J/mol
è 2ø
Note that because the temperature rise is specified, we do not need to know if the process was
reversible.
7. http://webbook.nist.gov/
8. Perry, R.M., Green, D.W., 2008. Chemical Engineer’s Handbook, 8th ed., New York: McGraw-Hill.
9. Poling, B.E., Prausnitz, J.M., O’Connell, J.P., 2001. The Properties of Gases and Liquids, 5th ed., New York: McGraw-
Hill.
62 Unit I First and Second Laws
along the entire saturation curve. Complete property tables for some other compounds are available in
the literature or online, however, most textbooks present charts to conserve space, and we follow that
trend. In the cases where tables or charts are available, their use is preferred for phase transitions
away from the normal boiling point, although a hypothetical path that passes through the normal
boiling point can usually be constructed easily.
The energy of vaporization is more difficult to find than the enthalpy of vaporization. It can
be calculated from the enthalpy of vaporization and the P-V-T properties. Since U H PV,
Far from the critical point, the molar volume of the vapor is much larger than the molar volume of
the liquid. Further, at the normal boiling point (the saturation temperature at 1.01325 bar), the ideal
gas law is often a good approximation for the vapor volume,
where Tr is reduced temperature Tr = T/Tc, is the acentric factor (to be described in Chapter 7),
also available on the back flap. If accurate vapor pressures are available, the enthalpy of vaporiza-
tion can be estimated far from the critical point (i.e., Tr < 0.75) by the Clausius–Clapeyron equa-
tion:
sat
vap d ln P
H – R ------------------- (Tr < 0.75) (ig) 2.46
d 1 T
The background for this equation is developed in Section 9.2. Vapor pressure is often represented
by the Antoine equation, log10Psat = A B/(T(°C)+ C). If Antoine parameters are available, they
may be used to estimate the derivative term of Eqn. 2.46,
sat sat 2
d ln P 2.3026d log 10P – 2.3026 B T C + 273.15
------------------- = ------------------------------------------
- = --------------------------------------------------------------------- (ig)
d 1 T d 1 T T C +C
2
where T is in °C, and B and C are Antoine parameters for the common logarithm of pressure. For
Antoine parameters intended for other temperature or pressure units, the equation must be carefully
converted. The temperature limits for the Antoine parameters must be carefully followed because
the Antoine equation does not extrapolate well outside the temperature range where the constants
have been fit. If Antoine parameters are unavailable, they can be estimated to roughly 10% accu-
racy by the shortcut vapor pressure (SCVP) model, discussed in Section 9.3,
10. Pitzer, K.S., Lippmann, D.Z., Curl Jr., R.F., Huggins, C.M., Petersen, D.E. 1955. J. Am. Chem. Soc., 77:3433.
Section 2.11 Reference States 63
11. In the most detailed calculations, absolute zero is used as a reference state to create some thermodynamic tables. This is
based on a principle known as the third law of thermodynamics, that states that entropy goes to zero for a perfect crystal at
absolute zero. The difficulties in the rigorous calculations are mentioned above, and although the principles are straightfor-
ward, the actual calculations are beyond the scope of this book.
64 Unit I First and Second Laws
As you will notice in the following problems, reference states are not necessary when working
Reference with a pure fluid in a closed system or in a steady-state flow system with a single stream. The
states are not re-
quired for steady- numerical values of the changes in internal energy or enthalpy will be independent of the reference
state flow systems state.
with only a few
streams, but are When multiple components are involved, or many inlet/outlet streams are involved, definition
recommended
when many streams
of reference states is recommended since flow rates of the inlet and outlet streams will not necessar-
are present. ily match one-to-one. The reference state for each component may be different, so the reference
temperature, pressure, and state of aggregation must be clearly designated.
For unsteady-state open systems that accumulate or lose mass, reference states are imperative
Reference when values of U or H changes of the system or surroundings are calculated as the numerical
states are required values depend on the reference state. It is only when the changes for the system and surroundings
for unsteady-state are summed together that the reference state drops out for unsteady-state open systems.
changes in U, H
when mass accu-
mulation is present. Ideal Gas Properties
For an ideal gas, we must specify only the reference T and P.12 An ideal gas cannot exist as a liquid or
solid, and this fact completely specifies the state of our system. In addition, we need to set HR or UR
(but not both!) equal to zero.
T
ig ig
U =
òT R
C V dT + U R (ig) 2.49
T
ig ig
H = òT R
C P dT + H R (ig) 2.50
ig ig ig
Also at all states, including the reference state, U H – PV = H – RT so URig HRig RTR.
The ideal gas approximation is reliable when contributions from intermolecular potential energy
are relatively small. A convenient guideline is, in term of reduced temperature Tr = T/Tc, and
reduced pressure Pr = P/Pc, where Pc is the critcal pressure.
Assume ideal gas behavior if P < Psat and Tr > 0.5 + 2Pr (ig) 2.51
12. For calculation of ideal gas U and H, only a reference temperature is required; however, for the entropy introduced in
the next chapter, a reference pressure is needed, so we establish the P requirement now.
Section 2.11 Reference States 65
Figure 2.6 Illustrations of state pathways to calculate properties involving liquid/vapor phase changes.
The examples are representative, and modified paths would apply for states above the normal
boiling point, Tb. Similar pathways apply for solid/liquid or solid/vapor transformations. Note
that a generalized correlation is used for Hvap which differs from the normal boiling point
value. The method is intended to be used at subcritical conditions. Pressure corrections are not
illustrated for any paths here.
Solution: Heat capacity constants are available in Appendix E. For all cases, 20°C is 293.15K,
90°C is 363.15K, and the normal boiling point is Tb = 329.15K.
(a) HL = 0 because the liquid is at the reference state. The vapor enthalpy is calculated analogous
to Fig. 2.6, pathway (a). The three terms of pathway (a) are HV = 4639 + 30200 + 2799 = 37,638
J/mol. The difference in enthalpy is H = 37,638 J/mol.
(b) HL will use a path analogous to Fig. 2.6, pathway (b). The three terms of pathway (b) are
HL = 2366 – 30200 – 4638 = –32472 J/mol. HV is calculated using Eqn. 2.50, HV = 5166 J/mol.
The difference is H = 5166 + 32472 = 37,638 J/mol, same as part (a).
(c) HL will use a path analogous to Fig. 2.6, pathway (c). The generalized correlation of Eqn.
2.45 predicts a heat of vaporization at Tb of 29,280 J/mol, about 3% low. At 20°C, the heat of
vaporization is predicted to be 31,420 J/mol. The two steps in Fig. 2.6 (c) are HL = –365 – 31420
= –31785 J/mol. The enthalpy of vapor is the same calculated in part (b), HV = 5166 J/mol. The
enthalpy difference is H = 5166 + 31785 = 36,950 J/mol, about 2% low relative to part (b).
66 Unit I First and Second Laws
Solution: The properties of water and steam can be found from the saturated steam tables, inter-
polating between 20 C and 25 C. For saturated water or steam being heated from 24 C to 25 C,
and for vaporization at 25 C:
Of these values, the values for U of steam are lowest and can serve as the benchmark. How
much would kinetic and potential energy of a system have to change to be comparable to 1000 J?
Kinetic energy: If KE = 1000 J, and if the kg is initially at rest, then the velocity change must be,
2 2 1000J
v = ---------------------- or v 44.7 m/s
1kg
Velocity and This corresponds to a velocity change of 161 kph (100 mph). A velocity change of this order of
height changes magnitude is unlikely in most applications except nozzles (discussed below). Therefore, kinetic
must be large to be
significant in the en- energy changes can be neglected in most calculations when temperature changes occur.
ergy balance when
temperature Potential energy: If PE 1000 J, then the height change must be,
changes also occur.
1000 J
z = ------------------------------- = 102 m
N
1 æ 9.8066 ------ö
è kgø
This is equivalent to about one football field in position change. Once again this is very unlikely
in most process equipment, so it can usually be ignored relative to heat and work interactions.
Further, when a phase change occurs, these changes are even less important relative to heat and
work interactions.
Example 2.9 demonstrates that kinetic and potential energy changes of a fluid are usually neg-
ligible when temperature changes by a degree or more. Moreover, kinetic and potential energy
Section 2.12 Kinetic and Potential Energy 67
changes are closely related to one another in the design of piping networks because the temperature
changes are negligible. The next example helps illustrate the point.
Solution: A boundary will be placed around the expansion section of the piping. The system is
·
fixed volume, ( W EC = 0 ), adiabatic without shaft work. The open steady-state system is under
steady-state flow, so the left side of the energy balance is zero.
1 2
2 in 2 out
v gz in v gz out · · ·
0 = å H + -------- + ----- m· – å H + -------- + ----- m· + Q + W EC + W S
2g c g c 2g c g c
inlets outlets
2
– v
Simplifying: H = -----------------
2g c
Liquid water is incompressible, so the volume (density) does not change from the inlet to the
·
outlet. Letting A represent the cross-sectional area, and letting D represent the pipe diameter, V
v1A1 v2A2 Þ v2 v1(A1 / A2),
2 D1
4
2
v = v 1 æ ------ö – 1
è D 2ø
2
–v1 D1 4
H = -------- æ ------ö – 1
2g c è D 2ø
To calculate the temperature rise, we can relate the enthalpy change to temperature since they
are both state properties. From Eqn. 2.42, neglecting the effect of pressure,
H = CP T (*)
J 18.00 J kg - = 0.004K
C P = 4184 ----------- Þ T = -------------------------------------- (*)
kgK 4184 J kgK
68 Unit I First and Second Laws
Example 2.10 shows that the temperature rise due to velocity changes is very small. In a real
system, the measured temperature rise will be slightly higher than our calculation presented here
because irreversibilities are caused by the velocity gradients and swirling in the region of the sud-
den enlargement that we haven’t considered. These losses increase the temperature rise. In fluid
mechanics, irreversible losses due to flow are characterized by a quantity known as the friction
factor. The losses of a valve, fitting, contraction, or enlargement can be characterized empirically
by the equivalent length of straight pipe that would result in the same losses. We will introduce
these topics in Section 5.7. However, we conclude that from the standpoint of the energy balance,
the temperature rise is still small and can be neglected except in the most detailed analysis such as
the design of the piping network. In Example 2.10 the velocity decreases, and enthalpy increases
due to greater flow work on the inlet than the outlet. Note that the above result for a liquid does not
depend on whether the enlargement is rapid or gradual. A gradual taper will give the same temper-
ature change since the energy balance relates the enthalpy change to the initial and final velocities,
but not on the manner in which the change occurs.
Applications where kinetic and potential energy changes are important include solids such as
projectiles, where the temperature changes of the solids are negligible and the purpose of the work
is to cause accelerate or elevate the system. One example of this application is a steam catapult
used to assist in take-off from aircraft carriers. A steam-filled piston + cylinder device is expanded,
and the piston drags the plane to a velocity sufficient for the jet engines to lift the plane. While the
kinetic and potential energy changes for the steam are negligible, the work done by the steam
causes important kinetic energy changes in the piston and plane because of their large masses.
2 in 2 out
v - + gz in v - + gz out · · ·
0 = H + ------- ----- m· – H + ------- ----- m· + Q + W EC + W S
2g c g c 2g c g c
Changes in kinetic and potential energy are small relative to changes in enthalpy as we just dis-
cussed. When in doubt, the impact of changes in velocity can be evaluated as described in Example
2.9. The amount of heat transfer is negligible in a throttle. The boundaries are not expanding, and
there is also no mechanical device for transfer of work, so the work terms vanish. Therefore, a
throttle is isenthalpic:
Section 2.13 Energy Balances for Process Equipment 69
Nozzles
Nozzles are specially designed devices utilized to convert pressure drop into kinetic energy. Com-
mon engineering applications involve gas flows. An example of a nozzle is a booster rocket. Noz-
zles are also used on the inlets of impulse turbines to convert the enthalpy of the incoming stream to
a high velocity before it encounters the turbine blades.13 u is significant for nozzles. A nozzle is
designed with a specially tapered neck on the inlet and sometimes the outlet as shown schematically in
Fig. 2.7. Nozzles are optimally designed at particular velocities/pressures of operation to minimize
viscous dissipation.
The energy balance is written for a boundary around the nozzle. Any accumulation of energy
in the nozzle is neglected since it is small relative to flow through the device and zero at steady
state. Velocity changes are significant by virtue of the design of the nozzle. However, potential
energy changes are negligible. Heat transfer and work terms are dropped as justified in the discus-
sion of throttles. Reducing the energy balance for a nozzle shows the following:
2 in 2 out
v gz in v gz out · · ·
0 = H + -------- + ----- m· – H + -------- + ----- m· + Q + W EC + W S
2g c g c 2gc g c
Properly designed nozzles cause an increase in the velocity of the vapor and a decrease in the
enthalpy. A nozzle can be designed to operate nearly reversibly. Example 4.12 on page 162
describes a typical nozzle calculation.
Throttles are much more common in the problems we will address in this text. The meaning of
“nozzle” in thermodynamics is much different from the common devices we term “nozzles” in
everyday life. Most of the everyday devices we call nozzles are actually throttles.
Assessing when simplifications are justified requires testing the implications of eliminating
assumptions. For example, to test whether a particular valve is acting more like a throttle or a noz-
zle, infer the velocities before and after the nozzle and compare to the enthalpy change. If the
Flow Flow
13. Turbine design is a specialized topic. Introductions to the actual operation are most readily available in mechanical
engineering thermodynamics textbooks, such as Jones, J.B., Dugan, R.E. 1996. Engineering Thermodynamics. Upper Sad-
dle River, NJ: Prentice-Hall, pp. 734–745.
70 Unit I First and Second Laws
kinetic energy change is negligible relative to the enthalpy change then call it a throttle. Take note
of the magnitude of the terms in the calculation so that you can understand how to anticipate a sim-
ilar conclusion. For example, the volume change of a liquid due to a pressure drop is much smaller
than that of a gas. With less expansion, the liquids accelerate less, making the throttle approxima-
tion more reasonable. This kind of systematic analysis and reasoning is more important than mem-
orizing, say, that throttles are for liquids.
Heat Exchangers
Heat exchangers are available in a number of flow configurations. For example, in an industrial
heat exchanger, a hot stream flows over pipes that carry a cold stream (or vice versa), and the objec-
tive of operation is to cool one of the streams and heat the other. A generic tube-in-shell heat
exchanger can be illustrated by a line diagram as shown in Fig. 2.8. Tube-in-shell heat exchangers
consist of a shell (or outer sleeve) through which several pipes pass. (The figure just has one pipe
for simplicity.) One of the process streams passes through the shell, and the other passes through
the tubes. Stream A in our example passes through the shell, and Stream B passes through the tubes.
The streams are physically separated from one another by the tube walls and do not mix. Let’s sup-
pose that Stream A is the hot stream and Stream B is the cold stream. In the figure, both streams
flow from left to right. This type of flow pattern is called concurrent. The temperatures of the two
streams will approach one another as they flow to the right. With this type of flow pattern, we must
be careful that the hot stream temperature that we calculate is always higher than the cold stream
temperature at every point in the heat exchanger.14 If we reverse the flow direction of Stream A, a
countercurrent flow pattern results. With a countercurrent flow pattern, the outlet temperature of
the cold stream can be higher than the outlet temperature of the hot stream (but still must be lower
than the inlet temperature of the hot stream). The hot stream temperature must always be above the
cold stream temperature at all points along the tubes in this flow pattern also.
So far, our discussion has assumed that there are no phase transitions occurring in the heat
exchanger. If Stream A is a hot stream, and Stream B is converted from liquid at the inlet to vapor at
the outlet, we call the heat exchanger a boiler to bring attention to the phase transition occurring
inside. The primary difference in the operation of a boiler to that of a generic heat exchanger is that
the cold stream temperature change might be small or even zero. This is because the phase change
will occur isothermally at the saturation temperature of the fluid corresponding to the boiler pres-
sure, absorbing large amounts of heat. In a similar fashion, we could have Stream A be cooling
water and Stream B be an incoming vapor which is condensed. We would call this heat exchanger a
condenser, to clearly bring attention to the phase change occurring inside. In this case, the temper-
Tube
Stream B Out
Stream B In
Figure 2.8 Illustration of a generic heat exchanger with a concurrent flow pattern. The tube side
usually consists of a set of parallel tubes which are illustrated as a single tube for
convenience.
14. This may seem like common sense, but sometimes when calculations are performed, it is surprisingly easy to overlook
the fact that a valid mathematical result might be physically impossible to obtain.
Section 2.13 Energy Balances for Process Equipment 71
ature change of the hot stream might be small. Another type of heat exchanger that we will use in
Chapter 5 is the superheater. A superheater takes a vapor that is saturated and superheats it.
There are two more important points to keep in your mind as you perform thermodynamic cal-
culations. For the purposes of this text we will neglect pressure drops in the heat exchangers; the out-
let pressure will match the inlet pressure of Stream A, and a similar statement applies for Stream B.
Note that this does not imply that streams A and B are at the same pressure. Also, we neglect heat
transfer to or from the surroundings unless specified. Therefore, all heat transfer occurs inside the
heat exchanger, not at the boundaries of the heat exchanger and the surroundings.
There are other configurations of heat exchangers such as kettle-type reboilers and plate-and-
frame configurations; however, for thermodynamic purposes, only the flow pattern is important,
not the details of material construction that lead to the flow pattern. Thus, the tube-in-shell concepts
will be adequate for our needs.
The energy balance that we write depends on how we choose our system. Since the streams are
physically separated from one another, we may write a balance for each of the streams indepen-
dently, or we may place the system boundary around the entire heat exchanger and write a balance
for both streams. Let us take the system to be Stream B and let us suppose that Stream B is boiled.
In this case, there is just one inlet and outlet. There is no shaft work or expansion/contraction work.
Even though the process fluid is expanding as it evaporates, the system boundaries are not expand-
ing; expansion effects will be automatically included in the energy balance by the enthalpy terms
which have the flow work embedded in them. If the system is operating at steady state, the left-
hand side of the energy balance is zero,
2 in 2 out
v gz in v gz out · · ·
0 = H + -------- + ----- m· – H + -------- + ----- m· + Q + W EC + W S The energy
2g c gc 2g c g c balance for a heat
exchanger may be
written in several
which simplifies to ways.
·
0 = – Hm· + Q 2.54
·
where Q is the rate of heat transfer from the hot stream. On a molar (or mass) basis,
If we take the system boundaries to be around the entire heat exchanger, then there are multiple
streams, and all heat transfer occurs inside, resulting in
2 in 2 out
v gz in v gz out · · ·
0 = å H + -------- + ----- m· – å H + -------- + ----- m· + Q + W EC + W S
2g c g c 2g c g c
inlets outlets
which simplifies to
in in out out
m· – m·
0 =
åH å H 2.56
inlets outlets
72 Unit I First and Second Laws
Since Eqns. 2.55 and 2.57 look quite different for the same process, it is important that you under-
stand the placement of boundaries and their implications on the balance expression.
The energy balance for the turbine only involves the kinetic energy change for the entering and
exiting fluid, not for the changes occurring inside the turbine. Since the nozzles which cause large
kinetic energy changes are inside the turbine unit, these changes are irrelevant to the balance
around the unit. Recall from the development of our energy balance that we are only interested in
the values of enthalpy, kinetic, and potential energy for streams as they cross the boundaries of our
system. The energy balance for a steady-state turbine involving one inlet and one outlet is:
2 in 2 out
v - + gz in v - + gz out · · ·
0 = H + ------- ----- m· – H + ------- ----- m· + Q + W EC + W S
2g c g c 2g c gc
which becomes
in out · ·
Adiabatic tur- 0 = H –H m· + W S = – Hm· + W S 2.58
bine or expander.
Rotors
Stators
Inlet
Shaft
Outlet
Figure 2.9 Illustration of a turbine. The rotor (shaft) turns due to the flow of gas.
The blades connected to the shell are stationary (stators), and are
sometimes curved shapes to perform as nozzles. The stator blades are
not shown to make the rotors more clear.
Section 2.13 Energy Balances for Process Equipment 73
When we calculate values for the H and work, they will be negative values.
Adiabatic Compressors
Adiabatic compressors can be constructed in a manner qualitatively similar to adiabatic turbines
with stationary vanes (stators). This type of compressor is called an axial compressor. The main dif-
ferences between turbines and axial compressors are: 1) the details of the construction of the vanes
and rotors, which we won’t be concerned with; 2) the direction of flow of the fluid; and 3) the fact
that we must put work into the compressor rather than obtaining work from it. Thus, the energy bal-
ance is the same as the turbine (Eqns. 2.58 and 2.59). When we calculate values for the H and
work, they will be positive values, where they were negative values for a turbine. Compressors may
also be constructed as reciprocating (piston/cylinder) devices. This modification has no impact on
our energy balance, so it remains the same. Analogous to turbines, it is conventional to assume that
compressors are adiabatic unless otherwise noted.
Pumps
Pumps are used to move liquids by creating the pressure necessary to overcome the resistance to
flow. They are in principle just like compressors, except the liquid will not change density the way a
gas does when it is compressed. Again, the energy balance will be the same as a turbine or compres-
sor (Eqns. 2.58 and 2.59). The primary difference we will find in application of the energy balance is Compressors
that tabulated enthalpies are difficult to find for compressed liquids. Therefore, if we want to calcu- and pumps usually
late the work needed for a pump, we can find it from the energy balance after we have calculated or have the same
energy balance as
determined the enthalpy change. turbines.
out
WS = òin V dP reversible pump, compressor, turbine 2.60
L out in P
WS V P –P = ------- liquid pump 2.61
When the work is to calculated, the adiabatic shaft work may, in principle, be analyzed using H or
Eqn 2.60. For gases, it is usually easier to use other constraints to find the enthalpy change and then
calculate the work by equality, though in principle the integral can be evaluated. For the special
case of liquids, Eqn 2.60 can be replaced by Eqn. 2.61.
74 Unit I First and Second Laws
1. Choose system boundaries; decide whether this boundary location will make the system
open or closed.
2. Identify all given state properties of fluids in system and crossing boundaries. Identify
which are invariant with time. Identify your system as steady or unsteady state. (For
unsteady-state pumps, turbines, or compressors, the accumulation of energy within the
device is usually neglected.) For open, steady-state systems, write the mass balance and
solve if possible.
3. Identify how many state variables are unknown for the system. Recall that only two state
The phase rule variables are required to specify the state of a pure, single-phase fluid. The number of
is important in unknowns will equal the number of independent equations necessary for a solution.
determining the re-
quired number of (Remember in a system of known total volume V, that if n is known, the state variable V is
equations. known.)
4. Write the mass balance and the energy balance. These are the first equations to be
used in the solution. Specify reference states for all fluids if necessary. Simplify energy
balance to eliminate terms which are zero for the system specified in step 1.15 Combine
the mass balance and the energy balance for open systems.
For unsteady-state problems:
(a) Identify whether the individual terms in the energy balance may be integrated directly
Always con- without combining with other energy balance terms. Often the answer is obtained most
sider the overall easily this way. This is almost always possible for closed-system problems.
balance.
(b) If term-by-term integration of the energy balance is not possible, rearrange the equation
to simplify as much as possible before integration.
5. Look for any other information in the problem statement that will provide additional
Look for key equations if unknowns remain. Look for key words such as adiabatic, isolated, throt-
words. tling, nozzle, reversible, and irreversible. Using any applicable constraints of throttling
devices, nozzles, and so on, relate stream properties for various streams to one another and
to the system state properties. Constraints on flow rates, heat flow, and so on. provide addi-
tional equations. With practice, many of these constraints may be recognized immediately
before writing the energy balance in steps 3 and 4.
6. Introduce the thermodynamic properties of the fluid (the equation of state). This provides all
equations relating P, V, T, U, H, CP, and CV. The information will consist of either 1) the
Use the same ideal gas approximation; 2) a thermodynamic chart or table; or 3) a volumetric equation of
property method state (which will be introduced in Chapter 7). Using more than one of these sources of infor-
throughout the prob- mation in the same problem may introduce inconsistencies in the properties used in the solu-
lem if possible.
tion, depending on the accuracy of the methods used.
Combine the thermodynamic information with the energy balance. Work to minimize the
number of state variables which remain unknown. Many problems are solved at this point.
15. Section 2.17 on page 85 may be helpful for details on interpreting each term of the balance for new applications.
Section 2.15 Closed and Steady-State Open Systems 75
7. Do not hesitate to move your system boundary and try again if you are stuck. Do not forget
to try an overall balance (frequently, two open systems can be combined to give an overall Try different
system boundaries.
closed system, and strategy 4a can be applied). Make reasonable assumptions.
8. After an answer is obtained, verify assumptions that were made to obtain the solution. Make sure
answers seem
reasonable.
2.15 CLOSED AND STEADY-STATE OPEN SYSTEMS
Several types of systems are quite common in chemical engineering practice. You need to be famil-
iar with the results of their analysis and benefit if you memorize these results for rapid recall. You
must simultaneously recall the assumptions underlying each simple model, however, to avoid
incorrect applications.
Solution: First consider the energy balance. The system will be the gas in the cylinder. The sys-
tem will be closed. Since a basis is not specified, we can choose 1 mole. Since there is no mass
flow, heat transfer, or shaft work, the energy balance becomes:
2
g
v - + ----
d U + ------- - z = dQ + dW S + dW EC
2g c g c
dU = – PdV
In this case, as we work down to step 4 in the strategy, we see that we cannot integrate the sides
independently since P depends on T. Therefore, we need to combine terms before integrating.
dV CV R
C V dT = – RT ------- which becomes ------- dT = – --- dV (ig) 2.62
V T V
The technique that we have performed is called separation of variables. All of the temperature
dependence is on the left-hand side of the equation and all of the volume dependence is on the
right-hand side. Now, if we assume a constant heat capacity for simplicity, we can see that this
integrates to
CV T i
------- ln ------i = ln V
----- (*ig)
R T V
T-ö C V R
æ ----- V
i
i = ------ (*ig) 2.63 These boxed
èT ø V equations relate
state variables for
adiabatic reversible
changes of an ideal
gas in a closed
system.
76 Unit I First and Second Laws
CV
------
Tö R
æ ----- i
T P
ç -i ÷ = ------i --- (*ig)
èT ø P T
Rearranging,
T-ö C V R T
æ ----- T CV R + 1 P
i ------i = æ ------iö = -----i (*ig)
èT ø T èT ø P
which becomes
T ö CP R
æ ----- P
- = -----i (*ig) 2.64
è T iø P
We may also insert the ideal gas law into Eqn. 2.63 to convert to a formula relating P and V.
Using T PV/R,
PV ö C V R
æ ---------- V
i
i
-i = ----- (*ig)
èP V ø V
R CV R CV + 1 CP C V
P Vi Vi Vi Vi
-----i = æ -----ö ----- = æ -----ö = æ -----ö (*ig)
P è Vø V è Vø è Vø
CP CV
PV = const (*ig) 2.65
The analysis of a piston+cylinder implied the assumption of a closed system. This might be a
reasonable approximation for a single stroke of a combustion engine, but most chemical engineer-
ing applications involve continuous operation. Nevertheless, we can apply the lessons learned from
the analysis of the closed system when extending to steady-state systems, as exemplified below.
H = Q + WS 2.66
dW S = dH = VdP
In this case, as we work down to step 4 in the strategy, we see that we cannot integrate the sides
independently since P depends on T. Therefore, we need to combine terms before integrating.
dP CP R
C P dT = RT ------- which becomes ------- dT = --- dP (ig) 2.67
P T P
Once again, we have performed separation of variables. The rest of the derivation is entirely
analogous to Example 2.11, and, in fact, the resultant formula is identical.
R CP
T2 æ P 2ö
----- = ç ------÷ (*ig) 2.68
è P 1ø
Steady-state
T1 adiabatic, revers-
ible processing of
Note that this formula comes up quite often as an approximation for both open and closed systems. an ideal gas
results in the
Making the appropriate substitutions, same relations
as Example 2.11.
2
---
25 7
T 2 = 298 æ ------ö = 472 K
è 5ø
ig
ig ig ig ig P R CP
WS = H = CP T = CP T1 æ -----2-ö –1 (*ig) 2.69
è P 1ø Steady-state
adiabatic, revers-
ible compression of
an ideal gas.
Substituting, WS 3.5·8.314·(472 298) 5063 J/mole
At the given flow rate, and reiterating that this problem statement specifies a reversible process:
We have systematically extended our analysis from a single step of a closed system, to a con-
tinuous system with no heat loss. Let’s consider isothermal operation.
Solution: Let’s return to the perspective of the section ‘Understanding Enthalpy and Shaft
Work’ on page 54 and analyze the EC work and flow work for an ideal gas packet of unit mass.
The WEC is,
WEC = – òPdV = – ò(RT/V)dV = – RTln(V2/V1) (*ig)
For an isothermal, ideal gas, V2/V1 = P1/P2. Noting the reciprocal and negative logarithm,
This is the work to isothermally squeeze an ideal gas packet of unit mass to a given pressure.
The flow work performed on an ideal gas packet of unit mass is,
Therefore, the total requirement for isothermally compressing an ideal gas packet of unit mass is,
WS = RTln(P2/P1) (*ig)
æ P 2ö J
W S = RT ln ç ------÷ = 8.314 298 ln 5 = 3987 --------- (ig) 2.70
P
è 1ø mol
At the given flow rate, and reiterating that this problem statement specifies a reversible process,
Compared to adiabatic compression, the isothermal compressor requires less work. This hap-
pens because cooling withdraws energy from the system. It is difficult to achieve perfectly adia-
batic or isothermal operation in practice, but adiabatic operation is usually a better approximation
because compression is so rapid that there is insufficient time for heat transfer. Usually fluids are
cooled after compression as we discuss in later chapters.
Here is a brain teaser. Suppose the process fluid had been steam instead of an ideal gas. How
ig
would you have solved the problem then? Note, for steam V RT/P and dH C P dT . We illustrate
the answer in Example 4.16 on page 173.
Section 2.15 Closed and Steady-State Open Systems 79
Solution: First draw a schematic. Designate boundaries. Both System A and System B are open
steady-state systems.
Stream (1)
System A
3.5 MPa
350 C
1100 kg/h
The mass balance gives m· 3 = 990 kg/h . Next, determine which streams are completely speci-
fied: Streams (1), (2), and (4) are fully specified. Since Stream (3) is saturated, the temperature
and pressure and specific enthalpies of the saturated vapor and liquid can be found, but the qual-
ity needs to be calculated to determine the overall molar enthalpy of the stream. From the steam
tables we find H1 directly. For H2 we use linear interpolation. The value H(1.5 MPa, 225 C) is
not available directly, so we need to first interpolate at 1.4 MPa between 200 C and 250 C to
find H(1.4 MPa, 225 C) and then interpolate between this value and the value at 1.6 MPa:
1
H 1.4MPa 225 C = --- 2803.0 + 2927.9 = 2865.5 kJ/kg
2
23.6
H 1.6MPa 225 C = 2792.8 + ---------- 2919.9 – 2792.8 = 2854.5 kJ/kg
48.6
Then to find H2: H2 0.5·(2865.5 + 2854.5) 2860.0 kJ/kg. For H4 we can interpolate in the
superheated steam tables:
20
H 4 = 2675.8 + ------ 2776.6 – 2675.8 = 2716.1 kJ/kg
50
80 Unit I First and Second Laws
Stream 1 2 3, 4
H(kJ/kg) 3104.8 2860.0 2716.1
·
The energy balance for System A gives, using W S = – 100 kW given in the problem statement,
· · · ·
0 = H 1 m· 1 – H 2 m· 2 – H 3 m· 3 + Q + W S 3104.8(1100) 2860.0(110) 2716.1(990) Q + WS
At 0.79 MPa from the sat’d P table, HL 718.5 kJ/kg and HV 2049 kJ/kg.
2716.1 – 718.5
q = ------------------------------------ = 0.975
2049
The energy balance for a non-adiabatic turbine is identical to the balance for an isothermal com-
pressor, but the conclusions are entirely different. In the compressor, we want to minimize work, so
the heat loss works to our advantage. For the turbine, we want to maximize work, so any loss of
energy should be avoided.
The examples in this section comprise several important common scenarios, but they also illustrate
a procedure for analyzing systems with systematically increasing sophistication. In the context of
certain simplifying assumptions, like the ideal gas model, we can derive final working equations
applicable to process calculations. When those assumptions are invalid, however, we can still apply
the energy balance, but we are more careful in the generality of the results we obtain. Processes
involving steam, for example, require something more than the ideal gas model, and additional
tools are required to develop a general analysis.
An ideal gas is leaking from an insulated tank. Relate the change in temperature to the change in
pressure for gas leaking from a tank. Derive an equation for U for the tank.
Solution: Let us choose our system as the gas in the tank at any time. This will be an open,
unsteady-state system. There is no inlet stream and one outlet stream. The mass balance gives
dn dnout.
We can neglect kinetic and potential energy changes. Although the gas is expanding, the system
size remains unchanged, and there is no expansion/contraction work. The energy balance
becomes (on a molar basis):
in in out out
d nU = H dn – H dn + dQ + dW EC + dW S
Since the enthalpy of the exit stream matches the enthalpy of the tank, Hout H.
d nU = – H out dn out = Hdn . Now H depends on temperature, which is changing, so we are
not able to apply hint 4(a) from the problem-solving strategy. It will be necessary to combine
terms before integrating. By the product rule of differentiation, the left-hand side expands to
d nU = ndU + Udn . Collecting terms in the energy balance,
ndU = H – U dn
Performing some substitutions, the energy balance can be written in terms of T and n,
C V dT dn
H – U = PV = RT ; dU = C V dT; Þ ------- ------ = ------ (ig)
R T n
CV T n-
------- ln ----- = ln --- (*ig)
R i i
T n
i CV T
n PT T P
ln ---- = ln --- ----- = – ln ----- + ln ----- = ------- ln ----- (ig)
i T Pi i i R i
n T P T
substituting,
CP T
æC T P
------V- + 1ö ln ----- = ------- ln ----- = ln ----- (*ig)
èR ø i R i i
T T P
82 Unit I First and Second Laws
èP ø
i
T è ø
Through the ideal gas law (PV = RT), we can obtain other arrangements of the same formula.
Pi 1 1 –1 –1
V V i = æ -----ö ; P i P = V V i = Ti T ; T i T = V Vi (*ig) 2.72
è Pø
The numerical value for the change in internal energy of the system depends on the reference
state because the reference state temperature will appear in the result:
f f i i
U = n CV T – TR + UR – n CV T – TR + UR (*ig)
At first glance, one might expect to use the same equation for a filling tank, but simply change
the pressure ratio. Careful analysis shows that the energy balance is similar, but the final result is
quite different.
Solution: The system will be the gas inside the tank at any time. The system will be an open,
unsteady-state system. The mass balance is dn dnin. The energy balance becomes:
in in out out
d nU = H dn – H dn + dQ + dW EC + dW S
We recognize that Hin will be constant throughout the tank filling. Therefore, by hint 4a from the
problem-solving strategy, we can integrate terms individually. We need to be careful to keep the
superscript since the incoming enthalpy is at a different state than the system. The right-hand
side of the energy balance can be integrated to give
f f
in in in f i in f
òi H dn = H
òi dn = H n –n = H n
Un = U f n f – U i n i = U f n f
Section 2.16 Unsteady-State Open Systems 83
f in in in in in
U = H = U + PV = U + RT (ig) 2.73
And with our definition of heat capacity, we can find temperatures:
U = C V T f – T in = RT in Þ T f = T in R + C v C v = T inC p C v (*ig)
Note that the final temperature is independent of pressure for the case considered here.
You should not get the impression that unsteady, open systems are limited to ideal gases.
Energy balances are independent of the type of operating fluid.
Solution: There are some similarities with the solution to Example 2.15 on page 81; however,
we can no longer apply the ideal gas law. The energy balance reduces in a similar way, but we
note that the exiting stream consists of only vapor; therefore, it is not the overall average
enthalpy of the tank:
out out V
d mU = – H dm = H dm
The sides of the equation can be integrated independently if the vapor enthalpy is constant.
Looking at the steam table, the enthalpy changes only about 10 kJ/kg out of 2800 kJ/kg (0.3%)
along the saturation curve down to 195 C. Let us assume it is constant at 2795 kJ/kg making the
integral of the right-hand side simply HV m. Note that this procedure is equivalent to a numeri-
cal integration by trapezoidal rule as given in Appendix B on page 822. Many students forget
that analytical solutions are merely desirable, not absolutely necessary. The energy balance then
can be integrated using hint 4a on page 74.
f f i i f i
U = m U – m U = 2795 m – m = 2795 –25 = – 69 875 kJ
The quantity m f = 475, and miU i will be easy to find, which will permit calculation of U f. For
each m3 of the original saturated mixture at 215 C,
V
t m3 - = 0.00233 m 3 kg
= -------------------------------------
423.4 + 5.28 kg
So the tank volume, quality, and internal energy are:
500 kg m3 3
V T = --------------------------------------- = 1.166 m
428.63
To guide our first guess, we need U L U f 841.0 kJ/kg. Our first guess is T f 195 C. Values for
the properties from the steam tables are shown in the table below. This initial guess gives U fcalc
845 kJ/kg; no further iteration is necessary. The HV at this state is 2789; therefore, our assumption
of Hout constant is valid.
The universe frequently consists of three subsystems, as illustrated in Fig. 2.10. The container (Sys-
tem 2) is frequently combined with System 1 (designated here as System (1 2)) or System 3 (des-
ignated here as System (2 3)). For every balance, all variables are of the system for which the
balance is written,
2 2 in
d- æ v gzö æ v gzö in
---- m ç U + -------- + -----÷ =
å ç H + -------- + -----÷ m·
dt è 2g c g cø è 2g c g cø
inlets
2.74
2 out
æ v gzö · out Q· + W
· ·
–
å ç H + -------- + -----÷
è 2g c g cø
m +
å EC + W S
outlets surfaces
where superscripts “in” and “out” denote properties of the streams which cross the boundaries,
which may or may not be equal to properties of the system.
1. Non-zero heat interactions of Systems 1 and 3 are not equal unless the heat capacity or
mass of System 2 is negligible.
2. Hout, Hin account for internal energy changes and work done on the system due to flow
across boundaries.
Adiabatic wall ·
System boundary Q ·
WS
Heat-conducting wall
System 1 (Fluid)
m· in m· out
System 2 System 3
(Container) (Surroundings)
Piston
·
W EC
· F boundaryö dV system
3. W EC = æ – ------------------------ --------------------- represents work done on the system due to expansion or
è A ø dt
boundary
contraction of system size. Fboundary is the absolute value of the system force at the bound-
ary. (By Newton’s third principle, the forces are equal but act in opposite directions at any
boundary.) Note that system 1 and system 3 forces are not required to be equal. Unequal
forces create movement (acceleration) of any movable barrier (e.g., piston head).
·
4. W S represents the work done on the system resulting from mechanical forces at the surface
of the system except work due to expansion/contraction or mass flow across boundaries.
Turbines and compressors are part of system 2; thus they are involved with work · interac-
tions with the fluid in system 1. Note that piston movement is· calculated as W EC for sys-
tems 1, 3, (1 2), or (2 3), but the movement is calculated as W S for system 2 alone. When
a balance for system 2 is considered, the movement of the piston is technically shaft work,
even though no shaft is involved. (The piston is a closed system, and it does not expand or
contract when it moves.) As another example of the general definition of shaft work as it
relates to forces at the surface of a system, consider the closed system of Fig. 2.3 on page 48
being raised 150 m or accelerated to 75 m/s. There is a work interaction at the surface of
the system required for these energy changes even though there is not a “rotating” shaft.
· ·
5. Non-zero W EC or W S interactions of systems 1 and 3 are not equal unless changes · in
kinetic and potential
· energy of the moving portion of system 2 (e.g., piston head for W EC
or turbine for W S ) are negligible and the movement is reversible.
6. Frictional forces, if present, must be attributed to one of the systems shown above. Irrevers-
ibility due to any cause does not require additional energy balance terms because energy is
always conserved, even when processes are irreversible.
7. Electrical and magnetic fields have not been included.
8. On the left-hand side of the equation, kinetic and potential energy changes are calculated
based on movement of the center of mass. In a composite system such as (1+2), they may
be calculated for each subsystem and summed.
2.18 SUMMARY
We are trying to be very careful throughout this chapter to anticipate every possibility that might
arise. As a result, the verbiage gets very dense. Think of the complete energy balance as a checklist,
reminding you to consider whether each term may contribute significantly to a given problem, and
learning to translate key terms like “frictionless,” “reversible,” “continuous,” and “steady state”
into meaningful reductions of the balance.
Important Equations
If we relax the formality, we can summarize most of this chapter casually as follows:
U = Q + W EC closed systems
in in out out
d nU = H dn – H dn + dQ + dW EC + dW S open, unsteady-state systems
Section 2.18 Summary 87
Naturally, it is best to appreciate how these equations result from simplifications. Remember to
check the general energy balance for terms that may be significant in exceptional situations.
A summary of expansion/contraction work relations for ideal gases is also presented here,
however it is recommended that you become proficient in the manipulations leading to these for-
mulas. Section 2.4 summarizes factors that may make a process irreversible. The following formu-
las represent reversible work done when the system pressure is inserted for the isothermal process.
The isobaric formula is the only one that can be used to directly calculate work done on the sur-
roundings in an irreversible process, and in that case the surrounding pressure is used instead of the
system pressure.
Isobaric W EC = – ò P dV = – P V 2 – V 1
dV
W EC = – ò P dV = – ò const --------------------
CP CV
- (*ig)
V
or
Adiabatic and reversible U = C V T 2 – T1 = W EC (*ig)
R CP R CV
T2 æ P 2ö æ V 1ö
----- = ç ------÷ = ç -----
-÷ (*ig)
T1 è P 1ø è V 2ø
This last equation (for T2/T1) recurs frequently as we examine processes from various perspectives
and simplify them to ideal gases for preliminary consideration. You should commit it to memory
soon and learn to recognize when it is applicable.
Test Yourself
1. Write the energy balance without looking at the book. To help remember the terms, think
about the properties the terms represent rather than memorizing the symbols.
2. In the presentation of the text, which side of the balance represents the system and which
terms represent interactions at the boundaries?
3. Explain the terms “closed-system,” “open-system,” and “steady state” to a friend of the
family member who is not an engineer.
4. Explain how a reference state helps to solve problems. Select a reference state for water
that is different from the steam table reference state. Create a path starting from saturated
liquid below the normal boiling point, through the normal boiling point, and cooling down
to saturated vapor at the initial temperature. Use heat capacities and the latent heat at the
normal boiling point to estimate the heat of vaporization and compare it with the steam
table value.
5. Write a MATLAB, Excel, or calculator routine that will enable you to calculate heat capac-
ity integrals easily.
6. Think of as many types of paths as you can from memory (isothermal, adiabatic, etc.) and
try to derive the heat and work flow for a piston/cylinder system along each path.
88 Unit I First and Second Laws
P2.1 A pot of water is boiling in a pressure cooker when suddenly the pressure relief valve
becomes stuck, preventing any steam from escaping. System: the pot and its contents after
the valve is stuck. ANS. d mU dt = Q·
P2.2 The same pot of boiling water as above. System: the pot and its contents before the valve is
·
stuck. ANS. d mU = inH V + Q
P2.3 An gas home furnace has been heating the house steadily for hours. System: the furnace.
·
ANS. H = Q (gas furnace)
P2.4 An gas home furnace has been heating the house steadily for hours. System: the house and
· ·
all contents. ANS. d mU dt = Q Heat + QLoss (gas furnace)
P2.5 A child is walking to school when he is hit by a snowball. He stops in his tracks. System:
2
the child. ANS. mU + mv 2gc = m snow v 2 2g c snow
P2.6 A sealed glass bulb contains a small paddle-wheel (Crookes radiometer). The paddles are
painted white on one side and black on the other. When placed in the sun, the paddle wheel
begins to turn steadily. System: the bulb and its contents. (ANS. U = 0 )
P2.7 A sunbather lays on a blanket. At 11:30 A.M., the sunbather begins to sweat. System: the
sunbather at noon. ANS. d mU dt = m· H V + Q·
P2.8 An inflated balloon slips from your fingers and flies across the room. System: balloon and
out
its contents. ANS. d mU + mv 2 balloon 2g c dt = H + v 2 2g c dm dt + W EC
B. Numerical Problems
P2.9 Consider a block of concrete weighing 1 kg.
(a) How far must it fall to change its potential energy by 1 kJ? (ANS. 100 m)
(b) What would be the value of its velocity at that stage? (ANS. 44.7 m/s)
P2.10 A block of copper weighing 0.2 kg with an initial temperature of 400 K is dropped into 4 kg
of water initially at 300 K contained in a perfectly insulated tank. The tank is also made of
copper and weighs 0.5 kg. Solve for the change in internal energy of both the water and the
block given CV 4.184 J/g-K for water and 0.380 J/g-K for copper. (ANS. 7480 J, 7570 J)
P2.11 In the preceding problem, suppose that the copper block is dropped into the water from a
height of 50 m. Assuming no loss of water from the tank, what is the change in internal
energy of the block? (ANS. 7570 J)
P2.12 In the following take CV 5 and CP 7 cal/mol-K for nitrogen gas:
(a) Five moles of nitrogen at 100 °C is contained in a rigid vessel. How much heat must
be added to the system to raise its temperature to 300 °C if the vessel has a negligible
heat capacity? (ANS. 5000 cal) If the vessel weighs 80 g and has a heat capacity of
0.125 cal/g-K, how much heat is required? (ANS. 7000 cal)
Section 2.19 Practice Problems 89
(a) Using the steam tables, compute the flow of heat into the process equipment per kg of
steam. (ANS. 288 kJ/kg)
(b) Compute the value of enthalpy at the inlet conditions, H in, relative to an ideal gas at the
same temperature, Hi g. Consider steam at 1 bar and 600°C as an ideal gas. Express
your answer as (H in H ig)/RT in. (ANS. 0.305)
P2.17 A 700 kg piston is initially held in place by a removable latch above a vertical cylinder.
The cylinder has an area of 0.1 m2; the volume of the gas within the cylinder initially is 0.1
m3 at a pressure of 10 bar. The working fluid may be assumed to obey the ideal gas equa-
tion of state. The cylinder has a total volume of 0.25 m3, and the top end is open to the sur-
rounding atmosphere at 1 bar.
(a) Assume that the frictionless piston rises in the cylinder when the latches are removed
and the gas within the cylinder is always kept at the same temperature. This may seem
like an odd assumption, but it provides an approximate result that is relatively easy to
obtain. What will be the velocity of the piston as it leaves the cylinder? (ANS. 13.8 m/s)
(b) What will be the maximum height to which the piston will rise? (ANS. 9.6 m)
(c) What is the pressure behind the piston just before it leaves the cylinder? (ANS. 4 bar)
(d) Now suppose the cylinder was increased in length such that its new total volume is
0.588 m3. What is the new height reached by the piston? (ANS. ~13 m)
(e) What is the maximum height we could make the piston reach by making the cylinder
longer? (ANS. ~13 m)
P2.18 A tennis ball machine fires tennis balls at 40 mph. The cylinder of the machine is 1 m long;
the installed compressor can reach about 50 psig in a reasonable amount of time. The ten-
nis ball is about 3 inches in diameter and weighs about 0.125 lbm. Estimate the initial vol-
ume required in the pressurized firing chamber. [Hint: Note the tennis ball machine fires
horizontally and the tennis ball can be treated as a frictionless piston. Don’t be surprised if an
iterative solution is necessary and ln (V2 / V1) ln(1 V/V1)]. (ANS. 390 cm3)
P2.19 A 700 kg piston is initially held in place by a removable latch inside a horizontal cylinder.
The totally frictionless cylinder (assume no viscous dissipation from the gas also) has an
area of 0.1 m2; the volume of the gas on the left of the piston is initially 0.1 m3 at a pressure
of 8 bars. The pressure on the right of the piston is initially 1 bar, and the total volume is
0.25 m3. The working fluid may be assumed to follow the ideal gas equation of state. What
would be the highest pressure reached on the right side of the piston and what would be the
90 Unit I First and Second Laws
position of the piston at that pressure? (a) Assume isothermal; (b) What is the kinetic
energy of the piston when the pressures are equal?16 (ANS. (a) 12.5 MPa; (b) 17 kJ)
2.2 One mole of an ideal gas (CP 7R/2) in a closed piston/cylinder is compressed from T i
100 K, P i 0.1 MPa to P f 0.7 MPa by the following pathways. For each pathway, calcu-
late U , H , Q, and W EC : (a) isothermal; (b) constant volume; (c) adiabatic.
2.3 One mole of an ideal gas (CP 5R/2) in a closed piston/cylinder is compressed from T i 298 K,
P i 0.1 MPa to P f 0.25 MPa by the following pathways. For each pathway, calculate
U , H , Q, and W EC : (a) isothermal; (b) constant volume; (c) adiabatic.
2.4 One mole of an ideal gas (CP 7R/ 2) in a closed piston/cylinder is expanded from T i 700 K,
P i 0.75 MPa to P f 0.1 MPa by the following pathways. For each pathway, calculate
U , H , Q, and W EC : (a) isothermal; (b) constant volume; (c) adiabatic.
2.5 One mole of an ideal gas (CP 5R/2) in a closed piston/cylinder is expanded from T i 500 K,
P i 0.6 MPa to P f 0.1 MPa by the following pathways. For each pathway, calculate U ,
H , Q, and W EC : (a) isothermal; (b) constant volume; (c) adiabatic.
2.6 (a) What is the enthalpy change needed to change 3 kg of liquid water at 0 C to steam at
0.1 MPa and 150 C?
(b) What is the enthalpy change needed to heat 3 kg of water from 0.4 MPa and 0 C to
steam at 0.1 MPa and 150 C?
(c) What is the enthalpy change needed to heat 1 kg of water at 0.4 MPa and 4 C to steam
at 150 C and 0.4 MPa?
(d) What is the enthalpy change needed to change 1 kg of water of a water-steam mixture
of 60% quality to one of 80% quality if the mixture is at 150 C?
(e) Calculate the H value for an isobaric change of steam from 0.8 MPa and 250 C to sat-
urated liquid.
(f) Repeat part (e) for an isothermal change to saturated liquid.
(g) Does a state change from saturated vapor at 230 C to the state 100 C and 0.05 MPa
represent an enthalpy increase or decrease? A volume increase or decrease?
(h) In what state is water at 0.2 MPa and 120.21 C? At 0.5 MPa and 151.83 C? At 0.5
MPa and 153 C?
(i) A 0.15 m3 tank containing 1 kg of water at 1 MPa and 179.88 C has how many m3 of
liquid water in it? Could it contain 5 kg of water under these conditions?
(j) What is the volume change when 2 kg of H2O at 6.8 MPa and 93 C expands to 1.6 bar
and 250 C?
(k) Ten kg of wet steam at 0.75 MPa has an enthalpy of 22,000 kJ. Find the quality of the
wet steam.
2.7 Steam undergoes a state change from 450 C and 3.5 MPa to 150 C and 0.3 MPa. Deter-
mine H and U using the following:
(a) Steam table data.
(b) Ideal gas assumptions. (Be sure to use the ideal gas heat capacity for water.)
2.8 Five grams of the specified pure solvent is placed in a variable volume piston. What are the
molar enthalpy and total enthalpy of the pure system when 50% and 75% have been evap-
orated at: (i) 30 C, (ii) 50 C? Use liquid at 25 C as a reference state.
(a) Benzene ( L 0.88 g/cm3)
(b) Ethanol ( L 0.79 g/cm3)
L
(c) Water without using the steam tables ( 1 g/cm3)
(d) Water using the steam tables
2.9 Create a table of T, U, H for the specified solvent using a reference state of H 0 for liquid
at 25 C and 1 bar. Calculate the table for: (i) liquid at 25 C and 1 bar; (ii) saturated liquid
at 1 bar; saturated vapor at 1 bar; (iii) vapor at 110 C and 1 bar. Use the Antoine equation
(Appendix E) to relate the saturation temperature and saturation pressure. Use the ideal gas
law to model the vapor phase.
(a) Benzene
(b) Ethanol
(c) Water without using the steam tables
(d) Water using the steam tables
2.10 One kg of methane is contained in a piston/cylinder device at 0.8 MPa and 250 C. It under-
goes a reversible isothermal expansion to 0.3 MPa. Methane can be considered an ideal gas
under these conditions. How much heat is transferred?
2.11 One kg of steam in a piston/cylinder device undergoes the following changes of state. Cal-
culate Q and W for each step.
(a) Initially at 350 kPa and 250 C, it is cooled at constant pressure to 150 C.
(b) Initially at 350 kPa and 250 C, it is cooled at constant volume to 150 C.
2.12 In one stroke of a reciprocating compressor, helium is isothermally and reversibly com-
pressed in a piston cylinder from 298 K and 20 bars to 200 bars. Compute the heat
removal and work required.
2.13 Air at 30°C and 2MPa flows at steady state in a horizontal pipeline with a velocity of 25 m/s.
It passes through a throttle valve where the pressure is reduced to 0.3 MPa. The pipe is the
same diameter upstream and downstream of the valve. What is the outlet temperature and
velocity of the gas? Assume air is an ideal gas with a temperature-independent CP 7R / 2,
and the average molecular weight of air is 28.8.
2.14 Argon at 400 K and 50 bar is adiabatically and reversibly expanded to 1 bar through a tur-
bine in a steady process. Compute the outlet temperature and work derived per mole.
2.15 Steam at 500 bar and 500 C undergoes a throttling expansion to 1 bar. What will be the
temperature of the steam after the expansion? What would be the downstream temperature
if the steam were replaced by an ideal gas, CP /R 7/2?
92 Unit I First and Second Laws
2.16 An adiabatic turbine expands steam from 500°C and 3.5 MPa to 200 C and 0.3 MPa. If the
turbine generates 750 kW, what is the flow rate of steam through the turbine?
2.17 A steam turbine operates between 500°C and 3.5 MPa to 200 C and 0.3 MPa. If the turbine
generates 750 kW and the heat loss is 100 kW, what is the flow rate of steam through the
turbine?
2.18 Valves on steam lines are commonly encountered and you should know how they work.
For most valves, the change in velocity of the fluid flow is negligible. Apply this principle
to solve the following problems.
(a) A pressure gauge on a high-pressure steam line reads 80 bar absolute, but temperature
measurement is unavailable inside the pipe. A small quantity of steam is bled out
through a valve to atmospheric pressure at 1 bar. A thermocouple placed in the bleed
stream reads 400 C. What is the temperature inside the high-pressure duct?
(b) Steam traps are common process devices used on the lowest points of steam lines to
remove condensate. By using a steam trap, a chemical process can be supplied with so-
called dry steam, i.e., steam free of condensate. As condensate forms due to heat losses
in the supply piping, the liquid runs downward to the trap. As liquid accumulates in the
steam trap, it causes a float mechanism to move. The float mechanism is attached to a
valve, and when the float reaches a control level, the valve opens to release accumu-
lated liquid, then closes automatically as the float returns to the control level. Most
steam traps are constructed in such a way that the inlet of the steam trap valve is always
covered with saturated liquid when opened or closed. Consider such a steam trap on a
7 bar (absolute) line that vents to 1 bar (absolute). What is the quality of the stream that
exits the steam trap at 1 bar?
2.19 An overall balance around part of a plant involves three inlets and two outlets which only
contain water. All streams are flowing at steady state. The inlets are: 1) liquid at 1MPa,
25 C, m· = 54 kg/ min; 2) steam at 1 MPa, 250 C, m· 35 kg/min; 3) wet steam at 0.15
MPa, 90% quality, m· = 30 kg/min. The outlets are: 1) saturated steam at 0.8 MPa, m· = 65
kg/min; 2) superheated steam at 0.2 MPa and 300 C, m· = 54 kg/min. Two kW of work are
being added to the portion of the plant to run miscellaneous pumps and other process
equipment, and no work is being obtained. What is the heat interaction for this portion of
the plant in kW? Is heat being added or removed?
2.20 Steam at 550 kPa and 200 C is throttled through a valve at a flow rate of 15 kg/min to a
pressure of 200 kPa. What is the temperature of the steam in the outlet state, and what is
the change in specific internal energy across the value, U out – U in ?
2.21 A 0.1 m3 cylinder containing an ideal gas (CP /R 3.5) is initially at a pressure of 10 bar and
a temperature of 300 K. The cylinder is emptied by opening a valve and letting pressure
drop to 1 bar. What will be the temperature and moles of gas in the cylinder if this is accom-
plished in the following ways:
(a) Isothermally.
(b) Adiabatically. (Neglect heat transfer between the cylinder walls and the gas.)
2.22 As part of a supercritical extraction of coal, an initially evacuated cylinder is fed with
steam from a line available at 20 MPa and 400 C. What is the temperature in the cylinder
immediately after filling?
Section 2.20 Homework Problems 93
2.23 A large air supply line at 350 K and 0.5 MPa is connected to the inlet of a well-insulated
0.002 m3 tank. The tank has mass flow controllers on the inlet and outlet. The tank is at 300 K
and 0.1 MPa. Both valves are rapidly and simultaneously switched open to a flow of
0.1 mol/min. Model air as an ideal gas with CP 29.3 J/mol-K, and calculate the pressure
and temperature as a function of time. How long does it take until the tank is within 5 K of
the steady-state value?
2.24 An adiabatic tank of negligible heat capacity and 1 m3 volume is connected to a pipeline
containing steam at 10 bar and 200 C, filled with steam until the pressure equilibrates, and
disconnected from the pipeline. How much steam is in the tank at the end of the filling pro-
cess, and what is its temperature if the following occurs: