You are on page 1of 64

Advanced Thermodynamics

Exergy / Availability:
A Measure of Work Potential
Definitions

• Exergy (also called Availability or Work Potential): the


maximum useful work that can be obtained from a system at a
given state in a given environment; in other words, the most
work you can get out of a system
• Surroundings: outside the system boundaries
• Environment: the area of the surroundings not affected by the
process at any point (For example, if you have a hot turbine,
the air next to the turbine is warm. The environment is the
area of the surroundings far enough away that the
temperature isn’t affected.)
• Dead State: when a system is in thermodynamic equilibrium
with the environment, denoted by a subscript zero; at this
point no more work can be done
The energy content of the universe is constant, just as its mass
content is. Yet at times of crisis we are bombarded with speeches
and articles on how to “conserve” energy. As engineers, we know
that energy is already conserved. What is not conserved is exergy,
which is the useful work potential of the energy. Once the exergy is
wasted, it can never be recovered. When we use energy (to heat
our homes, for example), we are not destroying any energy; we are
merely converting it to a less useful form, a form of less exergy.

Exergy and the Dead State

The useful work potential of a system is the amount of energy we


extract as useful work. The useful work potential of a system at the
specified state is called exergy. Exergy is a property and is
associated with the state of the system and the environment. A
system that is in equilibrium with its surroundings has zero exergy
and is said to be at the dead state. The exergy of the thermal
energy of thermal reservoirs is equivalent to the work output of a
Carnot heat engine operating between the reservoir and the
environment.
Example

• A coal-fired furnace is used in a power plant. It


delivers 5000 kW at 1000 K. The environment is at
300 K. What is the exergy of the added heat? You
can use two steps to solve this problem.
• Determine the maximum percentage of the
heat that can be converted to work.
• Using your answer from the first part,
determine the maximum work possible.
• This is the maximum work output possible
between the given state and the dead state, i.e.,
the heat’s exergy. In this case, 30% of the 5000
kW is unavailable energy—it can’t be converted to
work.
Why Study Exergy?
• In the last several decades, exergy analysis has begun to be
used for system optimization.
• By analyzing the exergy destroyed by each component in
a process, we can see where we should be focusing our
efforts to improve system efficiency.
• It can also be used to compare components or systems
to help make informed design decisions.
Reversible Work

• Wrev (reversible work): the maximum amount of work it’s possible


to produce (or minimum necessary to input) in a process
between given initial and final states. Note that this is different
from an isentropic process where we were given an inlet state
and solved for the exit state using s2=s1. Since the exit and inlet
states are both fixed, the process is not necessarily isentropic.
• What two conditions will cause a process to be isentropic?
Irreversibilities, I

• Irreversibility, I: exergy destroyed; wasted work potential. It


represents energy that could have been converted into work but
was instead wasted
• What are some sources of I?
• To have high system efficiency, we want I to be as small as
possible.
I, cont.

• I=Wrev, out –Wu, out (work output device, like a turbine) OR


• I=Wu, in –Wrev, in (work input device, like a pump)
• Wu: useful work; the amount of work done that can actually be
used for something desirable
• Wu=W-Wsurr where W=actual work done
Surroundings Work, Wsurr

Here some work is used to


push the atmospheric air
(the surroundings) out of
the way; that work can’t be
used for other purposes.

W surr  P0 dV  P0 V 2  V1  positive

Surroundings Work, Wsurr, cont.

• Here Patm helps push the


piston in; this is gained
work. In a process where
the piston goes in and out
continually, the
surrounding work values
cancel out.
Wsurr  P0 V2  V1  negative

• What is Wsurr for a control


volume?
Second Law Efficiency, ηII

• Thermal efficiency tells us what we get out compared to what we


put in.
• The second law efficiency tells us how much we get out
compared to the maximum possible we could get out, given the
inlet and exit conditions.
Second Law Efficiency, cont.

• hth,max=1-TL/TH=1-300/800=0.635
• Say hth=0.45
• hII=0.45/0.625=0.72
• We want a high hth and hII
• Another way to look at this: for a work
output device hII=Wu/Wrev
Second Law Efficiency, cont.

• A general definition:

exergy recovered (what' s availableafter the process)


h II 
exergy supplied (what' s availableat the beginning)
exergy destroyed (I)
 1
exergy supplied
Three Efficiency Definitions

• The second two are defined for work OUTPUT


devices
Wnet
Thermal h th 
Qin
Wactual
Isentropich s 
Wisentropic
nd Wu
2 Law h II 
Wrev
Example

• A freezer is maintained at -10 oC by removing


heat from it at a rate of 3 kW. The power input to
the freezer is 2.5 kW, and the surrounding air is
30 oC. Determine a) the reversible power, b) the
irreversibility, an c) the second-law efficiency of
this freezer.

Ref: Cengel & Boles, Thermodynamics, An Engineering Approach, 4th edition, Mc-Graw Hill, 2002.):
Exergy, X
• We can calculate the exergy, X (work potential) at a given state.
The work potential is a function of the total energy of the system.

XX
• (remember
+X
KE that in +X
PE a control
internal energy
mass,
+X flow
there will bework
no flow work)
• XKE (exergy due to kinetic energy): V2/2 (on a per unit mass basis
• XPE: gZ
• Xinternal energy: u-uo+Po(v-vo)-To(s-s0)
• To see a derivation of this last equation, see the appendices on
the web site. The “o” stands for the dead state (atmospheric
conditions). If a piston is at atmospheric pressure and
temperature (the dead state), it can’t do any work.
Exergy of a Closed System

• Exergy of a closed system, per unit mass j, can be found be


adding all the terms
2
V
f   u  uo  + Po  v  vo   To  s  so  + + gZ
2
• This gives us the maximum work we could possibly get out of a
system.
• Usually we will be more interested in the change in exergy from
the beginning to end of a process.
• For a closed system,
f2  f1  Df  ?
Example 8-1

Oxygen gas is compressed in a piston-cylinder device from an initial state of 0.8


m3/kg and 25oC to a final state of 0.1 m3/kg and 287oC. Determine the reversible
work input and the increase in the exergy of the oxygen during this process. Assume
the surroundings to be at 25oC and 100 kPa.

We assume that oxygen is an ideal gas with constant specific heats. From Table A-2,
R = 0.2598 kJ/kgK. The specific heat is determined at the average temperature
T1 + T2 (25 + 287) oC
Tav    156oC
2 2
 (156 + 273)K  429K
Table A-2(b) gives Cv, ave = 0.690 kJ/kgK.
The entropy change of oxygen is
T  v 
s2  s1  Cv, ave ln  2  + R ln  2 
 T1   v1 
 m3 
kJ  (287 + 273) K  kJ  0.1 
 kg 
 0.690 ln   + 0.2598 ln
kg  K  (25 + 273) K  kg  K  m3 
 0.8 kg 
 
kJ
 0.105
kg  K

We calculate the reversible work input, which represents the minimum work input
Wrev,in in this case, from the exergy balance by setting the exergy destruction equal to
zero.
X in  X out  X destroyed  DX system
14 2 43 14 2 43 14 2 43
Net exergy transfer Exergy Change
by heat, work, and mass destruction in exergy

Wrev,in  X 2  X 1
Therefore, the change in exergy and the reversible work are identical in this case.
Substituting the closed system exergy relation, the reversible work input during this
process is determined to be
wrev,in  f2  f1
 (u2  u1 ) + P0 (v2  v1 )  T0 ( s2  s1 )
 Cv,ave (T2  T1 ) + P0 (v2  v1 )  T0 ( s2  s1 )
kJ m3 kJ
 0.690 (287  25)K + 100 kPa(0.1  0.8)
kg  K kg m3kPa
kJ
 (25 + 273)K( 0.105 )
kg  K
kJ
 142.1
kg
The increase in exergy of the oxygen is
kJ
x2  x1  f2  f1  wrev, in  142.1
kg
For a control volume

• Xcv=Xclosed+Xflow work
• y=Xcv/m (exergy per unit mass)
• Xflow work=Wflow-Wagainst atmosphere=Pv-Pov
V2
y cv   u  u o  + Po v  Po vo  To  s  so  + + gz + Pv  Po vo
2
• Now combine terms: u+Pv=h; 2
uo+Povo=ho
V
y cv   h  ho   To  s  so  + + gz
2
Change in exergy

• If we only have one fluid stream


V22  V12
y 2 y 1  Dy   h2  h1   To  s 2  s1  + + g  z 2  z1 
2

• If we have multiple streams


 V 2   V 2 
Dy   m 2  h2  To s 2 +


2
2
+ gz2  


 m 1  h1  To s1 +


1
2
+ gz1 


Exergy Balance

• We will use these equations in an exergy balance to solve for


such quantities as reversible work or exergy destroyed.
• Xin-Xout-Xdestroyed=DXsys

• Xdestroyed is potential work that was destroyed due to


irreversibilities like friction.
• Exergy can be transferred (Xin-Xout) by heat, work, and mass
flow
Exergy Transfer by Heat Transfer

• As we add heat to a system, we increase


its ability to do work.

 To 
Wmax  X heat  QH h max  QH 1  
 TH 
Exergy Transfer by Work and Mass Flow

• If we do work on a system, we increase its ability to


do work.
• Xwork=W-Wsurr for boundary work
• Xwork=W for all other kinds of work
• Remember Wsurr  P0 V2  V1 
• Xmass=my
Xdestroyed

• Xdestroyed=I=ToSgen
• Recall
• DSsys=Sin-Sout+Sgen
Entropy Generated, Sgen

• For a steady-state control volume, this leads us to


Q k
S gen   m s   m s  
out
e e
in
i i
Tk
• For a control mass, this becomes
Qk
S gen  S 2  S1   Tk

• Here Tk is the temperature of the heat source or heat sink (not


the system temperature).
Final Equation for DXsys for control mass

 To 
 1 
 T
 k
Qk  W  Po V2  V1    To S gen  X 2  X 1


• Terms in [ ] are W-Wsurr=Wu
• If we want to find Wrev, then ToSgen=0 and
Wu=Wrev
• Note that if heat transfer is to/from the
surroundings, the Q term drops out.
Example

• A 12-f3 rigid tank contains R-134a at 30 psia and 40%


quality. Heat is transferred now to the refrigerant from a
source at 120°F until the pressure rises to 60 psia.
Assuming the surroundings to be at 75°F, determine a)
the amount of heat transfer between the source and
the refrigerant and b) the exergy destroyed during the
process.

Ref: Cengel and Boles


Final Equation for DXsys for control
volume

For multiple fluid streams, unsteady flow:


 To 
 1 
 T
 k
Qk  W  Po V2  V1    To S gen +


my  m y
i i e e  X 2  X1

For one fluid streams, steady flow:


 To 
 1 
 T
 k
Qk  W  To S gen + m y i y e   0


To find Wrev, set Sgen=0. If adiabatic, Q=0.
Example 8-2

Steam enters an adiabatic turbine at 6 MPa, 600C, and 80 m/s and leaves at 50
kPa, 100C, and 140 m/s. The surroundings to the turbine are at 25C. If the power
output of the turbine is 5MW, determine
(a)the power potential of the steam at its inlet conditions, in MW.
(b) the reversible power, in MW.
(c)the second law efficiency.

We assume steady-flow and neglect changes in potential energy.


The mass flow rate of the steam is determined from the steady-flow energy equation
applied to the actual process,
0 (steady)
E&in  E&out  DE&systems
14 2 43 14 2 43
Rate of net energy transfer Rate of change
by heat, work, and mass of energy
r2 r2
V V
m&1 (h1 + 1 )  m&2 (h2 + 2 )  W&out  0
2 2
Conservation of mass for the steady flow gives
m&in  m&out  Dm&system
14 2 43 14 2 43
Rate of net mass transfer Rate of change
of mass

m&1  m&2  0
m&1  m&2  m&
The work done by the turbine and the mass flow rate are
r2 r2
  V1 V2  
&
Wout  m (h1  h2 ) + 
&  
  2 2  
W&out
m&
(h1  h2 )  Dke
where
r r
 V22 V12 
Dke    
 2 2 
(140 m/s) 2  (80 m/s) 2  1kJ/kg 
  2 2 
2  1000 m /s 
kJ
 6.6
kg
From the steam tables:
 kJ
h  3658.8
P1  6 MPa   1 kg
 
T1  600o C   kJ
s1  7.1693
 kg  K
 kJ
h  2682.4
P2  50 kPa   2
kg

T2  100o C   kJ
s2  7.6953
 kg  K
 kJ
h  h  104.83
P0  100 kPa   0 o
f@25 C
kg

T0  25o C   kJ
s0  sf@25oC  0.3672
 kg  K
W&out
m&
(h1  h2 )  Dke
5 MW  1000 kJ/s 

kJ  MW 
(3658.8  2682.4  6.6)
kg
kg
 5.16
s
The power potential of the steam at the inlet conditions is equivalent to its exergy at
the inlet state. Recall that we neglect the potential energy of the flow.
r2 0
&  V1 
1  m&y 1  m& (h1  h0 )  T0 ( s1  s0 ) + + gz1 
 2 
 kJ kJ 
 (3658.8  104.83) kg  (298 K )(7.1693  0.3672) kg  K 
&  5.16 kg 
 
s  (80m/s)  kJ/kg  
1 2

 +  2 2  
 2  1000 m /s  
kg  kJ   MW 
 5.16  1533.3   
s  kg  1000 kJ/s 
 7.91MW
The power output of the turbine if there are no irreversibilities is the reversible power
and is determined from the rate form of the exergy balance applied on the turbine and
setting the exergy destruction term equal to zero.
0 0 (steady)
X&in  X&out  X&destroyed  DX&system
14 2 43 14 2 43 14 2 43
Rate of net exergy transfer Rate of exergy Rate of change
by heat, work, and mass destruction of exergy

X&in  X&out
m&y  W&
1 rev, out + m&y 2
W&rev, out  m&(y 2 y 1 ) 0

 m& (h1  h2 )  T0 ( s1  s2 )  Dke  Dpe 


 kJ kJ 
 (3658.8  2682.4) kg  (298 K )(7.1693  7.6953) kg  K 
kg
W&rev, out  5.16  
s  kJ 
 6.6 
 kg 
kg  kJ   MW 
 5.16 1126.5   
s  kg   1000 kJ/s 
 5.81MW
The second-law efficiency is determined from
W& 5 MW
hII  &   86.1%
W
rev 5.81MW
Set up the following problems.

1. Refrigerant at T1 and P1 is throttled to a pressure of P2. Find the


reversible work and exergy destroyed during this process. The
atmosphere has a temperature of To.
2. Air at T1 and P1 with a velocity of V1 enters a nozzle and exits at
P2 and T2 with a velocity of V2. There is a heat loss Q from the
nozzle to the surroundings at To. Find the exergy destroyed
during this process.
3. Air enters a compressor at ambient conditions (To and Po) and
leaves at P2 and T2. The compressor is deliberately cooled, and
there is a rate of heat loss of Q to the surroundings. The power
input to the compressor is PWR. Find the rate of irreversibility,
I, for this process.
Exergy Analysis for a Cycle, 1 fluid stream, steady flow

I  To S gen
for a component :

S  m  s  s   Q
gen e i
Tk
Sgen  S + S
gen, pump gen,boiler + S gen,turbine + S gen,cond.
 Q pump   Qboiler   Qturbine   Qcond., in 

 m  s 2  s1   +  s3  s 2   +  s 4  s3   +  s1  s 4  

To   
Tcomb.chamber    To     Tlake 

 Qout Qin 
I  T m 
o 

 T k ,out


Tk ,in 
Second Law Efficiency for a Cycle

W net ,actual W net , actual


h II  
W net , reversible W net , actual + I
Exergy Forms
Exergy of kinetic energy

Kinetic energy is a form of mechanical energy and can be converted directly into
work. Kinetic energy itself is the work potential or exergy of kinetic energy
independent of the temperature and pressure of the environment.
r
V2
Exergy of kinetic energy: xke  ke  (kJ/kg)
2
Exergy of potential energy

Potential energy is a form of mechanical energy and can be converted directly into
work. Potential energy itself is the work potential or exergy of potential energy
independent of the temperature and pressure of the environment.

Exergy of potential energy: xpe  pe  gz (kJ/kg)


Useful Work

The work done by work producing devices is not always entirely in a useable form.
Consider the piston-cylinder device shown in the following figure.

The work done by the gas expanding in the piston-cylinder device is the boundary work and can
be written as
 W  P dV  ( P  P0 ) dV + P0 dV
  Wb, useful + P0 dV

The actual work done by the gas is


W  Wb, useful +  P0 dV
 Wb, useful + P0 (V2  V1 )
The word done on the surroundings is

Wsurr   P0 dV  P0 (V2  V1 )
Any useful work delivered by a piston-cylinder device is due to the pressure above the
atmospheric level.
Wu  W  Wsurr
Reversible Work

Reversible work Wrev is defined as the maximum amount of useful work that can be
produced (or the minimum work that needs to be supplied) as a system undergoes a
process between the specified initial and final states. This is the useful work output
(or input) obtained when the process between the initial and final states is executed in
a totally reversible manner.

Irreversibility

The difference between the reversible work Wrev and the useful work Wu is due to the
irreversibilities present during the process and is called the irreversibility I. It is
equivalent to the exergy destroyed and is expressed as
I  X destroyed  T0 Sgen  Wrev, out  Wu, out  Wu, in  Wrev, in
where Sgen is the entropy generated during the process. For a totally reversible
process, the useful and reversible work terms are identical and thus irreversibility is
zero.

Irreversibility can be viewed as the wasted work potential or the lost opportunity to do
work. It represents the energy that could have been converted to work but was not.

Exergy destroyed represents the lost work potential and is also called the wasted
work or lost work.

Second-Law Efficiency

The second-law efficiency is a measure of the performance of a device relative to the


performance under reversible conditions for the same end states and is given by
hth Wu
hII  
hth , rev Wrev
for heat engines and other work-producing devices and
COP Wrev
hII  
COPrev Wu
for refrigerators, heat pumps, and other work-consuming devices.

In general, the second-law efficiency is expressed as


Exergy recovered Exergy destroyed
h II   1
Exergy supplied Exergy supplied
Exergy of change of a system

Consider heat transferred to or from a closed system whenever there is a


temperature difference across the system boundary. The exergy for a system may be
determined by considering how much of this heat transfer is converted to work
entirely. Let’s take a second look at the following figure.
Taking the heat transfer to be from the system to its surroundings, the conservation of
energy is
 Ein   Eout  dEsystem
0   Q   W  dU
The work is the boundary work and can be written as
 W  P dV  ( P  P0 ) dV + P0 dV
  Wb, useful + P0 dV
Any useful work delivered by a piston-cylinder device is due to the pressure above the
atmospheric level.

To assure the reversibility of the process, the heat transfer occurs through a reversible
heat engine.
T0 Q
 WHE  hth Q  (1  ) Q   Q  T0
T T
 Qnet  Q
dS  
T T
 WHE   Q + T0 dS
 Q   WHE  T0 dS
   WHE  T0 dS     Wb, useful + P0 dV   dU
 Wtotal useful   Wb, useful +  WHE
  dU  P0 dV + T0 dS
Integrating from the given state (no subscript) to the dead state (0 subscript), we have
Wtotal useful  (U 0  U )  P0 (V0  V ) + T0 ( S0  S )
 (U  U 0 ) + P0 (V  V0 )  T0 ( S  S 0 )
This is the total useful work due to a system undergoing a reversible process from a
given state to the dead state, which is the definition of exergy.

Including the kinetic energy and potential energy, the exergy of a closed system is
r2
V
X  (U  U 0 ) + P0 (V  V0 )  T0 (S  S 0 ) + m + mgz
2
on a unit mass basis, the closed system (or nonflow) exergy is
r2
V
f  (u  u0 ) + P0 (v  v0 )  T0 ( s  s0 ) + + gz
2
 (e  e0 ) + P0 (v  v0 )  T0 ( s  s0 )
Here, u0, v0, and s0 are the properties of the system evaluated at the dead state. Note
that the exergy of the internal energy of a system is zero at the dead state is zero
since u = u0, v = v0, and s = s0 at that state.

The exergy change of a closed system during a process is simply the difference
between the final and initial exergies of the system,
DX  X 2  X 1  m(f2  f1 )
 ( E  E0 ) + P0 (V  V0 )  T0 ( S  S 0 )
r r
V22  V12
 (U  U 0 ) + P0 (V  V0 )  T0 ( S  S0 ) + m + mg ( z2  z1 )
2
On a unit mass basis the exergy change of a closed system is
Df  (f2  f1 )
 (e  e0 ) + P0 (v  v0 )  T0 ( s  s0 )
r2 r2
V  V1
 (u  u0 ) + P0 (v  v0 )  T0 ( s  s0 ) + 2 + g ( z2  z1 )
2
Exergy of flow

The energy needed to force mass to flow into or out of a control volume is the flow
work per unit mass given by (see Chapter 3).
wflow  Pv

The exergy of flow work is the excess of flow work done against atmospheric air at P0
to displace it by volume v. According to the above figure, the useful work potential
due to flow work is
wflow, energy  Pv  P0 v
Thus, the exergy of flow energy is
xflow energy  Pv  P0 v  ( P  P0 )v
Flow Exergy

Since flow energy is the sum of nonflow energy and the flow energy, the exergy of
flow is the sum of the exergies of nonflow exergy and flow exergy.
x flowing fluid  xnonflowing fluid + xflow exergy
r2
V
 (u  u0 ) + P0 (v  v0 )  T0 ( s  s0 ) + + gz + ( P  P0 )v
2
r2
V
 (u + Pv )  (u0 + P0 v0 )  T0 ( s  s0 ) + + gz
2
r2
V
 (h  h0 )  T0 ( s  s0 ) + + gz
2
The flow (or stream) exergy is given by
r2
V
y  (h  h0 )  T0 ( s  s0 ) + + gz
2
The exergy of flow can be negative if the pressure is lower than atmospheric
pressure.

The exergy change of a fluid stream as it undergoes a process from state 1 to state 2
is r2 r2
V  V1
Dy  y 2 y 1  (h2  h1 )  T0 ( s2  s1 ) + 2 + g ( z2  z1 )
2
Exergy Transfer by Heat, Work, and Mass

Exergy can be transferred by heat, work, and mass flow, and exergy transfer
accompanied by heat, work, and mass transfer are given by the following.

Exergy transfer by heat transfer

By the second law we know that only a portion of heat transfer at a temperature above
the environment temperature can be converted into work. The maximum useful work is
produced from it by passing this heat transfer through a reversible heat engine. The
exergy transfer by heat is

 T 
Exergy transfer by heat: X heat   1  0  Q
 T 
Note in the above figure that entropy generation is always by exergy destruction and
that heat transfer Q at a location at temperature T is always accompanied by entropy
transfer in the amount of Q/T and exergy transfer in the amount of (1-T0/T)Q.

Note that exergy transfer by heat is zero for adiabatic systems.


Exergy transfer by work

Exergy is the useful work potential, and the exergy transfer by work can simply be
expressed as
W  Wsurr (for boundary work)
Exergy transfer by work: X work  
W (for other forms of work)

where Wsurr  P0 (V2  V1 ), P0 is atmospheric pressure, and V1 and V2 are the initial and
final volumes of the system. The exergy transfer for shaft work and electrical work is
equal to the work W itself.

Note that exergy transfer by work is zero for systems that have no work.
Exergy transfer by mass

Mass flow is a mechanism to transport exergy, entropy, and energy into or out of a
system. As mass in the amount m enters or leaves a system the exergy transfer is
given by

Exergy transfer by mass: X mass  my


Note that exergy transfer by mass is zero for systems that involve no flow.

The Decrease of Exergy Principle and Exergy Destruction

The exergy of an isolated system during a process always decreases or, in the
limiting case of a reversible process, remains constant. This is known as the
decrease of exergy principle and is expressed as
DX isolated  ( X 2  X 1 )isolated  0
Exergy Destruction

Irreversibilities such as friction, mixing, chemical reactions, heat transfer through finite
temperature difference, unrestrained expansion, non-quasi-equilibrium compression,
or expansion always generate entropy, and anything that generates entropy always
destroys exergy. The exergy destroyed is proportional to the entropy generated as
expressed as
X destroyed  T0 Sgen
The decrease of exergy principle does not imply that the exergy of a system cannot
increase. The exergy change of a system can be positive or negative during a
process, but exergy destroyed cannot be negative. The decrease of exergy principle
can be summarized as follows:
 0 Irreversible proces

X destroyed  0 Reversible process
 0 Impossible process

Exergy Balances

Exergy balance for any system undergoing any process can be expressed as
 Total   Total   Total   Change in the 
       
 exergy 
  exergy 
  exergy 
  total exergy 
 entering   leaving   destroyed   of the system 
       
General:
X in  X out  X  DX
14 2 43 14destroyed
2 43 14 2system
43
Net exergy transfer Exergy Change
by heat, work, and mass destruction in exergy
General, rate
form:

X&in  X&out  X&destroyed  DX&system


14 2 43 14 2 43 14 2 43
Rate of net exergy transfer Rate of exergy Rate of change
by heat, work, and mass destruction of exergy
General, unit-mass
basis:
( xin  xout )  xdestroyed  Dxsystem
wh
ere
 T 
X&heat  1  0  Q&
 T 
X&work  W&useful
X&  m&y
mass

DX&system  dX system / dt

For a reversible process, the exergy destruction term, Xdestroyed, is zero.

Considering the system to be a general control volume and taking the positive
direction of heat transfer to be to the system and the positive direction of work
transfer to be from the system, the general exergy balance relations can be
expressed more explicitly as
 T0 
 1  T  Qk   W  P0 (V2  V1 )  +  miy i   mey e  X destroyed  X 2  X 1
 k 

 T0  &  & dVCV  dX CV


 1  T  Qk  W  P0
 dt 
+  &
miy i   &
mey e  &
X destroyed 
dt
 k 
where the subscripts are i = inlet, e = exit, 1 = initial state, and 2 = final state of the
system. For closed systems, no mass crosses the boundaries and we omit the terms
containing the sum over the inlets and exits.
Example 8-1

Oxygen gas is compressed in a piston-cylinder device from an initial state of 0.8


m3/kg and 25oC to a final state of 0.1 m3/kg and 287oC. Determine the reversible
work input and the increase in the exergy of the oxygen during this process. Assume
the surroundings to be at 25oC and 100 kPa.

We assume that oxygen is an ideal gas with constant specific heats. From Table A-2,
R = 0.2598 kJ/kgK. The specific heat is determined at the average temperature
T1 + T2 (25 + 287) oC
Tav    156oC
2 2
 (156 + 273)K  429K
Table A-2(b) gives Cv, ave = 0.690 kJ/kgK.
The entropy change of oxygen is
T  v 
s2  s1  Cv, ave ln  2  + R ln  2 
 T1   v1 
 m3 
kJ  (287 + 273) K  kJ  0.1 
 kg 
 0.690 ln   + 0.2598 ln
kg  K  (25 + 273) K  kg  K  m3 
 0.8 kg 
 
kJ
 0.105
kg  K

We calculate the reversible work input, which represents the minimum work input
Wrev,in in this case, from the exergy balance by setting the exergy destruction equal to
zero.
X in  X out  X destroyed  DX system
14 2 43 14 2 43 14 2 43
Net exergy transfer Exergy Change
by heat, work, and mass destruction in exergy

Wrev,in  X 2  X 1
Therefore, the change in exergy and the reversible work are identical in this case.
Substituting the closed system exergy relation, the reversible work input during this
process is determined to be
wrev,in  f2  f1
 (u2  u1 ) + P0 (v2  v1 )  T0 ( s2  s1 )
 Cv,ave (T2  T1 ) + P0 (v2  v1 )  T0 ( s2  s1 )
kJ m3 kJ
 0.690 (287  25)K + 100 kPa(0.1  0.8)
kg  K kg m3kPa
kJ
 (25 + 273)K( 0.105 )
kg  K
kJ
 142.1
kg
The increase in exergy of the oxygen is
kJ
x2  x1  f2  f1  wrev, in  142.1
kg
Example 8-2

Steam enters an adiabatic turbine at 6 MPa, 600C, and 80 m/s and leaves at 50
kPa, 100C, and 140 m/s. The surroundings to the turbine are at 25C. If the power
output of the turbine is 5MW, determine
(a)the power potential of the steam at its inlet conditions, in MW.
(b) the reversible power, in MW.
(c)the second law efficiency.

We assume steady-flow and neglect changes in potential energy.


The mass flow rate of the steam is determined from the steady-flow energy equation
applied to the actual process,
0 (steady)
E&in  E&out  DE&systems
14 2 43 14 2 43
Rate of net energy transfer Rate of change
by heat, work, and mass of energy
r2 r2
V V
m&1 (h1 + 1 )  m&2 (h2 + 2 )  W&out  0
2 2
Conservation of mass for the steady flow gives
m&in  m&out  Dm&system
14 2 43 14 2 43
Rate of net mass transfer Rate of change
of mass

m&1  m&2  0
m&1  m&2  m&
The work done by the turbine and the mass flow rate are
r2 r2
  V1 V2  
&
Wout  m (h1  h2 ) + 
&  
  2 2  
W&out
m&
(h1  h2 )  Dke
where
r r
 V22 V12 
Dke    
 2 2 
(140 m/s) 2  (80 m/s) 2  1kJ/kg 
  2 2 
2  1000 m /s 
kJ
 6.6
kg
From the steam tables:
 kJ
h  3658.8
P1  6 MPa   1 kg
 
T1  600o C   kJ
s1  7.1693
 kg  K
 kJ
h  2682.4
P2  50 kPa   2
kg

T2  100o C   kJ
s2  7.6953
 kg  K
 kJ
h  h  104.83
P0  100 kPa   0 o
f@25 C
kg

T0  25o C   kJ
s0  sf@25oC  0.3672
 kg  K
W&out
m&
(h1  h2 )  Dke
5 MW  1000 kJ/s 

kJ  MW 
(3658.8  2682.4  6.6)
kg
kg
 5.16
s
The power potential of the steam at the inlet conditions is equivalent to its exergy at
the inlet state. Recall that we neglect the potential energy of the flow.
r2 0
&  V1 
1  m&y 1  m& (h1  h0 )  T0 ( s1  s0 ) + + gz1 
 2 
 kJ kJ 
 (3658.8  104.83) kg  (298 K )(7.1693  0.3672) kg  K 
&  5.16 kg 
 
s  (80m/s)  kJ/kg  
1 2

 +  2 2  
 2  1000 m /s  
kg  kJ   MW 
 5.16  1533.3   
s  kg  1000 kJ/s 
 7.91MW
The power output of the turbine if there are no irreversibilities is the reversible power
and is determined from the rate form of the exergy balance applied on the turbine and
setting the exergy destruction term equal to zero.
0 0 (steady)
X&in  X&out  X&destroyed  DX&system
14 2 43 14 2 43 14 2 43
Rate of net exergy transfer Rate of exergy Rate of change
by heat, work, and mass destruction of exergy

X&in  X&out
m&y  W&
1 rev, out + m&y 2
W&rev, out  m&(y 2 y 1 ) 0

 m& (h1  h2 )  T0 ( s1  s2 )  Dke  Dpe 


 kJ kJ 
 (3658.8  2682.4) kg  (298 K )(7.1693  7.6953) kg  K 
kg
W&rev, out  5.16  
s  kJ 
 6.6 
 kg 
kg  kJ   MW 
 5.16 1126.5   
s  kg   1000 kJ/s 
 5.81MW
The second-law efficiency is determined from
W& 5 MW
hII  &   86.1%
W
rev 5.81MW

You might also like