Вы находитесь на странице: 1из 597

Lecture 1

Systems and Measurement

Sections 1.1 1.4

Why Thermodynamics?
And where these techniques applied?

infrastructure
ecosystems

transportation

To learn about each of these, we cant treat everything all at once!

boundary

System

surroundings/
environment

Types of systems

Isolated

Closed
(control mass)

insulated bath

piston in cylinder

Open
(control volume)

engine

Example
Are these systems isolated, closed or open?
candle in a dark room

electric car sliding on frictionless


ice, assume air drag is negligible,
boundary at surface of car

greenhouse, windows closed,


boundary just outside structure

dog, boundary at surface

How do you choose the boundary?


Depends on:
What you know
What you want to know
conventional car,
sliding on surface
with k = 0.1

air intake
exhaust gas
-> external wheel speed

What are we looking at, when looking at the system?


2 Regimes

Macroscopic

Large scale features


Temperature
Pressure
Volume

Microscopic

System of particles
Molecular, atomic, quantum
energy levels

Ways we describe a system


Properties
independent of history of the
system
have numerical value
can be measured or
computed by looking at the
system macroscopically

Examples:
pressure, temperature, mass, volume
A test if a descriptor of a system is a property
A property is a property if, and only if, its change in value between two states is
independent of the process.

Ways we describe a system


State
the set of particular values the
properties have at a particular time

P = 1.203 x 105 Pa
T = 385 K
V = 0.3 m3

Ways we describe a system


Process
how a system changes from one
state to another over time

steady state

properties arent changing with time

equilibrium

no net energy is transferred between system and surroundings

Watch the video on


Equilibrium vs steady state
http://tll.mit.edu/help/equilibrium-vs-steady-state

from 1:30 7:30

(Of course, you can watch the rest if you like. It does describe some
engineering thought experiments, measurement and analysis but some
of the concepts are beyond the scope of this lecture)

Intensive Properties

does not depend on the amount


can be location or time dependent

Temperature
Color
Pressure
Density

depends on the amount


not location dependent
can vary with time

Mass
Length
Volume
Shape
Energy

primary

Quantities, Units, and their Relationships

secondary
(derived)

=
SI
english

1 N = 1 kg m/s2
1 lbf = 1 lb * 32.1740 ft/s2

Example (P1.13)
At a certain elevation, the pilot of a balloon has a mass of 120 lb and a weight
of 119 lbf. What is the local acceleration of gravity, in ft/s2, at that elevation?

If the balloon drifts to another elevation where g = 32.05 ft/s2,what is her


weight in lbf, and mass in lb?

Work through For Example problems


in section 1.4

then

Work on homework
problem 1.10, 1.17

Lecture 3

Measurements and Units

Sections 1.5 1.8

Density (r) and Specific Volume (v)


From a macroscopic perspective, description
of matter is simplified by considering it to be
distributed continuously throughout a region.
When substances are treated as continua, it
is possible to speak of their intensive
thermodynamic properties at a point.
At any instant the density (r ) at a point is
defined as
m
r lim
V V ' V

(Eq. 1.6)

where V ' is the smallest volume for which a definite


value of the ratio exists.

Density (r) and Specific Volume (v)


Density is mass per unit volume.
Density is an intensive property that may
vary from point to point.
SI units are (kg/m3).
English units are (lb/ft3).

Density (r) and Specific Volume (v)


Specific volume is the reciprocal of
density: v = 1/r .
Specific volume is volume per unit mass.
Specific volume is an intensive property
that may vary from point to point.
SI units are (m3/kg).
English units are (ft3/lb).
Specific volume is usually preferred for
thermodynamic analysis when working with
gases that typically have small density values.

Molar Quantities of Matter (n)


We can quantify how much matter we have on a
molar basis instead:
n: mol, kmol, lbmol

m
n
M

(Eq. 1.8)

where M is molecular weight and


n (kmol)
n (lbmol)
n (mol, gmol)

m (kg)
m (lb)
m (g)

NA = 6.022 x 10^23 molecules/mol

M (kg/kmol)
M (lb/lbmol)
M (g/mol)

Example:
A) Is Avogadros number the same when written in terms of
# of molecules/lbmol? (1 lb = 453.6 g)

B) To the right is a list of molecular weights. If the table is in


g/mol, what would the values be in kg/kmol and lb/lbmol?

Notation: (for some variable x)


x: x in terms of mol, kmol, lbmol
V

(multiply both sides by M)

n
M

M
M V
m
1
V
V
n
n

(Eq. 1.9)

Pressure (p)
Consider a small area A passing through a point
in a fluid at rest.
The fluid on one side of the area exerts a
compressive force that is normal to the area, Fnormal.
An equal but oppositely directed force is exerted on
the area by the fluid on the other side.
The pressure (p) at the specified point is defined
as the limit
F

p lim normal
A
A A '

(Eq. 1.10)

where A' is the area at the point in the same limiting sense
as used in the definition of density.

Pressure Units
SI unit of pressure is the pascal:
1 pascal = 1 N/m2
Multiples of the pascal are frequently used:
1 kPa = 103 N/m2
1 bar = 105 N/m2
1 MPa = 106 N/m2

English units for pressure are:


pounds force per square foot, lbf/ft2
pounds force per square inch, lbf/in.2

Absolute Pressure
Absolute pressure: Pressure with respect to
the zero pressure of a complete vacuum.
Absolute pressure must be used in
thermodynamic relations.
Pressure-measuring devices often indicate
the difference between the absolute pressure of
a system and the absolute pressure of the
atmosphere outside the measuring device.

Gage and Vacuum Pressure


When system pressure is greater than
atmospheric pressure, the term gage
pressure is used.
p(gage) = p(absolute) patm(absolute)
(Eq. 1.14)

When atmospheric pressure is


greater than system pressure, the term
vacuum pressure is used.
p(vacuum) = patm(absolute) p(absolute)
(Eq. 1.15)

Work through For Example problem


in section 1.6

then

Work on homework problems


1.26, 1.40

Temperature (T)
If two blocks (one warmer than the other) are
brought into contact and isolated from their
surroundings, they would interact thermally with
changes in observable properties.
When all changes in observable properties cease,
the two blocks are in thermal equilibrium.
Temperature is a physical property that
determines whether the two objects are in thermal
equilibrium.

Thermometers
Any object with at least one measurable property
that changes as its temperature changes can be
used as a thermometer.
Such a property is called a thermometric
property.
The substance that exhibits changes in the
thermometric property is known as a thermometric
substance.

Thermometers
Example: Liquid-in-glass thermometer
Consists of glass capillary tube connected to a bulb filled
with liquid and sealed at the other end. Space above liquid
is occupied by vapor of liquid or an inert gas.
As temperature increases, liquid expands in volume and
rises in the capillary. The length (L) of the liquid in the
capillary depends on the temperature.
The liquid is the thermometric substance.
L is the thermometric property.

Other types of thermometers:


Thermocouples
Thermistors
Radiation thermometers and optical pyrometers

Temperature Scales
Kelvin scale: An absolute thermodynamic temperature
scale whose unit of temperature is the kelvin (K); an SI base
unit for temperature.
Rankine scale: An absolute thermodynamic temperature
scale with absolute zero that coincides with the absolute zero
of the Kelvin scale; an English base unit for temperature.
T(oR) = 1.8T(K)

(Eq. 1.16)

Celsius scale (oC):


T(oC) = T(K) 273.15 (Eq. 1.17)
Fahrenheit scale (oF):

T(oF) = T(oR) 459.67 (Eq. 1.18)

Work through For Example problem


in section 1.7

then

Work on homework problems


1.52, 1.56

Problem-Solving Methodology
Known: Read the problem, think about it, and
identify what is known.
Find: State what is to be determined.
Schematic and Given Data: Draw a sketch of
system and label with all relevant information/data.
Engineering Model: List all simplifying
assumptions and idealizations made.
Analysis: Reduce appropriate governing
equations and relationships to forms that will
produce the desired results.

Lecture 4

Energy and Work

Sections 2.1 2.3

(not 2.2.6, 2.2.7, 2.2.8)

Closed System Energy Balance


Energy is an extensive property that
includes the kinetic and gravitational potential
energy of engineering mechanics.
For closed systems, energy is transferred in
and out, across the system boundary, by two
means only: by work and by heat.
Energy is conserved
1st law of thermodynamics

Closed System Energy Balance


The closed system energy balance states:
The change in the
amount of energy
contained within
a closed system
during some time
interval

Net amount of energy


transferred in and out
across the system boundary
by heat and work during
the time interval

Lets look deeper into this energy balance,


including what we mean by energy change
and energy transfer.

Change in Energy of a System


In engineering thermodynamics the change
in energy of a system has three parts:
Kinetic energy
Gravitational potential energy
Internal energy

Change in Kinetic Energy


The change in kinetic energy is associated with
the motion of the system as a whole relative to
an external coordinate frame such as the surface
of the earth.
For a system of mass m the change in kinetic
energy from state 1 to state 2 is

1
2
2
m

V
V
KE = KE2 KE1 =
2
1
2
where

(Eq. 2.5)

V1 and V2 are the initial and final velocity magnitudes.


is the final value minus initial value.

Change in Gravitational Potential Energy


The change in gravitational potential energy is
associated with the position of the system in the
earths gravitational field.
For a system of mass m the change in potential
energy from state 1 to state 2 is
PE = PE2 PE1 = mg(z2 z1)
where

(Eq. 2.10)

z1 and z2 are the initial and final elevations relative to


the surface of the earth, respectively.
g is acceleration due to gravity.

Change in Internal Energy


The change in internal energy is associated with the
makeup of the system, including its chemical
composition.
There is no simple expression like Eqs. 2.5 and 2.10 for
finding internal energy change for a wide range of
applications. Usually, we will be able to use the data
from tables in appendices of the textbook.
Like kinetic and gravitational potential energy, internal
energy is an extensive property.
Internal energy is represented by U.
The specific internal energy on a mass basis is u.
The specific internal energy on a molar basis is u .

Change in Energy of a System


So, the total change in energy of a system from
state 1 to state 2 is

E2 E1 = (U2 U1) + (KE2 KE1) + (PE2 PE1)


(Eq. 2.27a)

E = U + KE + PE

(Eq. 2.27b)

Since an arbitrary value E1 can be given for the


energy of a system at state 1, no particular
significance can be attached to the value of
energy at state 1 or any other state. Only
changes in the energy of a system between
states have significance.

Energy Transfer by Work


Energy can be transferred to and from
closed systems by two means only:
Work
Heat

You have studied work in mechanics and


those concepts will stay the same in the study
of thermodynamics. However,
thermodynamics deals with phenomena not
included within the scope of mechanics, and
this requires a broader interpretation of work.

Illustrations of Work
When a spring is compressed,
energy is transferred to the spring by
work.
When a gas in a closed vessel is
stirred, energy is transferred to the
gas by work.
When a battery is charged
electrically, energy is transferred to
the battery contents by work.
The first two examples of work are familiar from
mechanics. The third example is an example of this
broader interpretation of work.

Energy Transfer by Work


The symbol W denotes an amount of energy
transferred across the boundary of a system by work.
Since engineering thermodynamics is often
concerned with internal combustion engines,
turbines, and electric generators whose purpose is to
do work, convention is that the work done by a
system as positive.
W > 0: work done by the system
W < 0: work done on the system

The same sign convention is used for the rate of


energy transfer by work called power, given by W .

Modeling Expansion and Compression Work


We will see, a useful example to study is a gas (or
liquid) undergoing an expansion (or compression)
process while confined in a piston-cylinder
assembly.

During the process, the gas exerts a normal force


on the piston, F = pA , where p is the pressure at the
interface between the gas and piston and A is the
area of the piston face.

Modeling Expansion and Compression Work


From mechanics, the work done by the gas as the
piston face moves from x1 to x2 is given by

W = Fdx =

pAdx

Since the product Adx = dV , where V is the volume


of the gas, this becomes

W=

V2

pdV

(Eq. 2.17)

V1

For compression, dV is negative and so is the


value of the integral, which keeps with the sign
convention for work.

Modeling Expansion and Compression Work


To do the integral in Eq. 2.17 we need to know a relationship
between gas pressure at the interface between the gas and
piston and the total gas volume. During an actual expansion of
a gas such a relationship may be difficult, or even impossible, to
obtain because of non-equilibrium effects (e.g. combustion in
the cylinder of an automobile engine).
Then, the work value must be obtained by experiment.
Eq. 2.17 can be used to find the work of idealized processes
when the pressure p is the pressure of the entire quantity of gas
undergoing the process and not only the pressure at the piston
face.
Usually then we can approximate the gas as undergoing a
sequence of equilibrium states, which we call a
quasiequilibrium process.

Quasiequilibrium Processes
In a quasiequilibrium
expansion, the gas moves
along a pressure-volume
curve, or path, as shown.
Using Eq. 2.17, the work
done by the gas on the
piston is given by the area
under the curve of pressure
versus volume.

Polytropic Processes
When we can write the path as a function, we can
evaluate the work integral and simplify the equation.
An example is a quasiequilibrium process where
the function can be written as pVn = constant , we call
the process a polytropic process (where n is a
constant called the polytropic index).
For the case n = 1, pV = constant and Eq. 2.17 gives
V2
W = (constant ) ln
V1

where constant = p1V1 = p2V2.

Remember: Work calculated using a quasiequilibrium model is an approximation.

Polytropic Indices
n=0
n<1
n=1

pV0 = p = const.
(isobaric)
systems with high T, high thermal energy
input
pV = ZnRT = const.
(isothermal)

1<n<k

quasi-adiabatic

n=k

adiabatic, reversible

n=

pV, V dominates -> V=const.


(isochoric)

adiabatic

no mass transfer
no heat transfer

(closed system)

k: specific heat ratio

Work through For Example problems


in sections 2.1 and 2.2,
and Example Problem 2.1
then

Work on homework problems


2.7, 2.14 (answer in SI),
2.23(a&b only use MATLAB for b)

Lecture 5

Heat and Cycles

Sections 2.4 & 2.6

Energy Transfer by Heat


The symbol Q denotes an amount of energy
transferred across the boundary of a system by heat.
Since engineering thermodynamics is often
concerned with steam engines, internal combustion
engines and turbines, and that heat will need to be
added to the system for it to function, convention is
that the heat transfer into a system as positive.
Q > 0: heat transfer to the system
Q < 0: heat transfer from the system

The same sign convention is used for the rate of


energy transfer by heat, given by Q .
Note: this is the opposite convention from work!

Modes of Heat Transfer


For any particular application, energy
transfer by heat can occur by one or more of
three modes:
conduction
radiation
convection

Conduction
Conduction is the transfer of energy
from more energetic particles of a
substance to less energetic adjacent
particles due to interactions between
them.
The time rate of energy transfer by
conduction is quantified by Fouriers
law.
An application of Fouriers law to a
plane wall at steady state is shown at
right.

Conduction
By Fouriers law, the rate of heat transfer across any
plane normal to the x direction, Q x, is proportional to the
wall area, A, and the temperature gradient in the x
direction, dT/dx,
dT

Qx kA
(Eq. 2.31)
dx
where
k is a proportionality constant, a property of the wall
material called the thermal conductivity.
The minus sign is a consequence of energy transfer in
the direction of decreasing temperature.

In this case, temperature varies linearly with x, and thus


dT T2 T1
T2 T1

( 0) and Eq. 2.31 gives Qx kA

dx
L
L

Thermal Radiation
Thermal radiation is energy transported by
electromagnetic waves (or photons). Unlike
conduction, thermal radiation requires no
intervening medium and can take place in a
vacuum.
The time rate of energy transfer by radiation is
quantified by expressions developed from the
Stefan-Boltzman law.

Thermal Radiation
An application involving net
radiation exchange between a
surface at temperature Tb and a
much larger surface at Ts (< Tb)
is shown at right.
Net energy is transferred in the direction of the arrow
and quantified by
(Eq. 2.33)
Q esA[T 4 T 4 ]
e

where
A is the area of the smaller surface,
e is a property of the surface called its emissivity,
s is the Stefan-Boltzman constant.

Convection
Convection is energy transfer between a solid
surface and an adjacent gas or liquid by the
combined effects of conduction and bulk flow
within the gas or liquid.
The rate of energy transfer by convection is
quantified by Newtons law of cooling.

Convection
An application involving
energy transfer by
convection from a transistor
to air passing over it is
shown at right.
Energy is transferred in the direction of the arrow and
quantified by

Q c hA[Tb Tf ]

(Eq. 2.34)

where
A is the area of the transistors surface and
h is an empirical parameter called the convection heat
transfer coefficient.

Work through For Example problem


in section 2.4.2, and
Example Problems 2.4, 2.5

then

Work on homework problem


2.51

Thermodynamic Cycles
A thermodynamic cycle is a sequence of
processes that begins and ends at the same
state.
Examples of thermodynamic cycles include
Power cycles that develop a net energy transfer by
work using an energy input by heat transfer from hot
combustion gases.
Refrigeration cycles that provide cooling for a
refrigerated space using an energy input by work in
the form of electricity.
Heat pump cycles that provide heating to a dwelling
using an energy input by work in the form of electricity.

Power Cycle
A system undergoing a power cycle is
shown at right.
The energy transfers by heat and work
shown on the figure are each positive in the
direction of the accompanying arrow. This
convention is commonly used for analysis
of thermodynamic cycles.
Wcycle is the net energy transfer by work from the system
per cycle of operation in the form of electricity, typically.
Qin is the heat transfer of energy to the system per cycle
from the hot body drawn from hot gases of combustion or
solar radiation, for instance.
Qout is the heat transfer of energy from the system per
cycle to the cold body discharged to the surrounding
atmosphere or nearby lake or river, for example.

Power Cycle
Applying the closed system energy balance to each
cycle of operation,

DEcycle = Qcycle Wcycle

(Eq. 2.39)

Since the system returns to its initial state after each


cycle, there is no net change in its energy: DEcycle = 0,
and the energy balance reduces to give

Wcycle = Qin Qout

(Eq. 2.41)

In words, the net energy transfer by work from the


system equals the net energy transfer by heat to the
system, each per cycle of operation.

Power Cycle
The performance of a system undergoing a power cycle
is evaluated on an energy basis in terms of the extent to
which the energy added by heat, Qin, is converted to a net
work output, Wcycle. This is represented by the ratio

Wcy cle
Qin

(power cycle)

(Eq. 2.42)

called the thermal efficiency.


Introducing Eq. 2.41, an alternative form is obtained

Qin Qout
Q
1 out
Qin
Qin

(power cycle)

(Eq. 2.43)

Power Cycle
Using the second law of thermodynamics (Chapter 5), we will
show that the value of thermal efficiency must be less than
unity: < 1 (< 100%). That is, only a portion of the energy
added by heat, Qin, can be obtained as work. The remainder,
Qout, is discharged.

Example: A system undergoes a power cycle while receiving


1000 kJ by heat transfer from hot combustion gases at a
temperature of 500 K and discharging 600 kJ by heat transfer
to the atmosphere at 300 K. Taking the combustion gases and
atmosphere as the hot and cold bodies, respectively, determine
for the cycle, the net work developed, in kJ, and the thermal
efficiency.
Substituting into Eq. 2.41, Wcycle = Qin Qout
Wcycle = 1000 kJ 600 kJ = 400 kJ.
Wcy cle

Then, with Eq. 2.42,


Qin
= 400 kJ/1000 kJ = 0.4 (40%).
Note the thermal efficiency is commonly reported on a
percent basis.

Refrigeration Cycle
A system undergoing a
refrigeration cycle is shown at right.
As before, the energy transfers
are each positive in the direction of
the accompanying arrow.
Wcycle is the net energy transfer by work to the system
per cycle of operation, usually in the form of electricity.
Qin is the heat transfer of energy to the system per
cycle from the cold body drawn from a freezer
compartment, for example.
Qout is the heat transfer of energy from the system
per cycle to the hot body discharged to the space
surrounding the refrigerator, for instance.

Refrigeration Cycle
Since the system returns to its
initial state after each cycle, there
is no net change in its energy:
DEcycle = 0, and the energy
balance reduces to give
Wcycle = Qout Qin

(Eq. 2.44)

In words, the net energy transfer by work to the


system equals the net energy transfer by heat
from the system, each per cycle of operation.

Refrigeration Cycle
The performance of a system undergoing a
refrigeration cycle is evaluated on an energy basis as
the ratio of energy drawn from the cold body, Qin, to
the net work required to accomplish this effect, Wcycle:

Qin
Wcy cle

(refrigeration cycle)

(Eq. 2.45)

called the coefficient of performance for the refrigeration


cycle.
Introducing Eq. 2.44, an alternative form is obtained
Qin

Qout Qin

(refrigeration cycle) (Eq. 2.46)

Heat Pump Cycle


The heat pump cycle analysis
closely parallels that given for the
refrigeration cycle. The same figure
applies:

But now the focus is on Qout,


which is the heat transfer of
energy from the system per
cycle to the hot body such as to
the living space of a dwelling.
Qin is the heat transfer of energy to the
system per cycle from the cold body drawn
from the surrounding atmosphere or the ground,
for example.

Heat Pump Cycle


As before, Wcycle is the net
energy transfer by work to
the system per cycle,
usually provided in the form
of electricity.

As for the refrigeration cycle, the energy balance


reads

Wcycle = Qout Qin

(Eq. 2.44)

Heat Pump Cycle


The performance of a system undergoing a heat
pump cycle is evaluated on an energy basis as the
ratio of energy provided to the hot body, Qout, to the
net work required to accomplish this effect, Wcycle:

Qout
Wcy cle

(heat pump cycle)

(Eq. 2.47)

called the coefficient of performance for the heat pump


cycle.
We can also write this as:
Qout

Qout Qin

(heat pump cycle) (Eq. 2.48)

Example: A system undergoes a heat pump cycle


while discharging 900 kJ by heat transfer to a dwelling
at 20oC and receiving 600 kJ by heat transfer from the
outside air at 5oC. Taking the dwelling and outside air
as the hot and cold bodies, respectively, determine for
the cycle, the net work input, in kJ, and the coefficient
of performance.
Substituting into Eq. 2.44, Wcycle = Qout Qin
Wcycle = 900 kJ 600 kJ = 300 kJ.
Then, with Eq. 2.47,

= 900 kJ/300 kJ = 3.0.

Qout
Wcy cle

Work through For Example problems


in section 2.6

then

Work on homework problems


2.71, 2.74

Lecture 7

Using Data Tables

Sections 3.4, 3.5, 3.6, 3.8, 3.8.1, 3.9, 3.10.1

Steam Tables
Tables of properties for different substances
are frequently set up in the same general
format. The tables for water, called the
steam tables, provide an example of this
format. The steam tables are in appendix
tables A-2 through A-5.
Table A-4 applies to water as a superheated
vapor.
Table A-5 applies to compressed liquid water.
Tables A-2 and A-3 apply to the two-phase,
liquid-vapor mixture of water.

Single-Phase Regions
Since pressure and temperature
are independent properties in
the single-phase liquid and
vapor regions, they can be used
to fix the state in these regions.
Tables A-4/A-4E (Superheated
Water Vapor) and A-5/A-5E
(Compressed Liquid Water)
provide several properties as
functions of pressure and
temperature, as considered
next.

Table A-5/A-5E

Table A-4/A-4E

Single-Phase Regions
Properties tabulated in Tables A-4 and A-5 include
Temperature (T)
Pressure (p)

Table A-5/A-5E

Specific volume (v)


Specific internal energy (u)
Specific enthalpy (h), which is a
sum of terms that often appears in
thermodynamic analysis:

Table A-4/A-4E

h = u + pv
Enthalpy is a property because it is defined in terms of
properties; physical significance is associated with it in Chapter 4.

Specific entropy (s), an intensive property developed in


Chapter 6

Example: Single-Phase Regions of Tables


What are the specific volume, specific enthalpy and specific
internal energy for superheated water vapor at 10 MPa and
400oC?

Linear Interpolation
When a state does not fall exactly on the grid of values provided
by property tables, linear interpolation between adjacent entries
is used.
Example: Specific volume (v) associated with superheated
water vapor at 10 bar and 215oC is found by linear interpolation
between adjacent entries in Table A-4.
(0.2275 0.2060) m3/kg (v 0.2060) m3/kg

slope =
=
(240 200)oC
(215 200)oC

v = 0.2141 m3/kg

Table A-4
T
C

v
m /kg
3

u
kJ/kg

h
kJ/kg

s
kJ/kgK

p = 10 bar = 1.0 MPa


(Tsat = 179.91oC)
Sat.
200
240

0.1944
0.2060
0.2275

2583.6
2621.9
2692.9

2778.1
2827.9
2920.4

6.5865
6.6940
6.8817

Two-Phase Liquid-Vapor Region


Tables A-2/A-2E
(Temperature Table) and A3/A-3E (Pressure Table)
provide
saturated liquid (f) data
saturated vapor (g) data
Table note: For saturated liquid specific volume, the table heading is vf103.
At 8oC, vf 103 = 1.002 vf = 1.002/103 = 1.002 103.
Table A-2
Temp
o
C

Press.
bar

.01
4
5
6
8

0.00611
0.00813
0.00872
0.00935
0.01072

Specific Volume

m /kg
3

Sat.
Liquid

Sat.
Vapor

vf103
1.0002
1.0001
1.0001
1.0001
1.0002

vg
206.136
157.232
147.120
137.734
120.917

Internal Energy
kJ/kg
Sat.
Sat.
Liquid
Vapor

Sat.
Liquid

Evap.

Sat.
Vapor

Entropy
kJ/kgK
Sat.
Sat.
Liquid
Vapor

uf
0.00
16.77
20.97
25.19
33.59

hf
0.01
16.78
20.98
25.20
33.60

hfg
2501.3
2491.9
2489.6
2487.2
2482.5

hg
2501.4
2508.7
2510.6
2512.4
2516.1

sf
0.0000
0.0610
0.0761
0.0912
0.1212

ug
2375.3
2380.9
2382.3
2383.6
2386.4

Enthalpy
kJ/kg

sg
9.1562
9.0514
9.0257
9.0003
8.9501

Temp
o
C
.01
4
5
6
8

Two-Phase Liquid-Vapor Region


The specific volume of a two-phase liquidvapor mixture can be determined by using
the saturation tables and quality, x.
The total volume of the mixture is the sum
of the volumes of the liquid and vapor
phases: V = V + V
liq

vap

Dividing by the total mass of the mixture, m, an average


specific volume for the mixture is:
V Vliq Vvap
v=

With Vliq = mliqvf , Vvap = mvapvg , mvap/m = x , and mliq/m = 1 x :

v = (1 x)vf + xvg = vf + x(vg vf) (Eq. 3.2)

V Vliq Vvap
v= =
+
m
m
m

Vliq = mliqvf , Vvap = mvapvg , mvap/m = x , mliq/m = 1 x

v = (1 x)vf + xvg = vf + x(vg vf)

Two-Phase Liquid-Vapor Region


Since pressure and temperature are NOT
independent properties in the two-phase liquidvapor region, they cannot be used to fix the state
in this region.
The property, quality (x), defined only in the twophase liquid-vapor region, and either temperature
or pressure can be used to fix the state in this
region.
v = (1 x)vf + xvg = vf + x(vg vf) (Eq. 3.2)
u = (1 x)uf + xug = uf + x(ug uf) (Eq. 3.6)
h = (1 x)hf + xhg = hf + x(hg hf) (Eq. 3.7)

Two-Phase Liquid-Vapor Region


Example: A system consists of a two-phase liquid-vapor
mixture of water at 6oC and a quality of 0.4. Determine the
specific volume, in m3/kg, of the mixture.
Solution: Apply Eq. 3.2, v = vf + x(vg vf)
Substituting values from Table 2: vf = 1.001103 m3/kg and
vg = 137.734 m3/kg:
v = 1.001103 m3/kg + 0.4(137.734 1.001103) m3/kg
v = 55.094 m3/kg
Table A-2
Temp
o
C

Press.
bar

.01
4
5
6
8

0.00611
0.00813
0.00872
0.00935
0.01072

Specific Volume

m /kg
3

Sat.
Liquid

Sat.
Vapor

vf103
1.0002
1.0001
1.0001
1.0001
1.0002

vg
206.136
157.232
147.120
137.734
120.917

Internal Energy
kJ/kg
Sat.
Sat.
Liquid
Vapor

Sat.
Liquid

uf
0.00
16.77
20.97
25.19
33.59

hf
0.01
16.78
20.98
25.20
33.60

ug
2375.3
2380.9
2382.3
2383.6
2386.4

Enthalpy
kJ/kg
Evap.

Sat.
Vapor

Entropy
kJ/kgK
Sat.
Sat.
Liquid
Vapor

hfg
2501.3
2491.9
2489.6
2487.2
2482.5

hg
2501.4
2508.7
2510.6
2512.4
2516.1

sf
0.0000
0.0610
0.0761
0.0912
0.1212

sg
9.1562
9.0514
9.0257
9.0003
8.9501

Temp
o
C
.01
4
5
6
8

Property Data Use in the


Closed System Energy Balance
Example: A piston-cylinder assembly contains 2 kg of
water vapor at 100oC and 1 bar. The water vapor is
compressed to a saturated vapor state where the
pressure is 2.5 bar. During compression, there is a heat
transfer of energy from the vapor to its surroundings
having a magnitude of 250 kJ. Neglecting changes in
kinetic energy and potential energy, determine the work,
in kJ, for the process of the water vapor.
T
State 1
T1 = 100oC
p1 = 1 bar

State 2

2 kg
of water
Q = 250 kJ

Saturated vapor
p2 = 2.5 bar

p2 = 2.5 bar
2
T1 = 100oC

p1 = 1 bar
1

Property Data Use in the


Closed System Energy Balance
Solution: An energy balance for the closed system is
0

KE + PE +U = Q W
where the kinetic and potential energy changes are neglected.

Thus

W = Q m(u2 u1)

State 1 is in the superheated vapor region and is fixed by


p1 = 1 bar and T1 = 100oC. From Table A-4, u1 = 2506.7 kJ/kg.
State 2 is saturated vapor at p2 = 2.5 bar. From Table A-3,
u2 = ug = 2537.2 kJ/kg.
W = 250 kJ (2 kg)(2537.2 2506.7) kJ/kg = 311 kJ
The negative sign indicates work is done on the system as
expected for a compression process.

A Small Primer on Partial Derivatives


A dependent variable can be a function of many independent
variables, for instance:

A( x, y, z ) = Bxy + Cx yz + Dyz + E
2

The partial derivative is the derivative of the function with respect


to one variable as if all the other variables were constants:

A
= By 2 (1) + Cyz (2 x) + 0 + 0 = By 2 + 2Cxyz
x
We can then label which variable(s) we held constant when we
were taking the partial derivative:

A

x y

partial derivative of A with respect to x,


holding y constant

Specific Heats
Three properties related to specific internal energy and specific
enthalpy having important applications are the specific heats cv
and cp and the specific heat ratio k.

u
cv =

T v

h
cp =

T p

(Eq. 3.8)

(Eq. 3.9)

k=

cp
cv

(Eq. 3.10)

In general, cv is a function of v and T (or p and T), and cp


depends on both p and T (or v and T).
Specific heat data are provided in Fig 3.9 and Tables A-19
through A-21.
Although cv and cp are referred to as specific heats, there is no
general relationship between them and the heat transfer term of
the energy balance denoted by Q.

Property Approximations for Liquids


Approximate values for v, u, and h at liquid states can be
obtained using saturated liquid data.
Since the values of v and u for liquids
change very little with pressure at a fixed
Saturated
liquid
temperature, Eqs. 3.11 and 3.12 can be used
to approximate their values.

v(T, p) vf(T)
u(T, p) uf(T)

(Eq. 3.11)
(Eq. 3.12)

An approximate value for h at liquid states can be obtained using


Eqs. 3.11 and 3.12 in the definition h = u + pv: h(T, p) uf(T) + pvf(T)
or alternatively

h(T, p) hf(T) + vf(T)[p psat(T)]

(Eq. 3.13)

where psat denotes the saturation pressure at the given temperature

When the underlined term in Eq. 3.13 is small

h(T, p) hf(T)

(Eq. 3.14)

Work through For Example problems


in sections 3.5, 3.6, ,
and Example Problems
3.1, 3.2, 3.3, 3.4
then

Work on homework problems


3.42, 3.54 (answer in SI)

Lecture 8

Using Data Tables

Sections 3.5 & 3.6

Example: What is the specific volume of water at a


state where p = 10 bar and T = 215oC?

liquid

vapor
Quality

Temperature

supercritical
fluid

isobaric lines
(constant pressure)

Specific Volume

saturated region
(vapor dome)

p = 10 bar and T = 215oC

liquid

vapor
Quality

Temperature

supercritical
fluid

isobaric lines
(constant pressure)

Specific Volume

saturated region
(vapor dome)

p = 10 bar and T = 215oC

p = 10 bar and T = 215oC

Linear Interpolation
(x2, y2)

(x3, y3)

(x1, y1)

(y2 - y1)

(y3 - y1)

(x2 - x1)

(x3 - x1)

(y2 - y1)

(y3 - y1)

(x2 - x1)

(x3 - x1)
y3 is our unknown spec. vol.

multiply both sides by (x3 - x1)

add y1 to both sides

p = 10 bar and T = 215oC

Make sure you understand all for example,


Sample and in-class problems to date, which
includes:
Which data table corresponds to which region of
a T-v diagram and/or which approximations hold
there.
What phase(s) exist in a region of the T-v
diagram.
How p changes in different regions of the T-v
diagram.

Lecture 9

Ideal Gas Law


Compressibility Factor

Sections 3.11, 3.12

Incompressibility of Liquids
Approximate values for v, u, and h at liquid states can be
obtained using saturated liquid data.
Since the values of v and u for liquids
change very little with pressure at a fixed
temperature, Eqs. 3.11 and 3.12 can be used
to approximate their values.

v(T, p) vf(T)
u(T, p) uf(T)
h(T, p) hf(T)

(Eq. 3.11)
(Eq. 3.12)
(Eq. 3.14)

Saturated
liquid

Generalized Compressibility Chart


The p-v -T relation for 10 common gases is
shown in the generalized compressibility chart.

Generalized Compressibility Chart


In this chart, the compressibility factor, Z, is plotted versus
the reduced pressure, pR, and reduced temperature TR,
pv
where
Z
pR = p/pc
TR = T/Tc
RT
(Eq. 3.23)
(Eq. 3.27)
(Eq. 3.28)

R is the universal gas constant


R

8.314 kJ/kmolK
1.986 Btu/lbmoloR
1545 ftlbf/lbmoloR

(Eq. 3.22)

The symbols pc and Tc denote the


temperature and pressure at the
critical point for the particular gas
under consideration. These values
are obtained from Tables A-1 and
A-1E.

Generalized Compressibility Chart


When p, pc, T, Tc, v , and R are used in consistent units, Z,
pR, and TR are numerical values without units.
Example: For air at 200 K, 132 bar, TR = 200 K/133 K = 1.5,
pR = 132 bar/37.7 bar = 3.5 where Tc and pc for air are from
Table A-1. With these TR, pR values, the generalized
compressibility chart gives Z = 0.8.
With v = Mv, an alternative
form of the compressibility
factor is
pv
Z=
(Eq. 3.24)
RT
where
R=

R
M

(Eq. 3.25)

Studying the Generalized Compressibility Chart


The solid lines labeled with TR values represent best fits to
experimental data. For the 10 different gases represented
there is little scatter in data about these lines.
At the lowest reduced
temperature value
shown, TR = 1.0, the
compressibility factor
varies between 0.2
and 1.0. Less
variation is observed
as TR takes higher
values.

Studying the Generalized Compressibility Chart


For each specified value of TR, the compressibility factor
approaches a value of 1.0 as pR approaches zero.

Studying the Generalized Compressibility Chart


Low values of pR, where Z 1, do not necessarily
correspond to a range of low absolute pressures.
For instance, if pR = 0.05, then p = 0.05pc. With pc values
from Table A-1
Water vapor
Ammonia
Carbon dioxide
Air

pc = 220.9 bar p = 11 bar


pc = 112.8 bar p = 5.6 bar
pc = 73.9 bar p = 3.7 bar
pc = 37.7 bar p = 1.9 bar

These pressure values range from 1.9 to 11 bar, which in


engineering practice are not normally considered as low
pressures.

Introducing the Ideal Gas Model


To recap, the generalized
compressibility chart shows that
at states where the pressure p
is small relative to the critical
pressure pc (where pR is small),
the compressibility factor Z is
approximately 1.
At such states, it can be assumed with reasonable
accuracy that Z = 1. Then

pv = RT

(Eq. 3.32)

Introducing the Ideal Gas Model


Three alternative forms of Eq. 3.32 can be derived
as follows:
With v = V/m, Eq. 3.32 gives

pV = mRT

(Eq. 3.33)

With v = v/M and R = R/M, Eq. 3.32 gives

pv R T

(Eq. 3.34)

Finally, with v = V/n, Eq. 3.34 gives

pV = nRT

(Eq. 3.35)

Introducing the Ideal Gas Model


While the ideal gas model does not provide an
acceptable approximation generally, in the
limiting case considered in the discussion of the
compressibility chart, it is justified for use, and
indeed commonly applied in engineering
thermodynamics at such states.
Appropriateness of the ideal gas model can be
checked by locating states under consideration
on one of the generalized compressibility charts
provided by appendix figures Figs. A-1 through
A-3.

Work Example Problems


3.7, 3.8

then

Work on homework problems


3.103, 3.113

Lecture 11

Control Volumes
Mass Rate Balance

Sections 4.1 4.3

Mass Rate Balance

time rate of change of


mass contained within the
control volume at time t

time rate of flow of


mass in across
inlet i at time t

dmcv
i m
e
m
dt

time rate of flow


of mass out across
exit e at time t

(Eq. 4.1)

Mass Rate Balance


In practice there may be several locations on the
boundary through which mass enters or exits.
Multiple inlets and exits are accounted for by
summing over all entrances and exits:

dmcv
m i m e
dt
i
e

(Eq. 4.2)

Eq. 4.2 is the mass rate balance for control


volumes with several inlets and exits.

Example: (Problem 4.2)


Liquid propane enters and initially empty cylindrical storage tank at a mass flow rate
of 10 kg/s. Flow continues until the tank is filled with propane at 20oC, 9 bar. The
tank is 25 m long and has a 4 m diameter. Determine the time, in minutes, to fill the
tank.

Mass Flow Rate


(One-Dimensional Flow)
Flow is normal to the boundary at locations where mass
enters or exits the control volume.
All intensive properties are uniform with position over each
inlet or exit area (A) through which matter flows.
V is velocity

V is volume

dm d ( V ) d ( Ax)
dx

A AV
dt
dt
dt
dt

AV
m
v

volume flow rate

(Eq. 4.4b)

where

V is velocity
v is specific volume

Example: (Problem 4.4)


Data are provided for the crude oil
storage tank. Determine:
a) The mass of oil in the tank, in kg,
after 24 hours, and
b) The volume of oil in the tank, in
m3, at 24 hours

Mass Rate Balance


(Steady-State Form)

Steady-state: all properties are unchanging


in time.
For steady-state control volume, dmcv/dt = 0.

dmcv
m i m e
dt
i
e

m i m e
i

(mass rate in)

(mass rate out)

(Eq. 4.6)

Work through
For Example problems in section 4.2
Example Problems 4.1 and 4.2

then

Work on homework problems


4.9, 4.16

Lecture 12

Energy Rate Balance

Sections 4.4 4.5

Mass Rate Balance

time rate of change of


mass contained within the
control volume at time t

time rate of flow of


mass in across
inlet i at time t

dmcv
= m i m e
dt

time rate of flow


of mass out across
exit e at time t

(Eq. 4.1)

Mass Rate Balance


In practice there may be several locations on the
boundary through which mass enters or exits.
Multiple inlets and exits are accounted for by
summing over all entrances and exits:

dmcv
= m i m e
dt
i
e

(Eq. 4.2)

Eq. 4.2 is the mass rate balance for control


volumes with several inlets and exits.

Energy Rate Balance

time rate of change


of the energy
contained within
the control volume
at time t

net rate at which


energy is being
transferred in
by heat transfer
at time t

net rate at which


energy is being
transferred out
by work at
time t

net rate of energy


transfer into the
control volume
accompanying
mass flow

2
2
V
dEcv
V
= Q W + m i (ui + i + gzi ) m e (u e + e + gz e ) (Eq. 4.9)
2
dt
2
control volume

interaction with surroundings

inlets and exits

Evaluating Work for a Control Volume


The expression for work is

W = Wcv + m e ( peve ) m i ( pi vi )

(Eq. 4.12)

where
cv accounts for work associated with rotating
W
shafts, displacement of the boundary, and electrical
effects.
m
e ( pe ve ) is the flow work at exit e.

m i ( pi vi ) is the flow work at inlet i.

Control Volume Energy Rate Balance


(One-Dimensional Flow Form)
Using Eq. 4.12 in Eq. 4.9
control volume
inlets and exits
interaction with surroundings with closed system defs.

2
2
dEcv
V
V
= Qcv Wcv + m i (ui + pi vi + i + gzi ) m e (ue + peve + e + gze )
dt
2
2

(Eq. 4.13)
For convenience substitute enthalpy, h = u + pv

2
2
dEcv
V
V
= Qcv Wcv + m i (hi + i + gzi ) m e (he + e + gze )
dt
2
2

(Eq. 4.14)

Control Volume Energy Rate Balance


(One-Dimensional Flow Form)
In practice there may be several locations
on the boundary through which mass enters
or exits. Multiple inlets and exits are
accounted for by introducing summations:
2
2
dEcv
V
V
= Qcv Wcv + m i (hi + i + gzi ) m e (he + e + gze )
dt
2
2
i
e

(Eq. 4.15)

Eq. 4.15 is the accounting balance for the


energy of the control volume.

Control Volume Energy Rate Balance


(Steady-State Form)
Steady-state: all properties are unchanging
in time.
For steady-state control volume, dEcv/dt = 0.
2
2
V
V
0 = Q cv Wcv + m i (hi + i + gzi ) m e (he + e + gze )
2
2
e
i

(Eq. 4.18)

Control Volume Energy Rate Balance


(Steady-State Form, One-Inlet, One-Exit)
Many important applications involve one-inlet,
one-exit control volumes at steady state.
The mass rate balance reduces to m1 = m 2 = m .
2
2

Eq.
(V
V
)

1
2
0 = Q cv Wcv + m (h1 h2 ) +
+ g ( z1 z2 )
2

4.20a

or dividing by mass flow rate


(V12 V22 )
Q cv Wcv
0=

+ (h1 h2 ) +
+ g ( z1 z2 )
m
m
2

Eq.
4.20b

Example: (Problem 4.26)


Air enters a horizontal, constant-diameter heating duct operating at steady state at
290 K, 1 bar, with a volumetric flow rate of 0.25 m3/s, and exits at 325 K, 0.95 bar.
The flow area is 0.04 m2. Assuming the ideal gas model with k = 1.4 for the air,
determine:
a) The mass flow rate, in kg/s, b) the velocity at the inlet and exit, and
c) the rate of heat transfer in kW.

Example: (Problem 4.28)


At steady state, air at 200 kPa, 52oC, and a mass flow rate of 0.5 kg/s enters a
horizontal insulated duct having differing inlet and exit cross-sectional areas. At the
duct exit, the pressure of the air is 100 kPa, the velocity is 255 m/s, and the crosssectional area is 2 x 10-3 m2. Assuming the ideal gas model, determine: a) the
temperature of the air at the exit, in C,
b) The velocity of the air at the inlet in m/s, and c) the inlet cross-sectional area, in m2.

Work on homework problems


4.23, 4.24

Lecture 13

Nozzles, Diffusers,

Turbines, Compressors,
and Pumps
Sections 4.6 4.8

Nozzles and Diffusers

1
1
2
2
P1 V1 gh1 P 2 V2 gh 2
2
2

Bernoullis equation

Nozzle: a flow passage of varying crosssectional area in which the velocity of a gas
or liquid increases in the direction of flow.
Diffuser: a flow passage of varying crosssectional area in which the velocity of a gas
or liquid decreases in the direction of flow.

Nozzle and Diffuser Modeling


2
2
Eq.

(V

V
)
1
2
g ( z1 z2 )
0 Q cv Wcv m (h1 h2 )
2
4.20a

Wcv 0.

If the change in potential energy from inlet to exit is


negligible, g(z1 z2) drops out.
If the heat transfer with surroundings is negligible,
Q cv drops out.
V2 V2
2
0 (h1 h2 ) 1

(Eq. 4.21)

Example: (Problem 4.31)


Steam enters a nozzle operating at steady state at 20 bar, 280oC, with a velocity of 80
m/s. The exit pressure and temperature are 7 bar and 180oC, respectively. The
mass flow rate is 1.5 kg/s. Neglecting heat transfer and potential energy, determine:
a) The exit velocity in m/s, and
b) the inlet and exit flow areas, in cm2.

Turbines

Turbine: a device in which power is


developed as a result of a gas or liquid
passing through a set of blades attached to
a shaft free to rotate.

Turbine Modeling
2
2
Eq.

(V

V
)
2 g(z z )
0 Q cv Wcv m (h1 h2 ) 1
1
2 4.20a
2

If the change in kinetic energy of flowing matter is


negligible, (V12 V22) drops out.
If the change in potential energy of flowing matter is
negligible, g(z1 z2) drops out.
If the heat transfer with surroundings is negligible,
Q cv drops out.

Wcv m (h1 h2 )

Compressors and Pumps


Compressors and Pumps:
devices in which work is done
on the substance flowing
through them to change the
state of the substance, typically
to increase the pressure and/or
elevation.
Compressor : substance is gas
Pump: substance is liquid

Compressor and Pump Modeling


2
2
Eq.

(V

V
)
2 g(z z )
0 Q cv Wcv m (h1 h2 ) 1
1
2 4.20a
2

If the change in kinetic energy of flowing matter is


negligible, (V12 V22) drops out.
If the change in potential energy of flowing matter is
negligible, g(z1 z2) drops out.
If the heat transfer with surroundings is negligible,
Q cv drops out.

Wcv m (h1 h2 )

Example: (Problem 4.41)


Steam enters a well-insulated turbine operating at steady state at 4 MPa with a specific
enthalpy of 3015.4 kJ/kg and a velocity of 10 m/s. The steam expands to the turbine exit
where the pressure is 0.07 MPa, specific enthalpy is 2431.7 kJ/kg, and the velocity is 90
m/s. The mass flow rate is 11.95 kg/s. Neglecting potential energy effects, determine
the power developed by the turbine, in kW.

Work through Example Problems


4.3, 4.4, 4.5 and 4.6
then

Work on homework problems


4.33, 4.43, 4.53

Lecture 14

Heat Exchangers and


Throttling Devices
Sections 4.9 and 4.10

Heat Exchangers

Direct contact: A mixing chamber in which hot and


cold streams are mixed directly.
Tube-within-a-tube counterflow: A gas or liquid
stream is separated from another gas or liquid by a
wall through which energy is conducted. Heat
transfer occurs from the hot stream to the cold
stream as the streams flow in opposite directions.

Heat Exchanger Modeling


2
2
V
V
0 = Q cv Wcv + m i (hi + i + gzi ) m e (he + e + gze )
2
2
e
i

(Eq. 4.18)
Wcv = 0.
If the kinetic energies of the flowing streams are
negligible, m i(Vi2/2) and m e (Ve2/2) drop out.
If the potential energies of the flowing streams are
negligible, m igzi and m e gze drop out.
If the heat transfer with surroundings is negligible,
Q cv drops out.

0 = m i hi m e he
i

m i = m e
i

for each connected flow volume

Problem 4.76 Steam enters a heat exchanger operating at steady state at 250 kPa and
a quality of 90% and exits as a saturated liquid at the same pressure. A separate stream
of oil with a mass flow rate of 29 kg/s enters at 20oC, and exits at 100oC with no
significant change in pressure. The specific heat of the oil is c = 2.0 kJ/kgK. Kinetic and
potential energy effects are negligible. If heat transfer from the heat exchanger to its
surroundings is 10% of the energy required to increase the temperature of the oil,
determine the steam mass flow rate, in kg/s.

Pause the video

Work through examples 4.7 and 4.8

then

Do homework problem 4.74

Throttling Devices

Throttling Device: a device that achieves


a significant reduction in pressure by
introducing a restriction into a line through
which a gas or liquid flows. Means to
introduce the restriction include a partially
opened valve or a porous plug.

Throttling Device Modeling


2
2

Eq.
(V
V
)

2 + g(z z )
0 = Q cv Wcv + m (h1 h2 ) + 1
1
2 4.20a
2

Wcv = 0.
If the change in kinetic energy of flowing matter
upstream and downstream of the restriction is
negligible, (V12 V22) drops out.
If the change in potential energy of flowing matter is
negligible, g(z1 z2) drops out.
If the heat transfer with surroundings is negligible,
Q cv drops out.
p <p
(Eq. 4.22)

h2 = h1

Problem 4.89 Ammonia enters the expansion valve of a refrigeration system at a


pressure of 10 bar and a temperature of 24oC and exits at 1 bar. If the refrigerant
undergoes a throttling process, what is the quality of the refrigerant exiting the
expansion valve?

Pause the video

Work through example 4.9

then

Try suggested problem 4.90


(well work on this in class)

Lecture 15

Second Law of Thermodynamics,


Entropy,
Reversible and Irreversible Processes,
and
the Kelvin-Plank Statement
Sections 5.1-5.4

spontaneous heat transfer

spontaneous expansion

falling mass
-> observed direction of these
processes

Aspects of the
Second Law of Thermodynamics
From conservation of mass and energy
principles, mass and energy cannot be
created or destroyed.
For a process, conservation of mass and
energy principles indicate the disposition of
mass and energy but do not infer whether the
process can actually occur.
The second law of thermodynamics
provides the guiding principle for whether a
process can occur.

Aspects of the
Second Law of Thermodynamics
The second law of thermodynamics has many
aspects, which at first may appear different in kind
from those of conservation of mass and energy
principles. Among these aspects are:
predicting the direction of processes.
establishing conditions for equilibrium.
determining the best theoretical performance of
cycles, engines, and other devices.
evaluating quantitatively the factors that prevent
achieving the best theoretical performance
level.

Aspects of the
Second Law of Thermodynamics
Other aspects of the second law include:
defining a temperature scale independent of the
properties of any thermometric substance.
developing means for evaluating properties
such as u and h in terms of properties that are
more readily obtained experimentally.
Scientists and engineers have found additional uses
of the second law and deductions from it. It also
has been used in philosophy, economics, and other
disciplines far removed from engineering
thermodynamics.

Second Law of Thermodynamics


Alternative Statements
There is no simple statement that captures all
aspects of the second law. Several
alternative formulations of the second law are
found in the technical literature. Three
prominent ones are:
Clausius Statement
Kelvin-Planck Statement
Entropy Statement

Second Law of Thermodynamics


Alternative Statements
The focus of Chapter 5 is on the Clausius and
Kelvin-Planck statements.
The Entropy statement is developed and applied
in Chapter 6.
Like every physical law, the basis of the second
law of thermodynamics is experimental
evidence. While the three forms given are not
directly demonstrable in the laboratory,
deductions from them can be verified
experimentally, and this infers the validity of the
second law statements.

Clausius Statement
of the Second Law
It is impossible for any system to operate in such
a way that the sole result would be an energy
transfer by heat from a cooler to a hotter body.

Thermal Reservoir
A thermal reservoir is a system that
always remains at constant temperature
even though energy is added or removed
by heat transfer.
Such a system is approximated by the
earths atmosphere, lakes and oceans, and
a large block of a solid such as copper.

Kelvin-Planck Statement
of the Second Law
It is impossible for any system to operate in a
thermodynamic cycle and deliver a net amount
of energy by work to its surroundings while
receiving energy by heat transfer from a single
thermal reservoir.

Entropy Statement
of the Second Law
Mass and energy are familiar examples of
extensive properties used in thermodynamics.
Entropy is another important extensive property.
How entropy is evaluated and applied is detailed
in Chapter 6.
Unlike mass and energy, which are conserved,
entropy is produced within systems whenever
non-idealities such as friction are present.
The Entropy Statement is:
It is impossible for any system to operate in
a way that entropy is destroyed.

Irreversibilities
One of the important uses of the second law of
thermodynamics in engineering is to determine
the best theoretical performance of systems.
By comparing actual performance with best
theoretical performance, insights often can be had
about the potential for improved performance.
Best theoretical performance is evaluated in terms
of idealized processes.
Actual processes are distinguishable from such
idealized processes by the presence of nonidealities called irreversibilities.

Irreversibilities Commonly
Encountered in Engineering Practice
Heat transfer through a finite temperature
difference
Unrestrained expansion of a gas or liquid to
a lower pressure
Spontaneous chemical reaction
Spontaneous mixing of matter at different
compositions or states
Friction sliding friction as well as friction in
the flow of fluids

Irreversibilities Commonly
Encountered in Engineering Practice
Electric current flow through a resistance
Magnetization or polarization with hysteresis
Inelastic deformation
All actual processes involve effects such as
those listed, including naturally occurring
processes and ones involving devices we
construct from the simplest mechanisms to
the largest industrial plants.

Irreversible and Reversible


Processes
During a process of a system,
irreversibilities may be present:
within the system, or
within its surroundings (usually the
immediate surroundings), or
within both the system and its
surroundings.

Irreversible and Reversible


Processes
A process is irreversible when
irreversibilities are present within the system
and/or its surroundings.
All actual processes are irreversible.
A process is reversible when no
irreversibilities are present within the system
and its surroundings.
This type of process is fully idealized.

Irreversible and Reversible


Processes
A process is internally reversible when no
irreversibilities are present within the
system. Irreversibilities may be present
within the surroundings, however.
An internally reversible process is a
quasiequilibrium process (see Sec.
2.2.5).

Example: Internally Reversible Process


Water contained within a piston-cylinder evaporates
from saturated liquid to saturated vapor at 100oC. As
the water evaporates, it passes through a sequence of
equilibrium states while there is heat transfer to the
water from hot gases at 500oC.
For a system enclosing the water there are no
internal irreversibilities, but
Such spontaneous
heat transfer is an
irreversibility in its
surroundings: an
external irreversibility.

Analytical Form of the


Kelvin-Planck Statement
For any system undergoing a
thermodynamic cycle while
exchanging energy by heat
transfer with a single thermal
reservoir, the net work, Wcycle,
can be only negative or zero
never positive:
Wcycle 0

< 0: Internal irreversibilities present


= 0: No internal irreversibilities

(Eq. 5.3)

NO!

single
reservoir

Lecture 16

The Second Law Applied to


Cycles (with Two Reservoirs)
Some Groundwork

Sections 5.5 5.7

Ideas Weve Developed About the Second Law


1. There is a tendency in all processes (and cyclic devices) to
produce unrecoverable energy (irreversibilities) in either doing
or consuming work.
2. To fully restore systems with irrevesibilities to their original
state after a process has taken place requires expenditure of
some work external to the system.
3. Machines that utilize all of the energy from a single
reservoir and convert it to work (even though the First Law is
satisfied) i.e. perpetual motion machines are not possible.
4. Heat will only flow from a high to low temperature reservoir
without the aid of externally supplied work (e.g heat pump).
A process is considered reversible when there are no
irreversibilities within the system as it undergoes the process.

Thermal Efficiency and Coefficient of Performance

,
(refrigeration, heat pump)

(power cycle)

Applications to Power Cycles Interacting


with Two Thermal Reservoirs
For a system undergoing a power cycle while
communicating thermally with two thermal
reservoirs, a hot reservoir and a cold reservoir,
the thermal efficiency of any such cycle is

Wcycle
QH

QC
(Eq. 5.4)
= 1
QH (Eq. 2.43)

Applications to Power Cycles Interacting


with Two Thermal Reservoirs
By applying the Kelvin-Planck statement of the second law,
Eq. 5.3, three conclusions can be drawn:
1. The value of the thermal efficiency must be less than
100%. Only a portion of the heat transfer QH can be obtained
as work and the remainder QC is discharged by heat transfer
to the cold reservoir.

Wcycle
QH

QC
= 1
QH

Two other conclusions, called the Carnot corollaries, are:

Carnot Corollaries
2. The thermal efficiency of an irreversible power
cycle is always less than the thermal efficiency of a
reversible power cycle when each operates between
the same two thermal reservoirs.
In other words, the reversible cycle between two reservoirs
tells us what the maximum efficiency of the cycle is. If there
are reversibilities, the efficiency can only be smaller than the
max.

Carnot Corollaries
3. All reversible power cycles operating between the
same two thermal reservoirs have the same thermal
efficiency.
In other words, once you have found one way to calculate the
efficiency for a reversible cycle, you know what the max
efficiency is.

A cycle is considered reversible when there are no


irreversibilities within the system as it undergoes the
cycle and heat transfers between the system and
reservoirs occur reversibly.

Applications to Power Cycles Interacting


with Two Thermal Reservoirs

Wcycle

QC
= 1
=
QH
QH

I < R ( any ) < 1


I: cycle with irreversibilities, R: reversible cycle

Applications to Refrigeration and Heat Pump


Cycles Interacting with Two Thermal Reservoirs
For a system undergoing a refrigeration cycle or heat
pump cycle while communicating thermally with two
thermal reservoirs, a hot reservoir and a cold
reservoir,
the coefficient of performance for the
refrigeration cycle is

QC
QC
(Eq. 5.5)
=
=
Wcycle QH QC

(Eq. 2.45)

and for the heat pump cycle is

QH
QH
=
=
(Eq. 5.6)
Wcycle QH QC

(Eq. 2.47)

Applications to Refrigeration and Heat Pump Cycles


Interacting with Two Thermal Reservoirs
By applying the Kelvin-Planck statement of the
second law, Eq. 5.3, three conclusions can be drawn:
1. For a refrigeration effect to occur a net work input
Wcycle is required. Accordingly, the coefficient of
performance must be finite in value.

QC
QC
=
=
Wcycle QH QC
Two other conclusions are:

Q
Q
H
H
=

W
Q

Q
cycle
H
C

Applications to Refrigeration and Heat Pump Cycles


Interacting with Two Thermal Reservoirs
2. The coefficient of performance of an irreversible
refrigeration cycle is always less than the coefficient
of performance of a reversible refrigeration cycle
when each operates between the same two thermal
reservoirs.
In other words, the reversible cycle between two reservoirs
tells us what the maximum coefficient of performance (COP)
of the cycle is. If there are reversibilities, the COP can only be
smaller than the max.

Applications to Refrigeration and Heat Pump Cycles


Interacting with Two Thermal Reservoirs
3. All reversible refrigeration cycles operating
between the same two thermal reservoirs have the
same coefficient of performance.
In other words, once you have found one way to calculate the
coefficient of performance for a reversible cycle, you know what
the max COP is.

All three conclusions also apply to a system


undergoing a heat pump cycle between hot and cold
reservoirs.

Applications to Power Cycles Interacting


with Two Thermal Reservoirs
QC
QC
=
=
Wcycle QH QC
QH
QH
=
=
Wcycle QH QC

COP I < COP R (any ) <


I: cycle with irreversibilities, R: reversible cycle

Lecture 17

The Kelvin Temperature Scale


and Theoretical Limits
of Performance of
Cycles Between Two Reservoirs

Sections 5.8 5.9

Kelvin Temperature Scale


Consider systems undergoing a power cycle and a
refrigeration or heat pump cycle, each while
exchanging energy by heat transfer with hot and cold
reservoirs:

The Kelvin temperature is defined so that


QC

QH

TC

=
TH
rev
cycle

(Eq. 5.7)

Kelvin Temperature Scale


In words, Eq. 5.7 states: When cycles are reversible,
and only then, the ratio of the heat transfers equals a
ratio of temperatures on the Kelvin scale, where TH is
the temperature of the hot reservoir and TC is the
temperature of the hot reservoir.
As temperatures on the Rankine scale differ from
Kelvin temperatures only by the factor 1.8:
T(oR)=1.8T(K), (i.e. it is also an absolute temperature
scale), Rankine would also be an acceptable
temperature scale to use in Eq. 5.7.

Maximum Performance Measures for Cycles


Operating between Two Thermal Reservoirs
Previous deductions from the Kelvin-Planck statement of the second law
include:
1. The thermal efficiency of an irreversible power
cycle is always less than the thermal efficiency of a
reversible power cycle when each operates between
the same two thermal reservoirs.
2. The coefficient of performance of an irreversible
refrigeration cycle is always less than the
coefficient of performance of a reversible
refrigeration cycle when each operates between the
same two thermal reservoirs.
3. The coefficient of performance of an irreversible
heat pump cycle is always less than the coefficient
of performance of a reversible heat pump cycle
when each operates between the same two thermal
reservoirs.

The reversible
process
determines
the maximum
theoretical
performance
for a cycle.

Maximum Performance Measures for Cycles


Operating between Two Thermal Reservoirs
Using Eq. 5.7 in Eqs. 5.4, 5.5, and 5.6, we get respectively:
QC

QH

= C
TH
rev
cycle

Power Cycle:

max

TC
= 1
TH

(Eq. 5.9)

Refrigeration Cycle:

TC
max =
TH TC

(Eq. 5.10)

Heat Pump Cycle:

TH
max =
TH TC

(Eq. 5.11)

where TH and TC must be on the Kelvin or Rankine scale.

Example: Power Cycle Analysis


A system undergoes a power cycle while
receiving 1000 kJ by heat transfer from a
thermal reservoir at a temperature of 500 K
and discharging 600 kJ by heat transfer to a
thermal reservoir at (a) 200 K, (b) 300 K, (c)
400 K. For each case, determine whether
the cycle operates irreversibly, operates
reversibly, or is impossible.

Hot Reservoir

TH = 500 K
QH = 1000 kJ
Wcycle

Power
Cycle

QC = 600 kJ

TC = (a) 200 K,
(b) 300 K,
(c) 400 K
Cold Reservoir

Solution: To determine the nature of the cycle, compare


actual cycle performance () to maximum theoretical cycle
performance (max) calculated from Eq. 5.9

Example: Power Cycle Analysis


Actual Performance: Calculate using the heat
transfers:
Q
600 kJ
= 1

QH

= 1

1000 kJ

= 0.4

Maximum Theoretical Performance: Calculate


max from Eq. 5.9 and compare to :

max

TC
200 K
= 1
= 0.6
TH
500 K

0.4 < 0.6

Irreversibly

(b) max

TC
300 K
= 1
= 1
= 0.4
500 K
TH

0.4 = 0.4

Reversibly

(c) max

TC
400 K
= 1
= 1
= 0.2
TH
500 K

0.4 > 0.2

Impossible

(a) max = 1

Problem 5.59
At steady state, a refrigeration cycle operating between hot and cold reservoirs at 300 K
and 275 K, respectively, removes energy by heat transfer from the cold reservoir at a
rate of 600 kW.
a) If the cycles coefficient of performance is 4, determine the power input required, in
kW.
b) Determine the minimum theoretical power require, in kW, for any such cycle.

Work through examples 5.1 5.3

then

Work on homework
problems 5.17, 5.19, 5.29, 5.45, 5.65

Lecture 18

Carnot Cycle
Clausius Inequality

Introducing Entropy

Sections 5.10, 5.11, 6.1, 6.2

Carnot Cycle
The Carnot cycle provides a specific
example of a reversible cycle that operates
between two thermal reservoirs. Other
examples are provided in Chapter 9: the
Ericsson and Stirling cycles.
In a Carnot cycle, the system executing the
cycle undergoes a series of four internally
reversible processes: two adiabatic
processes alternated with two isothermal
processes.

Carnot Power Cycles


The p-v diagram and schematic of a gas in a piston-cylinder
assembly executing a Carnot cycle are shown below:

Carnot Power Cycles


The p-v diagram and schematic of water executing a Carnot
cycle through four interconnected components are shown
below:

In each of these cases the thermal efficiency is given by

TC
max = 1
TH

(Eq. 5.9)

Carnot Refrigeration and Heat Pump Cycles


If a Carnot power cycle is operated in the opposite
direction, the magnitudes of all energy transfers
remain the same but the energy transfers are
oppositely directed.
Such a cycle may be regarded as a Carnot
refrigeration or heat pump cycle for which the
coefficient of performance is given, respectively, by
Carnot Refrigeration Cycle:

Carnot Heat Pump Cycle:

max

TC
=
TH TC

TH
max =
TH TC

(Eq. 5.10)

(Eq. 5.11)

Clausius Inequality
The Clausius inequality is developed from
the Kelvin-Planck statement of the second
law and can be expressed as:
Q

= cycle
T b

(Eq. 5.13)

where
indicates integral is to be performed over all parts of the
boundary and over the entire cycle.

b subscript indicates integrand is evaluated at the boundary


of the system executing the cycle.

Clausius Inequality
The Clausius inequality is developed from
the Kelvin-Planck statement of the second
law and can be expressed as:
Q

= cycle
T b

(Eq. 5.13)

The nature of the cycle executed is indicated by the value


of cycle:

cycle = 0 no irreversibilities present within the system


cycle > 0 irreversibilities present within the system
cycle < 0 impossible

Eq.
5.14

Example: Use of Clausius Inequality


A system undergoes a cycle while receiving
1000 kJ by heat transfer at a temperature of 500 K
and discharging 600 kJ by heat transfer at (a) 200
K, (b) 300 K, (c) 400 K. Using Eqs. 5.13 and 5.14,
what is the nature of the cycle in each of these
cases?
Solution: To determine the nature of the cycle,
perform the cyclic integral of Eq. 5.13 to each case
and apply Eq. 5.14 to draw a conclusion about the
nature of each cycle.

Example: Use of Clausius Inequality


Applying Eq. 5.13 to each cycle:
(a) cycle =

1000 kJ 600 kJ

= 1 kJ/K
500 K 200 K

Qin Qout
Q
= cycle
=

TC
T b TH
cycle = +1 kJ/K > 0

Irreversibilities present within system


(b) cycle =

1000 kJ 600 kJ

= 0 kJ/K
500 K 300 K

cycle = 0 kJ/K = 0

No irreversibilities present within system


(c) cycle

1000 kJ 600 kJ
=

= 0.5 kJ/K
500 K 400 K

cycle = 0.5 kJ/K < 0

Impossible

Clausius Inequality
The Clausius inequality considered next
provides the basis for developing the
entropy concept in Chapter 6.
The Clausius inequality is applicable to any
cycle without regard for the body, or bodies,
from which the system undergoing a cycle
receives energy by heat transfer or to
which the system rejects energy by heat
transfer. Such bodies need not be thermal
reservoirs.

Work through example after Eq 5.14

then

Work on homework
problems 5.85, 5.87

Introducing Entropy Change and the


Entropy Balance
The entropy change and entropy balance
concepts are developed using the Clausius
inequality expressed as:
Q

= cycle
T b

(Eq. 5.13)

where

cycle = 0 no irreversibilities present within the system


cycle > 0 irreversibilities present within the system
cycle < 0 impossible

Eq.
5.14

Defining Entropy Change


Consider two cycles, each composed
of two internally reversible processes,
process A plus process C and
process B plus process C, as shown
in the figure.
Applying Eq. 5.13 to these cycles gives,

where cycle is zero because the cycles are


composed of internally reversible processes.

Defining Entropy Change


Subtracting these equations:
Since A and B are arbitrary internally reversible
processes linking states 1 and 2, it follows that the
value of the integral is independent of the
particular internally reversible process and
depends on the end states only.

Defining Entropy Change


Recalling (from Sec. 1.3.3) that a quantity is a property if,
and only if, its change in value between two states is
independent of the process linking the two states, we
conclude that the integral represents the change in some
property of the system.
We call this property entropy and represent it by S. The
change in entropy is

(Eq. 6.2a)
where the subscript int rev signals that the integral is
carried out for any internally reversible process linking
states 1 and 2.

Defining Entropy Change


Equation 6.2a allows the change in entropy between
two states to be determined by thinking of an internally
reversible process between the two states. But since
entropy is a property, that value of entropy change
applies to any process between the states internally
reversible or not.
Entropy change is introduced by the integral of Eq. 6.2a
for which no accompanying physical picture is given.
Still, the aim of Chapter 6 is to demonstrate that entropy
not only has physical significance but also is essential for
thermodynamic analysis.

Entropy Facts
Entropy is an extensive property.
Just as mass and energy are accounted for by mass and
energy balances, entropy is accounted for by an entropy
balance.
Like mass and energy, entropy can be transferred across
the system boundary.
Like any other extensive property, the change in entropy
can be positive, negative, or zero:

By inspection of Eq. 6.2a, units for entropy S are kJ/K and


Btu/oR.
Units for specific entropy s are kJ/kgK and Btu/lb oR.

Entropy Facts
For problem solving, specific entropy values are provided in
Tables A-2 through A-18. Values for specific entropy are
obtained from these tables using the same procedures as
for specific volume, internal energy, and enthalpy, including
use of

(Eq. 6.4)
for two-phase liquid-vapor mixtures, and

(Eq. 6.5)
for liquid water, each of which is similar in form to
expressions introduced in Chap. 3 for evaluating v, u, and h.

Entropy Facts
For problem solving, states often are shown on
property diagrams having specific entropy as a
coordinate: the temperature-entropy and
enthalpy-entropy (Mollier) diagrams shown here

Expanded View of Enthalpy-Entropy Diagram

Work through examples in Sections


6.2.1, 6.2.2, 6.2.3, 6.2.5

Lecture 19

Calculating Entropy, and


Entropy Change in
Incompressible Substances,
Ideal Gases
Sections 6.3 6.5

Last Time:
Mathematical Form of the Second Law
The change in entropy is

where the subscript int rev signals that the integral is


carried out for any internally reversible process linking
states 1 and 2.
That the value of the integral is independent of the particular
internally reversible process and depends on the end states
only -> entropy is a property of the state of the system.

Entropy and Heat Transfer


By inspection of Eq. 6.2a, the defining equation for
entropy change on a differential basis is

(Eq. 6.2b)
Equation 6.2b indicates that when a closed system
undergoing an internally reversible process receives
energy by heat transfer, the system experiences an
increase in entropy. Conversely, when energy is removed
by heat transfer, the entropy of the system decreases.
From these considerations, we say that entropy transfer
accompanies heat transfer. The direction of the entropy
transfer is the same as the heat transfer.

Entropy and Heat Transfer


In an internally reversible, adiabatic process (no heat
transfer), entropy remains constant. Such a constantentropy process is called an isentropic process.
On rearrangement, Eq. 6.2b gives

Integrating from state 1 to state 2,

(Eq. 6.23)

Entropy and Heat Transfer


From this it follows that
an energy transfer by
heat to a closed system
during an internally
reversible process is
represented by an area
on a temperature-entropy
diagram:

Calculating Entropy Change


The property data provided in Tables A-2 through A-18,
similar compilations for other substances, and numerous
important relations among such properties are established
using the TdS equations. When expressed on a unit mass
basis, these equations are

(Q )intrev = dU + (W )intrev

First Law, Closed System

(Eq. 6.10a)
(Eq. 6.10b)

dH = dU + d ( pV ) = dU + pdV +Vdp
dU + pdV = dH Vdp

Calculating Entropy Change


As an application, consider a
change in phase from saturated
liquid to saturated vapor at
constant pressure.
Since pressure is constant, Eq.
6.10b reduces to give
dh
ds =
T
Then, because temperature is also constant during the
phase change

(Eq. 6.12)
This relationship is applied in property tables for
tabulating (sg sf) from known values of (hg hf).

Calculating Entropy Change


For example, consider water vapor at 100oC
(373.15 K). From Table A-2, (hg hf) = 2257.1 kJ/kg.
Then
(sg sf) = (hg hf) /T = (2257.1 kJ/kg)/373.15 K
= 6.049 kJ/kgK
which agrees with the value from Table A-2, as
expected.

Next, the TdS equations are applied to two


additional cases: substances modeled as
incompressible and gases modeled as ideal
gases.

Calculating Entropy Change of an


Incompressible Substance
The incompressible substance model assumes the specific
volume is constant and specific internal energy depends
solely on temperature: u = u(T). Then, du = c(T)dT, where
c denotes specific heat.
With these relations, Eq. 6.10a reduces to give

On integration, the change in specific entropy is

When the specific heat is constant

(Eq. 6.13)

Calculating Entropy Change of an Ideal Gas


The ideal gas model assumes pressure, specific volume
and temperature are related by pv = RT. Also, specific
internal energy and specific enthalpy each depend solely
on temperature: u = u(T), h = h(T), giving du = cvdT and
dh = cpdT, respectively.
Using these relations and integrating, the TdS equations
give, respectively

(Eq. 6.17)

(Eq. 6.18)

Calculating Entropy Change of an Ideal Gas


Since cv and cp are functions of temperature for ideal gases,
such functional relations are required to perform the
integration of the first term on the right of Eqs. 6.17 and 6.18.
For several gases modeled as ideal gases, including air,
CO2, CO, O2, N2, and water vapor, the evaluation of
entropy change can be reduced to a convenient tabular
approach using the variable so defined by

(Eq. 6.19)
where T ' is an arbitrary reference temperature.

Calculating Entropy Change of an Ideal Gas


Using so, the integral term of Eq. 6.18 can be expressed as

Accordingly, Eq. 6.18 becomes

(Eq. 6.18)

(Eq. 6.20a)
or on a per mole basis as

(Eq. 6.20b)
For air, Tables A-22 and A-22E provide so in units of
kJ/kgK and Btu/lboR, respectively. For the other gases
mentioned, Tables A-23 and A-23E provide s o in units of
kJ/kmolK and Btu/lbmoloR, respectively.

Calculating Entropy Change of an Ideal Gas


Example: Determine the change in specific entropy, in
kJ/kgK, of air as an ideal gas undergoing a process from
T1 = 300 K, p1 = 1 bar to T2 = 1420 K, p2 = 5 bar.
From Table A-22, we get so1 = 1.70203 and so2 = 3.37901,
each in kJ/kgK. Substituting into Eq. 6.20a
s2 s1 = (3.37901 1.70203)

kJ
kg K

kJ
8.314 kJ 5 bar

.
215
1
ln
=

1 bar
kg K
.
97
kg
K
28

Ideal Gas Properties of Air


T(K), h and u(kJ/ kg), so (kJ/ kgK)
when s = 0

Table A-22
T
250
260
270
280
285
290
295
300
305
310

h
250.05
260.09
270.11
280.13
285.14
290.16
295.17
300.19
305.22
310.24

u
178.28
185.45
192.60
199.75
203.33
206.91
210.49
214.07
217.67
221.25

so
1.51917
1.55848
1.59634
1.63279
1.65055
1.66802
1.68515
1.70203
1.71865
1.73498

pr
0.7329
0.8405
0.9590
1.0889
1.1584
1.2311
1.3068
1.3860
1.4686
1.5546

vr
979.
887.8
808.0
738.0
706.1
676.1
647.9
621.2
596.0
572.3

T
1400
1420
1440
1460
1480
1500
1520
1540
1560
1580

h
1515.42
1539.44
1563.51
1587.63
1611.79
1635.97
1660.23
1684.51
1708.82
1733.17

when s = 0
u
1113.52
1131.77
1150.13
1168.49
1186.95
1205.41
1223.87
1242.43
1260.99
1279.65

so
3.36200
3.37901
3.39586
3.41247
3.42892
3.44516
3.46120
3.47712
3.49276
3.50829

pr
450.5
478.0
506.9
537.1
568.8
601.9
636.5
672.8
710.5
750.0

vr
8.919
8.526
8.153
7.801
7.468
7.152
6.854
6.569
6.301
6.046

Problem 6.15
One-tenth kmol of carbon monoxide (CO) is a piston-cylinder assembly undergoes a
process from p1 = 150 kPa, T1 = 300 K to p2 = 500 kPa, T2 = 370 K. For the process,
W = -300 kJ. Employing the ideal gas model, determine
a) the heat transfer, in kJ, and
b) the change in entropy, in kJ/K.
c) Show the process on a T-s diagram.

Work through examples in


Sections 6.3 6.5

then

Work on homework
problems 6.1, 6.3, 6.10

Lecture 20

Entropy in Closed Systems


Entropy Rate in Closed Volumes

Sections 6.6, 6.7, 6.9, 6.10

Entropy and Heat Transfer


By inspection of Eq. 6.2a, the defining equation for
entropy change on a differential basis is

(Eq. 6.2b)
Equation 6.2b indicates that when a closed system
undergoing an internally reversible process receives
energy by heat transfer, the system experiences an
increase in entropy. Conversely, when energy is removed
by heat transfer, the entropy of the system decreases.
From these considerations, we say that entropy transfer
accompanies heat transfer. The direction of the entropy
transfer is the same as the heat transfer.

Entropy and Heat Transfer


In an internally reversible, adiabatic process (no heat
transfer), entropy remains constant. Such a constantentropy process is called an isentropic process.
On rearrangement, Eq. 6.2b gives

Integrating from state 1 to state 2,

(Eq. 6.23)

Entropy and Heat Transfer


From this it follows that
an energy transfer by
heat to a closed system
during an internally
reversible process is
represented by an area
on a temperature-entropy
diagram:

Problem 6.24
A gas within a piston-cylinder assembly undergoes an isothermal process at
400 K during which the change in entropy is -0.3 kJ/K. Assuming the ideal gas
model for the gas and negligible KE and PE effect, evaluate the work, in kJ.

Work through Example 6.1

then

Work on homework
problem 6.23

Entropy Balance for Closed Systems


The entropy balance for closed systems can be
developed using the Clausius inequality expressed as
Eq. 5.13 and the defining equation for entropy change,
Eq. 6.2a. The result is
where the subscript b
indicates the integral
(Eq. 6.24) is evaluated at the
system boundary.

In accord with the interpretation of cycle in the Clausius


inequality, Eq. 5.14, the value of in Eq. 6.24 adheres to the
following interpretation

= 0 (no irreversibilities present within the system)


> 0 (irreversibilities present within the system)
< 0 (impossible)

Entropy Balance for Closed Systems


That has a value of zero when there are no internal
irreversibilities and is positive when irreversibilities are
present within the system leads to the interpretation that
accounts for entropy produced (or generated) within the
system by action of irreversibilities.
Expressed in words, the entropy balance is
change in the amount
of entropy contained
within the system
during some
time interval

net amount of
amount of
entropy transferred in
entropy produced
across the system boundary + within the system
accompanying heat transfer
during some
during some time interval
time interval

Entropy Balance for Closed Systems


Example: One kg of water vapor contained
within a piston-cylinder assembly, initially at
5 bar, 400oC, undergoes an adiabatic
expansion to a state where pressure is 1 bar
and the temperature is (a) 200oC, (b) 100oC.
Using the entropy balance, determine the
nature of the process in each case.

Boundary

Since the expansion occurs adiabatically, Eq. 6.24


reduces to give
0

S 2 S1 =

+
T b

m(s2 s1) =

(1)

where m = 1 kg and Table A-4 gives s1 = 7.7938 kJ/kgK.

Entropy Balance for Closed Systems


(a) Table A-4 gives, s2 = 7.8343 kJ/kgK. Then
Eq. (1) gives
= (1 kg)(7.8343 7.7938) kJ/kgK = 0.0405 kJ/K
Since is positive, irreversibilities are present within
the system during expansion (a).

(b) Table A-4 gives, s2 = 7.3614 kJ/kgK. Then


Eq. (1) gives
= (1 kg)(7.3614 7.7938) kJ/kgK = 0.4324 kJ/K
Since is negative, expansion (b) is impossible: it
cannot occur adiabatically.

Entropy Balance for Closed Systems


More about expansion (b): Considering Eq. 6.24

<0 = <0 + 0
Since cannot be negative and
For expansion (b) S is negative, then
By inspection the integral must be negative and
so heat transfer from the system must occur in
expansion (b).

Entropy Rate Balance for Closed Systems


On a time rate basis, the closed system entropy
rate balance is
(Eq. 6.28)
where

dS
= the time rate of change of the entropy of the
dt system
Q j
Tj

= the time rate of entropy transfer through the

portion of the boundary whose temperature is Tj

= time rate of entropy production due to


irreversibilities within the system

Entropy Rate Balance for Closed Systems


Example: An inventor claims that the device shown
generates electricity at a rate of 100 kJ/s while receiving a
heat transfer of energy at a rate of 250 kJ/s at a temperature
of 500 K, receiving a second heat transfer at a rate of 350
kJ/s at 700 K, and discharging energy by heat transfer at a
rate of 500 kJ/s at a temperature of 1000 K. Each heat
transfer is positive in the direction of the accompanying
arrow. For operation at steady state, evaluate this claim.
T1 = 500 K

T2 = 700 K

Q
2

= 350 kJ/s

T3 = 1000 K

Entropy Rate Balance for Closed Systems


Applying an energy rate balance
at steady state

dE
= 0 = Q1 + Q 2 Q 3 W e
dt

Solving W e = 250 kJ/s + 350 kJ/s 500 kJ/s = 100 kJ/s


The claim is in accord with the first law of thermodynamics.
Applying an entropy rate balance
at steady state
Solving

dS
=0=
dt

Q1 Q 2 Q 3
+

+
T1 T2 T3

250 kJ/s 350 kJ/s 500 kJ/s


+

700 K
1000 K
500 K
kJ/s
kJ/s
= 0.5
= (0.5 + 0.5 0.5)
K
K

Since is negative, the claim is not in accord with the


second law of thermodynamics and is therefore denied.

Work through Example 6.2, 6.3, 6.4

then

Work on homework
problem 6.41

Entropy Rate Balance for Control Volumes


Like mass and energy, entropy can be transferred into or
out of a control volume by streams of matter.
Since this is the principal difference between the closed
system and control volume entropy rate balances, the
control volume form can be obtained by modifying the
closed system form to account for such entropy transfer.
The result is

(Eq. 6.34)
i si and m e se account, respectively, for rates of entropy
where m
transfer accompanying mass flow at inlets i and exits e.

Entropy Rate Balance for Control Volumes


For control volumes at steady state, Eq. 6.34 reduces to
give

(Eq. 6.36)
For a one-inlet, one-exit control volume at steady state,
Eq. 6.36 reduces to give

(Eq. 6.37)
where 1 and 2 denote the inlet and exit, respectively, and
is the common mass flow rate at these locations.

Entropy Rate Balance for Control Volumes


Example: Water vapor enters a valve at 0.7
bar, 280oC and exits at 0.35 bar. (a) If the
water vapor undergoes a throttling process,
determine the rate of entropy production
within the valve, in kJ/K per kg of water
p1 = 0.7 bar
vapor flowing. (b) What is the source of
T1 = 280oC
entropy production in this case?
(a) For a throttling process, there is no significant
heat transfer. Thus, Eq. 6.37 reduces to
0
Q j
(s1 s2 ) + cv
0=
+ m (s1 s2 ) + cv 0 = m
Tj

p2 = 0.35 bar
h2 = h1

Entropy Rate Balance for Control Volumes


cv

Solving

= s 2 s1

From Table A-4, h1 = 3035.0 kJ/kg, s1 = 8.3162 kJ/kgK.


For a throttling process, h2 = h1 (Eq. 4.22). Interpolating
in Table A-4 at 0.35 bar and h2 = 3035.0 kJ/kg,
s2 = 8.6295 kJ/kgK.
Finally

cv
m

= (8.6295 8.3162) kJ/kgK = 0.3133 kJ/kgK

(b) Selecting from the list of irreversibilities provided in


Sec. 5.3.1, the source of the entropy production here is
the unrestrained expansion to a lower pressure undergone
by the water vapor.

Entropy Rate Balance for Control Volumes


Comment: The value of the entropy production for a single
component such as the throttling valve considered here often
does not have much significance by itself. The significance of
the entropy production of any component is normally
determined through comparison with the entropy production
values of other components combined with that component to
form an integrated system. Reducing irreversibilities of
components with the highest entropy production rates may
lead to improved thermodynamic performance of the
integrated system.

Work through Example 6.6, 6.7, 6.8

then

Work on homework
problem 6.41

Lecture 21

Isentropic Processes
Turbines, Compressors
and Pumps

Sections 6.11, 6.12

Calculating Entropy Change of an Ideal Gas


Tables A-22 and A-22E provide
additional data for air modeled
as an ideal gas. These values,
denoted by pr and vr, refer only
to two states having the same
specific entropy. This case has
important applications, and is
shown in the figure.

Calculating Entropy Change of an Ideal Gas


When s2 = s1, the following equation relates T1, T2, p1,
and p2
p2 pr (T2 )
=
(s1 = s2, air only) (Eq. 6.41)
p1
pr (T1 )
where relative pressure pr(T ) is read from Table A-22
or A-22E, as appropriate.
Ideal Gas Properties of Air
T(K), h and u(kJ/ kg), so (kJ/ kgK)
when s = 0

Table A-22
T
250
260
270
280
285
290
295
300
305
310

h
250.05
260.09
270.11
280.13
285.14
290.16
295.17
300.19
305.22
310.24

u
178.28
185.45
192.60
199.75
203.33
206.91
210.49
214.07
217.67
221.25

so
1.51917
1.55848
1.59634
1.63279
1.65055
1.66802
1.68515
1.70203
1.71865
1.73498

pr
0.7329
0.8405
0.9590
1.0889
1.1584
1.2311
1.3068
1.3860
1.4686
1.5546

vr
979.
887.8
808.0
738.0
706.1
676.1
647.9
621.2
596.0
572.3

T
1400
1420
1440
1460
1480
1500
1520
1540
1560
1580

h
1515.42
1539.44
1563.51
1587.63
1611.79
1635.97
1660.23
1684.51
1708.82
1733.17

when s = 0
u
1113.52
1131.77
1150.13
1168.49
1186.95
1205.41
1223.87
1242.43
1260.99
1279.65

so
3.36200
3.37901
3.39586
3.41247
3.42892
3.44516
3.46120
3.47712
3.49276
3.50829

pr
450.5
478.0
506.9
537.1
568.8
601.9
636.5
672.8
710.5
750.0

vr
8.919
8.526
8.153
7.801
7.468
7.152
6.854
6.569
6.301
6.046

Calculating Entropy Change of an Ideal Gas


Example: Air undergoes a process from T1 = 620 K, p1 = 12 bar
to a final state where s2 = s1, p2 = 1.4 bar. Employing the ideal
gas model, determine the final temperature T2, in K. Solve using
(a) pr data from Table A-22.
(a) With Eq. 6.41 and pr(T1) = 18.36 from Table A-22
p2
1.4 bar

= 18.36
pr (T2 ) = pr (T1 )
= 2.142

12 bar
p1

Interpolating in Table A-22, T2 = 339.7 K


Ideal Gas Properties of Air
T(K), h and u(kJ/ kg), so (kJ/ kgK)
when s = 0

Table A-22
T
315
320
325
330
340
350

h
315.27
320.29
325.31
330.34
340.42
350.49

u
224.85
228.42
232.02
235.61
242.82
250.02

s
1.75106
1.76690
1.78249
1.79783
1.82790
1.85708

pr
1.6442
1.7375
1.8345
1.9352
2.149
2.379

vr
549.8
528.6
508.4
489.4
454.1
422.2

T
600
610
620
630
640
650

h
607.02
617.53
628.07
638.63
649.22
659.84

when s = 0
u
434.78
442.42
450.09
457.78
465.50
473.25

s
2.40902
2.42644
2.44356
2.46048
2.47716
2.49364

pr
16.28
17.30
18.36
19.84
20.64
21.86

vr
105.8
101.2
96.92
92.84
88.99
85.34

Calculating Entropy Change of an Ideal Gas


When s2 = s1, the following equation relates T1, T2, v1,
and v2
v 2 v r (T2 )
=
(s1 = s2, air only) (Eq. 6.42)
v1 v r (T1 )
where vr(T ) is read from Table A-22 or A-22E, as
appropriate.
Ideal Gas Properties of Air
T(K), h and u(kJ/ kg), so (kJ/ kgK)
when s = 0

Table A-22
T
250
260
270
280
285
290
295
300
305
310

h
250.05
260.09
270.11
280.13
285.14
290.16
295.17
300.19
305.22
310.24

u
178.28
185.45
192.60
199.75
203.33
206.91
210.49
214.07
217.67
221.25

so
1.51917
1.55848
1.59634
1.63279
1.65055
1.66802
1.68515
1.70203
1.71865
1.73498

pr
0.7329
0.8405
0.9590
1.0889
1.1584
1.2311
1.3068
1.3860
1.4686
1.5546

vr
979.
887.8
808.0
738.0
706.1
676.1
647.9
621.2
596.0
572.3

T
1400
1420
1440
1460
1480
1500
1520
1540
1560
1580

h
1515.42
1539.44
1563.51
1587.63
1611.79
1635.97
1660.23
1684.51
1708.82
1733.17

when s = 0
u
1113.52
1131.77
1150.13
1168.49
1186.95
1205.41
1223.87
1242.43
1260.99
1279.65

so
3.36200
3.37901
3.39586
3.41247
3.42892
3.44516
3.46120
3.47712
3.49276
3.50829

pr
450.5
478.0
506.9
537.1
568.8
601.9
636.5
672.8
710.5
750.0

vr
8.919
8.526
8.153
7.801
7.468
7.152
6.854
6.569
6.301
6.046

Entropy Change of an Ideal Gas


Assuming Constant Specific Heats
When the specific heats cv and cp are assumed constant,
Eqs. 6.17 and 6.18 reduce, respectively, to
(Eq. 6.17)

(Eq. 6.18)

(Eq. 6.21)

(Eq. 6.22)

These expressions have many


applications. In particular, they can be
applied to develop relations among T,
p, and v at two states having the same
specific entropy as shown in the figure.

Entropy Change of an Ideal Gas


Assuming Constant Specific Heats
Since s2 = s1, Eqs. 6.21 and
6.22 become
With the ideal gas relations
where k is the specific ratio
These equations
can be solved,
respectively, to give
Eliminating the
temperature ratio gives

(Eq. 6.43)
(Eq. 6.44)

(Eq. 6.45)

Calculating Entropy Change of an Ideal Gas


Example: Air undergoes a process from T1 = 620 K, p1 = 12 bar
to a final state where s2 = s1, p2 = 1.4 bar. Employing the ideal
gas model, determine the final temperature T2, in K. Solve using
(b) a constant specific heat ratio k evaluated at 620 K from Table
A-20: k = 1.374. Comment.
(b) With Eq. 6.43
p2

T2 = T1
p1

(k 1) / k

1.4 bar
= 620 K

12 bar

(0.374 / 1.374 )

T2 = 345.5 K
Comment: The approach of (a) accounts for variation of
specific heat with temperature but the approach of (b) does not.
With a k value more representative of the temperature interval,
the value obtained in (b) using Eq. 6.43 would be in better
agreement with that obtained in (a) with Eq. 6.41.

Work through Example 6.9, 6.10

then

Work on homework
problems 6.123, 6.128

Isentropic Turbine Efficiency


For a turbine, the energy rate
balance reduces to

2
2

(V

V
)
1
2

0 = Qcv Wcv + m (h1 h2 ) +


+ g ( z1 z2 )
2

If the change in kinetic energy of flowing matter is negligible,


(V12 V22) drops out.
If the change in potential energy of flowing matter is
negligible, g(z1 z2) drops out.
If the heat transfer with surroundings is negligible, Q cv drops
out.

W cv
= h1 h2
m

where
the left side is work developed per unit of mass flowing.

Isentropic Turbine Efficiency


For a turbine, the entropy rate
balance reduces to

If the heat transfer with surroundings is negligible, Q j


drops out.

cv
m

= s2 s1 0

Isentropic Turbine Efficiency


Since the rate of entropy production
cannot be negative, the only turbine exit
states that can be attained in an adiabatic
expansion are those with s2 s1. This is
shown on the Mollier diagram to the right.
The state labeled 2s on the figure would be attained only in an
isentropic expansion from the inlet state to the specified exit
pressure that is, 2s would be attained only in the absence of
internal irreversibilities. By inspection of the figure, the
maximum theoretical value for the turbine work per unit of mass
flowing is developed in such an internally reversible, adiabatic
expansion:
W cv

m = h1 h2s
s

Isentropic Turbine Efficiency


The isentropic turbine
efficiency is the ratio of the
actual turbine work to the
maximum theoretical work,
each per unit of mass flowing:

(Eq. 6.46)

Isentropic Turbine Efficiency


Example: Water vapor enters a turbine
at p1 = 5 bar, T1 = 320oC and exits at p2 1
= 1 bar. The work developed is
measured as 271 kJ per kg of water
vapor flowing. Applying Eq. 6.46,
determine the isentropic turbine
efficiency.
From Table A-4, h1 = 3105.6 kJ/kg, s1 = 7.5308 kJ/kg.
With s2s = s1, interpolation in Table A-4 at a pressure of
1 bar gives h2s = 2743.0 kJ/kg. Substituting values into
Eq. 6.46
W cv / m
271 kJ/kg
t =
=
= 0.75 (75%)
h1 h2s (3105.6 2743.0 ) kJ/kg

Isentropic Compressor and Pump Efficiencies


For a compressor the energy rate
balance reduces to

1
2

2
2

(V

V
)
1
2

0 = Qcv Wcv + m (h1 h2 ) +


+ g ( z1 z2 )
2

If the change in kinetic energy of flowing matter is negligible,


(V12 V22) drops out.
If the change in potential energy of flowing matter is
negligible, g(z1 z2) drops out.
If the heat transfer with surroundings is negligible, Q cv drops
out.

W cv

= h2 h1

where
the left side is work input per unit of mass flowing.

Isentropic Compressor and Pump Efficiencies


For a compressor the entropy
rate balance reduces to

If the heat transfer with surroundings is negligible, Q j


drops out.

cv
m

= s2 s1 0

Isentropic Compressor and Pump Efficiencies


Since the rate of entropy production
cannot be negative, the only
compressor exit states that can be
attained in an adiabatic compression
are those with s2 s1. This is shown on
the Mollier diagram to the right.
The state labeled 2s on the figure would be attained only in an
isentropic compression from the inlet state to the specified exit
pressure that is, state 2s would be attained only in the
absence of internal irreversibilities. By inspection of the figure,
the minimum theoretical value for the compressor work input
per unit of mass flowing is for such an internally reversible,
adiabatic compression:
W cv

= h2s h1

Isentropic Compressor and Pump Efficiencies


The isentropic compressor
efficiency is the ratio of the
minimum theoretical work
input to the actual work input,
each per unit of mass flowing:

(Eq. 6.48)
An isentropic pump efficiency is defined similarly.

Work through Examples


6.11, 6.12, 6.13, 6.14a

then

Work on homework
problems 6.136, 6.153

Lecture 22

Internally Reversible,
Steady-State Processes
and
Statistical Forms of Entropy

Sections 6.13, 6.8

Heat Transfer and Work in Internally Reversible


Steady-State Flow Processes
Consider a one-inlet, oneexit control volume at
steady state:
Compressors, pumps, and
other devices commonly
encountered in engineering
practice are included in this
class of control volumes.

2
W cv

The objective is to introduce expressions for the heat


transfer rate Q cv / m and work rate W cv / m in the absence of
internal irreversibilities. The resulting expressions have
important applications.

Heat Transfer and Work in Internally Reversible


Steady-State Flow Processes
In agreement with the discussion of energy transfer by
heat to a closed system during an internally reversible
process (Sec. 6.6.1), in the present application we have

(Eq. 6.49)
where the subscript int rev signals that the expression applies
only in the absence of internal irreversibilities.

As shown by the figure, when the states


visited by a unit mass passing from inlet to
exit without internal irreversibilities are
described by a curve on a T-s diagram, the
heat transfer per unit of mass flowing is
represented by the area under the curve.

Heat Transfer and Work in Internally Reversible


Steady-State Flow Processes
Neglecting kinetic and potential energy effects, an energy
rate balance for the control volume reduces to
W cv

V12 V22
Q cv

+ g (z1 z 2 )

=
int m int + (h1 h2 ) +

rev

rev

With Eq. 6.49, this becomes

(1)

Since internal irreversibilities are assumed absent, each unit


of mass visits a sequence of equilibrium states as it passes
from inlet to exit. Entropy, enthalpy, and pressure changes
are therefore related by the TdS equation, Eq. 6.10b:

Heat Transfer and Work in Internally Reversible


Steady-State Flow Processes
Integrating from inlet to exit:
With this relation Eq. (1) becomes
(Eq. 6.51b)
If the specific volume remains approximately constant,
as in many applications with liquids, Eq. 6.51b becomes
(Eq. 6.51c)
This is applied in the discussion of vapor power cycles
in Chapter 8.

Heat Transfer and Work in Internally Reversible


Steady-State Flow Processes
As shown by the figure, when the states visited
by a unit mass passing from inlet to exit without
internal irreversibilities are described by a curve
on a p-v diagram, the magnitude of vdp is
shown by the area behind the curve.

Heat Transfer and Work in Internally Reversible


Steady-State Flow Processes
Example: A compressor operates at
steady state with natural gas entering at
at p1, v1. The gas undergoes a polytropic
process described by pv = constant and
exits at a higher pressure, p2.

(a) Ignoring kinetic and potential energy effects, evaluate


the work per unit of mass flowing.
(b) If internal irreversibilities were present, would the
magnitude of the work per unit of mass flowing be less
than, the same as, or greater than determined in part (a)?

Heat Transfer and Work in Internally Reversible


Steady-State Flow Processes
(a) With pv = constant, Eq. 6.51b gives

The minus sign indicates that the compressor


requires a work input.

(b) Left for class discussion.

Work through examples


at the end of sections
and Example 6.15

then

Work on homework
problem 6.163

Increase in Entropy Principle


For any process in an isolated system, the energy
balance equation must say that energy is conserved,
as there is no energy transfer with the surroundings:

E ]isolated = 0
If there is no energy transfer with the surroundings,
there is no change in energy of the surroundings
during that process as well:

E ]isolated + E ]surroundings = 0

Increase in Entropy Principle


Looking at entropy for the isolated system during this process:

S ]isolated =

+ isolated
T b

Entropy is produced in the system during the process, but not


in the surroundings, so the total entropy change everywhere
is:

S ]isolated + S ]surroundings = isolated

Since there is some amount of entropy production for all real


processes, the increase of entropy principle says that
processes only occur in the direction of total entropy change
of the system plus surroundings increases.

Statistical Interpretation of Entropy


We start with a box half filled
with particles in equilibrium,
kept there by a partition down
the middle.
Then we remove the partition,
let the particles come to
equilibrium again, U and T
havent changed, but volume
is twice as large and we have
less information about where
the particles are. We can use:
to calculate the entropy of the
system now.

(Eq. 6.31)

Statistical Interpretation of Entropy


There are many more possibilities of locations (and sometimes
momenta) for these particles. There is a relationship between
the number of possible combinations of these locations and
momenta, or microstates, available to the particles in a system
and entropy.
(Eq. 6.32)
where
S:entropy,
N:number of particles,
k:Boltzmanns constant
w: thermodynamic probability (based on probability of each state possible)

Any process that increases the number of possible microstates


increases entropy. An increased number of possible
microstates means we know less about the system, so we can
look at an increase in entropy is an increase in disorder.

Example of Microstates
16 different ways to arrange 4 particles

Work through Example 6.5


and read the grey
Breaking the Second Law insert

Lecture 23

Power Generation,
Vapor Power Plants
and
the Rankine Cycle

Sections 8.1, 8.2.1, 8.2.2

Introducing Power Generation


To meet our national power needs there are
challenges related to
Declining economically
recoverable supplies of
nonrenewable energy resources.
Effects of global climate change
and other environmental and human
health and safety issues.
Rapidly increasing demand for
power owing to increasing
population.

Today we are heavily dependent on coal, natural


gas, and nuclear, all of which are nonrenewable.

Introducing Power Generation


While coal, natural gas, and nuclear will continue to play
important roles in years ahead, contributions from wind power,
solar power, and other renewable sources are expected to be
increasingly significant up to mid-century at least.
Table 8.2

Introducing Power Generation


Table 8.2 also shows that thermodynamic cycles
are a fundamental aspect of several power plant
types that employ renewable or nonrenewable
sources.
In Chapters 8 and 9, vapor power systems, gas
turbine power systems, and internal combustion
engines are studied as thermodynamic cycles.
Vapor power systems in which a working fluid is
alternately vaporized and condensed is the focus of
Chapter 8. The basic building block of vapor power
systems is the Rankine cycle.

Introducing Vapor Power Plants


The components of four alternative vapor power
plant configurations are shown schematically in Fig.
8.1. They are
Fossil-fueled vapor power plants.
Pressurized-water reactor nuclear vapor power
plants.
Concentrating solar thermal vapor power plants.
Geothermal vapor power plants.

In each of the four types of vapor power plant, a


working fluid is alternately vaporized and
condensed. A key difference among the plants is the
origin of the energy required to vaporize the working
fluid.

Vapor Power Plants - General Design


A: Component for delivering energy required
for vaporization.
B-D remain the same.

Vapor Power Plants: Fossil Fuel


In fossil-fueled plants, the energy required for
vaporization originates in combustion of the fuel.

Vapor Power Plants: Nuclear


In nuclear plants, the energy required for
vaporization originates in a controlled nuclear
reaction.

Vapor Power Plants: Solar


In solar plants, energy required for vaporization
originates in collected and concentrated solar
radiation.

Vapor Power Plants: Geothermal


In geothermal plants, the energy required for
vaporization originates in hot water and/or steam
drawn from below the earths surface.

Introducing Vapor Power Plants


The fossil-fueled vapor power plant of Fig. 8.1(a) will be
considered as representative. The overall plant is broken into
four major subsystems identified by A, B, C, and D. Water is
the working fluid.

Introducing Vapor Power Plants


Subsystem A provides the heat transfer of energy needed to
vaporize water circulating in subsystem B. In fossil-fueled
plants this heat transfer has its origin in the combustion of the
fuel.

Introducing Vapor Power Plants


In subsystem B, the water vapor expands through the
turbine, developing power. The water then condenses and
returns to the boiler.

Introducing Vapor Power Plants


In subsystem C, power developed by the turbine drives an
electric generator.

Introducing Vapor Power Plants


Subsystem D removes energy by heat transfer arising from
steam condensing in subsystem B.

Introducing Vapor Power Plants


Each unit of mass of water periodically undergoes a
thermodynamic cycle as it circulates through the components
of subsystem B. This cycle is the Rankine cycle.

Power Cycle Review


The first law of thermodynamics
requires the net work developed by a
system undergoing a power cycle to
equal the net energy added by heat
transfer to the system:

Wcycle = Qin Qout


The thermal efficiency of a power
cycle is
W cycle
=
Q
in

Power Cycle Review


The second law of thermodynamics requires the thermal
efficiency to be less than 100%. Most of todays vapor power
plants have thermal efficiencies ranging up to about 40%.
Thermal efficiency tends to increase as the average
temperature at which energy is added by heat transfer
increases and/or the average temperature at which energy is
rejected by heat transfer decreases.
Improved thermodynamic performance of power cycles, as
measured by increased thermal efficiency, for example, also
accompanies the reduction of irreversibilities and losses.
The extent of improved power cycle performance is limited,
however, by constraints imposed by thermodynamics and
economics.

The Rankine Cycle


Each unit of mass of water circulating through the
interconnected components of Subsystem B of Fig. 8.1(a)
undergoes a thermodynamic cycle known as the Rankine cycle.
There are four principal
control volumes involving
these components:
Turbine
Condenser
Pump
Boiler
All energy transfers by work and heat are taken as positive in
the directions of the arrows on the schematic and energy
balances are written accordingly.

The Rankine Cycle


The processes of the Rankine cycle are
Process1-2: vapor expands through the turbine
developing work
Process 2-3: vapor condenses to liquid through
heat transfer to cooling water
Process 3-4: liquid is pumped into the boiler
requiring work input
Process 4-1: liquid is heated to saturation and
evaporated in the boiler through
heat transfer from the energy source

The Rankine Cycle


Engineering model:
Each component is analyzed as a
control volume at steady state.
The turbine and pump operate
adiabatically.
Kinetic and potential energy changes
are ignored.

The Rankine Cycle


Applying mass and energy rate balances

Turbine

W t
= h1 h2
m

(Eq. 8.1)

Condenser

Q out
= h2 h3
m

(Eq. 8.2)

Pump
Boiler

W p

= h4 h3

(Eq. 8.3)

Q in
= h1 h4
m

(Eq. 8.4)

The Rankine Cycle


Performance parameters

Thermal Efficiency
Wcycle W t / m W p / m (h1 h2 ) (h4 h3 )
= =
=

Q
Q / m
(h h )
in

in

Back Work Ratio

(Eq. 8.5a)

W p / m (h4 h3 )
bwr =
=

Wt / m (h1 h2 )

(Eq. 8.6)

Back work ratio is characteristically low for vapor


power plants. For instance, in Example 8.1, the
power required by the pump is less than 1% of the
power developed by the turbine.

The Rankine Cycle


Provided states 1 through 4 are fixed,
Eqs. 8.1 through 8.6 can be applied to
determine performance of simple vapor
power plants adhering to the Rankine cycle.
Since these equations are developed
from mass and energy balances, they apply
equally when irreversibilities are present
and for idealized performance in the
absence of such effects.

Ideal Rankine Cycle


The ideal Rankine cycle provides a simple
setting to study aspects of vapor power plant
performance. The ideal cycle adheres to
additional modeling assumptions:
Frictional pressure drops are absent during
flows through the boiler and condenser. Thus,
these processes occur at constant pressure.
Flows through the turbine and pump occur
adiabatically and without irreversibility. Thus,
these processes are isentropic.

Ideal Rankine Cycle


The ideal Rankine cycle consists of four internally
reversible processes:
Process1-2: Isentropic expansion
through the turbine
Process 2-3: Heat transfer from the
working fluid passing through the
condenser at constant pressure
Process 3-4: Isentropic compression in the pump
Process 4-1: Heat transfer to the working fluid passing
through the boiler at constant pressure
Possible variations of this cycle include: turbine inlet state
is superheated vapor, State 1; pump inlet, State 3, falls in
the liquid region.

Ideal Rankine Cycle


Since the ideal Rankine cycle involves
internally reversible processes, results from
Sec. 6.13 apply.
Applying Eq. 6.51c, the pump work input per
unit of mass flowing is evaluated as follows
W p

v ( p p ) (Eq. 8.7b)
3 4
3

where v3 is the specific volume at the pump


inlet and the subscript s signals the isentropic
process of the liquid flowing through the pump.

Ideal Rankine Cycle


Applying Eq. 6.49, areas under process lines on
the T-s diagram can be interpreted as heat transfers,
each per unit of mass flowing:
Area 1-b-c-4-a-1 represents
heat transfer to the water flowing
through the boiler
Area 2-b-c-3-2 represents heat
transfer from the water flowing
through the condenser.
Enclosed area 1-2-3-4-a-1
represents the net heat input or
equivalently the net work
developed by the cycle.

Carnot vs. Rankine Cycle

Internally reversible
processes.
Highest thermal efficiency.
Very difficult to build
practically (limited temperatures
and phases) and robustly (pump
wants X=0 or damage)

Internally reversible
processes (ideal).
Lower thermal efficiency.
Can build in many different
ways: freedom in the
cost/power-output landscape.

Work through Example 8.1

then

Work on homework
problem 8.2
(Well do Problem 8.7 in class)

Bubbles: Cavitation-Induced Damage


Cavitation bubble seen on
high-speed camera

http://www.scs.illinois.edu/suslick/sonoche
mbrittanica.html

Damage on metal due


to cavitation

Excerpt from:
https://www1.eere.energy.gov/manufacturing/tech_assistance/pdfs/pump.pdf

The damaging aspect of cavitation occurs when these vapor


bubbles return to liquid phase in a violent collapse. During this
collapse, high-velocity water jets impinge onto surrounding
surfaces. The force of this impingement often exceeds the
mechanical strength of the impacted surface, which leads to
material loss. Over time, cavitation can create severe erosion
problems in pumps, valves, and pipes.

Lecture 24

Pressure Losses and Irreversibilities


in the Rankine Cycle
and Considerations for
Improving Cycle Performance

Sections 8.2.3, 8.2.4, 8.3

Using the Ideal Rankine Cycle to Study the Effects


on Performance of Varying Condenser Pressure
The figure shows two cycles having the same boiler
pressure but different condenser pressures:
one is at atmospheric pressure and
the other is at less than atmospheric pressure.

Using the Ideal Rankine Cycle to Study the Effects


on Performance of Varying Condenser Pressure
Since the temperature of heat rejection for cycle 1-2-3-4-1
is lower than for cycle 1-2-3-4-1, the cycle condensing below
atmospheric pressure has the greater thermal efficiency.
Reducing condenser pressure tends to increase thermal
efficiency.

max

T < 100o C

TC
= 1
TH

Using the Ideal Rankine Cycle to Study the Effects


on Performance of Varying Condenser Pressure
Power plant condensers normally operate with steam condensing at a
pressure well below atmospheric. For heat rejection to the
surroundings the lowest feasible condenser pressure is the saturation
pressure corresponding to the ambient temperature.
A primary reason for including condensers in vapor power plants is
the increase in thermal efficiency realized when the condenser operates
at a pressure less than atmospheric pressure. Another is that the
condenser allows the working fluid to flow in a closed loop and so
demineralized water that is less corrosive than tap water can be used
economically.
As steam condenses, energy is discharged by heat transfer to cooling
water flowing separately through the condenser. Although the cooling
water carries away considerable energy, its utility is extremely limited
because cooling water temperature is increased only by a few degrees
above the ambient. Such warm cooling water has little thermodynamic
or economic value.

Using the Ideal Rankine Cycle to Study the Effects


on Performance of Varying Boiler Pressure
The figure shows two cycles having the same condenser pressure
but different boiler pressures:
one is at a given pressure and
the other is at higher than the
given pressure.

Since the average temperature of heat addition for cycle 1-2-3-4-1 is


greater than for cycle 1-2-3-4-1, the higher pressure cycle has the greater
thermal efficiency. Increasing the boiler pressure tends to increase
thermal efficiency.

Using the Ideal Rankine Cycle to Study the Effects


on Performance of Varying Boiler Pressure
However, an increase in boiler pressure also results in a
reduction of steam quality for the expansion through the
turbine compare the lower quality state 2 with state 2.
If the quality of the expanding steam becomes too low, the
impact of liquid droplets in the steam can erode the turbine
blades and possibly diminish performance.

x2 x2

Using the Ideal Rankine Cycle to Study the Effects


on Performance of Varying Boiler Pressure
Rankine cycle modifications known as superheat and
reheat permit advantageous operating pressures in the
boiler and condenser while avoiding low-quality steam in
the turbine expansion. Another such modification is the
supercritical cycle. All three are considered next.

Ideal Rankine Cycle with Superheat


The figure shows a Rankine cycle with superheated vapor
at the turbine inlet: cycle 1-2-3-4-1.
The cycle with superheat has a
higher average temperature of
heat addition than cycle 1-2-3-4-1
without superheat, and thus has a
higher thermal efficiency.

Tavg(4-1)
Tavg(4-1)

x2

x2

Moreover, with superheat the quality at the turbine exit,


state 2, is greater than at state 2, the turbine exit without
superheat.
A combined boiler-superheater is called a steam
generator.

Ideal Reheat Cycle


The reheat cycle is another
Rankine cycle modification
that takes advantage of the
increased thermal efficiency
resulting from a high average
temperature of heat addition
and yet avoids low-quality
steam during expansion
through the turbine.
With reheat, the steam
quality at the turbine exit, state
4, is greater than at state 4,
the turbine exit without reheat.

xx44 x4

Ideal Reheat Cycle


Departures of the reheat
cycle from the simple vapor
power cycle considered thus
far include
Process 1-2. Steam
expands through the highpressure turbine stage.
Process 2-3. Steam is
reheated in the steam
generator.
Process 3-4. Steam
expands through the lowpressure turbine stage.

Ideal Reheat Cycle


When computing the
thermal efficiency of a reheat
cycle it is necessary to
account for the
Total work developed in
the two turbine stages:
processes 1-2 and 3-4.
Total heat added during
vaporization/superheating
and reheating: processes
6-1 and 2-3.
W t HP + W t LP W p
=
Q + Q
6-1

2-3

Supercritical Ideal Reheat Cycle


Steady improvements over many decades in
materials and fabrication methods have permitted
significant increases in maximum allowed temperatures
and steam generator pressures.
These efforts are embodied in the supercritical
reheat cycle.

Supercritical Ideal Reheat Cycle


Steam generation occurs at a
pressure greater than the critical
pressure of water: Process 6-1.
No phase change occurs during
process 6-1. Instead, water flowing
through tubes is heated from liquid to
vapor without the bubbling associated
with boiling.
Todays supercritical plants achieve
thermal efficiencies up to 47%.
Ultrasupercritical plants capable of
operating at still higher temperatures
and steam generator pressures will
have thermal efficiencies exceeding
50%.

Principal Irreversibilities
Irreversibilities are associated with each of the
four subsystems designated by A, B, C, and D on
the figure.

Principal Irreversibilities
For a fossil-fueled vapor power
plant the most significant site of
C
B
irreversibility by far is subsystem
D
A. There, irreversibility is
associated with combustion of the
fuel and subsequent heat transfer
from hot combustion gases to
A
vaporize water flowing through
the closed loop of subsystem B, which is the focus of discussion.
As such irreversibilities occur in the surroundings of subsystem B,
they are classified as external irreversibilities for that subsystem.
Irreversibilities in subsystems C and D are also external
irreversibilities of subsystem B but they are minor relative to the
combustion and heat transfer irreversibilities occurring in the boiler.
Combustion irreversibility is studied in Chapter 13.

Principal Irreversibilities
Irreversibilities associated with subsystem B are called
internal. They include the expansion of steam through the
turbine and the effects of friction in flow through the pump,
boiler, and condenser.
The most significant internal irreversibility associated with
subsystem B is expansion through the turbine.

Principal Irreversibilities
Isentropic turbine efficiency, introduced in Sec. 6.12.1,
accounts for the effects of irreversibilities within the turbine in
terms of actual and isentropic turbine work, each per unit of
mass flowing through the turbine.
work developed in the actual
expansion from turbine inlet state
to the turbine exit pressure

(W t / m ) (h1 h2 )
t =
=
(Wt / m )s (h1 h2s )

(Eq. 8.9)

work developed in an isentropic


expansion from turbine inlet
state to exit pressure

Principal Irreversibilities
While pump work input is much less than turbine work
output, irreversibilities in the pump affect net power output of
the vapor plant.
Isentropic pump efficiency, introduced in Sec. 6.12.3,
accounts for the effects of irreversibilities within the pump in
terms of actual and isentropic pump work input, each per unit
of mass flowing through the pump.
work input for an isentropic process
from pump inlet state to exit pressure

(W p / m ) s (h4s h3 )
(Eq. 8.10a)
p =
=
(Wp / m ) (h4 h3 )
work input for the actual process from pump
inlet state to the pump exit pressure

Principal Irreversibilities
For simplicity, frictional effects resulting in
pressure reductions from inlet to exit for the boiler
and condenser are ignored in Chapter 8.
Flow through these components is assumed to
occur at constant pressure.

Work through Example 8.2, 8.3, 8.4

then

Work on homework
problem 8.25
(well work on problem 8.22 in class)

Lecture 25

Regenerative and Cogeneration


Rankine Cycles

Sections 8.4, 8.5

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
Another method for increasing thermal efficiency
of vapor power plants is by application of
regenerative feedwater heaters.
Regenerative feedwater heating increases the
average temperature of heat addition, thereby
increasing the thermal efficiency.
In vapor power plants both open and closed
feedwater heaters are typically used. Each
feedwater heater type is considered in Sec. 8.4.

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
A vapor power cycle using a single open
feedwater heater (a mixing chamber) is shown in
the figure.

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
In this cycle, water passes isentropically through each of
two turbine stages, 1-2 and 2-3, and each of two pumps, 4-5
and 6-7.
Flow through the steam generator, 7-1, and condenser, 3-4,
occur at constant pressure.
Mixing occurs within the
feedwater heater, thereby
introducing a source of
irreversibility into an
otherwise ideal cycle.
The feedwater heater is
assumed to operate
adiabatically.

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
Referring to the T-s diagram, heat transfer to the cycle takes
place from state 7 to state 1, rather than from state a to state 1
as would be the case if there were no open feedwater heater.
Accordingly, when an open feedwater heater is included, the
average temperature of heat addition to the cycle increases;
thus, thermal efficiency increases.
Tavg(7-1)
Tavg(a-1)

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
Referring to the schematic, let us follow a unit of mass as it
moves through the cycle. The unit of mass is denoted in
parentheses by (1).
The unit of mass (1) enters the
turbine at state 1 and expands to
(1)
(1-y)
state 2 where a fraction (y) is
(y)
diverted into the open feedwater
heater.
The remaining fraction (1-y)
expands through the second
turbine stage to state 3, is
condensed to state 4, and then
pumped to state 5 where it enters
the open feedwater heater.

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
The fractions (y) and (1-y)
entering the feedwater heater
at states 2 and 5, respectively,
mix within that component,
giving a single stream at state
6 and recovering the unit of
mass (1).
The unit of mass (1) at state
6 is pumped to state 7 where it
enters the steam generator
and returns to state 1 after
heating.

(y)
(1)
(1-y)

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
Applying steady-state mass and energy rate
balances to a control volume enclosing the
feedwater heater, the fraction of the total flow y is
(Eq. 8.12)
With fraction y known, mass and energy rate
balances applied to control volumes around the
other components yield the following expressions,
each on the basis of a unit of mass entering the
first turbine stage.

Regenerative Vapor Power Cycle Using


an Open Feedwater Heater
For the turbine stages

(Eq. 8.13)
For the pumps

(Eq. 8.14)
For the steam generator
(Eq. 8.15)

For the condenser


(Eq. 8.16)

Problem 8.40
A power plant operates on a regenerative vapor power cycle with one open feed water
heater. Steam enters the first turbine stage at 12 MPa, 520oC and expands to 1 MPa,
where some of the steam is extracted and diverted to the open feed water heater
operating at 1 MPa. The remaining steam expands through the second stage to the
condenser pressure of 6 kPa. Saturated liquid exits the open feed water heater at
1 MPa. For isentropic processes in the turbines and pumps, determine for the cycle:
a) the thermal efficiency
b) mass flow rate in the first
stage of the turbine in kg/h
for a net power output of
330 MW.

Work through Example 8.5

then

Work on homework
problem 8.49, 8.57, 8.74

Cogeneration Systems
Are integrated systems that simultaneously yield
two valuable products, electricity and steam (or hot
water) from a single fuel input.
Typically provide cost savings relative to producing
power and steam (or hot water) in separate systems.
Are widely deployed in industrial plants, refineries,
food processing plants, and other facilities requiring
process steam, hot water, and electricity.
Can be based on vapor power plants, gas turbine
power plants, internal combustion engines, and fuel
cells.

Cogeneration Systems
An application of cogeneration based on vapor
power plants is district heating providing steam or
hot water for space heating together with electricity
for domestic, commercial, and industrial use.
Electricity provided
to the community

Steam exported to
the community

Cogeneration Systems
Exporting useful steam to the community limits the electricity
that also can be provided from a given fuel input, however.
For instance, to produce saturated vapor at 100oC (1 atm) for
export to the community water circulating through the power
plant will condense at a higher temperature and thus at a
higher pressure.
In such an operating mode thermal efficiency is less than
when condensation occurs at a pressure below 1 atm, as in a
plant fully dedicated to power production.
T > 100oC
p > 1 atm

Work through Example 8.6

Lecture 26

Introduction to
Internal Combustion Engines
and the
Air-Standard Otto Cycle

Sections 9.1, 9.2

Area Interpretations for


Work and Heat Transfer
Ideal cycles formed from internally
reversible processes are used in Chapter 9 to
further understanding of reciprocating internal
combustion engines and gas turbine power
systems.
Closed systems involving expansion and
compression work are used to model
reciprocating engines. For these applications,
the following area interpretations apply for
internally reversible processes:

Area Interpretations for


Work and Heat Transfer
W
=
m int
rev

Q
= Tds
m int

pdv

rev

Observe that these expressions give work


and heat transfer per unit of mass contained
within the closed system.

Area Interpretations for


Work and Heat Transfer
One-inlet, one exit control volumes at steady state
are used to model gas turbine power plants. For these
applications, the following area interpretations apply for
internally reversible processes:
p

W
= vdp
m int

rev

Q
= Tds
m int

rev

Observe that these expressions give work and heat


transfer per unit of mass flowing through the control
volume.

Ideal Gas Model Review


Elementary thermodynamic analyses of reciprocating
internal combustion engines and gas turbines use ideal
model principles, as reviewed in Table 9.1.

Ideal Gas Model Review

Ideal Gas Model Review

Considering Reciprocating
Internal Combustion Engines
What are reciprocating internal combustion engines?
They are reciprocating engines commonly used in
automobiles, trucks, and buses.
How do reciprocating internal combustion engines
differ from the vapor power plants considered in
Chapter 8 and the gas turbines considered in later
sections of Chapter 9?
Processes occur within reciprocating pistoncylinder arrangements rather than by mass flowing
through a series of interconnected components.

Considering Reciprocating Internal


Combustion Engines Two Types
Spark-ignition
A mixture of fuel and air is ignited by a spark
plug.
This type is
advantageous for applications up to about
300 hp (225 kW).
lightweight and relatively low cost.
predominantly used by automobiles in the
U.S.

Considering Reciprocating Internal


Combustion Engines Two Types
Compression-ignition
Air is compressed to a high pressure and
temperature.
Combustion occurs spontaneously when fuel is
injected.
This type is
preferred for high-power applications and
when fuel economy is required.
used in heavy trucks and buses, locomotives
and ships, and auxiliary power units.

Introducing Engine Terminology


Displacement volume: volume
swept by piston when it moves from
top dead center to bottom dead
center
Top dead center
Stroke
Bottom dead center

Compression ratio, r : volume


at bottom dead center divided by
volume at top dead center

Introducing Engine Terminology


Four-stroke cycle
Four strokes of the piston
for every two revolutions of
the crankshaft
Intake stroke
With the intake valve open,
piston stroke draws a fresh
charge into the cylinder.
For spark-ignition
engines, the charge
includes fuel and air.
For compression-ignition
engines, the charge is air
alone.

Introducing Engine Terminology


Compression stroke
With both valves closed,
piston compresses charge,
raising the pressure and
temperature, and requiring
work input from the piston to
the cylinder contents.
For spark-ignition
engines, combustion is
initiated by the spark plug.
For compressionignition engines,
combustion is initiated
by injecting fuel into the hot
compressed air.

Introducing Engine Terminology


Power stroke
The gas mixture expands
and work is done on the
piston as it returns to bottom
dead center.

Exhaust stroke
The burned gases are
purged from the cylinder
through the open exhaust
valve.

Suck

Squeeze

Bang

Blow

Introducing Engine Terminology


Smaller engines operate on two-stroke cycles
with intake, compression, expansion, and exhaust
accomplished in one revolution of the crankshaft.
Internal combustion engines undergo
mechanical cycles, but the cylinder contents do
not execute a thermodynamic cycle matter is
introduced at one composition and is later
discharged at a different composition.

Introducing Engine Terminology


Mean effective pressure, mep, is an important
performance parameter.
mep is a theoretical constant pressure that if it
acted on the piston during the power stroke
would produce the same net work as actually
developed in one cycle.
(Eq. 9.1)

For two engines of equal displacement


volume, the one with a higher mep would
produce the greater net work, and if the engines
run at the same speed, greater power.

Introducing Engine Terminology


indicated mean effective pressure (imep) is a
calculated mean pressure from measurement of a
pressure trace from an engine.

Simulating Reciprocating
Internal Combustion Engines
Detailed study of performance of reciprocating
internal combustion engines requires consideration
of complexities including:

Combustion processes occurring within the cylinder.


The effects of irreversibilities related to combustion,
heat transfer, and friction.
Heat transfer between the gases in the cylinder and
the cylinder walls.
The work required to charge the cylinder and
exhaust the products of combustion.

Accurate analyses of reciprocating internal


combustion engines normally requires computer
simulation.

Air-Standard Analysis of Reciprocating


Internal Combustion Engines
To conduct elementary analyses of reciprocating internal
combustion engines, simplifications are required. Although
highly idealized, an air-standard analysis can provide insights
and qualitative information about actual performance.
An air-standard analysis has the following elements:

A fixed amount of air modeled as an ideal gas is the


working fluid. Ideal gas relations are reviewed in Table 9.1.
The combustion process is replaced by heat transfer from
an external source. Combustion is studied in Chapter 13.
There are no intake and exhaust processes. The cycle is
completed by a constant-volume heat transfer process while
the piston is at bottom dead center.
All processes are internally reversible.
In a cold air-standard analysis, the specific heats are
assumed constant at their ambient temperature values.

Air-Standard Analysis of Reciprocating


Internal Combustion Engines
For reciprocating internal combustion engines,
three cycles that adhere to air-standard cycle
idealizations are the Otto, Diesel, and Dual cycles.
These cycles differ only in the way the heat addition
process that replaces combustion in the actual cycle
is modeled:
Otto cycle: Heat addition at constant volume.
Diesel cycle: Heat addition at constant
pressure.
Dual cycle: Heat addition at constant volume
followed by heat addition at constant pressure.

Air-Standard Otto Cycle


The Otto cycle consists of four internally reversible
processes in series:
Process 1-2: isentropic compression.
Process 2-3: constant-volume heat addition to the air
from an external source.
Process 3-4: isentropic expansion.
Process 4-1: constant-volume heat transfer from the
air.

The Otto cycle


compression ratio is:
V1 V4
r=
=
V2 V3

Air-Standard Otto Cycle


Ignoring kinetic and potential energy effects,
closed system energy balances for the four
processes of the Otto cycle reduce to give
W34
W12
= u2 u1 ,
= u3 u 4
m
m
(Eq. 9.2)
Q23
Q41
= u3 u 2 ,
= u4 u1
m
m

The thermal efficiency is the ratio of the net


work to the heat added:
(Eq. 9.3)

Air-Standard Otto Cycle


Since the air-standard Otto cycle is composed of
internally reversible processes, areas on the T-s and
p-v diagrams can be interpreted as heat and work,
respectively:
On the T-s diagram, heat transfer per unit of
mass is Tds. Thus,
Area 2-3-a-b-2 represents
heat added per unit of mass.
Area 1-4-a-b-1 is the heat
rejected per unit of mass.
The enclosed area is the net
heat added, which equals the
net work output.

Air-Standard Otto Cycle


On the p-v diagram, work per unit of mass is
pdv. Thus,
Area 1-2-a-b-1 represents
work input per unit of mass
during the compression
process.
Area 3-4-b-a-3 is the work
done per unit of mass in the
expansion process.
The enclosed area is the net
work output, which equals the
net heat added.

Air-Standard Otto Cycle


The compression ratio, r = V1/V2, is an important
operating parameter for reciprocating internal combustion
engines as brought out by the following discussion
centering on the T-s diagram:
An increase in the compression ratio
changes the cycle from 1-2-3-4-1 to
1-2-3-4-1.
Since the average temperature of heat
addition is greater in cycle 1-2-3-4-1,
and both cycles have the same heat
rejection process, cycle 1-2-3-4-1 has
the greater thermal efficiency.
Accordingly, the Otto cycle thermal
efficiency increases as the
compression ratio increases.

Work through Example 9.1

Lecture 27

Air-Standard
Diesel and Dual
Cycles

Sections 9.3, 9.4

Air-Standard Diesel Cycle


The Diesel cycle consists of four internally
reversible processes in series:
Process 1-2: isentropic compression.
Process 2-3: constant-pressure heat addition to the
air from an external source.
Process 3-4: isentropic expansion.
Process 4-1: constant-volume heat transfer from
the air.

The Diesel cycle


has a two-step
power stroke:
process 2-3 followed
by process 3-4.

Air-Standard Diesel Cycle


V1
The Diesel cycle compression ratio is: r =
V2
V3
The Diesel cycle cut-off ratio is: rc =
V2

Air-Standard Diesel Cycle


Process 2-3 is heat addition at constant pressure.
Accordingly, the process involves both heat and work.
The work is given by

(Eq. 9.9)

Introducing Eq. 9.9 into the closed system energy balance


for process 2-3 and solving for Q23/m gives
(Eq. 9.10)
Note: Enthalpy appears only for notational convenience and
does not signal use of control volume concepts.

The thermal efficiency is the ratio of the net work to the


heat added:
(Eq. 9.11)
Like the Otto cycle, thermal efficiency increases with
increasing compression ratio.

Air-Standard Diesel Cycle


As for the Otto cycle, areas on the T-s and p-v
diagrams of the Diesel cycle can be interpreted as
heat and work, respectively:
On the T-s diagram, heat transfer per unit of
mass is Tds. Thus,
Area 2-3-a-b-2 represents
heat added per unit of mass.
Area 1-4-a-b-1 is the heat
rejected per unit of mass.
The enclosed area is the net
heat added, which equals the
net work output.

Air-Standard Diesel Cycle


On the p-v diagram, work per unit of mass is
pdv. Thus,
Area 1-2-a-b-1 represents work
input per unit of mass during the
compression process.
Area 2-3-4-b-a-2 is the work
done per unit of mass in the
two-step power stroke: process
2-3 followed by process 3-4.
The enclosed area is the net
work output, which equals the
net heat added.

Air-Standard Dual Cycle


By considering heat transfer to the air
undergoing the power cycle as occurring in two
steps: constant volume followed by constant
pressure, the air-standard Dual cycle aims to
mimic the pressure-volume variation of actual
internal combustion engines more closely than
achievable with the Otto and Diesel cycles.

Air-Standard Dual Cycle


The air-standard Dual cycle consists of five internally
reversible processes in series:
Process 1-2: isentropic compression.
Process 2-3: constant-volume heat addition to the air
from and external source.
Process 3-4: constant-pressure heat addition to the air
from an external source.
Process 4-5: isentropic expansion.
Process 5-1: constant-volume heat transfer from the air.
As for the Diesel
cycle, the Dual cycle
has a two-step
power stroke:
process 3-4 followed
by process 4-5.

Air-Standard Dual Cycle


Using closed system energy balances for each of
the processes, the following expression for thermal
efficiency for the air-standard Dual Cycle can be
developed:
(Eq. 9.14)

Note: As for the Diesel cycle, enthalpy appears only for


notational convenience and does not signal use of control
volume concepts.

Like the Otto and Diesel cycles, thermal efficiency


increases with increasing compression ratio.

Air-Standard Dual Cycle


The specific internal energies and temperatures
at each principal state are determined using
methods similar to those used for the Otto and
Diesel Cycles.
Areas on the T-s and p-v diagrams of the Dual
cycle can be interpreted as heat and work,
respectively, as in the cases of the Otto and Diesel
cycles.

Actual Reciprocating Internal


Combustion Engines
As implied by the discussion of the Otto, Diesel,
and Dual cycles, it is advantageous for actual
reciprocating internal combustion engines to have
high compression ratios.
However, since the temperature of the fuel-air
mixture being compressed in spark-ignition
engines also increases with compression ratio,
the possibility of autoignition or knock limits
the compression ratio of such engines to the
range 9.5-11.5, when fueled with unleaded
gasoline.

Actual Reciprocating Internal


Combustion Engines
Since only air is compressed in the cylinder,
compression-ignition engines do not experience
engine knock due to premature autoignition of fuel.
Accordingly, such engines can

operate at higher compression ratios than sparkignition engines.


use less refined fuels having higher ignition
temperatures than the volatile fuels required by sparkignition engines.

Work through Examples 9.2, 9.3

then

Work on homework problems


9.3, 9.20, 9.38

Lecture 28

Gas Turbine Power Plants


and
The Brayton Cycle

Sections 9.5, 9.6

Considering Gas Turbine Power Plants


Gas turbine power plants are more quickly
constructed, less costly, and more compact than the
vapor power plants considered in Chapter 8.
Gas turbines are suited for stationary power
generation as well as for powering vehicles,
including aircraft propulsion and marine power
plants.
Gas turbines are
increasingly used for large-scale power
generation, and
for such applications fueled primarily by
natural gas, which is relatively abundant today.

Considering Gas Turbine Power Plants


Gas turbines may operate on an open or closed basis, as
shown in the figures.
The open gas turbine is more commonly used and is the
main focus of our study of gas turbines.
Study of the individual components of these configurations
requires the control volume forms of the mass, energy, and
entropy balances.
Open to the atmosphere

Closed

Considering Gas Turbine Power Plants


The open mode gas turbine is an internal combustion
power plant.
Air is continuously drawn into
the compressor where it is
compressed to a high pressure.
Air then enters the combustion
chamber (combustor) where it
mixes with fuel and combustion
occurs.
Combustion products exit
at elevated temperature and
pressure.
Part of the
turbine work
Combustion products
is used to
expand through the turbine
drive the
and then are discharged to the compressor.
surroundings.

The remainder is
available as net work
output to drive an
electric generator, to
propel a vehicle, or
for other uses.

Considering Gas Turbine Power Plants


The closed gas turbine operates as follows:
A gas circulates through four components: turbine,
compressor, and two heat exchangers at higher and lower
operating temperatures, respectively.
The turbine and compressor play the same roles as in the
open gas turbine.
As the gas passes through the
higher-temperature heat
exchanger, it receives energy by
heat transfer from an external
source.
The thermodynamic cycle is
completed by heat transfer to the
surroundings as the gas passes
through the lower-temperature
heat exchanger.

Considering Gas Turbine Power Plants


The heat transfer associated with the highertemperature heat exchanger of the closed gas
turbine originates from an external source, which
may include
External combustion of
biomass, municipal solid
waste, fossil fuels such as
natural gas, and other
combustibles.
Waste heat from industrial
processes.
Solar thermal energy.
A gas-cooled nuclear
reactor.

Air-Standard Analysis of
Open Gas Turbine Power Plants
To conduct elementary analyses of open gas turbine power
plants, simplifications are required. Although highly idealized,
an air-standard analysis can provide insights and qualitative
information about actual performance.
An air-standard analysis has the following elements:

The working fluid is air which behaves as an ideal gas.


Ideal gas relations are reviewed in Table 9.1.
The temperature rise that would be brought about by
combustion is accomplished by heat transfer from an
external source.
With an air-standard analysis, we avoid the complexities of
the combustion process and the change in composition
during combustion, which simplifies the analysis
considerably. Combustion is studied in Chapter 13.
In a cold air-standard analysis, the specific heats are
assumed constant at their ambient temperature values.

Air-Standard Brayton Cycle


The schematic of a simple open air-standard gas turbine
power plant is shown in the figure.
The energy transfers by heat and work are in the
directions of the arrows.
Air circulates through the components:
At state 1, air is drawn into the
compressor from the surroundings.
Process 1-2: the air is
compressed from state 1 to
state 2.
Process 2-3: The
temperature rise that would be
achieved in the actual power
plant with combustion is
realized here by heat transfer, Q in .

Air-Standard Brayton Cycle


Process 3-4: The high-pressure, high-temperature air
expands through the turbine. The turbine drives the
compressor and develops net power, W cycle .
Air returns to the
surroundings at state 4 with a
temperature typically much
greater than at state 1.
After interacting with the
surroundings, each unit of mass
returns to the same condition as
the air entering at state 1,
thereby completing a
thermodynamic cycle.

Air-Standard Brayton Cycle


Process 3-4: The high-pressure, high-temperature air
expands through the turbine from state 3 to state 4. The
turbine drives the compressor and develops net power, W cycle .
Air returns to the
surroundings at state 4 with a
temperature typically much
greater than at state 1.
After interacting with the
surroundings, each unit of mass
returns to the same condition as
the air entering at state 1,
thereby completing a
thermodynamic cycle.
We imagine process 4-1 being
achieved by a heat exchanger, as
shown by the dashed line in the figure.

Air-Standard Brayton Cycle


Cycle 1-2-3-4-1 is called the Brayton cycle.
The compressor pressure ratio, p2/p1, is a key
Brayton cycle operating parameter.

Air-Standard Brayton Cycle


Analyzing each component as a control
volume at steady state, assuming the
compressor and turbine operate
adiabatically, and neglecting kinetic and
potential energy effects, we get the following
expressions for the principal work and heat
transfers, which are positive in accord with
our convention for cycle analysis.
Turbine
Heat addition
(Eq. 9.15)
Compressor

(Eq. 9.17)
Heat rejection

(Eq. 9.16)

(Eq. 9.18)

Air-Standard Brayton Cycle


The thermal efficiency is
(Eq. 9.19)
The back work ratio is
(Eq. 9.20)
Note: A relatively large portion of the work developed by the
turbine is required to drive the compressor. For gas turbines,
back work ratios range from 20% to 80% compared to only 1-2%
for vapor power plants.

Since Eqs. 9.15 through 9.20 have been developed from mass
and energy balances, they apply equally when irreversibilities
are present and in the absence of irreversibilities.

Ideal Air-Standard Brayton Cycle


The ideal air-standard Brayton cycle provides an
especially simple setting for study of gas turbine power
plant performance. The ideal cycle adheres to additional
modeling assumptions:
Frictional pressure drops are absent during flows through
the heat exchangers. These processes occur at constant
pressure. These processes are isobaric.
Flows through the turbine and pump occur adiabatically
and without irreversibility. These processes are isentropic.
Accordingly, the ideal Brayton cycle consists of two
isentropic processes alternated with two isobaric
processes. In this respect, the ideal Brayton cycle is
similar to the ideal Rankine cycle, which also consists of
two isentropic processes alternated with two isobaric
processes (Sec. 8.2.2).

Ideal Air-Standard Brayton Cycle

The ideal air-standard Brayton cycle consists of four


internally reversible processes:
Process1-2: Isentropic compression of air flowing through the
compressor.
Process 2-3: Heat transfer to the air as it flows at constant pressure
through the higher-temperature heat exchanger.
Process 3-4: Isentropic expansion of the air through the turbine.
Process 4-1: Heat transfer from the air as it flows at constant
pressure through the lower-temperature heat exchanger.

Ideal Air-Standard Brayton Cycle


Since the ideal Brayton cycle involves internally
reversible processes, results from Sec. 6.13 apply.
On the p-v diagram, the work per unit of mass
flowing is vdp. Thus on a per unit of mass flowing
basis,
Area 1-2-a-b-1
represents the
compressor work input.
Area 3-4-b-a-3
represents the turbine
work output.
Enclosed area 1-2-3-4-1
represents the net work
developed.

integrate

Ideal Air-Standard Brayton Cycle


On the T-s diagram, the
heat transfer per unit of
mass flowing is Tds. Thus,
on a per unit of mass
flowing basis,
Area 2-3-a-b-2 represents the
heat added.
Area 4-1-b-a-4 represents the
heat rejected.
Enclosed area 1-2-3-4-1
represents the net heat added or
equivalently, the net work
developed.

integrate

Effects of Compressor Pressure Ratio on


Brayton Cycle Performance
That the compressor pressure ratio, p2/p1, is an
important operating parameter for gas turbines is
brought out simply by the following discussions
centering on the T-s diagram:

Effects of Compressor Pressure Ratio on


Brayton Cycle Performance
Increasing the compressor pressure ratio from p2/p1 to
p2/p1 changes the cycle from 1-2-3-4-1 to 1-2-3-4-1.
Since the average temperature of heat
addition is greater in cycle 1-2-3-4-1, and
both cycles have the same heat rejection
process, cycle 1-2-3-4-1 has the greater
thermal efficiency.

Effects of Compressor Pressure Ratio on


Brayton Cycle Performance
Increasing the compressor pressure ratio from p2/p1 to
p2/p1 changes the cycle from 1-2-3-4-1 to 1-2-3-4-1.

60
th (%)

Since the average temperature of heat


addition is greater in cycle 1-2-3-4-1, and
both cycles have the same heat rejection
process, cycle 1-2-3-4-1 has the greater
thermal efficiency.
Accordingly, the Brayton cycle thermal
efficiency increases as the compressor
pressure ratio increases.

2 4 6 8 10
Compressor
Pressure Ratio

Effects of Compressor Pressure Ratio on


Brayton Cycle Performance
Increasing the compressor pressure ratio from p2/p1 to
p2/p1 changes the cycle from 1-2-3-4-1 to 1-2-3-4-1.
Since the average temperature of heat
addition is greater in cycle 1-2-3-4-1, and
both cycles have the same heat rejection
process, cycle 1-2-3-4-1 has the greater
thermal efficiency.
Accordingly, the Brayton cycle thermal
efficiency increases as the compressor
pressure ratio increases.
The turbine inlet temperature also
increases with increasing compressor
ratio from T3 to T3.

Effects of Compressor Pressure Ratio on


Brayton Cycle Performance
However, there is a limit on the maximum
temperature at the turbine inlet imposed by
metallurgical considerations of the turbine blades.
Lets consider the effect of increasing compressor
pressure ratio on Brayton cycle performance when
the turbine inlet temperature is held constant.
This is investigated using the T-s diagram as
presented next.

Effects of Compressor Pressure Ratio on


Brayton Cycle Performance
The figure shows the T-s diagrams of two ideal
Brayton cycles having the same turbine inlet temperature
but different compressor pressure ratios.
Cycle A has the greater
compressor pressure ratio and
thus the greater thermal efficiency.
Cycle B has the larger enclosed
area and thus the greater net work
developed per unit of mass flow.
For Cycle A to develop the same
net power as Cycle B, a larger
mass flow rate would be required
and this might dictate a larger
system.

Effects of Compressor Pressure Ratio on


Brayton Cycle Performance
Accordingly, for turbine-powered vehicles, where
size and weight are constrained, it may be
desirable to operate near the compressor pressure
ratio for greater net work per unit of mass flow and
not the pressure ratio for greater thermal efficiency.

Gas Turbine Power Plant Irreversibility


The most significant irreversibility by far is the
irreversibility of combustion. This type of irreversibility is
considered in Chap. 13, where combustion fundamentals
are developed.
Irreversibilities related to flow through the turbine and
compressor also significantly impact gas turbine
performance. They act to
decrease the work developed by the turbine and
increase the work required by the compressor,
thereby decreasing the net work of the power plant.

W net W t W c
=

m
m
m
marked decrease in net
work of the power plant

irreversiblities increase
compressor work
irreversibilites decrease
turbine work

Gas Turbine Power Plant Irreversibility


Isentropic turbine efficiency, introduced in Sec. 6.12.1,
accounts for the effects of irreversibilities within the turbine in
terms of actual and isentropic turbine work, each per unit of
mass flowing through the turbine.
work developed in the actual
expansion from turbine inlet state
to the turbine exit pressure

(W t / m )
(h3 h4 )
=
t =
(Wt / m ) s (h3 h4s )
work developed in an isentropic
expansion from turbine inlet
state to exit pressure

Gas Turbine Power Plant Irreversibility


Isentropic compressor efficiency, introduced in Sec.
6.12.3, accounts for the effects of irreversibilities within the
compressor in terms of actual and isentropic compressor work
input, each per unit of mass flowing through the compressor.

work input for an isentropic process from


compressor inlet state to exit pressure

(W c / m ) s (h2s h1 )
=
c =
(h2 h1 )
(Wc / m )
work input for the actual process from compressor
inlet state to the compressor exit pressure

Work through Examples 9.4, 9.5, 9.6

then

Work on homework problem


9.54

Lecture 29

Regeneration, Reheat
and Intercooling in
Gas Turbines

Sections 9.7, 9.8

Regenerative Gas Turbines


The hot turbine exhaust can be utilized with a preheater
called a regenerator.
The regenerator allows air exiting the
compressor to be preheated, process 2-x,
as the turbine exhaust gas cools, process
4-y.
Preheating reduces the heat added per
unit of mass flowing (and thus the amount
of fuel that must be burned):
With Regeneration Without Regeneration
Q in
= (h3 hx )
m

Q in
= (h3 h2 )
m

The net work per unit of mass flowing is not altered with the
inclusion of a regenerator. Accordingly, since the heat added is
reduced, thermal efficiency increases.

Regenerator Effectiveness
Since a finite temperature difference must exist
between the two streams of the regenerator for heat
transfer to take place between the streams, the coldside exiting temperature, Tx, must be less than the
hot-side entering temperature, T4.
As the stream-to-stream
temperature difference becomes
small Tx approaches T4, but
cannot exceed it. Accordingly,
T x T 4.
As the enthalpy of the air
varies only with temperature, we
also have hx h4.

T4

Regenerator Effectiveness
The regenerator effectiveness is defined as
the ratio of the actual enthalpy increase of the air
flowing through the cold side of the regenerator,
hx h2, to the maximum theoretical enthalpy
increase, h4 h2.
(Eq. 9.27)

Regenerator Effectiveness
In practice, regenerator effectiveness values
range from 60-80%, approximately. So, the
temperature Tx at the combustor inlet is invariably
below the temperature T4 at the turbine exit.
Selection of a regenerator is largely an
economic decision.
With regeneration less fuel is consumed by the
combustor but another component, the
regenerator, is required.
When considering use of a regenerator, the
trade-off between fuel savings and regenerator
cost must be weighed.

Gas Turbines with Reheat and Regeneration


A modification of the Brayton cycle that increases
the net work developed is multistage expansion
with reheat.
The figure shows a cycle with two turbine stages
and a reheat combustor between the stages.

Gas Turbines with Reheat and Regeneration


The ideal Brayton cycle with reheat is 1-2-3-a-b-4-1.
The ideal Brayton cycle without reheat is 1-2-3-4-1.
The reheat cycle has a larger enclosed area than
the cycle without reheat and thus a greater net work
developed per unit of mass flowing, which is the aim.
Cycle without reheat

Cycle with reheat

Gas Turbines with Reheat and Regeneration


The figure also shows that the temperature at the exit
of the second-stage turbine, state 4, is greater than at
the exit of the single turbine of the cycle without reheat,
state 4. Accordingly, with reheat the potential for
regeneration is also enhanced.
When reheat and regeneration are used together, the
thermal efficiency can increase significantly over that for
the cycle without reheat.

T4
T4

Gas Turbines with


Intercooling and Regeneration
Another modification of the Brayton cycle that
increases the net work developed is compression
with intercooling.
The figure shows two compressor stages and an
intercooler between the stages.

Gas Turbines with


Intercooling and Regeneration

The accompanying p-v diagram


shows the processes for internally
reversible operation:

Process 1-c. Isentropic


compression from state 1, where
pressure is p1, to state c, where
pressure is pi.
Process c-d. Constant-pressure
cooling from temperature Tc to
temperature Td.
Process d-2. Isentropic
compression to state 2, where
pressure is p2.

Isentropic compression without intercooling is


represented by process 1-c-2.

Gas Turbines with


Intercooling and Regeneration
Recalling that for such internally reversible processes the
work input per unit of mass flowing is given by vdp, the
following area interpretations apply, each per unit of mass
flowing:
With intercooling, area 1-c-d-2-a-b-1
represents the work input.
Without intercooling, area 1-2-a-b-1
represents the work input.
The cross-hatched area c-d-2-2-c
represents the reduction in work
achieved with intercooling.

If the total turbine work remains the same, a reduction in


compressor work results in an increase in the net work
developed, which is the aim.

Gas Turbines with


Intercooling and Regeneration
While compression with and without intercooling each
bring the air to the same final pressure, p2, the final
temperature with intercooling, T2, is lower than the final
temperature without intercooling, T2.
Comparing states 2 and 2 on the T-s diagram, T2 < T2.
The lower temperature at the compressor exit with
intercooling enhances the potential for regeneration.

T2
T2

Gas Turbines with


Intercooling and Regeneration
When compression with intercooling is used together with
regeneration, the thermal efficiency can increase significantly
over that for the cycle without intercooling.
The T-s diagram also shows that for cooling to the
surroundings the temperature Td at the intercooler exit
cannot be less than T1, the temperature of the air entering
the compressor from the surroundings: Td T1.

Td
T1

Regenerative Gas Turbine


with Reheat and Intercooling
Shown here is a regenerative gas turbine that
incorporates reheat and intercooling.
With these modifications to the basic Brayton cycle:
The net work
output is
increased.
The thermal
efficiency is
increased.

Regenerative Gas Turbine


with Reheat and Intercooling
Applying mass and energy rate
balances at steady state, we
obtain the following expressions,
each per unit of mass flowing:
Total turbine work:
W t
= (h6 h7) + (h8 h9) = t1(h6 h7s) + t2(h8 h9s)

m
where t1 and t2 denote the isentropic efficiencies of turbines 1 and 2,
respectively.

Total compressor work:


W c
= (h2 h1) + (h4 h3) = (h2s h1)/c1 + (h4s h3)/c2
m
where c1 and c2 denote the isentropic efficiencies of compressors 1
and 2, respectively.

Regenerative Gas Turbine


with Reheat and Intercooling
Applying mass and energy rate
balances at steady state, we
obtain the following expressions,
each per unit of mass flowing:
Total heat added:
Q in
= (h6 h5) + (h8 h7)

In this application, the regenerator effectiveness is:

reg = (h5 h4)/(h9 h4)


For cooling to the surroundings, the temperature at the
exit of the intercooler, T3, cannot be less than the
temperature of the air entering the compressor from the
surroundings: T3 T1.

Theoretical Cycle Limits: Ericcson Cycle


The Ericcson
cycle is the limit of:
adding or rejecting
heat at constant
pressure,
all heat rejected during intercooling goes to reheat stages,
infinite reheat and intercooling steps so that the cycle is
compressing or expanding at constant temperature, TC
and TH, respectively.
The efficiency of the Ericcson Cycle is the same as the
Carnot cycle (or any reversible cycle with heat rejection at
TC and heat addition at TH):

Theoretical Cycle Limits: Sterling Cycle


The Sterling cycle
has the same model
(infinite intercooling
and reheat cycles),
except that adding
and rejecting heat is
done at constant
volume.
The efficiency of the Sterling Cycle is the same as the
Carnot and Ericcson cycles:

Ideal Gas (Reversible) Power Cycles


control volume

Carnot
2 isothermal heat transfer
2 adiabatic (isentropic) work
transfer
Ericcson
2 isobaric heat transfer
2 isothermal work transfer
Sterling
2 constant volume
(isochoric) heat transfer
2 isothermal work transfer
Brayton

2 isobaric heat transfer


2 adiabatic (isentropic) work
transfer

closed system
Otto
2 isochoric heat transfer
2 adiabatic (isentropic) work
transfer

Diesel
1 isobaric, 1 isochoric heat
transfer
2 isothermal work transfer
Duel
1 isochoric+isobaric,
1 isochoric heat transfer
2 isothermal work transfer
Rankine [Vapor]
2 isobaric heat transfer
2 adiabatic (isentropic) work
transfer

Lecture 30

Gas Turbines for


Aircraft Propulsion

Sections 9.11, 9.12.1, 9.12.2

Gas Turbines for Aircraft Propulsion


Because of their favorable power-to-weight ratio, gas
turbines are well suited for aircraft propulsion. The
turbojet engine is commonly used for this purpose.
The figure provides the schematic of a turbojet engine.

Va

V5

Momentum Equation for


Steady One-Dimensional Flow
Recall Newtons law of motion for a closed system: F = ma

where F is the resultant force acting on a system of mass m


and a is the acceleration.
We also need a form of Newtons second law of motion
appropriate for the study of control volumes.
For a control volume such as the one-inlet, one-exit control
volume shown here, momentum is carried in and out at the inlets
and exits, respectively, according to

Rate of
momentum
transfer in

Rate of
momentum
transfer out

(Eq. 9.30)

Momentum Equation for


Steady One-Dimensional Flow
In words, Newtons second law for a control volume is:
(= 0 at steady state)
time rate of change
resultant force
of momentum contained = acting on the +
within the control volume
control volume

The form of the momentum


equation for a one-inlet, one-exit
control volume at steady state is:

(Eq. 9.31)

net rate at which momentum is


transferred into the control
volume accompanying mass flow

Gas Turbines for Aircraft Propulsion


The increase in velocity from diffuser inlet, Va, to nozzle exit,
V5, gives rise to the thrust developed by the engine in accord
with Newtons second law of motion:
(Eq. 9.31)

In harmony with air-standard analysis, we assume air


modeled as an ideal gas flows through the engine shown in
the schematic and the temperature rise that would be obtained
with combustion is achieved by heat transfer from an external
source.
Va

V5

Gas Turbines for Aircraft Propulsion


If the air flows through the components of the turbojet engine
without irreversibilities and stray heat transfer, air undergoes
the five processes shown on the T-s diagram:
Process a-1: Air at velocity Va enters the diffuser and
decelerates isentropically, while experiencing an increase in
pressure.
Process 1-2: The air experiences a further increase in
pressure isentropically, owing to work done by the compressor.

Va

V5

Gas Turbines for Aircraft Propulsion


If the air flows through the components of the turbojet engine
without irreversibilities and stray heat transfer, air undergoes
the five processes shown on the T-s diagram:
Process 2-3: The temperature of the air increases at constant
pressure as it receives a heat transfer from an external source.
Process 3-4: The high-pressure, high-temperature air
expands isentropically through the turbine, driving the
compressor.

Va

V5

Gas Turbines for Aircraft Propulsion


If the air flows through the components of the turbojet engine
without irreversibilities and stray heat transfer, air undergoes
the five processes shown on the T-s diagram:
Process 4-5: The air continues to expand isentropically
through the nozzle, achieving a velocity, V5, at the engine exit
much greater than the velocity, Va, at the engine inlet, and
thereby developing thrust.

Va

V5

Review: Nozzle and Diffuser Modeling


The one-inlet, one-exit energy rate balance at
steady state reads:
2
2

(V

V
)
i
e
0 = Q cv W cv + m (hi he ) +
+ g ( zi z e )
2

For a control volume enclosing a nozzle or diffuser,


Wcv = 0.
If the change in potential energy from inlet to exit is
negligible, g(zi ze) drops out.
If the heat transfer with surroundings is negligible,
Q cv drops out.
Vi2 Ve2

0 = (hi he ) +

Gas Turbines for Aircraft Propulsion


The energy rate balance applicable to the diffuser
takes the form
2
2
0 = (hi he ) +

Vi Ve

For the diffuser, i = a and e = 1. Then,


ha
Va

h1
V1 0

Va2 V12
0 = (ha h1 ) +

Since exit velocity is negligible, the energy rate


balance reduces to
Va2
h1 = ha +
2

Gas Turbines for Aircraft Propulsion


The energy rate balance applicable to the nozzle
takes the form
V2 V2
0 = (hi he ) +

For the nozzle, i = 4 and e = 5. Then,


h5
V5

h4
V4 0

V42 V52

0 = (h4 h5 ) +

5
4

Since inlet velocity is negligible, the energy rate


balance reduces to
2

V5
h4 = h5 +
V5 = 2(h4 h5 )
2

Gas Turbines for Aircraft Propulsion


Since the final expressions obtained for the
diffuser and nozzle are deduced from mass and
energy rate balances, they apply equally when
irreversibilities are present and in the absence of
irreversibilities.

Considering Compressible Flow


In many applications of engineering interest, gases
move at relatively high speeds and exhibit significant
changes in specific volume (density). They include
Flows through the nozzles and diffusers of jet
engines.
Flows through wind tunnels, shock tubes, and steam
ejectors.

These flows are known as compressible flows.

Velocity of Sound
A sound wave is a small pressure disturbance that
propagates through a gas, liquid, or solid at a velocity
c that depends on the properties of the medium.
Analyses using mass and momentum equations
supported by experimental data reveal that the
relation between pressure and specific volume
across a sound wave is nearly isentropic, and that its
velocity c called the velocity of sound is given by

(Eq. 9.36a)

(Eq. 9.36b)

Velocity of Sound
The special case of an ideal gas with constant
specific heats is used extensively in Chapter 9. For
this case, the relationship between pressure and
specific volume for fixed entropy is pvk = constant
where k is the specific heat ratio. Using this
relationship, Eq. 9.36b becomes
(Eq. 9.37)

The speed of sound is an intensive property


whose value depends on the state of the medium
through which sound propagates. While sound
propagates nearly isentropically, the medium itself
may be undergoing any process.

Mach Number
In subsequent discussions, the ratio of velocity V at
a state in a flowing fluid to the value of sonic velocity c
at the same state plays an important role. This ratio is
called the Mach number, M.
(Eq. 9.38)

Several important
terms associated with
Mach number are
shown in the table.

Mac h Number

T erm

M<1

S ubs onic

M=1

S onic

M>1

S upers onic

M >> 1

H ypers onic

M near 1

T rans onic

Work through Example 9.13

then

Work on homework problems


9.87, 9.90

Lecture 31

Vapor and Vapor-Compression


Refrigeration Cycles

Sections 10.1, 10.2

Carnot Refrigeration Cycle


Turbine and compressor
dont like operating
inside the vapor dome

Carnot Coefficient of Performance

Vapor-Compression
Refrigeration Cycle
Most common refrigeration cycle in use today
There are four principal
control volumes involving
these components:
Evaporator
Compressor
Condenser
Two-phase
Expansion valve
liquid-vapor mixture
All energy transfers by work and heat are taken as positive in
the directions of the arrows on the schematic and energy
balances are written accordingly.

The Vapor-Compression
Refrigeration Cycle
The processes of this cycle are
Process 4-1: two-phase liquid-vapor
mixture of refrigerant is evaporated
through heat transfer from the
refrigerated space.
Process 1-2: vapor refrigerant is
compressed to a relatively high
temperature and pressure requiring
work input.
Two-phase
liquid-vapor mixture
Process 2-3: vapor refrigerant
condenses to liquid through heat
transfer to the cooler surroundings.
Process 3-4: liquid refrigerant
expands to the evaporator pressure.

The Vapor-Compression
Refrigeration Cycle
Engineering model:
Each component is analyzed as a control
volume at steady state.
Dry compression is presumed: the
refrigerant is a vapor.
The compressor operates adiabatically.
The refrigerant expanding through the
valve undergoes a throttling process.
Kinetic and potential energy changes are
ignored.

The Vapor-Compression
Refrigeration Cycle
Applying mass and energy rate balances
Evaporator

Q in
= h1 h4
m

(Eq. 10.3)

The term Q in is referred to as the


refrigeration capacity, expressed in kW
in the SI unit system or Btu/h in the
English unit system.
A common alternate unit is the ton of
refrigeration which equals 200 Btu/min
or about 211 kJ/min.

The Vapor-Compression
Refrigeration Cycle
Applying mass and energy rate balances
Compressor
Assuming adiabatic
compression
Condenser
Expansion valve
Assuming a throttling
process

Wc
= h2 h1
m

(Eq. 10.4)

Q out
= h2 h3 (Eq. 10.5)
m

h4 = h3

(Eq. 10.6)

The Vapor-Compression
Refrigeration Cycle
Performance parameters
Coefficient of Performance (COP)
(Eq. 10.7)
Carnot Coefficient of Performance
(Eq. 10.1)
This equation represents the maximum theoretical
coefficient of performance of any refrigeration cycle
operating between cold and hot regions at TC and TH,
respectively.

Features of
Actual Vapor-Compression Cycle
Heat transfers between refrigerant and cold and
warm regions are not reversible.
Refrigerant temperature
in evaporator is less than
TC.
Refrigerant temperature
in condenser is greater
than TH.
Irreversible heat
transfers have negative
effect on performance.

Features of
Actual Vapor-Compression Cycle
The COP decreases primarily due to increasing
compressor work input as the
temperature of the
Trefrigerant
refrigerant passing
through the evaporator is
reduced relative to the
temperature of the cold
region, TC.
temperature of the
Trefrigerant
refrigerant passing
through the condenser is increased relative to the
temperature of the warm region, TH.

Features of
Actual Vapor-Compression Cycle
Irreversibilities during the compression process are
suggested by dashed line from state 1 to state 2.
An increase in specific
entropy accompanies an
adiabatic irreversible
compression process. The
work input for compression
process 1-2 is greater than
for the counterpart isentropic
compression process 1-2s.
Since process 4-1, and thus the refrigeration capacity,
is the same for cycles 1-2-3-4-1 and 1-2s-3-4-1, cycle
1-2-3-4-1 has the lower COP.

Isentropic Compressor Efficiency


The isentropic compressor efficiency is the ratio of
the minimum theoretical work input to the actual
work input, each per unit of mass flowing:
work required in an isentropic
compression from compressor inlet
state to the exit pressure

(Eq. 6.48)

work required in an actual


compression from compressor
inlet state to exit pressure

Actual Vapor-Compression Cycle


Example: The table provides steady-state operating
data for a vapor-compression refrigeration cycle
using R-134a as the working fluid. For a refrigerant
mass flow rate of 0.08 kg/s, determine the
(a) compressor power, in kW,
(b) refrigeration capacity, in tons,
(c) coefficient of performance,
(d) isentropic compressor efficiency.
State

2s

h (kJ/kg) 241.35 272.39 280.15 91.49 91.49

Actual Vapor-Compression Cycle


State

2s

h (kJ/kg) 241.35 272.39 280.15 91.49 91.49

(a) The compressor power is

W c = m (h2 h1 )
kg
kJ 1 kW

Wc = 0.08 (280.15 241.35)


= 3.1 kW
s
kg 1 kJ/s

(b) The refrigeration capacity is

Q in = m (h1 h4 )
kg
kJ
1 ton
60 s

= 3.41 tons
Qin = 0.08 (241.35 91.49)
s
kg 211 kJ/min min

Actual Vapor-Compression Cycle


State

2s

h (kJ/kg) 241.35 272.39 280.15 91.49 91.49

(c) The coefficient of performance is

(h1 h4 )
=
(h2 h1 )

(241.35 91.49)kJ/kg
= 3.86
(280.15 241.35)kJ/kg

Actual Vapor-Compression Cycle


State

2s

h (kJ/kg) 241.35 272.39 280.15 91.49 91.49

(d) The isentropic compressor


efficiency is

(
W c / m )s (h2 s h1 )
=
=

c =

W c / m

(h2 h1 )

(272.39 241.35)kJ/kg
= 0.8 = 80%
(280.15 241.35)kJ/kg

p-h Diagram
The pressure-enthalpy (p-h) diagram is a
thermodynamic property diagram commonly used
in the refrigeration field.

Extended p-h Diagrams

T (temperature) constant
s (entropy) constant
X (quality in vapor dome) constant

Extended p-h Diagrams

(water)

Work through Examples


10.1, 10.2, 10.3a-c

then

Work on homework problems


10.2, 10.16

Lecture 32

Vapor-Compression Applications,
Absorption Refrigeration
and Refrigerant Choice

Sections 10.3, 10.4, 10.5

Selecting Refrigerants
Refrigerant selection is based on several factors:

Q in
= h1 h4
m
http://en.wikipedia.org/wiki/List_of_refrigerants

refrigeration
capacity

Refrigerant Types and Characteristics


Chlorofluorocarbons (CFCs) and Hydrochlorofluorocarbons
(HCFCs) are early synthetic refrigerants each containing chlorine.
Because of the adverse effect of chlorine on Earths stratospheric
ozone layer, use of these refrigerants is regulated by international
agreement.
Hydrofluorocarbons (HFCs) and HFC blends are chlorine-free
refrigerants. Blends combine two or more HFCs. While these
chlorine-free refrigerants do not contribute to ozone depletion, with
the exception of R-1234yf, they have high GWP levels.
Natural refrigerants are nonsynthetic, naturally occurring
substances which serve as refrigerants. These include carbon
dioxide, ammonia, and hydrocarbons. These refrigerants feature
low GWP values; still, concerns have been raised over the toxicity
of NH3 and the safety of the hydrocarbons.

Refrigerant Types and Characteristics

Global Warming Potential (GWP) is a simplified index that estimates the potential
future influence on global warming associated with different gases when released
to the atmosphere.

Read journal article:


Experimental Comparison of the
Performance of Refrigerants R134a and
R32/R134a in Dry Expansion and Liquid
Recirculation Refrigerating Systems

(on D2L with this lecture)

Cascade Cycles

flash chamber

Multistage Compression
with Intercooling

saturated liquid

saturated vapor

Ammonia-Water Absorption Refrigeration


Absorption
refrigeration systems
have important
commercial and
industrial applications.
The principal
components of an
ammonia-water
absorption system are
shown in the figure.

Absorber
coolant

Ammonia-Water Absorption Refrigeration


The left-side of the
schematic includes
components familiar
from the discussion of
the vapor-compression
system: evaporator,
condenser, and
expansion valve.
Only ammonia flows
through these
components.

Absorber
coolant

Ammonia-Water Absorption Refrigeration


The right-side of the
schematic includes
components that
replace the compressor
of the vaporcompression
refrigeration system:
absorber, pump, and
generator. These
components involve
liquid ammonia-water
solutions.

Absorber
coolant

Ammonia-Water Absorption Refrigeration


A principal
advantage of the
absorption system is
that for comparable
refrigeration duty the
pump work input
required is intrinsically
much less than for the
compressor of a
vapor-compression
system.

Absorber
coolant

Ammonia-Water Absorption Refrigeration


Specifically, in the
absorption system ammonia
vapor coming from the
evaporator is absorbed in
liquid water to form a liquid
ammonia-water solution.
The liquid solution is then
pumped to the higher
operating pressure. For the
same pressure range,
significantly less work is
required to pump a liquid
solution than to compress a
vapor (see discussion of Eq.
6.51b).

Absorber
coolant

Ammonia-Water Absorption Refrigeration


However, since only
ammonia vapor is
allowed to enter the
condenser, a means must
be provided to retrieve
ammonia vapor from the
liquid solution.
This is accomplished
by the generator using
heat transfer from a
relatively hightemperature source.

Absorber
coolant

Ammonia-Water Absorption Refrigeration


Steam or waste heat
that otherwise might go
unused can be a costeffective choice for the
heat transfer to the
generator.
Alternatively, the heat
transfer can be provided
by solar thermal energy,
burning natural gas or
other combustibles, and
in other ways.

Absorber
coolant

Lecture 33

Properties of
Mixtures of Ideal Gases

Sections 12.1 12.3

Describing Mixture Composition

Consider a system consisting of a


number of gases within a container of
volume V. The temperature and
pressure of the gas mixture are T and
Gas 1: n1, m1
p, respectively.
Gas 2: n2, m2
The composition of the mixture can
Gas j: nj, mj
be described by giving the mass mi or
Sum: n m
the number of moles ni for each
component present.
The mass mi, number of moles ni, and molecular weight Mi
of component i are related by
ni is in kmol when mi is in
mi
kg and Mi is in kg/kmol.
ni =
(Eq. 12.1)
ni is in lbmol when mi is in
Mi
lb and Mi is in lb/lbmol.

Describing Mixture Composition


The mass fraction is the relative amount of each
component in the mixture. The mass fraction mfi of
component i is

mi
mf i =
m

(Eq. 12.3)

where m is the total mass of mixture.


The sum of the mass fractions of all components
in a mixture equals unity.

1 = mf i
i

Describing Mixture Composition


Alternatively, the mole fraction can
be used to describe the relative amount
of each component in the mixture. The
mole fraction yi of component i is

ni
yi =
n
(Eq. 12.6)

where n is the total moles of mixture.


The sum of the mole fractions of all
components equals unity.
The apparent (or average) molecular
weight M of a mixture is determined by
the mole-fraction average of the
component molecular weights:

1 = yi
i

M = yi M i
i =1

(Eq. 12.9)

Describing Mixture Composition


Example: The molar analysis of a gas mixture is
50% N2, 35% CO2, and 15% O2. Determine
(a) the apparent molecular weight of the mixture and
(b) the analysis in terms of mass fractions.

Solution:
(a) The apparent molecular weight of the mixture is
found using molecular weights (rounded) from Table A-1
j

M = yi M i
i =1

M = 0.50(28) + 0.35(44) + 0.15(32) = 34.2 kg/kmol

Describing Mixture Composition


(b) The actual amount of mixture is not known, the calculations
can be based on any convenient amount. We use 1 kmol of
mixture.
Then, the amount ni of each component, in kmol, is equal to its
mole fraction, as shown in column (ii).
Column (iii) lists the respective molecular weights.
Column (iv) gives the mass mi of each component, in kg per
kmole of mixture, obtained using mi = niMi (Eq. 12.1).
The mass fractions, listed as percentages in column (v), are
obtained by dividing the values in column (iv) by the column total
and multiplying by 100
i.e. (100%)mi/m = mfi (Eq. 12.3).
(i)
Component

(ii)
ni

(iii)
Mi

(iv)
mi

(v)
mfi %

N2

0.50

28

14

40.94

CO2

0.35

44

15.4

45.03

O2

0.15

32

4.8

14.04

34.2

100

1.00

Engineering Applications of Ideal Gas


Mixtures
We encounter ideal gas mixtures in many
important areas of application. Two of these are:
1. Systems involving chemical reactions and, in
particular, combustion. For these applications we
typically work on a molar basis. Combustion
systems are considered in Chapter 13.
2. Systems for air-conditioning and other
applications requiring close control of water vapor in
gas mixtures. For these applications we typically
work on a mass basis. Systems of this type are
considered in the second part of Chapter 12.

Relating p, V, and T for Ideal Gas Mixtures

Gas 1: n1, m1
Gas 2: n2, m2

Many systems of practical


interest involve mixtures where the
overall mixture and each of its
components can be modeled as
ideal gases. For such mixtures the
Dalton mixture model is
commonly used.

Gas j: nj, mj
Sum:

n m

The overall mixture is considered an ideal gas

nR T
p=
V

(Eq. 12.10)

The Dalton model also assumes each component


behaves as an ideal gas as if it were alone at
temperature T and volume V.

Relating p, V, and T for Ideal Gas Mixtures

ni R T
pi =
V

Gas 1: n1, m1
Gas 2: n2, m2

Then, with the Dalton model the


individual components do not exert the
mixture pressure p but rather a partial
pressure denoted by pi:

(Eq. 12.11)

Gas j: nj, mj
Sum:

n m

By combining Eqs. 12.10 and 12.11 the partial pressure


pi can be determined alternatively from

pi = yi p

(Eq. 12.12)

where the sum of the partial pressures equals the mixture


pressure
j

p = pi
i =1

(Eq. 12.13)

Evaluating U, H, and S for Ideal Gas Mixtures


(Molar Basis)
Then, when working on a molar basis expressions for U, H,
and S of a mixture consisting of several components are:

Evaluating U, H, and S for Ideal Gas Mixtures


(Molar Basis)
The mixture specific heats c v and c p are molefraction averages of the respective component
specific heats.

(Eq. 12.23)

(Eq. 12.24)

See Sec. 12.4 for applications using these


expressions for U, H, S, and the specific heats.

Evaluating U, H, and S for Ideal Gas Mixtures


(Mass Basis)
When working on a mass basis the expressions for U, H, S,
and specific heats of a mixture consisting of two components
a binary mixture are:
Table 12.2

Work through Examples


12.1 and 12.2

then

Work on homework problem 12.2

Lecture 34

Mixtures of Ideal Gases


and Psychrometric Applications

Sections 12.4, 12.5

Psychrometric Applications
The remainder of this chapter centers on systems
involving moist air. A condensed water phase may
also be present in such systems.
The term moist air refers to a mixture of dry air
and water vapor in which the dry air is treated as a
pure component.
The Dalton model applies to moist air.
By identifying gas 1 with dry air and gas 2
with water vapor, Table 12.2 gives moist air
property relations on a mass basis.
The study of systems involving moist air is known
as psychrometrics.

Table 12.2

Moist Air
Consider a closed system
consisting of moist air occupying
a volume V at mixture pressure p
and mixture temperature T.
In moist air the amount of
water vapor present is much
less than the amount of dry air:
mv << ma

nv << na.

The Dalton model applies to the mixture of dry air


and water vapor:

Moist Air
1. The overall mixture and each component, dry air
and water vapor, obey the ideal gas equation of state.
2. Dry air and water vapor within the mixture are
considered as if they each exist alone in volume V at
the mixture temperature T while each exerts part of
the mixture pressure.
3. The partial pressures pa and pv of dry air and
water vapor are, respectively
pa = ya p

pv = yv p

(Eq. 12.41b)

where ya and yv are the mole fractions of the dry air and
water vapor, respectively. These moist air expressions
conform to Eqs. (c) of Table 12.2.

Moist Air
4. The mixture pressure is the sum of the partial
pressures of the dry air and the water vapor:
p = pa + pv
Mixture pressure, p

5. A typical state of water


vapor in moist air is fixed
using partial pressure pv
and the mixture
temperature T.
The water vapor is
superheated at this state.

,
T

Typical state of
the water vapor
in moist air

Moist Air
6. When pv corresponds
to pg at temperature T,
the mixture is said to be
saturated.

Mixture pressure, p

,
T

7. The ratio of pv and pg


is called the relative
humidity, :
p
= v
pg

T , p

(Eq. 12.44)

Relative humidity is usually expressed as a percent and


dry air only
saturated air
ranges as
0 100%
(pv = 0)

(pv = pg)

Humidity Ratio
The humidity ratio of a moist air sample is the
ratio of the mass of the water vapor to the mass of
the dry air.
mv
(Eq. 12.42)
=
ma
Since mv << ma, the value of is typically << 1.
Using the ideal gas equation of state and the
relationship pa = p pv
18.02/28.97 = 0.622
mv
M v p vV / R T M v p v M v p v

=
=
=
=

M a pa M a p p v
M a p aV / R T
ma
pv
= 0.622
p pv

(Eq. 12.43)

Mixture Enthalpy
Values for U, H, and S for moist air can be found by
adding contributions of each component.
For example, the enthalpy H is

H = H a + H v = ma ha + mv hv (Eq. 12.45)
which conforms to Eq. (d) in Table 12.2.
Dividing by ma and introducing , the mixture enthalpy
per unit mass of dry air is
mv
H
hv = ha + hv
= ha +
ma
(Eq. 12.46) ma
For moist air, the enthalpy hv is very closely given by the
saturated vapor value corresponding to the given
temperature.
h h T
v

( )

(Eq. 12.47)

Heating Moist Air in a Duct


Example: Moist air enters a duct at 10oC, 80%
relative humidity, is heated as it flows through the
duct, and exits at 30oC. No moisture is added or
removed and the mixture pressure remains constant
at 1 bar. For steady-state operation and ignoring
kinetic and potential energy changes, determine
(a) the humidity ratio, 2, and
(b) the rate of heat transfer, in kJ per kg of dry air.

Heating Moist Air in a Duct


Solution:
(a) At steady state, mass rate balances for the dry air
and water vapor read:

m a1 = m a2 (dry air)
m v1 = m v2 (water vapor)
Since the mass flow rates of the dry air and water
vapor do not change from inlet to exit, they are
denoted for simplicity as m a and m v. Moreover, since
no moisture is added or removed, the humidity ratio
does not change from inlet to exit: 1 = 2. The
common humidity ratio is denoted by .
m

m a

Heating Moist Air in a Duct


The humidity ratio is evaluated using data at the
inlet:
The partial pressure of the water vapor at the inlet,
pv1, can be evaluated from the given inlet relative
humidity 1 and the saturated pressure pg1 at 10oC
from Table A-2:

pv1 = 1pg1 = 0.8(0.01228 bar) = 0.0098 bar


The humidity ratio can be found from:
pv
kg (vapor)
0.0098
= 0.622
= 0.622
= 0.00616
kg (dry air)
p pv
1 0.0098

Heating Moist Air in a Duct


(b) The steady-state form of the energy rate balance
reduces to:
0

0 = Q cv W cv + (m a ha1 + m v hv1 ) (m a ha2 + m v hv2 )

Solving for Qcv


Q cv = m a (ha2 ha1 ) + m v (hv2 hv1 )

Re-writing the humidity ratio, mv = ma, we get


Q cv
= (ha2 ha1 ) + (hv2 hv1 )
m a

Heating Moist Air in a Duct


Q cv
= (ha2 ha1 ) + (hv2 hv1 )
m a
For the dry air, ha1 and ha2
are obtained from ideal gas
table Table A-22 at 10oC and
30oC, respectively.

For the water vapor, hv1 and hv2 are


obtained from steam table Table A-2
at 10oC and 30oC, respectively, using
hv h g

Q cv
kJ
= (303.2 283.1)
+
kg (dry air)
m a

kg (vapor)
kJ
(2556.3 2519.8)
0.00616
kg (dry air)
kg (vapor)

kJ
Q cv
kJ
= 20.32
= (20.1 + 0.22)
m a
kg (dry air)
kg (dry air)

The
contribution of
the water vapor
to the heat
transfer
magnitude is
relatively
minor.

Work through Examples


12.3, 12.4, 12.5, 12.6

then

Work on homework problems


12.15, 12.31

Dew Point Temperature


When moist air is cooled, partial condensation of the water
vapor initially present can occur. This is observed in
condensation of vapor on window panes, pipes carrying cold
water, and formation of dew on grass.
An important special case is cooling of moist air at constant
mixture pressure, p.
The figure shows a sample of moist air, initially at State 1,
where the water vapor is superheated. The accompanying T-v
diagram locates
states of water.
Lets study this
system as it is
cooled in stages
from its initial
temperature.

Dew Point Temperature


In the first part of the cooling process, the mixture pressure
and water vapor mole fraction remain constant.
Since pv = yv p, the partial pressure of the water vapor
remains constant.
Accordingly, the water vapor cools at constant pv from
state 1 to state d, called the dew point.
The temperature at state d is called the dew point
temperature.
As the system cools
below the dew point
temperature, some of
the water vapor initially
present condenses.
The rest remains a
vapor.

Dew Point Temperature


At the final temperature, the system consists of the dry air
initially present plus saturated water vapor and saturated liquid.
Since some of the water vapor initially present has
condensed, the partial pressure of the water vapor at the final
state, pg2, is less than the partial pressure initially, pv1.
The amount of water that condenses, mw, equals the
difference in the initial and final amounts of water vapor:

mw = mv1 mv2

Dew Point Temperature


Using mv = ma and the fact that the amount of dry air
remains constant, the amount of water condensed per unit
mass of dry air is

mw
= 1 2
ma

(mw = mv1 mv2 )

where

p v1

1 = 0.622

p
p
v1

pg2
2 = 0.622
p pg2

and p denotes the mixture pressure, which remains constant


while cooling occurs.

Work through Examples


12.7, 12.8, 12.9

then

Work on homework problem 12.59

Lecture 35

Wet Bulb, Dry Bulb Temps,


Psychrometric Charts
and Air-Conditioning
Applications

Sections 12.6 12.8

Dry-bulb Temperature and


Wet-bulb Temperature
In engineering applications involving moist air, two
readily-measured temperatures are commonly used:
the dry-bulb and wet-bulb temperatures.
The dry-bulb temperature, Tdb, is simply the
temperature measured by an ordinary
thermometer placed in contact with the moist air.
The wet-bulb temperature, Twb, is the
temperature measured by a thermometer whose
bulb is enclosed by a wick moistened with water.

Dry-bulb Temperature and


Wet-bulb Temperature
The figure shows wet-bulb and dry-bulb
thermometers mounted on an instrument called a
psychrometer. Flow of moist air over the two
thermometers is induced by a battery-operated fan.
Owing to evaporation
from the wet wick to the
moist air, the wet-bulb
temperature reading is
less than the dry-bulb
temperature: Twb < Tdb.
Each temperature is
easily read from its
respective thermometer.

Moist
Air in

Psychrometric Chart
Graphical representations of moist-air data are provided by
psychrometric charts.
Psychrometric charts in SI and English units are given in
Figs. A-9 and A-9E, respectively. These charts are constructed
for a moist air mixture pressure of 1 atm.
Several important features of the psychrometric chart are
discussed in Sec. 12.7, including

Psychrometric Chart
Dry-bulb temperature, Tdb.

Moist air
state

Tdb

Psychrometric Chart
Humidity ratio, .

Moist air
state

Psychrometric Chart
Relative humidity, .

Moist air
state

Psychrometric Chart
Wet-bulb temperature, Twb.
Lines of constant wet-bulb temperature are
approximately lines of constant mixture enthalpy.

Twb
Moist air
state

Psychrometric Chart
Dew point temperature, Tdp.
Since the dew point is the state where moist air
becomes saturated when cooled at constant pressure, the
dew point for a given state is determined from the chart by
following a line of constant (constant pv) to the
saturation line where = 100%.

Moist air
state

Tdp

Psychrometric Chart
Mixture enthalpy per unit mass of dry air, (ha + hv).
The value of (ha + hv) is calculated using

ha = cpaT
Fig. 12.9: T in oC, cpa = 1.005 kJ/kg-K
Fig. 12.9E: T in oF, cpa = 0.24 Btu/lb-R
(ha + hv)

Moist air
state

Psychrometric Chart
Volume per unit mass of dry air, V/ma.
Lines giving V/ma can be interpreted as the volume of
dry air or of water vapor (each per unit mass of dry air)
because in keeping with the Dalton model each
component is considered to fill the entire volume.

Moist air
state

V/ma

Psychrometric Chart
Example: Using Fig. A-9, determine relative humidity,
humidity ratio, and mixture enthalpy, in kJ/kg (dry air)
corresponding to dry-bulb and wet-bulb temperatures of 30oC
and 25oC, respectively.

Psychrometric Chart
Solution:
(ha + hv) = 76 kJ/kg dry air

= 67%
25oC

= 0.0181 kg water/kg dry air

Analyzing Air-Conditioning Systems


The next series of slides demonstrates the
application of mass and energy rate balances
together with property data to typical air-conditioning
systems using the psychrometric principles
introduced thus far.
The applications we will explore include
Dehumidification
Humidification
Mixing of two moist air streams

Dehumidification
The aim of a dehumidifier is to remove some of
the water vapor in the moist air passing through
the unit.
This is achieved by allowing the moist air to
flow across a cooling coil carrying a refrigerant at
a temperature low enough that some water vapor
condenses.

Dehumidification
The figure shows a control volume enclosing a
dehumidifier operating at steady state.
Moist air enters at state 1.
As the moist air flows
2
1
over the cooling coil, some
= 100%,
water vapor condenses.
m , T ,
T <T ,
<
Saturated moist air exits
at state 2 (T2 < T1).
3
Condensate exits as
m
saturated liquid at state 3.
T =T
Here, we take T3 = T2.
2

Dehumidification
For the control volume, let us evaluate
The amount of
condensate exiting per unit
mass of dry air: m w/m a and
1
The rate of heat transfer
between the moist air and
m , T ,
cooling coil, per unit mass

of dry air: Qcv/ma.
3
a

m w
T3 = T2

2
2 = 100%,
T2 < T1,
2 < 1

Dehumidification
Mass rate balances. At steady state, mass rate balances
for the dry air and water are, respectively
m a1 = m a 2
(dry air)
m v1 = m w + m v 2 (water)

Solving for the mass flow rate of


the condensate
m w = m v1 m v 2

2
2 = 100%,

m a, T1, 1

Then, with m v1 = 1m a and m v2 = 2m a, where


m a denotes the common mass flow rate of
the dry air, we get the following expression
for the amount of water condensed per unit
mass of dry air
m w
= 1 2
(1)

ma

T2 < T1,
2 < 1

3
m w
T3 = T2

Dehumidification

Energy rate balance. With W cv = 0 and no significant


kinetic and potential energy changes, the energy rate
balance for the control volume reduces at steady state to
0 = Q cv + (m a ha1 + m v1hv1 ) m w hw (m a ha2 + m v 2 hv2 )

(2)

With m v1 = 1m a, m v2 = 2m a, and Eq. (1), Eq. (2) becomes


Q cv
= (ha + hv ) 2 (ha + hv )1 + (1 2 )hw
m a

(3)

Since heat transfer occurs from the moist air to the cooling

coil, Qcv/m a will be negative in value.

Dehumidification
Q cv
= (ha + hv ) 2 (ha + hv )1 + (1 2 )hw
m a

(3)

For the condensate, hw = hf (T2), where hf is obtained from


Table A-2.
Options for evaluating the underlined terms of Eq. (3) include
1 and 2 are known. Since T1 and
T2 are also known, ha1 and ha2 can be
obtained from ideal gas table Table
A-22, while hv1 and hv2 can be
obtained from steam table Table A-2
(ha + hv)2
using hv = hg.
Alternatively, using the respective
temperature and humidity ratio
values to fix the states, (ha + hv) at
states 1 and 2 can be read from a
psychrometric chart.

(ha + hv)1

T2

T1

Humidification
The aim of a humidifier is to increase the
amount of water vapor in the moist air passing
through the unit.
This is achieved by injecting steam or liquid
water.

Humidification
The figure shows a control volume enclosing a
humidifier operating at steady state.
Moist air enters at state 1.
Steam or liquid water is injected.
Moist air exits at state 2 with greater humidity
ratio, 2 > 1.
W cv = 0, Q cv = 0

ma1
3
h3, m 3

Humidification
For adiabatic operation, the accompanying
psychrometric charts show states 1 and 2 for each
case.
With relatively high-temperature steam injection, the
temperature of the moist air increases.
With liquid injection the temperature of the moist air may
decrease because the liquid is vaporized by the moist air
into which it is injected.
W cv = 0, Q cv = 0

ma1
3
h3, m 3

Humidification
For the control volume, let us evaluate
The humidity ratio, 2, and
The temperature, T2.

W cv = 0, Q cv = 0

ma1
3
h3, m 3

Humidification
Mass rate balances. At steady state, mass rate balances
for the dry air and water are, respectively
m a1 = m a 2
(dry air)
m v1 + m 3 = m v 2 (water)

Then, since m v1 = 1m a and m v2 = 2m a, where m a denotes


the common mass flow rate of the dry air, we get
m 3
2 = 1 +
m a

(1)

Since 1, m a, and m 3 are


specified, the humidity ratio 2
can be calculated from Eq. (1)

Humidification
Energy rate balance. With no significant kinetic and
potential energy changes, the energy rate balance for the
control volume reduces to
0 = Q W + (m h + m h ) + m h (m h + m h )
cv

cv

a a1

v1 v1

3 3

a a2

Since W cv and Q cv are each zero in this case

v 2 v2

0 = (m a ha1 + m v1hv1 ) + m 3 h3 (m a ha2 + m v2 hv2 )

(2)

With m v1 = 1m a and m v2 = 2m a, Eq. (2) becomes


m 3
0 = (ha1 + 1hv1 ) + ( )h3 (ha2 + 2 hv2 )
m a

(3)

Solving Eq. (3)


m 3
(ha2 + 2 hv2 ) = (ha1 + 1hv1 ) + ( )h3
m a

(4)

Humidification
m 3
(ha2 + 2 hv2 ) = (ha1 + 1hv1 ) + ( )h3
m a

(4)

Options for determining T2 from Eq. (4) include


Use the psychrometric chart:
The first term on the right side
of Eq. (4) can be read from the
chart using T1 and 1 to fix the
state.
Since the second term on the
right is known, the value of
(ha2 + 2hv2) can be calculated.
This value together with 2 fixes
the exit state, which allows T2 to
be determined by inspection.

(ha2 + 2hv2)
(ha1 + 1hv1)

2
1

T1 T2

Adiabatic Mixing of Two Moist Air Streams


In air-conditioning systems, a frequent
component is one that mixes moist air streams as
shown in the figure:

For the case of adiabatic mixing, let us consider


how the following quantities at the exit of the control
volume, m a3, 3, and T3, can be evaluated knowing
the respective quantities at the inlets.

Adiabatic Mixing of Two Moist Air Streams


Mass rate balances. At steady state, mass rate balances
for the dry air and water vapor are, respectively
m a1 + m a 2 = m a 3 (dry air)
m v1 + m v 2 = m v3 (water vapor)

With m v = m a, these equations combine to give

1m a1 + 2 m a2 = 3 (m a1 + m a2 )
Alternatively

m a1 3 2
=
m a2 1 3

(1)

These equations can be solved for 3 using known values of


1, 2, m a1, and m a2.

Adiabatic Mixing of Two Moist Air Streams


Energy rate balance. Ignoring the effects of kinetic and
potential energy, the energy rate balance for the control
volume reduces at steady state to
0 = Q cv W cv + (m a1ha1 + m v1hv1 ) + (m a2 ha2 + m v 2 hv2 ) (m a3ha3 + m v3 hv3 )
Since W and Q are each zero in this case
cv

cv

m a1 (ha1 + 1hv1 ) + m a2 (ha2 + 2 hv2 ) = m a3 (ha3 + 3 hv3 ) (Eq. 12.56c)

The enthalpies of the water vapor are evaluated using hv = hg.


With m a3 = m a1 + m a2, Eq. 12.56c can be solved to give an
expression with the same form as Eq. (1)
m a1 (ha3 + 3 hg3 ) (ha2 + 2 hg2 )
=
(2)

ma2 (ha1 + 1hg1 ) (ha3 + 3 hg3 )


Using known data, this equation can be solved for (ha + hg)3,
from which T3 can be evaluated.

Adiabatic Mixing of Two Moist Air Streams


From study of Eqs. (1) and (2) we conclude that on
a psychrometric chart state 3 lies on a straight line
connecting states 1 and 2, as shown in the figure
m a1 3 2
=
m a2 1 3

(1)

m a1 (ha3 + 3 hg3 ) (ha2 + 2 hg2 )


=
(2)

ma2 (ha1 + 1hg1 ) (ha3 + 3 hg3 )

Adiabatic Mixing of Two Moist Air Streams


Example: For adiabatic mixing of two moist air
streams with the data provided in the table below,
use the psychrometric chart to determine
(a) 3, in kg (vapor)/kg (dry air), and
(b) T3 in oC.
State
1
2

(ha + hg)*

T
m
a
(oC) (kg (dry air)/kg (vapor)) (kg (dry air)/min) (kJ/kg (dry air))
0.0094
48
24
497
0.002
10
5
180

values of (ha + hg) are read from Fig. A-9 using the respective
temperature and humidity ratio values.
*The

Adiabatic Mixing of Two Moist Air Streams


Solution:
(a) Inserting known values in Eq. (1),
497 3 0.002
=
180 0.0094 3

we get 3 = 0.0074 kg (vapor)/kg (dry air).


(b) Then from Fig. A-9

T3 = 19oC

Work on homework problem 12.70

Вам также может понравиться