Вы находитесь на странице: 1из 104

ENGRMAE 91 Introduction to Thermodynamics

Dong Hyeon Kim


donghk8@uci.edu

Summer, 2016

1
Contents
Introduction and Preliminaries 3
Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Classical vs. Statistical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Temperature Scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
The Zeroth Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

Properties of a Pure Substance 11


The Pure Substance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
The Phase Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
The P-V-T Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
Two Phase States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
State Postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Thermodynamic Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Compressed Liquid Table . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Equations of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
The Compressibility Factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

The First Law of Thermodynamics of a Control Mass 33


The Definition of Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Work Done at the Moving Boundary of a Simple Compressible System . . . . . . . . . . . 34
Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Sign Convention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
The Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
The First Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Internal Energy - A Thermodynamic Property . . . . . . . . . . . . . . . . . . . . . . . . 41
Enthalpy - A Thermodynamic Property . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
The Constant-Volume and Constant-Pressure Specific Heats . . . . . . . . . . . . . . . . . 43
The Internal Energy, Enthalpy, and Specific Heat of Ideal Gases . . . . . . . . . . . . . . 44

The First Law of Thermodynamics of a Control Volume 58


Conservation of Mass and the Control Volume . . . . . . . . . . . . . . . . . . . . . . . . 58
The Energy Equation For a Control Volume . . . . . . . . . . . . . . . . . . . . . . . . . . 60
The Steady-State Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
The Transient Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

The Second Law of Thermodynamics 70


Heat Engines and Refrigerators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
The Second Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
The Reversible Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
The Carnot Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Two Propositions Regarding the Efficiency of a Carnot cycle . . . . . . . . . . . . . . . . 75
The Thermodynamic Temperature Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

1
Ideal Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Property Relations in a Carnot Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Ideal Versus Real Machines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

The Second Law of Thermodynamics of a Control Mass 83


The Inequality of Clausius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Entropy - A Property of a System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Entropy of a Pure Substance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Entropy Change in Reversible Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
The Thermodynamic Property Relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Entropy Change of a Solid or Liquid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Entropy Change of an Ideal Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Entropy Change of a Control Mass During an Irreversible Process . . . . . . . . . . . . . 93

The Second Law of Thermodynamics of a Control Volume 98


The Steady-State Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
The Transient Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
Engineering Applications; Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

Review Video 103

2
Introduction and Preliminaries
Thermodynamics is the branch of physical sciences that deals with the relations between heat and
other types of energy (e.g. chemical, mechanical, etc.)

Since this is the introduction section, there will be little explanations for each topic.

Definitions
Let’s first start by defining some terms and concepts before the more complicated parts begins.

• System: a quantity of matter (i.e. the mass you want to study).

– Closed system (control mass): mass does not cross the boundary.
– Open system (control volume): mass does cross the boundary.
– Isolated system: nothing (mass and energy) crosses the boundary (i.e. no interaction
between the system and its surroundings).

• Surroundings: rest of the universe that interacts with the system (i.e. ”everything else”).

• Boundary (control surface): separates systems from the surroundings. Also note that a
boundary may be real -tangible- or abstract.

In this picture, the mass of the gas is the system; the dashed lines constitute the boundary.

• Thermodynamic Properties: the characteristics that help describe the state of a system
as a whole or locally.

– Extensive Property: varies directly with amount of mass (e.g. mass, volume, energy)
– Intensive Property: independent of the amount of mass (e.g. temperature, pressure,
specific energy, specific volume).

• Specific: per unit mass and uses lowercase letters to define quantities.

– Example: specific energy will be


E
e=
m

3
• Thermodynamic States: the specific conditions a system is in when determined by its
macroscopic properties (i.e. If you have certain values (e.g. pressure) for the properties,
those values will determine the state of the system).

• Equilibrium: a condition of balance characterized by the absence of driving potentials.

– Example: Heat transfer occurs when there is a temperature difference to reach a balance
– hot air transfers to cold air for both to reach same temperature.

• Thermodynamic Process: a transition from one state to another state. Process occurs
when the properties of a system change.

– Isothermal: constant temperature.


– Isobaric or Isopiestic: constant pressure.
– Isochoric or Isometric: constant volume.
– Cycle: a system in which it returns to the initial state.
– Isentropic: constant entropy.
– Isenthalpic: constant enthalpy.
– Adiabatic: no heat transfer.
– Reversible: process that can be reversed and brings the system and the surrounding
to the original state.
– Irreversible: All real processes where the system and the surrounding cannot be
brought back to the original state.

• Property Diagrams: representations of the processes.

– Temperature vs. Pressure


– Temperature vs. Volume
– Pressure vs. Volume
– Temperature vs. Entropy
– Enthalpy vs. Entropy

• Quasi Equilibrium Process: a process in which each intermediate state is only infinitesi-
mally removed from equilibrium. In other words, if the change in system is small (infinitesi-
mal) and requires a large period of time to complete, it is said to take place in quasi equilib-
rium.

4
In this picture, imagine the weights are infinitesimally small and taken off one by one.
Then the process would become quasi equilibrium because the piston would rise slowly.

Classical vs. Statistical


In thermodynamics, there are two approaches to study thermodynamics: classical and statistical.

• Classical

– Macroscopic view
∗ Temperatures
∗ Pressures

• Statistical

– Microscopic view
∗ Atoms
∗ Molecules
∗ Molecular interactions, which produce measurable results such as temperatures and
pressures.

Energy
Total energy is the sum of all the possible energies in the microscopic approach and can be defined
with the letter E:

E = U + KE + P E + ...

Where U is internal energy, KE is kinetic energy, P E is potential energy.

• Internal Energy associated with atomic and molecular energies.

• Potential Energy associated with intermolecular forces.

5
• Kinetic Energy associated with translational velocities.

• Vibrational Energy associated with molecular structure.

Let’s recall extensive and intensive properties; extensive deals with mass and intensive does not.
In order to deal with intensive internal energy, divide the internal energy by mass, which becomes
specific internal energy:

U
u=
m

Specific internal energy will be commonly used throughout this course.

Pressure
Let’s first recall pressure.

dFn
P = lim
dA→0 dA

Where Fn is the normal component of the force and dA is the varying area.

Now, let’s quickly recall a different equation for the pressure, but in relations with a ambient
pressure and certain depth:
P = P0 + ρgh
For pressure there are several units to consider. The most common one is the SI unit pascal, but
there are other common ones that are frequently used. There is the unit reference on the next page:

N
1 Pa = 1
m2
1 bar = 105 P a = 0.1 M P a = 100 KP a

Now, let’s learn the difference between absolute pressure and gauge pressure. Have a look at this
useful diagram on the next page:

6
The absolute zero pressure -a vacuum- is denoted as 0.
The absolute pressures are denoted as Pabs,1 and Pabs,2 .
The gauge pressure is the difference between Pabs,1 and Patm .
The vacuum gauge pressure is the difference between Patm and Pabs,2 .

In essence, gauge pressure is explained as the change in pressure or the difference of pressure:

∆ = P − P0 = ρgh

Temperature Scales
Be familiarized with the absolute temperature scales: Rankine and Kelvin.

F arenheit = 1.8 × Celsius + 32

Kelvin = Celsius + 273.15


Rankine = F arenheit + 459.67
Rankine = 1.8 × Kelvin

7
The Zeroth Law of Thermodynamics
The zeroth law of Thermodynamics says that if two bodies are each in thermal equilibrium with
some third body, then they are also in equilibrium with each other.

If TA = TC and TB = TC , then TA = TB

Let’s do some example problems now.

Example (1)
kg
Two cylinders are filled with liquid water, ρ = 1000 m 3 , and connected by a line with a

closed valve, as shown in the figure below. A has 100 kg and B has 500 kg of water, their
cross-sectional areas are AA = 0.1 m2 and AB = 0.25 m2 , and the height h is 1 m. Find the
pressure on either side of the valve.

Solution (1)
kg
Given: MA = 100 kg, MB = 500 kg, AA = 0.1 m2 , AB = 0.25 m2 , h = 1 m, ρ = 1000 m 3 , and

P0 = 101325 P a

Since there are multiple pressures interacting, this is a problem that deals with gauge pres-
sure (P = P0 + ρgh). Therefore, find out which variables are missing. In this problem, the
height of each water tanks must be found in order to find the pressures at the valve.

8
Let’s first denote hA and hB the height of the water in the respective water tanks. Now
recall that volume is the cross sectional area multiplied by its height. Next recall that den-
sity is equal to mass divided by volume. Knowing and manipulating the variables of these
two equations, the heights of each water tanks can be related and written as such:

VA (MA )
hA = = (1)
AA (ρ)(AA )

VA (MB )
hB = = (2)
AA (ρ)(AB )

Next let’s denote PA as the pressure on the right side of the valve and PB as the pressure
on the left. Since the water tanks are open, it can be concluded that there is atmospheric
pressure. Since there is a difference of pressure on the right and left side of the valve, the
gauge pressures for tank A and B can be written as such:

(MA )(g)
PA = Po + ρghA = P0 + (3)
(AA )
PA = 111135 P a

(MB )(g)
PB = Po + ρg(hB + h) = P0 + ρgh + (4)
(AB )
PB = 130755 P a

Now what happens when the valve is opened? The answer is that the system will no longer
be in equilibrium because the water fluid will start to flow and eventually reach a new state
of equilibrium.

Example (2)

A 5 kg piston in a cylinder with a diameter of 100 mm is loaded with a linear spring


and the outside atmospheric pressure is 100 kP a, as shown in the figure below. The spring
exerts no force on the piston when it is at the bottom of the cylinder, and the spring con-
N
stant is 50 mm . For the state shown, the volume is 0.4 L. Find the pressure inside the cylinder.

9
Solution (2)
N
Given: m = 5 kg, diameter = 100 mm → A = π(502 ) mm2 , P0 = 100 kP a, k = 50 ,
mm
and V = 0.4 L

Let’s analyze what is going on for the state. In the shown state, the piston is assumed
to be in equilibrium. Therefore, the forces acting on the piston are balanced; the pressure
acting on the bottom of the piston, the pressure acting on top of the piston, gravitational
weight, and the spring force. There is a spring force because the problem states that when
the piston is at the bottom of the cylinder, the spring exerts no force. In the figure, the
spring is compressed and the spring force directs downward. Let’s write the balanced equa-
tion while denoting up as positive and down as negative:

Fnet −→ P1 A − P0 A − mg − kx = 0

Note that being in equilibrium does not mean that the piston is not moving at all. It can
move at some constant velocity. Now if the piston is moving at a very slow constant velocity,
the system is in quasi equilibrium state. Okay, now the only unknown variable is the x.
However, x can be substituted by the volume divided by its cross sectional area:
V
x=
A
.
V
So after substituting the A
into the net force equation, the new equation is:
mg kV
P = P0 + + 2
A A

After plugging in everything:


P = 100kP a

10
Notice how the equation
mg kV
P = P0 + + 2
A A

Can be written as

P = C1 + C2 V

mg kV
Where C1 and C2 are constants, and is P0 + A
and is A2
, respectively.

Properties of a Pure Substance


This section focuses on Chapter 2 in the Fundamentals of Thermodynamics 8th Edition by
Borgnakke. Let’s get started on the topic of properties of a pure substance.

The Pure Substance


• Pure substance is one that has a homogeneous and invariable chemical composition.
It may exist in more than one phase, but the chemical composition is the same in all
phases.
– Example: Liquid water, a mix of liquid water and water vapor, and a mix of ice
and liquid water are all pure substances because every phase has the same chemical
composition.
– Example: A mix of liquid air and gas air is not a pure substance because the
composition of the liquid phase is a different from that of a gas phase.
– Example: Air can be treated as a pure substance even though air is made up of
different gases. If air is put into a cylinder and the air has same composition of
different molecules everywhere, then it can be treated as a pure substance.
• Saturation temperature designates the temperature at which vaporization takes
place at a given pressure.
• Saturation pressure designates the pressure at which vaporization takes place at a
given temperature.

The saturation temperature and the saturation pressure are linked at the moment of va-
porization. For example, when given a graph and a value of a saturation temperature, the
saturation pressure can be found from the graph. Vice versa. We will see more of this later.

11
The Phase Diagram
Let’s define some definitions and introduce the P-T graph.

The temperature at which vaporization takes place at a given pressure is commonly known
as boiling temperature. If it is the opposite, in which freezing takes place at a given
pressure, then it is called freezing point.

Let’s look at a water phase diagram.

Fusion line is the border between the solid phase and the liquid phase, whereas the right
one is called vaporization curve. The sublimation line separates the solid phase and
the vapor phase. The point where the curves meet is called the triple point and is the only
P, T combination in which all three phases (S,L,V) can coexist.

A phase diagram shows the distinct set of saturation properties for which it is possible
to have two phases in equilibrium with one another. For a high pressure, the vaporization
curve stops at a point called critical point. At the critical point, there will be no boiling
and the heat of a liquid will produce a vapor at a smooth transition. Also note that some
substances can make a phase transition from a solid to a vapor without going through the
liquid phase and this process is called sublimation.

The P-V-T Graphs


Let’s consider an experiment. Consider that there is liquid water at room temperature in
a cylinder with an attached massless piston, and the pressure outside can be controlled.
Now, heat is added slowly -quasi equilibrium- to the system. What happens? Temperature
increases, volume increases only slightly, and pressure stays constant. When the temperature
finally reaches 99.6◦ C, a phase change will result and vapor will form on top of liquid water,
due to water being denser than gas. Check (b) in the figure below. During this phase change,
volume increases considerably, while temperature remains the same -this is due to the water
molecules absorbing all the heat energy to break their bonds. Further heating creates more
and more vapor and eventually it results in a large increase in volume until all of the liquid
water is vaporized. The temperature will start to change once every drop of liquid has been
vaporized. Check out (c) in the figure below.

12
When the total volume, mass (specific volume), and temperature of water can be measured,
this experiment can be plotted as temperature versus volume. Let’s consider the experiment
again. Follow along the proceeding paragraphs along with the T − V graph below the para-
graphs.

Add heat slowly; the temperature goes up and volume expands slightly. This starts from
state A and goes towards state B. When state B is reached, the liquid water is at 99.6◦ C,
which is called saturated liquid, and there is a first sign of vapor. This line in the graph is
called saturated liquid line and is steep because liquid water cannot expand significantly.
Further heating increases the volume at a constant temperature (bonds breaking). Eventu-
ally, all the liquid is vaporized, which is called saturated vapor and is labeled as state C.
This line in the graph is called saturated vapor line and is less steep because water vapor
expands more significantly in the cylinder. Further heating will produce vapor at higher
temperatures in states called superheated vapor, where the temperature is higher than
the saturated temperature for the given pressure. The difference between a given T and the
saturated temperature for the same pressure is called degree of superheat. Note, that to
the left of the dome shaped curve, the region is called compressed liquid.

For higher pressures, the saturated temperature is also higher; for example, state F has
a saturated temperature of 179.9◦ C and a pressure of 1 MPa. State J has a saturated tem-
perature of 311◦ C and a pressure of 10 MPa and so forth. At the critical pressure or
critical point, there is a smooth transition from a liquid state to a vapor state without
going through the constant temperature vaporization line; for example, at 22.09 MPa, the
heating goes from state M to state N to state O in a smooth transition.

Heating at lower pressure (lower saturation temperature), two phases can be presented si-
multaneously, which is represented inside the blue, dome shaped curve. Heating at higher
pressures, two phases are not presented simultaneously; for example, at 22.09 MPa, the line
does not enter the dome. The region where the substance clearly changes from a liquid
state to a vapor state is called dense fluid. The states where the saturation temperature is
reached for the liquid (B, F, J) are saturated liquid states forming the saturated liquid line.
The states along the other two-phase region (N, K, G, C) are saturated vapor states forming
the saturated vapor line.

13
For typical substances, there are combinations of possible P − V − T graphs. There are
T − V graph or P − V graph or P − T graph or all of them in a 3D representation:

14
The figures above are the P − V , P − T , and P − V − T graphs. They show a substance
such as water that expands, or increases in volume, during freezing, so the solid surface had
a higher specific volume than a liquid surface. Note: The P-V-T diagram shows that the
constant pressure line (typical solidification) increases in volume in the saturated
solid-liquid region when going from liquid to solid.

15
The figures above are the P − V , P − T , and P − V − T graphs. They show a substance
that contracts, or decreases in volume, upon freezing, which is a common condition. Note:
The P-V-T diagram shows a reduction in volume of the constant pressure line when going
from liquid to solid.

16
The extended property diagram show:
• Triple point on the (P vs. T), whereas the triple point becomes a triple line on the (T
vs. V or P vs. V or P-V-T surface).
• Solid.
• Solid-liquid mixture.
• Solid-vapor mixture.
• The temperature, pressure, and volume diagram.
If you look at the 2 three-dimensional surfaces, the P-T phase diagram can be seen when
looking at the surface parallel with the volume axis; the liquidvapor surface is flat in that
direction, so it becomes the vaporization curve. The same happens with the solidvapor sur-
face, which is shown as the sublimation line, and the solidliquid surface becomes the fusion
line. For these three surfaces, it cannot be determined where on the surface a state is by
having the (P, T) coordinates alone. The two properties are not independent; they are a
pair of saturated P and T. A property like volume is needed to indicate where on the surface
the state is for a given T or P.

If the surface is viewed from the top down parallel with the pressure axis, you can see
the TV diagram without the complexities of the solid phase. Sometimes, one of the three
surfaces can show one property better than the others. One example is that the P-V diagram
with the constant temperature line (viewing parallel with the temperature axis) can show a
liquid-vapor region, similar to the T-V graph.

In the T-V graph, the pressure lines are constant, and they enter from the bottom left
and exit through the top right. Whereas, in the P-V diagram, the temperature lines are
constant, and they enter from the top and exit through the bottom right.

17
Two Phase States
Let’s look at the T − V graph again:

Recall that the inside the dome shaped curve, two phases of vapor and liquid can exist si-
multaneously.

If there was a less water vapor and more liquid water, the point would be closer to the
saturated liquid line. If there was more water vapor and less liquid water, the point would
be closer to the saturated vapor line. However, if given just the temperature and the pressure
values, a point cannot be found inside the dome shaped curve. For example, if given the
pressure value of 0.1 MPa and temperature value of 99.6◦ C, the point can be either be closer
to state B or closer to state C. To answer this problem, let’s move on to the next topic. See
the graph below:

18
By convention, the subscript f is defined as a property of a saturated liquid and the subscript
g is for a property of a saturated vapor. A saturation condition involving part liquid and part
vapor can be displayed on a T −v coordinates as in the figure above. All of the liquid present
is at state f with a specific volume vf and all of the vapor present is at state g with specific
volume vg . The total volume can be expressed as the sum of liquid volume and vapor volume:

V = Vliq + Vvap = mliq vf + mvap vg

To get the average specific volume of the system v, just divide by the total mass:

V mliq mvap
v= = vf + vg = (1 − x)vf + xvg
mtotal mtotal mtotal

The ratio of the mass of the vapor and the total mass of vapor and liquid can be defined as
quality:
mvap mvap
x= =
mtotal mvap + mliq

Quality is only defined and used for a mixture of saturated vapor and saturated liquid, which
is only in the dome region. So, quality does not apply in compressed liquid or a superheated
vapor because compressed liquid region is definitely all liquid and superheated vapor region
is all vapor. Note that quality can also be expressed in percentage. Finally, the quality x can
(v − vf )
be viewed as the fraction of the distance between saturated liquid and saturated
vf g
vapor as shown in the graph above.

19
State Postulate
The concept of state postulate tells us how many properties we need to define a state of a
certain system. For a simple, pure substance (a.k.a. simple compressible system), the state
is defined by two independent, intensive properties. Note: a simple compressible system is
a system where electrical, magnetic, gravitational, motion and surface-tension effects are
absent. The electromagnetic effects are considered as an internal force.

To understand the term independent property, consider the saturated-liquid and saturated-
vapor states of a pure substance. These two states have the same pressure and the same
temperature, but they are definitely not the same state. Therefore, in a saturation state,
pressure and temperature are not independent properties. Two independent, intensive prop-
erties are such as pressure and specific volume or pressure and quality, and they are required
to specify a saturation state of a pure substance.

Let me simplify what was just explained, so take a look at this graph again.

Let’s say I gave you a temperature value of 100◦ C and a pressure value of 0.1 MPa, then, on
the graph, you can easily plot a point in the superheated vapor region, which is a specified
state. Also, if gave you a temperature value of 300◦ C and a pressure value of 10 MPa, then
you can plot a point in the compressed liquid region, which is also a specified state. However,
if I gave you a temperature value of 179.9◦ C and a pressure value of 1 MPa, then the point
is inside the dome and a specific state cannot be exactly found.

Therefore, in the saturation state, pressure and temperature are not independent prop-
erties. So if you are given specific volume (volume is specific because the property must be
intensive) and quality or temperature and quality, then the saturation state can be found.
As long as the two properties are independent, then it is okay.

20
Typical sets of independent properties include:
• P, v (Pressure, Specific volume)
• T, v (Temperature, Specific volume)
• T, x (Temperature, Quality)
• P, x (Pressure, Quality)
• T, s (Temperature, Specific entropy)
• h, s (Specific enthalpy, Specific entropy)
Again, to determine the state of a pure substance, two independent and intensive ther-
modynamic properties are required. Then, the remaining properties can be obtained from
state laws (constitutive equations), tables or charts. For more complex systems, additional
properties are required. For example, if gravity considered, or velocity if motion is present.

Thermodynamic Tables
Thermodynamic tables are categorized by substance and phase. The information is normally
provided for:
• Compressed liquid.
• Saturated liquid-vapor mixture (pressure or temperature entry).
• Saturated solid-vapor mixture.
• Superheated vapor.
Since I am using Fundamentals of Thermodynamics 8th Edition by Borgnakke as my text-
book, the thermodynamic tables are all located in the appendices. So refer to the appendices
in the textbook whenever the specific values are needed.
Water: Appendix B.1

Ammonia: Appendix B.2

CO2 : Appendix B.3

R-410a: Appendix B.4

R-134a: Appendix B.5

N2 : Appendix B.6

21
Compressed Liquid Table
The table in the textbook has limited data for the compressed liquid. It has limited amount
of values for each properties. Thus, in the absence of a compressed liquid table, you can
approximate properties using the saturated liquid-vapor table. Because the properties are
more sensitive to changes in temperature, it is better to approximate to the saturated liquid
values at the requested temperature.

For example, imagine you want to find v(T, P ) and you noticed that it is a compressed
liquid. Then you will read vf (T ) and use that value as an approximation; v(T, P ) ≈ vf (T ).

Equations of State
Given property 1 and property 2, which are both intensive and independent, the state pos-
tulate tells us that those properties allow us to find any different property.

P roperty3 = f (P roperty1 , P roperty2 )

This could be then be expressed as tabulated data or as an equation, instead of that generic
function of properties.

In essence, instead of the ideal gas model to represent the gas behavior, or even the gen-
eralized compressibility chart, which is an approximate, it is desirable to have an equation
of state that accurately represents the P-v-T behavior for a particular gas over the entire
superheated vapor region. One equation is the cubic equations of state

RT a
P = − 2
v − b v + cbv + db2

in terms of the four parameters a, b, c, and d. If these four parameters are zero, then the
equation becomes the ideal gas model. The ideal gas law equation is a combination of
Boyle’s law, Charles’ law, and Gay-Lussac/Amonton’s law.

P V = mRT

Where the gas constant is R = R̄/Mm , the universal gas constant is R̄ = 8.314 kJ/kmol · K
and the molar mass is Mm .
These all basically reduces the equation to

P V = nR̄T

where n is the number of moles. Also to write the ideal gas law equation in terms of intensive
properties then it is
P v = RT
where R = R̄/M and M is the molecular weight.

22
The ideal gas law equation can be used if the density is low enough so that the intermolecular
distance is large.
P
ρ=
RT
As you can see, the density is low if the temperature is high or the pressure is low.

The Compressibility Factor


Again, the gas can be considered as ideal when the density is low. Therefore, only use the
ideal gas with these cases:
1. If the given pressures are significantly lower than the critical pressure, then the given
temperatures can be moderately low relative to the critical temperature.
2. If the given pressures are lower than the critical pressure, then the given temperatures
cannot be too low relative to the critical temperature.
3. If the given temperatures are significantly higher than the critical temperature, then
the given pressures can be moderately high relative to the critical pressure.
4. If the given temperatures are higher than the critical temperature, then the given
pressures cannot be too high relative to the critical pressure.
But what happens if these indicators of ideal gas fail? If we have the thermodynamic tables,
then looking at the tables is the best option, but what if we do not have the tables or the
specific substance is not in the tables?

We have several options. We can assume the gas is ideal and just use the ideal gas law
equation to get the best approximate value or we can use the cubic equation of state and
possibly find the four parameters (a,b,c,d) needed for the equation. The best option is to
use the compressibility factor equation.

Pv
Z=
RT
or
P v = ZRT
Note that for an ideal gas Z = 1, and the deviation of Z from unity is a measure of the
deviation of the actual relation from the ideal gas equation of state. So if pressure goes up,
then Z will go up. If the temperature goes down, then Z will also go up.

23
This figure shows a skeleton compressibility chart for nitrogen. From this chart, we make
three observations. The first is that at all temperatures Z → 1 as P → 0. That is as the
pressure approaches zero, the P-v-T behavior closely approaches that predicted by the ideal
gas equation of state. Second, at temperatures of 300 K and above (that is, room tempera-
ture and above), the compressibility factor is near unity up to a pressure of about 10 MPa.
This means that the ideal gas equation of state can be used for nitrogen (and of course, air)
over this range with considerable accuracy.

Third, at lower temperatures or at very high pressures, the compressibility factor devi-
ates significantly from the ideal gas value. Moderate-density forces of attraction tend to
pull molecules together, resulting in a value of Z < 1, whereas very-high-density forces of
repulsion tend to have the opposite effect.

If we examine, the compressibility diagrams for other substances, we find that the dia-
grams are all similar to that of the nitrogen. To put all the substances on a common basis,
we ”reduce” the properties with respect to the values at the critical point. The reduced
properties are defined as

P
reduced pressure = Pr = , Pc = critical pressure
Pc

T
reduced temperature = Tr = . Tc = critical temperature
Tc

24
This figure shows the generalized compressibility chart. If we have already calculated
the reduced pressure, then we go to the chart, find the reduced temperature, and find the
intersection with the compressibility factor.

Let’s do some example problems now.

Example (1)

Determine the state when (use Table B.1.2).

(a) P = 10 MPa and v = 0.003 m3 /kg

Solution (1)

Using the textbook, go to the Appendix B and find the table B.1.2. Notice how the pressure
in the first column is given by kilo pascals and not mega pascals. So convert 10 MPa to kPa,
which is 10,000 kPa.

25
From reading the table, 10,000 kPa has a temperature value of 311.06◦ C, but how can we
determine the state?

Using the given value of the specific volume, v = 0.003 m3 /kg, we can compare it to the
values of the specific volume of the saturation liquid and vapor. 0.003 m3 /kg is greater than
the specific volume of the saturation liquid, which is 0.001452 m3 /kg, and less than the
specific volume of the saturation vapor, which is 0.01803 m3 /kg.

Therefore, the state is in saturation state.

Example (2)

A piston arrangement is loaded with a linear spring and the outside atmosphere. It contains
water at 5 MPa, 400◦ C, with the volume being 0.1 m3 , as shown below. If the piston is at
the bottom, the spring exerts a force such that Plif t = 200 kPa. The system now cools until

26
the pressure reaches 1200 kPa. Find the mass of water and the final state (T2 , v2 ) and plot
the P-v diagram for the process.

Solution (2)

Given: P1 = 5 M P a, T1 = 400◦ C, V = 0.1 m3 , Plif t = 200 kP a with the piston at the


bottom, Process cools until P2 = 1200 kP a.

Find: mwater and (T2 , v2 ) and plot the P-v diagram.

First, let’s look at the table. We can start with the saturated table or the superheated
table. Let’s look at the superheated table first and then find the sub-table that corresponds
with the given pressure, which is 5 MPa (5000 kPa).

As you can see from the sub-table, next to the pressure of 5000 kPa, there is a temperature in
parenthesis, which is 263.99◦ C. This is the saturation temperature for a pressure of 5000 kPa.

27
With the saturation temperature found from the table, now we can find out what state
the system is before it cools. The temperature before the cooling process is 400◦ C, which is
higher than the saturation temperature of 263.99◦ C. Thus, the initial state of the system is
in the superheated vapor region of the P-v diagram.

Now, let’s find the value of the specific volume of the initial state. Go to the table again,
and go to the row where the temperature is 400◦ C. You can see that at 400◦ C, the specific
volume is 0.05781 m3 /kg.

Let’s now look at the state of the system after it starts to cool. If the system cools down,
then the temperature and pressure decrease. As the pressure decreases, the piston descends,
which means that the specific volume must also decrease. These clues all point to the idea
that the line of the P-v graph is linear and decreasing because the system is a cylinder with
a linear spring and goes from a higher pressure and specific volume to lower pressure and
specific volume.

So what do we know so far? We know that the initial pressure is 5000 MPa and initial
specific volume is 0.05781 m3 /kg. We also know that the line is linear and decreases. We
also know that when the piston is at the bottom, the pressure is 200 kPa and the specific
volume is 0. Let’s graph with the information we have so far.

Recall this linear spring equation of pressure from the previous section:

P = C1 + C2 V

In this problem’s case, the equation becomes

P = P0 + Cv

This equation shows that the new specific volume can be found using the slope formula.
Let’s first find the slope of the entire line.
P1 − Plif t 5000 − 200
= = 83030.61
v1 − 0 0.05781 − 0

28
Then, find the final specific volume.
P1 − P 2 5000 − 1200
= = 83030.61
v1 − v2 0.05781 − v2
Thus,
v2 = 0.01204 m3 /kg
Note: The idea of using the slope to find the value of an unknown is called interpolation.
Look up some examples to get familiar with it because it is an important and useful concept
for thermodynamics. Now, to find the mass, recall that the volume is specific volume ×
mass. Thus,
V 0.1
mwater = = = 1.73 kg
v1 0.05781
Finally, to find the temperature after the state of the system cools, use the saturated table
in the textbook this time.

As you can see when the pressure is 1200 kPa, the temperature is 187.99◦ C.
Thus, T2 = 187.99◦ C.

Example (3)

A 1-m3 rigid tank with air at 1 MPa and 400 K is connected to an air line as shown in
the figure below. The valve is opened and air flows into the tank until the pressure reaches
5 MPa, at which point the valve is closed and the temperature inside is 450 K.
(a) What is the mass of air in the tank before and after the process?
(b) The tank eventually cools to room temperature, 300 K. What is the
pressure inside the tank then?

29
Solution (3)

Since the substance is air, which is a mixture, we can assume that the abundant substance
is nitrogen.

As you can see, the critical pressure of nitrogen is 3.39 MPa, and the critical temperature
is 126.2 K. Now look at the initial pressure is 1 MPa and the final pressure is 5 MPa. As
you can see, the final pressure is not lower than the critical pressure. One of the criteria to
assume if a gas is ideal does not fit.

So let’s look at the criteria regarding the temperature. The initial temperature is 400 K
and the final temperature is 450 K. Both of the temperatures are higher than the critical
temperature so the criteria, of assuming if a gas is ideal, regarding the temperature does
fit. The given temperatures are significantly higher than the critical temperature, so the
pressures can be moderately high.

Remember, in order to assume if a gas is ideal, the density must be low. In order for
the density for to be low, either the temperature has to be high or the pressure has to be

30
low.
P1 V 1000 × 1
mair1 = = = 8.711kg
RT1 0.287 × 400

P2 V 5000 × 1
mair2 = = = 8.711kg
RT2 0.287 × 450

To find (b), understand that the process from state 2 to state 3 (room temperature), the
volume and mass both stay constant during the cooling. We can use the ideal gas law equa-
tion again, but this time we have three constants, which are R, m, and V. So we can exclude
them in our equation.
P3 P2
=
T3 T2

T3 300
P3 = P2 × = 5000 × = 3.33 M P a
T2 450

Example (4)

Argon is kept in a rigid 5-m3 tank at -30◦ C and 3 MPa. Determine the mass using the
compressibility factor. What is the error (%) if the ideal gas model is used?

Solution (4)

This is a compressibility factor problem so we can safely assume that there is no such thing
as an Argon table; therefore, we use the generalized compressibility factor chart.

31
However, we first need to know the critical temperature and pressure, so we go to Appendix
A in the textbook, which shows all the critical constants of substances.

The critical temperature for Argon is 150.8 K and the critical pressure is 4.87 MPa. Then
find the reduced temperature and reduced pressure.
243.15
Tr = = 1.612
150.8
3000
Pr = = 0.616
4870
Now using the generalized chart, we can find that Z ≈ 0.96 Finally, using the compressibility
factor equation, we can find the mass
PV 3000 × 5
m= = = 308.75 kg
ZRT 0.96 × 0.2081 × 243.2

32
To find the error, we use the ideal gas law equation, which states that Z = 1
PV
m= = 296.4 kg
ZRT
Therefore, there is a 4% error.

The First Law of Thermodynamics of a Control Mass


The first law of thermodynamics and energy equation are alternative equations for the same
fundamental physical law. Later we will see the actual difference in the expression of the
first law and the energy equation and recognize that they are consistent with one another.
Our procedure will be to state the energy equation for a system (control mass) undergoing
a process with a change of state of the system with time. We then look at the same law
expressed for a complete cycle and recognize the first law of thermodynamics.

After the energy equation is formulated, we will use it to relate the change of state in-
side a control volume to the amount of energy that is transferred in a process as work or
heat transfer. We can also see a change in state without any work or heat transfer. The
energy equation relates the various forms of energy of the control mass to the transfers of
energy by heat or work.

The Definition of Work


The classical definition of the work is mechanical work defined as a force F acting through
a displacement x, so incrementally
δW = F dx
and the finite work is Z 2
1 W2 = F · dx
1
To evaluate the work, it is necessary to know the force F as a function (dot product) of x.

Work is energy in transfer and thus crosses the control volume boundary (heat is also a
form of work and thus crosses the control volume boundary as well). It can be counted as
mechanical work, electrical work, chemical work, etc. Look at the figure below with a simple
system of a battery, motor, and a pulley.

33
Depending upon the choice of control volume, the work crossing the surface, as in sections
A, B, or C, can be electrical through the wires, mechanical by a rotating shaft out of the
motor, or a force from the rope of the pulley.

The work is scalar with units of energy and is called the joule (J).
1J =1N ·m
Also, if we take the derivative of work, then we get the rate of work, which is called power.
W
Ẇ = P =
t
where P is power (do not confuse power and pressure) and t is the time. The unit of power
is given by watts: [ Js ] = [W ].

The traditional definition of work is not really useful for thermodynamics since we use
pressure, volume, and other thermodynamic-related quantities a lot. This brings us to the
next topic of work.

Work Done at the Moving Boundary of a Simple Compressible System


In this section, we will consider in some detail the work done at the moving boundary of a
simple compressible system during a quasi-equilibrium process. Look at the figure below.

The system has a gas contained in a cylinder and a piston. Remove one of the small weights
from the piston, which will cause the piston to move upward a distance of dL. We can
consider this quasi-equilibrium process and calculate the amount of work W done by the
system during this process. The total force on the piston is AP , where P is the pressure of
the gas and A is the area of the piston. Therefore, the work δW is
δW = AP dL
but AdL = dV , the change in volume of the gas. Therefore,
δW = P dV
The work done at the moving boundary during a given quasi-equilibrium process can be
found by integrating δW = P dV . However, the integration can only be performed only if
we know the relationship between P and V during this process. This relationship may be
expressed as an equation or shown on a P-V diagram. Consider the graph below.

34
We use the same system of a compressible process with a piston/cylinder. At the beginning
of the process, the piston is at position 1, and the pressure is low. At the end of the process,
the piston is at position 2. Let us assume the process was quasi-equilibrium and that during
the process, the system passed through the states shown by the line connecting states 1 and
2. The work done on the air during compression process can be found by integrating and
thus,
Z 2 Z 2
1 W2 = δW = P dV
1 1

1 W2

is the work done during the process from state 1 to state 2.


Z 2
P dV
1

is the area under the curve 1-2.

Further consideration of the P-V diagram, such a the figure below, leads to another im-
portant discussion about work.

35
It is possible to go from state 1 to state 2 along many different quasi-equilibrium paths, such
as A, B, and C. Since the area under each curve represents the work for each process, the
amount of work done during each process not only is a function of the end states of the
process but also depends on the path followed in going from one state to another. Thus,
work (and heat) is called a path function or path dependent or in a mathematical term,
an inexact differential.

This is why the Greek symbol for differential, δ, is used for work because work is dependent
on the path. Whereas, d, also for differential, is used for volume because volume is only
dependent on the initial and final position and not dependent on the path.

This concept leads to a discussion of point and path functions. Thermodynamic proper-
ties are point functions, or exact differentials, a name that comes from the fact that, for
a given point on a diagram or surface, the state is fixed, and thus, there is a definite value
for each property corresponding to this point. So work and heat are not point functions,
rather path functions, because work and heat are not fixed at a certain point and requires a
path to know a definitive value.

Thus, volume, as a point function, can be written as


Z 2
dV = V2 − V1
1

Work (and heat), as a path function, can be written as


Z 2
1 W2 = δW 6= W2 − W1
1
Z 2
1 Q2 = δQ 6= Q2 − Q1
1
We will talk more about heat later, but this is just to note that heat is path dependent.

Let’s now look at work in specific cases of different processes.


• Constant volume process (V = 0).
1 W2 =0

• Constant pressure process.


1 W2 = P (V2 − V1 )

• Linear P-V process.


1
1 W2 = (P1 + P2 )(V2 − V1 )
2

36
• Non-linear P-V process where P = P (V ) (P is not constant).
– Need to consider polytropic process, which is a specific function of pressure vs.
volume. Z 2
1 W2 = P dV
1
P2 V2 − P1 V1
1 W2 =
1−n
V2 V2
1 W2 = P1 V1 ln = P2 V2 ln
V1 V1
Polytropic process with pressure and volume as the two parameters, an exponent, and a
constant is written as
P V n = constant
The polytropic exponent n is indicative of the type of process, and it can vary as real num-
bers. When n = 0, we have a constant pressure process. When n → ±∞. we have a constant
volume process.

When n 6= 1, we calculate the integral as

P V n = constant = P1 V1n = P2 V2n


constant P1 V1n P2 V2n
P = = =
Vn Vn Vn
Z 2 Z 2
dV V −n+1 2
P dV = constant n
= constant( )
1 1 V −n + 1 1
2
P2 V2n V21−n − P1 V1n V11−n P2 V2 − P1 V1
Z
constant 1−n
P dV = (V2 − V11−n ) = =
1 1−n 1−n 1−n

When n = 1,
P V = constant = P1 V1 = P2 V2
Z 2 Z 2
dV V2 V2
P dV = constant = P1 V1 ln = P2 V2 ln
1 1 V V1 V1

Heat Transfer
Let’s discuss heat and heat transfer now. This section briefly discusses heat transfer and the
more of heat transfer will be talked about in the ENGRMAE 120 course, which is Heat and
Mass Transfer.

Heat transfer is the transport of energy due to a temperature difference between the amounts
of matter. Molecules of matter have translational (kinetic), rotational, and vibrational en-
ergy. Energy in these modes can be transmitted to the nearby molecules by interactions

37
(collisions) or by exchange of molecules such that energy is emitted by molecules that have
more on average (higher temperature) to those that have less on average (lower temperature).
This energy exchange between molecules is heat transfer by conduction, and it increases
with the temperature difference and the ability of the substance to make the transfer. This
is expressed in Fourier’s law of conduction.

THot − Tcold
Q̇ ≈ kA
L
or
∂T
q̇ 00 = −k
∂x
where q̇ 00 = q̇ 00 · n̂ is the heat flux or rate of heat transfer per unit area, and given by units
W W
of [ m 2 ]. k is the thermal conductivity of the material and given by units of [ m·K ].

A different mode of heat transfer takes place when a medium is flowing, called convec-
tive heat transfer. In this mode, the bulk motion of a substance moves matter with a
certain energy level over or near a surface with a different temperature. The heat transfer
by conduction is dominated by the manner in which the bulk motion brings the two sub-
stances in contact or close proximity. In other words, heat energy is transferred between
a surface and a moving fluid at different temperatures is known as convection. Convective
heat transfer may take the form of either forced or natural. Forced convection occurs when
a fluid flow is induced by an external force. Natural convection is caused by buoyancy forces
due to density differences caused by temperature variations in the fluid. The overall heat
transfer is defined by Newton’s law of cooling

Q̇ = Ah(THot − TCold )

or
q̇ 00 = h(THot − TCold )
where q̇ 00 is given in units of [ m
W
2 ]. h is the convective heat transfer coefficient and given in
W
units of [ m2 ·K ]

The final mode of heat transfer is radiation, which transmits energy as electromagnetic
waves in space. The transfer does not need any medium and can happen in empty space,
but the emission (generation) of the radiation and the absorption do require a substance to
be present. More of radiation will be studied in ENGRMAE 120.

Sign Convention
Here is a brief summary of the sign convention.
• Heat transfer is positive
– When the heat enters the system.

38
• Heat transfer is negative
– When the heat exits the system.
• Work is positive
– When done by the system.
• Work is negative
– When done on the system.
In essence, when the system outputs work, then work is positive. When the system inputs
work, then work is negative.

The Energy Equation


Recall that the total energy E can be written as

E = U + KE + P E + other energies

where U is the internal energy and we will be looking mostly at problems that have to do
with internal energy.

For a control volume with a constant mass, a control mass, we express the conservation
of energy as a basic physical principle in a mathematical equation. This principle states
that you cannot create or destroy energy within the limits of classical physics. Therefore, if
the control volume has a change in energy, the change must be due to an energy transfer into
or out of the mass. This can only occur as work or heat transfers, and not as mass transfers
(since we are dealing with control mass). Writing this change in energy as an instantaneous
moment in time, we get the instantaneous rate form for energy,

dE
= Ė = Q̇ − Ẇ
dt
This equation gives the rate of change of the stored total energy as equal to the rate at which
energy is added minus the rate at which energy is removed. A process can move energy from
one place to another, but it cannot change the total energy.

39
In many cases, we are interested in finite changes from the beginning to the end of a process
and not focusing on the instantaneous rate of at which the process takes place. For these
cases, we integrate the energy equation with time from the beginning the process t1 to the
end of the process t2 by multiplying dt to get

dE = dU + d(KE) + d(P E) + d(other energies) = δQ − δW

and integrating Z
dE = E(t2 ) − E(t1 ) = E2 − E1

The right hand side terms are integrated as


Z Z Z
[Q̇ − Ẇ ]dt = δQ − δW = 1 Q2 − 1 W2
path path

Here the integration depends not only on the starting and ending states but also on the
process path in between; thus δQ is used instead of dQ to indicate the exact differential.
Basically, what this is saying is that energy is independent of the path (a.k.a. point functions
or exact differentials), which means that energy is a thermodynamic property. Finally, the
finite or total changes for energy is

E2 − E1 = 1 Q2 − 1 W2

The First Law of Thermodynamics


Consider a control mass where the substance inside goes through a cycle, in which the initial
and final states are the same. As the substance returns to the original state, there is no net
change in the control volume’s total energy and the rate of change is zero. The equation for
first law of thermodynamics is given as
I I
δQ = δW

H
The δQ is called
H cyclic integral of the heat transfer and represents the net heat transfer
in the cycle. δW is the cyclic integral of the work and represents the net work transfer
in the cycle.

The equation above could also be written as rates, where the integrals imply the summation
over all the boundaries of the control volume as

Cycle : Q̇net in = Ẇnet out

40
The first law of thermodynamics is a version of the law of conservation of energy, which
was in the previous section. So the energy equations can be adapted as the first law of
thermodynamics equations.

Here is a summary of the energy equations regarding first law of thermodynamics for a
closed system:
• For a process from 1 to 2.
E2 − E1 = 1 Q2 − 1 W2

• For multiple processes from 1 to n.


n
X n
X
En − E1 = Qj − Wj
j=1 j=1

• Rate form. Z Z
dE
( )sys = δ Q̇ − δ Ẇ
dt A A

– If kinetic energy and potential energies are negligible,


dU
( )sys = Q̇in − Ẇout
dt

Internal Energy - A Thermodynamic Property


Internal energy is an extensive property because it depends on the mass of the system. Ki-
netic and potential energies are also extensive. Recall that U designates regular internal
energy, while u is the specific internal energy and is intensive. Since specific internal energy
is intensive, it can be one of the independent properties to specify the state of a pure sub-
stance. For example, if we have specific internal energy and pressure of a substance, we can
use the steam tables to find the temperature.

The steam tables in the back of the book tabulates the internal energy of saturated liq-
uid as uf , internal energy of saturated vapor as ug , and the difference between the internal
energy of saturated liquid and saturated vapor as uf g . Also values for internal energy are
found in the steam tables in the same manner as for specific volume. In the liquid-vapor
saturation region,
U = Uliq + Uvap
or
mu = mf uf + mg vg
Dividing by m and introducing quality x gives

u = (1 − x)uf + xug

41
or
u = uf + xuf g
Also, if the system has only internal energy working on it (no kinetic, potential, etc.), then
the total energy equation is
E2 − E1 = U2 − U1
So the first law of thermodynamics energy equations can be used by using internal energy
instead of total energy. For example, E2 − E1 = U2 − U1 = 1 Q2 − 1 W2

Enthalpy - A Thermodynamic Property


Let us consider a control mass undergoing a quasi-equilibrium constant-pressure process, as
shown below.

Assume that there are changes only in the internal energy and that the only work done
during the process is that associated with the boundary movement. Taking the gas as our
control mass and applying the energy, we have

E2 − E1 = U2 − U1 = 1 Q2 − 1 W2

The work is given by Z 2


1 W2 = P dV = P (V2 − V1 )
1

with the pressure being constant (quasi-equilibrium). Therefore,

1 Q2 = U2 − U1 + P (V2 − V1 )

∆Q = ∆U + P ∆V
In this restricted (constant pressure) case, the heat transfer during the process is given in
terms of the change in the quantity U + P V between the initial and final states. Because all
these quantities (U , P , and V ) are thermodynamic properties, their combination must also
have the same characteristic of a thermodynamic property. Therefore, we can define a new
extensive property, the enthalpy.

H = U + PV

42
and specific enthalpy is,
h = u + Pv
Finally, the relationship between heat transfer and enthalpy is given by

1 Q2 = H2 − H1 = U2 − U1 + P (V2 − V1 )

Note: this is only true if the pressure remains constant during the process. However, this
does not mean that the use of enthalpy is only used when pressure is constant. Other cases,
such as control volume, will appear later. Enthalpy is introduced now so that enthalpy values
in the tables in the appendices can be used now.

The enthalpy of a substance in a saturation state and with a given quality is found the
same way as specific volume and internal energy.

h = (1 − x)hf + xhg

or
h = hf + xhf g
Also for substances in which the compressed-liquid tables are not given, the enthalpy is taken
as that of saturated liquid at the same temperature.

The Constant-Volume and Constant-Pressure Specific Heats


Specific heat is the amount of heat required per unit mass to raise temperature by one
degree. The amount of heat needed to change per unit mass by one degree depends on the
process, specifically two separate special cases.

1. Constant volume, for which ∂U = δQ − δW , but δW = 0 at constant volume, so the


specific at constant volume is

1  δQ  1  ∂U   ∂u 
Cv = = =
m δT v m ∂T v ∂T v
2. Constant pressure, for which ∂U = δQ − pdV , therefore δQ = ∂H, so the specific heat
at constant pressure is

1  δQ  1  ∂H   ∂h 
Cp = = =
m δT p m ∂T p ∂T p

Notice how in each of the cases, the formulas contain only thermodynamic properties. For
specific heat at constant volume, we have specific internal energy on the numerator and
temperature on the denominator. For specific heat at constant pressure, we have specific
enthalpy on the numerator and temperature on the denominator. Since they are all thermo-
dynamic properties, we can conclude that specific heat is also a thermodynamic property.

43
Also, more amount of heat is needed in the constant pressure system because some of the
heat is transferred into the work of moving the piston in the cylinder. In the constant volume
system, the heat is just being transferred to the substance in the rigid tank. Thereby, we
can conclude that for an isobaric system, the specific heat of a substance is larger than the
specific heat of the same substance in an isochoric system.

3. There is a special case for solids and liquids. Let’s first look at the enthalpy equation.

h = u + Pv

Differentiating,
dh = du + d(P v) = du + P dv + v dP
Both liquid and solid phases are nearly incompressible, so dv ≈ 0. Also, both phases have
small specific volume because they do not expand as much as vapor, thus v dP << du.
Finally, the equation becomes
C dT ≈ dh ≈ du
where C is either the constant-volume or the constant-pressure specific heat, as the two
would be nearly the same. And the change of enthalpy or internal energy for a solid or liquid
is
C(T2 − T1 ) ≈ h2 − h1 ≈ u2 − u1
In many processes involving a solid or a liquid, we might further assume that the specific
heat is constant, unless the process occurs at super low temperatures or over a wide range
of temperatures.

The Internal Energy, Enthalpy, and Specific Heat of Ideal Gases


In general, for any substance, the internal energy u depends on the two independent prop-
erties specifying the state. For a low-density (ideal) gas, however, u only depends on the
temperature T and much less on a second property. The dependence of u on P is less at low
pressure and is much less at high temperature: as density decreases, so does the dependence
of u on P or v. This is the same for enthalpy. For an ideal gas, h also depends on T .
Therefore, both internal energy and enthalpy is a function of temperature, for an ideal gas.

The relation between u and T can be established by using the constant-volume specific
heat  ∂h 
Cv =
∂T v
Because the internal energy is not a function of volume, we can drop the v and equation
becomes
du
Cv0 =
dT
du = Cv0 dT

44
where the subscript 0 denotes the specific heat of an ideal gas. For a given mass m,

dU = mCv0 dT

For enthalpy, it is
 ∂h 
Cp =
∂T p
Since the enthalpy of an ideal gas is a function of temperature and independent of pressure,
it is given
dh
Cp0 =
dT
dh = Cp0 dT
For a given mass m,
dH = mCp0 dT
Because the internal energy and enthalpy of an ideal gas are functions of temperature only, it
also follows that the constant-volume and constant-pressure specific heats are also functions
of temperature only. It is,
Cv0 = f (T ), Cp0 = f (T )
Because all gases approach ideal gas behavior as the pressure → 0, the ideal gas specific heat
for a given substance is often called the zero-pressure specific heat, and the zero-pressure,
constant pressure specific heat is given as the symbol Cp0 . The zero-pressure constant volume
specific heat is given the symbol Cv0 . The figure below shows Cp0 as a function of temperature
for a number of substances.

45
A very important relation between the constant pressure and constant volume specific heats
of an ideal gas may be developed from the definition of enthalpy.
h = u + P v = u + RT
Because P v = RT . Differentiating and substituting the internal energy and enthalpy equa-
tions for ideal gases,
dh = du + Rd T
Cp0 dT = Cv0 dT + R dT
Therefore,
Cp0 − Cv0 = R
This tells us that the difference between the constant-pressure and constant-volume specific
heats of an ideal gas is always constant, though both are functions of temperature.

Okay, now we have to consider some possibilities of the internal energy and enthalpy specific
heat equations for the ideal gases. I know. I know.. This stuff is really confusing so I will
try my best to make everything clear.

As you can see from the graph on the previous page, we can see that Cp0 is not always
constant and varies with temperature. That is why we must consider different possibilities
for the equations. The first possibility is the simplest, in which the specific heat Cp0 is
actually constant. Then we can simply integrate the equation
dh = Cp0 dT
to get
h2 − h1 = Cp0 (T2 − T1 )
and integrate
du = Cv0 dT
to get
u2 − u1 = Cv0 (T2 − T1 )
What about when the specific heats are not constant? This is the second possibility, in
which the specific heats are not constant. Then we still integrate the same equations, but
instead, they become
Z 2
h2 − h1 = Cp0 dT
1

and
Z 2
u2 − u1 = Cv0 dT
1

where Cp0 = f (T ) and Cv0 = f (T ).

46
How do we know when the specific heats are constants? We can only assume Cp0 and
Cv0 are constants
1. When the gas is mono-atomic.
2. When the gas is not mono-atomic, then the temperature range must not be too large
and infinitesimally small (Check the graph on the previous page).
Hope you understood everything because we are going to do some example problems now. I
will try my best to explain the concepts as we do the problems so everything ties in together.

Example (1)

A cylinder containing 1 kg of ammonia has an externally loaded piston. Initially the ammo-
nia is at 2 MPa and 180◦ C. It is now cooled to saturated vapor and 40◦ C and then further
cooled to 20◦ C, at which point the quality is 50%. Find the total work for the process,
assuming a piecewise linear variation of P versus V .

Solution (1)

Given: m = 1 kg of ammonia, P1 = 2 MPa, T1 = 180◦ C, T2 = 40◦ C, T3 = 20◦ C, and


x3 = 50%

In general, we need the pressures and volumes to solve for work. In order to solve for
the total work, we need the pressures and volumes for every state. Therefore, we should look
at this problem in three different states. Let’s look at state 1 first.

In state 1, we know that P1 = 2 MPa and T1 = 180◦ C. We have two independent and
intensive properties, so we can look at our thermodynamic table for ammonia.

47
As you can see, when P1 = 2 MPa and T1 = 180◦ C, v1 = 0.10571 m3 /kg.

Let’s look at state 2 now. At state 2, the problem tells us that it is cooled to saturated
vapor and T2 = 40◦ C.

According to the table, the saturation pressure is 1554.9 kPa when the temperature is 40◦ C,
which means Psat = P2 = 1554.9 kPa. Remember the thermodynamic tells us the saturation
pressure for the given temperature entries. If given pressure, then the table tells us the
saturation temperature for the given pressure. Since state 2 is at saturated vapor, we know
that state 2 lies on the saturated liquid-vapor line.
We also know that x2 = 1, since it is all saturated vapor. There is no saturated liquid so
v2 = vg = 0.08313 m3 /kg, according to the table.

Let’s look at state 3 now. At state 3, we know T3 = 20◦ C and x3 = 0.5. According to the
table, when T3 = 20◦ C, then P3 = 857.5 kPa, vf = 0.001638 m3 /kg and vg = 0.14922 m3 /kg.
To find specific volume for state 3, we do
v3 = (1 − x3 )vf + (x3 )vg = (1 − 0.5)(0.001638) + (0.5)(0.14922) = 0.07543 m3 /kg
Now we know all the pressures and volumes for each states. To find work, we do
WT otal = 1 W2 + 2 W3
Since we know that this is a linear process,
1 1
WT otal = (P1 + P2 ) m(v2 − v1 ) + (P2 + P3 ) m(v3 − v2 )
2 2

48
1 1
(2000+1554.9) (1)(0.08313−0.10571)+ (857.5+1554.9) (1)(0.07543−0.08313) = −49.4 kJ
2 2

Example (2)

A 400-L tank A contains argon gas at 250 kPa and 30◦ C. Cylinder B, having a friction-
less piston of such mass that a pressure of 150 kPa will float it, is initially empty. The valve
is opened, and argon flows into B and eventually reaches a uniform state of 150 kPa and
30◦ C throughout. What is the work done by the argon?

Solution (2)

Given: VA = 400 L, PA1 = 250 kPa, TA1 = 30◦ C, Plif t = 150 kPa, VB1 = 0 m3 ,
PA2 = PB2 = P2 = 150 kPa, and TA2 = TB2 = T2 = 3◦ C

To find work, we must find pressures and volumes, specifically PB1 , PB2 , and VB2 .

The substance we are dealing with is a gas, so let’s try to see if we can use the ideal
gas law. The critical pressure for argon is 4.87 MPa and both PA1 and P2 << Pc ; therefore,
we can use the ideal gas law equation. Let’s consider state 1 first.
PA1 VA = mRTA1
Let’s look at state 2 now.
P2 (VA + VB2 ) = mRT2
VA at state 2 is the same as VA at state 1 because the tank A’s volume is fixed. Also, we
notice that mRTA1 = mRT2 because m and R stay constant, and TA1 = T2 = 30◦ C. Thus,
PA1 VA − P2 VA (250 × 0.4) − (150 × 0.4)
PA1 VA = P2 (VA + VB2 ) → VB2 = = = 0.2667 m3
P2 150

49
Let’s try to find the pressure now. We notice that this is an isobaric process. How is
this constant pressure you ask? Well, it can be hard to wrap your head around this concept
if not explained, so here is the explanation.

While the piston moves, we assume that the piston is in equilibrium at every instantaneous
moment. Remember, quasi-equilibrium process? Quasi-equilibrium says that the piston is
in equilibrium when the piston moves infinitely slowly. Thus, at every instance, the force
net equation is
Fnet : AP0 + mpiston g = APlif t
No matter where the piston is in the cylinder or what the piston’s height is, as long as the
piston is floating, A, P0 , mpiston , and g are all constants. Thus Plif t must stay the same as
well in order for the piston to be in equilibrium. Thus, this process is an isobaric process:
Plif t = PB1 = PB2 .

Also note that the piston rises if and only if the pressure is 150 kPa. Therefore, once
PB1 approaches to Plif t , the volume will change, and a change of volume results in the pro-
cess of work.

The work formula for a constant pressure is

1 W2 = Plif t (VB2 − VB1 ) = 150(0.2667 − 0) = 40 kJ

Example (3)

Water in a piston/cylinder is at 100◦ C, x = 0.5 with mass 1 kg, and the piston rests on
the stops. The equilibrium pressure that will float the piston is 300 kPa. The water is
heated to 300◦ C by an electrical heater. At what temperature would all the liquid be gone?
Find the final (P,v) and the work in the process.

Solution(3)

Given: Water, T1 = 100◦ C, x1 = 0.5 m = 1 kg, Pf loat = 300 kP a, T2 = 300◦ C, and all
the liquid is gone at the final state.

Let’s divide the problem into two different states: state 1 and state 2.

50
Let’s look at state 1 (when the piston is resting on the stops). We know that T1 = 100◦ C,
x1 = 0.5 and m = 1 kg. Since we know the quality, we know that the water is at saturated
liquid-vapor state. Therefore, we can also deduce that Psat = P1 .

Looking at the steam table, we find that P1 = 101.3 kPa when T1 = 100◦ C. We also find
that vf = 0.001044 m3 /kg and vg = 1.67290 m3 /kg. Therefore, we can find

v1 = (1 − x1 )vf + x1 vg = 0.8369 m3 /kg

Now let’s look at state 2 (when the piston is floating). In order for the piston to float, we
need the pressure of the system to be at least 300 kPa. Since we are assuming that the
piston is floating, that means P2 = Pf loat .

51
If we look at the steam table (B.1.2), we see that when P2 = 300 kPa, the Tsat = 133.55◦ C.
Since T2 > Tsat , we know that state 2 must be in the superheated water vapor state.

Also, Tsat is when all the water becomes superheated water vapor a.k.a. when all the liquid
is gone.

To find the specific volume when P2 = 300 kP a and T2 = 300◦ C, we look at the super-
heated water vapor table (B.1.3).

v2 = 0.87529 m3 /kg Now that we have pressures and volumes at each state, we can find
the work done by the system. Recall that work is only done if and only if there is volume
change. The volume does not change when P1 = 101.3 kP a. The piston rises only when
P1 = Plif t . Thereby, this is an isobaric process.

1 W2 = P (v2 − v1 )m = 11.5 kJ

52
Here is the P − v diagram and notice how the volume is not changing when the pressure is
changing from 101.3 kPa to 300 kPa.

Example (4)

Two kilograms of water at 120◦ C with a quality of 25% has its temperature raised 20◦
in a constant volume process as in the figure below. What are the heat transfer and work?

Solution (4)

Given: m = 2 kg of water, T1 = 120◦ C, x1 = 0.25, T2 = 140◦ C, and isochoric/isometric


process.

Let’s look at state 1 first. Since we know quality at state 1, we know that state is in
saturated-liquid-vapor state and we also know Psat = P1 .

53
Looking at table B.1.1, when T = 120◦ C, Psat = P1 = 198.5 kPa, vf = 0.001060 m3 /kg,
vg = 0.89189 m3 /kg, uf = 503.48 kJ/kg, ug = 2529.24 kJ/kg

v1 = (1 − x1 )vf + x1 vg = 0.22376 m3 /kg

u1 = (1 − x1 )uf + x1 ug = 1009.92 kJ/kg


Now, let’s look at state 2. When T = 140◦ C, vf = 0.001080 m3 /kg, vg = 0.50865 m3 /kg.
Since vf < v2 < vg , the state is saturated-liquid-vapor. Thus Psat = P2 = 361.3 kPa and
uf = 588.72 kJ/kg, ug = 2550.02 kJ/kg.

We do not know x2 , so we cannot find u2 . However, we can find x2 by using v2 , vf and


vf g .
v2 = vf + x2 vf g
v2 − vf
x2 = = 0.4385
vf g
Finally
u2 = (1 − x2 )uf + uf g = 1448.838 kJ/kg

54
To find heat transfer, we must first find work. However, this is an isochoric process so there
is no change in volume. Therefore, work is zero. The energy equation is given by
E2 − E1 = U2 − U1 = 1 Q2 − 1 W2
Since we know the internal energies and work, we can find heat transfer.
1 Q2 = m(u2 − u1 ) + 0 = 877.83 kJ

Example (5)

A 10-m high cylinder, with a cross sectional area of 0.1 m2 , has a massless piston at the
bottom with water at 20◦ C on top of it, as shown below. Air at 300 K, with a volume of 0.3
m3 , under the piston is heated so that the piston moves up, spilling the water out over the
side. Find the total heat transfer when all the water has been pushed out.

Solution (5)

Given: l = 10 m, A = 0.1 m2 , TW 1 = 20◦ C, TA1 = 300 K, VA1 = 0.3 m3

Let’s look at state 1 of the water. We know that the total volume of the cylinder is V = 1 m3
because V = lA. So we can find VW 1 = V − VA1 = 0.7 m3 . Using volume and the density of
the water, we can find the mass of the water
mW 1 = ρVW 1 = 700 kg
and specific volume
VW 1
vW 1 = = 0.001 m3 /kg
mW 1
Let’s now look at state 1 of the air. Since it is air, we can use the ideal gas law (air is almost
always in ideal gas conditions). However, in order to use the ideal gas law, we need to know
either mass or pressure because we already have volume and temperature. We can find the
pressure by using the force net equation at the piston.
Fnet : APA1 = AP0 + mW 1 g

55
mW 1 g
PA1 = P0 +
A
ρVW 1 g
PA1 = P0 + = 168.6 kP a
A
Using the ideal gas law,
PA1 VA1
mA1 = = 0.58745 kg
RTA1
We can also find uA1 by using the ideal gas table (A7.1).

Thus, uA1 = 214.36 kJ/kg Now let’s move onto state 2 when all the water is gone. If we
write out the force net equation you can see that the pressure of air at state 2 is 100 kPa.

Fnet : APA2 = AP0

PA2 = P0 = 100 kP a
The volume is now V = VA2 = 1 m3 . The mass of air did not change so it is mA1 = mA2 =
0.58745 kg Using the ideal gas equation, we can find temperature of air at state 2
PA2 VA2
TA2 = = 593.12 K
mA2 R
Since the ideal gas table (A7.1) does not have specific values for when T = 593.12 K, we
must interpolate between 580 K and 600 K

56
593.12 − 580
uA2 = 419.87 + (435.10 − 419.87)( ) = 429.86 kJ/kg
600 − 580
Now we have all the values needed to find work and heat transfer. Let’s first find work. If
you recall the force net equation
ρVW 1 g
PA1 = P0 +
A
It can be written as
PA1 = A + BVW 1
since all the quantities except volume of water are constants. This force net equation tells
us that the process is linear. Therefore, we use the linear work equation.
1
1 W2 = (PA1 + PA2 )(VA2 − VA1 ) = 94.01 kJ
2
Heat transfer can be found by

1 Q2 = (u2 − u1 )m + 1 W2

1 Q2 = (uA2 − uA1 )mA2 + 1 W2 = 220.61 kJ

57
The First Law of Thermodynamics of a Control Volume
In the previous sections, we considered the first law of thermodynamics for a control mass
undergoing a change of state. Recall that a control mass is a fixed quantity of mass. Now
we move on to the mass of a system changing when its energy changes.

Conservation of Mass and the Control Volume


The surface of this control volume is the control surface that completely surrounds the
volumes. Mass as well as heat and work can cross the control surface, and the mass with its
properties can change with time. The figure below shows a schematic diagram of a control
volume that includes heat transfer, shaft work, moving boundary work, accumulation of
mass within the control volume, and mass flows. It is important to identify and label each
flow of mass and energy and the parts of the control volume that can store (accumulate)
mass.

Now that we are dealing with mass, we need to keep track of the mass, and thus consider the
conservation of mass law as it relates to control volume. The law says that mass cannot
be created nor destroyed.

We must consider all the mass that flows into and out of the control volume and the net
increase of the mass within the control volume. Consider the figure below, in which a tank
with a cylinder and piston and two pipes attached.

58
The contents inside a control volume can change between two states depending on how much
mass enters or leaves the control volume. Therefore, the change of mass in a control volume
between state 1 and state 2, or between two times, is given by

m2 − m1 = ∆(m)c.v. = min − mexit

This equation is the transient or the variation form of the mass conservation. However,
we must also look at the specific or instantaneous moment in time where the change of mass
is happening. That can be done by using the rate form of mass conservation.

dmc.v.
= ṁc.v. = ṁin − ṁexit
dt
This states that if the mass inside the control volume changes with time, it is because we
add some mass or take some mass out. There are no other means by which the mass inside
the control volume could change. The conservation of mass equation is commonly referred
to as continuity equation.

Now consider a fluid flowing through a pipe, as illustrated below.

If the contents entering or leaving a control volume, there must be a relative velocity between
the control volume and the substance at the control surface. Consider that such a surface has
a differential area (dA)n and the n is the unit normal pointing outwards, while the relative
velocity is Vn = vf luid − vc.s. , where v is the velocity of the fluid and the control surface. All
these quantities can be quantified as the mass flow rate, which tells us how fast the mass
of a substance is flowing in or out at a given time.
Z
ṁ = ρ Vn dA
c.s.

59
Thus, the mass flow rate can be written as

1
dṁ = ρ Vn dA = Vn dA
v
where ρ is the density, v is the specific volume, Vn is the velocity normal to the surface, and
dA is the normal cross sectional area.

The Energy Equation For a Control Volume


Consider the figure below, whenever a fluid mass enters a control volume at state 1 (into)
or exits at state 2 (exits), there is boundary movement work associated with that process.

In simpler words, the surroundings must do work to push the mass inside the control volume.
Conversely, the system must do work do push the mass outside of the control volume. The
amount of work needed to make the mass flow during the process is called flow work.

The velocity and the area correspond to a certain volume per unit time entering the control
volume, enabling us to relate that to the mass flow rate and the specific volume at the state
of the mass going in. Thereby, we are able to express the flow work as
Z
Ẇf low = F Vn = P Vn dA = P V̇ = P v ṁ

For the flow that leaves the control volume, work is being done by the control volume,
Pe ve ṁe , and for the mass that enters, the surroundings do the rate of work, Pi vi ṁi . Finally
the flow work rate is given by

δ Ẇf low = P v dṁ

60
Now that we have work for an open system (control volume), let’s move onto the topic
of energy equation. Recall that the energy equation for a closed system (control mass) is
 dE 
= Ėc.m. = Q̇ − Ẇ
dt c.m.

This equation has to be modified for a control volume, but how? Well, we just found out
that for an open system, so the flow work is involved in the new equation. Anything else?
Well, for an open system mass crosses the boundary, so mass is involved too. Let’s take a
look at the energy equation for a control volume.

dEc.v.
= Q̇ − Ẇ + ṁi ei − ṁe ee + Pi vi ṁi − Pe ve ṁe
dt
We can simplify this equation. Recall that e is forms of energy added together.

ei = ui + kei + pei + ...

Thus
ṁi (ei + Pi vi ) = ṁi (ui + kei + pei + Pi vi + ...)
and
hi = ui + Pi vi
This is the enthalpy equation and thus

ṁi (hi + kei + pei + ...)

ṁe (he + kee + pee + ...)


In many of our problems, we can neglect all the other energies except enthalpy. Finally, we
can write the energy equation as

dEc.v. X X
= Q̇c.v. − Ẇc.v. + ṁi hi − ṁe he
dt
The energy equation for open systems are given either as flow work or enthalpy.

If other types of energies are involved in the system, such as kinetic and potential, then
the energy equation is

dEc.v. X 1 X 1
= Q̇c.v. − Ẇc.v. + ṁi (hi + V2i + gZi ) − ṁe (he + V2e + gZe )
dt 2 2
where V is the velocity of the fluid and Z is the height of the fluid with respect to a reference
point.

61
The Steady-State Process
We are going to study steady state processes using the control volume equations. This is
suitable analytical model for long term steady operation of devices such as turbines, com-
pressors, nozzles, boilers, and condensers. This type of process does not include the short
term transient startup or shutdown of devices, just in the steady operating period of time.
For example, consider an air compressor that operates with a constant mass rate of flow into
and out of the compressor, constant properties at each point across the inlet and exit ducts,
a constant rate of heat transfer to the surroundings, and a constant power input. At each
point in the compressor, the properties are constant with time.

In a steady state process, the process occurs continuously and does not change in time,
which leads us to two assumptions.
1. The energy inside the control volume is always the same amount.
2. The mass inside the control volume is always the same amount.
These two assumptions lead us to
dEc.v. dmc.v.
= 0 and =0
dt dt
Therefore, we can conclude that for the steady-state processes we can conclude and write,

dEc.v. X 1 X 1
= 0 = Q̇c.v. − Ẇc.v. + ṁi (hi + V2i + gZi ) − ṁe (he + V2e + gZe )
dt 2 2
and X X
ṁi = ṁe

The energy of the steady state process is zero and the mass that flows into is equal to the
mass that flows out of the system.

Now let’s look at some common steady state devices; these devices are key words in problems
that indicate what is happening during the process. For example, if a problem says anything
about compressors, then we can go ahead and assume that q = 0 and pressure increases.
• Heat Exchangers (condensers, evaporators, etc): Exchange heat.
• Turbines: Produce power.
• Pumps: Increase pressure.
• Compressors: Increase pressure.
• Nozzles: Increase kinetic energy.
• Diffusers: Decrease kinetic energy.

62
• Throttling Devices: Decrease pressure.
If you go to the end of chapter 4 of the textbook, there will be Table 4.1 with typical steady
state/flow devices and it gives you a rundown of each purposes, conditions, and assumptions.
Here is preview of that table. Make sure to use this table when doing problems because it
can push you into the right direction of finding the answer.

The Transient Process


Now we are going to study transient processes using the control volume equations. The
transient process causes the states and the conditions to change with time and thus involves
an unsteady flow. This can be, for example, the filling or emptying of a closed tank with a
fluid where the stored mass and its state in the control volume change with time. Another
example is when filling a flat tire with air, the mass of air and its pressure increase as the
process goes on, and the process stops when a desired pressure is reached.

Similar to the steady state process, we want to look at the conservation of mass equation
and the energy equation. Let’s first look at the conservation of mass equation.
dmc.v.
= ṁi − ṁe
dt
This is the rate form of the equation. If we integrate it with respect to time, we would get
the integrated form.
m2 − m1 = mi − me
This integrate form of the conservation of mass equation will be applicable to many tran-
sient process problems; we will look at the system at the initial conditions and at the final
conditions.

Now, the energy equation is


dEc.v.
= Q̇c.v. − Ẇc.v. + ṁi (hi + kei + pei ) − ṁe (he + kee + pee )
dt

63
When we integrate the rate form of energy equation with respect to time, we get
Z 2 Z 2
E2 − E1 = 1 Q2 − 1 W2 + ṁi (hi + ...)dt − ṁe (he + ...)dt
1 1

If (hi + ...) and (he + ...) vary during the process, the integration of these two quantities can
get complicated. However, if (hi + ...) and (he + ...) of the fluid coming into and out of the
control volume do not change with time during the process, then they can be considered
constants and be taken out of the integrals. Finally, we get
X 1 X 1
E2 − E1 = 1 Q2 − 1 W2 + mi (hi + V2i + gZi ) − me (he + V2e + gZe )
2 2
The first law equation for the transient process is
X 1 X 1
E2 − E1 = Qc.v. − Wc.v. + mi (hi + V2i + gZi ) − me (he + V2e + gZe )
2 2
Where the integration of mass flow rates ṁ becomes total mass m of coming in and out.
This equation only works when we can prove that the energies of both states (h + ke + pe)
do not change with time. Always verify first before using the equation.

Example (1)

Liquid water is at 180◦ C, 2000 kPa is throttled into a flash evaporator chamber having
a pressure of 500 kPa. Neglect any change in kinetic energy. What is the fraction of liquid
and vapor in the chamber?

Solution (1)

From here on, I will stop providing the picture of the steam tables because I will assume
that if you have gotten this far into this course, you already know how to read them. Also,
I am too lazy.

Given: Ti = 180◦ C, Pi = 2000 kPa, Pe = 500 kPa, and the key words are flash evaporator
and throttled.

A flash evaporator chamber is shown in the figure above. The liquid water will flow in from
the left and then in the chamber, it will be evaporated into both liquid and vapor. The
vapor will flow out of the chamber at the top and then the liquid water will flow out at the
bottom.

64
This process is taking in steady state conditions because the problem does not give any
time variations or conditions. We are going to look at the inlet and the outlet state. We
define the outlet state when the water is coming out of the valve and not out of the chamber.
This is because we must look at the state of the water when it is in the chamber. If we define
the exit as the substance exiting the chamber then we can’t study what is happening inside
the chamber. Let’s look at state inlet (into).

When T = 180◦ C, Psat = 1002.2 kPa.

Pi > Psat → Compressed Liquid

Usually, if the state is in compressed liquid, we would normally approximate v ≈ vf at


that temperature. However, for these pressure and temperature values, there is data at the
compressed liquid table (B.1.4).

vi = 0.001127 m3 /kg

hi = 763.71 kJ/kg
For a steady state, we know two assumptions
X X
ṁi = ṁe

and X 1 X 1
0 = Q̇c.v. − Ẇc.v. + ṁi (hi + V2i + gZi ) − ṁe (he + V2e + gZe )
2 2
and we can ignore potential energy and kinetic energy, so
X X
0 = Q̇c.v. − Ẇc.v. + ṁi hi − ṁe he

In the steady flow table (4.1), it says that a flash evaporator chamber does no work and heat
transfer, so X X
0=0−0+ ṁi hi − ṁe he
ṁi hi = ṁe he → hi = he = 763.71 kJ/kg
Since the problem is asking for the fraction of liquid and vapor in the chamber, we can safely
assume that the state of the chamber is in saturated liquid vapor. Therefore, when Pe = 500
kPa, hf = 64021 kJ/kg and hf g = 2108.47 kJ/kg (B.1.2). We can find

ue = uf + xe uf g
ue − uf
xe = = 0.0586
uf g
V apor : xe = 0.0586 Liquid : 1 − xe = 0.941
V apor : 5.86% Liquid : 94.1%

65
Example (2)

A compressor receives 0.05 kg/s R-410a at 200 kPa, −20◦ C and 0.1 kg/s R-410a at 400
kPa, 0◦ C. The exit flow is at 1000 kPa, 60◦ C, as shown below. Assume it is adiabatic,
neglect kinetic energies, and find the required power input.

Solution (2)

Given: ṁ1 = 0.05 kg/s, P1 = 200 kPa, T1 = −200◦ C, ṁ2 = 0.1 kg/s, P2 = 400 kPa,
T2 = 0◦ C, T3 = 60◦ C, P3 = 1000 kPa

Examine state 1, which is the first inlet. When P1 = 200 kPa, T1 = −200◦ C, the state
is superheated vapor. Therefore, we can look at the B.4.2 table and we find that
h1 = 278.72 kJ/kg
Next look at state 2, which is the second inlet. When P2 = 400 kPa, T2 = 0◦ C, the state is
also superheated. Therefore, we still look at the B.4.2 table and find that
h2 = 290.42 kJ/kg
Then, look at state 3, which is the exit. When T3 = 60◦ C, P3 = 1000 kPa, the state is also
superheated vapor. According to the B.4.2 table,
h3 = 335.75 kJ/kg
Finally, we look at the steady state assumptions.
ṁ1 + ṁ2 = ṁ3
and since we can ignore the kinetic and other types of energies, we can write
X X
0 = Q̇c.v. − Ẇc.v. + ṁi hi − ṁe he

There is no heat transfer because it is adiabatic. The equation becomes


X X
Ẇc.v. = ṁi hi − ṁe he

66
Ẇc.v. = (ṁ1 h1 + ṁ2 h2 ) − (ṁ3 h3 ) = −7.385 kW

Example (3)

A heat exchanger, shown below, is used to cool an air flow from 800 to 360 K, with both
states at 1 MPa. The coolant is a water flow at 15◦ C, 0.1 MPa. If the water leaves as
saturated vapor, find the ratio of the flow rates ṁwater /ṁair .

Solution (3)

Given: T1 = 800 K, P1 = 1 MPa, T2 = 360 K, P2 = 1 MPa, T3 = T4 = 15◦ C, P3 = P4 = 0.1


MPa, x4 = 1

State 1 is air flowing in. Since it is air and we have the temperature, we can look at
table A7.1, the ideal gas tables. Our given value of temperature leads us to
h1 = 822.20 kJ/kg
State 2 is the air flowing out, so we still use A7.1 and find that
h2 = 360.86 kJ/kg
State 3 is the water flowing in, and with the given temperature and pressure, we find that
the water flowing in is in the state of compressed liquid. There is a compressed liquid table
in the appendices, however, the data is limited with our given values. Thus, we have to
approximates h3 ≈ hf (15◦ C)
h3 = 62.98 kJ/kg (B.1.1)
State 4 is water leaving the compressor. It is given to us that state 4 is saturated vapor,
thus x4 = 1. Thus we must look at B.1.1 and find that
h4 = hg (99.62◦ C) = 2675.46 kJ/kg
Finally, we look at the steady state assumptions. First, our conservation of mass assumption
as
ṁ1 + ṁ3 = ṁ2 + ṁ4
But this can be reduced to,
ṁ1 = ṁ2 = ṁair and ṁ3 = ṁ4 = ṁwater

67
Next, we can ignore all energies except enthalpy because kinetic energy and potential energy
are not happening in our control volume; we write our energy equation as
X X
0 = Q̇c.v. − Ẇc.v. + ṁi hi − ṁe he

Work and heat transfer are zero because according to the Typical steady flow devices (Table
4.1), we see that w = 0 and Q = 0. Thus,
X X
ṁi hi = ṁe he

and substituting the conservation of mass equation

ṁ1 h1 + ṁ3 h3 = ṁ2 h2 + ṁ4 h4

ṁ1 (h1 − h2 ) = ṁ4 (h4 − h3 )


Finally,
ṁw ṁ4 (h1 − h2 )
= = = 0.1766
ṁa ṁ1 (h4 − h3 )

Example (4)

A 25-L tank, shown below, that is initially evacuated is connected by a valve to an air
supply line flowing air at 20◦ C, 800 kPa. The valve is opened, and air flows into the tank
until the pressure reaches 600 kPa. Determine the final temperature and mass inside the
tank, assuming the process id adiabatic.

Solution (4)

Given: V = 25 L, TL = 20◦ C, PL = 800 kPa, P2 = 600 kPa

We can see right off the bat that this is a transient process problem because the tank is
being filled up and nothing is exiting. In control volume problems, especially in transient
processes, it is best to define your control volume before starting the problem. In this prob-
lem, our control volume is pretty obvious, it is the inside of our tank. However, in some
problems, it is not so obvious. Always define your control volume first.

68
In state 1 of the control volume, there is nothing inside the tank.

Let’s look at state 2, we only know that P2 = 600 kPa and V2 = V = 25 L. There is
not much we can do here.

Let’s look at the inlet state. What is entering? The air in the line. Thus, Pi = 800
kPa, Ti = 20◦ C = 293.15 K. We can find the enthalpy by looking at the ideal gas table
(A7.1). However, there are no exact values for when T = 293.15 K. Therefore, we must
interpolate between T = 290K and T = 298.15 K.
293.15 − 290
hi = 290.43 + (298.62 − 290.43) = 293.595 kJ/kg
298.15 − 290
Let’s look at transient process assumptions. First, the conservation of mass equation,

m2 − m1 = mi − me −→ m2 = mi

and next, the energy equation


X X
E2 − E1 = 1 Q2 − 1 W2 + mi hi + me he

However, there is no work because this is a rigid tank. There is no heat transfer because
this is adiabatic. There are no kinetic and potential energies involved in this problem. So
the energy equation is
X
E2 − E1 = U2 − U1 = 0 − 0 + mi hi + 0

There is also no internal energy in state 1 because the tank is initially empty. Therefore,

U2 = mi hi −→ m2 u2 = mi hi −→ u2 = hi

u2 = hi = 293.595 kJ/kg
We can find the temperature of the system in state 2 by using u2 and the A7.1 table. However,
we must interpolate again because there is no exact values for when u = 293.595 kJ/kg.
Interpolating between u = 286.49 kJ/kg and u = 301.04 kJ/kg, we get
293.595 − 286.49
T2 = 400 + (420 − 400) = 410 K
301.04 − 286.94
We can find m2 by using the ideal gas law.
P2 V2
m2 = = 0.1275 kg
RT2

69
The Second Law of Thermodynamics
What we have been doing so far all pertains to the the conservation of energy, which is the
first law of thermodynamics. Now we look at the second law of thermodynamics, which can
be quite confusing and hard to grasp in the beginning because we are going to be looking at
a lot of cycles in this section. The second law of thermodynamics acknowledges that
processes proceed in a certain direction but not in the opposite direction. For instance, a
hot cup of coffee cools by heat transfer to the surroundings, but heat will not flow into the
cup from the cooler surroundings.

Heat Engines and Refrigerators


Consider a device and a reservoir of energy at a high temperature TH . Suppose that this
machine is operating in a cycle and we take certain amount of energy from the high temper-
ature reservoir QH and the device produces work W . However, not all of the energy that
entered the device will be used to produce work. Thus, the leftover energy QL is thrown
away into a reservoir of energy at a low temperature. This device is called a heat engine.

Now consider another device and a reservoir of energy at a low temperature TL . Suppose
that the device is operating in a cycle and it takes certain amount of energy from the low
temperature reservoir QL . In order to receive this energy, work must be done on the device.
Then the leftover energy QH is thrown away into the high temperature reservoir. This device
is called a refrigerator or heat pump.

These devices will be commonly used to examine the second law of thermodynamics.

70
A heat engine may be defined as a device that operates in a thermodynamic cycle and
does a certain amount of net positive work through the transfer of heat from a high temper-
ature body to a low temperature body.

A refrigerator or heat pump is defined as a device that operates in a thermodynamic


cycle and requires a certain amount of net negative work through the transfer of heat from
a low temperature body to a high temperature body.

A thermal reservoir is a body to which and from which heat can be transferred indef-
initely without change in the temperature of the reservoir. Thus, they always remain at
constant temperature. An ocean and the atmosphere are good approximate examples. Also,
a reservoir from which is transferred is called a source, and a reservoir to which heat is
transferred is called a sink.

Note that we have departed from using the sign connotation for heat because for a heat
engine, QL is negative when the working fluid is considered as the system. It will be advan-
tageous to use the symbol QH to represent the heat transfer to or from the high temperature
body and QL to represent the heat transfer to or from the low temperature body. The
direction of heat transfer will be evident from the context and indicated by arrows.

Now let’s introduce the concept of thermal efficiency of a heat engine. In general, the
efficiency is the ratio of output (the energy sought) to input (the energy that costs), but he
output and input must be clearly defined.

Wnet (energy sought) QH − QL QL


ηthermal = = =1−
QH (energy that costs) QH QH

Heat engines and refrigerators can also be considered in general as energy conversion de-
vices, as the energy coming in is conserved but comes out in a different form. The thermal
efficiency is then a conversion efficiency for the process of going from the necessary input to
the desired output.

Now let’s look at the thermal efficiency of a refrigerator. However, we do not call it an
efficiency and instead we define it as coefficient of performance (COP), which we des-
ignate with the symbol β. For a refrigerator, the energy sought is QL , which makes sense
because we want cold temperature to store our food, and the energy that costs is the net
work Wnet .
QL (energy sought) QL 1
β= = =
Wnet (energy that costs) QH − QL QH /QL − 1
The ideal situation for a heat engine is 1 because ideally, we want all of QH to convert to
work output and no energy lost. On the other hand, the ideal situation for a refrigerator is
∞ because ideally, we want all of QL to convert to QH without any work input.

Note that a refrigerator or heat pump cycle can be used with either of two objectives.

71
It can be used as a refrigerator, which case the primary objective is QL . It can also be used
as a heating system, referred to as a heat pump, the objective is QH . Thus, the coefficient
of performance (COP) of a heat pump is different

QH (energy sought) QH 1
β0 = = =
Wnet (energy that costs) QH − QL 1 − QL /QH

and it follows that for a given cycle β 0 − β = 1 Unless otherwise specified, the term COP
will always refer to a refrigerator.

The Second Law of Thermodynamics


Okay, let us actually get to the topic of the second law. There are two classical statements
of the second law, known as Kelvin-Planck statement and the Clausius statement.
• The Kelvin-Planck Statement: It is impossible to construct a device that will
operate in a cycle and delivers a net amount of work while exchanging heat (heat)
with a single reservoir. Therefore, it is impossible to build a device that has a thermal
efficiency of 100% or 1.

• The Clausius Statement: It is impossible to construct a device that operates in a


cycle and produces no effect other than the transfer of heat from a cooler reservoir to
a warmer reservoir. Notice, how in this case, the first law of thermodynamics is not
violated; the heats are actually conserved. Likewise to the heat engine, it is impossible
to build a device that has a coefficient of performance of ∞.

72
Three observations should be made about these two statements. The first observation is that
both are negative statements. It is, of course, impossible to prove these negative statements.
However, we can say that the second law of thermodynamics rests on experimental evidence.
Every relevant experiment that has been conducted, either directly or indirectly, verifies the
second law, and no experiment has ever been conducted that contradicts second law. The
basis of the second law is therefore experimental evidence.

The second observation is that these two statements of the second law are equivalent. Two
statements are equivalent if the truth of either statement implies the truth of the other or if
the violation of either statement implies the violation of the other. That a violation of the
Clausius statement implies a violation of the Kelvin-Planck statement may be shown. Look
at the device below.

First, there is a hot reservoir and a cold reservoir, and in between, there is a refrigerator on
the left and a heat engine on the right. The refrigerator is removing energy QL from the cold
reservoir and delivering the same amount of energy QL to the hot reservoir. On the right,
the heat engine is taking a certain amount of energy QH from the hot reservoir, delivering
net amount of work, and rejecting some amount of energy QL to the cold reservoir. As you
can see, the refrigerator is violating the Clausius statement because there is no work input.
Next, the system boundary is defined to be the two devices and the cold reservoir and shown
by the dotted lines. The new device, created by the new boundary, is receiving a certain
amount of net heat from the high temperature reservoir and also delivering a net amount of
work. However, the cold reservoir does not exist anymore, since it is bounded by the bound-
ary. The new device now becomes a heat engine that violates the Kelvin-Planck statement
since there is no low temperature reservoir. This schematic shows the second observation: if
one statement is violated, the other statement is implicitly violated.

The third observation is that frequently, the second law of thermodynamics has been stated
as the impossibility of constructing a perpetual-motion machine of the second kind. A
perpetual-motion machine of the first kind would create work from noting or create
mass or energy, thus violating the first law. A perpetual-motion machine of the sec-
ond kind would extract heat from a source and then convert this heat completely into other
forms of energy, thus violating the second law.

73
The Reversible Process
A reversible process for a system is defined as a process that, once having taken place,
can be reversed, returning both the system and the surroundings to their original states.
However, there is no such a thing in reality. All real processes are irreversible. Often, we can
take the system back to the original state, but not the surroundings. Surroundings will not
change and not return to the original state. Some factors of why there cannot be a reversible
process can include friction, unrestrained expansion, mixing, heat transfer through a finite
temperature differences, finite time. Generally, the faster a process takes place, the more
irreversible the process will be; think about a paper burning so infinitely slowly versus a
paper exploding into million pieces in a matter of seconds.

The Carnot Cycle


Having defined the reversible process and considered some factors that make processes irre-
versible, let us move to the topic of an ideal cycle, officially called the Carnot cycle.

First, the Carnot cycle is reversible; every process could be reversed. Second, many dif-
ferent working substances can be used, such as a gas or substance with a phase change.
Third, there are also various possible arrangements of machinery. Finally, it is emphasized
that the Carnot cycle can, in principle, be executed in many different ways. The important
point to be made is that the Carnot cycle, regardless of what the working substance may
be, always have the same four basic processes between the high temperature reservoir and
the cold temperature reservoir. These processes are:
1. A reversible isothermal process in which heat is transferred to or from the high tem-
perature reservoir.
2. A reversible adiabatic process in which the temperature of the working substance de-
creases from the high temperature to the low temperature.
3. A reversible isothermal process in which heat is transferred to or from the low temper-
ature reservoir.
4. A reversible adiabatic process in which the temperature of the working substance in-
creases from the low temperature to the high temperature.
Here is an example of a gaseous system operating on a Carnot cycle with the four basic
processes in one of many different ways.

74
From process 1 to 2, there is an isothermal expansion where the gas heats up constantly
and the piston slowly ascends. From process 2 to 3, there is an adiabatic expansion, where
the heat transfer stops but continues to expand. Then from process 3 to 4, there is an
isothermal compression, where the gas cools down constantly and the piston slowly descends.
From process 4 to 1, there is an adiabatic compression, where the heat transfer stops but
continues to compress. Then the cycle repeats. If we were to plot this process in a P − v
diagram, it would be this

As you can see, from 1 to 2 is on the isothermal line TL and from 3 to 4 is on the isothermal
line TH , which both correspond to the gas system. Although not shown on the graph, the
work done is the area between the top and bottom lines. The Carnot cycle will be used to
compare other cycles we will study later. Basically, the Carnot cycle is our reference cycle.

Two Propositions Regarding the Efficiency of a Carnot cycle


There are two important propositions regarding the efficiency of a Carnot cycle.
• Proposition 1: For given, fixed high temperature reservoir and low temperature reser-
voir, a reversible heat engine is more efficient than any irreversible heat engines.

ηrev ≥ ηany

• Proposition 2:For given, fixed high temperature reservoir and low temperature reser-
voir, all Carnot heat engine have the same efficiency.

ηrev1 = ηrev2

75
According to this figure, with a fixed high temperature reservoir and low temperature reser-
voir, the work done by the reversible engine is greater than the work done by the irreversible
engine (Wrev > Wirrev ), which makes sense because an ideal engine will produce more work
than that of a non-ideal engine because of its greater efficiency.

The Thermodynamic Temperature Scale


The two propositions of the Carnot cycle is implying that the efficiency of a Carnot cycle
depends only on the temperatures because the temperature reservoirs are fixed. Therefore,
we need provide a basis for temperature measurement. The temperature scale must be
defined in terms of a particular thermometer substance and device. A temperature scale
that is independent of any particular substance is called an absolute temperature scale.
The facts that a Carnot cycle is independent of the working substance and depends only on
the reservoir temperatures provide the basis for such an absolute temperature scale called the
thermodynamic scale. Since the efficiency of a Carnot cycle is a function only of temperature,
it is
QL
ηthermal = 1 − = 1 − ψ(TL , TH )
QH
which ψ designates a functional relation.

For simplicity, the thermodynamic scale is defined as


QH TH
=
QL TL
Substituting this definition of the thermodynamic scale into the thermal efficiency of a Carnot
cycle results in the thermal efficiency of a Carnot cycle and the absolute temperatures of the
two reservoirs
Wnet QL TL
ηthermal = =1− =1−
QH QH TH

76
If we make the same substitution of the thermodynamic scale into the COP refrigerator
equation, we get
QL 1 TL
β= = =
Wnet TH /TL − 1 TH − TL
The COP of the heat pump is given as well

QH TH
β0 = =
Wnet TH − TL

Note: the definition of the thermodynamic scale only holds for a Carnot cycle and not a
real cycle. Thus, the three derived equations listed above are only for Carnot heat engines,
Carnot refrigerators, and Carnot heat pumps.

Ideal Cycles
Here are some ideal and reversible cycles for you be familiarized with them. We will be
mostly looking at the Carnot and Refrigeration cycles in this course. The other two will be
studied in the later thermodynamics course, ENGRMAE 115.
• Carnot:
– 2 isothermal processes.
– 2 adiabatic processes.
• Brayton (Gas):
– 2 isobaric processes.
– 2 adiabatic processes.
• Rankine (Liquid-Vapor):
– 2 isobaric processes.
– 2 adiabatic processes.
• Refrigeration (Vapor Compression):
– 2 isobaric processes.
– 2 adiabatic processes.

77
Property Relations in a Carnot Cycle
This section is dedicated to proving that the definition of the thermodynamic scale
QH TH
=
QL TL
Consider that we have an ideal gas Carnot cycle.

We will be using the first law of thermodynamics and the ideal gas law for proof.
du = δq − δw
P v = RT
Using the internal energy change
du = Cv0 dT
Substituting the internal energy change and the ideal gas law equations into the first law,
we get
RT
δq = Cv0 dT + dv
v
We now proceed to integrate this first law equation for each of the four processes that make
up the Carnot cycle. For the isothermal heat addition process 1-2, we have
v2
qH = 1 q2 = 0 + RTH ln
v1
or
v2
qH = RTH ln
v1
For the adiabatic expansion process 2-3, we divide by T to get,
Z TL
Cv0 v3
0= dT + R ln
TH T v2

78
For the isothermal heat rejection process 3-4,
v4
qL = −3 q4 = −0 − RTL ln
v3
or
v3
qL = RTL ln
v4
and for the adibatic compression process 4-1 we divide by T to get,
Z TH
Cv0 v1
0= dT + R ln
TL T v4

From the two adiabatic processes (2-3) and (4-1), we get


Z TH
Cv0 v3 v1
= R ln = −R ln
TL T v2 v4

Therefore,
v3 v4 v3 v2
= or =
v2 v1 v4 v1
Finally,
qH RTH ln v2 /v1 TH
= =
qL RTL ln v3 /v4 TL
For a reversible (ideal) cycle.
qH TH
=
qL TL
or
QH TH
=
QL TL

Ideal Versus Real Machines


Any real heat engine, refrigerator, or heat pump will be less efficient than the Carnot counter-
parts.
• Real Heat Engine ≤ Carnot Heat Engine:

QL TL
ηreal thermal = 1 − ≤1−
QH TH
• Real Refrigerator ≤ Carnot Refrigerator

QL TH
βreal = ≤
QH − QL TH − TL

79
• Real Heat Pump ≤ Carnot Heat Pump

0 QH TH
βreal = ≤
QH − QL TH − TL
The less than equality sign is the definition with the use of the energy equation and thus is
always true. The equal to equality sign is valid only if the cycle is reversible (Carnot).

80
Example (1)

60 kg per hour of water runs through a heat exchanger, entering as saturated liquid at 20
kPa and leaving as saturated vapor. The heat is supplied by a heat pump operating from a
low temperature reservoir at 16◦ C with a COP of half that of the similar Carnot unit. Find
the rate of work into the heat pump.

Solution (1)

Given: ṁ = 60 kg/hr = 0.01667 kg/s, P1 = P2 200 kPa, State 1 → SAT-LIQ, State 2


→ SAT-VAP, TL = 16◦ C = 289 K, βreal
0 0
= βCarnot /2, Heat Pump, Heat exchanger.

Let’s draw the schematic for this problem first. It says that the water flows into and out of
a heat exchanger. Also, a heat pump supplies heat Q̇H to the heat exchanger. So, draw the
heat pump and heat exchanger together.

Next, let’s write out the steady state assumptions (constant mass and energy equations) of
the heat exchanger since mass is flowing in and out.

ṁi = ṁe −→ ṁ1 = ṁ2

0 = Q̇ − Ẇ + ṁ1 h1 − ṁ2 h2
Q̇ = ṁ2 h2 − ṁ1 h1
This heat transfer that is actually Q̇H because it is the heat transfer from the heat pump to
the heat exchanger.

We have to find the h1 and h2 , but notice how the pressure at state 1 is the same as
pressure at state 2 because it just goes from saturated liquid to saturated vapor. Therefore,
h1 = hf and h2 = hg

Q̇ = ṁ2 h2 − ṁ1 h1 = 0.01667(2706.63 − 504.68) = 36.7 kJ/s

81
Now, recall that the COP of a real and carnot heat pump are

0 Q̇H
βreal =
Q̇H − Q̇L
and
0 TH
βCarnot =
TH − TL
Therefore,
0 0
βreal = βCarnot /2

0 Q̇H TH /(TH − TL )
βreal = =
Ẇnet 2
Since pressure at state 1 and pressure at state 2 did not change, that would mean the
temperature did not change as well. Therefore, TH = 120.23◦ C.

Q̇H
Ẇ = TH /(TH −TL )
= 19.45 kW
2

Example (2)

A car engine operates with a thermal efficiency of 35%. Assume the air conditioner has
a COP of β = 3 working as a refrigerator cooling the inside using engine shaft work to drive
it. How much extra fuel energy should be spent to remove 1 kJ from the inside?

Solution (2)

Given: Heat engine, η = 35%, refrigerator, β = 3

Draw the schematic first. The fuel energy is burned and then used in the heat engine
to provide work to the air conditioner, so QH of the heat engine is Qf uel .

82
Let’s start with the heat engine first.
Qf uel − QL eng Wnet
0.35 = =
Qf uel Qf uel
Wnet = 0.35Qf uel
Now let’s do the AC.
QL QL
3= =
QH − QL Wnet
QL 1
Wnet = = kJ
3 3
Now that we have work, which is the same work supplied by the heat engine, we can find
the heat transfer from the fuel energy.
Wnet
Qf uel = = 0.9523 kJ
0.35

The Second Law of Thermodynamics of a Control Mass


Up to this point in our topic of second law of thermodynamics, we have dealt only with
thermodynamic cycles. However, we are often concerned with processes instead of cycles. In
the topic of the first law, properties, such as internal energy, enthalpy, etc., were introduced
and enabled us to use the first law quantitatively for processes. Similarly, entropy will be
used to treat the second law quantitatively.

The Inequality of Clausius


The first step in our consideration of the property, we call entropy, is to establish the in-
equality of Clausius. I
δQ
≤0
T
The inequality of Clausius is a result of the second law of thermodynamics. It will be
demonstrated to be valid for all possible cycles, including both reversible and irreversible
heat engines and refrigerators. Since any reversible cycle can be represented by a series of
Carnot cycles, in this analysis we consider only a Carnot cycle that leads to the inequality
of Clausius.

First, consider a reversible (Carnot) heat engine cycle operating between reservoirs at tem-
peratures TH and TL , as shown below.

83
H
For this cycle, the cyclic integral of the heat transfer δQ is greater than zero.
I
δQ = QH − QL > 0

Since TH and TL are constants, from the definition of the absolute temperature scale and
from the fact that this is a reversible cycle, it follows that
I
δQ QH QL
= − =0
T TH TL
Therefore, for Carnot (reversible) heat engines
I
δQ ≥ 0

and I
δQ
=0
T

Now consider an irreversible cyclic heat engine operating between the same TH and TL
as the reversible heat engine and receiving the same quantity of heat QH . Comparing the
irreversible cycle with the reversible one, we conclude from the second law that

Wirr < Wrev

Since QH − QL = W for both reversible and irreversible cycles, we conclude that

QH − QL irr < QH − QL rev

and therefore
QL irr > QL rev
As a result, I
δQ = QH − QL irr > 0

84
and I
δQ QH QL irr
= − <0
T TH TL
Finally we conclude that for irreversible (real) heat engines
H
δQ ≥ 0

and I
δQ
<0
T
H
If we combine the two δQ/T into one for all cycles, we get the inequality of Clausius
equation. I
δQ
≤0
T

Entropy - A Property of a System


By applying the inequality of Clausius equation, we can demonstrate that the second law
of thermodynamics leads to a property of a system called entropy. Let a system (control
mass) undergo a reversible process from state 1 to state 2 along a path A, and let the cycle
be completed along path B, which is also reversible.

Because this is a reversible cycle, we can write


I Z 2  Z 1 
δQ δQ δQ
=0= +
T 1 T A 2 T B
Now consider another reversible cycle, which proceeds first along path C and is then com-
pleted along path B. For this cycle, we can write
I Z 2  Z 1 
δQ δQ δQ
=0= +
T 1 T C 2 T B
Subtracting the second equation from the first, we get
Z 2  Z 2 
δQ δQ
=
1 T A 1 T C

85
H
Since δQ/T is the same for all reversible paths between states 1 and 2, we conclude that
this quantity is independent of the path and is a function of the end states only; therefore,
a property. This property is called entropy and is designated with a S. It follows that
entropy may be defined as a property of a substance in accordance with the relation
 δQ 
dS =
T rev
Entropy is an extensive property, and the intensive entropy is designated as s. It is important
to note that entropy is defined here in terms of a reversible process. The change in the entropy
of a system as it undergoes a change of state may be found by integrating and thus,
Z 2 
δQ
∆Sc.m. = S2 − S1 =
1 T rev

Entropy of a Pure Substance


Values of specific entropy are given in tables of thermodynamic properties in the same man-
ner as specific volume and specific enthalpy. The units of specific entropy in the steam tables,
refrigerant tables, and ammonia tables, are kJ/kgK.

In the saturation region, the entropy may be calculated using quality


s = (1 − x)sf + xsg = sf + xsf g
The entropy of a compressed liquid is tabulated in the same manner as the other properties.
These properties are primarily a function of the temperature and are not greatly different
from those for saturated liquid at the same temperature.

The thermodynamic properties of a substance are often shown on a temperature-entropy


diagram and on an enthalpy-entropy diagram, which is also called a Mollier diagram.

86
The T − s diagram is useful when we are dealing with mixture of liquid and vapor.

The h−s diagram is most useful when we are dealing with superheated vapor or high quality
mixtures.

Entropy Change in Reversible Processes


In this section, we consider the various processes. We will limit ourselves to systems that un-
dergo reversible processes and consider the Carnot cycle, reversible heat-transfer processes,
and reversible adiabatic processes.

Let the working fluid of a heat engine operating on the Carnot cycle make up the sys-
tem. The first process is the isothermal transfer to heat to the working fluid from the
high-temperature reservoir. For this process we write
Z 2 
δQ
S2 − S1 =
1 T rev
Since this is a reversible process in which the temperature of the working fluid remains
constant, the equation can be integrated to give
Z 2
1 1 Q2 QH
S2 − S1 = δQ = =
TH 1 TH TH
This process is shown in the figure below and the area under the line 1-2 represents the heat
transferred to the working fluid during the process.

87
The second process of a Carnot cycle is a reversible adiabatic one. From the definition of
entropy,
 δQ 
dS =
T rev
it is evident that the entropy remains constant in a reversible adiabatic process. A constant
entropy process is called an isentropic process (Note: a process is isentropic only if the
process is reversible and adiabatic). Line 2-3 represents this process, and this process is
concluded at state 3 when the temperature of the working fluid reaches TL .

The third process is the reversible isothermal process in which heat is transferred from
the working fluid to the low-temperature reservoir. For this process we write
Z 4 
δQ 3 Q4 QL
S4 − S3 = = =
3 T rev TL TL
because during this process, the heat transfer is negative (entropy decreases). Furthermore,
because the final process 4-1, which completes the cycle, is a reversible adiabatic process, it
is evident that the entropy decrease in process 3-4 must exactly equal the entropy increase in
process 1-2. The area under line 3-4 represents the heat transferred from the working fluid
to the low temperature reservoir.

Since the net work of a cycle is equal to the net heat transfer, area 1-2-3-4-1 must rep-
resent the net work of a cycle. The efficiency of the cycle may also be expressed in terms of
areas
Wnet area 1 − 2 − 3 − 4 − 1
ηT h = =
QH area 1 − 2 − b − a − 1

If the cycle is reversed, as shown in the figure below, we have a heat pump/refrigerator.

88
Notice that the entropy of the working fluid increases at TL , since heat is transferred to the
working fluid at TL . The entropy decreases at TH because of heat transfer from the working
fluid.

Let’s reiterate the formulas regarding the T − s diagram we just saw above.
• For some reversible process that transfers heat to the working fluid (heat addition).

QH = TH ∆S

• For some reversible process that transfers heat from the working fluid (heat rejection).

QL = TL ∆S
• The net work is the area delimited by the cycle path.

Wnet = (TH − TL )∆S

The Thermodynamic Property Relation


The first law for a change of state under these conditions can be written
δQ = dU + δW
For a reversible process of a simple compressible substance, we write
δQ = T dS and δW = P dV
Substituting these relations into the energy equation, we get

T dS = dU + P dV
We can call this a thermodynamic property relation because everything in this equation is
a property. The variables in this equation do not depend on any paths and instead only on

89
the initial and final states. Note that this equation was derived by assuming a reversible
process. This equation can therefore be integrated for any reversible process, for during such
a process the state of the substance can be identified at any point during the process. This
equation is also often applied to an irreversible process between two given states, but the
integration of the equation is performed along a path between the two same states.

The enthalpy is defined as


H = U + PV
and differentiating this is
dH = dU + P dV + V dP
Substituting these relations into the property relation equation above, we get

T dS = dH − V dP
These two expressions we derived are often called the Gibbs equations.

Entropy Change of a Solid or Liquid


In the specific heat section, we learned that it is possible to express internal energy and
enthalpy in terms of the specific heat. We can now use this result and the thermodynamic
property relation to calculate entropy change for a solid or liquid. (incompressible substance).
Using the first thermodynamic property relation,
T ds = du + P dv
Since incompressible substances cannot change in volume, we write
T ds = du
Dividing by the temperature,
du C
ds ' ' dT
T T
The specific heat constant remains constant for incompressible substances and integrating
this gives us
T2
s2 − s1 ' C ln
T1

Entropy Change of an Ideal Gas


Let’s now derive an expression for entropy change of an ideal gas. Let’s first start with the
thermodynamic relation equation.
T ds = du + P dv
For an ideal gas,
P R
du = Cv0 dT and =
T v

90
Therefore,
dT Rdv
ds = Cv0 +
T v
Z 2
dT v2
s2 − s1 = Cv0 + R ln
1 T v1
Similarly,
T ds = dh − vdP
For an ideal gas
v R
dh = Cp0 dT and =
T P
Therefore,
dT RdP
ds = Cp0 −
T P
Z 2
dT P2
s2 − s1 = Cp0 − R ln
1 T P1

If assume that the specific heat constants, Cv0 and Cp0 , are constants, then the two equations
will be
T2 v2
s2 − s1 = Cv0 ln + R ln
T1 v1
and
T2 P2
s2 − s1 = Cp0 ln − R ln
T1 P1

Another term to consider is standard entropy. Standard entropy refers to the result
of the integration from reference temperature T0 to any other temperature T for Cp0 as a
function of temperature. Z T
0 Cp0
ST =
T0 T

This function can be tabulated in the single entry (temperature) ideal-gas table, for example
air table A7 and other gas table A8. The entropy change between any two states 1 and 2 is
then given by
P2
s2 − s1 = s0T 2 − s0T 1 − R ln
P1

Let us now consider the case of an ideal gas undergoing an isentropic process. Recall that if
a process is both adiabatic and reversible, the process is isentropic.
 dQ 
ds =
T rev

91
Adiabatic → dQ = 0 → ds = 0. Therefore,
T2 v2
s2 − s1 = 0 = Cv0 ln + R ln
T1 v1
and
T2 P2
s2 − s1 = 0 = Cp0 ln − R ln
T1 P1
For the first entropy change expression, we can rewrite it as
 T Cv0  v R
2 2
ln + ln =0
T1 v1
Simplifying it further by taking the natural log of each terms,
 T Cv0  v R
2 2
=1
T1 v1
Also recall that for an ideal gas,
R = Cp0 − Cv0
and let
Cp0
k=
Cv0
Finally, we can write
T2  v1 k−1
=
T1 v2
Likewise, if we take the second entropy change of an ideal gas equation and do similar steps,
we get
T2  P2  k−1
k
=
T1 P1
If we were to set these two new expressions together, we would get a new expression in terms
of pressure and volume.
P2  v1 k
=
P1 v2
Does this expression look familiar? It should because it is the polytropic process.

(P2 v2 )k = (P1 v1 )k

P v k = constant
This is a polytropic process with n = k.

Here are P − v and T − s graphs with some of the common polytropic processes. For
example n = 0, n = ∞, n = 1, and now with entropy n = k.

92
Isobaric process: n = 0, P = constant

Isothermal process: n = 1, T = constant

Isentropic process: n = k, s = constant

Isochoric process: n = ∞, v = constant

You should memorize the k for air, and it is k = 1.4. Anyway, all the k values of ideal gases
are in table A.5.

Entropy Change of a Control Mass During an Irreversible Process


Consider a control mass that undergoes the cycles shown below.

The cycle made up of the reversible process A and B is a reversible cycle. Therefore, we can
write by applying the inequality of Clausius
I Z 2  Z 1 
δQ δQ δQ
= + =0
T 1 T A 2 T B

93
The cycle made up of the irreversible process C and the reversible process B is an irreversible
cycle. Therefore, for this cycle the inequality of Clausius may be applied, giving the result
I Z 2  Z 1 
δQ δQ δQ
= + <0
T 1 T C 2 T B
Subtracting the second equation from the first and rearranging, we have
Z 2  Z 2 
δQ δQ
>
1 T A 1 T C
Since path A is reversible, and since entropy is a property,
Z 2  Z 2
δQ
= dSA = S2 − S1
1 T A 1

Therefore, Z 2  δQ 
S2 − S1 >
1 T C

In order to make this inequality equal (> → =), we have to add something on the right hand
side. We add something called entropy generation.
Z 2  δQ 
S2 − S1 = + Sgen
1 T C

The entropy of a system can decrease if the heat being removed is sufficiently large such
R 2  δQ 
that the is 1 T + Sgen < 0. However, the entropy generation is always positive for real
C
(irreversible) processes and zero for ideal (reversible) processes.

Example (1)

A piston/cylinder setup contains air at 100 kPa, 400 K that is compressed to a final pressure
of 1000 kPa. Consider two different processes: (a) a reversible adiabatic process and (b) a
reversible isothermal process. Show both processes in P − v and a T − s diagram. Find the
final temperature and the specific work for both processes.

Solution (1)

Given: P1 = 100 kPa, T1 = 400 K, P2 = 1000 kPa

Let’s start with the reversible-adiabatic process. Recall that when a process is reversible
and adibatic, the process is isentropic.

Start with the first law.


E2 − E1 = U2 − U1 = 1 Q2 − 1 W2

94
mtot (u2 − u1 ) = mtot 1 q2 − mtot 1 w2
The process is adibatic so there no is q.

1 w2 = u1 − u2

Look at state 1 of the system. Since the working fluid is air, we can use the ideal gas tables
(A7.1) to find values for the given temperature.

u1 = 286.49 kJ/kg

s0T 1 = 7.15926 kJ/kg K


Look at state 2 of the system. Not much we can do about state 2 since we only know the
pressure.

Look at the entropy change of the system now.


P2
s2 − s1 = s0T 2 − s0T 1 − R ln
P1
We know all the values here, so we can find the standard entropy for state 2.
P2
s0T 2 = s0T 1 + R ln
P1
s0T 2 = 7.82010 kJ/kg K
We can use this standard entropy value in table A7.1 to find T2 and u2 . However, there is
no exact value for this standard entropy, so we must interpolate between 740 K and 760 K.
7.8201 − 7.80008
T2 = 740 + (760 − 740)
7.82905 − 7.80008
T2 = 753.8 K
and
7.8201 − 7.80008
u2 = 544.33 + (560.32 − 544.33)
7.82905 − 7.80008
u2 = 555.38 kJ/kg
Now we can finally find specific work.

1 w2 = u1 − u2 = −268.89 kJ/kg

Let’s look at the reversible-isothermal process. Recall that isothermal means the tempera-
ture is constant.

Start with the first law again.

E2 − E1 = U2 − U1 = 1 Q2 − 1 W2

95
mtot (u2 − u1 ) = mtot (1 q2 − 1 w2 )
Since the process is isothermal, the internal energy of the system stays constant.
u2 = u1
Therefore,
1 w2 =1 q2
Look at state 1 of the system. Again, we can find the values in the ideal gas table (A7.1)
for the given temperature value since the working fluid is air.
u1 = 286.49 kJ/kg
s0T 1 = 7.15926 kJ/kg K
Look at state 2 of the system.
T2 = T1 = 400 K
u2 = u1 = 286.49 kJ/kg
s0T 2 = s0T 1 = 7.15926 kJ/kg K
Look at the entropy change of the system.
P2
s2 − s1 = s0T 2 − s0T 1 − R ln
P1
P2
s2 − s1 = 0 − R ln
P1
s2 − s1 = −0.66084 kJ/kg K
We can use the entropy change to find heat. Recall this formula.
 δQ 
dS =
T rev
1 Q2 = T (S2 − S1 )
1 q2 = T (s2 − s1 ) = (400)(−0.66084) = −264.336 kJ/kg
Finally,
1 w2 =1 q2 = −264.336 kJ/kg
The P − v and T − s graphs are shown below.

96
Example (2)

Two rigid, insulated tanks are connected with a pipe and valve. One tank has 0.5 kg air at
200 kPa, 300 K and the other has 0.75 kg air at 100 kPa, 400 K. The valve is opened and
the air comes to a single uniform state, without external heat transfer. Find the final T and
P and the entropy generation.

Solution (2)

Given: mA = 0.5 kg, PA1 = 200 kPa, TA1 = 300 K, mb = 0.75 kg, PB1 = 100 kPa,
TB1 = 400 K, insulated tanks, air.

Start with the first law of the entire system.


E2 − E1 = U2 − U1 =1 Q2 −1 W2
There are no work and heat transfer, since they are rigid, insulated tanks.
U2 = U1
mtot u2 = mA uA1 + mb uB1
Look at state 1 of tank A. We can use the ideal gas table (A7.1) to find some values for the
given temperature.
uA1 = 214.36 kJ/kg
s0T A1 = 6.86926 kJ/kg K
and
mA RTA1
PA1 VA = mA RTA1 −→ VA = = 0.21525 m3
PA1
Look at state 1 of tank B.
uB1 = 286.49 kJ/kg
s0T B1 = 7.15926 kJ/kg K
and
mB RTB1
PB1 VB = mB RTB1 −→ VB = = 0.861 m3
PB1
Look at state 2 of the uniform system.
mA uA1 + mB uB1
u2 =
mtot

97
Plug all the values in to find that
u2 = 257.638 kJ/kg
Use this internal energy to find the temperature of the uniform system, but interpolate
between 360 K and 380 K. You know why we interpolate.
257.638 − 257.53
T2 = 360 + (380 − 360)
271.99 − 257.53
T2 = 360.15 K
Use the ideal gas law to find P2 .
mtot RT2 (mA + mB )RT2
P2 = =
Vtot (VA + VB )
P2 = 120.05 kP a
Now we must find entropy generation. Recall that the equation for entropy change, which
includes entropy generation is
Z 2 
δQ
S2 − S1 = + Sgen
1 T C
Since the process is adibatic, the equation becomes
S2 − S1 = Sgen
mtot s2 − (mA sA1 + mB sB1 ) = Sgen
(mA + mB )s2 − mA sA1 − mB sB1 = Sgen
Rearrange values so that the mA values are grouped together and mB values are grouped
together.
mA (s2 − sA1 ) + mB (s2 − sB1 ) = Sgen
T2 P2 T2 P2
mA (Cp0 ln − R ln ) + mB (Cp0 ln − R ln ) = Sgen
TA1 PA1 TB1 PB1
Plug all the values in.
Sgen = 0.0466 kJ/kg

The Second Law of Thermodynamics of a Control Volume


Just as we did the first law for a control volume, we will do the second law for a control
volume. We start with the second law expressed as a change of entropy for a control mass
in a rate form
 dS  X Q̇
= + Ṡgen
dt c.m. T
to which we now will add the contributions from the mass flow rates into and out of the
control volume. A simple example of such a situation is illustrated below.

98
The flow of mass does carry an amount of entropy, s, per unit mass flowing, but it does not
give rise to any other contributions. As a process may take place in the flow, entropy can
be generated, but this is attributed to the space it belongs to (either inside or outside of the
control volume).

The balance of entropy as an equation then states that the rate of change in total en-
tropy inside the control volume is equal to the net sum of fluxes across the control surface
plus the generation rate.
rate of change = +in − out + generation
or
 dS  X X X Q̇c.v.
= ṁi si − ṁe se + + Ṡgen
dt c.v. T
The devices we looked at from the problems regarding the first law of a control volume
will also apply here. This also means that we will look at two familiar processes regarding
the second law of control volume. They are the steady state processes and the transient
processes.

The Steady-State Process


For the steady-state process, we conclude that there is no change with time of the entropy
per unit mass at any point within the control volume, and therefore, the first term of the
equation equals zero.
 dS 
=0
dt c.v.
The properties do not change with time, which allows us to simplify the second law as
X X X Q̇c.v.
ṁe se − ṁi si = + Ṡgen
T

99
The various mass flows, heat transfer, and entropy generation rates, and states are all con-
stant with time.

If in a steady-state process there is only one area over which mass enters the control volume
at a uniform rate and only one area over which mass leaves the control volume at a uniform
rate, we can write
ṁi = ṁe
Thus,
˙ e − si ) = Q̇c.v.
m(s + Ṡgen
T
Dividing by the mass flow rate gives
qc.v.
se = si + + sgen
T
Recall that sgen is always greater than or equal to zero:
• sgen = 0 when the process is reversible.
• q = 0 when the process is adiabatic.
• se = si when the process is both reversible and adiabatic.

The Transient Process


For the transient process, the second law for a control volume can be written as
 dS  X X X Q̇c.v.
= ṁi si − ṁe se + + Ṡgen
dt c.v. T
If this is integrated over the time interval, t, we have
Z t 
dS
dt = (m2 s2 − m1 s1 )c.v.
0 dt c.v.
Z tX  X Z tX  X Z t
ṁi si dt = mi si , ṁe se dt = me se , Ṡgen dt = 1 S2 gen
0 0 0
Therefore, for this period of time t, we can write the second law for the transient process as,
Z tX
X X Q̇c.v.
(m2 s2 − m1 s1 )c.v. = mi si − me se + dt + 1 S2 gen
0 T
Since in this process the temperature is uniform throughout the control volume at any instant
of time, the integral on the right reduces to
Z tX Z t X Z t
Q̇c.v. 1 Q̇c.v.
dt = Q̇c.v. dt = dt
0 T 0 T 0 T
and therefore, the second law for the transient process can be written
Z t
X X Q̇c.v.
(m2 s2 − m1 s1 )c.v. = mi si − m e se + dt + 1 S2 gen
0 T

100
Engineering Applications; Efficiency
In general, we can say that to determine the efficiency of a machine in which a process
takes place, we compare the actual performance of the machine under given conditions to
the performance that would have been achieved in an ideal process. For example, a steam
turbine is intended to be an adiabatic machine. The only heat transfer is the unavoidable
heat transfer that takes place between the given turbine and the surroundings. We also note
that for a given steam turbine operating in a steady-state manner, the state of the steam
entering the turbine and the exhaust pressure are fixed. Therefore, the ideal process is a
reversible adiabatic process, which is an isentropic process, between the inlet state and the
turbine exhaust pressure.

The ideal turbine process would go from state i to state es because the process is isentropic.
However, in a real turbine process, there would be entropy generation. Thus, the process
would go from state i to e.

Denoting the work done in the real process i to e as w, and that done in the ideal isentropic
process from the same Pi , Ti to the same Pe as ws , we define the isnetropic efficiency of
the turbine as
w hi − he
ηthermal = =
ws hi − hes
The entropy generation of the ideal process and the entropy generation of the real turbine
process are as follows
• Ideal turbine: Sgen = 0 −→ ses = si
• Real turbine: Sgen > 0 −→ se > si
Thus, the entropy generation of a real turbine is

Ṡgen = ṁ(se − si )

101
Let’s analyze the pumps now. We normally assume compressors or pumps to be adiabatic.
In this case, fluid enters the compressor at Pi and Ti , the condition at which it exists, and
exits at the desired value of Pe , the reason for building the compressor. Thus, the ideal
process between the given inlet state i and the exit pressure would be an isentropic process
between state i and state es with a work input of ws .

The real process is irreversible and the fluid exits at a real state e with a larger entropy, and
a larger amount of work input w is required. The isentropic compressor or (pump in the
case of a liquid) efficiency is defined as

ws hi − hes
ηcomp = =
w hi − he

Thus, the entropy generation of a real compressor/pump is

Ṡgen = ṁ(se − si )

Another equation to know and be familiar with is the work of an ideal pump with an
incompressible liquid. There is a whole derivation process for this equation, but I am too
lazy and it’s the end of the course. Here is the equation.

Ẇs = ṁv(Pi − Pe )

102
Let’s do a brief rundown of another useful device, the nozzle.

The energy balance of a real nozzle is


1
hi − he − V2e = 0
2
The energy balance of an ideal nozzle is
1
hi − hes − V2es = 0
2
The isentropic efficiency of a nozzle is

ke 1/2V2e hi − he
ηnozz = = 2 =
kes 1/2Ves hi − hes

Entropy generation in a real steady-state nozzle is

Ṡgen = ṁ(se − si )

Review Video
Here is a comprehensive video of a few example problems.

https://youtu.be/Dwo3fuq31do?list=PLqOZ6FDR Q7lRjgvvbv9t92I − P 0i9 Dj

103

Вам также может понравиться