Vous êtes sur la page 1sur 84

CONCISE CLASSICAL THERMODYNAMICS

Reference Notes for CEEN 501


Professor Steven Rogak
Department of Mechanical Engineering, UBC
December 2015
Clean Energy Engineering

CONTENTS
Overview...................................................................................................................................................................................1
1.1 Purpose of These Notes ............................................................................................................................................1
1.2 Thermodynamics: Classical and Statistical Viewpoints ..............................................................................1
1.3 Organization of Notes and the Grand Concept Map ......................................................................................2
1.4 Learning Outcomes .....................................................................................................................................................4
1.5 Laws, Formulas and Definitions you should Memorize ..............................................................................5
Balances and Laws................................................................................................................................................................5
Definitions ................................................................................................................................................................................5
Processes ..................................................................................................................................................................................5
Property Models ....................................................................................................................................................................6
Systems, State and Measurable Properties ................................................................................................................7
2.1 Systems ............................................................................................................................................................................7
2.2 State and Properties (State Functions) ..............................................................................................................7
Mass m (kg) .............................................................................................................................................................................7
Volume V (m3) ........................................................................................................................................................................8
Pressure P (kPa) ....................................................................................................................................................................8
Temperature T (K) and the Zeroth Law ......................................................................................................................9
Processes, Work and Heat .............................................................................................................................................. 10
3.1 Processes ..................................................................................................................................................................... 10
3.2 Work .............................................................................................................................................................................. 12
Definition ............................................................................................................................................................................... 12
Boundary Work................................................................................................................................................................... 12
Other forms of Work ......................................................................................................................................................... 13
3.3 Heat Transfer ............................................................................................................................................................. 14
Definition ............................................................................................................................................................................... 14
Modes of Heat Transfer and Calculation .................................................................................................................. 14
The Adiabatic Process ...................................................................................................................................................... 15
3.4 Q,W and Inexact Differntials ................................................................................................................................ 15
Thermodynamic Laws for a Control Mass ............................................................................................................... 17
4.1 Statement of the 1st Law ........................................................................................................................................ 17
4.2 Physical Interpretation of Energy ..................................................................................................................... 18
4.3 Statement of the 2nd Law ....................................................................................................................................... 19
4.4 Physical Interpretation of Entropy ................................................................................................................... 20

ii
Clean Energy Engineering
4.5 The Reversible Path for Energy and Entropy Calculations ..................................................................... 20
Thermodynamic Property Relations.......................................................................................................................... 21
5.1 Extensive, Intensive and Specific Properties ................................................................................................ 21
5.2 Phase.............................................................................................................................................................................. 22
5.3 Simple Compressible Pure Substance.............................................................................................................. 22
5.4 PVT Surface for a SCPS ........................................................................................................................................... 24
5.5 Other Thermodynamic Surfaces ........................................................................................................................ 25
5.6 First and Second Laws for a Reversible Process of a SCS ........................................................................ 26
5.7 Special Property Models for Parts of PVT Surface ...................................................................................... 26
Gases........................................................................................................................................................................................ 28
Incompressible liquids and Solids .............................................................................................................................. 29
Two-Phase Mixtures ......................................................................................................................................................... 30
5.8 The Enthalpy and Heat Capacities ..................................................................................................................... 31
Why can we sometimes use Cp if pressure is not constant? (Student FAQ) .............................................. 32
5.9 Prove This! .................................................................................................................................................................. 33
Open System (Control Volume) Analysis ................................................................................................................. 34
6.1 Mass Balance .............................................................................................................................................................. 34
㲰ӟ
6.2 Energy Balance (1st Law) ...................................................................................................................................... 36
6.3 Entropy Balance (2nd Law) ................................................................................................................................... 38
6.4 Steady-State Steady-Flow (SSSF) Problems .................................................................................................. 39
6.5 Uniform State, Uniform Flow (USUF)............................................................................................................... 40
6.6 Prove This! .................................................................................................................................................................. 41
Reversibility Again ............................................................................................................................................................ 42
7.1 Tests for Irreversibility .......................................................................................................................................... 42
Process includes non-equilibrium states ................................................................................................................. 42
Friction+motion or heat transfer +∆T ....................................................................................................................... 42
The “movie” of an irreversible process looks silly in reverse ......................................................................... 42
Entropy is generated in an irreversible process ................................................................................................... 42
7.2 Examples of Irreversible Processes .................................................................................................................. 43
Heat Transfer Between Two Subsystems ................................................................................................................ 43
Shooting apples and explosions ................................................................................................................................... 44
Throttling or pipes with friction .................................................................................................................................. 45
7.3 Prove this.. ................................................................................................................................................................... 45

iii
Clean Energy Engineering
Heat Engines and Carnot Efficiency ........................................................................................................................... 46
8.1 Heat Engines ............................................................................................................................................................... 46
8.2 Reversible Heat Engine Efficiency ..................................................................................................................... 47
8.3 The Carnot Engine.................................................................................................................................................... 47
The 4 Processes .................................................................................................................................................................. 47
A Vapor Cycle Carnot Engine......................................................................................................................................... 48
8.4 Alternate Statements of the 2nd Law................................................................................................................. 49
8.5 Prove This! .................................................................................................................................................................. 49
Final Words: Problem Solving ...................................................................................................................................... 50
9.1 Solution Components.............................................................................................................................................. 50
Problem Statement and Sketch (ALWAYS First!) ................................................................................................. 50
System Definition ............................................................................................................................................................... 50
Use Definitions to Translate Variables ...................................................................................................................... 51
Property Models and State Diagrams ........................................................................................................................ 51
Process.................................................................................................................................................................................... 51
Formulate Equations (& Assumptions if Needed)................................................................................................ 51
Calculate................................................................................................................................................................................. 51
Ӡ
9.2 Better than Memorizing all the Examples ...................................................................................................... 52
9.3 Problem Solving in the Real World ................................................................................................................... 53
Exergy (a.k.a. availability)and irreversibility ......................................................................................................... 54
10.1 Qualitative Ideas ....................................................................................................................................................... 54
10.2 Reversible work for the USUF process ............................................................................................................ 54
10.3 Irreversibility for the USUF Process ................................................................................................................. 56
10.4 Conclusions ................................................................................................................................................................. 56
10.5 Prove This! .................................................................................................................................................................. 57
Thermodynamics of Ideal Gas Mixtures ................................................................................................................... 58
11.1 Pressure........................................................................................................................................................................ 58
11.2 Energy and Enthalpy ............................................................................................................................................... 58
11.3 Entropy ......................................................................................................................................................................... 59
11.4 Air-Water Model ....................................................................................................................................................... 59
11.5 Prove this!.................................................................................................................................................................... 61
Phase and Chemical Equilibrium – General Relations ........................................................................................ 62
12.1 Reading from Textbook ......................................................................................................................................... 62
12.2 Notation for Multicomponent, Multispecies Systems ............................................................................... 62

iv
Clean Energy Engineering
12.3 Equilibrium of a System at Specified T, P ....................................................................................................... 63
Chemical Reactions for Ideal Gas Mixtures ............................................................................................................. 65
13.1 Fuels and Review of Stoichiometry................................................................................................................... 65
13.2 Enthalpy of Formation and Enthalpy of Reaction ....................................................................................... 66
Example – Heating a house with a gas furnace ...................................................................................................... 67
13.3 Adiabatic Flame Temperature ............................................................................................................................ 69
13.4 Chemical Equilibrium ............................................................................................................................................. 69
13.5 Modern Method of Finding Equilibrium Composition .............................................................................. 72
Multicomponent Phase Equilibrium .......................................................................................................................... 73
14.1 Gibbs Phase Rule....................................................................................................................................................... 73
14.2 The Salt-Water-Air System ................................................................................................................................... 74
14.2.1 Volume: Gas....................................................................................................................................................... 75
14.2.2 Volume: Liquid ................................................................................................................................................. 75
14.2.3 Enthalpy: Gas .................................................................................................................................................... 76
14.2.4 Enthalpy: Liquid .............................................................................................................................................. 76
14.2.5 Entropy: Gas...................................................................................................................................................... 76
14.2.6 Entropy: Liquid Solution.............................................................................................................................. 78
ᓐӠ
14.2.7 Phase Boundaries ........................................................................................................................................... 79

v
Clean Energy Engineering

OVERVIEW
1.1 PURPOSE OF THESE NOTES
These notes summarize the core theory of classical thermodynamics for engineers. They are complete
but as they lack examples, you will need other activities to help you learn thermodynamics. These
notes cover entropy and temperature in the way I think they should be covered, consistent with the
classroom lectures.

1.2 THERMODYNAMICS: CLASSICAL AND STATISTICAL VIEWPOINTS


All scientific experiments and all engineering experience indicate that the laws of thermodynamics
(which relate heat, work, energy E, entropy S and temperature) are never violated. Thus,
thermodynamics applies to EVERYTHING, but we focus on engineering applications where we want to
convert one form of energy into another. More specifically, we focus on systems that contain gases and
liquids that can expand and do work when they are heated. The discovery of this type of conversion,
from heat to mechanical power, was central to the Industrial Revolution that changed the world and is
still at the core of most energy systems.

Thermodynamics becomes especially important when the internal (microscopic) state of things
change. These microscopic changes can be reflected in the average speed of the molecules (related to
temperature), the way the molecules are packed together (ie., density and phase), and chemical
composition. Starting with the principle of conservation
�ӟ of energy, and assuming that all allowed
microstates are equally probable, the theory of statistical mechanics can be used to derive the 2nd Law
of Thermodynamics, compute the entropy and the temperature of a system. This is appealing because
it provides a very precise and physical notion of entropy, but it is mathematically complicated even for
ideal gases. For engineering, it is more practical to use “classical thermodynamics”, which can handle
simple and complex substances equally easily.

Classical thermodynamics makes no use of the microscopic viewpoint. Instead, we start by assuming
the truth of the 1st and 2nd Laws of Thermodynamics, and then derive useful results for engineering
computation. The mathematics is not hard, but energy (E) and entropy (S) are abstract unmeasurable
quantities. These notes take the classical approach, with occasional reference to a microscopic
viewpoint as a visualization aid.

Concise Classical Thermodynamics 1


Clean Energy Engineering

1.3 ORGANIZATION OF NOTES AND THE GRAND CONCEPT MAP


The next page lays out the key concepts of the course, and their interconnections to the applications,
shown near the middle of the diagram. The 1st and 2nd Laws are central to this course, but before
stating these, key concepts must be established first. These concepts include the idea of state fixed by
measureable properties (PVT; Chapter 2), equilibrium, processes, heat, work introduced in Chapter 3.

The 1st and 2nd Laws are introduced together in Chapter 4. Introducing the 2nd Law and entropy so
early is not conventional but it is sensible. Often, the 2nd Law is introduced as a proposition about heat
engines (Kelvin-Planck box in map on the next page), from which some brilliant and difficult reasoning
is used to derive the relation for entropy change. In this course, I take the change of entropy as “the
second law”, an approach used more in chemistry, chemical engineering and physics. This approach
has several advantages. Firstly, the 1st and 2nd Laws are cast in a similar form, providing a way to
calculate the change in a property (E, S respectively). Secondly, we often use entropy in engineering;
we do not often use the Kelvin-Planck statement. Thirdly, it is easier to start from the change in
entropy and prove Kelvin-Planck (Chapter 7) than the other way around.

Beyond the 1st and 2nd Laws, we need data, models and relations between properties to solve real
problems. This vast area of study is introduced briefly in Chapter 5.

The thermodynamic laws are introduced for closed systems for simplicity, but many (most?)
engineering problems are better analyzed using “open” systems, with mass flowing in or out. For this,
we need to recast balance equations for mass, energy, and entropy (Chapter 6).
긐Ӣ
At the end of these notes, Chapter 9 provides guidance on how to integrate the theoretical components
into a consistent problem solving method.

Before starting into the theory, the learning objectives are stated in this chapter, followed by the key
information that you will want to commit to memory well before this course is finished.

Concise Classical Thermodynamics 2


Clean Energy Engineering

Concise Classical Thermodynamics 3


Clean Energy Engineering

1.4 LEARNING OUTCOMES

Memorize
• the fundamental balance equations (laws) for mass, energy and entropy .
• definitions of key thermodynamic properties and concepts.
• numerical values of key thermodynamic properties of common substances.
• shape of the Pressure-Volume-Temperature (PVT) surface for water.
Draw
• processes consistently on different state diagrams (Pressure-Volume (P-V), Temperature-
Entropy (T-S) etc).
• boundaries of open and closed systems that correspond with the balance equations used in
problem solving.
Compute
• all thermodynamic properties of a Simple Compressible Pure Substance (SCPS) given 2 other
properties.
Select

• among several substance models (ideal gas, solid, liquid, property tables) consistent with
information given in problem statements.
• boundaries of open and closed systems that help solve the given problem.
Solve

• engineering problems in which heat, work or thermodynamic states are desired, using all of
the skills listed above, for open and closed systems.
Assess

• the meaning of problem solutions in a simple design settings (i.e., explain how a particular
numerical result should affect a design choice).

Concise Classical Thermodynamics 4


Clean Energy Engineering

1.5 LAWS, FORMULAS AND DEFINITIONS YOU SHOULD MEMORIZE


A good working understanding requires a lot more than memorization, but if you don’t memorize
anything, then it will be very hard for you to follow the lectures and tutorials. Here I have summarized
everything you should memorize. Work on it every day. At first, you may understand a small part of
what you are memorizing, but by the end of the course, I hope these few pages will be trusty tools in
your brain, ready for application to problems (and don’t panic- this information is on the formula
sheet).

BALANCES AND LAWS


Closed System (Control Mass): A region in space (contiguous or not) with no mass flow in or out.
Open System (Control Volume CV): A region in space that exchanges matter with surroundings.
First Law (Energy E) Control mass: E2-E1=Q-W E=U+KE+PE

Control volume:
dE cv
dt inlets i
( 2

outlets j
)
= Q& − W& s − W& BDY + ∑ m& i hi + gz i + V2i − ∑ m& j h j + gz j + 2j
V2
( )
δQ
Second Law (Entropy S) Control mass dS ≥
T
 Q& 
 + ∑ m& i (si ) − ∑ m& j (s j ) + S& gen
dS cv
Control volume = ∑ 
dt k  Tcv  k inlets i outlets j

Consequences of 2nd Law:


• Kelvin Planck: It is impossible to construct a device that operates on a cycle, exchanges heat with a
single thermal reservoir, and has no other impact繐ӭexcept raising a weight.
• Zeroth Law: If systems A&B are in thermal equilibrium AND B&C are in thermal equilibrium, THEN
A&C are in thermal equilibrium
• Carnot engine efficiency W/QH=1-TL/TH

DEFINITIONS
 ≡ 1  ≡  +


ℎ ≡ ℎ − ℎ ∂u ∂h ∂u Cp
  
T≡ Cp ≡ Cv ≡ γ ≡k≡
∂s v ∂T P ∂T v Cv

PROCESSES
Adiabatic & reversible= isentropic
Q=0, P=constant, control mass: isenthalpic
Q=0, V=constant, control mass: iso-energy
Polytropic process Pvn=constant
P2V2 − P1V1 V 
For control masses, polytropic: W = if n ≠ 1 W = P1V1 ln 2  if n = 1
1− n  V1 
Internally reversible process for a control mass: δW=PdV and δQ=TdS
W& s 2
Reversible steady-state steady flow SSSF shaft work = − ∫ v dP
m& 1

Concise Classical Thermodynamics 5


Clean Energy Engineering

PROPERTY MODELS
SCPS Simple Compressible (no magnetic or electrical work) + Pure Substance (no composition
change)

U=f(V, S) and P=function of (V, T) state can be shown on planer diagram

du=Tds-Pdv dh=Tds+vdP

For 2 phases e.g. gas (g) and liquid (f): h=hf+xhfg (similar for all extensive properties)

Ideal gas: PV=nRT h,u are functions of T only; Cp-Cv=R

constant heat capacity: s2-s1=Cp ln(T2/T1)- R ln (P2/P1)

+ isentropic: Pvk = C1 Tv k −1 = C2 TP(1−k ) / k = C3

Real gas PV=ZnRT Z=f(T/Tc, P/Pc) Z1 for low density

Incompressible liquids and solids: v~constant and u=u(T)


T2
s 2 − s1 = C ln
T1

述Ө
Key Property Data

1 atm= 101.3 kPa=760 mm Hg=14.69 psi=1.013bar

1 kJ=0.95 BTU 1 cal=4.184 J 1 m=3.2808 ft

Gas Constant: For air R=287 m2/(s2K)=0.287 kJ/kg/K Universal: R=8.314 J/mol/K

Water: MW=18 g/mol Cp=Cv=4.18 kJ/kg/K for liquid water

latent heat of fusion: 333.7 kJ/kg enthalpy of vaporization at 1 bar is 2258 kJ/kg

Water vapor heat capacity ~1.91 kJ/kg/K

Concise Classical Thermodynamics 6


Clean Energy Engineering

SYSTEMS, STATE AND MEASURABLE PROPERTIES


Before discussing thermodynamics, we must define “systems”, the concept of “state”, and remind you
of the key measurable properties involved in thermodynamics: mass, volume, pressure and
temperature. Other key thermodynamic properties (U, S, H) are discussed in Chapters 4&5.

2.1 SYSTEMS
A “system” is an identified collection of atoms. If not otherwise specified, systems are “closed”,
meaning that no mass passes across the boundary of the system. Sometimes, we use “open systems”
(usually called “control volumes”) where mass enters and/or leaves.

In thermodynamics the selection (and indication with a dotted line) of the system boundary is as
critical to problem solving as the free-body diagram in statics and dynamics. The balance equations
that we will write for mass, energy and entropy have no meaning if the system is not defined.

Consider the following sketch of a piston-cylinder Moving piston


arrangement containing air. The piston moves upward.
a) Everything
Our analysis boundary might include a) everything, b) b) Piston
just the piston, c) just the air (with a moving boundary c) Cylinder contents
d) Boundary fixed in space
following the piston) or d) part of the air in a fixed
L
region of space.

These 4 systems do not contain the same amount of D


mass, energy or entropy. Three of them (a,b,c) are Fig. 2.1
closed systems (control masses); one (d) is a control
volume (or open system) because air will cross the boundary of the system as the piston moves.

2.2 STATE AND PROPERTIES (STATE FUNCTIONS)


Once we have defined our system, we can start to define its state by specifying various properties. The
state of a system is its “configuration”, as specified by as many variables as needed. At the microscopic
level, we would need to specify the position, type and velocity of every molecule in the system – an
absurdly large amount of information. Fortunately, the macroscopic state is adequately described by
only a few properties, which can be thought of as summarizing the microscopic information. These
properties include the mass, volume, energy, pressure, temperature and entropy. Properties are
sometimes called “state” or “point” functions because they depend only on the current configuration,
not the path leading to configuration.

Equilibrium is an important and subtle concept in thermodynamics, introduced here and revisited
later. A system is said to be in equilibrium if, when it is isolated (left alone), it has no spontaneous
tendency to evolve into a different macroscopic state. Typically, this means that pressure and
temperature are uniform throughout the system.

This course focusses on “simple compressible pure substances” (SCPS), whose state is fixed completely
by the relation between mass, pressure, volume and temperature, introduced next.

MASS M (KG)
Concise Classical Thermodynamics 7
Clean Energy Engineering
For a closed system (isolated or not), the mass is constant as long as we do not consider nuclear
reactions (which we will not). For open systems, mass balances are more complex (Chapter 6). Mass
is measureable with familiar methods such as a balance and calibrated weights.

VOLUME V (M3)
Volume is easily computed from the system geometry (e.g., for system c in Fig 2.1, V=πLD2/4). There is
no “volume conservation” principle, but we often encounter systems with constant volume (for
example, when gas is contained by a perfectly rigid tank). The most confusing thing about volume is
probably the notation! In this course, we also use V for velocity. Keeping track of the units prevent
any ambiguity. In some cases, instead of using specific volume (v≡V/M) we can use density ρ≡1/v.

PRESSURE P (KPA)
The pressure is the average normal compressive stress exerted on a surface. In terms of the normal
stresses introduced in your solid mechanics course,

σ x +σ y +σ z
P=− (2.3)
3

In static fluids (or those with negligible shear stresses), all normal stresses are the same and the
pressure completely determines the state of stress in the substance. In this case, force balances can be
used to determine P. In Fig. 2.1, if the piston moves slowly, the pressure P of system c would be

P=P0 + mp g/A 述Ө (2.4)

where P0 is the atmospheric pressure acting on the top of the piston, mp is the piston mass and g is the
gravitational acceleration. Modern pressure transducers employ other physical principles, but their
calibration can be traced back to the fundamental force balances as illustrated above. For a liquid or
gas at equilibrium, the pressure is nearly uniform (or varying gradually with elevation, as you know
from hydrostatics).

Concise Classical Thermodynamics 8


Clean Energy Engineering

TEMPERATURE T (K) AND THE ZEROTH LAW


Temperature characterizes the way energy is organized in a system, and is closely linked to entropy.
From statistical mechanics, T is related to the kinetic energy of the moving molecules1, but we never
need this idea in classical thermodynamics. Here is what you need to memorize:

1. T must be uniform throughout a system at thermal equilibrium. This is restatement of


the Zeroth Law of Thermodynamics.

2. We can measure temperature of unknown system A by bringing a thermometer into


thermal equilibrium with A. The thermometer is a system that responds to temperature
in a known way. For example, the pressure of an ideal gas in a fixed volume will be
proportional to T using T=P V/R/m.

3. The ideal gas thermometer implies that there is an absolute zero temperature (because
T  0 as P 0. In any problem involving ideal gases, entropy and Carnot efficiencies,
you MUST work in absolute temperature (e.g., Kelvin K).

Some treatments of thermodynamics skip the 0th Law. Such treatments either take it as self-evident
that there is a property T that is measured by thermometers, or introduce the entropy first and define
temperature from this.

�ӱ

1 For a system where the root-mean-square speed of molecules is vRMS, and molecules have mass m0, T=m0
vRMS2/3kB. The Boltzmann constant is kB. This microscopic picture is comforting, but we do not actually use this
idea in classical thermodynamics. Note that the internal energy U includes the translational kinetic energy term
m0 vRMS2/2, so we expect a relation between T and U. However, the energy includes other terms that are
unrelated to T (Chapter 4). Please do not think of T as a measure of U.

Concise Classical Thermodynamics 9


Clean Energy Engineering

PROCESSES, WORK AND HEAT


3.1 PROCESSES
A process is a set of states through which a system passes going from initial state 1 to final state 2. A
state diagram represents the state of the system by plotting key property values, for example pressure
and volume. This approach works well for simple systems (eg., all air, or all water) but gets
complicated for systems composed of diverse subsystems. In such cases, many variables would be
needed to represent the state, and the process is a path through higher dimensional space.

If the states along the path are close to equilibrium, then the process is internally reversible. For a
simple system, an internally reversible processes can be represented unambiguously on a “state
diagram” (with a solid curve) because at any instant it can be characterized by a single value of T and P
(recall T and P must be uniform throughout the system at equilibrium).

We may have equilibrium for the initial and final states (hence these points can be represented on a
state diagram), but the path may include non-equilibrium states, usually involving non-uniform T and
P. This is an internally irreversible process, shown with a dotted line - - - to indicate that the path
cannot be properly defined on the state diagram (and in fact the shape of the curve has no meaning,
but it does show that states 1 and 2 are connected). Reversibility is a sufficiently subtle concept that it
is covered in its own Chapter 7.

Consider the gradual expansion of a gas (or any simple compressible substance) in cylinder (Fig. 3.1,
述Ө
process (a)). A load is applied to the piston to exactly balance the changing pressure of the gas. This is
a reversible (internally and externally) process. In contrast, start with the gas in the same state 1 (held
in one half of a container by a thin wall). Suppose this wall breaks, and the gas rapidly fills the entire
container, which we have contrived to match the final volume of the cylinder V2. With suitable heat
transfer to the gas, we can force the pressure to match that of P2 in the cylinder. Thus, the terminal
states of processes (a) and (b) are the same, but process (b) is irreversible. There are infinitely many
processes that would take us from 1 to 2. One more (c, which happens to be reversible) is shown in
Figure 3.1 and will be discussed in the next section.

Concise Classical Thermodynamics 10


Clean Energy Engineering
Aside from its reversibility (or lack thereof), a process might be characterized by a particular property
held constant, or there might be a special constraint on the system. Here is a list of the more common
processes with some mechanisms that can cause them:

Adiabatic (no heat transfer, Q=0) – this is typically produced by very fast processes and/or a lot of
insulation around the system. This is not “iso-anything” because heat is not a property, as will
be emphasized later in this chapter.

Isobaric (constant pressure, e.g., first part of process c) - can arise if the load on a piston is held
constant.

Isochoric (constant volume, e.g., second part of process c) - can be obtained using a rigid container or
piston prevented from moving.

Isothermal (constant temperature) – often arises when the process is “slow” and the system is in
contact with a very large constant temperature reservoir.

Isenthalpic (constant enthalpy, Chapter 5) – typically arises when a system is insulated and isobaric.

Isoenergy (constant energy, Chapter 4) – guaranteed for isolated systems!

Isentropic (constant entropy, Chapter 4)- guaranteed by 2nd Law for processes that are both adiabatic
AND reversible. Theoretically, it is possible to contrive an isentropic process that is
irreversible and cooled at just the right rate.

Polytropic (PVn=constant; exponent n is constant) - this process simply provides the mathematical
shape of the process on a P-V diagram. For the special case n=0, this is the isobaric process.
Nothing is implied about the substance involved, but if the substance is an ideal gas, the
polytropic process has special meaning for particular values of n.

Cyclic Process – Any process that starts and ends at the same state. Cycles can be made up of
reversible OR irreversible processes.

Concise Classical Thermodynamics 11


Clean Energy Engineering

3.2 WORK
DEFINITION
A closed system can have only two types of interaction with its surroundings. One of these is a work
interaction, defined as “an interaction equivalent to raising a weight with no other effect on the
surroundings”.

This definition is equivalent to the notion of a


force moving through a distance, δW=F . dx
because you could always use a frictionless
mechanical system to harness this moving force in
the lifting of a weight. For example, imagine that
we have a gas in a cylinder (Fig. 3.2) with pressure
P in the left side less than pressure P0 on the right
side. The imbalanced pressure on the piston can
be used to pull on the rope and lift a weight. The
“system” has not actually lifted a weight (we must
be very picky about the system boundary!), but the interaction is equivalent to raising a weight.

The second part of the definition (“…with no other effect on the surroundings”) is important. For
example, we might produce the pressure imbalance on the piston by heating the gas on the right side.
In this case, the work interaction might have been the same, but heat (to be defined momentarily) was
transferred from the environment, so perhaps now䃰some
Ӵ object outside the system is now cooler.
Therefore, this is not a pure work interaction.

BOUNDARY WORK
In the example above, work was done as a gas expanded
against a moving boundary. This case merits a special
calculation method. Consider a substance in a cylinder
with uniform pressure P. Suppose the piston moves up a
very small distance dx, resulting in a larger volume dV. To
do this, the substance has had to apply a force F=P πD2/4
on the piston and move it through a distance. The work
done is δW=Fdx= P πD2/4 dx. At the same time, we can
compute the extra volume dV= πD2/4 dx, so we can
rewrite the work as

δW=P dV (3.1)

Concise Classical Thermodynamics 12


Clean Energy Engineering
For a process involving substantial movement, the pressure may not be constant, and the work is
obtained by integration:
V2

1W 2 = ∫ PdV (3.2)
V1

This integral is the area under the process curve on a P-V diagram. Thus, process (c) from Figure 3.1
resulted in more work than process (a), ie., 1W2c> 1W2a .

Caution: this equation for work is based on the pressure at the moving surface. If the pressure is not
uniform, then we cannot compute the work from any “average” pressure. For example, the area under
curve (b) is more than under curve (a), but it is NOT true that 1W2b> 1W2a . Instead, 1W2b=0, as can be
proven by considering either a system boundary that moves with edge of the gas (where P=0), or using
a non-moving system boundary with the final volume. In thermodynamics, the area under a
dashed (irreversible) process curve has no meaning.

OTHER FORMS OF WORK


Although we focus on Simple Compressible Substances that can ONLY do boundary work, we often add
other types of substances into the problems to be solved. For example, if a rotating shaft penetrates
the system boundary, there can be “shaft work”. A sliding rod under stress (in tension in Fig. 3.2) can
also do work. Magnetic or electric fields transmitted through the control surface and do work. Electric
current is a form of work. There is no thermodynamic limit to the efficiency of an electric motor, for
example, which can be used to lift weights.

From this, you might expect that gravitational forces are part of the work, but in fact we conventionally
include the effect of gravity in the potential energy term of the First Law, not the work term. This trick
works for systems in an inertial reference frame or even under constant acceleration. An example
where this does not work is the shaking of an air-water mixture contained by a constant-volume,
insulated thermos. A SCS in such a container should be isolated, but clearly we can transfer energy to
the fluid through vigorous shaking. We will not consider such cases further in this course.

Concise Classical Thermodynamics 13


Clean Energy Engineering

3.3 HEAT TRANSFER


DEFINITION
Heat transfer is the other (non-work) form of interaction. A more satisfying definition: “ heat transfer
is an energy transfer due solely to a difference in temperature”.

MODES OF HEAT TRANSFER AND CALCULATION


Direction of
Conduction occurs when fast molecules collide with their slower heat transfer
(low temperature) neighbors. On average, the molecules in the hot
object transfer energy to molecules in the cold object. Conduction Area A x
can occur in solids, liquids and gases. For conduction in 1-
Thermal
dimension, the rate of heat transfer (Watts=J/s) is
conductivity k
(W/m/K)
dT
Q& = − kA (3.3)
dx

Convection occurs when a fluid (temperature Tf) flows past or through an object of different
temperature (Tw). Engineers have developed special methods to compute convection heat transfer
coefficients as a function of flow and geometry. The Variation of
temperature
symbol “h” is conventionally used for the heat transfer Tf-Tw near
Tf surface
coefficient (although it has nothing to do with enthalpy, a Flow Boundary layer
property introduced later and also denoted by h):
�ӯ
Q& = hA(Tf − Tw ) (3.4) Solid object T w

Fundamentally, it is still conduction through a relatively


quiescent layer (thickness δ) of the fluid (conductivity kf) that occurs near the solid surface, and the
convection coefficient is approximately.

kf
h≈ (3.5)
δ
The “quiescent layer” is the boundary layer, which gets thinner in high velocity flows (hence the
“wind-chill effect”).

Radiation is the only heat transfer mode that occurs in a vacuum (though it occurs in substances too).
All matter radiates photons; the more energetic the molecules, generally the shorter the photon
wavelength. Thus, the hottest objects emit radiation with shorter (higher energy) photons. For two
black parallel plates at different temperatures, the heat transfer is:

(
Q& = σA T14 − T24 ) (3.6)

Here σ is the Stefan Boltzmann constant. Note that the radiation heat flux vanishes if the temperature
is uniform, as it must for any heat transfer mechanism. Also note that the radiation flux increases with

Concise Classical Thermodynamics 14


Clean Energy Engineering
the 4th power of temperature, and is therefore relatively more important in high-temperature
problems.

THE ADIABATIC PROCESS


The computation of heat transfer rates by conduction, convection and radiation can become very
involved. Beyond introducing these mechanisms, these notes are concerned primarily with two
aspects of heat transfer:

1. Using the 1st or 2nd Laws to compute Q or


2. Recognizing when a process is adiabatic (Q=0).

By definition, Q=0 if the system is at exactly the same temperature as its surroundings. However,
nearly uniform temperature is not the best sign of an adiabatic process. In fact, in the laboratory, the
best way to keep a system at constant temperature is to immerse it in a stirred fluid bath (water, oil or
fluidized sand). This results in a very high convection heat transfer coefficient so that even a small
temperature difference (say 0.1 oC) is enough to produce substantial heat transfer rates. This heat
transfer prevents the system temperature from deviating much from the bath temperature.

The two factors most important to determining whether a process is adiabatic or not are

a) insulation

b) time
兠Ӫ
Insulation materials have low conductivity and will impede radiative and convective heat transfer. In
this course, the phrase “well insulated” is “code” for “adiabatic”. It is left to another course to provide
the methods needed to decide how much insulation is enough.

The time over which a process occurs is equally important in practice (although hardly mentioned in
the textbook). Notice that the equations above predict the heat transfer rate (kJ/s), so a very rapid
process will not allow a significant amount of heat transfer (in kJ) to occur. Thus, air compressed in an
engine running at 4000 rev/min would be nearly adiabatic, producing a large air temperature
increase. Conversely, if the same engine is cranked at 1 rev/min, the air temperature will hardly be
affected by the compression.

3.4 Q,W AND INEXACT DIFFERNTIALS


Frequently, we need to integrate a quantity along the process; this is a path integral. For example, we
might add up all the infinitesimal temperature changes dT along the path:

2
T2 − T1 = ∫ dT (3.7)
1

The limits of integration should be read “from state 1 to state 2”. Because temperature is a state
function, the integral can be written as a difference of this function at the two states. However, if the
quantity to be integrated is the work done by the system, the result is subtly but critically different:

Concise Classical Thermodynamics 15


Clean Energy Engineering
2
W2 = ∫ δW
1 ≠ W2 − W1 (3.8)
1

The limits of integration again signify that the path runs from state 1 to state 2; the integral is still the
limit of a sum (now we add up the work δW done over each portion of the path), but the result
depends on the path taken, so there is no “potential function”. For example, it makes no physical sense
to speak of the “work at state 2”. Note that we used differential “dT” for a small change in temperature
(a property) but inexact differential “δW” for small amount of work. Similarly,

1 Q2 = ∫ δQ ≠ Q2 − Q1 (3.9)
1

The quantities W1, W2, Q1, Q2 are physically nonsensical. To speak of the “heat at some state” is
equivalent to saying that “the travel distance to Vancouver is 36,233 km” – one immediately wants to
ask “where were you travelling from and what route did you take?”

兠Ӫ

Concise Classical Thermodynamics 16


Clean Energy Engineering

THERMODYNAMIC LAWS FOR A CONTROL MASS


4.1 STATEMENT OF THE 1ST LAW
There is a property “E”, the total energy, whose change for a closed system is given by

   −  (4.1)

The sign convention for heat and work is defined in Figure 4.1 showing a simple piston-cylinder
arrangement, but equation 4.1 applies to ALL closed systems, filled with ANY substances. Here the use
of δ reminds us that Q and W are not state functions. For a finite change in state, the differential form
above is replaced by

 −     −   (4.2)

For an isolated system (the universe is a big example of an isolated system), the 1st Law reduces to the
principle of conservation of energy, dE=0.

兠Ӫ

Concise Classical Thermodynamics 17


Clean Energy Engineering

4.2 PHYSICAL INTERPRETATION OF ENERGY


One of the most disturbing things about energy is that it cannot be measured directly. Only changes in
energy can be determined (from the 1st Law). There is no such thing, and can never be, an “energy
meter”. Moreover, there can be no “absolute zero” of energy- numerical values are defined only with
respect to a conventional baseline. Nevertheless, a more physical interpretation of energy can be
comforting and even useful. The total energy E can be partitioned into many different forms, some of
which you already know how to compute:

E=KE+PE+U (4.3)

KE=macroscopic translational or rotational kinetic energy as covered in dynamics


PE=macroscopic potential energy, e.g. from gravity or stored in an elastic spring
U=internal energy =Utrans+Urotation+Uvibration+UPE
Utrans=translational kinetic energy of all the molecules in the system
Urotation=rotational kinetic energy for polyatomic molecules only
Uvibration=energy due to stretching and bending bonds in polyatomic molecules
UPE= energy of physical and chemical interactions between molecules, as well as
nuclear interactions between molecules.

Note that the components of the internal energy related to molecular motion have a natural zero – the
absence of any motion (at absolute 0 Kelvin). The microscopic potential energy, however, has no
natural zero. Gas molecules could condense into a lower energy state, but after this, the might react
for form new species with even lower energy; with nuclear fission and fusion we could get new
elements with much lower energy – far below the level set by convention as “zero”. Many students
come to thermodynamics with the incorrect idea that there can be no such thing as negative energy.

So far, we don’t have a way of measuring Q, but we have postulated that for a sufficiently insulated
system, Q=0. In such adiabatic cases, it is possible to determine ∆E from work transfers, which are
measurable. After doing enough of these adiabatic experiments, the energy (relative to some
conventional baseline) is a known function, allowing the heat transfer to be calculated for non-
adiabatic cases using the 1st Law. Joule (cover image of these notes) is famous for the first experiments
of this nature, see for example http://en.wikipedia.org/wiki/James_Prescott_Joule.

Concise Classical Thermodynamics 18


Clean Energy Engineering

4.3 STATEMENT OF THE 2ND LAW


There is a property of a system, the “entropy” (symbol S, units J/K), whose change is given by

 !
"
(4.4)

In Equation 4.4, we take the equality (=) when the process is reversible; for an irreversible process
dS>δQ/T. For a finite change of state (as illustrated in Figure 4.2), the change of entropy is
 
# − # ! $ "
(4.5)

As for energy, entropy is not measureable and only defined to within an arbitrary constant.

The inequality seems to imply an ambiguity or vagueness in the


entropy change- but this is misleading. The entropy change is exactly
determined by the state change 12. The uncertainty comes into the
relation between heat transfer and entropy: for reversible processes
we have a clear relation, but for irreversible processes, knowing the
heat transfer and temperature only allows us to compute a lower
bound on ∆S. Consider Figure 4.3, which illustrates two ways of
warming up 1 kg of water from, say 290 K to 291 K. Using the slow,
reversible heating method (a), the heat transfer would turn out to be
Q=4200 J, and the entropy change S2-S1 would be approximately 4.2/290.5=14.5 J/K, using equation
兠Ӫ
4.4. We could also get to the same final state by stirring the water (Joule`s experiment) in an insulated
container. Now Q=0, but the process is irreversible (you can put work into the water to warm it up;
you can`t extract work from this device and have the water cool down). So, we can only say from the
second law that the entropy change is greater than zero. In fact, we know the entropy change is 14.5
J/K, but only because we first did the reversible experiment (a).

The physical concept of reversibility is difficult and


later notes will cover this concept in more detail
(Chapter 7). We will also show that this definition
of entropy leads easily to predictions about
perpetual motion machines and the Carnot cycle
efficiency (Chapter 8).

Concise Classical Thermodynamics 19


Clean Energy Engineering

4.4 PHYSICAL INTERPRETATION OF ENTROPY


The concept of entropy (equation 4.4) was first introduced by Rudolf Clausius in 1850, 138 years after
the first practical heat engine was built by Thomas Newcomen. He did this using a mathematical
argument, with no conception of what might occur on a microscopic level. You may have heard that
entropy is a measure of disorder in a system. This is mostly correct, but does not lead to useful
calculation methods without a hard course in statistical mechanics. Long after Clausius, Boltzman
related entropy to “thermodynamic probability” Γ, a statistical measure of atomic disorder, as

#  %& ln Γ (4.6)

Here kB is the Boltzmann constant. Equation 4.6 is absolutely correct but useless for our purposes
because finding Γ is hard to impossible. Equation 4.6 implies that there is such a thing as “zero
entropy”, which is what the 3rd Law of Thermodynamics states. Again, we focus on equations 4.4 and
4.5 are useful to us because they involve physical quantities Q and T.

4.5 THE REVERSIBLE PATH FOR ENERGY AND ENTROPY CALCULATIONS


Two states (1,2) can be connected by an infinite number of different paths, not all reversible. The
concept of the reversible path can be useful in the analysis of problems even if the real process is not
reversible, because the end states are path independent. This means that the energy and entropy
changes can be computed for some convenient path, but the results will be valid for any path.

Concise Classical Thermodynamics 20


Clean Energy Engineering

THERMODYNAMIC PROPERTY RELATIONS


The properties P, V, T were introduced in Chapter 2, while E, U, S were introduced in Chapter 4. This is
the key cast of characters needed for thermodynamics. In this section we introduce some basic rules
for working with properties for real substances.

5.1 EXTENSIVE, INTENSIVE AND SPECIFIC PROPERTIES


Often we must break a system into subsystems. For example, the piston-cylinder arrangement (Fig.
3.1) consists of the piston (A), cylinder (B) and the contents (C, perhaps air).

How do we compute the properties of the entire system ABC from knowledge of the subsystems?
There is an important class of properties that are extensive, and for these, the “whole is the sum of the
parts”. Extensive properties introduced so far are mass, volume, total energy, internal energy and
entropy. Thus,

mABC = mA + mB + mC (5.1a)

VABC = VA + VB + VC (5.1b)

EABC = EA + EB + EC (5.1c)

UABC = UA + UB + UC (5.1d)

SABC = SA + SB + SC (5.1e)
兠Ӫ
This is simple but very powerful because it provides a basis for
breaking large complex systems into simpler parts.

Pressure P and temperature T are intensive properties. As discussed earlier, if the whole system is at
thermal equilibrium

TABC = TA = TB = TC (5.2)

If the whole system is NOT at thermal equilibrium, then the subsystem temperatures are unequal and
it is impossible to assign a temperature to the whole system ABC. Similarly, if the system is in
mechanical equilibrium,

PABC = PA = PB = PC (5.3)

but otherwise the system pressure has no single value.

Another class of properties is intensive properties manufactured by dividing extensive properties by


mass: v=V/m (m3/kg), e=E/m (kJ/kg), u=U/m (kJ/kg), s=S/m (kJ/K/kg). The sum vA+vB+vC has no
useful meaning (as for the other intensive properties) but we CAN assign meaningful specific
properties to the whole system as averages, for example:

vABC = (mA vA + mB vB + mC vC ) / mABC (5.4)

Concise Classical Thermodynamics 21


Clean Energy Engineering
Note that rather than specific volume, v, its inverse, density, ρ(kg/m3), is often used in engineering.

5.2 PHASE
A phase is region of matter with homogeneous (uniform in space) properties and composition. Ice
(solid), liquid water (liquid) and water vapor (gas) have different properties; if any combination of
these are present together in a system, we have a multiphase system. Ice cubes floating in liquid water
is a simple 2-phase system. Steam bubbles rising in a pot of boiling water is another example.
Multiphase systems do not have homogeneous properties.

5.3 SIMPLE COMPRESSIBLE PURE SUBSTANCE


The 1st and 2nd Laws apply to all substances, but every substance responds differently to changes in
volume or pressure. In solid mechanics, you used constitutive relations to relate stress and strain, with
material-dependent properties such as the Young’s Modulus. In thermodynamics, property models
can be very complex; the pressure P at equilibrium might be a function of many variables:

P=P(T, V, n1, n2, n3…, magnetic field, surface area, electric field…) (5.5)

Here n1, n2, n3… are the number of moles of the different chemical species that might be present. Some
materials might interact with magnetic and electric fields, or might be influence by surface tension. In
addition, we would have different functions for the internal energy U and entropy S of the substance.

In this course we will focus on Simple compressible pure substances (SCPS). ”Simple Compressible”
means that the only way a substance can do work on the surroundings is through movement of its
boundary (ie., expanding against a pressure). This contrasts with special magnetic or electrical
materials2.

“Pure Substance” means that the material has a single, unchanging chemical composition, or less
restrictively, it is a mixture with constant composition (air is often modeled in this way). Note SCPS
can, and often do, exist in multiple phases simultaneously in a system, for example liquid and solid
water at 0 oC, 1 bar.

For a SCPS, the state of the system at equilibrium is determined by T (an indicator of average
molecular speed) and v (an indicator of how much space each molecule has, which in turn affects
intermolecular forces). You should find it plausible that we can write

P=P(v, T) (5.6a)

u=u(v, T) (5.6b)

s=s(v, T) (5.6c)

2Strictly, this notion applies to equilibrium states. Moving fluids with significant velocity gradients contain shear
stress that can transfer work without volume changes. Thus, water in Joule’s paddle-wheel experiment is
actually not behaving as a SCS when work is done on it.

Concise Classical Thermodynamics 22


Clean Energy Engineering

兠Ӫ

Concise Classical Thermodynamics 23


Clean Energy Engineering

5.4 PVT SURFACE FOR A SCPS


The actual functions 5.6abc differ from one material to another, but qualitatively they are similar. Any
function of the form z=f(x,y) is a 3-dimensional surface. The function P(v, T) or “PVT” surface (Figure
5.2) has a critical role in thermodynamics. This surface covers all P, V, T, but we slice out a finite
region; the grey edges are not actually part of the PVT surface, but indicate where we have cut the
surface. On this surface, you can see regions of different phases – liquid, gas, solid. The liquid+gas and
solid+gas regions are shown with a speckled pattern to indicate that in this zone we really have 2
distinct phases mixed together.

You will need to know how to visualize this PVT surface as a 3-d object. If you are visualizing this
surface correctly, then it should be easy to sketch the P-v, P-T and T-v projections obtained by viewing
the surface along the T, v and P axis directions, respectively. Being able to sketch projections of the
PVT surface showing various constant-property contours is critical to problem solving.

兠Ӫ

Concise Classical Thermodynamics 24


Clean Energy Engineering

5.5 OTHER THERMODYNAMIC SURFACES


We can fix the state of a SCPS with T and v, but almost any two properties will do. For example, given
u and s, we could rearrange the functions above (5.6abc) to obtain T, v and P. Simple equations for P, u
and s exist only for special cases (e.g., gases). Engineers must work with fluids such as water and
refrigerants. Unfortunately, there are no simple equations for their properties, and we must rely on
sophisticated correlations built into computer programs or property tables (eg, at the back of your
textbook). Water is the most important thermodynamic substance; MATLAB functions to determine
properties are available at http://www.mathworks.com/matlabcentral/fileexchange/9817 . The
water properties are also displayed graphically on the temperature entropy diagram for water (Fig.
5.3). On this plot, T and s are the independent variables; P, v and h(enthalpy, to be defined in a few
pages) are determined by interpolating contours.

Concise Classical Thermodynamics 25


Clean Energy Engineering

5.6 FIRST AND SECOND LAWS FOR A REVERSIBLE PROCESS OF A SCS


If a process is reversible and the only work is boundary work, then the work is related to P and V:

δW=PdV (5.7)

Also, if a process is reversible, then from the 2nd Law,

δQ=TdS (5.8)

Finally, if the system does not change KE or PE in the process, dE=dU, and the First Law is
transformed:

dU=TdS-PdV (5.9)

Now here is the surprising part: this result is actually not restricted by the assumptions made in its
derivation! We have in the end a relation involving only state functions (properties). We do require
that the system be at an equilibrium state and that it be a SCS, but it does not matter how it got there.
Furthermore, it does not matter whether the SCS has kinetic or PE – the relationship tells us about dU,
independent of KE and PE changes.

Property models (and measurements of properties) must satisfy this relation to be valid; it helps
reconstruct complete thermodynamic data sets given partial information.

5.7 SPECIAL PROPERTY MODELS FOR PARTS OF PVT SURFACE


The P-T projection of the PvT surface (also known as the phase diagram, Figure 5.4) is particularly
useful because it shows exactly where we have gas, solid, vapor or 2-phase mixtures. For each of these
different regions of the P-T plane, different calculation methods are appropriate.

Concise Classical Thermodynamics 26


Clean Energy Engineering

兠Ӫ

Concise Classical Thermodynamics 27


Clean Energy Engineering

GASES
Some texts distinguish between “vapors” which are at a temperature below critical Tc, and “gases” at
higher temperatures. This is largely artificial and not useful. If you examine the PVT surface, you will
see that there are no sharp boundaries as you move up and around the critical point from liquid to gas
to vapor.

Ideal Gases

For ideal gases (special case of SCPS) with constant heat capacity Cv, there are simple equations for the
pressure, internal energy and entropy:

P= RT/v (5.10a)

u= u0 + Cv (T-T0) (5.10b)

s= s0 + Cv ln (T/T0) + R ln (v/v0) (5.10c)

The ideal gas constant R must be in the appropriate units3. The entropy and energy are determined
relative to some reference state T0, v0. Most introductory thermal physics courses cover these models
for ideal gases- this should not be new. In the case that the heat capacity is not constant, it must still be
a function of temperature and the equations for u(T) and s(V, T) are a little more complex.

Non-ideal or “real” gases

The ideal gas law is only valid at “sufficiently low density” (high specific volume). At high density, we
must introduce a non-unity “compressibility factor” Z into the equation:

P=Z RT/v (5.11)

Z can be determined from correlations (or a myriad of other calculation methods), and when Z=1, then
the gas is ideal. The ideal gas law is surprisingly good (few % error) for most cases with specific
volume 10 times the critical volume. For nitrogen, vc=0.0899 m3/kmol (0.00204 m3/kg or ρ=489
kg/m3). Thus, we expect near-ideal behavior if v>0.0204 m3/kg or ρ<49 kg/m3. For reference, the
density of air (mostly nitrogen) at sea level is about 1.1 kg/m3. The textbook appendices include
critical T, P, V for many common substances.

3The universal gas constant is 8.314 J/mol/K; for mass-based units, divide by the molecular weight of the
species involved. Note that specific entropy and heat capacity both have the same units as R. Some texts use
typefaces that distinguish between /kg and /mol quantities, but this is not an issue if you check all your
equations for dimensional homogeneity every time. Conversion from molar to mass units is messier in systems
with many chemical species – it is covered later in the course.

Concise Classical Thermodynamics 28


Clean Energy Engineering

INCOMPRESSIBLE LIQUIDS AND SOLIDS


Note in Figure 5.2 that the isotherms (constant T contours) for the liquid and solid portions of the PVT
diagram are very steep: the volume hardly changes at all even for enormous pressure increases. Often
we can model this behavior as “incompressible”. The terminology is odd, but we still call such a
material “simple compressible” if it cannot do magnetic, electrical or other forms of work that would
affect internal energy4. For incompressible substances, v changes slightly with temperature

v=v(T) (5.12a)

u= u0 + Cv (T-T0) (5.12b)

s= s0 + Cv ln (T/T0) (5.12c)

Here, u0, and s0 are the energy and entropy at the reference temperature T0 and Cv is the heat capacity.
Naturally, these are all material-dependent constants.

㡰ӳ

4Solids can contain elastic strain energy. By convention, we normally consider this as part of the potential
energy, for example in the case of springs. This is justified because normally thermal and elastic effects are not
strongly coupled. Notable exceptions would be shape memory alloys, beyond the scope of this course.

Concise Classical Thermodynamics 29


Clean Energy Engineering

TWO-PHASE MIXTURES
For a large part of the PVT surface, we may have 2
phases present. In the P-T projection, the 2-phase
regions appear as curves with no area. This implies
that P and T cannot be varied independently without
leaving the 2-phase region. It also implies that there is
relation (known as the “saturation curve”) Psat (Tsat)
that can help you find one property given the other.

Consider the vapor-liquid mixture in more detail. In


the P-v projection, the 2-phase region covers a large
area, with the volume ranging from the “saturated
liquid volume” vf5 to the “saturated vapor volume” vg. In Fig. 5.5, only the liquid, vapor and
liquid+vapor regions are shown, to avoid the complications of the solid portions of the diagram.
Consider a system that might be at states a, b, c or d all at the same pressure. As noted already, the
fixed pressure implies a fixed temperature, so the isotherm is horizontal in the 2 phase region (a, b, c, d
appear as a single point on the P-T diagram).

In a 2 phase mixture, we define quality x as the mass fraction of the vapor:

mg
x≡ (5.13)
mg + m f
㡰ӳ

The states a, b, c and d are mixtures of saturated liquid and vapor with different quality:
a) x=0, we have only satuated liquid
b) x~0.2, picture liquid with some large vapor bubbles
c) x~0.8, imagine a gas with small droplets of liquid
d) x=1, saturated vapor (a gas)

By considering the gas (subscript g) and liquid (subscript f) in the mixtures as two subsystems, you
should be able to prove that the mixture average specific properties are:

v= (1-x) vf + x vg (5.14a)

u= (1-x) uf + x ug (5.14b)

s= (1-x) sf + x sg (5.14c)

5 I guess the subscript “f” refers to “fluid”, even though gases are considered types of fluids in all fluid mechanics
textbooks. The subscript “g” presumably refers to “gas”, even though some themo texts say that a gas can only
exist above the critical temperature. This sloppy notation is an indication that we can use gas and vapor
interchangeably in this course.

Concise Classical Thermodynamics 30


Clean Energy Engineering
We are assuming that the mixture is in an equilibrium state, therefore the temperature and pressure
are uniform throughout the mixture.

5.8 THE ENTHALPY AND HEAT CAPACITIES


A function of properties is automatically a property, so we could invent an infinite number of new
thermodynamic properties, even though the properties introduced so far are sufficient to describe and
solve all thermodynamics problems. We define (symbol ≡ means “identically equal to”) new
properties that turn out to be convenient.

Enthalpy H≡U+PV, h=H/m≡u+Pv (5.15)

Heat capacities are another important class of derived properties, but the name is an unfortunate
historical relic. Systems do not “contain” heat, as has been known since the “caloric theory” of energy
has been dismissed. Instead, heat capacities are really defined in terms of properties.

∂h
Heat capacity at constant pressure C p ≡ (5.16a)
∂T P

∂u
Heat capacity at constant volume Cv ≡ (5.16b)
∂T v

The notation used in the derivatives specifies exactly what is held constant while temperature is being
changed. In your math class, you don’t usually specify what is held constant in a partial derivative
because math problems normally start out specifying what variables are involved. For example, if
z=f(x,y) , then the partial fx implies that y must be held constant. In thermodynamics we must be more
careful with the notation because we can represent variables in alternative ways. We can think of
enthalpy being a function of T and P, or of T and v, or T and S. Thus, the extra notation in equation 3.13
implies h(T,P) and u(T, v)

This strictly mathematical definition of heat capacity is actually consistent with experiments where we
transfer heat to a system held at constant pressure or constant volume. However, emphasizing the
experimental method of measuring Cp and Cv can lead to confusion… hence the next section.

Concise Classical Thermodynamics 31


Clean Energy Engineering

WHY CAN WE SOMETIMES USE CP IF PRESSURE IS NOT CONSTANT? (STUDENT FAQ)


Thermodynamics is full of surfaces, such as P(v, T), u(T, v) or h(T, P). In math notation, a surface is
z=f(x, y), and if we want to know the change in z for arbitrary changes in x, y we use

∂z ∂z
dz = dx + dy (5.17)
∂x y ∂y x

Thermodynamics always obeys the rules of mathematics, so applied to u(T,v) we get

would get

∂u ∂u
du = dT + dv (5.18)
∂T v ∂v T

Applied to the enthalpy expressed as a function of T and P, we get

∂h ∂h
dh = dT + dP (5.19)
∂T P ∂P T

These are convenient expansions because the partials with respect to T have just been defined:

∂u
du = Cv dT + dv 㡰ӳ (5.20)
∂v T

Or

∂h
dh = CP dT + dP (5.21)
∂P T

Now here is the important part: we can neglect the second term in the expansion if EITHER the
differential is zero (ie, that property is constant) OR if the partial derivative is zero.

For ideal gases, we have stated many times that u varies only with T, so we can neglect the second term,
no matter what is happening to P and v: we can use du=Cv dT.

Recalling that we can ALWAYS write h=u+Pv, then using Pv=RT for an ideal gas, we find h=u(T)+RT.
Thus, h only varies with temperature and we can use dh=Cp dT for ideal gases.

For liquids and solids, we can neglect the second term in Eq. 3.16a (for energy, du) because in most
problems, the change in volume (dv) of a liquid is exceedingly small. However, the second term for the
enthalpy expansion (involving dP) (3.16b) is normally retained.

Concise Classical Thermodynamics 32


Clean Energy Engineering

5.9 PROVE THIS!


1. It was shown that du = T ds-P dv. Use this to show that given equations 5.10a and 5.10b, then
5.10c must follow automatically.
2. For an incompressible liquid with constant heat capacity, h2-h1= C(T2-T1)+ v(P2-P1).
3. In the L-V two-phase region, the specific entropy is s=sf + x sfg where sfg ≡sg-sf.
4. For an ideal gas with known temperature-dependent Cv(T),
T

u − u0 = ∫ Cv (T ' )dT '


T0

5. For a SCS undergoing a reversible isobaric process, the heat transfer is Q=H2-H1
6. For an adiabatic, reversible process, S2=S1.
7. For a polytropic process, independent of the working fluid, the work is
P2V2 − P1V1
W = n ≠1
1− n
V 
W = P2V2 ln  2  n =1
 V1 
8. For an ideal gas, Cp-Cv=R
9. For an ideal gas with constant specific heats, an isentropic process is a special polytropic
process with n=k, where k ≡ Cp/Cv.
10. For an ideal gas, an iso-energy process (u=constant) is isothermal.
11. In the liquid-vapor 2-phase region of a SCS,㡰an
ӳ isothermal process is isobaric.

12. In the liquid-vapor region, show that an isobaric process is not iso-energy.
13. The temperature of an incompressible liquid does not change if we place it in an insulated
piston-cylinder arrangement and increase the pressure.

Concise Classical Thermodynamics 33


Clean Energy Engineering

OPEN SYSTEM (CONTROL VOLUME) ANALYSIS


So far, the notes have covered all the theory needed to solve problems involving closed systems. In
engineering, however, we often deal with machines that process fluids flowing through them, and so it
is convenient to take the machine as our analysis boundary. Such open systems are commonly called
“control volumes”. We would like to describe how mass, energy and entropy change in a control
volume.

6.1 MASS BALANCE


Mass conservation for a closed system is trivially satisfied, but we now consider the open system
(control volume) in Fig. 6.1 with several inlet and outlet pipes, as well as mechanisms for transferring
work. The dashed control surface cuts across each of the pipes (which have cross sectional area A).
Suppose we measure the velocity V and density ρ at each station 1, 2, 3. We could then compute the
mass flux for pipe 1 as

m& 1 = ρ1 A1V1 (6.1)

In most of what follows, we will work with the mass flux


(kg/s) rather than velocities, densities and areas, but it is
important to emphasize that the mass flux would be
determined by measurements made at the control
surface where mass crosses it. Dimensionally, this flux
兠Ӫ
looks like a time-derivative of “m”, but there is no state
function “m”. The relevant property in this problem is the
mass in the control volume MCV. You could determine this
quantity knowing the density at all locations x,y,z inside
the control volume:

M CV = ∫∫∫ ρ ( x, y, z ) dxdydz (6.2)


CV

In many practical situations, we do not know the spatial variations of density inside the control
volume, so it is useful to derive mass balances that do not require this. The mass fluxes affect MCV, of
course. In a small time interval ∆t, we expect MCV to change by

∆M CV = m& 1∆t + m& 2 ∆t − m& 3 ∆t


∆M CV dM CV (6.3)
lim = = m& 1 + m& 2 − m& 3
∆t →0 ∆t dt
The final result might seem obvious, but the definition of the derivative was used to emphasize that
the LHS (left-hand side) is in fact a time derivative, while the RHS is best thought of as a sum of ρAV
terms.

Concise Classical Thermodynamics 34


Clean Energy Engineering
It is very common to choose control surfaces that are fixed
in space, for example as shown in Fig. 6.2 (identical to 6.1
except that in this case, there is no moving piston as part of
the control surface). Further simplifications can arise if the
problem is steady (considered a few pages later), or if the
fluid has constant density.

In many engineering systems, the fluid (liquid water,


hydraulic fluid) has practically constant density. From the
volume-integral MCV (equation 6.2) it is clear that the
control mass cannot change with time, so

0 = m& 1 + m& 2 − m& 3 (6.4)

Note that these fluxes can vary with time, but if the flux in increases, then the flux out must increase
simultaneously. For example, in Fig. 6.3, the control volume is the inside of a bathtub. After the tub is
& o (t ) = m& i (t ) .
filled, we must have m

Figure 6.3

Concise Classical Thermodynamics 35


Clean Energy Engineering

6.2 ENERGY BALANCE (1ST LAW)


We will start with the same device as shown in Fig.6.1, but in 6.4 we have noted that the specific
energy e≡u+V2/2+ gz will be relevant to the balance. Also, we consider the possibility of analyzing the
device with two control surfaces (A, and B).

The mass fluxes for CVA and CVB are the same (except for the small amount of piston strut mass
moving through the top of CVB). The heat and work terms will be different.

Start by considering the energy ECV in CVA


carefully. Because we know how to analyze closed
systems, imagine that at time instant t, we take a
snapshot of all the fluid in CVA and call this “System
1” (Sys1).

Because the fluid is moving in pipes 1, 2, 3, Sys1 will


eventually get flushed out of the control volume. For
a very small time interval, from t to t+∆t, the
difference between the Sys1 and CVA will be confined
to the regions near the pipes. CVA will contain fluid I
and II (shaded region in Fig 6.3) that is not part of Sys
1. At the exit, Sys 1 will contain fluid III that is
outside CVA.

The masses I, II, III are determined from the fluxes at 1, 2 , 3 and the time interval, for example:

δmI= ρ1V1A1∆t (6.5)

The volume of these regions I, II, III would be v1δm1, v2 δm2, v3δm3 , respectively, where v is the specific
volume (=1/ρ).

Now apply the 1st Law to Sys 1 over the time interval ∆t. The electric heater produces flux Q& ; the
shaft work will be W& s . We will denote the power at the moving piston to be W& BDY (“boundary work”).
There are additional work terms due to the flow at 1, 2, 3. The fluid regions I, II, III are drawn
cartoonishly to remind you of pistons transferring work to or from Sys 1. Now the change in energy of
Sys 1 can be written

( )
Esys1 (t + ∆t ) − Esys1 (t ) = Q& −W& s − W& BDY ∆t + (P1v1m & 2 − P3v3m& 3 )∆t
& 1 + P2v2 m (6.6)

We are seeking the change in ECVA. At time t, ECVA=ESys1, simply because we have chosen Sys1 and CVA to
be coincident at that time. However, at t+∆t, they differ due to the energy contained in I, II and III.

E sys1 (t + ∆t ) = E CVA (t + ∆t ) − (e1 m& 1 + e2 m& 2 − e3 m& 3 )∆t (6.7)

Put equation 6.7 into equation 6.6 to get

Concise Classical Thermodynamics 36


Clean Energy Engineering

( )
ECVA(t + ∆t ) − ECVA(t ) = Q& − W& s − W& BDY ∆t + ((e1 + P1v1 )m & 3 )∆t
& + (e2 + P2 v2 )m& 2 − (e3 + P3v3 )m
(6.8)

Finally, dividing through by ∆t and taking the limit as the interval goes to zero,

dECVA
= Q& − W& s − W& BDY + (e1 + P1v1 )m& + (e2 + P2 v 2 )m& 2 − (e3 + P3 v3 )m& 3 (6.9)
dt

You should have no difficulty using the definitions of e and h to get the First Law for a control
volume:

dECVA  V 21   V 22   V 23 
= Q& − W& s − W& BDY +  h1 + + gz1 m& +  h2 + + gz 2 m& 2 −  h3 + + gz 3 m& 3
dt  2   2   2 

(6.10)

If the control volume has constant shape, there can be no W& BDY term. If electrical current passes
through the control surface, there is a W& elec term. For example, using CVB, the heat transfer term is
zero, replaced by a W& elec term of the same magnitude. The Steady-State, Steady Flow (SSSF) case is
important enough to warrant its own section later, and there are a few transient cases also worth a
separate section below.

Concise Classical Thermodynamics 37


Clean Energy Engineering

6.3 ENTROPY BALANCE (2ND LAW)


The entropy balance for a control volume can be
derived in the same manner as the energy
balance, so we will simply state the result and
explain some tricky issues.

Naturally, we are concerned with the flux of


entropy carried by the mass fluxes. We must
also be concerned with the heat fluxes and the
temperature at which these fluxes occur. To
illustrated this point, suppose that the heat flux
to the control volume occurs in two parts (i and
ii; the second parts happens to be due to the
electric heater). The 2nd Law for the control volume is:

dSCVA Q& i Q& ii


= + + s1 m& 1 + s 2 m& 2 − s3 m& 3 + S&Gen, A (6.11)
dt Ti Tii

The entropy generation term takes the place of the inequality used in the previous statements of the
Second Law (units kW/K). Entropy generation cannot be negative; normally it is very hard to
determine this except by knowing the other terms in the equation, or by assuming that it is zero
(reversibility assumption). ᶀӷ

As usual, the exact choice of control volume is critical. If the temperature of the control surface is 400
and 450 K at the two locations where heat transfer occurs, then these temperatures must be used in
the denominators of the balance (eq 6.11). However, if we are concerned with CVB, the entropy
balance would involve a different temperature,

dS CVB Q&
= i + s1 m& 1 + s 2 m& 2 − s3 m& 3 + S&Gen, B (6.12)
dt 500

The mass fluxes and specific entropies at pipes 1, 2, 3 would be the same, but the control volume
temperature is different and in fact the 2nd heat flux vanishes (because now the control volume is
importing electric work, which does not show up in the entropy balance). The entropy generation
term is larger due to the irreversibility of the heat transfer through the device walls and the electrical
resistance heating, which is now inside the control volume.

Use of the entropy balance requires more careful control surface definition than either the mass or
energy balances. It also requires solid understanding of reversibility, to be discussed later. Therefore,
you will not be expected to work with the entropy balance yet, but it is logical to include the theory in
this section for completeness.

Concise Classical Thermodynamics 38


Clean Energy Engineering

6.4 STEADY-STATE STEADY-FLOW (SSSF) PROBLEMS


It is often desirable to operate machinery and devices steadily. SSSF problems often arise naturally in
the analysis of pumps, compressors, turbines, valves, pipes, heat exchangers and rocket nozzles. The
SSSF condition means that no measurements made around the control surface would change with
time. The mass fluxes are constant (but typically non-zero). If the fluid properties inside the control
volume are steady, then triple integral for MCV must be independent of time, and the mass balance
reduces to:

0 = m& 1 + m& 2 − m& 3 (6.13)

We don’t require density to be uniform throughout the control volume. For example, pipes 1 and 2
might carry high-pressure high-density air into a turbine, and the exhaust air (3) might have lower
density.

The energy balance is simplified because the LHS must vanish, and because there can be no boundary
work (at least, I cannot see how the boundary of a control volume could expand while the machine
operates in steady state).

 V 21   V 22   V 23 
0 = Q& − W& s +  h1 + + gz1 m& 1 +  h2 + + gz 2 m& 2 −  h3 + + gz 3 m& 3 (6.14)
 2   2   2 

The entropy balance for CVA of Figure 6.4 (without boundary expansion) operated in SSSF is:

Q& i Q& ii
0= + + s1 m& 1 + s 2 m& 2 − s3 m& 3 + S&Gen, A (6.15)
Ti Tii

Concise Classical Thermodynamics 39


Clean Energy Engineering

6.5 UNIFORM STATE, UNIFORM FLOW (USUF)


Without the SSSF assumptions, problems are MUCH harder. Not only is there an extra non-zero term
in all the balances (LHS), but now we need an initial condition and we are concerned about finding the
state of the control volume at some particular time. By comparison, the SSSF problems are time
independent. There are only two classes of these transient problems that are suitable for
undergraduate homework or exams. The first is the case where all mass fluxes are zero  we are back
to closed system analysis as covered in the first part of the course. The second case that the control
volume has a uniform (but time-varying) state and the flow in or out has steady properties (but not
necessarily constant mass flow). This is the “Uniform State Uniform Flow (USUF)” situation. Because
the state inside the control volume is spatially uniform, at any instant we can compute the total
amount of some property as the product of mass and the uniform specific property, eg. SCV=msCV. This
means that we can integrate the equations from time t1 to time t2 and the left hand side of the balance
equation will be simple, involving the mass and property values at state 1 and state 2. The RHS of the
balance equations is also integrable because the properties at the inlet and exit do not vary with time,
so they can be brought outside the integral. For the USUF situation, we must be very careful with the
notation, and it works better to use subscripts 1, 2 for the initial and final states inside the control
volume, while i, e will refer to the inlet and exit states (understanding that we could have many inlet
and exit pipes by adding more similar terms). This notation is used in Fig. 6.6. At time t=t2, the picture
would look identical, but the entropy would be SCV2=m2s2.

兠Ӫ

CVB
CVA se, ρe, Ve,

si, ρi, Vi,

m1s1=SCV1
Shaft work
Tii=450 K
Ti=400 K Electric

500K Q& i =1000 W power to


heater

Fig. 6.6 Notation for CV undergoing USUF process

Concise Classical Thermodynamics 40


Clean Energy Engineering
For Fig 6.6, the USUF mass, energy and entropy balances are:

) − )  )* − )+
1
* +
) , − ) ,  - . /0 − 23 + )* 4ℎ* + + 67* 8 − )+ 4ℎ+ + + 67+ 8
1 2 2

1
. 1
) 9 − ) 9  - .
/0 + )* 9* − )+ 9+ + - #+< /0
1 :; 1

Internal, kinetic and potential energy are included in the specific energy “e”, and WCV includes
boundary and shaft work. The time integrals in the 2nd Law are hard, usually, but the time integral
in the the 1st Law is simply the total heat transfer transferred to the CV over the time interval.

6.6 PROVE THIS!


1. An adiabatic, no-work SSSF process is isenthalpic if the fluid velocities are not too high.

2. For liquid water flowing through a CV, show that a 1 oC temperature change affects the energy
balance more than a velocity change from 10 m/s to 20 m/s.

3. Think of a practical example that shows that the specific work in a SSSF process CANNOT be
proportional to the area under the P-v curve.
㡰ӳ

4. Show that for a control volume with the inlets and outlets blocked, we recover the mass,
energy and entropy balances introduced in Notes 1.

5. Generalize the control volume mass, energy and entropy balances for an arbitrary number of
inlet and outlet pipes.

Concise Classical Thermodynamics 41


Clean Energy Engineering

REVERSIBILITY AGAIN
An internally reversible process is a set of states through which the system can be taken in both
directions. This is a compact definition, but does not lead to practical “tests” that you can used to
decide whether or not a process is close to reversible or not. Some practical “tests” will be introduced
below, then it will be shown through several examples that these tests are actually equivalent.

7.1 TESTS FOR IRREVERSIBILITY


PROCESS INCLUDES NON-EQUILIBRIUM STATES
Our test for a system being in equilibrium is the same one introduced at the beginning of the notes: if
you isolate the system, does its state continue to evolve over timescales relevant to your problem?

This implies that we have uniform T, P and concentration at each state in the process. This is a little
too restrictive and we should specify that this must be true within communicating subsystems. For
example, suppose that one part of the system is a piston-cylinder system undergoing pressure and
temperature changes, while the other part is a large block of iron insulated from the cylinder. If there
is no heat transfer between the block and the cylinder (due to sufficiently good insulation), then we do
not require these subsystems to have the same temperature. Within the cylinder, however, we would
require uniform temperature.

FRICTION+MOTION OR HEAT TRANSFER +∆T


If fluid or mechanical friction is important for components
哰ӵ that have large relative motions, then the
process is irreversible. If there is heat transfer with a substantial temperature difference, then the
process is irreversible.

THE “MOVIE” OF AN IRREVERSIBLE PROCESS LOOKS SILLY IN REVERSE


The laws of mechanics are time-reversible, but for complex systems, entropy provides a direction to
time, and you can really see this reflected in the following absurd videos:

http://www.youtube.com/watch?v=9wnYdWfO77M

http://www.metatube.com/en/videos/17052/Reverse-Video/

In some cases, irreversibility may not be obvious in a simple video, but if you include additional
instrumentation, you can always tell when the movie is being run in the correct direction.

ENTROPY IS GENERATED IN AN IRREVERSIBLE PROCESS


If we can compute the entropy change in a process, and use this in the entropy balance for a system or
control volume, then the magnitude of the entropy generation term will tell us that the process is
irreversible.

Concise Classical Thermodynamics 42


Clean Energy Engineering

7.2 EXAMPLES OF IRREVERSIBLE PROCESSES


HEAT TRANSFER BETWEEN TWO SUBSYSTEMS
Consider blocks of metal (good conductors) A and B, surrounded by very good insulation, and given
initially different temperatures TA1 and TB1. They are separated by an imperfect insulator, so that
while we observe the blocks, they come to the same temperature. Without this layer between the
blocks, equilibrium would be reached faster, but there would be large temperature differences within
each block during the process; I want to consider the case where the temperature change occurs
almost completely within this layer (as seen in Fig. 7.2).

T A1 TA2 Final
Intermediate
S A1 A SA2 Very good initial A

insulation
T B1 TB2 Imperfect
position insulator
S B1 B SB2 Very poor
insulation
B

TB1 TA1
Temperature
Fig. 7.1 㡰ӳ
Fig 7.2

Equilibrium states? At any stage of the process, A and B have different temperatures, so the only
equilibrium state is the final one. The process is not reversible. Note that the subsystems A and B
undergo nearly reversible processes because they have uniform temperatures. At any stage of the
process, if you isolate A and B by making the imperfect insulator into a perfect insulator, the process
stops instantly, leaving A and B at their different (but internally uniform) temperatures.

Friction+motion or heat transfer + ∆T? There is no motion, but there is heat transfer across a large
temperature difference.

Movie played backwards? This process is not visually stunning, unless you can “see” temperature,
either because the hot object is glowing hot, or because we have thermometers placed in the field of
view. Watch it backwards: the objects start at uniform temperature and heat is spontaneously
transferred from B to A. However, the comedy (violation of the 2nd Law) of the situation is only
apparent if you film the entire system A and B. If the camera is focussed only on A OR B, we see
nothing wrong with the reverse movie unless we can see the tiny temperature gradients in each block.

Entropy Change? For the purposes of this problem, the “universe” is the 2-block system and the
insulating layer between them, which we will assume has negligible mass. For solids with constant
heat capacity, the change in entropy is

Concise Classical Thermodynamics 43


Clean Energy Engineering

T2
S 2 − S1 = mC ln (7.1)
T1

Suppose for simplicity, blocks A and B have mass 1 kg and heat capacity 0.7 kJ/kg/K. Further suppose
that the initial temperatures are 300 and 500K, so that the final temperature is 400K. The change in
entropy for the composite system (neglecting the thin insulator) is

T2 A T  400 400 
∆S AB = ∆S A + ∆S B = m AC A ln + mB CB ln 2 B = 0.7(kJ / K ) ln + ln  = 0.045kJ / K
T1A T1B  500 300 

(7.2)

The unconvinced student can try all combinations of initial temperature, mass and block properties.
Entropy must always go up in an isolated system.

SHOOTING APPLES AND EXPLOSIONS


There is a classic high-speed photo by Edgerton at MIT showing a bullet going
through an apple. If you have trouble imagining this, go to
http://webmuseum.mit.edu/browser.php?m=objects&kv=96483&i=75289
or follow the qr code here.

Equilibrium States? Imagine placing the apple and bullet in an isolating box
at the instant the photo was taken. Would the state 哰change?
ӵ Yes! The bullet
would bounce around inside the box until it stopped moving. Warm apple sauce would cover the
inside of the box. The final resting state is not what you see in the photo, so it cannot be in an
equilibrium state.

Friction+motion? Sure.

Movies? Hopefully, nobody expects a bullet shot backwards to


repair a damaged apple.

Entropy? The calculations here are not easy, but if you looked
at the apple and bullet before and after the impact, you would
find that the temperature has increased slightly, resulting in
increased entropy without any heat transfer. The fact that it
might be difficult to do this calculation should not be
discouraging; rather, it is the reason we need other methods of
determining reversibility.

You can follow the same set of steps in analyzing any explosion,
such as the 21 kiloton atomic test on Bikini Atoll conducted in
1946 (Figure 7.3)

Concise Classical Thermodynamics 44


Clean Energy Engineering

THROTTLING OR PIPES WITH FRICTION


Suppose water flows steadily to the right in the pipe of
Fig. 7.4. In the middle of the control volume, there is an
=P1-P2
obstruction with a small hole, forcing the fluid velocity to
increase and causing back eddies on the right side of the Back eddy
obstruction. The pressure on the upstream side (P1) is
higher, as can be seen by the liquid column height in the
manometer tube above the pipe. What are the clues that 1 2

this flow from 12 is irreversible?

Equilibrium states? Fig. 7.4

If the dashed control volume boundary suddenly becomes solid isolating wall, then we expect it to take
some time before the fluid motion dies away and pressures become uniform. The process is not
reversible.

Friction+motion? Yes

Movies? Any fluid mechanics student should know that the back eddies should occur on the
downstream side of the obstruction, and the flow should go from high pressure to low pressure. So,
we would know that the reverse movie is unrealistic.

Entropy? Suppose that the entire process is adiabatic. For this no-work, SSSF process with no
significant KE differences between 1 and 2, we get h1=h2 from the 1st Law. For the liquid flow, specific
volume v is constant, so from the definition of h we get, u1+P1v=u2+P2v . Using the relation between
internal energy and heat capacity for the liquid, C(T2-T1)=v(P1-P2). Therefore, T2>T1. The second law
reduces to:

& (s1 − s2 ) + S&Gen,


0=m (7.3)

Reviewing the relation between T and s for a liquid, we conclude that s2>s1, so the entropy generation
term is positive.

7.3 PROVE THIS..


1. Two masses m with initial temperature T1 and velocities V and –V collide head-on inelastically.
Each has a heat capacity C. Show that the entropy rise is

 V2 
S 2 − S1 = 2mC ln1 + 
 2T1C 

2. For the example above, use other tests for reversibility.

3. Apply the tests for reversibility to the rapid gas expansion of Fig. 4.1 b

Concise Classical Thermodynamics 45


Clean Energy Engineering

HEAT ENGINES AND CARNOT EFFICIENCY


In these notes we will use the 2nd Law (dS≥δQ/T) to calculate the efficiency of a reversible Carnot heat
engine. We will then use this to prove the Kelvin-Planck and Clausius statements of the 2nd Law. Many
textbooks work in the opposite direction, starting with Kelvin-Planck, deriving the Carnot efficiency,
and finally deducing the existence of entropy.

8.1 HEAT ENGINES


A heat engine (Fig. 8.1) can be defined as a Thermal Reservoir H at temperature TH is an
infinitely big system that does no work and has
device that operates on a cycle, exchanges
fixed mass. Transfer of heat δQH from H changes
heat with two thermal reservoirs, and UH by dUH = −δQH
produces work. A thermal reservoir can be
an object with temperature TH or TL that is
so large that heat exchange with the engine δQH
lifted
does not significantly affect the temperature
of the reservoir. For example, many power Heat engine E Mg
plants use the ocean as the low-temperature operates on a
reservoir (or heat sink). A heat exchanger cycle
kept at constant temperature (for example, δW=δQH-δQL
by continuous combustion of coal) might be
treated as a thermal reservoir, but for the δQL
following derivations, it is easier to think of Thermal Reservoir L at temperature TL is an
哰ӵ
the reservoir as a giant thermal mass. infinitely big system that does no work and has
fixed mass. Transfer of heat δQL TO L changes
It is absolutely essential that the engine UL by dUL= +δQL
operates on a cycle, so that over one
complete cycle, it cannot accumulate any
energy or entropy (state functions). Thus,
Fig. 8.1
δW=δQH-δQL (8.1)

The purpose of a heat engine is to produce work δW (which might be used to lift a weight Mg).
Typically, the heat transferred from the high temperature reservoir is expensive, because it involves
burning a fuel, building a solar collector, drilling into the hot earth, or controlling nuclear fission
reactions. Therefore, it is useful to characterize a heat engine by its thermal efficiency

δW δQ
η= = 1− L (8.2)
δQH δQH

Computing the best possible efficiency is the objective of this topic.

Concise Classical Thermodynamics 46


Clean Energy Engineering

8.2 REVERSIBLE HEAT ENGINE EFFICIENCY


Consider the ideal case of an engine E which is internally reversible and its interactions with the
reservoirs are also reversible. Our “universe” consists of the engine E, the hot reservoir H, the cold
reservoir L, and whatever weight gets lifted by the work we produce (Mg)

Apply the 2nd Law to H+E+L+MG together and separately for the process where the engine goes
through 1 cycle. This “universe” is isolated, so if the process is reversible, entropy should be
unchanged:

∆SHELMG= ∆SH +∆SE +∆SL +∆SMG = 0 (8.3)

Lifting a weight does not increase its entropy, so ∆SMG = 0.

Engine E operates on a cycle, so all properties return to their starting values; ∆SE = 0.

We are left with

∆SH +∆SL = 0 (8.4)

Because we are assuming reversibility, we can compute the entropy changes from the heat transfer,
using the sign convention of Fig. 8.1 that δQH is the heat transferred FROM H and δQL is the heat
transferred TO L.

-δQH /ΤΗ + δQL /ΤL = 0  δQL /δQH = TL /TH 铰ӵ (8.5)

Recalling our definition of efficiency, we obtain the Carnot Efficiency:

η=1-TL/TH (8.6)

Remember that in addition to reversibility, we assume that the engine exchanges heat with only two,
fixed-temperature reservoirs. Many ideal cycles, such as the Otto or Rankine cycles introduced later,
effectively exchange heat with a series of reservoirs at different temperatures. The efficiency for those
cycles can be thought of the weighted-average efficiency for many Carnot engines operating at
different temperatures.

8.3 THE CARNOT ENGINE


THE 4 PROCESSES
There are practical difficulties in building a machine that operates on the Carnot cycle, but it may be
useful to have a more tangible picture in mind – something that might be built with difficulty. All
Carnot engines use the following 4 processes:

1. Isothermal heat transfer from H.


2. Adiabatic, reversible expansion of the working fluid.
3. Isothermal heat transfer to L
4. Adiabatic, reversible compression of the working fluid.

Concise Classical Thermodynamics 47


Clean Energy Engineering

A VAPOR CYCLE CARNOT ENGINE


The device in Fig. 8.2 uses water in the 2-phase region to implement the Carnot cycle. Figure 8.3
shows how the cycle might appear on the T-S diagram for water. The exit state of the condenser is the
lower left corner of the cycle on the T-s diagram.

QH

Boiler (2-phase region, P H)

compressor
turbine
Shaft to drive compressor from turbine
W

Condenser (2-phase region, PL)


QL

Fig. 8.2

兠Ӫ

Concise Classical Thermodynamics 48


Clean Energy Engineering

8.4 ALTERNATE STATEMENTS OF THE 2ND LAW


Because everything in the derivation of the Carnot efficiency was reversible, our Carnot engine can be
used as a refrigerator, and still δQL /δQH = TL /TH. It would be very nice (in fact, too good to be true) to
have a refrigerator that requires no work input (δW=0). For such a device, using the 1st Law, δQH= -
δQL. Computing the entropy changes of H and L, we find

∆SHL= ∆SH +∆SL = δQH /ΤΗ − δQL /ΤL = δQH (1/ΤΗ − 1/ΤL )

This implies that the entropy of the universe goes down if TH>TL), which is forbidden by the 2nd Law.
The impossibility of this work-free refrigerator is the Clausius statement of the 2nd Law.

The Kelvin-Planck statement of the 2nd Law forbids the existence of a heat engine that can produce
work without rejecting any heat to a low temperature reservoir (proof is left as an exercise).

The work-free refrigerator and the 100% efficiency heat engine are both “perpetual motion machines
of the 2nd kind” because they violate the 2nd Law. Perpetual motion machines of the 1st kind violate the
1st Law.

8.5 PROVE THIS!


1. Given the Carnot efficiency, prove that it is not possible for a device to operate on a cycle,
exchange heat with a single reservoir, and produce work (Kelvin-Planck statement of 2nd Law).

2. No heat engine operating between two fixed-temperature reservoirs can have an efficiency
greater than the Carnot efficiency.

3. The Carnot cycle must have the shape of a rectangle on a T-S diagram, regardless of the
working fluid.

4. Build a Carnot engine using air in a piston cylinder arrangement. Make the air go around the 4
processes of the Carnot cycle. Pretend that you DON’T know the second Law, and compute the
work and heat transfer for each process using the various ideal gas relations of Chapter 4 in
the textbook. Show that the efficiency is the Carnot cycle.

Concise Classical Thermodynamics 49


Clean Energy Engineering

FINAL WORDS: PROBLEM SOLVING


You can read this section before you have completed the preceding sections, but I suspect it will make
most sense after you have already worked through a dozen problems.

9.1 SOLUTION COMPONENTS


Most thermodynamics textbook problems
sound something like this: A device with
<description here> goes through a process
with <specified constraints>. The initial state
is <1 or 2 properties given> and the final
state is <1 or 2 properties given>. Find
<missing information about states or Q or
W>.

Normally this means that you have to


complete the information in all of the boxes
of Fig. 9.1 so that terms in the 1st and 2nd
Laws can be specified. Normally you will be
asked to sketch the process on a state
diagram, and even if you aren’t, you should. Gluing these boxes together, we may need some
mathematics, some physical assumptions (only if the problem has too many unknowns as posed), and
a perfectly clear understanding of thermodynamics 铰terminology
ӵ and definitions. It is impossible to
provide sequential steps that will solve every problem, and the following steps might come in different
orders.

PROBLEM STATEMENT AND SKETCH (ALWAYS FIRST!)


Translate the problem statement into a sketch that incorporates the key features of the physical
configuration. If there is information specific to the system in states 1 and 2, you should illustrate the
device in both states. Include the data that is given and the data that is desired. DO NOT put the
system boundary (dotted line) on the diagram unless the question makes it clear where to draw the
boundaries.

SYSTEM DEFINITION
Review the information given and desired, and draw your control mass (or volume) boundaries. This
is an art, but here are some useful guidelines.

1. Recall that the balance equations (eg., 1st Law) reflect the system definition exactly. The left
hand side (eg., ∆E) is always the change of some extensive property within the system. The
right hand side is always a flux across the boundary, added up over the entire area of the
boundary. If our boundaries are located where we either know information or want
information, then the resulting equations might be useful.

Concise Classical Thermodynamics 50


Clean Energy Engineering
2. Sometimes several systems are needed to solve a problem. Don’t be afraid to use the additive
nature of extensive properties to analyze complex systems. If you choose the wrong boundary
at first, you may need to go back and try another one.

3. In complex problems, the largest system boundary often yields the simplest equations because
the interactions between the components do not show up in the equations for the whole
system (this is perfectly analogous to internal forces being irrelevant to the static equilibrium
of a freebody).

USE DEFINITIONS TO TRANSLATE VARIABLES


Problem data are often given in an inconvenient form. If you are given U but want H, you can always
use H≡U+PV to make the translation. It is very helpful to have memorized which formulas are always
true and which have special “strings attached”. Start with things that are always true!

PROPERTY MODELS AND STATE DIAGRAMS


Based on the given information, select your approach to the computation of properties. At the same
time, start the construction of appropriate state diagrams. The choice of diagram varies with the
problem and the way data is given. For example, if states are specified by P and V, then clearly you will
want to make a P-V sketch. If there is any doubt about the phases involved, start a P-T sketch as well.
These sketches will help you select the property models because they can help you determine whether
the material stays as a low density gas (in which case, ideal gas relations are ok) or goes through the 2-
phase region (in which case you will need property tables or equivalent software).

PROCESS
When you have selected the property models, defined your systems, and reviewed the constraints on
your system, you can start drawing the process(es) on the state diagrams. Be careful to use dashed
curves to indicate internally irreversible processes; this will have implications if you need to use the
2nd Law or compute boundary work.

FORMULATE EQUATIONS (& ASSUMPTIONS IF NEEDED)


You probably started with the First Law as soon as the system was defined, and your selection of the
property models gave other equations. Even the use of steam tables is equivalent to a mathematical
relation eg., u=u(x, Tsat).

When you think you have a complete set of equations, determine whether you have N equations for N
unknowns. If you have too many unknowns, it may be necessary to search the problem statement for
additional constraints or clues leading to assumptions. If you have too many equations, possibly you
have made some assumptions already that were unnecessary.

CALCULATE
This should be the last step! Resist the urge to start computing before you make your sketches.

Concise Classical Thermodynamics 51


Clean Energy Engineering

9.2 BETTER THAN MEMORIZING ALL THE EXAMPLES


For new types of problems, you really do need a systematic approach. For problems that you have
already done before (but they differ only in numerical values), you can take shortcuts. I hope the next
section discourages you from taking shortcuts.

Consider the devices, materials and properties mentioned in the notes so far, and consider how many
different problems might be created without changing the basic pattern.

(a) Main substance (>5 variants): ideal gas, real gas, 2-phase mixture, liquid, solid

(b) Secondary material involved in problem (8 variants): none, metal piston walls, spring,
secondary fluid system (ideal gas, real gas, 2-phase mixture, liquid)

(c) Process type and constraints (at least 8 variants; more if these variants are chained
together).

(d) What are we asked to find ( >4 variants) : final state, initial state, Q, W

(e) What variables are used to specify states (>4 variants): (T, V), (T,P), (U, S), (P, V).

These factors a-e are almost independent, so we can mix and match all combinations. If so, then the
number of different problems is the product of all the variants: 5x8x8x4x4=5120.

This only considers problems involving up to 2 subsystems! If you are fast (2 problems per hour), you
䅀ӳ
might do all these variations in 15 weeks, assuming you do not eat or sleep. Further, this does not
consider the open system (control volume) variations of the second half of the notes. This is great
news for anyone making up final exams in thermodynamics, but very bad news for students that
expect to memorize “template” problems, hoping that the exam problems will differ only in the
specified numerical values.

A smarter approach is start out understanding that these factors a-e are almost independent. If we
learn how to deal with all the phases of H2O in one problem, we do not have to do this for every variant
of factors b-e. In this smarter approach, the number of problems we need to try is less than the SUM of
factors a to e: 5+8+8+4+4=29. This is a much more manageable number!

It is an easy thing to recommend this smarter approach, but how do you actually do it? This gets into
some deep aspects of the way we think, and there is no established way to teach it, despite the fact that
it applies to all aspects of engineering (not just thermodynamics). It is something you will need to
practice, and here are some tips:

• Don’t neglect the theory. Memorize the Laws and definitions; practice key derivations so that
you know what conditions apply to each formula.

• Make sure your system diagram always matches your equations.

• Be sure you know how to visualize the PVT surface – it is the key to thermodynamic properties.

Concise Classical Thermodynamics 52


Clean Energy Engineering
• Be sure that your calculus is in good shape. Setting up integrals or differential equations is
often key to problem solving, and if you are shaky on the mathematical glue of the problem
solving, you will probably fall back on memorizing entire example problems.

• When you do sample problems, think about how the solution would change if each part of the
problem statement was changed. Doing 1 problem thoughtfully is like doing 10 problems
without reflection.

9.3 PROBLEM SOLVING IN THE REAL WORLD


As hard as it may seem to solve all of the textbook problems in this course, solving thermodynamics
problems in the real world is much, much harder. There are several reasons, which I list here starting
with the most important:

• You may be given lots of extraneous data, and missing key information needed to solve the
problem.
• The problem is not laid out like a textbook problem
• You don’t know what simplifying assumptions can be made
• You may need to solve a dynamics or fluid mechanics problem as part of the overall solution.
• The required theory is beyond that covered in your coursework

The real world certainly is very complex, and if you consider everything going on in a real situation, no
doubt it will involve phenomena that could keep a PhD student busy for 5 years. The key to useful
application of theory is knowing how to make appropriate
婐Ӫ simplifying assumptions, and at the same
time recognizing the limitations of your calculations. This is not easy, but if you keep working at it,
you will get better at it.

Concise Classical Thermodynamics 53


Clean Energy Engineering

EXERGY (A.K.A. AVAILABILITY)AND IRREVERSIBILITY


10.1 QUALITATIVE IDEAS
The key idea is that the best process (producing the most work, or requiring the least work) is one that
involves no irreversibilities (internally AND externally). Thus, we can use the reversible process as the
standard of comparison. Exergy (also known as Availability) will be a new property defining the
“work potential” of a system within a specified environment, and the Irrversibility (also known as Lost
Work”) tells us how much more work we could have obtained using the ideal reversible process.

10.2 REVERSIBLE WORK FOR THE USUF PROCESS


Start by considering a non-ideal USUF process for the control volumes of Fig. 10.1. Like all USUF
proceses, the inlet and outlet thermodynamic states WCV QCV
Environment constant P0, T0
(T, P etc) do not change with time, as illustrated by
using the same colors for time t1 (top) and t2
Inlet i State 1
(bottom). Within the inner control volume, the Exit
flow e
state is spatially uniform, but varies with time
(different colours shown top and bottom). The
Possible ∆V
temperature of the environment is a constant T0,
CV T, P may not be const
but the surface of the smaller control surface
changes with time (call it T=T1 in the initial state,
Inlet i Exit
Tk=T2 in the final state). In Fig. 10.1, the upper left State 2
flow e
portion of the control volume boundary moves, 铰 ӵ

increasing volume by volume ∆V and resulting in


boundary work. The work and heat transfer Fig. 10.1
crossing the boundary total WCV and QCV,
respectively, regardless of whether we consider the inner or outer control surface.

From chapter 6 of these notes, recall the First and Second Laws for the control volume (taken to be the
coloured box in Fig. 10.1):

3 >? 3 >A
) , − ) ,  23 − 23 + )* =ℎ* + + 67* @ − )+ =ℎ+ + + 67+ @
 
[10.1]

1 B.
) 9 − ) 9  $1 /0
1
. /0
+ )* 9* − )+ 9+ + $1 #+<
CD
[10.2]

Consider a hypothetical perfectly reversible process yielding the same states (1,2, i, e). We do not need
to know about the details of the wonderful new frictionless machinery that will be installed within the
inner control volume, but by definition the entropy generation will be zero in Eq10.2. The work done
by this control volume, WCV,rev is not actually the maximum useful work if the environment is at
pressure P0 and temperature T0. Why? The process might have involved expansion of the control
volume; part of the work produced by the control volume must be consumed by pushing back the
atmosphere. Furthermore, if the process requires heat delivered at temperature T>T0, this costs us at
least the work required to operate a Carnot heat pump.

Concise Classical Thermodynamics 54


Clean Energy Engineering
Fig. 10.2 shows how the useful reversible work is related to the reversible work done by the control
volume. The maximum useful work
Q
produced by the control volume Wrev, useful 0 T0
(with given states 1,2, i, e and Po∆V Carnot
environment P0, T0) is: WCarnot Heat pump
Wrev, useful=WCV, rev-WCarnot-P0∆V [10.3] WCV,rev QCV,rev T
Exit e
Inlet i State 1 (changes to
Now use Eq 10.1 for the first term on
State 2 as in Fig. 10.1)
the RHS and find WCarnot in terms of T,
T0 as follows:

B.E F1 B.GH,JA F1

CE C(1)
[10.4]
Fig. 10.2
Get the last term in terms of the
properties of the stuff in the CV. Thus the maximum useful work is

* +
M+N,OP+OQ  23,M+N + ) , − ) , + )* 4ℎ* + + 67* 8 − )+ 4ℎ+ + + 67+ 8
2 2

1 .
23,M+N
− 423,M+N − :R - /08
1 :(0)


R ()  − )  )
憐Ӻ
[10.5]

This big mess is simplified by noting that the integral is the same as in Eq. 10.2. For the reversible case
we are considering, the entropy generation term is absent from 10.2, so we find

M+N,OP+OQ  ) (, +
R  ) − ) (, +
R  ) +

3 >? 3 >A
)* =ℎ* + + 67* − :R 9* @ − )+ =ℎ+ + + 67+ − :R 9+ @
 
[10.6]

Finally we write this in terms of 2 new terms, the system exergy φ and flow exergy ψ :

M+N,OP+OQ  ) S − ) S + )* T* − )+ T+ [10.7]

Where these system and flow exergies are defined as

3>
S  U − :R 9 + 
+ 67 +
R  [10.8]


3>
T  ℎ − :R 9 +  + 67 [10.9]

Concise Classical Thermodynamics 55


Clean Energy Engineering
Note that these are not true state functions because they depend on the surrounding conditions
(temperature T0). Some textbooks define the exergy using “dead state” thermodynamic properties
(u0,s0,v0) subtracted from the properties used in Eqs 10.8 and 10.9, but these terms will ultimately
cancel out thanks to conservation of mass (m1-m2+mi-me=0). Calculating exergy relative to a dead
state can be slightly more cumbersome, but if the dead state is appropriate for the ambient conditions
of your problem, the exergy will be zero if there is no potential to do work.

10.3 IRREVERSIBILITY FOR THE USUF PROCESS


The irreversibility I of the process shown in Fig. 10.1 is found by comparing the useful work it
produces, Wuseful with the ideal useful work Wrev,useful (Eq. 10.6) process using the same state changes
and environment (Fig. 10.2).

Wuseful =Wrev,useful- I [10.10]

As discussed above, the “useful” work is the control volume work minus the minimum cost of
interacting with the environment: pushing back the atmosphere and pumping heat up from T0. The
appropriate picture is Fig 10.2 again, and the equation for Wuseful is identical to Eq 10.5, but when we
try to eliminate the integral using the 2nd Law, this time the entropy generation term remains. The
result is
1
.
V  :R $1 #+< /0  :R ∆#O<*N+MP+  XF+P1 [10.11]

Thus, the irreversibility (also known as exergy destroyed,


铰ӵ Xdest) is the “lost work” and it is also directly
related to the entropy generated in the universe by irreversible processes in the control volume. We
have done a complete exergy balance on the control volume, and it is given in Eq 10.10, but it is more
often expressed with the components specified in detail:

1 B.
 −
R ()  − )  ) − = − :R $1 C(1)
/0@  ) S − ) S + )* T* − )+ T+ − V [10.12]

The most important special case is the SSSF process, in rate form. This is obtained from [10.12] by
noting that m1=m2, v1=v2, and the control volume temperature is constant with time, but may vary
spatially. We can also generalize from the 1-inlet-1-outlet case of 10.12 to the case with multiple inlets
and outlets.

. − ∑; ;. =1 − CE @  ∑* )* T* − ∑+ )+ T+ − V
C
[10.13]
D

10.4 CONCLUSIONS
Now you know that the “potential to do work” is exergy (not energy as your Grade 9 teacher might
have said, misleadingly). How is exergy used in practice? The main advantage of doing an exergy
analysis of a process is to determine how much better you might do if you could eliminate
irreversibilities. This can be valuable, but in most cases, we understand qualitatively where the lost
work occurs beforehand: friction, heat transfer with large ∆T, and mixing. The exergy analysis

Concise Classical Thermodynamics 56


Clean Energy Engineering
quantifies the lost work, directing our efforts to the most important issues. However, Eq. 10.11 shows
that the lost work is directly related to entropy generation – something we can calculate with the 2nd
Law (Eq. 10.2), knowing nothing about exergy balances. Second-Law analysis of processes gives rise
to the “2nd Law Efficiency”, typically Wuseful/Wrev,useful (this is NOT the same as the thermal efficiency of a
cycle W/QH). Again, you do not need the exergy balances to compute 2nd-Law efficiencies, but you do
need to be able to calculate reversible work. In the final analysis, I believe that the main value of the
preceding material is that it reinforces the use of 1st and 2nd Laws, and that it introduces the concepts
of reversible work and lost work.

10.5 PROVE THIS!


Exergy is a potentially confusing topic because there are many ways of presenting the information,
with new variables that are combinations of old familiar ones. Therefore, it is essential that you spend
some time with pencil and paper on the derivations yourself. A few particularly useful ones are
suggested below:

1. Prove Eq. 10.11 by following the steps given above that equation.

2. Simplify all equations and diagrams for the steady-state steady flow process. You could simply
make a copy of these notes and cross out certain terms in each equation and diagram.

3. Simplify the equations for a closed system.

4. Simplify all equations for the case that the control volume is fixed in space (no volume change)
憐Ӻ
and the surface of the control volume is T0, the environment temperature.

5. Suppose we pump an incompressible fluid through a pipe of constant diameter at constant


temperature T. The surroundings are at a lower temperature T0. There is friction in the pipe,
so the outlet pressure P2 is less than inlet P1. Prove that the exergy destruction is:

. :R
).(
 −
 ) Z1 − [
:

Concise Classical Thermodynamics 57


Clean Energy Engineering

THERMODYNAMICS OF IDEAL GAS MIXTURES


So far, we have analyzed only substances with constant composition. Soon we turn to situations where
the composition can change throughout a process through chemical reactions or physical processes
that could remove or add one species to a mixture (eg. water condensation and evaporation into air).
The control volume balances developed so far (mass, energy, entropy, exergy) are completely valid for
problems with changing composition, but property calculation (e.g., getting a number for the h in the
First Law) requires new property models. This is generally very difficult, but fortunately, many
practical problems involve mixtures of ideal gases, which can be modeled easily. One of these cases is
the modeling of moist air, considered at the end of this chapter. The other important application is
gas-phase combustion, which is complex enough to warrant a separate chapter.

11.1 PRESSURE
The key to ideal gas mixtures is the mental
P1 N1, U1, S1
cartoon Fig. 11.1 Consider 2 different gases (1, 2) PT=P1+P2
in containers of identical sizes and temperature. NT=N1+N2,
PT
The pure gases will produce a pressure UT=U1+U2,
T0,V0 ST=S1+S2
\? ]CE

*  ^_` a  1,2
3E
[11.1]
T0,V0
P2 N2, U2, S2
The total pressure PT of a gas mixture is the sum
of partial pressures:
铰ӵ
(\b \> )]CE

C 
 +
 
T0,V0
3E
[11.2]

Here we are using the universal gas constant


Fig. 11.1
(8.314 J/mol/K) and the quantity of gas N is
specified in moles. It is often convenient to work with the “mole fraction” yi =Ni/NT of species i; you
should be able to prove that Pi=yiPT.

The underlying physics behind this additivity is that in the mixture, each species “ignores the
presence” of the other species. Strictly, the molecules do interact through very frequent collisions with
each other, but there is no energy stored by these interactions, resulting in “thermodynamic
independence”.

11.2 ENERGY AND ENTHALPY


Recall that the energy and enthalpy of a constant-composition ideal gas (pure gases, or mixtures such
as air with fixed composition) were a function of temperature alone. The spacing between the
molecules does not affect u(T) and h(T). For a mixture (say, O2 and N2) it should not be surprising,
then, that the energy of O2 is not affected by the N2. Therefore, we get the energy of the mixture by
adding up the parts.

Concise Classical Thermodynamics 58


Clean Energy Engineering

11.3 ENTROPY
Recall that the entropy of an ideal gas depends on volume (or pressure) as well as temperature. In a
mixture, we can add up the entropies of each components, but we have to remember that the
molecules of each species are spread over the entire system volume (or equivalently, we must
compute entropy using the partial pressure). For an ideal gas with constant heat capacity,
C i?
9* − 9M+,*  cd,* ef − hef
CJAg iJAg
[11.3]

For the mixture of Fig 11.1, the total entropy is

C ib C i>
#C  j 9 + j 9  j kcd, ef − hef + 9M+, l + j kcd, ef − hef + 9M+, l [11.4]
CJAg iJAg CJAg iJAg

As in the previous equations, it is easier to work in moles, but mass-based formulae can be derived using
the molecular weight of each species. Equations 11.1-11.4 generalize to any number of species simply
by including more terms in the sums.

11.4 AIR-WATER MODEL


Moist air (possibly in contact with liquid water) is a very important system in science and engineering.
The preceding sections are still valid, but we have new considerations if there is the possibility of
liquid being present. The key assumptions used to憐modelӺ
this mixture are:

The air has constant moles of all species (nitrogen, oxygen, argon,
Pressure maintained at P0
carbon dioxide) and does not dissolve in the liquid.
Air (N2, O2, CO2 , Ar..); fixed moles
Water vapor partial pressure Pv is affected by the presence of and
liquid water because it can be transferred to/from a liquid phase. H2O(v); moles exchanged with liquid

If there is liquid water in equilibrium with moist air, the water


vapor partial pressure is
Liquid water, no dissolved gas,
pressure P0, temperature T0
Pv=Psat(T)

Where Psat(T) is found from the saturated water tables using the
specified system temperature (T0 in Fig. 11.2). If the water vapor is Fig. 11.2
in equilibrium with liquid water, the relative humidity is
RH=100%. Usually it is less than this, and characterized by

RH=100 x Pv/Psat(T)

Recall that the mole fraction yv=Pv/PT, so a specified mole ratio does not fix the RH. At RH=100%, we
have Pv=0.87 kPa (yv=0.0087) @ 5C and Pv=7.38 kPa (yv=0.0738) @ 40C (confirm this with your steam
tables!)

Concise Classical Thermodynamics 59


Clean Energy Engineering
The energy, enthalpy and entropy of the water should be calculated using steam-table methods to
correctly account for phase change, if it occurs. The energy, enthalpy and entropy of the air (species
other than H2O) should be calculated using the air partial pressure. Typically, constant heat-capacity
can be assumed for most engineering applications. With the ideal gas model, steam tables, and dry air
thermodynamic data, you have all you need to compute the properties of moist air. A MATLAB
program could be written to do this quite easily, but in the days before computers, this information
was compiled into the “psychrometric chart” (next page). To read this chart, a few new terms must be
defined:

• Dry bulb temperature- the temperature you would measure with an ordinary thermometer
(the true temperature of the system).
• Wet-bulb temperature (~adiabatic saturation temperature)– if you put a thin moist sponge
around the tip of a thermometer, evaporative cooling will reduce the temperature (more so if
the air is dryer). The difference between dry and wet bulb temperatures decreases as
RH100%
• Humidity ratio w=(mass of water vapor)/(mass of dry air). It is directly related (using
molecular weights) to the mole fraction.

憐Ӻ

Produced by Arthur Ogawa, License under Creative Commons


https://en.wikipedia.org/wiki/Psychrometrics#/media/File:PsychrometricChart.SeaLevel.SI.svg

Concise Classical Thermodynamics 60


Clean Energy Engineering

11.5 PROVE THIS!


1. Humidity ratio w is related to water vapor mole fraction by: w=0.622 yv/(1-yv).
2. Moist air (RH<100%, mass flow 1.5 kg/s) flows steadily into a large insulated chamber.
The humidity ratio is w1 and the temperature is T1. The chamber is wet inside, as it is
continually supplied with a small flow of liquid water at temperature T2. The out flows out
of chamber with humidity ratio w2 and temperature T2 such that RH2=100%. Show that
the First Law (with the aid of mass conservation) can be reduced to:

ℎm + n ℎN + (n − n )ℎo  ℎm + n ℎN

where subscript A refers to dry air, v to vapor and L to liquid. Pressure is 1 bar throughout. By
using a range of T1, w1 values, show that T2 is actually the wet bulb temperature, using T1 as the
dry bulb temperature. Use the steam tables for H2O properties and constant Cp assumption for air.
You will may need to iterate, guessing values of T2 and checking the energy balance.

Concise Classical Thermodynamics 61


Clean Energy Engineering

PHASE AND CHEMICAL EQUILIBRIUM – GENERAL RELATIONS


12.1 READING FROM TEXTBOOK
Cengel and Boles (5th, 6th Ed) have a chapter 13 on ideal gas mixtures, with a little on ideal solutions
and desalination. This will be useful for understanding moist air thermodynamics better; moist air is
really a special case of the material discussed in this set of notes.

The later part of these notes cover chemical equilibrium, which are covered in chapters 15 and 16 of
Cengel and Boles.

The material on the salt water system is not covered in any detail in the text and is optional reading
that will help you understand the desalination case study discussed later in the course.

12.2 NOTATION FOR MULTICOMPONENT, MULTISPECIES SYSTEMS


Now that we have a mixture of different types of molecules in a system, we compute the
extensive properties by adding the contributions from each type of molecule:

H=NA HA +NB HB
S=NA SA +NB SB
Here, NA,NB are the moles of A, B in the system (units of kmol) and HA, HB, SA, SB are the molar
enthalpies and entropies (units of kJ/kmol and kJ/kmol-K, respectively). Strictly, these are
partial derivatives keeping temperature pressure and the moles of the other species fixed:
Ӫ 兠
pq
m ≡ s
p\r \ ,C,i
t

This is a slightly odd nomenclature because we are using capital letters for the per-mole
properties (“specific” properties, for which we normally use lowercase), and some texts use an
overbar to denote the molar quantity. The nomenclature is not a big issue if you carefully note
the context of the symbol in the equation: the product NA HA must have units kJ. In fact, the
equations above could be rewritten in terms of mass and per-mass quantities.

Similar relations for internal energy and volume arise naturally:

U=NA UA +NB UB
V=NA VA +NB VB

If the system we consider is simply water vapor (A) and liquid (B), then the molar quantities
are easy to find. For vap/liq water at 100C, the steam tables give vf=0.001043 m3/kg and
vg=1.6720 m3/kg, and using MW=18 kg/kmol, we find VA=30.096 m3/kmol and VB=0.01877
m3/kmol. These specific volumes are independent of the presence of the other component – in
this case because A and B are not mixed on a microscopic level.

Concise Classical Thermodynamics 62


Clean Energy Engineering
In the more general (“non-ideal”) case, the molar volumes are influenced by the mixture ratio
of A, B (or whatever components are in the system). Whatever the merits of alcohol and water,
their mixture is non-ideal: one liter of H2O and 1 liter of pure ethanol produce a solution with
volume less than 2 liters after dissolving in each other. In such cases, we are careful to use VA,
VB to indicate the molar volumes in a particular mixture, while vA, vB are the volume of the pure
components at the same temperature and pressure. When VA≠vA, then the we have a “non-
ideal” solution or mixture, and it turns out that HA≠hA and there is non-zero “enthalpy of mixing”
and the solution will get hotter or colder than the ingredients from which it is prepared. We’ll
say more about computing the SA, HA etc. in a few pages.

12.3 EQUILIBRIUM OF A SYSTEM AT SPECIFIED T, P


The requirements for equilibrium of a non-reacting multiphase system or a chemically
reacting system are essentially the same:

1. Conservation of mass applies to all the atoms in a closed system.


2. There a constraints on the types of molecules that can exist (that is, only a limited
number of species)
3. At equilibrium, the entropy of the universe is maximized.
Consider a closed, multicomponent (possibly multiphase) system at a specified temperature
and pressure. It has been allowed to sit for a long time, so that the chemicals and phases have
arranged themselves in the most stable configuration.
憐Ӻ
We want to know if we can relate this stable composition to the
W thermodynamic properties for the individual components or species.
To keep the derivation clean, consider only two types of molecules in
the system: A and B. You could think of these as two different chemical
A<=>B
Q species (such as gas phase H2 and H) or two phases such as H2O(L) and
T, Pfixed H2O(v)). The essential feature is that A, B have very different enthalpies
and entropies even at the same T.

For constant T, P, the change in entropy of the system is reflected only


in the change in composition that might occur in the process, and these are quantified by the
changes in NA, NB:

dS=dNA SA+dNB SB

Recall that for a constant pressure heating of a simple compressible substance, the work is P∆V,
which can be combined with the energy to yield the heat transfer as

δQ=dH= dNA HA+dNB HB

Now, if there has been heat transfer δQ to our cylinder, then the surroundings have undergone
an entropy change

Concise Classical Thermodynamics 63


Clean Energy Engineering

dSsurr= - δQ/T

(imagine that this process occurs as slow as you like, keeping temperatures and pressures
spatially uniform, so that heat transfer can be considered reversible).

The relevant “universe” is the cylinder + surroundings. We can be sure that the entropy cannot
go down, but the 2nd Law does allow the entropy of the universe to go up (i.e., dSuniv= dS +dSsurr
>0). If some change in composition results in dS univ>0, then the reaction is spontaneous and
the system was not at equilibrium. We seek the condition for which dS univ=0:

dS univ=0= dNA SA+dNB SB - δQ/T= dNA SA+dNB SB – (dNA HA+dNB HB)/T

Since the LHS of the equation is zero, we are free to multiply by (-T) to get

0= dNA(HA - T SA)+ dNB(HB - T SB)

The quantities in the parentheses are the molar Gibbs Free Energies (GA, GB ) also called the
“chemical potential” (µ). In general, the Gibbs Free Energy of a system is simply H-TS, and I will
state (without proving) that maximizing Suniv is equivalent to minimizing G at fixed T and P. Our
condition at equilibrium can be written as

dNA µA + dNB µB=0


憐Ӻ
If we had 7 different types of molecules in the system,

dNA µA +dNB µB + dNC µC+ dND µD + dNE µE+ dNF µF + dNG µG=0

This isn’t immediately helpful unless we have methods to calculate chemical potentials (these
depend on the substance and property models we choose) and a way to relate the dNA, dNB etc.
We’ve been using water liquid+vapor as an illustrative example; in this case dNA= - dNB because
if 1 mole of vapor is formed in the system, then we must have lost 1 mole of liquid (total moles
are conserved if there is no chemical reaction).

dNA µA + (-dNA )µB= dNA (µA - µB )=0

Go to your steam tables now and confirm that all along the saturation curve, µliq = µvap!

The arguments given above apply equally to phase and chemical equilibrium, but the mechanics
of calculations are quite different in the two cases.

Concise Classical Thermodynamics 64


Clean Energy Engineering

CHEMICAL REACTIONS FOR IDEAL GAS MIXTURES


When chemical reactions occur in a system, the component atoms can “shuffle around” between
different chemical species until Suniv is maximized. Conceptually, this is the same as the
situation for phase equilibrium, but the changes in the numbers of moles are linked in a more
complicated way. We need to review reaction stoichiometry carefully.

13.1 FUELS AND REVIEW OF STOICHIOMETRY


Reaction stoichiometry must be specified before other analysis. For example, methane and

oxygen may react as follows:

CH4 + 2 O2  1 CO2 + 2 H2O

In general, we can think of a chemical equation involving species A1, A2… Ai as a sum:

∑ν i Ai = 0
i

In the methane oxidation,

νCH4 = 1
㡰ӳ
νO2 = 2

νCO2 = -1

νH2O = -2

Recall that the number of each type of atom must be conserved in a chemical reaction. The
reaction above describes combustion with the stoichiometric amount of oxygen. Often, fuels are
burned with more (and occasionally less) than this amount. The equivalence ratio ϕ is the
amount by which the fuel exceeds the stoichiometric amount of fuel. When fuels are burned in
air, the stoichiometry is:

CH4 +νO2 (O2 +3.76 N2) νCO2 CO2 + νH2O H2O + 3.76 νO2 N2

Note that methane and oxygen can react to form different products:

CH4 + x O2  y ([1-ε] CO2 +ε CO)+ z H2O

In this case, we can find x,y,z if we know ε, but ε is determined by chemical kinetics and
equilibrium. Small amounts of other compounds (soot, nitrogen oxides) can also be formed.
These side reactions are very important for environmental reasons, but often the
thermodynamics of the heat release are well approximated by “complete combustion”.

Concise Classical Thermodynamics 65


Clean Energy Engineering

13.2 ENTHALPY OF FORMATION AND ENTHALPY OF REACTION


We need a consistent way of defining thermodynamic properties when reaction is possible:

Reference State: a pure element in its usual state at 0.1MPa and 298K has an
“enthalpy of formation hf0” of zero.

hf0 of other molecules is determined experimentally, conceptually in an isothermal flow reactor


at 298K.

Reaction chamber
Species of
Elements in interest e.g. CO2
e.g. C(s), O2
Heat transfer needed to keep
everything at 298K

The “standard enthalpy of reaction” ∆HR is the憐heat


Ӻ transfer to the reactor when products and
reactants are at standard state. As you know, the reaction of carbon with oxygen to form CO2 is
exothermic, so the heat transfer is negative, and hf0 for CO2 is -393,522 kJ/kmol. If reactants and
products are not at standard state, the first law still applies, but we need to know how
properties for each species vary with temperature and pressure.

The follow example has some blank spaces that you should be able to work through. The
answers will be posted later- try it on your own first!

Concise Classical Thermodynamics 66


Clean Energy Engineering

EXAMPLE – HEATING A HOUSE WITH A GAS FURNACE.

Suppose that the heat loss from a house is 5 kW. If a gas furnace (80% efficiency) is used, what
is the required natural gas flow rate?

Take the house as a control volume and apply the 1st Law:

Now use the 1st Law on a control volume around the furnace itself, neglecting the blower motor
power:

We must consider the chemical reaction in the furnace. Natural gas is not pure methane, and
there are products other than water and carbon
兠Ӫ dioxide (including pollutants), but the main
reaction is:

CH4 +_ (O2 +3.76 N2)  CO2 + __H2O + __(3.76)N2

The 1st Law above was written in terms of mass flow, but we can easily work in moles and get
the heat transfer for 1 kmol of natural gas (assumed to be methane) burned. For simplification,
assume the reactants and products are at 25C. From the tables at the back of the text, we get
thermodynamic properties.

Species hf kJ/kmol Mole reacted n nhf


CH4 -74,873 1
O2 0
N2 0
H2O(v) -241,827
CO2 -393,522
N2 0

Sum

Concise Classical Thermodynamics 67


Clean Energy Engineering
Thus, the enthalpy of reaction (heat transfer from the isothermal combustor) is ___________kJ/kmol. The
molecular weight of methane is 16 kg/kmol, so the heating value on a mass basis is 50,000 kJ/kg. If
the furnace is “80% efficient”, 20% of the heating value goes up the chimney.

Thus, the natural gas flow rate needed is:

憐Ӻ

Concise Classical Thermodynamics 68


Clean Energy Engineering

13.3 ADIABATIC FLAME TEMPERATURE


So far, we have applied the first law to SSSF systems
maintained at constant temperature by heat transfer
(exactly equal to the heat of reaction). What happens Initial state Final State
H(T) for reactants
if the reactor is insulated and no work is done?
Applying the SSSF version of the First Law to this
insulated control volume, you find that the enthalpy
per unit mass of the reactants must be equal to the
products. This is represented in the sketch by having H ∆HReaction
the intial and final states along a horizontal line in an
H-T plot. To do the calculations, we break the H(T) for products
combustion into two imaginary processes. First, the
reaction occurs isothermally with heat transfer
∆Hreaction. Next, the product mixture is heated up until
we get back to the original enthalpy (moving along the
H(T) curve). 298K T
T adiabatic

13.4 CHEMICAL EQUILIBRIUM


Now we turn back to chemical equilibrium. In兠Ӫtheory we have already solved the problem of
chemical equilibrium. Recall that if there are k chemical species involved in a particular
chemical reaction going on in the system, maximizing Suniv requires that
;

u /j* v*  0
*w

Now the mole changes are not the same for each species, but rather linked by stoichiometry of
the reaction through dNi=νi dζ, where z is a “progress variable for the reaction”. No need to
worry about z much because the main ideas is that subbing this back into the condition above
gives:
;

u y* v*  0
*w

We’ll have an equation like this for each chemical reaction that can occur. The stoichiometric
coefficients will be known, but what about the chemical potentials? Generally µι(T,P, x1, x2,
x3…..)

Concise Classical Thermodynamics 69


Clean Energy Engineering
where xi is the mole fraction in whatever phases are present. (These mole fractions are the
quantities that we want to obtain at the end of the day). In general, relating chemical potential
to composition is a hard problem for non-ideal mixtures. Now we consider ideal gas mixtures
to simplify computations. For ideal gas mixtures,

T  0 T
c  x P 
µj = h 0
f,j + ∫ c pj dT ' − T  s f , j + ∫ pj dT ' − R ln j 
T0  T0 T '  P0 

The first 2 terms on the right give the molar enthalpy (Hj) according to the nomenclature introduced
at the beginning. Ideal gas mixtures are ideal in the sense discussed at the beginning of this
chapter: Hj=hj (the enthalpy per mole of each species is unaffected by the presence of the other
species). The baseline for enthalpy was an arbitrary convention.

The big term on the RHS is TSj. Think of the 3 terms in [ ] as follows:

• Entropy at standard state for the pure species

• Correction for temperature not at T0=298K

• Correction for partial pressure xiP not at P0=1 atm.

The baseline for entropy is determined by the憐Third


Ӻ Law of Thermodynamics.6

Notice that only the very last term for the chemical potential depends on the species
concentration; other terms can be combined as a single term and we can simplify by using
partial pressure Pi=xiP and using pressure in atmospheres (so that P0) need not be written.

v* (:,
* )  v* (:,
R ) + h:ef(
* )

v* (:,
* )  v* z (:) + h:ef(
* )

Now our condition for equilibrium is

6 The entropy is zero for a perfect crystal at absolute zero temperature.

The 3rd law is relevant to

1. Chemical reactions and thermodynamic properties at very low temperatures.

2. Establishing the baseline absolute entropies tabulated, for example, in your texts.

The odds of you needing this in an engineering career are astronomically low, but you might use the Third Law to
impress people at cocktail parties.

Concise Classical Thermodynamics 70


Clean Energy Engineering
;

u y* {v* z (:) + h:ef(


* )|  0
*w

Or
; ;
y* v* z (:)
, } ~− u   €
* ?
h:
*w *w

Suppose our reaction is CO2CO+1/2 O2. The equilibrium expression for this reaction would
be:
z
v2… z − v2… z − v…
2…
1/ƒ+„  , } ~− 
h:

R.‡ …>

At room temperature, carbon dioxide is strongly favoured, so the partial pressures of CO and O2 would
be near zero. This is consistent with the free energy of CO2 being very low relative to CO and O2: the
LHS is the exponential of a positive term (>>1).

The LHS of this equation is the equilibrium constant 1/Keq (clearly, this “constant” varies with
temperature!). For the RHS, the partial pressures must be linked together – we don’t really have 3
unknowns. This linkage comes from the setup of the problem and the initial conditions. For example,
憐Ӻ
if we start with 1 mole of CO2 and we form z moles of CO, then the moles of each species will be:

NCO2= 1-z NCO= z NO2= z/2 Ntotal=1-z +z +1/2 z= 1+1/2 z

If the total pressure in the system is P, then the partial pressures will be
ˆ‰ ‰ ‰/

2… 
;
2… 
;
… 

‰/ ‰/ ‰/

Putting these into the expression for equilibrium constant, we get


z
v2… z − v2… z − v… 1−7 7
1/ƒ+„  , } ~−   R.‡ (1 + )R.‡
h: 77 2

Finally, 1 equation in 1 unknown (z). Unfortunately, it is a non-linear algebraic equation. The


combustion of methane involves dozens of chemical species and hundreds of chemical
reactions. Computers are necessary!

Concise Classical Thermodynamics 71


Clean Energy Engineering

13.5 MODERN METHOD OF FINDING EQUILIBRIUM COMPOSITION


Other old texts will have many examples of how you do equilibrium calculations by hand, like
the one on the previous page. This is one area where computer methods are far superior.
Download the free GASEQ at http://www.arcl02.dsl.pipex.com/ and try the following exercises:

1. Problem Type: adiabatic T, constant P; methane/air flame (a standard setup)

• Program will automatically calculate 2226 adiabatic flame temperature


for 300K input

• T0=300 K (hit calculate) 2367 K output (note changed composition of


products)

• Try changing the amount of O2 or N2 – what happens?

2. Carbon, hydrogen and oxygen atoms are introduced into a reaction vessel at temperature T
and pressure P. Although the ingredients are added in the form of methane, oxygen and carbon
monoxide, the final composition depends only on the C:H:O ratio and the species allowed in the
mix. For T=2000K, P=10 atm, find the final composition if the initial composition is a ratio of
C:H:O =100:10:10 The possible species in the final mixture are CO, CO2, H2, C(s), H2O.
憐Ӻ
3. Try the hydrogen dissociation example of the last page.

Concise Classical Thermodynamics 72


Clean Energy Engineering

MULTICOMPONENT PHASE EQUILIBRIUM


Multiphase, multicomponent systems are common in nature. Consider the 3-component
system (C=3) consisting of water, salt and air (modeled as pure N2 in our cartoon below). We
could easily have a gas phase, liquid solution, solid ice and solid salt present in equilibrium (P
=4). In the most general case, any of the components could be present in trace quantities in
any of the phases, although in practice the solid phases might be relatively pure (as shown in
the cartoon). Certainly, there can be many salt molecules dissolved in the water and many
water molecules present as water vapor. The green arrows  indicate that the molecules
can jump between phases. For example, there is the sublimation “reaction”:

H2O(s)  H2O(v)

There will be a net H2O


exchange between the two N2

phases until the entropy of the GAS


universe is maximized. Recall H O(s) that
2
this means minimizing the Gibbs
energy G for the system, and this in
LIQUID
turn imposes a condition on the
chemical potential: µS = µV . NaCl
兠Ӫ
Now, if the solid phase is pure
water, then adding or
subtracting water to it does not NaCl(s)
change the chemical potential µS.
However, the vapor phase entropy is affected by the amount of water vapor present in the box.
Recall that the entropy varies as –R ln(Pi), so as Pi increases, the entropy per mole decreases.
G=H-TS, so decreasing entropy means higher G (and µ). So, as water moves to the vapor phase,
µV will increases until it matches the ice chemical potential. The microscopic kinetic picture
of this may be more appealing to you. There is always a tendency for molecules of ice to become
ejected into the gas phase. If there is no water in the gas phase, then none can come back and
land on the solid surface. However, as the concentration of water vapor increases, the reverse
reaction (vs) becomes faster, until finally forward and reverse reaction rates are balanced.

While the ice and water vapor are busy equilibrating, all the other components and
phases are also evolving towards the same “goal”: increasing Suniv.

14.1 GIBBS PHASE RULE


Gibbs considered the equilibrium of the most general multiphase system. Suppose there are C
components (chemical species) in a system with P phases co-existing. The conditions of equilibrium
are:

Concise Classical Thermodynamics 73


Clean Energy Engineering
T is the same for all phases (P-1 equations)

P1=P2=..PP (P-1 equations)

µi1= µi1 =….µiP C(P-1 equations)

For each phase, the state would be fixed by T,P and C-1 mole fractions (C+1) variables. The
number of intensive properties that can be varied independently is:

F=P(C+1)-(C+2)(P-1)=2+C-P

This is the Gibbs phase rule, and all phase diagrams must conform!
For a single component system, at most 3 phases can coexist (at the triple point; here both T
and P are already fixed and F =0). The Gibb’s phase rule is closely related to the number of
properties needed to fix the state of a system – but it is slightly different. For example, at the
triple point of water, T and P are fixed, but the relative amounts of liquid, water and vapor
would still need to be specified (2 variables – as always for a SCPS).
For a 2-phase steam mixture, Gibbs says F =1, meaning that if we specify T, then P would be
fixed. However, the quality or some other information about the relative amounts of gas and
liquid would still need to be specified.
Multicomponent systems can contain distinct 憐 solid
Ӻ phases, immiscible liquid phases as well as
a gas phase. Generally, each time a component is added, the number of properties needed to
fix the state goes up by 1. Returning to the salt-water-nitrogen system, we expect F=2+C-
P=2+3-4=1. This means that you could vary a single property (say, temperature)
independently, but all of the other properties would be fixed. For example, at 20 C, there
would be some definite pressure at which all 4 phases could exist, and the concentrations of
components in each phase would be fixed as well.

14.2 THE SALT-WATER-AIR SYSTEM


The thermodynamics of the 3 component system consisting of NaCl, H2O and “air” (pseudo pure
gas) controls, to a large extent, the climate of earth (atmosphere and ocean!) and is also of great
technological interest. Desalination of seawater, for example, is an energy intensive process
that involves this system. In order to predict the energy requirements of desalination, we need
a way of calculating the thermodynamic properties (eg., h, s, v as a function of T, P, composition)
and the phase boundaries that result from this property model.

Concise Classical Thermodynamics 74


Clean Energy Engineering

Figure 14.1 Saltwater phase diagram. Source: Wikipedia http://en.wikipedia.org/wiki/File:WatNaCl.png

We will work with a highly simplified model for this system, based on these assumptions:

• The vapour phase is ideal and contains only H2O(v) and pseudopure air.
• The liquid phase contains only H2O(L) and dissolved NaCl. No air is dissolved in the
liquid, and the solution is considered “ideal”.
• The model does not describe solid water (ice) or salt. This restricts are calculations to
roughly T>0 oC and xs<0.263 (see diagram above). If anyone is interested, we could
probably extend the model.

14.2.1 VOLUME: GAS


As an ideal mixture, the volume of the gas can be determined from the ideal gas law knowing
the total pressure P and the total moles (the sum of the 2 components in the mixture):

(j‹ + jm )h:R


Recall that we can work in mass rather than moles, using the molecular weights MW:
)m
(Ž
Œ
+ )h:R
m



If you desired specific volume, v=V/(mw+mA) and with a little algebra you get a relation
involving humidity ratio w and the specific volumes that would be computed by the ideal gas
law for pure components.
1−n
(Ž
‹
+ )h:R
m
  nŽ + (1 − n)m


14.2.2 VOLUME: LIQUID

Concise Classical Thermodynamics 75


Clean Energy Engineering
The solution of salt (mass fraction xs) will be nearly incompressible but ideal in the sense that
specific volume will vary linearly with xs.

(:,
, P )  (1 − P )‹ + P P

Here, vw is the specific volume of pure water at the specified T and P; it can be determined
from the steam tables in the usual way or from Xsteam. The specific volume of pure salt can
be taken as a constant vs=0.000348 m3/kg (ρ=2875 kg/m3). Actually, NaCl(s) has a density of
2165 kg/m3, but the dissolved salt is dissociated and the ions fit nicely next to water
molecules, so the solution behaves as if the salt has a higher density.

14.2.3 ENTHALPY: GAS


We model the specific enthalpy h [kJ/kg] as is normal for HVAC moist air calculations. You
could use psychrometric charts, but for integration into spreadsheet models, it is more
convenient to specify the equation:

ℎ  nℎŽ + (1 − n)ℎm

For the water vapour (first term), again use Xsteam with the desired T and partial pressure.
Pressure should not affect hW, but if you mistakenly specify P>Psat(T), Xsteam will tell you
that you have liquid rather than vapour. For the air, the enthalpy can be modelled simply as
hA=Cp(T-T0) where Cp=1.005 kJ/kg/K and T0=0 oC (hA=CpT if T in oC).

14.2.4 ENTHALPY: LIQUID


The ideal solution enthalpy will will vary linearly with xs.

ℎ(:,
, P )  (1 − P )ℎ‹ + P ℎP

Here, hw is the specific volume of pure water at the specified T and P; it can be determined
from Xsteam. The specific enthalpy of pure salt can be modeled as a constant volume
material:

ℎP  cP : +
P

Here, Cs is the effective heat capacity of the salt in solution and vs is the pure salt effective
volume in solution (from above). The heat capacity of NaCl(s)~0.86 kJ/kg/C is probably quite
different from the dissolved salt, but we’ll take Cs=0.86 kJ/kg/C out of shear laziness (and it
will probably not have a big influence on calculations). Use T in 0C to set a consistent (but
arbitrary) reference level for the salt enthalpy.

If you wanted internal energies, they could be computed from u=h-Pv, now that we have
complete models for h and v for any mixture, T, P.

14.2.5 ENTROPY: GAS

Concise Classical Thermodynamics 76


Clean Energy Engineering
This is an ideal gas mixture of air and water vapour. It is somewhat easier to write the
relations in terms mole rather than mass fractions. Let yw=water mole fraction. The entropy
per mole of gas s can be written in terms of the molar entropies of the water and air.

9  ‘‹ #‹ + (1 − ‘‹ )#m

Caution: Sw and SA are NOT the entropies that we would compute for the pure
components at the given T, and total pressure P. Instead, these must be evaluated at the
correct partial pressure for each species. We have already stated that we will model the air
with constant Cp=(1.005 kJ/kg/K)(0.018 kg/mol) so to be consistent,
C (ˆ”Œ )i
#m  cd ln ’C s −R ln ’ iE
s
E

The reference state conditions are arbitrary, but as for the enthalpy, we need to be
consistent in all calculations. Choose T0=273.15 K (0 oC) and P0=1 atm for So=0. For the water
vapour entropy, we can use the Xsteam function s_pT(ywP, T) to get the entropy in kJ/kg/K;
multiply by 0.018 kg/mole to get SW in the correct units.

憐Ӻ

Concise Classical Thermodynamics 77


Clean Energy Engineering

14.2.6 ENTROPY: LIQUID SOLUTION


Surprisingly, the ideal salt solution is similar to an ideal gas mixture in several important
respects. Firstly, the volumes of the pure components can be added (weighted by their
fraction of the mixture) to find the volume of the mixture at the same T, P. Secondly, the
entropy of the mixture increases with dilution (lower mole fraction) in the same manner as
the ideal gas mixture. For the whole mixture,

9  ‘P #P + (1 − ‘P )#‹

Here, ys is the mole fraction of the NaCl in solution.

#Ž  9‹,dOM+ − hln•‘‹ –

#—  9P,dOM+ − hln•‘P –

This is not exactly correct because NaCl is dissociated in solution; we have ions in solution Na+
and Cl-. With 2 moles of ions for every mole of NaCl that goes in the brine, the concentrations
in the entropy equations should be corrected. After some algebra that there is no need to
worry about here, a better model of entropy is:

1 − ‘P
#Ž  9‹,dOM+ − hln Z [
1 + ‘P

2‘P
#—  9P,dOM+ − hln Z [
1 + ‘P

As in other parts of the model, the water properties will be computed with Xsteam:

Sw,pure= s_pT(P,T)*0.018 kg/mol

The salt entropy will be modeled with the constant heat capacity introduced earlier Cs=0.86
kJ/kg/C * (0.0584 kg/mol), taking the arbitrary reference state of zero entropy at 0 oC.
C
9P,dOM+  cP ln ’˜™.‡šs

Concise Classical Thermodynamics 78


Clean Energy Engineering

14.2.7 PHASE BOUNDARIES


The preceding equations for v, h, s as a function of T, P and composition provide constitute a
complete thermodynamic model for the salt-water-air system in that everything else we
might want could be computed from the equations above (given enough expertise in
thermodynamics). For convenience, some additional useful relations will be provided,
starting with the phase boundaries. You should know from experience that salty water boils
at a higher temperature (equivalently, its vapour pressure is reduced by the presence of salt).
Also, water evaporates into dry air at a temperature far below the boiling temperature. Both
effects come into play in our salt-water-air system (if both liquid and vapour phases are
present). At equilibrium, the water vapour partial pressure is given by
ˆ”
‘‹
 ”›
Pœ1 (:)
›

The LHS of this equation involves the vapour-phase concentration of water; the RHS
describes the reduction in water concentration in liquid, which reduces the vapour pressure
from the pure-water value of Psat(T). This equation is valid for the situation with no air
(yw=1), in which case the boiling pressure would be determined. This equation is also valid
for the air-water system with no salt (ys=0) – we get the relation we had in the “moist air”
lectures of the course.

憐Ӻ

Concise Classical Thermodynamics 79

Vous aimerez peut-être aussi