0% found this document useful (0 votes)
54 views33 pages

Thermodynamics of Gas Turbine Cycles

1. Thermodynamics is the science that deals with energy, work, heat, and their transformation between a system and its surroundings. It was developed in the 19th century to study steam engines. 2. A system is defined along with its boundary and surroundings. A system can be closed, open, isolated, or adiabatic depending on whether matter and energy are transferred across its boundary. 3. Thermodynamic properties like temperature and pressure must be uniform throughout a system in thermodynamic equilibrium for the system properties to be well-defined. A thermodynamic process is a change in a system from one equilibrium state to another.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
54 views33 pages

Thermodynamics of Gas Turbine Cycles

1. Thermodynamics is the science that deals with energy, work, heat, and their transformation between a system and its surroundings. It was developed in the 19th century to study steam engines. 2. A system is defined along with its boundary and surroundings. A system can be closed, open, isolated, or adiabatic depending on whether matter and energy are transferred across its boundary. 3. Thermodynamic properties like temperature and pressure must be uniform throughout a system in thermodynamic equilibrium for the system properties to be well-defined. A thermodynamic process is a change in a system from one equilibrium state to another.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 33

Thermodynamics of Gas Turbine Cycles

1.1 Introduction

Thermodynamics is the science that deals with energy production, storage,


transfer and conversion. It studies the effects of work, heat and energy on
a system. Despite the fact it is a very broad subject that affects most fields
of science including biology and microelectronics, we will concern mostly
with large scale observations. Small scale interactions will be described in
the kinetic theory of gases.
Historically, thermodynamics was born in the 19th century as scientists
were first discovering how to build and operate steam engines. Particularly
through the work of French physicist Nicolas Léonard Sadi Carnot who
introduced the concept of the heat-engine cycle and the principle of
reversibility in 1824. Scottish physicist Lord Kelvin was the first to
formulate a concise definition of thermodynamics in 1854. Carnot’s work
concerned the limitations on the maximum amount of work that can be
obtained from a steam engine operating with a high-temperature heat
transfer as its driving force. In later years the laws of thermodynamics
were developed. Thermodynamics is principally based on a set of four
laws which are universally valid when applied to systems that fall within
the constraints implied by each.

Our goal here will be to introduce thermodynamics as the energy


conversion science, to introduce some of the fundamental concepts and
definitions that are used in the study of engineering thermodynamics.
These fundamental concepts and definitions will be further applied to
energy systems and finally to thermal or gas power plant.

1
1.2. Thermodynamic Systems
1.2.1. System, Boundary and Surroundings
System. A system is a finite quantity of matter or a prescribed region of
space, see figure (1-1).

Boundary.Theactual or hypothetical envelope enclosing the system is the


boundary of the system. The boundary may be fixed or it may move, as
and when a system containing a gas is compressed or expanded. The
boundary may be real or imaginary. It is not difficult to envisage a real
boundary but an example of imaginary boundary would be one drawn
around a system consisting of the fresh mixture about to enter the cylinder
of an I.C. engine together with the remnants of the last cylinder charge after
the exhaust process, as shown in figure (1-2) .

Fig.1-1: The system Fig.1-2: The real and imaginary boundaries

1.2.2. Closed System


Refer to figure (1-3). If the boundary of the system is impervious to the
flow of matter, it is called a closed system. An example of this system is
mass of gas or vapour contained in an engine cylinder, the boundary of
which is drawn by the cylinder walls, the cylinder head and piston crown.
Here the boundary is continuous and no matter may enter or leave.

2
Fig.1-3: Closed system . Fig.1-4: Open system.

1.2.3. Open System


An open system is one in which matter flows into or out of the system.
Most of the engineering systems are open, as shown in figure (1-4).

1.2.4. Isolated System


An isolated system is that system which exchanges neither energy nor
matter with any other system or with environment.

1.2.5. Adiabatic System


An adiabatic system is one which is thermally insulated from its
surroundings. It can, however, exchange work with its surroundings. If it
does not, it becomes an isolated system.

Phase. A phase is a quantity of matter which is homogeneous throughout


in chemical composition and physical structure.

1.2.6. Homogeneous System


A system which consists of a single phase is termed as homogeneous
system. Examples: Mixture of air and water vapour, water plus nitric acid
and octane plus heptane.

1.2.7. Heterogeneous System


A system which consists of two or more phases is called a heterogeneous
system. Examples: Water plus steam, ice plus water and water plus oil.

3
1.4 Pure Substance
A pure substance is one that has a homogeneous and invariable chemical
composition even though there is a change of phase. In other words, it is a
system which is (a) homogeneous in composition, (b) homogeneous in
chemical aggregation. Examples: Liquid, water, mixture of liquid water
and steam, mixture of ice and water. The mixture of liquid air and gaseous
air is not a pure substance.

1.5 Thermodynamic Equilibrium


A system is in thermodynamic equilibrium if the temperature and pressure
at all points are same; there should be no velocity gradient; the chemical
equilibrium is also necessary.

Systems under temperature and pressure equilibrium but not under


chemical equilibrium are sometimes said to be in metastable equilibrium
conditions. It is only under thermodynamic equilibrium conditions that the
properties of a system can be fixed.

Thus for attaining a state of thermodynamic equilibrium the following


three types of equilibrium states must be achieved:

1. Thermal equilibrium. The temperature of the system does not


change with time and has same value at all points of the system.

2. Mechanical equilibrium. There are no unbalanced forces within the


system or between the surroundings. The pressure in the system is same at
all points and does not change with respect to time.

3. Chemical equilibrium. No chemical reaction takes place in the


system and the chemical composition which is same throughout the system
does not vary with time.

4
1.6 State
State is the condition of the system at an instant of time as described or
measured by its properties. Or each unique condition of a system is called
a state.

It follows from the definition of state that each property has a single value
at each state.

Stated differently, all properties are state or point functions. Therefore, all
properties are identical for identical states.

On the basis of the above discussion, we can determine if a given variable


is property or not by applying the following tests:

— A variable is a property, if and only if, it has a single value at each


equilibrium state.

— A variable is a property, if and only if, the change in its value between
any two pre-scribed equilibrium states is single-valued.

Therefore, any variable whose change is fixed by the end states is a


property.

1.7 Thermodynamic Process


A Process is a change of state of a system from an initial to a final state
due to an energy interaction (work or heat) with its surroundings. For
example in the following Process diagrams plotted by employing
thermodynamic properties as coordinates are very useful in visualizing the
processes. Some common properties that are used as coordinates are
temperature T, pressure P, and volume V (or specific volume v).

Figure (1-5) shows the P-V diagram of a compression process of a gas.

5
Note that the process path indicates a series of equilibrium states through
which the system passes during a process and has significance for quasi-
equilibrium processes only.

Fig.1-5: Process in the piston-cylinder device

The Process Path defines the type of process undergone. Typical process
paths are:

1.7.1 Isothermal Process


A process is said to be isothermal if the temperature of the system remains
constant during each stage of the process. Please note that a process being
isothermal does not imply anything about the heat transferred or work
done, i.e. heat transfer may take place during an isothermal process. An
isothermal process implies that the product of the volume and the pressure
is constant for an ideal gas.

1.7.2 Isochoric Process


A process is said to be Isochoric when the volume is held constant. In any
Isochoric process, the work done by the system is always zero. For any two
dimensional system, the heat energy transferred to that system is absorbed
by it as its internal energy. The other name of this process is isometric
process. Example – When we heat any empty container, the air inside gains

6
internal energy which can be felt due to increase in pressure and
temperature.

1.7.3 Isobaric Process


A process is said to be isobaric if the pressure of the system remains
constant during each step of the process. The heat transferred to the system
does work but also changes the internal energy of the system.

1.7.4 Adiabatic Process


A process is said to be adiabatic process. An adiabatic process is also
known as isocaloric process which is a thermodynamic process in which
no heat is transferred to or from the working fluid.if no heat enters or leaves
the system during any step of the

Conversely, a process that involves heat transfer (addition or loss of heat


to the surroundings) is generally called diabatic.

1.8 Entropy
The ‘thermodynamic flow’ quantity to be considered in heat engines has
𝑄̇
the value of . This quantity is conserved in a reversible engine; and is
𝑇

referred to as the ‘entropy’ flow rate. If the heat engine is ‘irreversible’


there is a production of entropy as the energy passes through the machine.
There is no tangible analogue between this entropy production and the flow
of ‘energy’ in the hydraulic machine. A direct consequence of this is the
Clauses inequality which has no counterpart in any other field of physical
science.

𝛿𝑄
∮ ≤0 (1-1)
𝑇

The significance of the inequality sign is that:

𝛿𝑄
for reversible processes ∮ =0 (1-2)
𝑇

7
𝛿𝑄
While; for irreversible processes ∮ <0 (1-3)
𝑇

Consideration of a cycle for a reversible engine leads to the conclusion


𝛿𝑄𝑅
that ∮ =0 (1-4)
𝑇

1.9 Thermodynamic Cycles

A thermodynamic cycle consists of a series of thermodynamic processes


transferring heat and work, while varying pressure, temperature, and other
state variables, eventually returning a system to its initial state.In the
process of going through this cycle, the system may perform work on its
surroundings, thereby acting as a heat engine.

State quantities depend only on the thermodynamic state, and cumulative


variation of such properties adds up to zero during a cycle. Process
quantities (or path quantities), such as heat and work are process
dependent, and cumulative heat and work are non-zero.

The thermodynamic cycles can be divided into two primary classes:

- Power cycles Thermodynamic power cycles are the basis for the
operation of heat engines, which supply most of the world's electric
power and run almost all motor vehicles. Power cycles can be
divided according to the type of heat engine they seek to model. The
most common cycles that model internal combustion engines are the
Otto cycle, which models gasoline engines and the Diesel cycle,
which models diesel engines. Cycles that model external combustion
engines include the Brayton cycle, which models gas turbines, and
the Rankine cycle, which models steam turbines.
- Heat pump cycles
Thermodynamic heat pump cycles are the models for heat pumps
and refrigerators. The difference between the two is that heat pumps
8
are intended to keep a place warm while refrigerators are designed
to cool it. The most common refrigeration cycle is the vapor
compression cycle, which models systems using refrigerants that
change phase. The absorption refrigeration cycle is an alternative
that absorbs the refrigerant in a liquid solution rather than
evaporating it.

1.9.1 Brayton Cycle


The ideal cycle for gas turbine is Brayton Cycle or Joule Cycle. This cycle
is of the closed type using a perfect gas with constant specific heats as a
working fluid. This cycle is a constant pressure cycle and is shown in
Figure (1-6). On P-V diagram and in Fig(1-7) on T-φ diagram. This cycle
consists of the following processes:

Fig .1-6: Brayton cycle on P-V diagram Fig .1-7: Brayton cycle on T-φ diagram

The cold air at 3 is fed to the inlet of the compressor where it is compressed
along 3-4 and then fed to the combustion chamber where it is heated at
constant pressure along 4-1. The hot air enters the turbine at 1 and expands
adiabatically along 1-2 and is then cooled at constant pressure along 2-3.

Heat supplied to the system = 𝐾𝑝 (𝑇1 − 𝑇4 ) (1-5)

Heat rejected from the system = 𝐾𝑝 (𝑇2 − 𝑇3 ) (1-6)

where Kp is Specific heat at constant pressure,

9
Work done = Heat supplied – Heat rejected
=𝐾𝑝 (𝑇1 − 𝑇4 ) − 𝐾𝑝 (𝑇2 − 𝑇3 ) (1-7)
Thermal efficiency (η) of Brayton Cycle :
𝑊𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 [𝐾𝑝 {(𝑇1 −𝑇4 )−𝐾𝑝 (𝑇2 −𝑇3 )}]
𝜂= =
𝐻𝑒𝑎𝑡 𝑆𝑢𝑝𝑝𝑙𝑖𝑒𝑑 𝐾𝑝 (𝑇1 −𝑇4 )

(𝑇 −𝑇3 )
𝜂 = 1 − (𝑇2 (1-8)
1 −𝑇4 )

For expansion 1-2


(𝛾−1)⁄
𝑇1 𝑃1 𝛾
=( )
𝑇2 𝑃2
(𝛾−1)⁄
𝑃1 𝛾
𝑇1 = 𝑇2 [( ) ] (1-9)
𝑃2

For compression 3-4


(𝛾−1)⁄ (𝛾−1)⁄
𝑇4 𝑃4 𝛾 𝑃1 𝛾
=( ) =( )
𝑇3 𝑃3 𝑃2
(𝛾−1)⁄
𝑃1 𝛾
𝑇4 = 𝑇3 [( ) ] (1-10)
𝑃2

Substituting the values of Tl and T4 in equation (1-8), we get:

(𝑇2 −𝑇3 )
𝜂 =1− (𝛾−1)
⁄𝛾
(1-11)
𝑃1
(𝑃 ) (𝑇2 −𝑇3 )
2

1.9.2 Rankine Cycle


Steam engine and steam turbines in which steam is used as working
medium follow Rankine cycle. This cycle can be carried out in four pieces
of equipment joint by pipes for conveying working medium as shown in
figure (1-8). The cycle is represented on Pressure Volume P-V and S-T
diagram as shown in figure (1-9) and (1-10) respectively.

11
Fig.1-8: Rankine cycle in steam turbine

Fig.1-9: Rankine cycle Pressure Volume P-V Fig .1-10: Rankine cycle on S-T diagram

The Rankine cycle is the ideal cycle for vapor power plants; it includes the
following four reversible processes:

b - c Water enters the pump as state 1 as saturated liquid and is compressed


isentropically to the operating pressure of the boiler.

c - d Saturated water enters the boiler and leaves it as superheated vapor at


state 3

d - a Superheated vapor expands isentropically in turbine and produces


work.

a - b High quality steam is condensed in the condenser .

11
The thermal efficiency of the cycle is determined from:

(𝐻1 − 𝐻2 )
⁄(𝐻 − 𝐻 ) (1-12)
1 𝑊2

Where,

Hl = Total heat of steam at entry pressure

H2 = Total heat of steam at condenser pressure (exhaust pressure)

Hw2 = Total heat of water at exhaust pressure

1.9.3 Carnot Cycle


This cycle is of great value to heat power theory although it has not been
possible to construct a practical plant on this cycle. It has high
thermodynamics efficiency.

It is a standard of comparison for all other cycles. The thermal efficiency


(η) of Carnot cycle is as follows:

(𝑇1 −𝑇2 )
𝜂= (𝑇1 )
(1-13)

Where,

T1 = Temperature of heat source

T2 = Temperature of receiver

12
1.10 Thermodynamic lows

There are four laws of thermodynamics that define fundamental physical


quantities (temperature, energy, and entropy) and that characterize
thermodynamic systems at thermal equilibrium. These are considered as
one of the most important laws in all of physics. The laws are as follows:

1.10.1 Zeroth law of thermodynamics

This law provides a definition and method of defining temperatures,


perhaps the most important intensive property of a system when dealing
with thermal energy conversion problems. Temperature is a property of a
system that determines whether the system will be in thermal equilibrium
with other systems.

Consider three thermodynamic systems, A, B, and C ,as shown in figure


(1-11) . If system A is in thermal equilibrium with (i.e., is the same
temperature as) system C and system B is in thermal equilibrium with
system C, then system A is in thermal equilibrium with system B.

Fig.1-11: Zeroth law of thermodynamics diagram

“If two systems are both in thermal equilibrium with a third then they
are in thermal equilibrium with each other“.

This statement is known as the zeroth law of thermodynamics.

13
1.10.2 First law of thermodynamics

It is observed that when a system is made to undergo a complete cycle then


network is done on or by the system. Consider a cycle in which net work
is done by the system. Since energy cannot be created, this mechanical
energy must have been supplied from some source of energy. Now the
system has been returned to its initial state: Therefore, its intrinsic energy
is unchanged, and hence the mechanical energy has not been provided by
the system itself. The only other energy involved in the cycle is the heat
which was supplied and rejected in various processes. Hence, by the law
of conservation of energy, the net work done by the system is equal to the
net heat supplied to the system.

The First Law of Thermodynamics can, therefore, be stated as follows:

“When a system undergoes a thermodynamic cycle then the net heat


supplied to the system from the surroundings is equal to net work done
by the system on its surroundingsˮ.

The boundary and the interactions that are present at the boundary play
roles in the analysis devoted to solving a problem. One feature is whether
the boundary is crossed by the flow of mass. A system defined by a
boundary impermeable to mass flow is a closed system . Conversely,
systems whose defining boundaries can be crossed by the flow of mass are
open systems, or flow systems.

The thermodynamics of open systems often relies on a special terminology;


for example, the thermodynamic system itself is referred to as the control
volume, the system boundary is the control surface, and the particular
patches of the boundary that are crossed by mass flow are the inlet and
outlet ports.

14
1.10.2`.1 First law of thermodynamics for closed system (Control Mass):

Consider the closed system shown schematically in figure (1-12): If this


system experiences a change of state from the initial state (1) to the final
state (2), the first law of thermodynamics requires that:

𝑄1−2 − 𝑊1−2 = 𝐸1 − 𝐸2 (1-14)

Heat Work Energy


transfer transfer change
(Property)

Energy interactions
(Non properties)

Fig.1-12: Graphic statements of the first law of thermodynamics for closed systems.

The difference between the net heat input Q1–2 and the net work output
W1–2 represents the change in the thermodynamic property called energy.
The first law proclaims the existence of energy as a thermodynamic
property. It defines the property “energy change.”

The energy change E2−E1 depends only on the end states, whereas the
energy interactions Q1–2 and W1–2 depend on the end states and the path of

15
the process that links the end states. This important distinction is stressed
under each term appearing in equation (1-14). Another way to stress this
difference is to use a different notation for the infinitesimal increments in
work and heat transfer relative to the exact differential notation that applies
to the infinitesimal change in E.

For this reason, the first law for a process between two states situated
infinitely close to one another is written as:

𝛿𝑄 − 𝛿𝑊 = 𝑑𝐸 (1-15)

In the same notation, the energy interactions on the left side of equation
(1-14) are:
2 2
𝑄1−2 = ∫1 𝛿𝑄 and 𝑊1−2 = ∫1 𝛿𝑊 (1-16)

The peculiar notation 𝛿 may not be the ideal way to emphasize the
difference between energy interactions and energy change. The alternative
is to introduce the concept of time in the description of the process, see the
figure (1-12) .

In this new description, state 1 is the condition of the system at time t1,
state 2 is the condition at time t2 , and the net energy interactions Ql–2 and
W1–2 are the time integrals :
𝑡 𝑡
𝑄1−2 = ∫𝑡 2 𝑄̇ 𝑑𝑡 , 𝑊1−2 = ∫𝑡 2 𝑊̇ 𝑑𝑡 (1-17)
1 1

Quantities 𝑄̇ and 𝑊̇ are the instantaneous heat transfer rate and the
mechanical power output, respectively. The first law of thermodynamics
for a closed system can be written on a per-unit-time basis as:

𝑑𝐸
𝑄̇ − 𝑤̇ = (1-18)
𝑑𝑡

16
Another way of stressing the path dependence of Q1–2 and W1–2 is
presented graphically in figure (1-13). The system can proceed from state
1 to state 2 along an infinity of paths.

Fig .1-13: Path dependence of the energy interactions Q1–2 and W1–2

Although the difference Q1–2 , W1–2 matches E2−E1 along both paths, the
sizes of Q1–2 and W1–2 vary from one path to the next. If the process
executed by the closed system is a cycle, the first law (1-14) reduces to:

∮ 𝛿𝑄 − ∮ 𝛿𝑊 = 0 (1-19)

1.10.2.2 First law of thermodynamics for open system (Control Volume):

The major difference between a Control Mass and Control Volume is that
mass crosses the system boundary of a control volume.

Figure (1-14) shows the main features of an open system: heat interactions
per unit time, 𝑄̇; work interactions per unit time 𝑊̇ ; and portions of the
boundary that are crossed by the flow of mass. For simplicity, the figure
shows only one of each type of boundary crossing, one inlet port labeled
“ in,” and one outlet port labeled “out.”

17
The open system, or the control volume, is the region contained between
the inlet and outlet ports; in other words, the stationary dashed lines labeled
in and out belong to the boundary of the open system. The work transfer
ratė 𝑊̇ refers to any mode or combination of work modes, P (dV/dt) ,̇ 𝑊𝑠ℎ
̇ ;
̇
𝑊𝑒𝑙𝑒𝑐𝑡𝑟𝑖𝑐𝑎𝑙 ̇
; 𝑊𝑚𝑎𝑔𝑛𝑒𝑡𝑖𝑐 , and so on .

Fig .1-14: Flow of a closed system (shaded area) through the space occupied by an
open system and the conversion of the first law for closed systems into a statement
valid for open systems.

Because the first-law statement (1-14) applies strictly to closed systems,


we seek a system with a fixed mass inventory that is related to the open
system of interest. If Mopen is the mass inventory of the open system at a
certain point in time t, we can think of the fixed mass inventory Mclosed that
flows at time t through the control volume. According to Fig. 1.7, the
relationship between Mopen and Mclosed is:

𝑀𝑐𝑙𝑜𝑠𝑒𝑑 𝐶𝑜𝑛𝑠𝑡𝑎𝑛𝑡 = 𝑀𝑜𝑝𝑒𝑛,𝑡 + ∆𝑀𝑖𝑛 = 𝑀𝑜𝑝𝑒𝑛,(𝑡+∆𝑡) + ∆𝑀𝑜𝑢𝑡 (1-20)

For the process from state 1 (time t) to state 2 (time t + Δt) executed by the

closed system, the first law of thermodynamics (1-14) reads :

𝐸𝑐𝑙𝑜𝑠𝑒𝑑,(𝑡+∆𝑡) − 𝐸𝑐𝑙𝑜𝑠𝑒𝑑 ,𝑡 = 𝑄̇ ∆𝑡 − 𝑊̇ ∆𝑡 + (𝑃 ∆𝑉)𝑖𝑛 − (𝑃 ∆𝑉)𝑜𝑢𝑡 (1-21)

18
The last two terms on the right side account for the P dV type of work
transfer associated with the deformation of the closed system from time t
to time t+Δt.

Note that in each term P is the local pressure: that is, the pressure in the
immediate vicinity of the port. Relations similar to eq. (1-20) express the
relative size of the energy inventories of the closed and open systems:

𝐸𝑐𝑙𝑜𝑠𝑒𝑑,𝑡 = 𝐸𝑜𝑝𝑒𝑛,𝑡 + ∆𝐸𝑖𝑛 (1-22)

𝐸𝑐𝑙𝑜𝑠𝑒𝑑,𝑡+∆𝑡 = 𝐸𝑜𝑝𝑒𝑛 ,(𝑡+∆𝑡) + ∆𝐸0𝑢𝑡 (1-23)

The ΔE’s and ΔV’s can be rewritten in terms of their per-unit mass as e
and v:

(∆𝐸)𝑖𝑛,𝑜𝑢𝑡 = (𝑒 ∆𝑀)𝑖𝑛.𝑜𝑢𝑡 and (∆𝑉)𝑖𝑛,𝑜𝑢𝑡 = (𝑣 ∆𝑀)𝑖𝑛.𝑜𝑢𝑡 (1-24)

Like the port pressure P, the specific energy and volume (e and v,
respectively) are properties of the intensive state of the fluid that crosses
the boundary at time t.

Combining equations (1-21) and (1-23) for the purpose of eliminating the
terms that refer to the energy inventory of the closed system (Eclosed), we
obtain:
1 ∆𝑀 ∆𝑀
(𝐸𝑜𝑝𝑒𝑛,(𝑡+∆𝑡) − 𝐸𝑜𝑝𝑒𝑛 ,𝑡 = 𝑄̇ − 𝑊̇ + [(𝑒 + 𝑃𝑣 ) ] − [(𝑒 + 𝑃𝑣 ) ]
∆𝑡 ∆𝑡 𝑖𝑛 ∆𝑡 𝑜𝑢𝑡

(1-25)

Invoking the limit Δt →0, writing 𝑚̇ for the mass flow rate ΔM∕Δt ,
dropping the subscript “open” from the energy inventory of the control
volume, and assuming that more than one inlet port and outlet port exist,
we arrive at the most general statement of the first law of thermodynamics
for an open system :
𝑑𝐸
= 𝑄̇ − 𝑊̇ + ∑𝑖𝑛 𝑚̇ (𝑒 + 𝑃𝑣) − ∑𝑜𝑢𝑡 𝑚̇(𝑒 + 𝑃𝑣) (1-26)
𝑑𝑡

19
What makes this statement more general than the per-unit-time first law
for closed systems equation (1-18) are the terms 𝑚̇(𝑒 + 𝑃𝑣) : These terms
represent the energy transfer associated with the flow of mass across the
system boundary.

Finally, in the absence of macroscopic forms of energy storage other than


kinetic and gravitational, the specific energy e can be decomposed into
1
(𝑢 + 𝑉 2 + 𝑔𝑧). The result of this decomposition is that the specific
2

enthalpy:

ℎ = 𝑢 + 𝑃𝑣 (1-27)

Shows up explicitly in the terms accounting for energy transfer via mass
flow:

𝑑𝐸 1 1
= 𝑄̇ − 𝑊̇ + ∑𝑖𝑛 𝑚̇ (𝑢 + 𝑉 2 + 𝑔𝑧) − ∑𝑜𝑢𝑡 𝑚̇(𝑢 + 𝑉 2 + 𝑔𝑧) (1-28)
𝑑𝑡 2 2

In the fields of gas dynamics and compressible fluid mechanics, the group
1
(𝑢 + 𝑉 2 + 𝑔𝑧) is recognized as the local stagnation enthalpy of the
2

flowing fluid.

The Δt →0 limit can be invoked in connection with the second of equation


(1-20) to yield the mass conservation statement:
𝑑𝑀
= ∑𝑖𝑛 𝑚̇ − ∑𝑜𝑢𝑡 𝑚̇ (1-29)
𝑑𝑡

This equation spells out the difference between open systems and closed
systems (in the latter, the 𝑚̇ values are all zero and the mass inventory M
is a constant).

21
1.10.3 Second law of thermodynamics

The second law of thermodynamics is a general principle, that goes beyond


the limitations imposed by the first law of thermodynamics. The first law
is used to relate and to evaluate the various energies involved in a process.
However, no information about the direction of the process can be obtained
by the application of the first law. The second law of thermodynamics
places constraints upon the direction of heat transfer and sets an upper limit
to the efficiency of conversion of heat to work in heat engines. So the
second law is directly relevant for many important practical problems.

One of the areas of application of the second law of thermodynamics is the


study of energy-conversion systems. For example, it is not possible to
convert all the energy obtained from a coal in coal-fired power plant or
from a nuclear reactor in a nuclear power plant into electrical energy. There
must be losses in the conversion process.

The second law of thermodynamics may be expressed in many specific


ways. Each statement expresses the same law. Listed below are three that
are often encountered.

- Clausius Statement: “It is impossible to construct a device which


operates on a cycle and whose sole effect is the transfer of heat from a
cooler body to a hotter body”.

- Kelvin-Planck Statement: “It is impossible to construct a device


which operates on a cycle and produces no other effect than the
production of work and the transfer of heat from a single body”.

21
1.10.3.1 Second law of thermodynamics for closed system

The essence of the second law is that in the Rumford and Joule experiments
the apparatuses received work and rejected heat. It was never the other way
around; in fact, all the attempts to construct a heat engine that would
operate cyclically as a closed system while in possible contact with a single
temperature reservoir have failed.

By recalling the heat engine sign convention of figure (1-12), the


impossibility described by Kelvin and Planck reduces to:

∮ 𝛿𝑊 ≤ 𝑜 (1-30)

or, in view of the first law for a cycle executed by a closed system equation
(1-19) :

∮ 𝛿𝑄 ≤ 𝑜 (1-31)

As we proceed with the generalization of these statements to cover systems


and processes of increasing complexity, it is important to keep in mind that
equations (1-30) and (1-31) apply only to a closed system that executes an
integral number of cycles while in communication with a single
temperature reservoir.

The next step in the direction of more complex cycles is made using the
concept of are eversible cycle, introduced by Sadi Carnot in 1824. This
concept became known as the Carnot cycle.

The modern graphical interpretation of the Carnot cycle is shown in figure


(1-15) either as an ideal gas working in the P−v plane, which is one of the
drawings made by Clapeyron (the other was done in the T−v plane), or in
terms of an unspecified working fluid on the T−s diagram .

22
Fig.1-15: Closed system executing a reversible cycle while in communication with two
temperature reservoirs T1 and T2. No assumption is made regarding the relative
magnitude of temperaturesT1 andT2 and the sense of the cycle on the P−v and T−s
planes.

It is important that each state visited by the system during the cycle is one
of uniform pressure, temperature, and specific volume. This means that the
system expands and contracts at such a slow rate that at each point in time
the state of the system is one of equilibrium.
An isothermal volume change is “slow enough” when the system is in
thermal equilibrium with the temperature reservoir. Under such conditions,
the system can execute the same cycle in the reverse sense; that is, it can
pass through the same sequence of equilibrium states in reverse order.
Therefore, if WC is the net work produced by the Carnot cycle and Q1C and
Q2 Care the respective heat interactions with reservoirs T1 and T2 during
the same cycle, the energy interactions experienced by the reversed Carnot
cycle are:
( 𝑊𝐶 , 𝑄1𝐶 , 𝑄2𝐶 )𝑟𝑒𝑣𝑒𝑟𝑠𝑒𝑑 = − ( 𝑊𝐶 , 𝑄1𝐶 , 𝑄2𝐶 ) (1-32)

To extend the second law represented by eq. (1-30) to the cyclical operation
of closed systems in contact with two temperature reservoirs. Figure (1-16)
shows a closed system (A) executing an unspecified cycle while in contact

23
with reservoirs T1 and T2, with T1≠T2. The net energy interactions
experienced by this unspecified cycle, Q1,and Q2, are related through the
first law:

𝑄1 + 𝑄2 = 𝑊 (1-33)

Fig 1-16 closed system that executes a cycle while in communication with two
temperature reservoirs

The Kelvin–Planck version of the second law places a constraint on the


signs of the two heat transfer interactions (Q1, Q2). For clarity, assume that
W is positive:

𝑊>0 (1-34)

Returning to the most general cycle executed in contact with two


temperature reservoirs figure(1-16), now we know that regardless of the
actual sign of T2−T1 the heat interactions(Q1, Q2)cannot have the same
sign ; in other words, let us designate that subscripts 1 and 2 identify the
signs of the heat interaction

𝑄1 > 0 and 𝑄2 < 0 (1-35)

24
Consider now the arrangement in figure (1-17), the unspecified cycle (A)
and the Carnot cycle (C) of figure (1-16) share reservoirs T1 and T2.

𝑄1 + 𝑄1𝐶 = 0 (1-36)

Which means that the reservoir T1 also executes a cycle. Because the
composite system ( A+C+T1 ) completes a cycle while making contact
with only one reservoir (T2), the second law (1-31) requires :

𝑄2 + 𝑄2𝐶 ≤ 0 (1-37)

Dividing this inequality through Q1 or − Q1C, which are both positive


according to equations. (1-35) and (1-36), we obtain an inequality between
two positive ratios:

−𝑄2 𝑄2𝐶
≥ (1-38)
𝑄1 𝑄1𝐶

At this stage, equation (1-38) allows us to conclude that the ratio obtained
by dividing the absolute value of the negative heat inter-action by the value
of the positive heat interaction cannot be smaller than a certain value. The
limiting case corresponds to the equal sign in equation (1-38):

−𝑄2 𝑄2𝐶
= (1-39)
𝑄1 𝑄1𝐶

which in combination with equation (1-36) and the first-law statements


W=Q1+Q2 and WC=Q1C+Q2C translates into :

( 𝑊𝐶 , 𝑄1𝐶 , 𝑄2𝐶 ) = − ( 𝑊 , 𝑄1 , 𝑄2 ) (1-40)

The equal sign in the second law (1-39) is associated with any reversible
cycle that is executed by the unspecified system (A). Therefore, the second
law can be stated by making reference only to the system of interest (A):

−𝑄2 −𝑄2
≥( ) (1-41)
𝑄1 𝑄1 𝑟𝑒𝑣

25
Where the subscript rev stands for “reversible”: in this case, for a reversible
cycle executed by an arbitrary closed system in contact with two
temperature reservoirs.

The limiting value (−Q2 ∕ Q1)rev can only be a function of parameters T1 and
T2, because the existence of two different temperature reservoirs was
assumed in the derivation of equation (1-41), Writing:

−𝑄2
( ) = 𝑓(𝑇1 , 𝑇2 ) (1-42)
𝑄1 𝑟𝑒𝑣

The unknown function f (T1,T2) has a special property that becomes visible
while invoking the definition (1-42) for two additional closed systems that
execute cycles while in contact with pairs of temperature reservoirs as
figure (1-17) .

Fig.1-17: Two closed systems that execute cycles while in contact with pairs of
temperature reservoirs
−𝑄3
( ) = 𝑓(𝑇1 , 𝑇3 ) (1-43)
𝑄1 𝑟𝑒𝑣
−𝑄3
( ) = 𝑓(𝑇2 , 𝑇3 ) (1-44)
𝑄2 𝑟𝑒𝑣

26
Dividing equation (1-43) by equation (1-44), along with using equation
(1-42) to eliminate (−Q2∕Q1)rev , yields :
𝑓(𝑇1 ,𝑇3 )
𝑓(𝑇1 , 𝑇2 ) = (1-45)
𝑓(𝑇2 ,𝑇3 )

The left side of eq. (1-45) does not depend on the constant T3 ; therefore,
the analytical form of f (T1,T2) is :

𝜓 (𝑇1 )
𝑓(𝑇1 , 𝑇2 ) =
𝜓(𝑇2 )
(1-46)

or, letting 𝜙=1∕𝜓 :

−𝑄3 𝜙𝑇2
( ) = (1-47)
𝑄1 𝑟𝑒𝑣 𝜙𝑇1

Equation (1-47) can be rewritten to express the ratio between the heat
interactions of a reversible cycle that absorbs one unit of energy (Q0) from
a reference reservoir of empirical temperature 𝜃0 and rejects the amount
–Q to an arbitrary reservoir (𝜃):
−𝑄
𝜙(𝜃) = 𝜙(𝜃0 ) ( ) (1-48)
𝑄0 𝑟𝑒𝑣

The numerical values obtained in this manner for function 𝜙 constitute the
thermodynamic temperature scale (symbol T); in other words, ≡ T or :
−𝑄
𝑇 = 𝑇0 ( ) (1-49)
𝑄0

One of the properties of the ideal gas Carnot cycle sketched in the P−v
plane is :
−𝑄2𝐶 𝜃2,𝐼𝑑𝑒𝑎𝑙 𝑔𝑎𝑠
= (1-50)
𝑄1𝐶 𝜃1,𝐼𝑑𝑒𝑎𝑙 𝑔𝑎𝑠

Where 𝜃ideal gas is the ideal gas temperature defined by:


𝑃𝑣
𝜃𝑖𝑑𝑒𝑎𝑙 𝑔𝑎𝑠 = (1-51)
𝑅

27
Comparing equation (1-47) with equation (1-50), we see that 𝜙 = 𝜃ideal gas :
𝜃𝑖𝑑𝑒𝑎𝑙 𝑔𝑎𝑠 = 𝑇 (1-52)
The reward for introducing the concept of a thermodynamic temperature
scale is that (−Q2∕Q1)rev =T2∕T1 and the second law (1-41) becomes :
𝑄1 𝑄2
+ ≤0 (1-53)
𝑇1 𝑇2

This new statement is general despite the assumption W > 0 made early in
the derivation of equation (1-38).

We see a sequence that leads to the second law for any cycle executed in
contact with any number of temperature reservoirs:
𝑄1 𝑄2 𝑄𝑛
+ + ⋯+ ≤0 (1-54)
𝑇1 𝑇2 𝑇𝑛

Assume that the statement corresponding to n heat reservoirs equation


(1-54) is correct. If, based on this last assumption, we can show that the
statement for n+1 heat reservoirs is correct:

𝑄𝑖
∑𝑛+1
𝑖=1 ≤0 (1-55)
𝑇𝑖

Then the assumed equation (1-55) is valid. Consider, for this purpose, a
system A that executes a cycle while in thermal communication withn+1
temperature reservoirs ,T1,T2,...,Tn,Tn+1 . No assumption is made about
the direction (sign) of each of the heat transfer interactions,
Q1,Q2, . .., Qn,Qn+1.Letus return reservoirTn+1 to its original state by
placing it in contact with a reversible cycle (C) such that :

𝑄𝑛+1 + 𝑄𝑛+1,𝐶 = 0 (1-56)

28
The second law for the composite system A+Tn+1+C, which
completes a cycle while in contact with temperature reservoirs,
we obtain :
𝑄𝑖 𝑄𝑛,𝐶
∑𝑛𝑖=1 + ≤0 (1-57)
𝑇𝑖 𝑇𝑛

Noting that cycle C is reversible (Qn,C ∕ Tn + Q (n+1),C∕Tn+1 = 0) and using


equation (1-56), we write in order :

𝑄𝑖 𝑄(𝑛+1),𝐶
∑𝑛𝑖=1 − ≤0 (1-58)
𝑇𝑖 𝑇(𝑛+1)

𝑄𝑖 𝑄(𝑛+1),𝐶
∑𝑛𝑖=1 + ≤0 (1-59)
𝑇𝑖 𝑇(𝑛+1)

The final step in this line of generalization is to consider a continuous


variation of system boundary temperature as the cycle is executed in
contact with an infinite sequence of temperature reservoirs, each
contributing a heat interaction of size 𝛿Q or ̇Q dt. For such cases, equation
(1-54) is written as :

𝛿𝑄
∮ ≤0 (1-60)
𝑇

With the understanding that T represents the Kelvin or Rankine


thermodynamic temperature of that portion of the system boundary that is
instantly crossed by the heat transfer 𝛿Q. The equal sign of equation (1-61)
refers to reversible cycles:

𝛿𝑄𝑟𝑒𝑣
∮ ≤0 (1-61)
𝑇

The concept of entropy as a thermodynamic property follows directly from


equation (1-62): If the net change in 𝛿Qrev ∕ T is zero at the end of a

29
reversible cycle, 𝛿Qrev ∕ T represents the change in a thermodynamic
property S:

δQrev
dS = (1-62)
T

This new property was named entropy by Clausius in 1865; however, the
same property was discovered and used earlier by Rankine , who called it
thermodynamic function, labeled it 𝜙, instead of S, and regarded equation
(1-62) as the “general equation of thermodynamics.”

Thinking of only that family of paths that are reversible (a path is reversible
if it can be a part of a reversible cycle), we integrate equation (1-62) and
obtain the definition of entropy change:

2 𝛿𝑄𝑟𝑒𝑣
𝑆1 − 𝑆2 = ∫1 (1-63)
𝑇

The arbitrary process 1→2 can be a part of a cycle 1→2→1, where the
return process 2→1 takes place along a reversible path; rewriting
equation(1-60) for this cycle and using the definition (1-63), we have, in
order,

2 𝛿𝑄 1 𝛿𝑄𝑟𝑒𝑣
∫1 + ∫2 ≤0 (1-64)
𝑇 𝑇

2 𝛿𝑄
∫1 𝑇
≤ 𝑆2 − 𝑆 1 (1-65)

Entropy Entropy
transfer change
(Nonproperty) (Property)

The second law of thermodynamics for a process equation (1-65) states


that, algebraically, the entropy transfer never exceeds the entropy change.
A measure of the strength of the inequality sign in equation (1-65) ,

31
(i.e., a new definition) is the entropy generation or entropy production Sgen
, a quantity that is never negative:
2 𝛿𝑄
𝑆𝑔𝑒𝑛 = 𝑆2 − 𝑆1 − ∫1 ≥0 (1-66)
𝑇

Like the entropy transfer, the entropy generation is path dependent.

1.10.3.2 Second law of thermodynamics for open system

The final step in the generalization of the second-law statement is based on


the closed system–open system transformation shown in Figure (1-14)
With reference to this drawing, we write:

𝑆𝑐𝑙𝑜𝑠𝑒𝑑,𝑡 = 𝑆𝑜𝑝𝑒𝑛,𝑡 + ∆𝑆𝑖𝑛 (1-67)

𝑆𝑐𝑙𝑜𝑠𝑒𝑑,(𝑡+∆𝑡) = 𝑆𝑜𝑝𝑒𝑛,(𝑡+∆𝑡) + ∆𝑆𝑜𝑢𝑡 (1-68)

and

(∆𝑆)𝑖𝑛,𝑜𝑢𝑡 = (𝑠∆𝑀)𝑖𝑛,𝑜𝑢𝑡 = (𝑠𝑚̇)𝑖𝑛,𝑜𝑢𝑡 ∆𝑡 (1-69)

Labeling the entropy generated from time t to t+Δt as ΔSgen , equation


(1- 66) yields :

𝑄̇𝑖
∆𝑆𝑔𝑒𝑛 = 𝑆𝑜𝑝𝑒𝑛 ,(𝑡+∆𝑡) − 𝑆𝑜𝑝𝑒𝑛 ,𝑡 − ∆𝑡 + (𝑚𝑠
̇ )𝑜𝑢𝑡 ∆𝑡 − (𝑚𝑠
̇ )𝑖𝑛 ∆𝑡 ≥ 𝑜
𝑇𝑖

(1-70)

Invoking the limit Δt →0, dropping the subscript “open,” and assuming the
existence of any number of spots with heat transfer i and mass flow (in,
out) on the control surface, we obtain :

𝑑𝑠 𝑄̇𝑖
̇
𝑆𝑔𝑒𝑛 = − ∑𝑖 + ∑𝑜𝑢𝑡 𝑚𝑠 ̇ 0
̇ − ∑𝑖𝑛 𝑚𝑠 ≥ (1-71)
𝑑𝑡 𝑇𝑖

Entropy Rate of Entropy Net entropy


generation entropy transfer flow rate out
rate accumulation rate of the control volume
inside the
control volume

31
1.10.4 Third law of thermodynamics

Figure (1- 18) shows the shape of the calculated S (T , ℋ) curves for a
paramagnetic salt such as gadolinium sulfate or gadolinium gallium garnet
at temperatures below 1 K . An extensive comparison of the various
properties of paramagnetic materials contemplated for magnetic
refrigeration systems has been provided by Barclay and Steyert.

Fig.1-18: Constant ℋ lines on the (T–S) plane of a paramagnetic salt and the process

of adiabatic demagnetization.

The geometry of the two constant-ℋ lines shown in figure (1-18) is an


excellent opportunity for reviewing the physical evidence that supports the
third law of thermodynamics. We can theorize that processes of type
0→1→2 can be repeated in the direction of lower temperatures, as shown
by the steps nested between the ℋ=0 and ℋ=ℋ1 lines in figure (1-19).

The assumption made here is that during each subsequent isothermal


magnetization process (e.g., 2→1′) the heat rejected by the “working”
paramagnetic salt is absorbed by a sufficiently large thermal mass of
temperature T2 ,the latter being the result of the first magnetic refrigeration
stage 0→1→2.

32
Fig.1-19: Unattainability of absolute zero by means of a finite sequence of isothermal
magnetization and adiabatic demagnetization processes of the type sown in Fig.1-18

From the construction of figure (1-19) is that the temperature T=0 cannot
be reached through a finite sequence of operations. Because it stems from
in the fact that the constant-ℋ lines intersect at T=0. The same geometric
feature can be expressed as:

lim (∆𝑆) 𝑇 = 0 (1-72)


𝑇→0

Where (ΔS)T is shorthand for the magnitude of the isothermal entropy


change experienced by the substance: for example, (ΔS) T=|S0−S1|.

The statement that summarizes the observation made above is the


following:

It is impossible by any procedure, no matter how idealized, to reduce


any system to the absolute zero in a finite number of operations.

33

You might also like