Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
0% found this document useful (0 votes)
71 views74 pages

Lecture MMC301 Up To 05-09-2023 PDF

Download as pdf or txt
Download as pdf or txt
Download as pdf or txt
You are on page 1/ 74

Metallurgical Thermodynamics and

Kinetics (MMC301)

Dr. Manas Kumar Mondal


Associate Professor,
Department of Metallurgical and Materials Engineering
National Institute of Technology, Durgapur
FIRST LAW OF THERMODYNAMICS
Work
Work(w) can be described as an interaction between a system and its surrounding. The interaction involves
exchange of energy .
Mechanical, Surface tension, Electrical, Magnetic work etc., but we are concerned only with Mechanical work.
Work of expansion: Consider the expansion of hot gases in the cylinder of an automobile engine. The hot gases push
the moving piston outward. Thus, work is done by the gases on the surroundings. This type of
work is known as the work of expansion.
Work of compression: Another example deals with the compression of air by means of an air compressor, or this air
is forced into a given volume with the help of the compressor. In this example, work is done
by the surroundings on the system. This type of work is known as the work of compression.
If system has done work on the surroundings then w is positive, whereas work is done on the system, the W is
negative
Let us, suppose that a gas enclosed in a cylinder (Figure 1) filled with a smoothly
moving frictionless piston is the system under consideration. The system is at a
pressure (P), volume (V) and temperature (T) and there are n moles of gas.
Assumptions:
• At every instant the applied or external pressure Pe is very nearly equal to the
pressure of gas.
• During a reversible process, the pressure, temperature and density of the gas
are uniform throughout the cylinder. Figure 1: Schematic representation
of reversible compression of gas
The mechanical work done (W) = force × displacement = APe (x) =Pe dV
dV = Adx is the volume change during the small displacement.
For a reversible process, Pe  P and W = P dV ……….. (1)

The molar work done (W) = P dV ……….. (1)

Where V is the molar volume of the gas at the particular instant indicated in Figure 1.

Suppose that the reversible compression process was carried out from an
initial thermodynamic state designated 1 to a final state designated 2.The
path of the process is depicted by a curve on the P-V diagram (Figure 2)
that joins the end points 1 and 2.

The small shaded strip under the curve represents the work done W = P dV.
𝑽
For the entire process stretching from 1 to 2, we have 𝑾 = ‫𝑽𝑑 𝑃 𝟐 𝑽׬‬..........(3).
𝟏

W is equal to the total area under the curve 1-2.


A system can change from one state to another in many different
ways. In other words, there are several different paths that the
system can take changing state 1 to state 2, two of which are shown
in figure 3. These two paths are labelled 1 A 2 and 1 B 2.

Path 1 A 2 consists of two distinct parts:


During the first part the system is compressed at constant pressure until
the volume become equal to VA (= V2);
During the second part the gas is heated at constant volume its the
pressure and temperature rises to P2 and T2 , respectively.

Path 1 B 2 also consists of two parts:


In the first part gas is heated at constant volume unit the pressure until the pressure rises to PB (= P2).
In the second part the gas is compressed at constant pressure until the volume becomes equal to V2.
Heat
The amount of heat, Q, involved in a process is an important thermodynamic quantity.

Heat flows from the system to its surroundings, or vice versa, by virtue of a temperature difference, however
small, between the system and the surroundings.

The convention is that the heat is positive when it is absorbed by the system from the surroundings and negative
when heat is lost by the system to do the surroundings.

In adiabatic processes, no heat is exchanged between the system and its surroundings. In other words, Q = 0 for an
adiabatic process.

W is dependent on the path taken by the system between two given thermodynamic States.

Similarly, Q is not a state function either, that is ,it can not be expressed as a function of the thermodynamic
coordinates (P, T, etc).
FIRST LAW OF THERMODYNAMICS
The statement of the law of conservation of energy.

E = Q  W ……………. (1)

Q and W are not state functions, but their difference Q  W = E is a state function.
In other words, E can be expressed as a function of thermodynamic variable such as P,T and so forth.

In this particular cycle, steps 1 and 3 are constant volume steps and 2 and 4 are constant pressure steps.

On applying the first low of each of these operations,


(E)1 = Q1  W1, (E)2 = Q2  W2, (E)3 = Q3  W3 and (E)4 = Q4  W4

For the entire cycle, (E)cycle = Qi  Wi ……. (2)

Where Qi = Q1 + Q2 + Q3 + Q4 and Wi = W1 + W2 + W3 + W4 (E)cycle either  or = or  0

Therefore, (E)cycle = 0

The first law applies whether the process is reversible or not. For an infinitesimal process, the
first law statement can be expressed as follows: dE = Q  W
REVERSIBLE AND IRREVERSIBLE PROCESSES
Reversible process:
A process is said to be reversible if it is performed in such a way that at the completion of the process, both the system and its
surroundings can be restored to their respective initial states. Because the system and the surroundings together constitute the
universe, it is clear that a reversible process is one that is performed in such a way that the universe is not able altered.
Reversible isothermal process:
The reversible isothermal process that is carried out reversibly keeping the temperature constant.
Reversible adiabatic process:
A reversible adiabatic process is carried out reversibly with the system thermally insulated from the surroundings.
The adiabatic free expansion of an ideal gas into an evacuated vessel in a typical example of an irreversible process.
For work to be done, the gas should expand against an opposing force
that is very nearly equal to the force exerted by the gas on the face of the
piston. In this example, the external force is either nil or substantially
less than the force exerted by the gas. The evacuated compartment offers
no resistance.
W = 0 and W = 0, and Q = 0 (as the process is adiabatic)
Q = E + W = 0 or E =0
Examples of irreversible process are i. Spontaneous flow of heat from a
system to cooler surroundings, ii. spontaneous chemical reaction, iii.
Spontaneous solidification of a super cool liquid metal and iv. Mixing of
substances to form a homogeneous solution .
HEAT AND WORK CHANGES IN REVERSIBLE PROCESSES:
ISOTHERMAL PROCESS:

From the ideal gas law, PV = RT0 = k0, where k0 is constant.

Similarly, PV = RT = k, where k is constant. For the example T > T0 and k > k0


Consider the P-V curve at T0 shown in figure 6. For an infinitesimally small change,
the work done is given by W = P dV
For an ideal gas, P = RT0/V. Therefore, W = RT0 d(lnV)
𝑽 𝑽
W = ‫ 𝑽׬‬2 𝑅𝑇0 𝑑 𝑙𝑛 𝑽 = 𝑅𝑇0 𝑙𝑛(𝑽𝟐 )
1 𝟏

This is for 1 mole of gas. If the system consists of n moles of ideal gas, then W = nRT0 ln(V2/V1) = n RT0 ln(P1/P2)
Q = W = RT0 ln(V2/V1), As E = 0 for an ideal gas.

CONSTANT VOLUME PROCESS:


Both the pressure and temperature of the system are varied, whereas its volume is held
constant an P-V diagram (figure 7). This process represented by a vertical line parallel to
the P axis.
W=0
From the first law of thermodynamics Q = E
ISOBARIC or CONSTANT PRESSURE PROCESS
In this process the pressure of the system is maintained constant. Its temperature and volume
are try to very.
On a P-V diagram (Figure 8) the isobaric process is simply a horizontal line

If the system is an ideal gas ,then PV = RT for 1 mole of gas or V = (R/P)T

At constant pressure, V is directly proportional to T.


For the small strip shown cross-hatched in figure 8. The corresponding work done is W = P dV.

2 2
W = ‫׬‬1 𝑃𝑑𝑉 = 𝑃 ‫׬‬1 𝑑𝑉 = 𝑃(𝑉2 − 𝑉1)

The interact energy change is E = E2 – E1

The heat exchanged, Qisobaric or QP , can readily be determined by applying the first law of thermodynamics:

QP = E + W = E2 – E1 + PV2 – PV1 = (E2 + PV2) – (E1 + PV1) = H2 – H1 = H

Where H, a new state property, known as the enthalpy (or heat content) is defined as follows: H = E + PV
For condensed phase, the product PV is rather small and for these substances H  E.

For an ideal gas, on the other hand PV is quite applicable.


HEAT AND WORK EFFECTS IN A CYCLIC PROCESS:
Consider 1 mole of an ideal gas enclosed in cylinder filled with a smoothly moving piston. In
thermodynamics parlance, this is our system. This system is now subjected to a series of
processes. Entire cycle is shown in the figure 9.
Initial state: P1, V1, T1 and E1 Process 12: reversible isothermal process.
Process 23: reversible isobaric process. Process 31: reversible isovolumic process.
Process 12: from P1, V1, T1, E1 to P2, V2, T1, E2

Work done, W12 = RT1 ln (V2/V1) Change in internal E12= E2 – E1


Heat exchanged, Q12= E2 – E1 + RT1 ln (V2/V1)

Process 23: from P2, V2, T1, E2 to P2, V1, T2, E3


Work done, W23 = P2(V1 - V2) = P2V1 - P2V2 Change in internal E23= E3 – E2 Heat exchanged, Q23= E3 – E2 + P2V1 - P2V2

Process 31: from P2, V1, T2, E3 to P1, V1, T1, E1


Work done, W31 = 0 Change in internal E31= E1 – E3 Heat exchanged, Q31= E1 – E3
For the entire cyclic process:
Net work done: Wcycle = RT1 ln (V2/V1) + P2V1 – P2V2 Change in internal Ecycle= 0

Net heat exchanged: Qcycle = RT1 ln (V2/V1) + P2(V1 – V2)


The net work done by the system in a cyclic process is equal to the net heat absorbed by the system.
HEAT CAPACITY
The heat capacity ‘C’ of a system can be defined as the quantity of heat required to raise the temperature of the system by 1 degree.
A system is subjected to an infinitesimal process. During this process, heat is transferred to the system, some work is done by the
system and there is some change in its internal energy Q = dE + W
The process considered is not isothermal but involves a change (dT) in temperature. We thus defined the heat capacity
C = Q/dT or C = 𝑙𝑖𝑚 (∆𝑄/∆𝑇)
∆𝑇→0
(1) heat capacity at constant pressure CP and (2) heat capacity at constant volume CV

CP = QP/dTP and CV = QV/dTV

For a constant pressure process, QP = enthalpy change, HP and QP = dHP
For a constant volume process, QV = internal energy change, EV and QV = dEV

𝐶𝑃 = 𝛿𝑄𝑃 𝑑𝑇𝑃 = 𝜕𝐻 𝜕𝑇 P
𝐶𝑉 = 𝛿𝑄𝑉 𝑑𝑇𝑉 = 𝜕𝐸 𝜕𝑇 V

Heat capacity can be expressed in calories or joules per degree kelvin or in Btu per degree Rankine. Molar heat capacity, C , is
heat capacity per mole.

The heat capacity of a substance is markedly influenced by temperature. At low temperature, CP approaches CV and at
temperature nearing absolute zero, both CP and CV tend to zero. At higher temperature, the Dulong and petit rule gives a
reasonable estimate of the heat capacity. According to this rule, the heat capacity of a solid approaches a value of 3R or 24.94
J/g.atom.K at high temperature.
LOW TEMPERATURE HEAT CAPACITIES:
The Debye theory of heat capacities provides a useful framework to analyzing low-temperature heat capacity data. Debye
equation is as follows: CV =𝐾 𝑇  3 Where K is a constant equal to 1944 J/g.mole.K and  is known as Debye temperature.
Limitation: The Debye equation neglects the electronic contribution to the heat capacity. This contribution is quite significant
for metallic substances.
The modified equation for heat capacity at low temperature is CV =𝐾 𝑇  3 + T …………… (1)
both  and  can be determined from low temperature heat capacity measurements. On dividing both sides of equation(1) by
T we have CV /T= 𝐾 𝜃 3 𝑇 2 + 𝛾
A plot of CV/T versus T2 yields a straight line with a slope equal to K/3 and the intercept at T2 = 0 gives the value of  .

HIGH TEMPERATURE HEAT CAPACITIES:


The temperature dependence of heat capacity at moderate to high temperatures is often expressed by means of an empirical
relationship. In fact several different kinds empirical relationship have been in use for representing CP -T behavior. One of the
more frequently used relationship is
CP = a + bT + cT2 J/g.mole.K.

where a, b and c are empirical constants to be determined from experimental measurement of CP.
On the other hand, Kelley prefers a slightly different relationship

CP = a + bT - cT2 J/g.mole.K.
Relationship between CP and CV for an ideal gas
The heat capacity at constant volume is 𝐶𝑉 = 𝜕𝐸 𝜕𝑇 𝑉 ………….. (1)

For an ideal gas, E is solely a function of T and independent of both P and V . Accordingly, for an ideal gas
𝑑𝐸
𝐶𝑉 = 𝑑𝑇 = 𝜕𝐸 𝜕𝑇 𝑃 = 𝜕𝐸 𝜕𝑇 𝑉 …………… (2)

The heat capacity at constant pressure is 𝐶𝑃 = 𝜕𝐻 𝜕𝑇 𝑃 ………….. (3)

As H = E + PV …… (4)
𝜕𝐻 𝜕𝑇 𝑃 = 𝜕𝐸 𝜕𝑇 𝑃 + 𝑃 𝜕𝑉 𝜕𝑇 𝑃 ………. 5

On combining equations (1), (2), (3) and (5) we obtain

𝐶𝑃 = 𝐶𝑉 + 𝑃 𝜕𝑉 𝜕𝑇 𝑃 ………. 6

𝑅
For an ideal gas, PV = RT 𝜕𝑉 𝜕𝑇 𝑃 = 𝑃 … … … . (7)

𝑅
Hence 𝐶𝑃 = 𝐶𝑉 + 𝑃 𝜕𝑉 𝜕𝑇 𝑃 = 𝐶𝑉 + 𝑃 ×  𝐶𝑃 − 𝐶𝑉 = 𝑅
𝑃
CALCULATION OF ENTHALPY CHANGES
The enthalpy change associated with change in temperature of a substances can be calculated from the heat capacity data a
follows.
𝑇
At constant pressure dH = CP dT On integration H = ‫ 𝑇׬‬2 𝐶𝑃 𝑑𝑇
2
𝑇 𝑏 1 1
On the substitution for CP, H = ‫ 𝑇׬‬2(𝑎 + 𝑏𝑇 + 𝐶𝑇 −2 )𝑑𝑇 → ∆𝐻 = 𝑎 𝑇2 − 𝑇1 + 𝑇22 − 𝑇12 − 𝑐 −𝑇
2 2 𝑇2 1

HEAT FORMATION OF SUBSTANCES:


It is impossible to measure the absolute value of a thermodynamic property such as the enthalpy of a substance. In practice
this difficulty is overcome by choosing a reference a standard state and measuring the enthalpy relative to the enthalpy of the
standard state. The heat or the enthalpy of formation of an element or compound will depend on the standard state chosen. For
convenience, the following standard states are defined.
Solids: For solid element or compounds, pure substance at 1 atm pressure and specified temperature is the standard state.
Liquids: For liquid elements or compound, pure liquid at 1 atm pressure and specified temperature is the standard state.

Gases: For gases elements and compounds, pure, ideally behaving gas at 1 atm pressure and specified temperature is the
standard state.

The standard state chosen depends primarily on temperature

The titanium metal has two allotropes: Ti () and Ti ().


1155𝐾
Ti () Ti (), absorbs 3350 J
Hess’s law and heat of reaction:
The amount of heat absorbed or liberated during a chemical reaction depends on the conditions under which the reaction is
taking place and also an the quantities of substances partaking in the reaction.
1
2CuFeS 2  6 O2  2CuO  Fe2O3  4 SO2
2
The quantity of heat liberated when 2 moles of chalcopyrite combine with 13/2 moles of oxygen to yield 2 moles cupric
oxide, 1 mole of ferric oxide and 4 moles of sulfur dioxide at constant temperature and pressure is known as the heat of
reaction at constant pressure or simply the heat of reaction.
The enthalpy is a state property. It depends an only the initial and final states.
According to Hess's law, the enthalpy change associated with a given chemical reaction is the same either it takes place in
one or several stages.
The reaction between methane and oxygen to produce carbon dioxide and water vapour can proceed by two distinctly
different paths. Direct oxidation or three stage process:
CH4 + 2O2 = 2CO2 + 2H2O, H1
Indirect oxidation or three stage process
CH4 = C + 2H2, H2
C + O2 = CO2, H3
2H2 + O2 = 2H2O, H4

Involving Hess’s law, we find that H1 = H2 + H3 + H4


Variation of heat of reaction with temperature: Kirchhoff’s law
The heat of reaction varies with temperature. Kirchhoff’s law gives the temperature
dependence of the heat of reaction in quantitative terms.
aA + bB + cC + …………. = lL + mM + nN +……… (endothermic reaction) ….(1)
0
The standard heat of reaction is ∆𝐻298 . ∆𝐻𝑇01 =?
Enthalpy is a state property. So, there will be no change in the enthalpy of the system in
a cyclic process.
Consider the cycle IJY X I in Figure 10
 H 298
0
 H Y  H J  H T01  H I  H X  0........(2)
T1 T1
For the reactants H Y  H J   aC
298
PA  bC PB  ...dT .......(3) For the products  H I  H X   lC PL  mCPM  .....dT .......(4)
298

Substitution and rearrangement of items yields


T1 T1

 H   aC  bC PB  ...dT  H   lC  mCPM  ...dT  0


0 0
298 PA T PL
298 298

T1

H T01  H 298   lC  mC PM  ...  aC PA  bC PB  ...dT ......(5)


0
PL
298

Differential form of Kirchhoff’s law 𝑑∆𝐻 𝑑𝑇 𝑃 = ∆𝐶𝑃 …… (6) C P   C P , products   C P ,reac tan ts
Problem
Find the H0 at 1200 K for the following reaction
Fe2O3 (S) + 3 CO (g) = 2 Fe(s) + 3CO2 (g)
Following data are given
Transformation Transformation Heat of transformation (J)
temperature (K)

(Fe2O3) (Fe2O3) 953 670

(Fe2O3) (Fe2O3) 1053 -


(Fe) (Fe) 1033 5105
(Fe) (Fe) 1187 670
Heat of formation and heat capacities
∆𝑯𝟎𝒇 𝒂𝒕 𝟐𝟗𝟖K (kJ) 𝑪𝑷 (J/mole.K)
Fe2O3 ()  821.3  = 98.28 + (77.28 × 103T)  14.85 × 105 𝑇2
 = 150.62
 = 132.63 + (7.36 × 103T)
Fe () 0.0  = 17.49 + (24.77 × 103T)
 = 37.66
 = 7.70 + (19.5 × 103T)
CO (g) 110.5 28.41 + (4.10 × 103T)  0.46 × 105 𝑇2
CO2 (g) 393.5 44.14 + (9.04 × 103T)  8.54 × 105 𝑇2
Solution: The heat of reaction at 298K for the reaction Fe2O3 (S) + 3 CO (g) = 2 Fe(s) + 3CO2 (g)
H 298
0
 2  0  3   393.3   821.3  3   110.5  27.7 KJ  27700 J

Fe2O3  ,298  3Cog ,298  2 Fe ,298  3Co2 g ,298 : H1


Fe2 O3  ,298  Fe2 O3  ,953 : H 2 Fe2 O3  ,953  Fe2 O3  ,953 : H 3 Fe2 O3  ,953  Fe2 O3  ,1053 : H 4

Fe2 O3  ,1053  Fe2 O3  ,1053 : H 5 Fe2 O3  ,1053  Fe2 O3  ,1200 : H 6 2Fe ,298  2Fe ,1033 : H 7

2 Fe ,1033  2Fe ,1033 : H 8 2Fe ,1033  2Fe ,1187 : H 9 2Fe ,1187  2Fe ,1187 : H10

2Fe ,1187  2Fe ,1200 : H11 3Cog,298  3Cog,1200 : H12 3Co2 g ,298  3Co2 g ,1200 : H13

For the reaction at 1200K, Fe2 O3  ,1200  3Cog ,1200  2Fe ,1200  3Co2 g ,1200

The heat of reaction 𝐻1200


0
is given by H 1200  H 1  H 2  ...  H 6   H 7  ...  H 11   H 12  H 13
0

H 1  H 298
0
 27700 J

 98.28  77.28 10 T   14.85 10 T dT  92831J


953 953

C 
3 2
H 2  P  Fe2 O3 dT  5

298 298

H 3  670 J H 4  15062J H 5  0 H 6  20716J H 7  49942J H 8  10210 J H 9  11600 J

H 10  1340 J H11  1340 J H12  84,840 J H 13  131300 J


Substitutions gives
H 1200
0
 27700  129274  73898  84840  131300  36621J
REVERSIBLE ADIABATIC PROCESS
For an infinitesimal reversible process undergone by an ideal gas, the first law is Q = dE + P dV

and heat capacity at constant volume is given by: 𝐶𝑉 = (𝜕𝐸/𝜕𝑇)𝑉 … … … . (1)

For the ideal gas, E is a function of T only. Therefore, the partial deviation of E with respect to T can be replaced by the
𝜕𝐸 𝑑𝐸
total deviation . 𝐶𝑉 = 𝜕𝑇 𝑉 = 𝑑𝑇 → 𝑑𝐸 = 𝐶𝑉 dT for ideal gas

On combining this result with the first law, we have Q  CV dT  PdV


In an adiabatic process, Q = 0 and P dV =  CV dT ……………. (2)

Furthermore, for 1 mole of substance: PV = RT  P = RT/V

Substitute into equation (2) gives RT  dV 


dV  CV dT  RT    CV dT
V  V 
 dV  dT
 R   CV .
 V  T
 R.d ln V  CV .d ln T
R

 d ln V CV
 d ln T
R

 d ln V CV
 d ln T  0
 CR 
 d ln V V .T   0
 
 
On integration: 𝑉 𝑅 𝐶𝑉 𝑇 = 𝑘 … … . . (3)
R CP R
 1     1.......(4) as CP –CV = R
CV CV CV
R

Combining equation (3) and equation (4) V CV


.T  K  TV  1  k ...........(5)
Equation (5) gives the relationship between temperature and volume for adiabatic process.
A pressure-volume relationship can be deduced as follows.
𝑃𝑉 (𝛾−1) 𝑃𝑉 (𝛾−1) 𝑃𝑉 𝛾
T= , 𝑇𝑉 = 𝑉 = 𝑅 = 𝑘 → 𝑃𝑉 𝛾 = C
𝑅 𝑅

WORK DONE IN AN ADIABATIC PROCESS FOR AN IDEAL GAS:


For an infinitesimal change, work done is given by W = P dV
𝑑𝑉
Eliminating P in favour of V, we can write 𝛿𝑊 = 𝐶 … … … … 6 [𝑎𝑠 𝑃𝑉 𝛾 = C]
𝑉𝛾

Let us suppose, that the initial and final states of the system are (P1, V1, T1) and (P2, V2, T2) respectively. An integrating
equation (6). 1−𝛾 1−𝛾
𝑊 = 𝐶 Τ 1 − 𝛾 𝑉2 − 𝑉1 … … . . (7)
𝛾
𝐴𝑔𝑎𝑖𝑛, 𝐶 = 𝑃2𝑉2𝛾 = P1𝑉1 =P𝑉 𝛾 … … … … … … … … . (8)

On combining equation (7) and (8) we obtain W = (P1V1 – P2V2)/( - 1) …………. (9)

For a ideal gas P1V1 = RT1 and P2V2 = RT2

𝑇ℎ𝑖𝑠 𝑙𝑒𝑎𝑑𝑠 𝑡𝑜 𝑊 = 𝑅Τ  − 1 𝑇1 − 𝑇2 = 𝐶𝑉 (𝑇1 − 𝑇2 )


POLYTROPIC PROCESS
Polytropic process as a general process , this process is a thermodynamic process that obey the reaction for polytrophic
process: 𝑃𝑉 𝑛 = 𝐶, 𝑤ℎ𝑒𝑟𝑒 𝑛 = 𝑝𝑜𝑙𝑦𝑡𝑟𝑜𝑝𝑖𝑐 𝑖𝑛𝑑𝑒𝑥 𝑎𝑛𝑑 𝑛

1. if n = 0 then PV 0  C  P  C  Constant pressure process or isobaric process.

2. if n =1 then PV 1  C  PV  C  Constant temperature process or isothermal process.

3. if n =  then PV   C if multiply both side power 1/


1  1
P V C
 

 V  CONSTANT  Constant volume process or isochoric process


4. if n =1.4=  𝑡ℎ𝑒𝑛 𝑃𝑉 𝛾 = 𝐶𝑜𝑛𝑐𝑡𝑎𝑛𝑡  Reversible adiabatic process or Isentropic process.
For an infinitesimal change, work done is given by W = P dV
𝑑𝑉
Eliminating P in favour of V, we can write 𝛿𝑊 = 𝐶 … … … … 1 [𝐹𝑜𝑟 𝑝𝑜𝑙𝑦𝑡𝑟𝑜𝑝𝑖𝑐 𝑝𝑟𝑜𝑐𝑒𝑠𝑠 𝑃𝑉 𝑛 = C]
𝑉𝑛
Let us suppose, that the initial and final states of the system are (P1, V1, T1) and (P2, V2, T2) respectively. An integrating
equation (1).
𝑊 = 𝐶 Τ 1 − 𝑛 𝑉21−𝑛 − 𝑉11−𝑛 … … . . (2)

𝑀𝑜𝑟𝑒𝑜𝑣𝑒𝑟, 𝐶 = 𝑃2𝑉2𝑛 = P1𝑉1𝑛 =P𝑉 𝑛 … … … … … … … … . (3)


On combining equation (2) and (3) we obtain W = (P1V1 – P2V2)/(n- 1) …………. (4)
For a ideal gas P1V1 = RT1 and P2V2 = RT2
𝑇ℎ𝑖𝑠 𝑙𝑒𝑎𝑑𝑠 𝑡𝑜 𝑊 = 𝑅Τ 𝑛 − 1 𝑇1 − 𝑇2

Again change in internal energy E2 – E1 = CV (T2 – T1)

From first law of thermodynamic


P1V1  P2V2 
E2  E1  Q  W  E2  E1  Q 
n 1

 CV T2  T1   Q 
P1V1  P2V2 
n 1
We know that CP – CV = R  CP/CV – 1 = R/CV  CV = R/( - 1) [As CP/CV = ]

Heat exchanged
Q
R
T2  T1   R .T1  T2 
 1 n 1
 1 1 
 Q  RT1  T2   
 n  1   1

Q
R
T1  T2    n   W .   n
n 1   1   1
Problem
Air (ideal gas with  = 1.4 ) at 1 bar and 300K is compressed till the final volume is one-sixteenth of the original volume,
following a polytrophic process PV1.25 = constant.
Calculate (a) The final pressure and temperature of the air, (b) Work done and (c) The energy transferred as heat per mole of air.
Solution:
(a) Initial volume V1, and final volume V2

According to polytropic process P1V11.25 = P2V21.25 = c


1.25
V 
P2  P1  1   1  16   32bar P1V1  P2V2  T  P2V2 .T  32  V1  300  600K
1.25

 V2  T1 T2
2
P1V1
1
1 16  V1

8.314  300  600


.T1  T2  
R
(b) W   9976.8 J mole  9.977 KJ mole
n 1 1.25  1

 n 1.4  1.25
(c) Q W   9.977   3.741 KJ mole
 1 1.4  1
Problem
Calculate the heat lost to the surrounding when 1g sphere of zinc sulphide is roasted to zinc oxide at 900⁰K. The following dada
are given:
Hf0 at 298K (kcal) CP (cal/g.mole.K)
ZnS (S)  49.05 12.16 + (1.24 × 103 T) – (1.36 × 105 × T2)
ZnO (S)  83.80 11.71 + (1.22 × 103 T) – (2.18× 105 × T2)
O2 (g) 0 7.16 + (1.0 × 103 T) – (0.4 × 105 × T2)
SO2 (g) 70.94 10.38 + (2.54 × 103 T) – (1.42 × 105 × T2)

Solution: Zinc sulphide roasting reaction given below


ZnS  1.5O2  ZnO  SO2 ...............(1)

At 298⁰K the heat of reaction is given by H 298


0
 83.80  49.05  70.94  105.94kcal
900

Applying Kirchhoff’s law H 900  H 298   C


0 0
P dT
298


For reaction (1) C P  11.71  10.38  12.16  1.5  7.16   1.22  10  2.54  10  1.24  10  1.5  10 T
3 3 3 3


  2.18 105  1.42 105  1.36 105  0.6 105 T  2 
   
 C P  0.81  1.02 10 3 T  1.62 105 T 2 cal g.mole. K
H 900
0
 105690  488  368  182  550  106178cal

This figure represented that amount of heat lost when 1g.mole ZnS is roasted at steady state.

The molecular weight of ZnS = 97.43


106178
Amount of heat lost =  1090 cal g
97.43

Note:
In this problem it is implied that the reactants, ZnS and O2 as well as the products ZnO & SO2 are at a steady
state temperature of 900⁰K.
It is also clear that maintenance of the steady state condition require that all heat generated by the roasting
process be delivered to the surrounding.
SECOND LAW OF THERMODYNAMICS
The first law of thermodynamics essentially and generalization of the principle of conservation of energy.

Prolonged
period of time

Work can be converted completely into heat, the reverse is not true with the help of a suitable device.

The second law, on the other hand, is capable of expressing the inherent limit to the efficiency of converting heat into work

Sadi Carnot (1796-1832), one of the outstanding workers in the history of thermodynamics, intensively studied the
fundamental reasons for the inherent limit to the efficiency.

The Clausius statement of the second law: No process is possible whose sole result is the transfer of heat from a colder body
to a hotter body.
The Kelvin-Plank statement of the second law: no process is possible whose sole result is the complete conversion of heat into
work.
EFFICIENCY OF A HEAT ENGINE
A heat engine is a device that converts heat into work.
The heat engine uses working substances (i.e. ideal gas, real gas, steam etc) and suffers no
permanent change during operation.

The working substances a system is taken through a series of processor that together from a cycle
the heat engine will absorbed heat during one part of the cycle, convert a portion of it into work
and reject the reminder during another part of the cycle.

No heat engine is 100% efficient, because a portion , however small, of the heat is observed is always
rejected

The efficiency of the heat engine () = W/Q2

From first law of consideration, for cyclic process: W = Q1 + Q2

Where Q2 is positive (because heat is absorbed by the system) and Q1 is negative (because heat is rejected by the system)

Q1
  1
Q2
The Carnot cycle
One of the most important cyclic processes is the Carnot cycle, first described by Sadi Carnot in 1824.

Figure: Equipment for carrying out a Carnot cycle.


Step A: Initially the system is in the thermodynamic state defined by the state
variables P1, V1, T1. The cylinder is there insulated and the gas is compressed
adiabatically and reversibly until its temperature became T2.
The changes occurring the steps are as follows QA = 0 (adiabatic process)

E A  CV (T2  T1 )
The work done is given by WA  E A  CV (T2  T1 )

Further more for an adiabatic process, T1V1 1  T2V2 1 where   C P CV Figure: The Carnot cycle shown in P-
V coordinates
T2 T1  V1 V2  ...............(1)
 1

Step B: The insulation jacket is removed and the system is contact with the high temperature reservoir at T2. It is allowed to
exposed isothermally and reversibly. Heat is absorbed by this system during the expansion and useful work is done on the
surrounding
EB = 0 (ideal gas), WB  RT2 ln V2' V2  and QB  Q2  WB  RT2 ln V2' V2  
Step C: The system is taken out of contact with the high temperature reservoir and is insulated it is allowed to expand
adiabatically and reversibly unit 1 the temperature of the system decline to T1
QC  0, EC  CV T1  T2  and WC  EC  CV T1  T2 

T2 V2   T1 V1 or T2 T1  V1 V2 


 1  1  1
...................(2)
Step D: The insulation is removed, and the system is placed in contact with the low temperature reservoir at T1. The system is
the compressed isothermally and reversibly until it reaches the initial state.
ED  0, WD  RT1 ln V1 V1 and QD  Q1  RT1 ln V1 V1

Heat is rejected by the system during the step (i.e, QD or Q1 is negative).

Enet  E A  EB  EC  ED  0

Wnet  WB  WD  RT2 lnV2 V2   RT1 lnV1 V1


 1  1
 V1   V1'  V2' V1'
From equations (1) and equation (2) we can get     '  
 V2   V2  V2 V1

   
Wnet  RT2 ln V2' V2  RT1 ln V1 V1'
 Wnet  RT V V   RT ln V V 
2 2
'
2 1 1
'
1

 
Wnet  RT ln V2' V2 T2  T1   Q2  Q1
 Wnet
 

Q2  Q1 R ln V2' V2 T2  T1  
 T2  T1  T2
The efficiency of a heat engine operating on the Carnot Cycle is given by
Q2 Q2 RT 2 ln V2' V1  
T1 Q1
  1  1 ...........(6  8)
T2 Q2
Carnot Refrigerator

Figure: The Carnot cycle shown in P-V coordinates Figure: Schematic diagram of Carnot Refrigerator

𝑉1′ൗ
𝑄1′ = heat absorbed = 𝑅𝑇1 ln( 𝑉1) = −𝑄1

𝑄2′ = heat rejected = 𝑅𝑇2 ln(𝑉2Τ𝑉 2) = −𝑄2


  RT1 ln V1 V1   RT2 ln V2 V2 


Wnet

For equation (1) and (2) V2 V2  V1 V1


'
Wnet  RT1 lnV1 V1   RT2 lnV2 V2    R lnV1 V1 T2  T1 

The coefficient of performance for a refrigerator is defined by,

 RT1 ln V1 V1  T1
C   Q1 Wnet
  
 R ln V1 V1 T2  T1  T2  T1

The coefficient of performance for a refrigerator, unlike the efficiency of heat engine, can be considerably greater
than unity.
The concept of entropy

The efficiency of a Carnot heat engine operating between a hot reservoir at T2 and cold reservoir at T1 is given by,

T1 Q Q Q
  1  1  1  1  2  0...............(1)
T2 Q2 T1 T2

All four steps in the Carnot cycle are reversible. Thus, the heat from appearing in equation (1) refer to heats exchanged in
reversible process.
Q1,r Q2,r
  0..............(2)
T1 T2

Where the subscript r denotes a reversible process. This can be written as  Qi ,r Ti  0.......(3)

Suppose we have consider a Carnot cycle in which T1 and T2 differ by only an infinitesimal amount. In such a cycle, Qi,r
represents the heat exchanged reversibly at temperature Ti.

For the entire cycle  cycle Qi ,r Ti  0


Q1,r Q2,r
For the Carnot cycle abcd  0
T1 T2

Q3,r Q4,r
For the Carnot cycle efgh  0
T3 T4

Q1,r Q2,r Q3,r Q4,r


For the two cycles    0
T1 T2 T3 T4

For the entire reversible cycle  cycle Qi ,r Ti  0

Although our proceeding discussions focused on irreversible cycles, the


arrangements can be applied to any generalized, reversible cycle process. Such
a process can be approximated closely by a series of cannot cycles, much as in Figure 1: An arbitrary reversible cycle that
𝛿𝑄
Figure 1 and the ratio 𝑖,𝑟ൗ𝑇𝑖 summed over the entire cycle can be shown to be is approximated by a series of Carnot
equal to zero. cycle.

The quantity 𝛿𝑄𝑟ൗ𝑇 has the attributes to state property. Its value does not change in a cyclic process and any change occurring in
it during a process is independent of the process path and depends only on the initial and final states of the system. The
quantity 𝛿𝑄𝑟ൗ𝑇 is designated dS, where S is a new thermodynamics quantity call entropy.
Qr
dS  ........................(3)
T
 dS  0..............................(4)
cycle

A system undergoes a change from an initial thermodynamic state (P1, V1, T1) to a final thermodynamic state (P2, V2, T2), the
2 𝛿𝑄
associated entropy change is given by ∆𝑆 = 𝑆2 − 𝑆1 = ‫׬‬1 𝑇 𝑟
where S1 and S2 are determined solely by the state variable.

Entropy Changes in Typical Reversible Process


(1) Reversible adiabatic process:
The initial and final states are given by (P1, V1, T1) and (P2, V2, T2) respectively. Because no heat leaves or enters the system,
Qr = 0. Hence S = 0
Entropy does not change in reversible adiabatic processes and because of this, they are also known as isentropic process.
However, entropy can change in an irreversible adiabatic process.

(2) Reversible isothermal process (ideal-gas system):


𝑉2
The initial and final states are given by (P1, V1, T1) and (P2, V2, T1). The heat absorbed, 𝑄2,𝑟 = 𝑅𝑇𝑙𝑛 . The temperature =T.
𝑉1

𝑄2,𝑟 𝑉2
The entropy change, ∆𝑆 = σ  ∆𝑆 = 𝑅 𝑙𝑛
𝑇 𝑉1
(3) Reversible isochoric or isovolumic process:

In such a process, no work is done and the heat exchanged is given by 𝛿𝑄𝑟 = 𝑑𝐸 = 𝐶𝑉 𝑑𝑇
𝛿𝑄𝑟
By definition 𝑑𝑆 = = 𝐶𝑉 𝑑𝑙𝑛𝑇
𝑇

If a system is subjected to an isochoric process defined by the initial state (P1, V1, T1) and (P2, V1, T2) the final state the
entropy change would be ∆𝑆 = ‫𝑇𝑛𝑙 𝑑 𝑉𝐶 ׬ = 𝑆𝑑 ׬‬

𝑇2
If CV is independent of temperature ∆𝑆 = 𝐶𝑉 𝑙𝑛 𝑇1

(4) Reversible isobaric process:


The heat exchanged in the process is equal to the change in enthalpy. That is 𝛿𝑄𝑟 = 𝑑𝐻 = 𝐶𝑃 𝑑𝑇
𝛿𝑄𝑟
Hence 𝑑𝑆 = = 𝐶𝑃 𝑑 𝑙𝑛𝑇
𝑇

For the process defined by the initial state (P1, V1, T1) and final state (P2, V1, T2) the final state.

∆𝑆 = 𝑆2 − 𝑆1 = න 𝐶𝑃 𝑑 ln 𝑇

For a substance with CP defined by CP = a + bT + c/T2


𝑇2 𝑐 𝑐
The entropy change is given by ∆𝑆 = 𝑎 𝑙𝑛 + 𝑏 𝑇2 − 𝑇1 − 2 𝑇22 + 2 𝑇12
𝑇1
𝑇𝑚
(5) Fusion of Solids: Consider the fusion of a solid at its normal melting point. 𝐴 𝑆 𝐴 𝑙

The fusion process proceeds at a constant temperature at 1 atm pressure.


0
Heat absorbed by the system =𝑄𝑟 = ∆𝐻𝑚

𝑄𝑟 ∆𝐻𝑚0
Entropy of fusion = 0
∆𝑆𝑚 = =
𝑇𝑟 𝑇𝑚

The entropy of the liquid is greater than that of the solid.

Example: Copper melts at 1083⁰C and its heat of fusion is 12970 J/mole. Calculate the entropy of fusion of copper.

Solution: H m0  12970 J mole

Tm  1083  273  1356K


12970
S m0   9.565 J mole. K
1356
(6) Evaporation of a Liquid:
𝑇𝑣
The process of evaporation 𝐴 𝑙 → 𝐴 𝑔 occurs at a constant temperature, Tv and constant pressure. During the process, heat
is absorbed and the liquid is said to boil. The temperature at which it boils is known as the boiling point. The heat absorbed is
labelled the heat of vaporization, designated ∆𝐻𝑣0 .
∆𝐻𝑣0
Heat absorbed by the system = ∆𝐻𝑣0 . The entropy of vaporization ∆𝑆𝑣0 = 𝑇𝑣
Example: Copper boils off at 2575⁰C and the heat vaporization is reported as 304.4 kJ/mole. Calculate the entropy of
vaporization.

Solution: H v0  304400 J mole Tv  2575  273  2848K

304400
S v0   106.9 J mole. K
2848

(7) Polymorphic Changes:

Solid iron can exit in three different allotropic modifications, for example, face-entered-cubic γ-iron is transformed to body-
cantered-cubic δ- iron at 1391⁰C. The transformation of γ- iron to δ- iron at 1391⁰C, the so called transition temperature. Tf , is
accompanied by absorption of heat. The transformation occurs at constant temperature and constant pressure.

Heat absorbed = Q = ∆𝐻𝑡0

Temperature of transformation = Tf

Entropy of transformation = ∆𝑆𝑓0

S 0f  H 0f T f
(8) Entropy changes in some irreversible process:

a. Free expansion of an ideal gas: This is an irreversible adiabatic process. (P1, V1, T1) represents the initial state and (P2, V2,
T1) denotes the final state after free expansion( as indicated earlier, the temperature of an ideal gas remains unaltered during
free expansion).

To calculate the entropy change associated with process we consider a single reversible process as a combination of reversible
process whose net result is to induce the same changes as the irreversible process.

A reversible isothermal process from (P1, V1, T1) to (P2, V2, T1) produces the same changes in the state variables as does
free expansion

Heat absorbed in the isothermal process = Qr = RT1 ln (V2/V1)

Temperature = T1 and entropy change, Sfe = R ln (V2/V1)

A combination consisting of a reversible isobaric process followed by a reversible isochoric (isovolume) process can also be
used to simulate the given free expansion process.
In the free-expansion process, although no heat is exchanged, there is a net gain in the entropy of the system.

For an irreversible process TdS  Q


(2) Solidification of super cooled liquids

Liquid metal can be super cooled to temperature considerably below their normal solidification temperatures.
Solidification of such liquids occurs spontaneously ( or irreversibly).

1g.mole of liquid silver that has been super cooled to 940⁰C is the system under consideration. The super cooled liquid is
allowed to solidify at 940⁰C.
Sirreversible = ?
The calculation is facilitated by constructing reversible cycle path between the
initial and the final state of the system. Such a reversible cyclic path, consist of
four sub processes, is shown in figure- 1.

I. Heat the super cooled liquid silver reversibly and isobarically to 960.5⁰C (the
normal solidification temperature)
II. Allow solidification of liquid silver reversibly at 960.5⁰C at 1 atm pressure.
III. Cool the solid silver reversibly and isobarically to a temperature at 940⁰C.
IV. Allow melting of the solid silver reversibly at 940⁰C at 1 atm pressure.
Figure 1: A cyclic path consisting
of four subprocesses all of which
SI, SII , SIII and SIV denote the entropy changes associated with the are reversible
four reversible sub processes.
 dS  0  S I  S II  S III  S IV  0
 S IV  S I  S II  S III .....................(1)
Note that the quantity on the left side of equation-(1) represents the entropy change associated with solidification of
liquid silver at 940⁰C.
The following thermochemical date are taken from Kubaschewski and Alcock
 10 5 
C Pl liquid silver   30.5 J mole.K
 
C PS solid silver   21.3  8.54  10 T  1.51 2  J mole.K
3

 T 
H s0  H m0  11090 J mole
1233.5
Sub process I: dS  C Pe T dT  S I   30.5 T dT  0.51 J mole.  K
1213

Sub process II: S II  H S  H l  Tm  H S 1233.5  8.99 J mole. K


0

1213
Sub process III: dS  C PS T dT  S III   C PS T dT  0.53 J mole.  K
1233.5

 S 4  9.01 J mole.K

In other words, when super cooled liquid silver at 940⁰C is allowed to solidify at that temperature, its entropy
decreases by 9.01 J/g.mole.K.
Entropy change of the universe in an irreversible process
Incorporation of an irreversible step into a Carnot cycle produces changes in the thermodynamic efficiency and the entropy
of the universe. The conventional Carnot cycle consists of four reversible steps:
Adiabatic compression (A), isothermal expansion (B), adiabatic expansion (C), and isothermal compression (D).

We shall consider a modified cycle in which an irreversible step B’ is


interposed between step B and C. The details of the modified cycle are
shown in P-V diagram in Figure 1.

There may be an increase in the entropy of the universe each time the
working substance is taken through the modified cycle. For simplicity, 1
mole of ideal gas is considered as the working substance.

Figure 1: Modified Carnot cycle with one


Figure 2: (a) at the beginning of the irreversible free expansion (step B’) and irreversible step (B’)
(b) at the completion of the irreversible free expansion (step B’)
The additional process introduced here is an adiabatic free expansion this step is clearly irreversible. The practical details of
how one might carry out step B’ in the context of the modified cycle are elaborated in Figure 2.

Fig-2(a), at the completion of step B all the gas is on the right of this diagram that separates it from an evacuated chamber on
the left.
After the cylinder and piston are thoroughly insulated, suppose that an opportune moment the diaphragm is ruptured, where
upon the gas rushes into the evacuated chamber.
At the conclusion of step B’, then this gas will occupy the entire space in the cylinder as shown in Figure 2(b).

Because there is no change in the temperature of an ideal gas during an adiabatic free expansion at the completion of step B’
the thermodynamic state of the system is defined as (P3’, V3’, T2).

The remainder the modified cycle consists of two reversible adiabatic (A and C’) and two reversible isothermals ( B & D ).

Step A : Heat absorbed QA = 0, Entropy change SA = 0 …………… (1)


and work done WA = EA = CV(T1-T2)

Step B : Heat absorbed QB = Q2 = RT2 ln(V3/V2)

Entropy change SB = R ln(V3/V2) …………………(2)

work done WB = Q2 = RT2 ln(V3/V2),


Step B’ : Heat absorbed QB’ = 0, work done WB’ = 0

The entropy change can be calculated by assuming a reversible isothermal process between thermodynamic state 3 and 3′.

Entropy change SB’ = R ln(V3’/V3) …………….. (3)

Step C′ : Heat absorbed QC’ = 0, Entropy change SC’ = 0 …………… (4)

and work done WC’ = EC’ = CV(T2 - T1)

Step D : Heat absorbed QD = Q’1 = RT2 ln(V1/V’4 )

Entropy change SD = R ln(V1/V’4) …………………(5)

work done WD = Q’1 = RT2 ln(V1/V’4),

For the entire cyclic process, W  W  W


i A  WB  WB  WC  WD
 CV T1  T2   RT2 ln V3 V2   0  W T2  T1   RT1 ln V1 V4 
 RT2 ln V3 V2   RT1 ln V1 V4 ...........................................................(6).

S S  S A  S B  S B  SC   S D  S s  0  R ln V3 V2   R ln V3 V3   0  R ln V1 V4 


 R ln V3 V2   R ln V1 V4   R ln V3 V3 .............................(7).
The two reversible adiabatic steps, A and C′
 1  1
T2V3'  T1V4'
T1V1 1  T2V2 1
 1
 V2 
 1 V ' 
 T1 T2   
and 3
T1 T2    V ' 
 V1   4 
By considering this two reversible adiabatic steps, A and C’, we can show that

  
....................8
V2 V3 V3 V4
  
V1 V  V2 V1
4
  V   ln V  
   R 3 
V
An substituting equation (8) into equation (7), we obtain S s  R ln   R ln
3 1
V  V    V3 
 2  4   
V V V   V V
 R ln  3
 1
 3 
 R ln 1  3 
 V2 V  V3   V  V2 
 4   4 

V V   
V3 V4
S S  R ln  1  4   R ln 1  0........................(9)
V4 V1 

V2 V1
This result is entirely reasonable considering the fact that we are dealing with the entropy change in a cyclic in a process.
By combining equations (8) and (6), we find
 V3  V 
Wnet  RT2 ln   RT1 ln 1 
 V2  V  
 4 
 V3  V 
 RT2 ln   RT1 ln 2 
 V2  V  
 3 
 V3  V V 
 RT2 ln   RT1 ln 2  3 
 V2   V3 V  
 3 

 V3   V3  V 
 RT2    RT1 ln   RT1 ln 3 
 V2   V2  V  
 3 
  V3  V  
  R ln  T2  T1   RT1 ln 3 ............................(9)
 V2   V3 
  
The efficiency of the modified cycle is given by


R lnV3 V2 T2  T1   RT1 lnV3 V3 
   Wnet Q2   
RT2 ln V3 V2 
T2  T1  
  RT1 ln(V3 V3 ) RT2 ln(V3 V2 )
T2  
T1  
 1  RT1 lnV3 V3  RT2 ln V3 V2 .................................10 
T2   

Because V3 is greater than either V3 or V2 , the third term on the right hand side is a negative quantity. Hence, it
follows that  is less than the Carnot efficiency. The diminution (decrease) in the efficiency of the modified cycle is
directly linked to the presence of the irreversible step B’, the greater the departure of V3 from V3, the longer is
decrease in efficiency.
In order to determine the entropy change of the universe due to the modified cycle, we need to calculate the entropy change of
the surroundings. The higher temperature reservoir “losses” an amount of thermal energy equal to Q2 . The low-temperature
reservoir “grains” an amount of heat equal to Q1’.

 V3 
  R ln ..................11
Q2
Entropy change of high-temperature reservoir = 
T2  V2 
 V 
  R ln 1 ................................12 
Entropy change of the low- temperature reservoir =  Q1
T1 V  
 4 

 V3  V 
Sl   R ln   R ln 1 ......................................13
 V2  V  
 4 
On combining equation (13) with equation (8), we find that

 V3  V   V V 
Sl   R ln   R ln 2 ......................................14  as 3  4 
 V2  V    V2 V1 
 3  
V  V V 
S l   R ln  3   R ln  2  3 
 V2   V3 
 V3 

V  V  V 
  R ln  3   R ln  3   R ln  3 
 V2   V2   
 V3 
V  
 R ln  3 
 V3 
 
Because V3’ is greater than V3, this quantity is positive the entropy change of the universe is given by

V  
Su  Sl  R ln 3 .......................15
 V3 
 
It can be said that the presence of an irreversible step in an otherwise reversible cyclic process will reduce the efficiency of a
heat engine operating on the cycle, furthermore, there will be an increase in the entropy of the universe each time the cycle is
completed.

Equation (9) and (15) can now be combined into a single expression Su  0 or dSu  0 ………….. (16)
The clausius inequality
The efficiency of a heat engine operating on a modified cyclic process containing an extra irreversible step is less than the
Carnot-cycle efficiency.
 RT ln V  V   
 T1   3 
 Q2  Q1
...........................17 
1 3
   1     


T2  RT2 ln V3 V2  
Q2
 
  
T Q
Because V3’  V3  V2 it is clear that 1  1   2  Q Q Q '

 T 
1
 0  2
 1

 2  Q2 T2 T1

It is important to note that Q2 and Q1 are the heat exchanged reversibly by the system with the surroundings.

Q2  Q2,r  heat absorbed reversibly by the system at T2



Q1  Q1,r  heat rejected reversibly by the system at T1

Q2,r Q1,r
For the cyclic process containing are irreversible step   0   Qi ,r Ti  0...............(18)
T2 T1

For the strict Carnot cycle consisting of four reversible step, Q i ,r Ti  0.............................19
Equations (18) and (19) can be combined to yield the following general expression:

Q i ,r Ti  0.............................20

In other words, wherever a system is taken around a cycle consisting of several steps, the sum of the heat exchanged
reversible Qi,r divided by the reservoir temperature, Ti for each step is less than equal to zero. This known as the
Clausius in equality.

Combined statement of first and second law


The mathematical formulation of the first law of thermodynamics, in differential form, is as follows:
Q  dE  W ....................................1
In general, the work term W is composed of two parts:- Wm, the mechanical (or P-V-T) work and Wn, the non
mechanical (i.e., magnetic, electrical etc) work, therefore

Q  dE  Wm  Wn ............................2


For a system for which only the mechanical work need to be considered, Wn = 0 and Q = dE + Wm ……… (3)

The mechanical work done in a reversible process is Wm = P dV


Equation (2) and (3) together yield Q  dE  PdV ...........................4

The usefulness of the first law equation (4) is severely limited by the fact that is not a complete differential, that it is
dependent on the both of the process. This short coming can be overcome by combining this relationship with the second law
statement. According to the second law, for a reversible process spanning two equilibrium states, the entropy change , dS, is
equal to the heat exchanged, Qr, divided by this temperature, T.
Qr
dS   Qr  TdS ...............................5
T

It is clear that for any reversible process involving only mechanical work,

TdS  dE  PdV .............................6

For the general case in which several different terms of work are involved,

TdS  dE  PdV  wn ,r .............................7 


Equation (6) and (7) are formulations of the combined first and second laws for a P-V-T system and a generalized system,
respectively.
Some useful thermodynamic relationships
(1) T and P are chosen as independent variables.

TdS  dE  PdV
dE  P 
 dS    dV
T T 
Considering E and V as function of T and P
 E   E 
dE    dP    dT ............................................1
 P T  T  P

 V   V 
dV    dP    dT ............................................2
 P T  T  P

Substitution gives  E 
1  E   P  V   V  
dS    dP    dT     dP    dT 

T  P P  T P  T  P T  T P 
1  E   V   1  E   V  
    P    dP     P    dT ...........................3
T  P T  P T  T  T  P  T  P 
Expressing S as a function of P and T

 S   S 
dS    dP    dT ............................................4
 P T  T  P

Comparing the coefficient of dP and dT in equation (3) and (4)

 S  1  E   V  
      P  ...........................................5
 P T T  P T  T T 

 S  1  E   V  
      P  ...........................................6
 T P T  T P  T P 

Differentiating equation (5) with respect to T at constant P,

2S 1  2E   2V   1   E   V  
   P    2     P   ..............................7 
TP T  TP  TP   T   P T  P T 

Differentiating equation (6) with respect to P at constant T

2S 1  2E   2V   V  
   P     ..............................8
PT T  TP  TP   T  P 
The mixed second-order differentials S appearing on the left side of the equation (7) and (8) are equal.
1 2E   2V   1   E   V    1    E
2
  2V   V  
  P    2     P       P     
T T P  T P    
T  P T   P T    
T T P  T P   T P 
 1   E   V   1  V 
 0   2     P    
  
T P T  P T  T  T P
 1   E   V    1  V  
 0   2     P       
 T    P  T  P  T  T  T P 
1  E   V    V 
    P     
T  P T  P T   T P
 E   V   V 
   T    P  .................................9 
 P T  T  P  P T
 E   V   V 
   P   T   .................................10 
 P T   P T  T P
 S  1  E   V  
From equation (5) and (10)       P   .........................................5
 P T T  P T  P T 

 S   V 
      ......................................11
 P T  T  P
From equation (6)  S  1  E   V  
      P  
 T P T   T P  T P 
1   E  PV   1  H 
      .............................12 
T T  P T  T  P

 V  1  H 
As a result we have dS    dP    dT .....................................13
 T  P T  T  P

From the definition of CP , the heat capacity at constant pressure, we note that

 1  H  C
    P
 T  T  P T
We can now define a quantity called the coefficient of thermal expansion, , as follows
 1  V 
     .....................(14)
 V  T  P
Thus,  is the volume change per unit volume per degree change in temperature at constant pressure.
 V 
   V
 T  P
C 
dS  VdP   P dT .......................................15
 T 
 V 
TdS  C P dT  T   dP........................................16
 T  P

(2) T and V are chosen as independent variables

TdS  dE  PdV .................................6


 S   S 
dS    dV    dT .......................................17 
 V T  T V

Considering E and V as a function T and V


 E   E   V   V 
dE    dT    dV and dV    dT    dV
 T V  V T  T V  V T

 1  E    1  E 
dS      P dV     dT .................................18
 
T V T   
T T V

Because dV and dT are independent, comparison of the coefficient of these deferential yields.
 S   1   E  
 S   1   E          ...................................................20
        P ...................................................19  T V  T   T V 
 V T  T   V T 

Differentiating equation (19) with respect to T at constant V,

2S 1   2 E  P   1  E  
        P ............................21
TV T  TV  T V  T 2  V  T 
Differentiating equation (20) with respect to V at constant T

2S 1 2E
  ....................(22)
VT T VT

Setting equation (21) and (22) equal and cancelling the mixed second-order differentials,

 1   E  P    1   E   1 2E
2
  V      2     P  . T
 T   TV  T V   T   V  T  T VT

1  P   1   E  
0    2     P 
T  T V  T   V  T 
 E   P 
   P  T  ............................................................23
 V  T  T V
 S   P 
As a result     ..............................24
 V T  T V

 S   1  E    E  
 .....................25
C
and        V  As C  
 V T  T  T V T 
V
  T V 

 P  C 
Put the value equation in (17) dS    dV   V dT ..................................26
 T V  T 

The partial differential 𝜕𝑃Τ𝜕𝑇 𝑉 can be expressed in terms of measurable quantities like the coefficient of thermal
expansion and isothermal compressibility. Later its defined as follows

1  V 
  Isothermal Compressibility     ...............................27 
V  P T
 is defined on the change in volume per unit volume per unit change in pressure under isothermal conditions are dividing
coefficient of thermal expansion  (equation 14) with isothermal compressibility  (equation (27))
 1  V 
    
 V  T  P

 1  V   V 
    
  
    P     P ................................28
V T T
  1  V   V 
    
 
V  P T   P T
Let us consider a constant volume process, for which equation (2) reduce to

 V   V 
dV    dPV    dTV  0...............................29
 P T  T P

The subscript V indicates that the changes occur in a constant volume process. Rearranging term in equation (29).

 V   V 
  dPV    dTV
 P T  T  P
 V 
 
dPV  P   T  P ....................................30 
   
dTV  T V  V 
 
  P T
 P  
Comparing equation (28) and (30), we obtain    .................................31
 T V 
C   P 
Substitute equation (31) into equation (27) dS   V dT    dV
 T   T V
C  
  V dT  dV ..........................................32 
 T  
T
 TdS  CV dT  dV .........................................33

(3) P and V are chosen as independent variables

Expressing S and E as functions of P and V, we can show that

 S  1  E 
     .....................................34
 P V T  P V

 S  1  E 
  P.....................................35
1
   
 V  P T  V  P T

Adopting the same procedure as before, it can be shown that

 S   C  T   C   
    V     V  ..........................36
 P V  T  P V  T   
 S   C  T   C 
    V     V ..........................37 
 V  P  T  V  P  TV 

 T   T 
TdS  CV   dP  C P  dV .................................38
 P V  V  P
Problem
Calculate the coefficient of thermal expansion and isothermal compressibility for an (a) Ideal gas (b) a Van der Waals
gas. By defining
 1  V   1  V 
     and     
 V  T  P  V  P T

Solution
(a) For 1 mole of an ideal gas, the equation of state is PV = RT
 1  V 
    
 
V  T P
 V  R  1
    .
1 R
T 
PV

R

 T  P P V P  R PV T 
R 1
 
VR T

 V  RT
   2
 P T P

Substitutions into the expression for  gives as


 1  V 
    
 V  P  T
 1    RT 
     2 
V   P 
RT  PV R 1
 R  T  
VP 2  PV T 
T 1
 
TP P
 a 
(b) For 1 mol of Van der Waals gas, the equation of state is  P  2 
V  b   RT
 V 
Differentiating of the equation of state with respect V at constant P yields.

 T 
  

RTV 3 2aV  b 
2

 V P RV 3 V  b 

Again  V 
  
1  V 
  

RV 3 V  b  
 T  P  T     P RTV  2aV  b 
3 2
  T
 V  P
Substitute in equation of   1  V  1 RV 3 V  b 
       2
 
V T P V  RTV 3
 2a V  b  

 RV 2 V  b 
  2
 RTV 3
 2a V  b  

 P  
To find the value of , we make use of equation   
 T V 
 P  R 
   
 V V V  b 
 V  b 

R


RV V  b V  b 
2

RTV 3  2aV  b 
2

 
 V
V  b 
22

RTV 3  2aV  b 
2
Difference between heat capacities CP and CV

From first law consideration alone it can be shown that

 E   V 
C P  CV     P   ..............................................1
 V T  T  P

 E   P 
Again we derived    P  T 
 V T  T V
 P   V 
C P  CV  T     ..............................................2
 T V  T  P

 1  V 
Again coefficient of thermal expansion     
 V  T  P

 1  V 
and isothermal compressibility     
 V  P T  V  R
  
 T  P P
For 1mole of ideal gas PV = RT 1 R 1
   
V P T
 V  RT
   2
 P  T P
 1   RT  1
        2  
V   P  P

 V   P  
   V and   
 T  P  T V 
 T 2V
C P  CV  TV . 
 
1 PV  1 1
 C P  CV  TV  .P     and  
T2 T T P 

 Rthe gas constant 


PV
 C P  CV 
T
From the kinetic theory of the gases, we know that for a monatomic gas CV = 3/2 R and CP = 5/2 R with  = CP/CV =
1.667 .

The measured values of CP, CV and  for monatomic gases at room temperature are in good agreement with those
values. For a diatomic gas, the values of CV, CP and  are 5/2 R, 7/2R and 1.40, respectively.
Problem: Show that for an ideal gas the internal energy is solely function of temperature

Solution:
Because the internal energy E is a state function, it can be expressed as a function of state variables such as V and T.
 E   E 
E = f(V, T) and dE    dT    dV
 T V  V T

 E   P 
   P  T 
 V  T  T V
 E   P 
   T  P
 V  T  T V
 P  R
For 1 mole of an ideal gas, PV = RT. Differentiating at constant volume   
 T V V

 E  TR  TR 
   P0  here  P 
 V T V V

 E    E  
dE    dT  CV dT  as    CV E = f(T) for an ideal gas
 T V   T V 
Entropy of an ideal gas
From useful thermodynamic correlation T and P are independent variables.
Let us begin with equation according to the equation for 1 mole of an ideal gas

 V  dT  V 
TdS  C P dT  T   dP  dS  C P   dP.................................1
 T  P T  T  P
Let Sr, Pr and Tr denote the entropy, pressure and temperature of the substance in a reference state. Frequently Tr = 298 K
and Pr = 1 atm defined the chosen reference state integration of equation yields.

S T P  V 
Sr dS  Tr P
C d ln T  Pr  T  P dP
T P  V 
 S  S r   C P d ln T     dP
Pr T
Tr
 P
P  V 
 dP.........................2 
T
 S  S r   C P d ln T   
Pr T
Tr
 P

Equation (2) is valid for any substance, including an ideal


 V  R
Equation for 1 mole of an ideal gas PV = RT, Integrating with constant pressure   
 T  P P
T P
 S  S r   C P d ln T   Rd ln P
Tr Pr
Because CP is independent of temperature for an ideal gas,
T  P
S  S r  C P ln   R ln .......................3
 Tr   Pr 

This gives the entropy as a function of temperature and pressure to calculate the entropy as a function of T and V .

C   
dS   V  dT   dV
 T   

For an ideal gas,  = 1/T and  = 1/P, thus / = P/T = R/V [As PV = RT  P/T = R/V]

dV
dS  CV d ln T  R  CV d ln T  R ln V
V
T  V 
S  S r  CV ln   R ln ........................................4
 Tr   Vr 

Similarly, we can calculate the entropy as a function of P and V

 T   T 
TdS  CV   dP  C P   dV

 V
P  V P
CV  T  C  T 
dS    dP  P   dV
T  P V T  V  P
CV  C
 . dP  P dV
T  TV
C T CP
 V  dP  dV
T P 1
 TV
T
 CV d ln p  C P d ln V

P V 
S  S r  CV ln    C P ln  .................................6.97 
 Pr   Vr 
Absolute value of entropy of a substance

The absolute value for entropy, unlike that for enthalpy or internal energy can be determined. There are numerous stand
order compilations, such as that by kubaschewski and Alcock, listing absolute entropy values for substances at a standard
temperature of 298.15K (or 298⁰K, for short). The entropy change under isobaric condition the equation is expressed as

 CP 
dS  VdP   dT
 T 
For isobaric dP = 0

 CP   S 0 
  P ........................................1
C
dS   dT  
T   T  P T
CP
 dS 0  dT
T

 CP 
dT ..............................................2 
298
0
S 298  S 00   
0
 T 
The degree sign denotes standard state at 1 atm. Here, S00 denotes the entropy of the substances at absolute zero. The
quantity of the left side of equation (2) that is the difference between the entropies at 298⁰K and 0⁰K. It Can be determined
by contracting a plot CP/T against T and then measuring the area under the curve. If we put S00= 0
dT ..................................3
298 CP
S 0
298 
0 T

At low temperature, specially near absolute zero, data on heat capacities are lacking for many substances. This lack of data
is overcome by making extrapolations to lower temperature. In this regards, this following relationship for heat capacities
CP has proved useful at low temperature.

C P  aT 3  T  CV .........................................4 

Where a and  are constants. Over this range of applicability of equation (4), the entropy ST0 at T can be obtained from
equations (3) and (4).
ST0   aT 3  T 
 dT 

 T 
aT 3
S   T ..............................................5
0
T
3
Over the range of validity of equation (4), because  is small, the entropy ST0 at TK tends to be less than CP at the same
temperature. At low temperature, (but not T < 1K), can be neglected and

T 3 CP
S a
0
T  ............................................6 
3 3

With equation (6), the extrapolation down to 0⁰K is not satisfactory, because deviations occur from observations in the
range T < 1K.

You might also like