<<

LECTURE NOTES ON INTERMEDIATE

Joseph M. Powers

Department of Aerospace and Mechanical University of Notre Dame Notre Dame, Indiana 46556-5637 USA

updated 28 January 2019, 3:07pm 2

CC BY-NC-ND. 28 January 2019, J. M. Powers. Contents

Preface 7

1 Review 9

2 Cycle analysis 17 2.1 Carnot...... 17 2.1.1 Analysis for a calorically perfect ideal ...... 17 2.1.2 ...... 24 2.2 Rankine ...... 28 2.2.1 Classical...... 28 2.2.2 Reheat...... 33 2.2.3 Regeneration ...... 33 2.2.4 Losses ...... 37 2.2.5 ...... 37 2.3 Airstandardcycles ...... 37 2.4 Brayton ...... 38 2.4.1 Classical...... 38 2.4.2 Regeneration ...... 43 2.4.3 EricssonCycle...... 46 2.4.4 Jetpropulsion...... 48 2.5 Reciprocating cycles ...... 53 2.6 Otto ...... 55 2.7 Diesel ...... 59 2.8 Stirling...... 64 2.9 ...... 65 2.9.1 -compression ...... 65 2.9.2 Airstandard...... 69

3 Gas mixtures 71 3.1 Somegeneralissues...... 71 3.2 Idealandnon-idealmixtures...... 75 3.3 Idealmixturesofidealgases ...... 75 3.3.1 Daltonmodel ...... 76

3 4 CONTENTS

3.3.1.1 Binary mixtures ...... 79 3.3.1.2 of mixing ...... 82 3.3.1.3 Mixturesofconstantmassfraction ...... 85 3.3.2 Summary of properties for the Dalton mixture model ...... 86 3.3.3 Amagatmodel ...... 92 3.4 Gas-vapormixtures...... 92 3.4.1 Firstlaw...... 95 3.4.2 Adiabaticsaturation ...... 97 3.4.3 Wet-bulbanddry-bulbtemperatures ...... 99 3.4.4 Psychrometricchart ...... 99

4 Mathematical foundations of thermodynamics 101 4.1 Exact differentials and state functions ...... 101 4.2 Twoindependentvariables...... 106 4.3 Legendretransformations...... 109 4.4 Heatcapacity ...... 114 4.5 VanderWaalsgas ...... 118 4.6 Redlich-Kwonggas ...... 124 4.7 and generalized charts ...... 125 4.8 Mixtures with variable composition ...... 126 4.9 Partialmolarproperties ...... 129 4.9.1 Homogeneousfunctions...... 129 4.9.2 Gibbsfreeenergy ...... 129 4.9.3 Otherproperties ...... 130 4.9.4 Relation between mixture and partial molar properties ...... 132 4.10 Irreversible in a closed ...... 133 4.11 Equilibrium in a two-component system ...... 136 4.11.1 equilibrium ...... 136 4.11.2 : introduction ...... 139 4.11.2.1 Isothermal, isochoric system ...... 140 4.11.2.2 Isothermal, isobaric system ...... 145 4.11.3 Equilibrium condition ...... 147

5 of a single reaction 149 5.1 Molecularmass ...... 149 5.2 ...... 151 5.2.1 Generaldevelopment ...... 151 5.2.2 -airmixtures ...... 159 5.3 First law analysis of reacting ...... 161 5.3.1 Enthalpyofformation ...... 161 5.3.2 and internal of combustion ...... 163 5.3.3 Adiabatic flame in isochoric stoichiometric systems . . . 164

CC BY-NC-ND. 28 January 2019, J. M. Powers. CONTENTS 5

5.3.3.1 Undiluted, cold mixture ...... 164 5.3.3.2 Dilute, cold mixture ...... 165 5.3.3.3 Dilute, preheated mixture ...... 167 5.3.3.4 Dilute, preheated mixture with minor species ...... 168 5.4 Chemical equilibrium ...... 169 5.5 Chemical kinetics of a single isothermal reaction ...... 175 5.5.1 Isochoricsystems ...... 175 5.5.2 Isobaricsystems...... 183 5.6 Some conservation and evolution equations ...... 189 5.6.1 Total conservation: isochoric reaction ...... 189 5.6.2 Element mass conservation: isochoric reaction ...... 190 5.6.3 : adiabatic, isochoric reaction ...... 191 5.6.4 Energy conservation: adiabatic, isobaric reaction ...... 193 5.6.5 Entropy evolution: Clausius-Duhem relation ...... 196 5.7 Simple one-step kinetics ...... 199

6 Thermochemistry of multiple reactions 203 6.1 Summary of multiple reaction extensions ...... 203 6.2 Equilibrium conditions ...... 210 6.2.1 Minimization of G via Lagrange multipliers ...... 210 6.2.2 Equilibration of all reactions ...... 216 6.3 Concise reaction rate law formulations ...... 218 6.3.1 Reaction dominant: J (N L)...... 219 ≥ − 6.3.2 Species dominant: J < (N L)...... 220 − 6.4 Adiabatic, isochoric kinetics ...... 220

7 Kinetics in some more detail 223 7.1 Isothermal, isochoric kinetics ...... 226 7.1.1 O O dissociation...... 226 − 2 7.1.1.1 Pair of irreversible reactions ...... 226 7.1.1.1.1 ...... 226 7.1.1.1.2 Example calculation ...... 232 7.1.1.1.2.1 Species concentration versus time ...... 233 7.1.1.1.2.2 versus time ...... 234 7.1.1.1.2.3 form ...... 236 7.1.1.1.3 Effectoftemperature...... 240 7.1.1.2 Single reversible reaction ...... 241 7.1.1.2.1 Mathematical model ...... 241 7.1.1.2.1.1 Kinetics ...... 241 7.1.1.2.1.2 Thermodynamics ...... 243 7.1.1.2.2 Example calculation ...... 244 7.1.2 Zel’dovich mechanism of NO production ...... 247

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6 CONTENTS

7.1.2.1 Mathematicalmodel ...... 247 7.1.2.1.1 Standardmodelform...... 248 7.1.2.1.2 Reducedform...... 249 7.1.2.1.3 Example calculation ...... 252 7.1.2.2 Stiffness, time scales, and numerics ...... 258 7.1.2.2.1 Effectoftemperature...... 258 7.1.2.2.2 Effect of initial pressure ...... 261 7.1.2.2.3 Stiffness and numerics ...... 263 7.2 Adiabatic, isochoric kinetics ...... 264 7.2.1 Thermalexplosiontheory ...... 265 7.2.1.1 One-step reversible kinetics ...... 265 7.2.1.2 First law of thermodynamics ...... 266 7.2.1.3 Dimensionless form ...... 269 7.2.1.4 Example calculation ...... 270 7.2.1.5 High activation energy asymptotics ...... 272 7.2.2 Detailed H O N kinetics ...... 276 2 − 2 − 2 Bibliography 281

CC BY-NC-ND. 28 January 2019, J. M. Powers. Preface

These are lecture notes for AME 50531 Intermediate Thermodynamics (AME 54531 for students in our London Programme), the second of two undergraduate courses in thermody- namics taught in the Department of Aerospace and at the Univer- sity of Notre Dame. Most of the students in this course are juniors or seniors majoring in aerospace or mechanical engineering. The objective of the course is to survey both practical and theoretical problems in classical thermodynamics. The notes draw heavily on the text specified for the course, Borgnakke and Sonntag’s (BS) Fundamentals of Thermodynamics, Eighth Edition, John Wiley, New York, 2013, es- pecially Chapters 8-14. In general the nomenclature of BS is used, and much of the notes follow a similar structure as that text. In addition, Abbott and van Ness’s Thermodynamics, McGraw-Hill, New York, 1972, has been used to guide some of the mathematical develop- ments. Many example problems have been directly taken from BS and other texts; specific citations are given where they arise. Many of the unique aspects of these notes, which focus on thermodynamics of reactive with detailed finite rate kinetics, have arisen due to the author’s involvement in research supported by the National Science Foundation under Grant No. CBET-0650843. The author is grateful for the support. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author and do not necessarily reflect the views of the National Science Foundation. These notes emphasize both problem-solving as well as some rigorous undergraduate-level development of the underlying classical theory. It should also be remembered that practice is essential to the learning process; the student would do well to apply the techniques presented here by working as many problems as possible. The notes, along with information on the course, can be found on the world wide web at https://www3.nd.edu/ powers/ame.50531. At this stage, anyone is free to make copies for their own use. I would∼ be happy to hear from you about suggestions for improvement. Joseph M. Powers [email protected] https://www3.nd.edu/ powers ∼ Notre Dame, Indiana; USA CC BY: $\ = 28 January 2019 The content of this book is licensed under Creative Commons Attribution-Noncommercial-No Derivative Works 3.0.

7 8 CONTENTS

CC BY-NC-ND. 28 January 2019, J. M. Powers. Chapter 1

Review

Review BS, Chapters 1-7.

Here is a brief review of some standard concepts from a first course in thermodynamics. property: characterizes the thermodynamics state of the system • – extensive: proportional to system’s extent, upper case variables V : total , SI-based units are m3, ∗ U: total , SI-based units are kJ, ∗ H: total enthalpy, SI-based units are kJ, ∗ S: total entropy, SI-based units are kJ/K. ∗ – intensive: independent of system’s extent, lower case variables (exceptions are temperature T and pressure P , which are intensive). Intensive properties can be on a mass basis or a molar basis: v: specific volume, SI-based units are m3/kg, ∗ u: specific internal energy, SI-based units are kJ/kg, ∗ h: specific enthalpy, SI-based units are kJ/kg, ∗ s: specific entropy, SI-based units are kJ/kg/K. ∗ v: molar specific volume, SI-based units are m3/kmole, ∗ u: molar specific internal energy, SI-based units are kJ/kmole, ∗ h: molar specific enthalpy, SI-based units are kJ/kmole, ∗ s: molar specific entropy, SI-based units are kJ/kmole/K. ∗ : ρ =1/v, mass per unit volume, SI-based units are kg/m3. • equations of state: relate properties • – Calorically Perfect (CPIG) has P v = RT and u u = − o c (T T ), v − o 9 10 CHAPTER 1. REVIEW

– Calorically Imperfect Ideal Gas (CIIG) has P v = RT and u uo = T − c (Tˆ) dTˆ. To v – RNon-ideal state equations have P = P (T, v), u = u(T, v). Any intensive thermodynamic property can be expressed as a function of at most two • other intensive thermodynamic properties (for simple compressible systems).

– P = RT/v: thermal for ideal gas, SI-based units are kP a.

– c = √kP v: sound speed for CPIG, SI-based units are m/s; k = cP /cv, the ratio of specific .

energy: E = U + (1/2)m(v v)+ mgz, total energy = internal+kinetic+potential. • · first law: dE = δQ δW ; if kinetic and potential are ignored, dU = δQ δW ; with • − − δQ = TdS and δW = PdV , we get the Gibbs equation dU = TdS PdV . − second law: dS δQ/T . • ≥ process: moving from one state to another, in general with accompanying trans- • fer and .

cycle: process which returns to initial state. • specific reversible work: w = 2 P dv; δw = P dv. • 1 2 1 specific reversible :R q = 2 Tds; δq = Tds. • 1 2 1 Figure 1.1 gives an example of an isothermal thermodynamicR process going from state 1 to state 2 in various thermodynamic planes. Figure 1.2 gives a sketch of a .

Example 1.1 Consider an ideal gas in the T s plane. Compare the slope of an isochore to that of an isobar at a given point. −

Recall the Gibbs equation for a simple compressible substance:

T ds = du + P dv. (1.1)

We have for the ideal gas

du = cvdT, if ideal gas. (1.2)

This holds for all ideal gases, be they calorically perfect or imperfect. Thus, the Gibbs equation can be rewritten as

T ds = cvdT +P dv. (1.3) =du | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. 11

P T

v1

1 v2 P1 1 T = T 2 1 2 2 P 2 T1 = T2

v s s s v1 2 v 1 2 2 2 w = ∫ P dv u - u = q - w q = ∫ T ds 12 1 2 1 12 12 12 1 P v v1 P2 P 1 1 v2 2 v2

P1

P2 2 v 1 1

T = T 1 2 T = T T T 1 2

Figure 1.1: Sketch of isothermal .

On an isochore, v is constant, so dv = 0. So on an isochore, we have

T ds = cvdT, on an isochore. (1.4) or, using the partial derivative notation,

∂T T = . (1.5) ∂s c v v

Next recall the definition of enthalpy, h: h = u + P v. (1.6)

We can differentiate Eq. (1.6) to get

dh = du + P dv + vdP. (1.7)

Substitute Eq. (1.7) into the Gibbs equation, (1.1), to eliminate du in favor of dh to get

T ds = dh P dv vdP +P dv, (1.8) − − =du = dh vdP. (1.9) | − {z }

CC BY-NC-ND. 28 January 2019, J. M. Powers. 12 CHAPTER 1. REVIEW

P T

s v

q cycle = w cycle

Figure 1.2: Sketch of thermodynamic cycle.

For an ideal gas, dh = cP dT , where cP is at most a function of T , so T ds = c dT vdP. (1.10) P − =dh Now for the isobar, dP = 0. Thus on an isobar,| {z } we have

T ds = cP dT, on an isobar. (1.11) And so the slope on an isobar in the T s plane is − ∂T T = . (1.12) ∂s c P P

Since cP > cv, (recall for an ideal gas that cP (T ) cv(T )= R > 0), we can say that the slope of the isochore is steeper than the isobar in the T s plane.− −

Example 1.2 Consider the following for air, modeled as a CPIG, from state 1 to state 2. P1 = 100 kPa, T1 = 300 K, T2 = 400 K. Show the second law is satisfied.

Since the process is isobaric, P = 100 kPa describes a straight line in the P v and P T planes and P = P = 100 kPa. Since we have an ideal gas, we have for the v T plane:− − 2 1 − R v = T, straight lines! (1.13) P  

kJ 0.287 (300 K) 3 RT1 kg K m v1 = = =0.861 , (1.14) P1  100 kPa kg kJ 0.287 (400 K) 3 RT2 kg K m v2 = = =1.148 . (1.15) P2  100 kPa kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 13

Since the gas is ideal:

du = cvdT. (1.16)

Since our ideal gas is also calorically perfect, cv is a constant, and we get

u2 T2 du = cv dT , (1.17) Zu1 ZT1 u u = c (T T ), (1.18) 2 − 1 v 2 − 1 kJ = 0.7175 (400 K 300 K), (1.19) kg K −   kJ = 71.750 . (1.20) kg Also we have T ds = du + P dv, (1.21)

T ds = cvdT + P dv, (1.22) RT R RT from ideal gas : v = : dv = dT dP, (1.23) P P − P 2 RT T ds = c dT + RdT dP, (1.24) v − P dT dP ds = (c + R) R , (1.25) v T − P dT dP = (c + c c ) R , (1.26) v P − v T − P dT dP = c R , (1.27) P T − P s2 T2 dT P2 dP ds = c R , (1.28) P T − P Zs1 ZT1 ZP1 T P s s = c ln 2 R ln 2 . (1.29) 2 − 1 P T − P  1   1  (1.30) Now since P is a constant, T s s = c ln 2 , (1.31) 2 − 1 P T  1  kJ 400 K = 1.0045 ln , (1.32) kg K 300 K     kJ = 0.2890 . (1.33) kg K Then one finds v2 v2 1w2 = P dv = P dv, (1.34) Zv1 Zv1 = P (v v ) , (1.35) 2 − 1 m3 m3 = (100 kPa) 1.148 0.861 , (1.36) kg − kg   kJ = 28.700 . (1.37) kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 14 CHAPTER 1. REVIEW

Now

du = δq δw, (1.38) − δq = du + δw, (1.39) 2 2 2 δq = du + δw, (1.40) Z1 Z1 Z1 q = (u u )+ w , (1.41) 1 2 2 − 1 1 2 kJ kJ q = 71.750 + 28.700 , (1.42) 1 2 kg kg kJ q = 100.450 . (1.43) 1 2 kg Now in this process the gas is heated from 300 K to 400 K. One would expect at a minimum that the surroundings were at 400 K. Check for second law satisfaction.

1q2 s2 s1 ? (1.44) − ≥ Tsurr kJ 100.450 kJ/kg 0.2890 ? (1.45) kg K ≥ 400 K

kJ kJ 0.2890 0.2511 , yes. (1.46) kg K ≥ kg K

CC BY-NC-ND. 28 January 2019, J. M. Powers. 15

P T T = 300 K 1 v T2 = 400 K 1

1 2 v P = P = 100 kPa 2 1 2 T2 2 1

T1

v s1 s s v1 2 v 2 2 2 w = ∫ P dv u - u = q - w q = ∫ T ds 12 1 2 1 12 12 12 1 P v v1 v 2 P = P = 100 kPa 1 2 1 2 P = P = 100 kPa v 1 2 2 2

v 1 1

T T T T T 1 2 T 1 2

Figure 1.3: Sketch for isobaric example problem.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 16 CHAPTER 1. REVIEW

CC BY-NC-ND. 28 January 2019, J. M. Powers. Chapter 2

Cycle analysis

Read BS, Chapters 5, 8, 9, 10. 2.1 Carnot

The is the most well-known thermodynamic cycle. It is a useful idealization, but is difficult to realize in practice. Its real value lies in serving as a standard to which other cycles can be compared. These are usually fully covered in introductory courses. The cycle can be considered as follows 1 2: isentropic compression, • → 2 3: isothermal expansion, • → 3 4: isentropic expansion, and • → 4 1: isothermal compression. • → This forms a rectangle in the T s plane The P v plane is more complicated. Both are shown in Figure 2.1. − −

2.1.1 Analysis for a calorically perfect ideal gas For this discussion, consider a calorically perfect ideal gas. The isotherms are then straight- forward and are hyperbolas described by 1 P = RT . (2.1) v The slope of the isotherms in the P v plane is found by differentiation: − ∂P 1 = RT , (2.2) ∂v − v2 T P = . (2.3) − v 17 18 CHAPTER 2. CYCLE ANALYSIS

P T 2

i s o t h isotherm e 3 r 2 m

i s e n t r o p isentrope e 3 isentrope

is is ot e he n 1 rm t ro 1 isotherm 4 pe 4

v s

wcycle = ∫ P dv = q cycle = ∫ T ds

Figure 2.1: Sketch of P v and T s planes for a CPIG in a Carnot cycle. − −

The slope of the isentrope is found in the following way. Consider first the Gibbs equation, Eq. (1.1): Tds = du + P dv. (2.4)

Because the gas is calorically perfect, one has

du = cvdT, (2.5)

so the Gibbs equation, Eq. (2.4), becomes

Tds = cvdT + P dv. (2.6)

Now for the ideal gas, one has

P v = RT, (2.7) P dv + vdP = RdT, (2.8) P dv + vdP = dT. (2.9) R

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.1. CARNOT 19

So then P dv + vdP Tds = c +P dv, (2.10) v R   =dT c c Tds = v +1 P dv + v vdP, Take ds =0, (2.11) R| {z }R c  c ∂P 0 = v +1 P + v v , (2.12) R R ∂v s  cv  ∂P +1 P = R , (2.13) ∂v − cv v s R cv +1 cP −cv P = , (2.14) − cv v cP −cv c + c c P = v P − v , (2.15) − cv v c P = P , (2.16) − cv v P = k . (2.17) − v Since k > 1, the magnitude of the slope of the isentrope is greater than the magnitude of the slope of the isotherm: ∂P ∂P > . (2.18) ∂v ∂v s T

Example 2.1 (Adapted from BS, 7.94, p. 274) Consider an ideal gas Carnot cycle with air in a piston cylinder with a high temperature of 1200 K and a heat rejection at 400 K. During the heat addition, the volume triples. The gas is at 1.00 m3/kg before the isentropic compression. Analyze.

Take state 1 to be the state before the compression. Then

m3 T = 400 K, v =1.00 . (2.19) 1 1 kg

By the

0.287 kJ (400 K) RT1 kg K 2 P = = 3 =1.148 10 kP a. (2.20) 1 v  m × 1 1.00 kg

Now isentropically compress to state 2. By the standard relations for a CPIG, one finds

−1 k−1 k T v P k 2 = 1 = 2 . (2.21) T v P 1  2   1 

CC BY-NC-ND. 28 January 2019, J. M. Powers. 20 CHAPTER 2. CYCLE ANALYSIS

So 1 1 T k−1 m3 400 K 1.4−1 m3 v = v 1 = 1.00 =0.06415 . (2.22) 2 1 T kg 1200 K kg  2      Note that v2 < v1 as is typical in a compression. The isentropic relation between pressure and volume can be rearranged to give the standard

k k P2v2 = P1v1 . (2.23)

Thus, one finds that

3 1.4 k 1.00 m v1 2 kg 3 P2 = P1 = (1.148 10 kPa) m3 =5.36866 10 kP a. (2.24) v2 × 0.06415 ×   kg ! Note the pressure has increased in the isentropic compression. Check to see if the ideal gas law is satisfied at state 2:

0.287 kJ (1200 K) RT2 kg K 3 P = = 3 =5.36866 10 kP a. (2.25) 2 v   m × 2 0.06415 kg

This matches. Now the expansion from state 2 to 3 is isothermal. This is the heat addition step in which the volume triples. So one gets

m3 m3 v =3v =3 0.06415 =0.19245 . (2.26) 3 2 kg kg  

T3 = T2 = 1200 K. (2.27) The ideal gas law then gives

0.287 kJ (1200 K) RT3 kg K 3 P = = 3 =1.78955 10 kP a. (2.28) 3 v   m × 3 0.19245 kg

Process 3 to 4 is an isentropic expansion back to 400 K. Using the isentropic relations for the CPIG, one gets 1 1 T k−1 m3 1200 K 1.4−1 m3 v = v 3 = 0.19245 =3.000 . (2.29) 4 3 T kg 400 K kg  4      3 1.4 k 0.19245 m v3 3 kg 1 P4 = P3 = (1.78955 10 kPa) m3 =3.82667 10 kP a. (2.30) v4 × 3.000 ×   kg ! Check: 0.287 kJ (400 K) RT4 kg K 1 P = = 3 =3.82667 10 kP a. (2.31) 4 v  m × 4 3.000 kg A summary of the states is given in Table 2.1. Now calculate the work, heat transfer and efficiency. Take the adiabatic exponent for air to be k =1.4. Now since cP k = , cP cv = R, (2.32) cv −

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.1. CARNOT 21

m3 T (K) P (kP a) v ( kg ) 1 400 1.148 102 1.000 2 1200 5.36886 × 103 0.06415 × 3 1200 1.78955 103 0.19245 4 400 3.82667 × 102 3.000 × Table 2.1: State properties for Carnot cycle.

one gets

R + c k = v , (2.33) cv kcv = R + cv, (2.34)

cv(k 1) = R, (2.35) − kJ R 0.287 kJ c = = kg K =0.7175 . (2.36) v k 1 1.4 1 kg K − −

Recall the first law:

u u = q w . (2.37) 2 − 1 1 2 − 1 2

Recall also the caloric equation of state for a CPIG:

u u = c (T T ). (2.38) 2 − 1 v 2 − 1

Now process 1 2 is isentropic, so it is also adiabatic, hence q = 0, so one has → 1 2

u u = q w , (2.39) 2 − 1 1 2 −1 2 =0 c (T T ) = w , (2.40) v 2 − 1 −|{z}1 2 kJ 0.7175 (1200 K 400 K) = w , (2.41) kg K − −1 2   kJ w = 5.7400 102 . (2.42) 1 2 − × kg

The work is negative as work is being done on the system in the compression process.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 22 CHAPTER 2. CYCLE ANALYSIS

Process 2 3 is isothermal, so there is no internal energy change. The first law gives → u u = q w , (2.43) 3 − 2 2 3 − 2 3 c (T T ) = q w , (2.44) v 3 − 2 2 3 − 2 3 =0 0 = q w , (2.45) | {z } 2 3 2 3 − v3 2q3 = 2w3 = P dv, (2.46) Zv2 v3 RT = dv, (2.47) v Zv2 v3 dv = RT , (2.48) 2 v Zv2 v3 = RT2 ln , (2.49) v2 3 m kJ 0.19245 kg = 0.287 (1200 K) ln 3 , (2.50) kg K 0.06415 m   kg kJ = 3.78362 102 . (2.51) × kg

The work is positive, which is characteristic of the expansion process. Process 3 4 is adiabatic so q = 0. The first law analysis gives then → 3 4 u u = q w , (2.52) 4 − 3 3 4 −3 4 =0 c (T T ) = w , (2.53) v 4 − 3 −|{z}3 4 kJ 0.7175 (400 K 1200 K) = w , (2.54) kg K − −3 4   kJ w = 5.74000 102 . (2.55) 3 4 × kg

Process 4 1 is isothermal. Similar to the other , one finds → u u = q w , (2.56) 1 − 4 4 1 − 4 1 c (T T ) = q w , (2.57) v 1 − 4 4 1 − 4 1 =0 0 = q w , (2.58) | {z } 4 1 4 1 − v1 4q1 = 4w1 = P dv, (2.59) Zv4 v1 RT = dv, (2.60) v Zv4 v1 = RT4 ln , (2.61) v4 3 m kJ 1.0000 kg = 0.287 (400 K) ln 3 , (2.62) kg K 3.0000 m   kg kJ = 1.26121 102 . (2.63) − × kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.1. CARNOT 23

kJ kJ kJ Process ∆u kg q kg w kg 1 2 5.74000  102  0 5.74000  102 → × − × 2 3 0 3.78362 102 3.78362 102 3 → 4 5.74000 102 × 0 5.74000 × 102 → − × × 4 1 0 1.26121 102 1.26121 102 → − × − × Total 0 2.52241 102 2.52241 102 × × Table 2.2: First law parameters for Carnot cycle.

Table 2.2 summarizes the first law considerations. The cycle work is found by adding the work of each individual process:

wcycle = 1w2 + 2w3 + 3w4 + 4w1, (2.64) kJ = ( 5.74+3.78362+ 5.74 1.26121) 102 =2.52241 102 . (2.65) − − × × kg

The cycle heat transfer is

qcycle = 1q2 + 2q3 + 3q4 + 4q1, (2.66) kJ = (0+3.78362+ 0 1.26121) 102 =2.52241 102 . (2.67) − × × kg

Note that

wcycle = qcycle. (2.68) Check now for the cycle efficiency. Recall that the thermal efficiency is

what you want w η = = cycle . (2.69) what you pay for qin

Here this reduces to kJ w 2.52241 102 cycle × kg η = = kJ =0.6667. (2.70) 2q3 3.78362 102 × kg Recall that the efficiency for any Carnot cycle should be

T η =1 low . (2.71) − Thigh

So general Carnot theory holds that the efficiency should be

400 K 2 η =1 = 0.6667. (2.72) − 1200 K 3 ∼ Recall further that for a Carnot cycle, one has

q T low = low . (2.73) qhigh Thigh

CC BY-NC-ND. 28 January 2019, J. M. Powers. 24 CHAPTER 2. CYCLE ANALYSIS

For this problem then one has q T −4 1 = low , (2.74) 2q3 Thigh kJ 1.26121 102 400 K × kg = , (2.75) 3.78362 102 kJ 1200 K × kg 1 1 = . (2.76) 3 3 Indeed, the appropriate relation holds.

2.1.2 Exergy Here we will introduce the concept of exergy, which relies on a Carnot cycle for its motivation. This concept is widely used in some industrial design applications; its use in the fundamental physics literature is not extensive, likely because it is not a thermodynamic property of a material, but also includes mechanical properties. This system quantity is one measure of how much useful work can be extracted from a system which is brought into equilibrium with a so-called reference rest state. First let us imagine that the surroundings are at a reference temperature of To and a reference pressure of Po. This yields a reference enthalpy of ho and a reference entropy of so. We also take the surroundings to be at rest with a velocity of v = 0, and a reference height of zo. Consider the sketch of Fig. 2.2. Let us consider a steady flow into and out of a . The equation yields dm cv =m ˙ m˙ . (2.77) dt in − out For steady flow conditions, we have d/dt = 0, so we recover

m˙ in =m ˙ out =m. ˙ (2.78) The first law of thermodynamics for the control volume yields dE 1 cv = Q˙ W˙ +m ˙ h + v v + gz dt cv − cv in in 2 in · in in   1 m˙ h + v v + gz . (2.79) − out out 2 out · out out   We will soon relate Q˙ cv/m˙ to qH,Carnot as sketched in Fig. 2.2; this will introduce a sign convention problem. For now, we will maintain the formal sign convention associated with Q˙ cv connoting heat transfer into the control volume. For steady flow conditions, we get 1 0= Q˙ W˙ +m ˙ h h + (v v v v )+ g(z z ) . (2.80) cv − cv in − out 2 in · in − out · out in − out   CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.1. CARNOT 25

control volume h + (1/2) v . v + g z ho + gz o

qH,Carnot

Carnot w engine Carnot

q L,Carnot

To

Figure 2.2: Sketch of control volume balance for exergy discussion.

Now take the “in” state to be simply a generic state with no subscript, and the “out” state to be the reference state, so

1 0= Q˙ W˙ +m ˙ h h + (v v)+ g(z z ) . (2.81) cv − cv − o 2 · − o   Let us next scale bym ˙ so as to get 1 0= q w + h h + v v + g(z z ). (2.82) cv − cv − o 2 · − o Here we have taken q Q˙ /m˙ and w = W˙ /m˙ . cv ≡ cv cv cv Now, let us insist that wcv = 0, and solve for qcv to get 1 q = h h + v v + g(z z ) . (2.83) cv − − o 2 · − o   Now, we imagine the working fluid to be at an elevated enthalpy, velocity, and height relative to its rest state. Thus in the process of bringing it to its rest state, we will induce qcv < 0. By our standard sign convention, this means that is leaving the system. This energy which leaves the system can be harnessed, in the best of all possible worlds, by a Carnot engine in contact with a thermal reservoir at temperature To to generate useful work. It is this work which represents the available energy, also known as the exergy. For the Carnot engine, we have w = q q . (2.84) Carnot H,Carnot − L,Carnot

CC BY-NC-ND. 28 January 2019, J. M. Powers. 26 CHAPTER 2. CYCLE ANALYSIS

Note the standard sign convention for work and heat transfer is abandoned for the Carnot analysis! It is wCarnot which gives the availability or exergy, which we define as ψ: ψ = q q . (2.85) H,Carnot − L,Carnot Now let us require that the heat input to the Carnot engine be the heat associated with the heat transfer from the control volume. Because of the inconsistency in sign conventions, this requires that q = q . (2.86) cv − H,Carnot Now for Carnot cycles, we know that

q = T (s s ), (2.87) H,Carnot − o q = T (s s ). (2.88) L,Carnot o − o So, by substituting Eqs. (2.86) and (2.88) into Eq. (2.85), the availability ψ is

ψ = q T (s s ). (2.89) − cv − o − o

Now use Eq. (2.83) to eliminate qcv from Eq. (2.89) to get 1 ψ = h h + v v + g(z z ) T (s s ), (2.90) − o 2 · − o − o − o  

1 ψ = h T s + v v + gz (h T s + gz ) . (2.91) − o 2 · − o − o o o   So the exergy, that is, the ability to useful work, is enhanced by high enthalpy h, which for ideal gases implies high temperature T , • high velocity v, • high height z, and • low entropy (or high order or structure) s. •

Example 2.2 Find the exergy for a CPIG.

For a CPIG, we have

h h = c (T T ), (2.92) − o P − o T P s s = c ln R ln . (2.93) − o P T − P  o   o 

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.1. CARNOT 27

So we get

T P 1 ψ = c (T T ) T c ln R ln + v v + g(z z ). (2.94) P − o − o P T − P 2 · − o   o   o 

For T To P Po, we can use Taylor series to simplify this somewhat. First, we recall the general Taylor∼ series for∼ a log function near unity. Consider

y(x) = ln x, (2.95)

for x 1. For a Taylor series near x = 1, we have ∼ dy 1 d2y y(x) y(1) + (x 1)+ (x 1)2 + .... (2.96) ∼ dx − 2 dx2 − x=1 x=1

Now for y = ln x, we have

dy 1 d2y 1 = , = , (2.97) dx x dx2 −x2

and thus

dy d2y y(1) = ln(1) = 0, =1, = 1 (2.98) dx dx2 − x=1 x=1

So 1 y(x) = ln x 0 + (x 1) (x 1)2 + ... (2.99) ∼ − − 2 −

We then expand ψ via the following steps:

T P 1 ψ = c (T T ) c T ln + RT ln + v v + g(z z ), (2.100) P − o − P o T o P 2 · − o  o   o  T T k 1 P 1 = c T 1 ln + − ln + v v + g(z z ), (2.101) P o T − − T k P 2 · − o  o   o   o  T T 1 T 2 k 1 P = c T 1 1 1 + ... + − 1+ ... P o T − − T − − 2 T − k P −  o   o   o  !  o ! 1 + v v + g(z z ), (2.102) 2 · − o 1 T 2 k 1 P 1 = c T 1 + ... + − 1+ ... + v v + g(z z ). (2.103) P o 2 T − k P − 2 · − o  o  !  o ! Note that in the neighborhood of the ambient state, relative pressure differences are more effective than relative temperature differences at inducing high exergy.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 28 CHAPTER 2. CYCLE ANALYSIS

2.2 Rankine

2.2.1 Classical The forms the foundation for the bulk of power generating devices which utilize as a working fluid. The ideal cycle is described by

1 2: isentropic pumping process in the , • → 2 3: isobaric heat transfer in the , • → 3 4: isentropic expansion in the , and • → 4 1: isobaric heat transfer in the condenser. • → Note that to increase cycle efficiency one can

lower the condenser pressure (increases in turbine), • superheat the steam, or • increase the pressure during heat addition. • A schematic for the Rankine cycle and the associated path in the T s plane is shown in Figure 2.3. −

3 T

3 turbine boiler

2 2 4 1 4

pump condenser 1 s

Figure 2.3: Schematic for Rankine cycle and the associated T s plane. −

Example 2.3 (adopted from Moran and Shapiro, p. 312) Consider steam in an ideal Rankine cycle. Saturated vapor enters the turbine at 8.0 MPa. Saturated liquid exits the condenser at P = 0.008 MPa. The net power output of the cycle is 100 MW . Find thermal efficiency •

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.2. RANKINE 29

back work ratio • mass flow rate of steam • rate of heat transfer Q˙ to the fluid in the boiler • in rate of heat transfer Q˙ in the condenser • out mass flow rate of condenser cooling water if the cooling water enters at 15 ◦C and exits at 35 ◦C. • Use the steam tables to fix the state. At the turbine inlet, one has P3 =8.0 MPa, and x3 = 1 (saturated steam). This gives two properties to fix the state, so that kJ kJ h = 2758 , s =5.7432 . (2.104) 3 kg 3 kg K

State 4 has P4 =0.008 MPa and s4 = s3 =5.7432 kJ/kg/K, so the state is fixed. From the saturation tables, it is found then that

5.7432 kJ 0.5926 kJ s4 sf kg K − kg K x4 = − = kJ =0.6745. (2.105) sg sf  7.6361   − kg K Note the quality is 0 x 1, as it must be. The enthalpy is then ≤ 4 ≤ kJ kJ kJ h = h + x h = 173.88 + (0.6745) 2403.1 = 1794.8 . (2.106) 4 f 4 fg kg kg kg     State 1 is saturated liquid at 0.008 MPa, so x1 = 0, P1 = 0.008 MPa. One then gets h1 = hf = 3 173.88 kJ/kg, v1 = vf =0.0010084 m /kg. Now state 2 is fixed by the boiler pressure and s2 = s1. But this requires use of the sparse compressed liquid tables. Alternatively, the pump work is easily approximated by assuming an incom- pressible fluid so that W˙ h = h + = h + v (P P ), (2.107) 2 1 m˙ 1 1 2 − 1 kJ m3 kJ h = 173.88 + 0.0010084 (8000 kPa 8 kPa) = 181.94 . (2.108) 2 kg kg − kg     The net power is W˙ cycle = W˙ t + W˙ p. (2.109) Now the first law for the turbine and pump give W˙ W˙ t = h h , p = h h . (2.110) m˙ 3 − 4 m˙ 1 − 2 The energy input that is paid for is Q˙ in = h h . (2.111) m˙ 3 − 2 The thermal efficiency is then found by W˙ + W˙ (h h ) + (h h ) η = t p = 3 − 4 1 − 2 , (2.112) Q˙ h3 h2 in − kJ kJ kJ kJ 2758 kg 1794.8 kg + 173.88 kg 181.94 kg = − − , (2.113)    2758 kJ  181.94 kJ    kg − kg = 0.371.     (2.114)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 30 CHAPTER 2. CYCLE ANALYSIS

By definition the back work ratio bwr is the ratio of pump work to turbine work:

W˙ bwr = p , (2.115) ˙ Wt

h h = 1 − 2 , (2.116) h h 3 − 4 kJ kJ 173.88 kg 181.94 kg = − , (2.117)  kJ   kJ  2758 1794.8 kg − kg

= 0.00837.    (2.118)

The desired mass flow can be determined since we know the desired net power. Thus

W˙ m˙ = cycle , (2.119) (h h ) + (h h ) 3 − 4 1 − 2 100 103 kW = × , (2.120) 2758 kJ 1794.8 kJ + 173.88 kJ 181.94 kJ kg − kg kg − kg  kg       = 104.697 , (2.121) s kg s kg = 104.697 3600 = 3.769 105 . (2.122) s hr × hr    

The necessary heat transfer rate in the boiler is then

Q˙ =m ˙ (h h ), (2.123) in 3 − 2 kg kJ kJ = 104.697 2758 181.94 , (2.124) s kg − kg       = 269706 kW , (2.125) = 269.7 MW. (2.126)

In the condenser, one finds

Q˙ =m ˙ (h h ), (2.127) out 1 − 4 kg kJ kJ = 104.697 173.88 1794.8 , (2.128) s kg − kg       = 169705 kW, (2.129) − = 169.7 MW. (2.130) −

Note also for the cycle that one should find

W˙ = Q˙ + Q˙ = (269.7 MW ) (169.7 MW ) = 100 MW. (2.131) cycle in out −

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.2. RANKINE 31

For the condenser mass flow rate now perform a mass balance:

dE cv = Q˙ W˙ +m ˙ (h h ) +m ˙ (h h ), (2.132) dt cv − cv c in − out 4 − 1 =0 =0 =0 0 =m|{z} ˙ (h |{z}h ) +m ˙ (h h ), (2.133) | {z } c in − out 4 − 1 m˙ (h h ) m˙ = 4 − 1 , (2.134) c − h h in − out kg kJ kJ 104.697 s 1794.8 kg 173.88 kg = − , (2.135) − 62.99 kJ 146.68kJ  kg − kg kg    = 2027.79 , (2.136) s kg 3600 s = 2027.79 , (2.137) s hr     kg = 7.3 106 . (2.138) × hr

The enthalpy for the cooling water was found by assuming values at the saturated state at the respective of 15 ◦C and 35 ◦C.

Example 2.4 Compute the exergy at various points in the flow of a Rankine cycle as considered in the previous example problem. For that example, we had at state 1, the pump inlet that

kJ kJ h = 173.88 , s =0.5926 . (2.139) 1 kg 1 kg K

After the pump, at state 2, we have

kJ kJ h = 181.94 , s =0.5926 . (2.140) 2 kg 2 kg K

After the boiler, at state 3, we have

kJ kJ h = 2758 , s =5.7432 . (2.141) 3 kg 3 kg K

After the turbine, at state 4, we have

kJ kJ h = 1794.8 , s =5.7432 . (2.142) 4 kg 4 kg K

Now for this example, kinetic and contributions to the exergy are negligible, so we can say in general that

ψ = (h T s) (h T s ). (2.143) − o − o − o o

CC BY-NC-ND. 28 January 2019, J. M. Powers. 32 CHAPTER 2. CYCLE ANALYSIS

Now for this problem, we have To = 298.15 K, ho = 104.89 kJ/kg, so = 0.3674 kJ/kg/K. We have ◦ estimated ho and so as the enthalpy and entropy of a saturated liquid at 25 C = 298.15 K. So the exergies are as follows. At the pump entrance, we get ψ = (h T s ) (h T s ) , (2.144) 1 1 − o 1 − o − o o kJ kJ = 173.88 (298.15 K) 0.5926 kg − kg K     kJ kJ 104.89 (298.15 K) 0.3674 , (2.145) − kg − kg K     kJ = 1.84662 . (2.146) kg After the pump, just before the boiler, we have ψ = (h T s ) (h T s ) , (2.147) 2 2 − o 2 − o − o o kJ kJ = 181.94 (298.15 K) 0.5926 kg − kg K     kJ kJ 104.89 (298.15 K) 0.3674 , (2.148) − kg − kg K     kJ = 9.90662 . (2.149) kg So the exergy has gone up. Note here that ψ ψ = h h because the process is isentropic. Here 2 − 1 2 − 1 ψ2 ψ1 = h2 = h1 = 181.94 173.88=8.06 kJ/kg. −After the boiler, just before− the turbine, we have ψ = (h T s ) (h T s ) , (2.150) 3 3 − o 3 − o − o o kJ kJ = 2758 (298.15 K) 5.7432 kg − kg K     kJ kJ 104.89 (298.15 K) 0.3674 , (2.151) − kg − kg K     kJ = 1050.32 . (2.152) kg Relative to the pump, the boiler has added much more exergy to the fluid. After the turbine, just before the condenser, we have ψ = (h T s ) (h T s ) , (2.153) 4 4 − o 4 − o − o o kJ kJ = 1794.8 (298.15 K) 5.7432 kg − kg K     kJ kJ 104.89 (298.15 K) 0.3674 , (2.154) − kg − kg K     kJ = 87.1152 . (2.155) kg

Note that ψ4 ψ3 = h4 h3 because the process is isentropic. The actual exergy (or available work) at the exit of− the turbine− is relative low, even though the enthapy state at the turbine exit remains at an elevated value.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.2. RANKINE 33

2.2.2 Reheat In a Rankine cycle with reheat, the steam is extracted from an intermediate stage of the turbine and reheated in the boiler. It is then expanded through the turbine again to the condenser pressure. One also avoids liquid in the turbine with this strategy. This generally results in a gain in cycle efficiency. Geometrically, the behavior on a T s diagram looks more like a Carnot cycle. This is often covered in an introductory thermodynamics− class, so no formal example will be given here. A schematic for the Rankine cycle with reheat and the associated T s diagram is shown in Figure 2.4. − T 3 Turbine 3 5 Boiler 4

5 2 4 2 6 1 6

Pump Condenser 1 s

Figure 2.4: Schematic for Rankine cycle with reheat along with the relevant T s diagram. −

2.2.3 Regeneration In a Rankine cycle with regeneration, some steam is extracted from the turbine and used to pre-heat the liquid which is exiting the pump. This can lead to an increased thermal efficiency, all else being equal. The analysis is complicated by the need to take care of more complex mass and energy balances in some components. A schematic for the Rankine cycle with regeneration and open feedwater heating is shown in Figure 2.5.

Example 2.5 (BS, Ex. 11.4, pp. 438-440) Steam leaves boiler and enters turbine at 4 MPa, 400 ◦C. After expansion to 400 kPa, some stream is extracted for heating feedwater in an open . Pressure in feedwater heater is 400 kPa, and water leaves it at a saturated state at 400 kPa. The rest of the steam expands through the turbine to 10 kPa. Find the cycle efficiency. 1 2: compression through Pump P 1, • → 2&6 3: mixing in open feedwater heater to saturated liquid state, • → 3 4: compression through Pump P 2, • → 4 5: heating in boiler, • →

CC BY-NC-ND. 28 January 2019, J. M. Powers. 34 CHAPTER 2. CYCLE ANALYSIS

5

Turbine Boiler

7 6

4 Feedwater Heater Condenser

2 Pump P2 3 Pump 1 P1

Figure 2.5: Schematic for Rankine cycle with regeneration and open feedwater heating.

5 6: partial expansion in turbine, • → 5 7: completion of turbine expansion, and • → 7 1: cooling in condenser. • → From the tables, one can find

kJ kJ kJ kJ kJ h = 3213.6 , h = 2685.6 , h = 2144.1 , h = 191.8 , h = 604.73 , 5 kg 6 kg 7 kg 1 kg 3 kg

m3 m3 v =0.00101 , v =0.001084 . (2.156) 1 kg 3 kg

First consider the low pressure pump.

h = h + v (P P ), (2.157) 2 1 1 2 − 1 kJ m3 = 191.8 + 0.00101 ((400 kPa) (10 kPa)) , (2.158) kg kg −     kJ = 192.194 (2.159) kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.2. RANKINE 35

The pump work is w = v (P P ), (2.160) P 1 1 1 − 2 m3 = 0.00101 ((10 kPa) (400 kPa)) , (2.161) kg −   kJ = 0.3939 . (2.162) − kg Note, a sign convention consistent with work done by the fluid is used here. At this point the text abandons this sign convention instead. Now consider the turbine dm cv =m ˙ m˙ m˙ , (2.163) dt 5 − 6 − 7 =0 m˙ =m ˙ +m ˙ , (2.164) | {z }5 6 7 m˙ m˙ 1 = 6 + 7 , (2.165) m˙ 5 m˙ 5 dE cv = Q˙ W˙ +m ˙ h m˙ h m˙ h , (2.166) dt cv − cv 5 5 − 6 6 − 7 7 =0 =0 W˙ =|{z}m ˙ h m˙ h m˙ h . (2.167) | {zcv} 5 5 − 6 6 − 7 7 On a per mass basis, we get,

W˙ cv m˙ 6 m˙ 7 wt = = h5 h6 h7, (2.168) m˙ 5 − m˙ 5 − m˙ 5 m˙ m˙ = h 6 h 1 6 h , (2.169) 5 − m˙ 6 − − m˙ 7 5  5  m˙ m˙ = h h + h 6 h 1 6 h , (2.170) 5 − 6 6 − m˙ 6 − − m˙ 7 5  5  m˙ m˙ = h h + h 1 6 1 6 h , (2.171) 5 − 6 6 − m˙ − − m˙ 7  5   5  m˙ = h h + 1 6 (h h ). (2.172) 5 − 6 − m˙ 6 − 7  5  Now consider the feedwater heater. The first law for this device gives dE cv = Q˙ W˙ +m ˙ h +m ˙ h m˙ h , (2.173) dt cv − cv 2 2 6 6 − 3 3 =0 =0 =0 |{z}m˙ 2 |{z}m˙ 6 | {zh3} = h2 + h6, (2.174) m˙ 3 m˙ 3 m˙ 7 m˙ 6 = h2 + h6, (2.175) m˙ 5 m˙ 5 m˙ m˙ = 1 6 h + 6 h , (2.176) − m˙ 2 m˙ 6  5  5 kJ m˙ kJ m˙ kJ 604.73 = 1 6 192.194 + 6 2685.6 , (2.177) kg − m˙ kg m˙ kg    5    5   m˙ 6 = 0.165451. (2.178) m˙ 5

CC BY-NC-ND. 28 January 2019, J. M. Powers. 36 CHAPTER 2. CYCLE ANALYSIS

Now get the turbine work m˙ w = h h + 1 6 (h h ), (2.179) t 5 − 6 − m˙ 6 − 7  5  kJ kJ = 3213.6 2685.6 kg − kg     kJ kJ + (1 0.165451) 2685.6 2144.1 , (2.180) − kg − kg     kJ = 979.908 . (2.181) kg Now get the work for the high-pressure pump w = v (P P ), (2.182) P 2 3 3 − 4 m3 = 0.001084 ((400 kPa) (4000 kPa)) , (2.183) kg −   kJ = 3.9024 . (2.184) − kg Now h = h + v (P P ), (2.185) 4 3 3 4 − 3 kJ m3 = 604.73 + 0.001084 ((4000 kPa) (400 kPa)) , (2.186) kg kg −     kJ = 608.6 . (2.187) kg Now get the net work

W˙ net =m ˙ 5wt +m ˙ 1wP 1 +m ˙ 5wP 2, (2.188) m˙ 1 wnet = wt + wP 1 + wP 2, (2.189) m˙ 5 m˙ 7 = wt + wP 1 + wP 2, (2.190) m˙ 5 m˙ = w + 1 6 w + w , (2.191) t − m˙ P 1 P 2  5  kJ kJ kJ = 979.908 + (1 0.165451) 0.3939 3.9024 , (2.192) kg − − kg − kg       kJ = 975.677 . (2.193) kg Now for the heat transfer in the boiler, one has q = h h , (2.194) h 5 − 4 kJ kJ = 3213.6 608.6 , (2.195) kg − kg     kJ = 2605.0 . (2.196) kg Thus, the thermal efficiency is kJ w 975.677 η = net = kg = 0.375. (2.197) q kJ h 2605.0 kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.3. AIR STANDARD CYCLES 37

This does represent an increase of efficiency over a comparable Rankine cycle without regeneration, which happens to be 0.369.

2.2.4 Losses Turbine: These are typically the largest losses in the system. The turbine efficiency is • defined by w h h η = t = 3 − 4 . (2.198) w h h ts 3 − 4s Here h4s and wts are the enthalpy and work the working fluid would have achieved had the process been isentropic. Note this is for a control volume.

Pump: Pump losses are usually much smaller in magnitude than those for . • The pump efficiency for a control volume is defined by

w h h η = ps = 2s − 1 . (2.199) w h h p 2 − 1 Piping: pressure drops via viscous and turbulent flow effects induce entropy gains • in fluid flowing through pipes. There can also be heat transfer from pipes to the surroundings and vice versa.

Condenser: Losses are relatively small here. • Losses will always degrade the overall thermal efficiency of the cycle.

2.2.5 Cogeneration Often steam is extracted after the boiler for alternative uses. A good example is the Notre Dame power plant, where steam at high pressure and temperature is siphoned from the turbines to heat the campus in winter. The analysis for such a system is similar to that for a system with regeneration. A schematic for cogeneration cycle is shown in Figure 2.6.

2.3 Air standard cycles

It is useful to model several real engineering devices by what is known as an air standard cycle. This is based on the following assumptions:

CPIG, •

CC BY-NC-ND. 28 January 2019, J. M. Powers. 38 CHAPTER 2. CYCLE ANALYSIS

5

Steam High pressure Low pressure Generator turbine . . turbine W Q (e.g. ND power T H plant)

6 7

Thermal process 4 steam load (e.g. heating the Condenser ND campus) . Q L (e.g. to St. Joseph’s Lake) Liquid 1 Liquid 8

2 3 Pump Pump Mixer P2 P1

. . W P1 W P2

Figure 2.6: Schematic for Rankine cycle with cogeneration.

no inlet or exhaust stages, • combustion process replaced by heat transfer process, • cycle completed by heat transfer to surroundings (not exhaust), and • all process internally reversible. • In some cycles, it is common to model the working fluid as a fixed mass. In others, it is common to model the system as a control volume.

2.4 Brayton

2.4.1 Classical The is the air standard model for . It is most commonly modeled on a control volume basis. It has the following components:

1 2: isentropic compression, • →

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.4. BRAYTON 39

2 3: isobaric heat transfer to combustion chamber, • → 3 4: isentropic expansion through turbine, and • → 4 1: isobaric heat exchange with surroundings. • → A schematic for the Brayton cycle is shown in Figure 2.7. Diagrams for P v and T s for − − Fuel

Combustion Chamber 3 2

Compressor Turbine w

1 Air Products 4

environmental exhaust return

Figure 2.7: Schematic for Brayton cycle. the Brayton cycle are shown in Figure 2.8.

P T

2 3 3

isentrope isobar

2 isentrope 4 isobar 1 4 1

v s

Figure 2.8: P v and T s diagrams for the Brayton cycle. − − The efficiency of the air standard Brayton cycle is found as w η = net . (2.200) qH

CC BY-NC-ND. 28 January 2019, J. M. Powers. 40 CHAPTER 2. CYCLE ANALYSIS

For the cycle, the first law holds that

w = q = q q . (2.201) net net H − L One notes for an isobaric control volume combustor that dE cv = Q˙ W˙ +mh ˙ mh˙ . (2.202) dt cv − cv in − out =0 =0 |{z} So | {z }

Q˙ =m ˙ (h h ), (2.203) cv out − in Q˙ cv = h h , (2.204) m˙ out − in q = h h , (2.205) out − in q = c (T T ). (2.206) P out − in So the thermal efficiency is

T4 T1 1 w q q q c (T T ) T1 η = net = H − L =1 L =1 P 4 − 1 =1 − . (2.207) qH qH − qH − cP (T3 T2) −  T3  T2 1 − T2 −   Now because of the definition of the process, one also has

P P 3 = 2 . (2.208) P4 P1 And because 1 2 and 3 4 are isentropic, one has → → P T k/(k−1) P T k/(k−1) 2 = 2 = 3 = 3 . (2.209) P T P T 1  1  4  4  So one then has T T 2 = 3 . (2.210) T1 T4 Cross-multiplying, one finds T T 3 = 4 . (2.211) T2 T1 Subtracting unity from both sides gives

T T 3 1= 4 1. (2.212) T2 − T1 −

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.4. BRAYTON 41

So the thermal efficiency takes the form T 1 1 η =1 1 =1 =1 . (2.213) T2 (k−1)/k − T2 − − P T1 2 P1   Note that if a Carnot cycle were operating between the same temperature bounds that ηCarnot =1 T1/T3 would be greater than that for the Brayton cycle. Relative− to the Rankine cycle, the Brayton cycle has a large fraction of work. So the backwork ratio bwr is larger. One must also account for deviations from ideality. These effects are summarized in component efficiencies. The relevant efficiencies here, assuming a control volume approach, are

h2s h1 ηc = − , (2.214) h2 h1 h − h η = 3 − 4 . (2.215) t h h 3 − 4s For a CPIG, one gets then

T2s T1 ηc = − , (2.216) T2 T1 T − T η = 3 − 4 . (2.217) t T T 3 − 4s

Example 2.6 (adopted from Moran and Shapiro, Example 9.6) Air enters the compressor of an air-standard Brayton cycle at 100 kPa, 300 K with a volumetric flow rate is 5 m3/s. The pressure ratio in the compressor is 10. The turbine inlet temperature is 1400 K. Find the thermal efficiency, the back work ratio and the net power. Both the compressor and turbine have efficiencies of 0.8. First calculate the state after the compressor if the compressor were isentropic.

P T k/(k−1) 2 = 2s = 10, (2.218) P T 1  1  (k−1)/k T2s = T1(10) , (2.219) = (300 K)(10)(1.4−1)/1.4, (2.220) = 579.209 K. (2.221) Now T T η = 2s − 1 , (2.222) c T T 2 − 1 T2s T1 T2 = T1 + − , (2.223) ηc (579.209 K) (300 K) = (300 K)+ − , (2.224) 0.8 = 649.012 K. (2.225)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 42 CHAPTER 2. CYCLE ANALYSIS

Now state three has T = 1400 K. Recall that 2 3 involves heat addition in a combustion chamber. 3 → Now calculate T4s for the ideal turbine. Recall that the expansion is to the same pressure as the compressor inlet so that P3/P4 = 10. The isentropic relations give

T k/(k−1) P 4s = 4 , (2.226) T P  3  3 P (k−1)/k T = T 4 , (2.227) 4s 3 P  3  1 (1.4−1)/1.4 = (1400 K) , (2.228) 10   = 725.126 K. (2.229) Now account for the actual behavior in the turbine: T T η = 3 − 4 , (2.230) t T T 3 − 4s T = T η (T T ), (2.231) 4 3 − t 3 − 4s = (1400 K) (0.8)((1400 K) (725.126 K)), (2.232) − − = 860.101 K. (2.233) Now calculate the thermal efficiency of the actual cycle. w q q q c (T T ) T T η = net = H − L =1 L =1 P 4 − 1 =1 4 − 1 , (2.234) q q − q − c (T T ) − T T H H H P 3 − 2 3 − 2 860.101 K 300 K = 1 − , (2.235) − 1400 K 649.012 K − = 0.254181. (2.236) If the cycle were ideal, one would have w q q q c (T T ) T T η = net = H − L =1 L =1 P 4s − 1 =1 4s − 1 , (2.237) ideal q q − q − c (T T ) − T T H H H P 3 − 2s 3 − 2s (725.126 K) (300 K) = 1 − , (2.238) − (1400 K) (579.209 K) − = 0.482053. (2.239) Note also that for the ideal cycle one also has

T1 300 K ηideal =1 =1 =0.482053. (2.240) − T2s − 579.209 K The back work ratio is w bwr = comp , (2.241) wturb c (T T ) = P 2 − 1 , (2.242) c (T T ) P 3 − 4 T T = 2 − 1 , (2.243) T T 3 − 4 (649.012 K) (300 K) = − , (2.244) (1400 K) (860.101 K) − = 0.646439. (2.245)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.4. BRAYTON 43

If the process were ideal, the back work ratio would have been

wcomp bwrideal = , (2.246) wturb c (T T ) = P 2s − 1 , (2.247) c (T T ) P 3 − 4s T T = 2s − 1 , (2.248) T T 3 − 4s (579.209 K) (300 K) = − , (2.249) (1400 K) (725.126 K) − = 0.413721. (2.250)

Now get the net work. First get the mass flow rate from the volume flow rate. This requires the specific volume. kJ 0.287 (300 K) 3 RT1 kg K m v1 = = =0.861 . (2.251) P1  100 kPa kg

Now 3 ˙ m V 5 s kg m˙ = ρ V˙ = = 3 =5.8072 . (2.252) 1 v m s 1 0.861 kg

Recall cP = kcv =1.4(717.5 J/kg/K) = 1004.5 J/kg/K =1.0045 kJ/kg/K. Now

W˙ =mc ˙ ((T T ) (T T ), (2.253) cycle P 3 − 4 − 2 − 1 kg kJ = 5.8072 1.0045 ((1400 K) (860.101 K) (649.012 K) + (300 K))(2.254), s kg K − −     = 1113.51 kW. (2.255)

If the cycle were ideal, one would have

W˙ =mc ˙ ((T T ) (T T ), (2.256) ideal P 3 − 4s − 2s − 1 kg kJ = 5.8072 1.0045 ((1400 K) (725.126 K) (579.209 K)+ (300 K))(2.257), s kg K − −     = 2308.04 kW. (2.258)

Note that the inefficiencies had a significant effect on both the back work ratio and the net work of the engine.

2.4.2 Regeneration One can improve cycle efficiency by regeneration. The hot gas in the turbine is used to preheat the gas exiting the compressor before it enters the combustion chamber. In this cycle, one takes 1 2: compression, • → 2 x: compression exit gas goes through regenerator (), • →

CC BY-NC-ND. 28 January 2019, J. M. Powers. 44 CHAPTER 2. CYCLE ANALYSIS

x 3: combustion chamber, • → 3 4: expansion in turbine, and • → 4 y: turbine exit gas goes through regenerator (heat exchanger). • → A schematic for the Brayton cycle is shown in Figure 2.9.

y Regenerator

x Combustion 4 2 chamber 3 1

Compressor Turbine w

Figure 2.9: Schematic for Brayton cycle with regeneration.

For this to work, the gas at the turbine exit must have a higher temperature than the gas at the compressor exit. If the compression ratio is high, the compressor exit temperature is high, and there is little benefit to regeneration. Now consider the regenerator, which is really a heat exchanger. The first law holds that

dE cv = Q˙ W˙ +mh ˙ mh˙ +mh ˙ mh˙ , (2.259) dt cv − cv 2 − x 4 − y =0 =0 =0 0 = h|{z}2 hx|{z}+ h4 hy, (2.260) | {z } − − = c (T T )+ c (T T ), (2.261) P 2 − x P 4 − y = T T + T T . (2.262) 2 − x 4 − y Now consider the inlet temperatures to be known. Then the first law constrains the outlet temperatures such that

T2 + T4 = Tx + Ty. (2.263)

If the heat exchanger were an ideal co-flow heat exchanger, one might expect the outlet temperatures to be the same, that is, Tx = Ty, and one would have Tx = Ty = (1/2)(T2 +T4). But in fact one can do better. If a counter-flow heat exchanger is used, one could ideally expect to find

Tx = T4, Ty = T2. (2.264)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.4. BRAYTON 45

The thermal efficiency is w η = net , (2.265) qH w w = t − c , (2.266) qH c (T T ) c (T T ) = P 3 − 4 − P 2 − 1 , (2.267) c (T T ) P 3 − x (T T ) (T T ) = 3 − 4 − 2 − 1 , (2.268) T T 3 − x (T T ) (T T ) = 3 − 4 − 2 − 1 , (2.269) T3 T4 T T− = 1 2 − 1 , (2.270) − T T 3 − 4 T2 T1 T 1 = 1 1 − , (2.271) −  T4  T3 1 − T3  (k−1)/k P2 T1 1 P1 − = 1  , (2.272)   (k−1)/k − P4 T3 1 − P3   (k−1)/k P2 T1 1 P1 − = 1  , (2.273)   (k−1)/k − P1 T3 1 − P2     (k−1)/k (k−1)/k 1 P1 T P P2 = 1 1 2 − , (2.274) − T P  (k−1)/k 3  1  1 P1 − P2   T P (k−1)/k = 1 1 2 . (2.275) − T P 3  1  Note this is the Carnot efficiency moderated by the pressure ratio. As the pressure ratio rises, the thermal efficiency declines. The regenerator itself will not be perfect. To achieve the performance postulated here would require infinite time or an infinite area heat exchanger. As both increase, viscous losses increase, and so there is a trade-off. One can summarize the behavior of an actual regenerator by an efficiency defined as h h actual η = x − 2 = . (2.276) reg h′ h ideal x − 2 CC BY-NC-ND. 28 January 2019, J. M. Powers. 46 CHAPTER 2. CYCLE ANALYSIS

If we have a CPIG, then T T η = x − 2 . (2.277) reg T ′ T x − 2

Example 2.7 (extension for Moran and Shapiro, Example 9.6) How would the addition of a regenerator affect the thermal efficiency of the isentropic version of the previous example problem?

One may take the rash step of trusting the analysis to give a prediction from Eq. (2.275) of

T P (k−1)/k η = 1 1 2 , (2.278) − T P 3  1  300 K = 1 (10)(1.4−1)/1.4 = 0.586279. (2.279) − 1400 K   Without regeneration, the thermal efficiency of the ideal Brayton cycle had a value of 0.482053. Had the engine used a Carnot cycle between the same temperature limits, the efficiency would have been 0.785714.

2.4.3 If one used isothermal compression and expansion, which is slow and impractical, in place of isentropic processes in the Brayton cycle, one would obtain the Ericsson cycle. Diagrams for P v and T s for the Ericsson cycle are shown in Figure 2.10. − −

P T

2 3 3 4

isotherm isobar isobar

isotherm

1 4 2 1

v s

Figure 2.10: P v and T s diagram for the Ericsson cycle. − − One can outline the Ericsson cycle as follows:

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.4. BRAYTON 47

1 2: isothermal compression, • → 2 3: isobaric heating in the combustion chamber, • → 3 4: isothermal expansion in turbine, and • → 4 1: isobaric heat transfer in surroundings. • → One also takes the control volume approach. First consider the work in the isothermal compressor. For the control volume, one recalls that the work is given by the enthalpy difference, and that dh = Tds + vdP = δqc δwc, with δq = Tds and δw = vdP , so − c c − 2 2 dP P w = v dP = RT = RT ln 2 . (2.280) c − − 1 P − 1 P Z1 Z1 1

Note that P2/P1 > 1, so wc < 0; work is done on the fluid in the compressor. For the turbine, again for the control volume, the key is the enthalpy difference and dh = Tds + vdP = δq δw , with δq = Tds and δw = vdP . The turbine work would be t − t t t − 4 4 dP P w = v dP = RT = RT ln 4 . (2.281) t − − 3 P − 3 P Z3 Z3 3

But because P2 = P3 and P1 = P4, the turbine work is also

P1 P2 wt = RT3 ln = RT3 ln . (2.282) − P2 P1

Note that because P2/P1 > 1, the turbine work is positive. In the combustion chamber, one has from the first law

h h = q (2.283) 3 − 2 C c (T T ) = q . (2.284) P 3 − 2 C For the isothermal turbine, one has dh = c dT =0= q w , so q = w . The heat P t − t t t transfer necessary to keep the turbine isothermal is

P2 qt = wt = RT3 ln . (2.285) P1

So the total heat that one pays for is

P2 qH = qC + qt = cP (T3 T1)+ RT3 ln . (2.286) − P1

CC BY-NC-ND. 28 January 2019, J. M. Powers. 48 CHAPTER 2. CYCLE ANALYSIS

So the thermal efficiency is w + w η = t c , (2.287) qH P2 R(T3 T1) ln η = − P1 , (2.288) P2 cP (T3 T1)+ RT3 ln − P1 T1 P2 1 T ln P = − 3 1 , (2.289) k  T  P 1 1 + ln 2 k−1 − T3 P1   (k−1)/k T1 P2 1 T ln P = − 3 1 , (2.290)     (k−1)/k 1 T1 + ln P2 − T3 P1     T 1 = 1 1   , (2.291) 1− T1 − T3 T3   1+ (k−1)/k   ln P2    P1     T 1 T1 1 1 1 − T3 + ... . (2.292) ∼ − T  − (k−1)/k   3  ln P2 P1 Carnot     correction  | {z } This is expressed as a Carnot efficiency modified| by a correction{z which} degrades the efficiency. Note the efficiency is less than that of a Carnot cycle. For high temperature ratios and high pressure ratios, the efficiency approaches that of a Carnot engine. Now if one used multiple staged intercooling on the and multiple staged expansions with reheat on the turbines, one can come closer to the isothermal limit, and better approximate the Carnot cycle.

2.4.4 Jet propulsion If one modifies the Brayton cycle so that the turbine work is just sufficient to drive the compressor and the remaining enthalpy at the turbine exit is utilized in expansion in a nozzle to generate thrust, one has the framework for jet propulsion. Here the control volume approach is used. A schematic for a is shown in Figure 2.11. Diagrams for P v − and T s for the jet propulsion Brayton cycle are shown in Figure 2.12. An important component− of jet propulsion analysis is the of the flow. In the entrance region of the engine, the flow is decelerated, inducing a ram compression effect. For high velocity applications, this effect provides sufficient compression for the cycle and no compressor or turbine are needed! However, such a device, known as a , is not self-starting, and so is not practical for many applications.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.4. BRAYTON 49

Burner 3 2

1 a 4 Di!user Compressor Turbine Nozzle

5

environmental exhaust return

Figure 2.11: Schematic for jet propulsion.

P T 3 2 3 4 isobar isentrope 4 2 5 isentrope a a 1 isobar 1 5

v s

Figure 2.12: P v and T s diagrams for jet propulsion Brayton cycle. − − One can outline the ideal jet propulsion cycle as follows: 1 a: deceleration and ram compression in diffuser, • → a 2: isentropic compression in the compressor, • → 2 3: isobaric heating in the combustion chamber, • → 3 4: isentropic expansion in turbine, • → 4 5: isentropic conversion of thermal energy to kinetic energy in nozzle, and • → 5 1: isobaric heat transfer in surroundings. • → The goal of the jet propulsion cycle is to produce a thrust force which is necessary to balance a fluid-induced aerodynamic drag force. When such a balance exists, the system is in steady state. One can analyze such a system in the reference frame in which the engine is stationary. In such a frame, Newton’s second law gives d (m v ) = F + P A i P A i +m ˙ v m˙ v . (2.293) dt CV CV 1 1 − 5 5 1 − 5 thrust force small =0 |{z} | {z } | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. 50 CHAPTER 2. CYCLE ANALYSIS

Solving for the thrust force, neglecting the small differences in pressure, one gets,

F =m ˙ (v v ) . (2.294) 5 − 1

The work done by the thrust force is the product of this force and the air speed, v1. This gives W˙ = F v = m˙ (v v ) v . (2.295) | p| | · 1| | 5 − 1 · 1| Now the efficiency of the cycle is a bit different. The net work of the turbine and compressor is zero. Instead, the propulsive efficiency is defined as

Propulsive Power W˙ p ηp = = | | , (2.296) Energy Input Rate Q˙ | H | m˙ v (v v ) = | 1 · 5 − 1 |. (2.297) m˙ (h h ) | 3 − 2 | Note the following unusual behavior for flight in an ideal inviscid in which the flow always remains attached. In such a flow D’Alembert’s paradox holds: there is no aerodynamic drag. Consequently there is no need for thrust generation in steady state operation. Thrust would only be needed to accelerate to a particular velocity. For such an engine then the exit velocity would equal the entrance velocity: v5 = v1 and F = 0. Moreover, the propulsive efficiency is ηp = 0.

Example 2.8 ft (adopted from C¸engal and Boles, p. 485). A flies with a velocity of 850 s at an altitude where the air is at 5 psia and 40 F . The compressor has a pressure ratio of 10, and the temperature − lbm at the turbine inlet is 2000 F . Air enters the compressor at a rate ofm ˙ = 100 s . Assuming an air standard with CPIG air, find the temperature and pressure at the turbine exit, • velocity of gas at nozzle exit, and • propulsive efficiency. • For the English units of this problem, one recalls that

lbm ft Btu ft2 1 Btu = 777.5 ftlbf, g = 32.2 , 1 = 25037 . (2.298) c lbf s2 lbm s2 For air, one has

Btu ft lbf ft2 Btu c =0.240 , R = 53.34 = 1717.5 =0.0686 , k =1.4. (2.299) P lbm ◦R lbm ◦R s2 ◦R lbm ◦R First one must analyze the ram compression process. An energy balance gives v v v v h + 1 · 1 = h + a · a . (2.300) 1 2 a 2 small | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.4. BRAYTON 51

It is usually sufficient to neglect the kinetic energy of the fluid inside the engine. While it still has a non-zero velocity, it has been slowed enough so that enthalpy dominates kinetic energy inside the engine. So one takes v v h = h + 1 · 1 , (2.301) a 1 2 v v 0 = h h + 1 · 1 , (2.302) 1 − a 2 v v 0 = c (T T )+ 1 · 1 , (2.303) P 1 − a 2 v1 v1 Ta = T1 + · . (2.304) 2cP

The ambient temperature is T = 40+460 = 420 ◦R. (2.305) 1 − So after ram compression, the temperature at the inlet to the compressor is

2 ft 850 Btu ◦ s 1 lbm ◦ Ta = (420 R)+ Btu ft2 = 480 R. (2.306) 2 0.240  ◦ lbm R 25037 s2

Now get the pressure at the compressor inlet via the isentropic relations:

T k/(k−1) 480 ◦R 1.4/(1.4−1) P = P a = (5 psia) =8.0 psia. (2.307) a 1 T 420 ◦R  1    Now consider the isentropic compressor.

P2 = 10Pa = 10(8.0 psia)=80 psia. (2.308)

Now get the temperature after passage through the compressor.

P (k−1)/k T = T 2 = (480 ◦R) (10)(1.4−1)/1.4 = 927 ◦R. (2.309) 2 a P  a  The temperature at the entrance of the turbine is known to be 2000 F = 2460 ◦R. Now the turbine work is equal to the compressor work, so

wc = wt, (2.310) h h = h h , (2.311) 2 − a 3 − 4 c (T T ) = c (T T ), (2.312) P 2 − a P 3 − 4 T T = T T , (2.313) 2 − a 3 − 4 T = T T + T , (2.314) 4 3 − 2 a = 2460 ◦R 927 ◦R + 480 ◦R, (2.315) − = 2013 ◦R. (2.316)

Use the isentropic relations to get P4:

T k/(k−1) 2013 ◦R 1.4/(1.4−1) P = P 4 = (80 psia) = 39.7 psia. (2.317) 4 3 T 2460 ◦R  3   

CC BY-NC-ND. 28 January 2019, J. M. Powers. 52 CHAPTER 2. CYCLE ANALYSIS

Expansion through the nozzle completes the cycle. Assume an isentropic nozzle, therefore

P (k−1)/k T = T 5 , (2.318) 5 4 P  4  P (k−1)/k = T 1 , (2.319) 4 P  4  5 psia (1.4−1)/1.4 = =(2013 ◦R) , (2.320) 39.7 psia   = 1114 ◦R. (2.321)

Now consider the energy balance in the nozzle:

v v v v h + 5 · 5 = h + 4 · 4 , (2.322) 5 2 4 2 small v2 0 = h h| {z+ }5 , (2.323) 5 − 4 2 v v = c (T T )+ 5 · 5 , (2.324) P 5 − 4 2 v = 2c (T T ), (2.325) 5 P 4 − 5 p ft2 Btu 25037 2 = 2 0.240 (2013 ◦R 1114 ◦R) s , (2.326) v lbm ◦R − Btu u 1 lbm ! u   t ft = 3288 . (2.327) s

Now find the thrust force magnitude

lbm ft ft ft lbm F =m ˙ (v v )= 100 3288 850 = 243800 , (2.328) | | 5 − 1 s s − s s2     ft lbm 243800 s2 = lbm ft , (2.329) 32.2 lbf s2 = 7571.4 lbf. (2.330)

The power is the product of the thrust force and the air speed, so

ft Btu W˙ = F v = (7571.4 lbf) 850 , (2.331) | P | | · 1| s 777.5 ft lbf     Btu = 8277 , (2.332) s Btu 1.415 hp = 8277 , (2.333) s Btu   s ! = 11711 hp. (2.334)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.5. POWER CYCLES 53

Now for the heat transfer in the combustion chamber

Q˙ =m ˙ (h h ), (2.335) H 3 − 2 =mc ˙ (T T ), (2.336) P 3 − 2 lbm Btu = 100 0.240 (2460 ◦R 927 ◦R), (2.337) s lbm ◦R −     Btu = 36792 . (2.338) s So the propulsion efficiency is

W˙ 8277 Btu η = P = s = 0.225. (2.339) p ˙ Btu QH 36792 s The remainder of the energy is excess kinetic energy and excess thermal energy, both of which ultimately dissipate so as to heat the atmosphere.

Other variants of the turbojet engine include the very important engine in which a large cowling is added to the engine and a an additional in front of the compressor forces a large fraction of the air to bypass the engine. The turbofan engine is used in most large passenger jets. Analysis reveals a significant increase in propulsive efficiency as well as a reduction in jet noise. Other important variants include the turboprop, propfan, ramjet, scramjet, jet with , and .

2.5 Reciprocating engine power cycles

Here are some common notions for engines that depend on pistons driving in cylinders. The piston has a bore diameter B. The piston is connected to the crankshaft, and the stroke S of the piston is twice the radius of the crank, Rcrank:

S =2Rcrank. (2.340)

See Figure 2.13 for an illustration of this geometry for a piston-cylinder arrangement in two configurations. The total displacement for all the cylinders is

V = N (V V )= N A S. (2.341) displ cyl max − min cyl cyl Note that B2 A = π . (2.342) cyl 4 The compression ratio is not the ratio of ; it is the ratio of :

Vmax rv = CR = . (2.343) Vmin

CC BY-NC-ND. 28 January 2019, J. M. Powers. 54 CHAPTER 2. CYCLE ANALYSIS

B

S

2R crank

Figure 2.13: Piston-cylinder configurations illustrating geometric definitions.

For these engines the volumes are closed in expansion and compression, so the net work is

w = P dv = P (v v ). (2.344) net meff max − min I

Here one has defined a mean effective pressure: Pmeff . The net work per cylinder per cycle is W = mw = P m(v v )= P (V V ). (2.345) net net meff max − min meff max − min Now consider the total power developed for all the cylinders. Assume the piston operates at a frequency of ν cycles/s:

W˙ = N P (V V )ν. (2.346) net cyl meff max − min s It is more common to express ν in revolutions per minute: RP M = ν(60 min ), so

˙ RP M Wnet = Pmeff Ncyl(Vmax Vmin) s , (2.347) − 60 min =Vdispl RP M | {z } = Pmeff Vdispl s . (2.348) 60 min

This applies for a two-stroke engine. If the engine is a four stroke engine, the net power is reduced by a factor of 1/2.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.6. OTTO 55

2.6 Otto

The air-standard approximates the gasoline engine. It employs a fixed mass approach. Diagrams for P v and T s for the Otto cycle are shown in Figure 2.14. − −

P T

3 3

isentrope isochore

2 4 2 isentrope 4 1 isochore 1

v s

Figure 2.14: P v and T s diagrams for the Otto cycle. − −

One can outline the Otto cycle as follows:

1 2: isentropic compression in the compression stroke, • → 2 3: isochoric heating in the combustion stroke during spark ignition, • → 3 4: isentropic expansion in power stroke, and • → 4 1: isochoric rejection of heat to the surroundings. • → Note for isochoric heating, such as 2 3, in a fixed mass , the first law gives →

u3 u2 = 2q3 2w3, (2.349) − − v3 u3 u2 = 2q3 P dv, but v2 = v3, (2.350) − − v2 Z v2 u u = q P dv, (2.351) 3 − 2 2 3 − Zv2 =0 q = u u , (2.352) 2 3 3 − |2 {z } q = c (T T ), if CPIG. (2.353) 2 3 v 3 − 2

CC BY-NC-ND. 28 January 2019, J. M. Powers. 56 CHAPTER 2. CYCLE ANALYSIS

The thermal efficiency is found as follows: w η = net , (2.354) qH q q = H − L , (2.355) qH q = 1 L , (2.356) − qH c (T T ) = 1 v 4 − 1 , (2.357) − c (T T ) v 3 − 2 T T = 1 4 − 1 , (2.358) − T T 3 − 2 T4 T1 T 1 = 1 1 − . (2.359) −  T3  T2 1 T2 − Now one also has the isentropic relations:   T V k−1 2 = 1 , (2.360) T V 1  2  T V k−1 3 = 4 . (2.361) T V 4  3  But V4 = V1 and V2 = V3, so T V k−1 T 3 = 1 = 2 . (2.362) T V T 4  2  1 Cross multiplying the temperatures, one finds T T 3 = 4 . (2.363) T2 T1 Thus the thermal efficiency reduces to T η =1 1 . (2.364) − T2

In terms of the compression ratio r = V1 , one has v V2

1−k 1 η =1 rv =1 k−1 . (2.365) − − rv Note if the compression ratio increases, the thermal efficiency increases. High compression ratios introduce detonation in the fuel air mixture. This induces strong pressure waves in the cylinder and subsequent engine knock. It can cause degradation of piston walls. Some deviations of actual performance from that of the air-standard Otto cycle are as follows:

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.6. OTTO 57

specific heats actually vary with temperature, • combustion may be incomplete (induces pollution and lowers fuel efficiency), • work of inlet and exhaust ignored, and • losses of heat transfer to engine walls ignored. •

Example 2.9 (adopted from Moran and Shapiro, p. 363). The temperature at the beginning of the compression process of an air-standard Otto cycle with a compression ratio of 8 is 540 ◦R, the pressure is 1 atm, and the cylinder volume is 0.02 ft3. The maximum temperature is 3600 ◦R. Find temperature and pressure at each stage of the process, • thermal efficiency, and • P in atm. • meff For the isentropic compression,

V k−1 T = T 1 , (2.366) 2 1 V  2  = (540 ◦R)(8)1.4−1, (2.367) = 1240.69 ◦R. (2.368)

One can use the ideal gas law to get the pressure at state 2:

P V P V 2 2 = 1 1 , (2.369) T2 T1 V1 T2 P2 = P1 , (2.370) V2 T1 1240.69 ◦R = (1 atm)(8) , (2.371) 540 ◦R   = 18.3792 atm. (2.372)

◦ Now V3 = V2 since the combustion is isochoric. And the maximum temperature is T3 = 3600 R. This allows use of the ideal gas law to get P3:

P V P V 3 3 = 2 2 , (2.373) T3 T2 V2 T3 P3 = P2 , (2.374) V3 T2 =1 3600 ◦R = (18|{z}.3792 atm)(1) , (2.375) 1240.59 ◦R   = 53.3333 atm. (2.376)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 58 CHAPTER 2. CYCLE ANALYSIS

One uses the isentropic relations for state 4:

T V k−1 3 = 4 , (2.377) T V 4  3  T V k−1 4 = 3 , (2.378) T V 3  4  V k−1 T = T 2 , (2.379) 4 3 V  1  V 1−k T = T 1 , (2.380) 4 3 V  2  = (3600 ◦R)(8)1−1.4, (2.381) = 1566.99 ◦R. (2.382)

For the pressure at state 4, use the ideal gas law:

P V P V 4 4 = 3 3 , (2.383) T4 T3 V3 T4 P4 = P3 , (2.384) V4 T3 V2 T4 = P3 , (2.385) V1 T3 1 1566.99 ◦R = (53.3333 atm) , (2.386) 8 3600 ◦R   = 2.90184 atm. (2.387)

The thermal efficiency is

1 η = 1 k−1 , (2.388) − rv 1 = 1 , (2.389) − 81.4−1 = 0.564725. (2.390)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.7. DIESEL 59

Now get the mean effective pressure Pmeff .

W = P (V V ), (2.391) net meff max − min W P = net , (2.392) meff V V max − min m(u u )+ m(u u ) = 3 − 4 1 − 2 , (2.393) V V 1 − 2 c T T + T T = m v 3 4 1 2 , (2.394) − V2 − V1 1 − V1 P V c T T + T T = 1 1 v 3 4 1 2 , (2.395) − V2 − RT1 V1 1 − V1 P c T T + T T = 1 v 3 4 1 2 , (2.396) − V2 − T1 R 1 − V1 P 1 T T + T T = 1 3 4 1 2 , (2.397) − V2 − T1 k 1 1 − − V1 1 atm 1 3600 ◦R 1566.99 ◦R + 540 ◦R 1240.59 ◦R = − − , (2.398) 540 ◦R 1.4 1 1 1 − − 8 = 7.04981 atm. (2.399)

2.7 Diesel

The air standard approximates the behavior of a . It is modeled as a fixed mass system. Here the compression is done before injection, so there is no danger of premature ignition due to detonation. No spark plugs are used. Diagrams for P v and T s for the Diesel cycle are shown in Figure 2.15. One can outline the Diesel− cycle as follows:−

1 2: isentropic compression in the compression stroke, • →

2 3: isobaric heating in the combustion stroke, • →

3 4: isentropic expansion in power stroke, and • →

4 1: isochoric rejection of heat to the surroundings. • →

CC BY-NC-ND. 28 January 2019, J. M. Powers. 60 CHAPTER 2. CYCLE ANALYSIS

P T

2 3 3 isobar isentrope 2 4 isentrope 4 1 isochore

1

v s

Figure 2.15: P v and T s diagrams for the Diesel cycle. − −

The thermal efficiency is found as follows: w η = net , (2.400) qH q q = H − L , (2.401) qH q = 1 L , (2.402) − qH u u = 1 4 − 1 , (2.403) − h h 3 − 2 c (T T ) = 1 v 4 − 1 , (2.404) − c (T T ) P 3 − 2 1 T T4 1 = 1 1 T1 − . (2.405) T3 − k T2 1 T2 − All else being equal, the Otto cycle will have higher efficiency than the Diesel cycle. However, the Diesel can operate at higher compression ratios because detonation is not as serious a problem in the compression ignition engine as it is in the spark ignition.

Example 2.10 (adopted from BS, Ex. 12.8 pp. 501-502). An air standard Diesel cycle has a compression ratio of 20, and the heat transferred to the working fluid has 1800 kJ/kg. Take air to be an ideal gas with ◦ variable specific heat. At the beginning of the compression process, P1 = 0.1 MPa, and T1 = 15 C. Find Pressure and temperature at each point in the process, • Thermal efficiency, • P , the mean effective pressure. • meff

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.7. DIESEL 61

The and at the reference pressure for variable specific heat air are tabulated as functions of temperature in Table A7.1. Recall that

T h(T ) = href + cP (Tˆ) dT.ˆ (2.406) ZTref Recall further that T ds = dh vdP, (2.407) − dh v ds = dP. (2.408) T − T

Now for an ideal gas it can be shown that dh = cP (T )dT ; one also has v/T = R/P . Making these substitutions into Eq. (2.408) gives c (T )dT dP ds = P R . (2.409) T − P Integrating gives

T cP (Tˆ) P s(T, P ) = sref + dTˆ R ln , (2.410) Tˆ − Pref ZTref o =sT |o {z P } s(T, P ) = sT (T ) R ln . (2.411) − Pref

Here the reference pressure is Pref = 0.1 MPa. Note that this happens to be the inlet state, which is a coincidence; so at the inlet there is no correction to the entropy for pressure deviation from its reference value. At the inlet, one has T1 = 15+273.15 = 288.15 K. One then gets from interpolating Table A7.1 that kJ kJ kJ h = 288.422 , so = s =6.82816 , u = 205.756 . (2.412) 1 kg T1 1 kg K 1 kg For the isentropic compression, one has kJ 1 s = s =6.82816 , V = V . (2.413) 2 1 kg K 2 20 1 Now from the ideal gas law, one has P V P V 2 2 = 1 1 , (2.414) T2 T1 P T V 2 = 2 1 , (2.415) P1 T1 V2 P T 2 = 20 2 . (2.416) Pref 288 K Now

o P2 s2(T2, P2) = sT2 R ln , (2.417) − Pref kJ kJ T 6.82816 = so 0.287 ln 20 2 , (2.418) kg K T2 − kg K 288 K       kJ T kJ 0 = so 0.287 ln 20 2 6.82816 F (T ). (2.419) T2 − kg K 288 K − kg K ≡ 2      

CC BY-NC-ND. 28 January 2019, J. M. Powers. 62 CHAPTER 2. CYCLE ANALYSIS

o kJ kJ T2 (K) sT2 kg K F (T2) kg K 900 8.01581 8.57198 10−4 × 1000 8.13493 8.97387 10−2 800 7.88514 9.60091 × 10−2 − × 850 7.95207 4.64783 10−2 899.347 8.01237 −2.37449 × 10−3 − ×

Table 2.3: Iteration for T2.

This then becomes a trial and error search in Table A7.1 for the T2 which satisfies the previous equa- o tion. If one guesses T2 900 K, one gets sT2 = 8.01581 kJ/kg/K, and the right side evaluates to 0.000857198 kJ/kg/K.∼ Performing a trial and error procedure, one finds the results summarized in

Table 2.3. Take T2 = 900 K as close enough! Then

T2 V1 P2 = P1 , (2.420) T1 V2 900 K = (0.1 MPa) (20), (2.421) 288 K = 6.25 MP a. (2.422)

Now at state 2, T2 = 900 K, one finds that kJ kJ h = 933.15 , u = 674.82 . (2.423) 2 kg 2 kg The heat is added at constant pressure, so

q = h h , (2.424) H 3 − 2 h3 = h2 + qH , (2.425) kJ kJ = 933.15 + 1800 , (2.426) kg kg     kJ = 2733.15 . (2.427) kg

o One can then interpolate Table A7.1 to find T3 and sT3

2733.15 kJ 2692.31 kJ kg − kg T3 = (2350 K)+ ((2400 K) (2350 K)), (2.428)  2755 kJ  2692 .31 kJ  − kg − kg = 2382.17 K,     (2.429) kJ so = 9.16913 T3 kg K   kJ kJ 2733.15 kg 2692.31 kg kJ kJ + − 9.19586 9.16913 , (2.430)  2755 kJ  2692 .31 kJ  kg K − kg K kg − kg      kJ    = 9.18633 . (2.431) kg K

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.7. DIESEL 63

o kJ kJ T4 (K) sT4 kg K F (T4) kg K 400 7.15926 9.34561 10−1 − × 800 7.88514 4.07614 10−1 1400 8.52891 −7.55464 × 10−2 × 1200 8.34596 6.31624 10−2 1300 8.44046 −8.36535 × 10−3 1250 8.39402 2.68183 × 10−2 − × 1288 8.42931 1.23117 10−4 − ×

Table 2.4: Iteration for T4.

One also has P3 = P2 = 6.25 MP a. Now get the actual entropy at state 3:

o P3 s3(T3, P3) = sT3 R ln , (2.432) − Pref kJ kJ 6.25 MPa = 9.18633 0.287 ln , (2.433) kg K − kg K 0.1 MPa       kJ = 7.99954 . (2.434) kg K

Now 3 4 is an isentropic expansion to state 4, which has the same volume as state 1; V1 = V4. So the ideal→ gas law gives

P V P V 4 4 = 1 1 , (2.435) T4 T1 P T V 4 = 4 1 , (2.436) P1 T1 V4 P T V 4 = 4 1 . (2.437) Pref T1 V4

Now consider the entropy at state 4, which must be the same as that at state 3:

o P4 s4(T4, P4) = sT4 R ln , (2.438) − Pref

o T4 V1 s3 = sT4 R ln   , (2.439) − T1 V4    =1  kJ  kJ  T 7.99954 = so 0.287 |{z} ln 4 , (2.440) kg K T4 − kg K 288 K       kJ T kJ 0 = so 0.287 ln 4 7.99954 = F (T ). (2.441) T4 − kg K 288 K − kg K 4      

Using Table A7.1, this equation can be iterated until T4 is found. So T4 = 1288 K. At this temperature, one has kJ kJ h = 1381.68 , u = 1011.98 . (2.442) 4 kg 4 kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 64 CHAPTER 2. CYCLE ANALYSIS

Now one can calculate the cycle efficiency: u u η = 1 4 − 1 , (2.443) − h h 3 − 2 kJ kJ 1011.98 kg 205.756 kg = 1 − , (2.444) −  2733.15 kJ  933.15 kJ  kg − kg = 0.552098 .    (2.445) Now one has kJ kJ kJ q = u u = 205.756 1011.98 = 806.224 . (2.446) L 1 − 4 kg − kg − kg     So, kJ kJ kJ w = q + q = 1800 806.224 = 993.776 . (2.447) net H L kg − kg kg     So

Wnet Pmeff = , (2.448) Vmax Vmin w− = net , (2.449) vmax vmin w − = net , (2.450) v1 v2 −w = net , (2.451) RT1 RT2 P1 − P2 993.776 kJ = kg , (2.452) 0.287 kJ 288 K 900 K kg K 100 kPa − 6250 kPa = 1265.58 kP a.  (2.453)

2.8 Stirling

Another often-studied air-standard engine is given by the . This is similar to the Otto cycle except the adiabatic processes are replaced by isothermal ones. The efficiency can be shown to be equal to that of a Carnot engine. Stirling engines are difficult to build. Diagrams for P v and T s for the Stirling cycle are shown in Figure 2.16. One can outline− the Stirling− cycle as follows: 1 2: isothermal compression in the compression stroke, • → 2 3: isochoric heat transfer in the combustion stroke, • → 3 4: expansion in power stroke, and • → 4 1: isochoric rejection of heat to the surroundings. • →

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.9. REFRIGERATION 65

P T

3 3 4

isotherm isochore isochore

2 isotherm 4 2 1 1

v s

Figure 2.16: P v and T s diagrams for the Stirling cycle. − − 2.9 Refrigeration

A simple way to think of a refrigerator is a cyclic operating in reverse. Rather than extracting work from heat transfer from a high temperature source and rejecting heat to a low temperature source, the refrigerator takes a work input to move heat from a low temperature source to a high temperature source.

2.9.1 Vapor-compression

A common refrigerator is based on a vapor-compression cycle. This is a Rankine cycle in reverse. While one could employ a turbine to extract some work, it is often impractical. Instead the high pressure gas is simply irreversibly throttled down to low low pressure. One can outline the vapor-compression refrigeration cycle as follows:

1 2: isentropic compression • → 2 3: isobaric heat transfer to high temperature reservoir in condenser, • → 3 4: adiabatic expansion in throttling , and • → 4 1: isobaric (and often isothermal) heat transfer to low temperature reservoir in • .→

A schematic for the vapor-compression refrigeration cycle is shown in Figure 2.17. A T s diagram for the vapor-compression refrigeration cycle is shown in Figure 2.18. − The efficiency does not make sense for a refrigerator as 0 η 1. Instead a coefficient ≤ ≤

CC BY-NC-ND. 28 January 2019, J. M. Powers. 66 CHAPTER 2. CYCLE ANALYSIS

Condenser 3 2

Compressor w Expansion Compressor valve

4 Evaporator

1

Figure 2.17: Schematic diagram for the vapor-compression refrigeration cycle.

of performance, β, is defined as

what one wants β = , (2.454) what one pays for q = L . (2.455) wc

Note that a is effectively the same as a refrigerator, except one desires qH ′ rather than qL. So for a heat pump, the coefficient of performance, β , is defined as

q β′ = H . (2.456) wc

It is possible for both β and β′ to be greater than unity.

Example 2.11 (from Moran and Shapiro, p. 442) R 12 is the working fluid in an ideal vapor-compression refrigeration cycle that communicates thermally− with a cold region at 20 ◦C and a warm region at 40 ◦C. Saturated vapor enters the compressor at 20 ◦C and saturated liquid leaves the condenser at 40 ◦C. The mass flow rate of the is 0.008 kg/s. Find compressor power in kW , • refrigeration capacity in ton, • coefficient of performance, and • coefficient of performance of an equivalent Carnot cycle. •

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.9. REFRIGERATION 67

T 2

3

4 1

s

Figure 2.18: T s diagram for the vapor-compression refrigeration cycle. −

At compressor inlet for the ideal cycle, one has a saturated vapor.

◦ T1 = 20 C, x1 =1. (2.457)

The tables then give

kJ kJ h = 195.78 , s =0.6884 , P =5.6729 bar. (2.458) 1 kg 1 kg K 1

◦ Now at the end of the condenser (state 3), one has T3 = 40 C. The condenser is on an isobar, so the pressure is the saturation pressure at the temperature, which is

P3 = P2 =9.6065 bar. (2.459)

So, two properties at the end of the compression are known: pressure and entropy. The allows deter- mination of the enthalpy at the end of compression via the tables:

kJ h = 205.1 . (2.460) 2 kg

State 3 is at the end of the condenser, so x3 = 0, and one finds the enthalpy from the tables to be

kJ h = 74.59 . (2.461) 3 kg

Now the throttling device has constant enthalpy, so

kJ h = h = 74.59 . (2.462) 4 3 kg

However, the fluid has been throttled to a lower pressure: that of the evaporator, which is the same as state 1, the compressor inlet, so

P4 = P1 =5.6729 bar. (2.463)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 68 CHAPTER 2. CYCLE ANALYSIS

The compressor work is W˙ =m ˙ (h h ), (2.464) c 2 − 1 kg kJ kJ = 0.008 205.1 195.78 , (2.465) s kg − kg       = 0.075 kW. (2.466) Now one desires the heat which leaves the cold region to be high for a good refrigerator. This is the heat transfer at the low temperature part of the T s diagram, which here gives − Q˙ =m ˙ (h h ), (2.467) L 1 − 4 kg kJ kJ = 0.008 195.78 74.59 , (2.468) s kg − kg       = 0.9695 kW, (2.469) kJ 60 s 1 ton = 0.9695 , (2.470) s min 211 kJ     min ! = 0.276 ton. (2.471) The units ton is used for power and is common in the refrigeration industry. It is the power required to freeze one “short ton” of water at 0 ◦C in 24 hours. It is 12000 Btu/hr,3.516853 kW , or4.7162 hp. A short ton is a U.S ton, which is 2000 lbm. A long ton is British and is 2240 lbm. A metric ton is 1000 kg and is 2204 lbm. The coefficient of performance is Q˙ β = L , (2.472) W˙ c 0.9695 kW = , (2.473) 0.075 kW = 12.93. (2.474) The equivalent Carnot refrigerator would have

Q˙ L βmax = , (2.475) W˙ c Q˙ = L , (2.476) Q˙ Q˙ H − L 1 = , (2.477) Q˙ H ˙ 1 QL − 1 = , (2.478) TH 1 TL − T = L , (2.479) TH TL − 20 + 273.15 = , (2.480) 40+ 273.15 (20 + 273.15) − = 14.6575. (2.481)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 2.9. REFRIGERATION 69

2.9.2 Air standard This is effectively the inverse Brayton cycle. It is used in the liquefaction of air and other gases. It is also used in aircraft cabin cooling. It has the following components:

1 2: isentropic compression, • → 2 3: isobaric heat transfer to a high temperature environment • → 3 4: isentropic expansion through turbine, and • → 4 1: isobaric heat exchange with low temperature surroundings. • → A schematic for the air standard refrigeration cycle is shown in Figure 2.19. A T s diagram −

Q H

3 2

- w Expander Compressor

1

4

Q L

Figure 2.19: Schematic for air standard refrigeration cycle. for the air standard refrigeration cycle is shown in Figure 2.20.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 70 CHAPTER 2. CYCLE ANALYSIS

T 2

isobar

3

1

4 isobar

s

Figure 2.20: T s diagram for air standard refrigeration Brayton cycle. −

CC BY-NC-ND. 28 January 2019, J. M. Powers. Chapter 3

Gas mixtures

Read BS, Chapter 11. See for background Sandler, Chapter 7. See for background Smith, van Ness, and Abbott, Chapter 10. See for background Tester and Modell. One is often faced with mixtures of simple compressible substances, and it the thermody- namics of such mixtures upon which attention is now fixed. Here a discussion of some of the fundamentals of mixture theory will be given. In general, thermodynamics of mixtures can be a challenging topic about which much remains to be learned. In particular, these notes will focus on ideal mixtures of ideal gases, for which results are often consistent with intuition. The chemical engineering literature contains a full discussion of the many nuances associated with non-ideal mixtures of non-ideal materials.

3.1 Some general issues

Here the notation of BS will be used. There is no effective consensus on notation for mixtures. That of BS is more unusual than most; however, the ideas are correct, which is critical. Consider a mixture of N components, each a pure substance, so that the total mass and total number of moles are

N m = m + m + m + + m = m , mass, units= kg, (3.1) 1 2 3 ··· N i i=1 X N n = n + n + n + + n = n , moles, units= kmole. (3.2) 1 2 3 ··· N i i=1 X Recall 1 = 6.02214179 1023 molecule. The mass fraction of component i is defined × as ci: m c i , mass fraction, dimensionless. (3.3) i ≡ m

71 72 CHAPTER 3. GAS MIXTURES

The of component i is defined as yi:

n y i , mole fraction, dimensionless. (3.4) i ≡ n

Now the molecular mass of species i is the mass of a mole of species i. It units are typically g/mole. This is an identical unit to kg/kmole. Molecular mass is sometimes called “molec- ular weight,” but this is formally incorrect, as it is a mass measure, not a force measure. Mathematically the definition of Mi corresponds to

m kg g M i , = . (3.5) i ≡ n kmole mole i   Then one gets mass fraction in terms of mole fraction as m c = i , (3.6) i m n M = i i , (3.7) m niMi = N , (3.8) j=1 mj n M P i i = N , (3.9) j=1 njMj niMi = P n , (3.10) 1 N n j=1 njMj niMi = P n , (3.11) N nj Mj j=1 n y M P i i = N . (3.12) j=1 yjMj

Similarly, one finds mole fraction in termsP of mass fraction by the following: n y = i , (3.13) i n mi = Mi , (3.14) N mj j=1 Mj mi = P Mim , (3.15) N mj j=1 Mj m ci = P Mi . (3.16) N cj j=1 Mj P CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.1. SOME GENERAL ISSUES 73

The mixture itself has a mean molecular mass: m M , (3.17) ≡ n N m = i=1 i , (3.18) n P N n M = i i , (3.19) n i=1 XN

= yiMi. (3.20) i=1 X

Example 3.1 Air is often modeled as a mixture in the following molar ratios:

O2 +3.76N2. (3.21)

Find the mole fractions, the mass fractions, and the mean molecular mass of the mixture.

Take O2 to be species 1 and N2 to be species 2. Consider the number of moles of O2 to be

n1 =1 kmole, (3.22)

and N2 to be n2 =3.76 kmole. (3.23)

The molecular mass of O2 is M1 = 32 kg/kmole. The molecular mass of N2 is M2 = 28 kg/kmole. The total number of moles is

n =1 kmole +3.76 kmole =4.76 kmole. (3.24)

So the mole fractions are 1 kmole y = = 0.2101, (3.25) 1 4.76 kmole 3.76 kmole y = = 0.7899. (3.26) 2 4.76 kmole Note that N yi =1. (3.27) i=1 X That is to say y1 + y2 =0.2101+ 0.7899 = 1. Now for the , one has

kg m = n M = (1 kmole) 32 = 32 kg, (3.28) 1 1 1 kmole   kg m = n M = (3.76 kmole) 28 = 105.28 kg. (3.29) 2 2 2 kmole   So one has m = m1 + m2 = 32 kg + 105.28 kg = 137.28 kg. (3.30)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 74 CHAPTER 3. GAS MIXTURES

The mass fractions then are m 32 kg c = 1 = = 0.2331, (3.31) 1 m 137.28 kg m 105.28 kg c = 2 = = 0.7669. (3.32) 2 m 137.28 kg

Note that N ci =1. (3.33) i=1 X That is c1 + c2 =0.2331+ 0.7669 = 1. Now for the mixture molecular mass, one has

m 137.28 kg kg M = = = 28.84 . (3.34) n 4.76 kmole kmole

Check against another formula.

N kg kg kg M = y M = y M + y M = (0.2101) 32 + (0.7899) 28 = 28.84 . (3.35) i i 1 1 2 2 kmole kmole kmole i=1 X    

Now postulates for mixtures are not as well established as those for pure substances. The literature has much controversial discussion of the subject. A strong advocate of the axiomatic approach, C. A. Truesdell, proposed the following “metaphysical principles” for mixtures, which are worth considering.

1. All properties of the mixture must be mathematical consequences of properties of the constituents.

2. So as to describe the motion of a constituent, we may in imagination isolate it from the rest of the mixture, provided we allow properly for the actions of the other constituents upon it.

3. The motion of the mixture is governed by the same equations as is a single body.

Most important for the present discussion is the first principle. When coupled with fluid mechanics, the second two take on additional importance. The approach of mixture theory is to divide and conquer. One typically treats each of the constituents as a single material and then devises appropriate average or mixture properties from those of the constituents. The best example of this is air, which is not a single material, but is often treated as such.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.2. IDEAL AND NON-IDEAL MIXTURES 75

3.2 Ideal and non-ideal mixtures

A general extensive property, such as U, for an N-species mixture will be such that

U = U(T,P,n1, n2,...,nN ). (3.36)

A is a generalization of an intensive property, and is defined such that it is the partial derivative of an extensive property with respect to number of moles, with T and P held constant. For internal energy, then the partial molar internal energy is

∂U ui . (3.37) ≡ ∂ni T,P,nj,i=6 j

Pressure and temperature are held constant because those are convenient variables to control in an experiment. One also has the partial

∂V vi = . (3.38) ∂ni T,P,nj,i=6 j

It shall be soon seen that there are other natural ways to think of the volume per mole. Now in general one would expect to find

ui = ui(T,P,n1, n2,...,nN ), (3.39)

vi = vi(T,P,n1, n2,...,nN ). (3.40)

This is the case for what is known as a non-ideal mixture. An ideal mixture is defined as a mixture for which the partial molar properties ui and vi are not functions of the composition, that is

ui = ui(T,P ), if ideal mixture, (3.41)

vi = vi(T,P ), if ideal mixture. (3.42)

An ideal mixture also has the property that hi = hi(T,P ), while for a non-ideal mixture hi = hi(T,P,n1,...,nN ). Though not obvious, it will turn out that some properties of an ideal mixture will depend on composition. For example, the entropy of a constituent of an ideal mixture will be such that

si = si(T,P,n1, n2,...,nN ). (3.43)

3.3 Ideal mixtures of ideal gases

The most straightforward mixture to consider is an ideal mixture of ideal gases. Even here, there are assumptions necessary that remain difficult to verify absolutely.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 76 CHAPTER 3. GAS MIXTURES

3.3.1 Dalton model The most common model for a mixture of ideal gases is the Dalton model. Key assumptions define this model

Each constituent shares a common temperature. • Each constituent occupies the entire volume. • Each constituent possesses a which sums to form the total pressure of • the mixture.

The above characterize a Dalton model for any gas, ideal or non-ideal. One also takes for convenience

Each constituent behaves as an ideal gas. • The mixture behaves as a single ideal gas. • It is more convenient to deal on a molar basis for such a theory. For the Dalton model, additional useful quantities, the species mass concentration ρi, the mixture mass concentra- tion ρ, the species molar concentration ρi, and the mixture molar concentration ρ, can be defined. As will be seen, these definitions for concentrations are useful; however, they are not in common usage. Following BS, the bar notation, , will be reserved for properties which · are mole-based rather than mass-based. As mentioned earlier, the notion of a partial molal property is discussed extensively in the chemical engineering literature and has implications beyond those considered here. For the Dalton model, in which each component occupies the same volume, one has Vi = V. (3.44) The mixture mass concentration, also called the density is simply

m kg ρ = , . (3.45) V m3   The mixture molar concentration is n kmole ρ = , . (3.46) V m3   For species i, the equivalents are

m kg ρ = i , , (3.47) i V m3   n kmole ρ = i , . (3.48) i V m3   CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.3. IDEAL MIXTURES OF IDEAL GASES 77

One can find a convenient relation between species molar concentration and species mole fraction by the following operations n n ρ = i , (3.49) i V n n n = i , (3.50) n V = yiρ. (3.51) A similar relation exists between species molar concentration and species mass fraction via

ni m Mi ρi = , (3.52) V m Mi =mi m n M = i i , (3.53) V zmM}| {i =ci m 1 = ρ i , (3.54) m Mi z}|{c = ρ i . (3.55) Mi The specific volumes, mass and molar, are similar. One takes V V v = , v = , (3.56) m n V V vi = , vi = . (3.57) mi ni Note that this definition of molar specific volume is not the partial molar volume defined in the chemical engineering literature, which takes the form v = ∂V/∂n . i i|T,P,nj,i=6 j For the partial pressure of species i, one can say for the Dalton model

N

P = Pi. (3.58) i=1 X For species i, one has

PiV = niRT, (3.59) n RT P = i , (3.60) i V N N n RT P = i , (3.61) i V i=1 i=1 X X =P RT N | {zP} = n . (3.62) V i i=1 X =n

CC| {z BY-NC-ND.} 28 January 2019, J. M. Powers. 78 CHAPTER 3. GAS MIXTURES

So, for the mixture, one has

PV = nRT. (3.63)

One could also say

n P = RT = ρRT. (3.64) V =ρ |{z} Here n is the total number of moles in the system. Additionally R is the universal with value

kJ J R =8.314472 =8.314472 . (3.65) kmole K mole K

Sometimes this is expressed in terms of k , the , and , Avogadro’s B N number:

R = k , (3.66) BN molecule = 6.02214179 1023 , (3.67) N × mole J k = 1.380650 10−23 . (3.68) B × K molecule

Example 3.2 Compare the molar specific volume defined here with the partial molar volume from the chemical engineering literature.

The partial molar volume vi, is given by

∂V vi = . (3.69) ∂ni T,P,nj ,i=6 j

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.3. IDEAL MIXTURES OF IDEAL GASES 79

For the ideal gas, one has

N P V = RT nk, (3.70) Xk=1 RT N n V = k=1 k , (3.71) P PN ∂n ∂V RT k = k=1 ∂ni , (3.72) ∂ni P T,P,nj ,i=6 j P

RT N δ = k=1 ki , (3.73) P P =0 =0 =1 =0 RT δ + δ + + δ + + δ  1i 2i ··· ii ··· Ni  = z}|{ z}|{ z}|{ z}|{ , (3.74)  P  RT v = , (3.75) i P V = N , (3.76) k=1 nk V = P. (3.77) n Here the so-called Kronecker delta function has been employed, which is much the same as the identity matrix:

δ = 0, k = i, (3.78) ki 6 δki = 1, k = i. (3.79) Contrast this with the earlier adopted definition of molar specific volume V vi = . (3.80) ni So, why is there a difference? The molar specific volume is a simple definition. One takes the instantaneous volume V , which is shared by all species in the Dalton model, and scales it by the instantaneous number of moles of species i, and acquires a natural definition of molar specific volume consistent with the notion of a mass specific volume. On the other hand, the partial molar volume specifies how the volume changes if the number of moles of species i changes, while holding T and P and all other species mole numbers constant. One can imagine adding a mole of species i, which would necessitate a change in V in order to guarantee the P remain fixed.

3.3.1.1 Binary mixtures Consider now a binary mixture of two components A and B. This is easily extended to a general mixture of N components. First the total number of moles is the sum of the parts:

n = nA + nB. (3.81)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 80 CHAPTER 3. GAS MIXTURES

Now, write the ideal gas law for each component:

PAVA = nARTA, (3.82)

PBVB = nBRTB. (3.83)

But by the assumptions of the Dalton model, VA = VB = V , and TA = TB = T , so

PAV = nART, (3.84)

PBV = nBRT. (3.85)

One also has PV = nRT. (3.86)

Solving for n, nA and nB, one finds PV n = , (3.87) RT PAV nA = , (3.88) RT PBV nB = . (3.89) RT

Now n = nA + nB, so one has PV P V P V = A + B . (3.90) RT RT RT P = PA + PB. (3.91)

That is the total pressure is the sum of the partial pressures. This is a mixture rule for pressure One can also scale each constituent ideal gas law by the mixture ideal gas law to get

P V n RT A = A , (3.92) PV nRT P n A = A , (3.93) P n = yA, (3.94)

PA = yAP. (3.95)

Likewise

PB = yBP. (3.96) Now, one also desires rational mixture rules for energy, enthalpy, and entropy. Invoke Trues- dell’s principles on a mass basis for internal energy. Then the total internal energy U (with

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.3. IDEAL MIXTURES OF IDEAL GASES 81

units J) for the binary mixture must be

U = mu = mAuA + mBuB, (3.97) m m = m A u + B u , (3.98) m A m B = m (cAuA + cBuB) ,  (3.99)

u = cAuA + cBuB. (3.100)

For the enthalpy, one has

H = mh = mAhA + mBhB, (3.101) m m = m A h + B h , (3.102) m A m B = m (cAhA + cBhB) ,  (3.103)

h = cAhA + cBhB. (3.104)

It is easy to extend this to a mole fraction basis rather than a mass fraction basis. One can also obtain a gas constant for the mixture on a mass basis. For the mixture, one has

PV = nRT mRT, (3.105) ≡ PV mR = nR, (3.106) T ≡ = (nA + nB)R, (3.107) m m = A + B R, (3.108) M M  A B  R R = m + m , (3.109) A M B M  A B  = (mARA + mBRB) , (3.110) m m R = A R + B R , (3.111) m A m B R = (cARA + cBRB) .  (3.112)

For the entropy, one has

S = ms = = mAsA + mBsB, (3.113) m m = m A s + B s , (3.114) m A m B = m (cAsA + cBsB) ,  (3.115)

s = cAsA + cBsB. (3.116)

Note that sA is evaluated at T and PA, while sB is evaluated at T and PB. For a CPIG, one

CC BY-NC-ND. 28 January 2019, J. M. Powers. 82 CHAPTER 3. GAS MIXTURES has T P s = so + c ln R ln A , (3.117) A 298,A PA T − A P  o   o  o ≡sT,A T y P = s|o + c{z ln } R ln A . (3.118) 298,A PA T − A P  o   o  Likewise T y P s = so + c ln R ln B . (3.119) B 298,B P B T − B P  o   o  o ≡sT,B Here the “o” denotes some reference| state.{z As a} superscript, it typically means that the o property is evaluated at a reference pressure. For example, sT,A denotes the portion of the entropy of component A that is evaluated at the reference pressure Po and is allowed to vary with temperature T . Note also that sA = sA(T,P,yA) and sB = sB(T,P,yB), so the entropy of a single constituent depends on the composition of the mixture and not just on T and P . This contrasts with energy and enthalpy for which uA = uA(T ), uB = uB(T ), hA = hA(T ), o o hB = hB(T ) if the mixture is composed of ideal gases. Occasionally, one finds hA and hB used as a notation. This denotes that the enthalpy is evaluated at the reference pressure. o o However, if the gas is ideal, the enthalpy is not a function of pressure and hA = hA, hB = hB. If one is employing a calorically imperfect ideal gas model, then one finds for species i that y P s = so R ln i , i = A, B. (3.120) i T,i − i P  o  3.3.1.2

Example 3.3 Initially calorically perfect ideal gases A and B are segregated within the same large volume by a thin frictionless, thermally conducting diaphragm. Thus, both are at the same initial pressure and temperature, P1 and T1. The total volume is thermally insulated and fixed, so there are no global heat or work exchanges with the environment. The diaphragm is removed, and A and B are allowed to mix. Assume A has mass mA and B has mass mB. The gases are allowed to have distinct molecular masses, MA and MB. Find the final temperature T2, pressure P2, and the change in entropy.

The ideal gas law holds that at the initial state

mARAT1 mBRBT1 VA1 = , VB1 = . (3.121) P1 P1 At the final state one has

T1 V2 = VA2 = VB2 = VA1 + VB1 = (mARA + mBRB) . (3.122) P1

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.3. IDEAL MIXTURES OF IDEAL GASES 83

Mass conservation gives m2 = m1 = mA + mB. (3.123) One also has the first law

U U = Q W , (3.124) 2 − 1 1 2 − 1 2 U U = 0, (3.125) 2 − 1 U2 = U1, (3.126)

m2u2 = mAuA1 + mBuB1, (3.127)

(mA + mB)u2 = mAuA1 + mBuB1, (3.128) 0 = m (u u )+ m (u u ), (3.129) A A1 − 2 B B1 − 2 0 = m c (T T )+ m c (T T ), (3.130) A vA 1 − 2 B vB 1 − 2 mAcvAT1 + mBcvBT1 T2 = , (3.131) mAcvA + mBcvB

T2 = T1. (3.132)

The final pressure by Dalton’s law then is

P2 = PA2 + PB2, (3.133) m R T m R T = A A 2 + B B 2 , (3.134) V2 V2 m R T m R T = A A 1 + B B 1 , (3.135) V2 V2 (m R + m R ) T = A A B B 1 , (3.136) V2 substitute for V2 fromEq.(3.122) (3.137) (m R + m R ) T = A A B B 1 , (3.138) T1 (mARA + mBRB) P1

= P1. (3.139)

So the initial and final temperatures and pressures are identical. Now the entropy change of gas A is

T P s s = c ln A2 R ln A2 , (3.140) A2 − A1 PA T − A P  A1   A1  T y P = c ln 2 R ln A2 2 , (3.141) PA T − A y P  1   A1 1  T y P = c ln 1 R ln A2 1 , (3.142) PA T − A y P  1   A1 1  =0 y P = R |ln{z A}2 1 , (3.143) − A (1)P  1  = R ln y . (3.144) − A A2 Likewise

s s = R ln y . (3.145) B2 − B1 − B B2

CC BY-NC-ND. 28 January 2019, J. M. Powers. 84 CHAPTER 3. GAS MIXTURES

So the change in entropy of the mixture is ∆S = m (s s )+ m (s s ), (3.146) A A2 − A1 B B2 − B1 = m R ln y m R ln y , (3.147) − A A A2 − B B B2 R R = (nAMA) ln yA2 (nBMB) ln yB2, (3.148) − MA − MB =mA   =mB   =RA =RB = R| (n{z ln}y + n ln y ), | {z } (3.149) − A |A2 {z }B B2 | {z }

nA nB = R nA ln +nB ln  , (3.150) − nA + nB nA + nB         ≤0 ≤0    0. | {z } | {z } (3.151) ≥ We can also scale Eq. (3.149) by Rn to get

1 ∆S nA nB =  ln yA2 + ln yB2 , (3.152) R n − n n   =∆s =yA2 =yB2  ∆s   |{z} = (yA|{z}2 ln yA2 + yB|{z}2 ln yB2) , (3.153) R − = (ln yyA2 + ln yyB2 ) , (3.154) − A2 B2 = ln (yyA2 yyB2 ) . (3.155) − A2 B2

For an N-component mixture, mixed in the same fashion such that P and T are constant, this extends to N ∆S = R n ln y , (3.156) − k k k=1 XN nk = R nk ln N 0, (3.157) − i=1 ni ! ≥ Xk=1 P≤0 N nk | {z } = R n ln y , (3.158) − n k Xk=1 N n = Rn k ln y , (3.159) − n k Xk=1 m N = R y ln y , (3.160) − M k k Xk=1 CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.3. IDEAL MIXTURES OF IDEAL GASES 85

N = Rm ln yyk , (3.161) − k Xk=1 = Rm (ln yy1 + ln yy2 + + ln yyN ) , (3.162) − 1 2 ··· N = Rm ln (yy1 yy2 ...yyN ) , (3.163) − 1 2 N N yk = Rm ln yk , (3.164) − ! kY=1 Dividing by m to recover an intensive property and R to recover a dimensionless property, we get

N ∆s ∆s yk = = ln yk . (3.165) R R − ! Yk=1 Note that there is a fundamental dependency of the mixing entropy on the mole fractions. Since 0 y 1, the product is guaranteed to be between 0 and 1. The natural logarithm ≤ k ≤ of such a number is negative, and thus the entropy change for the mixture is guaranteed positive semi-definite. Note also that for the entropy of mixing, Truesdell’s third principle is not enforced. Now if one mole of pure N2 is mixed with one mole of pure O2, one certainly expects the resulting homogeneous mixture to have a higher entropy than the two pure components. But what if one mole of pure N2 is mixed with another mole of pure N2. Then we would expect no increase in entropy. However, if we had the unusual ability to distinguish N2 molecules whose origin was from each respective original chamber, then indeed there would be an entropy of mixing. Increases in entropy thus do correspond to increases in disorder.

3.3.1.3 Mixtures of constant mass fraction If the mass fractions, and thus the mole fractions, remain constant during a process, the equations simplify. This is often the case for common non-reacting mixtures. Air at moderate values of temperature and pressure behaves this way. In this case, all of Truesdell’s principles can be enforced. For a CPIG, one would have

u u = c c (T T )+ c c (T T ), (3.166) 2 − 1 A vA 2 − 1 B vB 2 − 1 = c (T T ). (3.167) v 2 − 1 where c c c + c c . (3.168) v ≡ A vA B vB Similarly for enthalpy

h h = c c (T T )+ c c (T T ), (3.169) 2 − 1 A PA 2 − 1 B P B 2 − 1 = c (T T ). (3.170) P 2 − 1

CC BY-NC-ND. 28 January 2019, J. M. Powers. 86 CHAPTER 3. GAS MIXTURES

where c c c + c c . (3.171) P ≡ A PA B P B For the entropy

s s = c (s s )+ c (s s ), (3.172) 2 − 1 A A2 − A1 B B2 − B1 T y P T y P = c c ln 2 R ln A 2 + c c ln 2 R ln B 2 , A PA T − A y P B P B T − B y P   1   A 1    1   B 1  T P T P = c c ln 2 R ln 2 + c c ln 2 R ln 2 ,(3.173) A PA T − A P B P B T − B P   1   1    1   1  T P = c ln 2 R ln 2 . (3.174) P T − P  1   1  The mixture behaves as a pure substance when the appropriate mixture properties are de- fined. One can also take c k = P . (3.175) cv

3.3.2 Summary of properties for the Dalton mixture model Listed here is a summary of mixture properties for an N-component mixture of ideal gases on a mass basis:

N

M = yiMi, (3.176) i=1 XN

ρ = ρi, (3.177) i=1 X1 1 v = = , (3.178) N 1 ρ i=1 vi N P u = ciui, (3.179) i=1 XN

h = cihi, (3.180) i=1 X =R R N N y M R R N R = = c R = i i i = y , (3.181) M i i N M i i=1 i=1 z }| { i=1 X X yjMj X =1 j=1 X =M | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. | {z } 3.3. IDEAL MIXTURES OF IDEAL GASES 87

N

cv = cicvi, (3.182) i=1 X c = c R, if ideal gas, (3.183) v P − N

cP = cicPi, (3.184) i=1 Xc k = P , (3.185) cv N

s = cisi, (3.186) i=1 X y M c = i i , (3.187) i M Pi = yiP, (3.188)

ρi = ciρ, (3.189) v 1 vi = = , (3.190) ci ρi V = Vi, (3.191)

T = Ti, (3.192) o hi = hi , if ideal gas, (3.193) Pi hi = ui + = ui + Pivi = ui + RiT , (3.194) ρi if ideal gas T o ˆ ˆ | {z } hi = h298,i + cPi(T ) dT , (3.195) Z298 if ideal gas T ˆ |o {zcPi(T ) }ˆ Pi si = s298,i + dT Ri ln , (3.196) 298 Tˆ − Po Z   o =sT,i | {z if ideal gas} y P P s = s|o R ln i {z = so R ln } i , (3.197) i T,i − i P T,i − i P  o   o  if ideal gas if ideal gas

RiT Pi = |ρiRiT ={zρRiTci =} | , {z } (3.198) vi if ideal gas | {z }

CC BY-NC-ND. 28 January 2019, J. M. Powers. 88 CHAPTER 3. GAS MIXTURES

N c RT P = ρRT = ρRT i = , (3.199) M v i=1 i X if ideal gas N T | o {z ˆ ˆ} h = cih298,i + cP (T ) dT , (3.200) i=1 298 X Z if ideal gas P h = u| + = u +{zP v = u + RT} , (3.201) ρ if ideal gas N T ˆ N cP (|T ){z } P y s = c so + dTˆ R ln R ln y i . i 298,i ˆ − P − i i=1 298 T o i=1 ! X Z   Y if ideal gas (3.202) | {z } These relations are not obvious. A few are derived in examples here.

Example 3.4 Derive the expression h = u + P/ρ.

Start from the equation for the constituent hi, multiply by mass fractions, sum over all species, and use properties of mixtures:

Pi hi = ui + , (3.203) ρi Pi cihi = ciui + ci , (3.204) ρi N N N P c h = c u + c i , (3.205) i i i i i ρ i=1 i=1 i=1 i X X X N N N ρ R T c h = c u + c i i , (3.206) i i i i i ρ i=1 i=1 i=1 i X X X =h =u N | {z } | {z } h = u + T ciRi, (3.207) i=1 X =R = u + RT, (3.208) | {z } P = u + . (3.209) ρ

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.3. IDEAL MIXTURES OF IDEAL GASES 89

Example 3.5 Find the expression for mixture entropy of the ideal gas.

T ˆ o cP i(T ) ˆ Pi si = s298,i + dT Ri ln , (3.210) 298 Tˆ − Po Z   T ˆ o cP i(T ) ˆ Pi cisi = cis298,i + ci dT ciRi ln , (3.211) 298 Tˆ − Po Z   N N N N T c (Tˆ) P s = c s = c so + c P i dTˆ c R ln i , (3.212) i i i 298,i i ˆ − i i P i=1 i=1 i=1 298 T i=1 o X X X Z X   N N N T c c (Tˆ) P = c so + i P i dTˆ c R ln i , (3.213) i 298,i ˆ − i i P i=1 298 i=1 T i=1 o X Z X X   T c (Tˆ) N P = so + P dTˆ c R ln i . (3.214) 298 ˆ − i i P 298 T i=1 o Z X   All except the last term are natural extensions of the property for a single material. Consider now the last term involving pressure ratios.

N N P P P P c R ln i = c R ln i + R ln R ln , (3.215) − i i P − i i P P − P i=1 o i=1 o o o ! X   X   N R P P P = R c i ln i + ln ln , (3.216) − i R P P − P i=1 o o o ! X   N R/M P P P = R c i ln i + ln ln , (3.217) − i N P P − P i=1 j=1 cj R/Mj o o o ! X   P N c /M P P P = R  i i ln i + ln ln  , (3.218) − N P P − P i=1 j=1 cj /Mj !  o  o o  X     P =yi   N  | P{z }P P = R y ln i + ln ln , (3.219) − i P P − P i=1 o o o ! X   N P yi P P = R ln i ln + ln , (3.220) − P − P P i=1 o o o ! X   N P yi P P = R ln i + ln o + ln , (3.221) − P P P i=1 o ! o ! Y   N P P yi P = R ln o i + ln , (3.222) − P P P i=1 o ! o ! Y   N Po 1 yi P = R ln N N (Pi) + ln , (3.223) − Pi=1 yi Pi=1 yi P P P i=1 ! o ! o Y CC BY-NC-ND. 28 January 2019, J. M. Powers. 90 CHAPTER 3. GAS MIXTURES

N P yi P = R ln i + ln , (3.224) − P P i=1 ! o ! Y   N y P yi P = R ln i + ln , (3.225) − P P i=1 ! o ! Y   N P = R ln yyi + ln . (3.226) − i P i=1 ! o ! Y So the mixture entropy becomes

T c (Tˆ) N P s = so + P dTˆ R ln yyi + ln , (3.227) 298 ˆ − i P 298 T i=1 ! o ! Z Y N T c (Tˆ) P = so + P dTˆ R ln R ln yyi . (3.228) 298 ˆ − P − i 298 T o i=1 ! Z Y classical entropy of a single body non−Truesdellian | {z } The extra entropy is not found in the theory for a single material,| and{z in fact} is not in the form suggested by Truesdell’s postulates. While it is in fact possible to redefine the constituent entropy definition in such a fashion that the mixture entropy in fact takes on the classical form of a single material via the ˆ o T cP i(T ) Pi definition si = s + dTˆ Ri ln + Ri ln yi, this has the disadvantage of predicting 298,i 298 Tˆ − Po no entropy change afterR mixing two pure substances.  Such a theory would suggest that this obviously is in fact reversible.

On a molar basis, one has the equivalents

N n ρ ρ = ρ = = , (3.229) i V M i=1 X 1 V 1 v = = = = vM, (3.230) N 1 n ρ i=1 vi N P u = yiui = uM, (3.231) i=1 XN

h = yihi = hM, (3.232) i=1 X

N

cv = yicvi = cvM, (3.233) i=1 X c = c R, (3.234) v P − N

cP = yicPi = cP M, (3.235) i=1 X CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.3. IDEAL MIXTURES OF IDEAL GASES 91

c k = P , (3.236) cv N

s = yisi = sM, (3.237) i=1 X ρi ρi = yiρ = , (3.238) Mi V v 1 vi = = = = viMi, (3.239) ni yi ρi ∂V V vi = = = v = vM, (3.240) ∂ni n P,T,nj if ideal gas

P = y P, (3.241) i i | {z } RT P = ρRT = , (3.242) v if ideal gas | {z RT} Pi = ρiRT = , (3.243) vi if ideal gas P h = |u + {z= u }+ P v = u + RT = hM, (3.244) ρ if ideal gas o hi = hi , if ideal gas, | {z } (3.245) Pi hi = ui + = ui + Pivi = ui + P vi = ui + RT = hiMi, (3.246) ρi if ideal gas T o ˆ ˆ | {z } hi = h298,i + cPi(T ) dT = hiMi, (3.247) Z298 if ideal gas T ˆ |o {zcPi(T ) }ˆ yiP si = s298,i + dT R ln , (3.248) 298 Tˆ − Po Z   o =sT,i | {z if ideal gas} y P s = |so R ln i {z= s M , } (3.249) i T,i − P i i  o  if ideal gas N T ˆ N | {z }cP (T ) P y s = y so + dTˆ R ln R ln y i = sM. i 298,i ˆ − P − i i=1 298 T o i=1 ! X Z   Y if ideal gas (3.250) | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. 92 CHAPTER 3. GAS MIXTURES

3.3.3 Amagat model The Amagat model is an entirely different paradigm than the Dalton model. It is not used as often. In the Amagat model, all components share a common temperature T , • all components share a common pressure P , and • each component has a different volume. • Consider, for example, a binary mixture of calorically perfect ideal gases, A and B. For the mixture, one has PV = nRT, (3.251) with n = nA + nB. (3.252) For the components one has

PVA = nART, (3.253)

PVB = nBRT. (3.254)

Then n = nA + nB reduces to PV PV PV = A + B . (3.255) RT RT RT Thus

V = VA + VB, (3.256) V V 1 = A + B . (3.257) V V 3.4 Gas-vapor mixtures

Next consider a mixture of ideal gases in which one of the components may undergo a to its liquid state. The most important practical example is an air-water mixture. Assume the following: The or liquid contains no dissolved gases. • The gaseous phases are all well modeled as ideal gases. • When the gas mixture and the condensed phase are at a given total pressure and tem- • perature, the equilibrium between the condensed phase and its vapor is not influenced by the other component. So for a binary mixture of A and B where A could have both gas and liquid components PA = Psat. That is the partial pressure of A is equal to its saturation pressure at the appropriate temperature.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.4. GAS-VAPOR MIXTURES 93

Considering an air mixture, one models the water vapor as an ideal gas and expresses the total pressure as

P = Pa + Pv. (3.258)

Here v denotes vapor and a denotes air. A good model for the enthalpy of the water vapor is to take

hv(T, lowP )= hg(T ). (3.259)

If T is given in degrees Celsius, a good model from the steam tables is

kJ kJ h (T ) = 2501.3 + 1.82 T. (3.260) g kg kg ◦C  

Some definitions:

absolute : ω, the mass of water present in a unit mass of dry air, also called • humidity ratio,

m ω v , (3.261) ≡ ma M n = v v , (3.262) Mana M PvV = v RT , (3.263) M PaV a RT M P = v v , (3.264) MaPa kg 18.015 kmole Pv = kg , (3.265) 28.97 kmole Pa P = 0.622 v ,  (3.266) Pa P = 0.622 v . (3.267) P P − v

dew point: temperature at which the vapor condenses when it is cooled isobarically. •

saturated air: The vapor in the air-vapor mixture is at the saturation temperature and • pressure.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 94 CHAPTER 3. GAS MIXTURES

relative humidity: The ratio of the mole fraction of the vapor in the mixture to the mole • fraction of vapor in a saturated mixture at the same temperature and total pressure:

nv n φ ng , (3.268) ≡ n n = v , (3.269) ng PvV = RT , (3.270) Pg V RT P = v . (3.271) Pg

Note, here the subscript g denotes saturated gas values. Combining, one can relate the relative humidity to the absolute humidity: ωP φ = a . (3.272) 0.622Pg

Example 3.6 (from C¸engal and Boles, p. 670) A 5 m 5 m 3 m room contains air at 25 ◦C and 100 kPa at a relative humidity of 75%. Find the × × partial pressure of dry air, • absolute humidity (i.e. humidity ratio), • masses of dry air and water vapor in the room, and the • dew point. • The relation between partial and total pressure is

P = Pa + Pv. (3.273)

Now from the definition of relative humidity,

Pv = φPg . (3.274)

◦ ◦ Here Pg is the saturation pressure at the same temperature, which is 25 C. At 25 C, the tables give

P ◦ =3.169 kP a. (3.275) g|25 C So Pv =0.75(3.169 kPa)=2.3675 kP a. (3.276) So from the definition of partial pressure

P = P P , (3.277) a − v = 100 kPa 2.3675 kP a, (3.278) − = 97.62 kP a. (3.279)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.4. GAS-VAPOR MIXTURES 95

Now for the absolute humidity (or specific humidity), one has

P ω = 0.622 v , (3.280) Pa 2.3675 kPa = 0.622 , (3.281) 97.62 kPa kg H O = 0.0152 2 . (3.282) kg dry air

Now for the masses of air and water, one can use the partial pressures:

PaV ma = Ma, (3.283) RT P V = a , (3.284) RaT (97.62 kPa) 75 m3 = kJ , (3.285) 8.314 kmole K kg (298 K ) 28.97 kmole   = 85.61 kg. (3.286)

PvV mv = Mv, (3.287) RT P V = v , (3.288) RvT (2.3675 kPa) 75 m3 = kJ , (3.289) 8.314 kmole K kg (298 K) 18.015 kmole   = 1.3 kg. (3.290)

Also one could get mv from

mv = ωma, (3.291) = (0.0152)(85.61 kg), (3.292) = 1.3 kg. (3.293)

Now the dew point is the saturation temperature at the partial pressure of the water vapor. With Pv =2.3675 kPa, the saturation tables give

◦ Tdew point = 20.08 C. (3.294)

3.4.1 First law The first law can be applied to air-water mixtures.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 96 CHAPTER 3. GAS MIXTURES

Example 3.7 (BS, Ex. 13.5, pp. 536-537.) An air-water vapor mixture enters the cooling coils of an air conditioner ◦ ◦ unit. The inlet is at P1 = 105 kPa, T1 = 30 C, φ1 =0.80. The exit state is P2 = 100 kPa, T2 = 15 C, ◦ φ2 =0.95. Liquid water at 15 C also exits the system. Find the heat transfer per kilogram of dry air.

Mass conservation for air and water give

m˙ =m ˙ m˙ , (3.295) a1 a2 ≡ a m˙ v1 =m ˙ v2 +m ˙ l2. (3.296)

Note at state 2, the mass flow of water is in both liquid and vapor form. The first law for the control volume give dE cv = Q˙ W˙ +m ˙ h +m ˙ h m˙ h m˙ h m˙ h , (3.297) dt cv − cv a a1 v1 v1 − a a2 − v2 v2 − l2 l2 =0 =0 Q˙ +m ˙ h +m ˙ h =m ˙ h |{z}+m ˙ h + (m ˙ m˙ )h , (3.298) cv a a1 |v1{zv1} a a2 v2 v2 v1 − v2 l2 Q˙ cv m˙ v1 m˙ v2 m˙ v1 m˙ v2 + ha1 + hv1 = ha2 + hv2 + − hl2, (3.299) m˙ a m˙ a m˙ a m˙ a

Q˙ cv + ha1 + ω1hv1 = ha2 + ω2hv2 + (ω1 ω2)hl2, (3.300) m˙ a −

Q˙ cv = ha2 ha1 ω1hv1 + ω2hv2 + (ω1 ω2)hl2, (3.301) m˙ a − − − = c (T T ) ω h + ω h + (ω ω )h . (3.302) pa 2 − 1 − 1 v1 2 v2 1 − 2 l2 Now at the inlet, one has from the definition of relative humidity

Pv1 φ1 = . (3.303) Pg1

◦ Here Pg is the saturated vapor pressure at the inlet temperature, T1 = 30 C. This is Pg1 =4.246 kPa. So one gets

Pv1 = φ1Pg1, (3.304) = (0.80)(4.246 kPa), (3.305) = 3.397 kP a. (3.306)

Now the absolute humidity (humidity ratio) is P ω = 0.622 v1 , (3.307) 1 P P 1 − v1 3.397 kPa = 0.622 , (3.308) 105 kPa 3.397 kPa − = 0.0208. (3.309)

At the exit temperature, the saturation pressure is Pg2 =1.705 kPa. So

Pv2 = φ2Pg2, (3.310) = (0.95)(1.705 kPa), (3.311) = 1.620 kP a. (3.312)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.4. GAS-VAPOR MIXTURES 97

Now the absolute humidity (humidity ratio) is P ω = 0.622 v2 , (3.313) 2 P P 2 − v2 1.620 kPa = 0.622 , (3.314) 100 kPa 1.620 kPa − = 0.0102. (3.315)

Then, substituting, one gets

Q˙ kJ kJ cv = 1.004 (15 ◦C 30 ◦C) 0.0208 2556.3 m˙ kg K − − kg a     kJ kJ +0.0102 2528.9 + (0.0208 0.0102) 62.99 , (3.316) kg − kg     kJ = 41.77 . (3.317) − kg dry air

3.4.2 Adiabatic saturation In an adiabatic saturation process, an air-vapor mixture contacts a body of water in a well insulated . If the initial humidity of the mixture is less than 100%, some water will evaporate and join the mixture. If the mixture leaving the duct is saturated, and the process is adiabatic, the exit temper- ature is the adiabatic saturation temperature. Assume the liquid water entering the system enters at the exit temperature of the mixture. Mass conservation for air and water and the first law for the control volume give dmair cv =m ˙ m˙ , air (3.318) dt a1 − a2 =0 dmwater | cv{z } =m ˙ +m ˙ m˙ , water (3.319) dt v1 l − v2 =0 dE | {z cv} = Q˙ W˙ +m ˙ h +m ˙ h +m ˙ h m˙ h m˙ h . (3.320) dt cv − cv a1 a1 v1 v1 l l − a2 a2 − v2 v2 =0 =0 =0 |{z} For steady state| {z results,} |{z} these reduce to 0 =m ˙ m˙ , air (3.321) a1 − a2 0 =m ˙ +m ˙ m˙ , water (3.322) v1 l − v2 0 =m ˙ h +m ˙ h +m ˙ h m˙ h m˙ h . (3.323) a1 a1 v1 v1 l l − a2 a2 − v2 v2

CC BY-NC-ND. 28 January 2019, J. M. Powers. 98 CHAPTER 3. GAS MIXTURES

Now mass conservation for air gives

m˙ =m ˙ m˙ . (3.324) a1 a2 ≡ a Mass conservation for water gives

m˙ =m ˙ m˙ . (3.325) l v2 − v1 Then energy conservation becomes

0 =m ˙ h +m ˙ h + (m ˙ m˙ )h m˙ h m˙ h , (3.326) a a1 v1 v1 v2 − v1 l − a a2 − v2 v2 m˙ m˙ m˙ m˙ 0 = h + v1 h + v2 v1 h h v2 h , (3.327) a1 m˙ v1 m˙ − m˙ l − a2 − m˙ v2 a  a a  a 0 = h + ω h +(ω ω ) h h ω h , (3.328) a1 1 v1 2 − 1 l − a2 − 2 v2 0 = h h + ω (h h )+ ω (h h ), (3.329) a1 − a2 1 v1 − l 2 l − v2 ω (h h ) = h h + ω (h h ), (3.330) − 1 v1 − l a1 − a2 2 l − v2 ω (h h ) = h h + ω (h h ), (3.331) 1 v1 − l a2 − a1 2 v2 − l ω (h h ) = c (T T )+ ω h . (3.332) 1 v1 − l pa 2 − 1 2 fg2

Example 3.8 (BS, 13.7E, p. 540). The pressure of the mixture entering and leaving the adiabatic saturater is 14.7 psia, the entering temperature is 84 F , and the temperature leaving is 70 F , which is the adiabatic saturation temperature. Calculate the humidity ratio and the relative humidity of the air-water vapor mixture entering.

The exit state 2 is saturated, so Pv2 = Pg2. (3.333)

The tables give Pv2 = Pg2 =0.363 psia. Thus one can calculate the absolute humidity by its definition:

Pv2 ω2 = 0.622 , (3.334) P2 Pv2 − 0.363 psia = 0.622 , (3.335) (14.7 psia) (0.363 psia) − lbm H O = 0.0157485 2 . (3.336) lbm dry air

The earlier derived result from the energy balance allows calculation then of ω1:

ω (h h ) = c (T T )+ ω h , (3.337) 1 v1 − l pa 2 − 1 2 fg2 c (T T )+ ω h ω = pa 2 − 1 2 fg2 , (3.338) 1 h h v1 − l 0.240 Btu ((70 F ) (84 F ))+0.0157485 1054.0 Btu = lbm F − lbm , (3.339) 1098.1 Btu 38.1 Btu lbm − lbm  lbm H2O   = 0.012895 . (3.340) lbm dry air

CC BY-NC-ND. 28 January 2019, J. M. Powers. 3.4. GAS-VAPOR MIXTURES 99

Here hv1 was estimated as the saturated vapor value at T = 84 F by interpolating the tables. In the absence of more information regarding the initial vapor state, this estimate is good as any. The value of hl is estimated as the saturated liquid value at T = 70 F . Now P ω = 0.622 v1 , (3.341) 1 P P 1 − v1 P1ω1 Pv1 = , (3.342) 0.622 + ω1 (14.7 psia)(0.0124895) = , (3.343) 0.622+0.0124895 = 0.28936 psia. (3.344) For the relative humidity

Pv1 φ1 = , (3.345) Pg1 0.28936 psia = , (3.346) 0.584 psia = 0.495. (3.347)

3.4.3 Wet-bulb and dry-bulb temperatures Humidity is often measured with a psychrometer, which has a wet bulb and dry bulb ther- mometer. The dry bulb measures the air temperature. • The wet bulb measures the temperature of a water soaked thermometer. • If the two temperatures are equal, the air is saturated. If they are different, some of • the water on the web bulb evaporates, cooling the wet bulb thermometer. The evaporative cooling process is commonly modeled (with some error) as an adiabatic • saturation process. These temperatures are also influenced by non-thermodynamic issues such as heat and • mass transfer rates, which induce errors in the device. Capacitance-based electronic devices are often used as an alternative to the traditional • psychrometer.

3.4.4 Psychrometric chart This well-known chart summarizes much of what is important for binary mixtures of air and water, in which the properties depend on three variables, e.g. temperature, pressure, and composition of the mixture.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 100 CHAPTER 3. GAS MIXTURES

CC BY-NC-ND. 28 January 2019, J. M. Powers. Chapter 4

Mathematical foundations of thermodynamics

Read BS, 12.2-12.4, 12.9, 14.1-14.4. Read Abbott and van Ness, Chapter 3. See Vincenti and Kruger, Chapter 3, for more background.

4.1 Exact differentials and state functions

In thermodynamics, one is faced with many systems of the form of the well-known Gibbs equation, Eq. (1.1): du = Tds P dv. (4.1) − This is known to be an exact differential with the consequence that internal energy u is a function of the state of the system and not the details of any process which led to the state. As a counter-example, the work, δw = P dv, (4.2) can be shown to be an inexact differential so that the work is indeed a function of the process involved. Here we use the notation δ to emphasize that this is an inexact differential.

Example 4.1 Show the work is not a . If work were a state function, one might expect it to have the form w = w(P, v), provisional assumption, to be tested. (4.3)

In such a case, one would have the corresponding differential form

∂w ∂w δw = dv + dP. (4.4) ∂v ∂P P v

101 102 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Now since δw = P dv from Newtonian mechanics, one deduces that

∂w = P, (4.5) ∂v P ∂w = 0. (4.6) ∂P v

Integrating Eq. (4.5), one finds w = P v + f(P ), (4.7) where f(P ) is some function of P to be determined. Differentiating Eq. (4.7) with respect to P , one gets ∂w df(P ) = v + . (4.8) ∂P dP v

Now use Eq. (4.6) to eliminate ∂w/∂P in Eq. (4.8) so as to obtain |v df(P ) 0 = v + , (4.9) dP df(P ) = v. (4.10) dP − Equation (4.10) cannot be: a function of P only cannot be a function of v. So, w cannot be a state property: w = w(P, v). (4.11) 6

Consider now the more general form

N ψ dx + ψ dx + + ψ dx = ψ dx . (4.12) 1 1 2 2 ··· N N i i i=1 X

Here ψi and xi, i = 1,...,N, may be thermodynamic variables. This form is known in mathematics as a Pfaff differential form. As formulated, one takes at this stage

x : independent thermodynamic variables, and • i ψ : thermodynamic variables which are functions of x . • i i Now, if the differential in Eq. (4.12), when set to a differential dy, can be integrated to form the function y = y(x1, x2,...,xN ), (4.13) the differential is said to be exact. In such a case, one has

N dy = ψ dx + ψ dx + + ψ dx = ψ dx . (4.14) 1 1 2 2 ··· N N i i i=1 X CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.1. EXACT DIFFERENTIALS AND STATE FUNCTIONS 103

Now, if the algebraic definition of Eq. (4.13) holds, what amounts to the definition of the partial derivative gives the parallel result that

∂y ∂y ∂y dy = dx1 + dx2 + + dxN . (4.15) ∂x1 ∂x2 ··· ∂xN xj ,j=16 xj ,j=26 xj ,j=6 N

Now, combining Eqs. (4.14) and (4.15) to eliminate dy, one gets

∂y ∂y ∂y ψ1dx1 + ψ2dx2 + + ψN dxN = dx1 + dx2 + + dxN . ··· ∂x1 ∂x2 ··· ∂xN xj ,j=16 xj ,j=26 xj ,j=6 N

(4.16)

Rearranging, one gets

∂y ∂y ∂y 0= ψ1 dx1 + ψ2 dx2 + + ψN dxN . ∂x1 − ∂x2 − ··· ∂xN − xj ,j=16 ! xj ,j=26 ! xj ,j=6 N !

(4.17)

Since the variables xi, i =1,...,N, are independent, dxi, i =1,...,N, are all indepen- dent in Eq. (4.17), and in general non-zero. For equality, one must require that each of the coefficients be zero, so

∂y ∂y ∂y ψ1 = , ψ2 = ,..., ψN = . (4.18) ∂x1 ∂x2 ∂xN xj ,j=16 xj ,j=26 xj ,j=6 N

So when dy is exact, one says that each of the ψi and xi are conjugate to each other. From here on out, for notational ease, the j = 1, j = 2,...,j = N will be ignored in the notation for the partial derivatives. It becomes6 especially6 con6 fusing for higher order derivatives, and is fairly obvious for all derivatives. If y and all its derivatives are continuous and differentiable, then one has for all i = 1,...,N, and k =1,...,N that ∂2y ∂2y = . (4.19) ∂xk∂xi ∂xi∂xk Now from Eq. (4.18), one has

∂y ∂y ψk = , ψl = . (4.20) ∂xk ∂xl xj xj

Taking the partial of the first of Eq. (4.20) with respect to xl and the second with respect to xk, one gets ∂ψ ∂2y ∂ψ ∂2y k = , l = . (4.21) ∂xl ∂xl∂xk ∂xk ∂xk∂xl xj xj

CC BY-NC-ND. 28 January 2019, J. M. Powers. 104 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Since by Eq. (4.19) the order of the mixed second partials does not matter, one deduces from Eq. (4.21) that ∂ψ ∂ψ k = l . (4.22) ∂xl ∂xk xj xj

This is a necessary and sufficient condition for the exact-ness of Eq. (4.12). It is a gen- eralization of what can be found in most introductory calculus texts for functions of two variables. For the Gibbs equation, (4.1), du = P dv + Tds, one has − y u, x v, x s, ψ P ψ T. (4.23) → 1 → 2 → 1 → − 2 → and one expects the natural, or canonical form of

u = u(v,s). (4.24)

Here, P is conjugate to v, and T is conjugate to s. Application of the general form of Eq. (4.22)− to the Gibbs equation (4.1) gives then

∂T ∂P = . (4.25) ∂v − ∂s s v

Equation (4.25) is known as a Maxwell relation . Moreover, specialization of Eq. (4.20) to the Gibbs equation (4.1) gives

∂u ∂u P = , T = . (4.26) − ∂v ∂s s v

N If the general differential dy = i=1 ψidxi is exact, one also can show

The path integral y y = PB N ψ dx is independent of the path of the integral. • B − A A i=1 i i The integral around a closedR contourP is zero: • N

dy = ψidxi =0. (4.27) i=1 I I X The function y can only be determined to within an additive constant. That is, there • is no absolute value of y; physical significance is only ascribed to differences in y. In fact now, other means, extraneous to this analysis, can be used to provide absolute values of key thermodynamic variables. This will be important especially for flows with reaction.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.1. EXACT DIFFERENTIALS AND STATE FUNCTIONS 105

Example 4.2 Show the heat transfer q is not a state function. Assume all processes are fully reversible. The first law gives

du = δq δw, (4.28) − δq = du + δw, (4.29) = du + P dv. (4.30)

Take now the non-canonical, although acceptable, form u = u(T, v). Then one gets

∂u ∂u du = dv + dT. (4.31) ∂v ∂T T v

So ∂u ∂u δq = dv + dT + P dv, (4.32) ∂v ∂T T v ∂u ∂u = + P dv + dT. (4.33) ∂v ∂T  T  v

≡ M ≡N

= M dv + N dT. (4.34) | {z } | {z } Now by Eq. (4.22), for δq to be exact, one must have

∂M ∂N = , (4.35) ∂T ∂v v T

(4.36)

This reduces to ∂2u ∂P ∂2u + = . (4.37) ∂T∂v ∂T ∂v∂T v

This can only be true if ∂P = 0. But this is not the case; consider an ideal gas for which ∂P = R/v. ∂T v ∂T v So δq is not exact.

Example 4.3 Show conditions for ds to be exact in the Gibbs equation.

du = T ds P dv, (4.38) − du P ds = + dv, (4.39) T T 1 ∂u ∂u P = dv + dT + dv, (4.40) T ∂v ∂T T  T v  1 ∂u P 1 ∂u = + dv + dT. (4.41) T ∂v T T ∂T  T  v

≡ M ≡N

| {z } | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. 106 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Again, invoking Eq. (4.22), one gets then ∂ 1 ∂u P ∂ 1 ∂u + = , (4.42) ∂T T ∂v T ∂v T ∂T  T v  vT 2 2 1 ∂ u 1 ∂u 1 ∂P P 1 ∂ u + = , (4.43) T ∂T∂v − T 2 ∂v T ∂T − T 2 T ∂v∂T T v 1 ∂u 1 ∂P P + = 0. (4.44) −T 2 ∂v T ∂T − T 2 T v

This is the condition for an exact ds . Experiment can show if it is true. For example, for an ideal gas, one finds from experiment that u = u(T ) and P v = RT , so one gets 1 R 1 RT 0+ = 0, (4.45) T v − T 2 v 0 = 0. (4.46) So ds is exact for an ideal gas. In fact, the relation is verified for so many gases, ideal and non-ideal, that one simply asserts that ds is exact, rendering s to be path-independent and a state variable.

4.2 Two independent variables

Consider a general implicit function linking three variables, x, y, z: f(x, y, z)=0. (4.47) In x y z space, this will represent a surface. If the function can be inverted, it will be possible− to− write the explicit forms x = x(y, z), y = y(x, z), z = z(x, y). (4.48) Differentiating the first two of the Eqs. (4.48) gives ∂x ∂x dx = dy + dz, (4.49) ∂y ∂z z y

∂y ∂y dy = dx + dz. (4.50) ∂x ∂z z x

Now use Eq. (4.50) to eliminate dy in Eq. (4.49):

∂x ∂y ∂y ∂x dx = dx + dz + dz, (4.51) ∂y ∂x ∂z ∂z z  z x  y

=dy

∂x ∂y ∂x ∂y ∂x 1 dx = | + {z dz, } (4.52) − ∂y ∂x ∂y ∂z ∂z  z z z x y!

∂x ∂y ∂x ∂y ∂x 0dx +0dz = 1 dx + + dz. (4.53) ∂y ∂x − ∂y ∂z ∂z  z z  z x y!

=0 =0

| {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. | {z } 4.2. TWO INDEPENDENT VARIABLES 107

Since x and z are independent, so are dx and dz, and the coefficients on each in Eq. (4.53) must be zero. Therefore from the coefficient on dx in Eq. (4.53)

∂x ∂y 1 = 0, (4.54) ∂y ∂x − z z ∂x ∂y = 1. (4.55) ∂y ∂x z z

So ∂x 1 = , (4.56) ∂y ∂y z ∂x z

and also from the coefficient on dz in Eq. (4.53)

∂x ∂y ∂x + = 0, (4.57) ∂y ∂z ∂z z x y

∂x ∂x ∂y = ,. (4.58) ∂z − ∂y ∂z y z x

So ∂x ∂y ∂z = 1. (4.59) ∂z ∂x ∂y − y z x

If one now divides Eq. (4.49) by a fourth differential, dw, one gets

dx ∂x dy ∂x dz = + . (4.60) dw ∂y dw ∂z dw z y

Demanding that z be held constant in Eq. (4.60) gives ∂x ∂x ∂y = , (4.61) ∂w ∂y ∂w z z z ∂x ∂w ∂x z = , (4.62) ∂y ∂y ∂w z z

∂x ∂w ∂x = . (4.63) ∂w ∂y ∂y z z z

If x = x(y,w), one then gets ∂x ∂x dx = dy + dw. (4.64) ∂y ∂w w y

CC BY-NC-ND. 28 January 2019, J. M. Powers. 108 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Divide now by dy while holding z constant so

∂x ∂x ∂x ∂w = + . (4.65) ∂y ∂y ∂w ∂y z w y z

These general operations can be applied to a wide variety of thermodynamic operations.

Example 4.4 Apply Eq. (4.65) to a standard P v T system and let − − ∂x ∂T = . (4.66) ∂y ∂v z s

So T = x, v = y, and s = z. Let now u = w. So Eq. (4.65) becomes ∂T ∂T ∂T ∂u = + . (4.67) ∂v ∂v ∂u ∂v s u v s

Now by definition ∂u c = , (4.68) v ∂T v

so ∂T 1 = . (4.69) ∂u c v v

Now by Eq. (4.26), one has ∂u = P , so one gets ∂v s − ∂T ∂T P = . (4.70) ∂v ∂v − c s u v

∂T For an ideal gas, u = u(T ). Inverting, one gets T = T (u), and so ∂v u = 0, thus ∂T P = . (4.71) ∂v −c s v

For an in an ideal gas, one gets dT P RT = = , (4.72) dv −cv − cvv dT R dv = , (4.73) T −cv v dv = (k 1) , (4.74) − − v T v ln = (k 1) ln o , (4.75) To − v T v k−1 = o . (4.76) To v  

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.3. LEGENDRE TRANSFORMATIONS 109

4.3 Legendre transformations

The Gibbs equation (4.1), du = P dv + Tds, is the fundamental equation of classical thermodynamics. It is a canonical− form which suggests the most natural set of variables in which to express internal energy u are s and v:

u = u(v,s). (4.77)

However, v and s may not be convenient for a particular problem. There may be other combinations of variables whose canonical form gives a more convenient set of independent variables for a particular problem. An example is the enthalpy:

h u + P v. (4.78) ≡

Differentiating the enthalpy gives

dh = du + P dv + vdP. (4.79)

Use now Eq. (4.79) to eliminate du in the Gibbs equation to give

dh P dv vdP = P dv + T ds. (4.80) − − − =du | {z } So

dh = Tds + vdP. (4.81)

So the canonical variables for h are s and P . One then expects

h = h(s,P ). (4.82)

This exercise can be systematized with the Legendre transformation,1 which defines a set of second order polynomial combinations of variables. Consider again the exact differential Eq. (4.14): dy = ψ dx + ψ dx + + ψ dx . (4.83) 1 1 2 2 ··· N N

1Two differentiable functions f and g are said to be Legendre transformations of each other if their first derivatives are inverse functions of each other: Df = (Dg)−1. With some effort, not shown here, one can prove that the Legendre transformations of this section satisfy this general condition.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 110 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

For N independent variables x and N conjugate variables ψ , by definition there are 2N 1 i i − Legendre transformed variables: τ = τ (ψ , x , x ,...,x )= y ψ x , (4.84) 1 1 1 2 3 N − 1 1 τ2 = τ2(x1, ψ2, x3,...,xN )= y ψ2x2, (4.85) . − . τ = τ (x , x , x ,...,ψ )= y ψ x , (4.86) N N 1 2 3 N − N N τ = τ (ψ , ψ , x ,...,x )= y ψ x ψ x , (4.87) 1,2 1,2 1 2 3 N − 1 1 − 2 2 τ = τ (ψ , x , ψ ,...,x )= y ψ x ψ x , (4.88) 1,3 1,3 1 2 3 N − 1 1 − 3 3 . . N τ = τ (ψ , ψ , ψ ,...,ψ )= y ψ x . (4.89) 1,...,N 1,...,N 1 2 3 N − i i i=1 X Each τ is a new dependent variable. Each τ has the property that when it is known as a function of its N canonical variables, the remaining N variables from the original expression (the xi and the conjugate ψi) can be recovered by differentiation of τ. In general this is not true for arbitrary transformations.

Example 4.5 Let y = y(x1, x2, x3). This has the associated differential form

dy = ψ1dx1 + ψ2dx2 + ψ3dx3. (4.90) Choose now a Legendre transformed variable τ z(ψ , x , x ): 1 ≡ 1 2 3 z = y ψ x . (4.91) − 1 1 Then ∂z ∂z ∂z dz = dψ + dx + dx . (4.92) ∂ψ 1 ∂x 2 ∂x 3 1 x2,x3 2 ψ1,x3 3 ψ1,x2

Now differentiating Eq. (4.91), one also gets

dz = dy ψ dx x dψ . (4.93) − 1 1 − 1 1 Elimination of dy in Eq. (4.93) by using Eq. (4.90) gives dz = ψ dx + ψ dx + ψ dx ψ dx x dψ , (4.94) 1 1 2 2 3 3 − 1 1 − 1 1 =dy = x dψ + ψ dx + ψ dx . (4.95) |− 1 1 {z2 2 3 } 3 Thus from Eq. (4.92), one gets ∂z ∂z ∂z x = , ψ = , ψ = . (4.96) 1 − ∂ψ 2 ∂x 3 ∂x 1 x2,x3 2 ψ1,x3 3 ψ1,x2

So the original expression had three independent variables x1, x2, x3, and three conjugate variables ψ1, ψ2, ψ3. Definition of the Legendre function z with canonical variables ψ1, x2, and x3 allowed determination of the remaining variables x1, ψ2, and ψ3 in terms of the canonical variables.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.3. LEGENDRE TRANSFORMATIONS 111

For the Gibbs equation, (4.1), du = P dv + Tds, one has y = u, two canonical variables, x = v and x = s, and two conjugates,−ψ = P and ψ = T . Thus N = 2, and one can 1 2 1 − 2 expect 22 1 = 3 Legendre transformations. They are − τ = y ψ x = h = h(P,s) = u + P v, enthalpy, (4.97) 1 − 1 1 τ = y ψ x = a = a(v, T ) = u T s, , (4.98) 2 − 2 2 − τ = y ψ x ψ x = g = g(P, T ) = u + P v T s, . 1,2 − 1 1 − 2 2 − (4.99)

It has already been shown for the enthalpy that dh = Tds + vdP , so that the canonical variables are s and P . One then also has ∂h ∂h dh = ds + dP, (4.100) ∂s ∂P P s

from which one deduces that ∂h ∂h T = , v = . (4.101) ∂s ∂P P s

From Eq. (4.101), a second Maxwell relation can be deduced by differentiation of the first with respect to P and the second with respect to s:

∂T ∂v = . (4.102) ∂P ∂s s P

The relations for Helmholtz and Gibbs free each supply additional useful relations including two new . First consider the Helmholtz free energy

a = u T s, (4.103) − da = du Tds sdT, (4.104) − − = ( P dv + Tds) Tds sdT, (4.105) − − − = P dv sdT. (4.106) − − So the canonical variables for a are v and T . The conjugate variables are P and s. Thus − − ∂a ∂a da = dv + dT. (4.107) ∂v ∂T T v

So one gets ∂a ∂a P = , s = . (4.108) − ∂v − ∂T T v

CC BY-NC-ND. 28 January 2019, J. M. Powers. 112 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

and the consequent Maxwell relation

∂P ∂s = . (4.109) ∂T ∂v v T

For the Gibbs free energy g = u + P v T s, (4.110) − =h = h T s, (4.111) | −{z } dg = dh Tds sdT, (4.112) − − = (Tds + vdP ) Tds sdT, (4.113) − − =dh = vdP sdT. (4.114) | −{z } So for Gibbs free energy, the canonical variables are P and T while the conjugate variables are v and s. One then has g = g(P, T ), which gives − ∂g ∂g dg = dP + dT. (4.115) ∂P ∂T T P

So one finds ∂g ∂g v = , s = . (4.116) ∂P − ∂T T P

The resulting Maxwell function is then

∂v ∂s = . (4.117) ∂T − ∂P P T

Example 4.6 Canonical Form

If P R/cP s h(s, P )= c T exp + (h c T ) , (4.118) P o P c o − P o  o   P  and cP , To, R, Po, and ho are all constants, derive both thermal and caloric state equations P (v,T ) and u(v,T ).

Now for this material

∂h P R/cP s = T exp , (4.119) ∂s o P c P  o   P  R/c −1 ∂h RT P P s = o exp . (4.120) ∂P P P c s o  o   P 

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.3. LEGENDRE TRANSFORMATIONS 113

Now since ∂h = T, (4.121) ∂s P ∂h = v, (4.122) ∂P s

one has P R/cP s T = T exp , (4.123) o P c  o   P  RT P R/cP −1 s v = o exp . (4.124) P P c o  o   P  Dividing Eq. (4.123) by Eq. (4.124) gives

T P = , (4.125) v R P v = RT, (4.126)

which is the thermal equation of state. Substituting from Eq. (4.123) into the canonical equation for h, Eq. (4.118), one also finds for the caloric equation of state

h = c T + (h c T ) , (4.127) P o − P o h = c (T T )+ h , (4.128) P − o o

which is useful in itself. Substituting in for T and To, P v P v h = c o o + h . (4.129) P R − R o   Using, Eq. (4.78), h u + P v, we get ≡ P v P v u + P v = c o o + u + P v . (4.130) P R − R o o o   So c c u = P 1 P v P 1 P v + u , (4.131) R − − R − o o o c    u = P 1 (P v P v )+ u , (4.132) R − − o o o c  u = P 1 (RT RT )+ u , (4.133) R − − o o u = (c R) (T T )+ u , (4.134) P − − o o u = (c (c c )) (T T )+ u , (4.135) P − P − v − o o u = c (T T )+ u . (4.136) v − o o So one canonical equation gives us all the information one needs. Often, it is difficult to do a single experiment to get the canonical form.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 114 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

4.4

Recall that ∂u c = , (4.137) v ∂T v ∂h c = . (4.138) P ∂T P

Then perform operations on the Gibbs equation

du = Tds P dv, (4.139) − ∂u ∂s = T , (4.140) ∂T ∂T v v ∂s c = T . (4.141) v ∂T v

Likewise,

dh = Tds + vdP, (4.142) ∂h ∂s = T , (4.143) ∂T ∂T P P ∂s c = T . (4.144) P ∂T P

One finds further useful relations by operating on the Gibbs equation:

du = Tds P dv, (4.145) − ∂u ∂s = T P, (4.146) ∂v ∂v − T T ∂P = T P. (4.147) ∂T − v

So one can then say

u = u(T, v), (4.148) ∂u ∂u du = dT + dv, (4.149) ∂T ∂v v T ∂P = c dT + T P dv. (4.150) v ∂T −  v 

For an ideal gas, one has

∂u ∂P R RT = T P = T , (4.151) ∂v ∂T − v − v T v  

= 0. (4.152)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.4. HEAT CAPACITY 115

Consequently, u is not a function of v for an ideal gas, so u = u(T ) alone. Since Eq. (4.78), h = u + P v, for an ideal gas reduces to h = u + RT

h = u(T )+ RT = h(T ). (4.153)

Now return to general equations of state. With s = s(T, v) or s = s(T,P ), one gets

∂s ∂s ds = dT + dv, (4.154) ∂T ∂v v T ∂s ∂s ds = dT + dP. (4.155) ∂T ∂P P T

Now using Eqs. (4.102, 4.117, 4.141, 4.144) one gets c ∂P ds = v dT + dv, (4.156) T ∂T v c ∂v ds = P dT dP. (4.157) T − ∂T P

Subtracting Eq. (4.157) from Eq. (4.156), one finds c c ∂P ∂v 0 = v − P dT + dv + dP, (4.158) T ∂T ∂T v P ∂P ∂v (c c )dT = T dv + T dP. (4.159) P − v ∂T ∂T v P

Now divide both sides by dT and hold either P or v constant. In either case, one gets ∂P ∂v c c = T . (4.160) P − v ∂T ∂T v P

Also, since ∂P/∂T v = (∂P/∂v T )(∂v/∂T P ), Eq. (4.160) can be rewritten as | − | | ∂v 2 ∂P c c = T . (4.161) P − v − ∂T ∂v  P  T

2 Now since T > 0, (∂v/∂T P ) > 0, and for all known materials ∂P/∂v T < 0, we must have | |

cP >cv. (4.162)

Example 4.7 For an ideal gas find c c . P − v

CC BY-NC-ND. 28 January 2019, J. M. Powers. 116 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

For the ideal gas, P v = RT , one has

∂P R ∂v R = , = . (4.163) ∂T v ∂T P v P

So, from Eq. (4.160), we have R R c c = T , (4.164) P − v v P R2 = T , (4.165) RT = R. (4.166)

This holds even if the ideal gas is calorically imperfect. That is

c (T ) c (T )= R. (4.167) P − v

For the ratio of specific heats for a general material, one can use Eqs. (4.141) and (4.144) to get

c T ∂s k = P = ∂T P , then apply Eq. (4.56) to get (4.168) c ∂s v T ∂T v

∂s ∂T = , then apply Eq. (4.58) to get (4.169) ∂T ∂s P v ∂s ∂P ∂T ∂v = , (4.170) − ∂P ∂T − ∂v ∂s  T s  s T  ∂v ∂s ∂P ∂T = . (4.171) ∂s ∂P ∂T ∂v  T T   s s

So for general materials

∂v ∂P k = . (4.172) ∂P ∂v T s

The first term can be obtained from P v T data. The second term is related to the isentropic sound speed of the material, which− − is also a measurable quantity.

Example 4.8 For a calorically perfect ideal gas with gas constant R and specific heat at constant volume cv, find expressions for the thermodynamic variable s and thermodynamic potentials u, h, a, and g, as functions of T and P .

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.4. HEAT CAPACITY 117

First get the entropy:

du = T ds P dv, (4.173) − T ds = du + P dv, (4.174)

T ds = cvdT + P dv, (4.175) dT P ds = c + dv, (4.176) v T T dT dv = c + R , (4.177) v T v dT dv ds = c + R , (4.178) v T v Z Z Z T v s s0 = cv ln + R ln , (4.179) − T0 v0 s s T R RT/P − 0 = ln + ln , (4.180) cv T0 cv RT0/P0 T T P R/cv = ln + ln 0 , (4.181) T T P  0   0  T 1+R/cv P R/cv = ln + ln 0 , (4.182) T P  0    T 1+(cp−cv )/cv P (cp−cv)/cv = ln + ln 0 , (4.183) T P  0    T k P k−1 = ln + ln 0 . (4.184) T P  0    So

T k P k−1 s = s + c ln + c ln 0 . (4.185) 0 v T v P  0    Now, for the calorically perfect ideal gas, one has

u = u + c (T T ). (4.186) o v − o For the enthalpy, one gets

h = u + P v, (4.187) = u + RT, (4.188) = u + c (T T )+ RT, (4.189) o v − o = u + c (T T )+ RT + RT RT , (4.190) o v − o o − o = u + RT +c (T T )+ R(T T ), (4.191) o o v − o − o =ho = h + (c + R)(T T ). (4.192) | o {z v } − o =cp

So | {z }

h = h + c (T T ). (4.193) o p − o

CC BY-NC-ND. 28 January 2019, J. M. Powers. 118 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

For the Helmholtz free energy, one gets

a = u T s. (4.194) − Thus,

T k P k−1 a = u + c (T T ) T s + c ln + c ln 0 . (4.195) o v − o − 0 v T v P  0    !

For the Gibbs free energy, one gets

g = h T s. (4.196) − Thus

T k P k−1 g = h + c (T T ) T s + c ln + c ln 0 . (4.197) o p − o − 0 v T v P  0    !

4.5 Van der Waals gas

A van der Waals gas is a common model for a non-ideal gas. It can capture some of the behavior of a gas as it approaches the vapor dome. Its form is

RT a P (T, v)= , (4.198) v b − v2 − where b accounts for the finite volume of the molecules, and a accounts for intermolecular forces.

Example 4.9 Find a general expression for u(T, v) if

RT a P (T, v)= . (4.199) v b − v2 − Proceed as before: First we have ∂u ∂u du = dT + dv, (4.200) ∂T ∂v v T

recalling that ∂u ∂u ∂P = c , = T P. (4.201) ∂T v ∂v ∂T − v T v

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.5. VAN DER WAALS GAS 119

Now for the van der Waals gas, we have

∂P R = , (4.202) ∂T v b v − ∂P RT T P = P, (4.203) ∂T − v b − v − RT RT a a = = . (4.204) v b − v b − v2 v2 −  −  So we have

∂u a = , (4.205) ∂v v2 T a u(T, v) = + f(T ). (4.206) −v

Here f(T ) is some as-of-yet arbitrary function of T . To evaluate f(T ), take the derivative with respect to T holding v constant:

∂u df = = c . (4.207) ∂T dT v v

Since f is a function of T at most, here cv can be a function of T at most, so we allow cv = cv(T ). Integrating, we find f(T ) as

T f(T )= C + cv(Tˆ)dT,ˆ (4.208) ZTo where C is an integration constant. Thus u is

T a u(T, v)= C + c (Tˆ)dTˆ . (4.209) v − v ZTo

Taking C = uo + a/vo, we get

T 1 1 u(T, v)= u + c (Tˆ)dTˆ + a . (4.210) o v v − v ZTo  o 

We also find

T 1 1 h = u + P v = u + c (Tˆ)dTˆ + a + P v, (4.211) o v v − v ZTo  o  T 1 1 RT v a h(T, v)= u + c (Tˆ)dTˆ + a + . (4.212) o v v − v v b − v ZTo  o  −

CC BY-NC-ND. 28 January 2019, J. M. Powers. 120 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Example 4.10 A van der Waals gas with J R = 200 , (4.213) kg K Pa m6 a = 150 , (4.214) kg2 m3 b = 0.001 , (4.215) kg J J c = 350 + 0.2 (T (300 K)), (4.216) v kg K kg K2 −     begins at T = 300 K, P =1 105 Pa. It is isothermally compressed to state 2 where P =1 106 Pa. 1 1 × 2 × It is then isochorically heated to state 3 where T3 = 1000 K. Find 1w3, 1q3, and s3 s1. Assume the surroundings are at 1000 K. −

Recall RT a P = , (4.217) v b − v2 − so at state 1

6 J Pa m 200 kg K (300 K) 150 2 105 Pa = kg . (4.218)   m3 − v2 v1 0.001 1  − kg Expanding, one gets   kg kg2 kg3 0.15+ 150 v 60100 v2 + 100000 v3 =0. (4.219) − m3 1 − m6 1 m9 1       This is a cubic equation which has three solutions: m3 v = 0.598 , physical, (4.220) 1 kg m3 v = 0.00125 0.0097i not physical, (4.221) 1 − kg m3 v = 0.00125+ 0.0097i not physical. (4.222) 1 kg

Now at state 2, P2 and T2 are known, so v2 can be determined:

6 J Pa m 200 kg K (300 K) 150 2 106 Pa = kg . (4.223)   m3 − v2 v2 0.001 2  − kg 3   The physical solution is v2 =0.0585 m /kg. Now at state 3 it is known that v3 = v2 and T3 = 1000 K. Determine P3:

6 200 J (1000 K) Pa m kg K 150 kg2 P3 = , (4.224)  m3  m3 − 3 2 0.0585 0.001 0.0585 m kg − kg kg = (3478261 Pa) (43831 Pa),   (4.225) − = 3434430 P a. (4.226)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.5. VAN DER WAALS GAS 121

Now w = w + w = 2 P dv + 3 P dv = 2 P dv since 2 3 is at constant volume. So 1 3 1 2 2 3 1 2 1 → v2 R RT Ra R w = dv, (4.227) 1 3 v b − v2 Zv1  −  v2 dv v2 dv = RT a , (4.228) 1 v b − v2 Zv1 − Zv1 v b 1 1 = RT ln 2 − + a , (4.229) 1 v b v − v  1   2 1  − 3 3 0.0585 m 0.001 m J kg − kg = 200 (300 K) ln 3 3 kg K  0.598 m  0.001 m    kg − kg       Pa m6 1 1 + 150 3 3 , (4.230) kg2 0.0585 m − 0.598 m   kg kg ! J J = 140408 + 2313 , (4.231) − kg kg     J = 138095 , (4.232) − kg kJ = 138.095 . (4.233) − kg

The gas is compressed, so the work is negative. Since u is a state property:

T3 1 1 u u = c (T )dT + a . (4.234) 3 − 1 v v − v ZT1  1 3  Now J J c = 350 + 0.2 (T (300 K)), (4.235) v kg K kg K2 −     J J = 290 + 0.2 T. (4.236) kg K kg K2     so T3 J J 1 1 u u = 290 + 0.2 T dT + a , (4.237) 3 − 1 kg K kg K2 v − v ZT1       1 3  J J 1 1 = 290 (T T )+ 0.1 T 2 T 2 + a , (4.238) kg K 3 − 1 kg K2 3 − 1 v − v      1 3  J J  = 290 ((1000 K) (300 K)) + 0.1 (1000 K)2 (300 K)2 kg K − kg K2 −      Pa m6 1 1 + 150 3 3 , (4.239) kg2 0.598 m − 0.0585 m   kg kg ! J J J = 203000 + 91000 2313 , (4.240) kg kg − kg       J = 291687 , (4.241) kg kJ = 292 . (4.242) kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 122 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Now from the first law

u u = q w , (4.243) 3 − 1 1 3 − 1 3 q = u u + w , (4.244) 1 3 3 − 1 1 3 kJ kJ q = 292 138 , (4.245) 1 3 kg − kg     kJ q = 154 . (4.246) 1 3 kg

The heat transfer is positive as heat was added to the system.

Now find the entropy change. Manipulate the Gibbs equation:

T ds = du + P dv, (4.247) 1 P ds = du + dv, (4.248) T T 1 a P ds = c (T )dT + dv + dv, (4.249) T v v2 T 1  a  1 RT a ds = c (T )dT + dv + dv, (4.250) T v v2 T v b − v2  −  c (T ) R  ds = v dT + dv, (4.251) T v b − T3 c (T ) v b s s = v dT + R ln 3 − , (4.252) 3 − 1 T v b ZT1 1 − 1000 J 290 kg K J v b = + 0.2 dT + R ln 3 − , (4.253)  T  kg K2  v b Z300   1 − J 1000 K J = 290 ln + 0.2 ((1000 K) (300 K)) kg K 300 K kg K2 −       3 3 0.0585 m 0.001 m J kg − kg + 200 ln 3 3 , (4.254) kg K  0.598 m  0.001 m    kg − kg J  J   J  = 349 + 140 468 , (4.255) kg K kg K − kg K       J = 21 , (4.256) kg K kJ = 0.021 . (4.257) kg K

Is the second law satisfied for each portion of the process?

First look at 1 2 → u u = q w , (4.258) 2 − 1 1 2 − 1 2 q = u u + w , (4.259) 1 2 2 − 1 1 2 T2 1 1 v b 1 1 q = c (T )dT + a + RT ln 2 − + a . (4.260) 1 2 v v − v 1 v b v − v ZT1  1 2 !   1 −   2 1 

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.5. VAN DER WAALS GAS 123

Recalling that T1 = T2 and canceling the terms in a, one gets

v b q = RT ln 2 − , (4.261) 1 2 1 v b  1  − 3 3 0.0585 m 0.001 m J kg − kg = 200 (300 K) ln 3 3 , (4.262) kg K  0.598 m  0.001 m    kg − kg J       = 140408 . (4.263) − kg

Since the process is isothermal,

v b s s = R ln 2 − (4.264) 2 − 1 v b  1  3 3− 0.0585 m 0.001 m J kg − kg = 200 ln 3 3 , (4.265) kg K  0.598 m  0.001 m    kg − kg     J = 468.0 . (4.266) − kg K

Entropy drops because heat was transferred out of the system.

Check the second law. Note that in this portion of the process in which the heat is transferred out of the system, that the surroundings must have T 300 K. For this portion of the process let us surr ≤ take Tsurr = 300 K.

q s s 1 2 ? (4.267) 2 − 1 ≥ T J J 140408 468.0 − kg , (4.268) − kg K ≥ 300 K J J 468.0 468.0 ok. (4.269) − kg K ≥ − kg K

Next look at 2 3 →

q = u u + w , (4.270) 2 3 3 − 2 2 3 T3 1 1 v3 q = c (T )dT + a + P dv , (4.271) 2 3 v v − v ZT2  2 3 ! Zv2  T3 since isochoric 2q3 = cv(T )dT, (4.272) ZT2 1000 K J J = 290 + 0.2 T dT, (4.273) kg K kg K2 Z300 K      J = 294000 . (4.274) kg

CC BY-NC-ND. 28 January 2019, J. M. Powers. 124 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Now look at the entropy change for the :

T3 c (T ) s s = v dT , (4.275) 3 − 2 T ZT2 J T3 290 kg K J = + 0.2 dT , (4.276)  T  kg K2  ZT2    J 1000 K  J = 290 ln + 0.2 ((1000 K) (300 K)), (4.277) kg K 300 K kg K2 −       J = 489 . (4.278) kg K

Entropy rises because heat transferred into system.

In order to transfer heat into the system we must have a different thermal reservoir. This one must have Tsurr 1000 K. Assume here that the heat transfer was from a reservoir held at 1000 K to assess the influence≥ of the second law. q s s 2 3 ? (4.279) 3 − 2 ≥ T J J 294000 489 kg , (4.280) kg K ≥ 1000 K J J 489 294 , ok. (4.281) kg K ≥ kg K

4.6 Redlich-Kwong gas

The Redlich-Kwong equation of state is

RT a P = . (4.282) v b − v(v + b)T 1/2 − It is modestly more accurate than the in predicting material behavior.

Example 4.11 For the case in which b = 0, find an expression for u(T, v) consistent with the Redlich-Kwong state equation. Here the equation of state is now

RT a P = . (4.283) v − v2T 1/2

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.7. COMPRESSIBILITY AND GENERALIZED CHARTS 125

Proceeding as before, we have

∂u ∂P = T P, (4.284) ∂v ∂T − T v R a RT a = T + , (4.285) v 2v2T 3/2 − v − v2T 1/2     3a = . (4.286) 2v2T 1/2 Integrating, we find 3a u(T, v)= + f(T ). (4.287) −2vT 1/2 Here f(T ) is a yet-to-be-specified function of temperature only. Now the specific heat is found by the temperature derivative of u:

∂u 3a df c (T, v)= = + . (4.288) v ∂T 4vT 3/2 dT v

Obviously, for this material, cv is a function of both T and v. Let us define cvo(T ) via

df c (T ). (4.289) dT ≡ vo Integrating, then one gets

T f(T )= C + cvo(Tˆ) dT.ˆ (4.290) ZTo 1/2 Let us take C = uo +3a/2/vo/To . Thus we arrive at the following expressions for cv(T, v) and u(T, v): 3a c (T, v) = c (T )+ , (4.291) v vo 4vT 3/2

T 3a 1 1 u(T, v)= u + c (Tˆ) dTˆ + . (4.292) o vo 2 1/2 − vT 1/2 ZTo voTo !

4.7 Compressibility and generalized charts

A simple way to quantify the deviation from ideal gas behavior is to determine the so-called compressibility Z, where P v Z . (4.293) ≡ RT

CC BY-NC-ND. 28 January 2019, J. M. Powers. 126 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

For an ideal gas, Z = 1. For substances with a simple molecular structure, Z can be tabulated as functions of the so-called reduced pressure Pr and reduced temperature Tr. Tr and Pr are dimensionless variables found by scaling their dimensional counterparts by the specific material’s temperature and pressure at the critical point, Tc and Pc: T P Tr , Pr . (4.294) ≡ Tc ≡ Pc Often charts are available which give predictions of all reduced thermodynamic properties. These are most useful to capture the non-ideal gas behavior of materials for which tables are not available.

4.8 Mixtures with variable composition

Consider now mixtures of N species. The focus here will be on extensive properties and molar properties. Assume that each species has ni moles, and the the total number of moles N is n = i=1 ni. Now one might expect the extensive energy to be a function of the entropy, the volume and the number of moles of each species: P

U = U(V,S,ni). (4.295)

The extensive version of the Gibbs law in which all of the ni are held constant is

dU = PdV + T dS. (4.296) − Thus ∂U ∂U = P, = T. (4.297) ∂V − ∂S S,ni V,ni

In general, since U = U(S,V,n ), one should expect, for systems in which the n are allowed i i to change that ∂U ∂U N ∂U dU = dV + dS + dni. (4.298) ∂V ∂S ∂ni S,ni V,ni i=1 S,V,nj X

Defining the new thermodynamics property, the µi, as ∂U µi , (4.299) ≡ ∂ni S,V,nj

one has the important Gibbs equation for multicomponent systems:

N dU = PdV + TdS + µ dn . (4.300) − i i i=1 X

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.8. MIXTURES WITH VARIABLE COMPOSITION 127

Obviously, by its definition, µi is on a per mole basis, so it is given the appropriate overline notation. In Eq. (4.300), the independent variables and their conjugates are

x = V, ψ = P, (4.301) 1 1 − x2 = S, ψ2 = T, (4.302)

x3 = n1, ψ3 = µ1, (4.303)

x4 = n2, ψ4 = µ2, (4.304) . . . .

xN+2 = nN , ψN+2 = µN . (4.305)

Equation (4.300) has 2N+1 1 Legendre functions. Three are in wide usage: the extensive − analog to those earlier found. They are

H = U + PV, (4.306) A = U T S, (4.307) − G = U + PV T S. (4.308) − A set of non-traditional, but perfectly acceptable additional Legendre functions would be

formed from U µ1n1. Another set would be formed from U + PV µ2n2. There are many more, but− one in particular is sometimes noted in the literature: −the so-called grand potential, Ω. The grand potential is defined as

N Ω U TS µ n . (4.309) ≡ − − i i i=1 X Differentiating each defined Legendre function, Eqs. (4.306-4.309), and combining with Eq. (4.300), one finds

N

dH = TdS + VdP + µidni, (4.310) i=1 XN dA = SdT PdV + µ dn , (4.311) − − i i i=1 XN dG = SdT + V dP + µ dn , (4.312) − i i i=1 XN dΩ = PdV SdT n dµ . (4.313) − − − i i i=1 X

Thus, canonical variables for H are H = H(S,P,ni). One finds a similar set of relations as

CC BY-NC-ND. 28 January 2019, J. M. Powers. 128 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

before from each of the differential forms:

∂U ∂H T = = , (4.314) ∂S ∂S V,ni P,ni

∂U ∂A ∂Ω P = = = , (4.315) − ∂V − ∂V − ∂V S,ni T,ni T,µi

∂H ∂G V = = , (4.316) ∂P ∂P S,ni T,ni

∂A ∂G ∂Ω S = = = , (4.317) − ∂T − ∂T − ∂T V,ni P,ni V,µi

∂Ω ni = . (4.318) − ∂µi V,T,µj

∂U ∂H ∂A ∂G µi = = = = (4.319) ∂ni ∂ni ∂ni ∂ni S,V,nj S,P,nj T,V,nj T,P,nj

Each of these induces a corresponding Maxwell relation, obtained by cross differentiation. These are

∂T ∂P = , (4.320) ∂V − ∂S S,ni V,ni

∂T ∂V = , (4.321) ∂P ∂S S,ni P,ni

∂P ∂S = , (4.322) ∂T ∂V V,ni T,ni

∂V ∂S = , (4.323) ∂T − ∂P P,ni T,ni

∂µi ∂S = , (4.324) ∂T − ∂ni P,nj V,nj

∂µi ∂V = , (4.325) ∂P ∂ni T,nj V,nj

∂µl ∂µk = , (4.326) ∂nk ∂nl T,P,nj T,P,nj

∂S ∂P = , (4.327) ∂V ∂T T,µi V,µi

∂ni ∂nk = (4.328) ∂µk ∂µi V,T,µj ,j=6 k V,T,µj ,j=6 i

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.9. PARTIAL MOLAR PROPERTIES 129

4.9 Partial molar properties

4.9.1 Homogeneous functions

In mathematics, a homogeneous function f(x1,...,xN ) of order m is one such that

m f(λx1,...,λxN )= λ f(x1,...,xN ). (4.329)

If m = 1, one has f(λx1,...,λxN )= λf(x1,...,xN ). (4.330) Thermodynamic variables are examples of homogeneous functions.

4.9.2 Gibbs free energy Consider an extensive property, such as the Gibbs free energy G. One has the canonical form G = G(T,P,n1, n2,...,nN ). (4.331)

One would like to show that if each of the mole numbers ni is increased by a common factor, say λ, with T and P constant, that G increases by the same factor λ:

λG(T,P,n1, n2,...,nN )= G(T,P,λn1, λn2,...,λnN ). (4.332)

Differentiate both sides of Eq. (4.332) with respect to λ, while holding P , T , and nj constant, to get

G(T,P,n1, n2,...,nN )= ∂G d(λn ) ∂G d(λn ) ∂G d(λn ) 1 + 2 + + N , (4.333) ∂(λn1) dλ ∂(λn2) dλ ··· ∂(λnN ) dλ nj ,P,T nj ,P,T nj ,P,T

∂G ∂G ∂G = n1 + n2 + + nN , (4.334) ∂(λn1) ∂(λn2) ··· ∂(λnN ) nj ,P,T nj ,P,T nj ,P,T

This must hold for all λ, including λ = 1, so one requires

∂G ∂G ∂G G(T,P,n1, n2,...,nN ) = n1 + n2 + + nN , ∂n1 ∂n2 ··· ∂nN nj ,P,T nj ,P,T nj ,P,T

(4.335)

N ∂G = ni. (4.336) ∂ni i=1 nj ,P,T X

Recall now the definition partial molar property , the derivative of an extensive variable with respect to species n holding n , i = j, T , and P constant. Because the result has units per i j 6

CC BY-NC-ND. 28 January 2019, J. M. Powers. 130 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

mole, an overline superscript is utilized. The partial molar Gibbs free energy of species i, gi is then ∂G gi , (4.337) ≡ ∂ni nj ,P,T

so that N

G = gini. (4.338) i=1 X Using the definition of chemical potential, Eq. (4.319), one also notes then that

N

G(T,P,n1, n2,...,nN )= µini. (4.339) i=1 X

The temperature and pressure dependence of G must lie entirely within µi(T,P,nj), which one notes is also allowed to be a function of nj as well. Consequently, one also sees that the Gibbs free energy per unit mole of species i is the chemical potential of that species:

gi = µi. (4.340)

Using Eq. (4.338) to eliminate G in Eq. (4.308), one recovers an equation for the energy:

N U = PV + TS + µ n . (4.341) − i i i=1 X 4.9.3 Other properties A similar result also holds for any other extensive property such as V , U, H, A, or S. One can also show that

N ∂V V = ni , (4.342) ∂ni i=1 nj ,A,T X N ∂U U = ni (4.343) ∂ni i=1 nj ,V,S X N ∂H H = ni , (4.344) ∂ni i=1 nj ,P,S X N ∂A A = ni , (4.345) ∂ni i=1 nj ,T,V X N ∂S S = ni . (4.346) ∂ni i=1 nj ,U,T X

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.9. PARTIAL MOLAR PROPERTIES 131

Note that these expressions do not formally involve partial molar properties since P and T are not constant. Take now the appropriate partial molar derivatives of G for an ideal mixture of ideal gases to get some useful relations:

G = H T S, (4.347) − ∂G ∂H ∂S = T . (4.348) ∂ni ∂ni − ∂ni T,P,nj T,P,nj T,P,nj

Now from the definition of an ideal mixture hi = hi(T,P ), so one has

N

H = nkhk(T,P ), (4.349) Xk=1 ∂H ∂ N = nkhk(T,P ) , (4.350) ∂ni ∂ni T,P,nj k=1 ! X N ∂nk = hk(T,P ), (4.351) ∂ni Xk=1 =δik N |{z} = δikhk(T,P ), (4.352) Xk=1 = hi(T,P ). (4.353)

Here, the Kronecker delta function δki has been again used. Now for an ideal gas one further has hi = hi(T ). The analysis is more complicated for the entropy, in which

N o Pk S = nk sT,k R ln , (4.354) − Po k=1    XN o ykP = nk sT,k R ln , (4.355) − Po k=1    XN o P nk = nk sT,k R ln R ln N , (4.356) − Po − n k=1   q=1 q !! XN N P P n = n so R ln R n ln k , (4.357) k T,k P k N − o − q=1 nq ! Xk=1    Xk=1 CC BY-NC-ND. 28P January 2019, J. M. Powers. 132 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

N ∂S ∂ o P = nk sT,k R ln ∂ni ∂ni − Po T,P,nj k=1    X N ∂ nk R nk ln N , (4.358) − ∂ni n T,P,nj k=1 q=1 q !! X N ∂nk o P P = sT,k R ln ∂ni − Po Xk=1    =δik N |{z}∂ nk R nk ln N , (4.359) − ∂ni n T,P,nj k=1 q=1 q !! X N o P ∂P nk = sT,i R ln R nk ln N . (4.360) − Po − ∂ni n    T,P,nj k=1 q=1 q !! X

Evaluation of the final term on the right side requires closer examination,P and in fact, after tedious but straightforward analysis, yields a simple result which can easily be verified by direct calculation: N ∂ nk ni nk ln N = ln N . (4.361) ∂ni n n T,P,nj k=1 q=1 q !! q=1 q ! X

So the partial molar entropy is in fact P P

∂S o P ni = sT,i R ln R ln N , (4.362) ∂ni − Po − n T,P,nj   q=1 q !

P = so R ln R ln y ,P (4.363) T,i − P − i  o  P = so R ln i , (4.364) T,i − P  o  = si (4.365) Thus, one can in fact claim for the ideal mixture of ideal gases that

g = h T s . (4.366) i i − i 4.9.4 Relation between mixture and partial molar properties A simple analysis shows how the partial molar property for an individual species is related to the partial molar property for the mixture. Consider, for example, the Gibbs free energy. The mixture-averaged Gibbs free energy per unit mole is G g = . (4.367) n

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.10. IRREVERSIBLE ENTROPY PRODUCTION IN A 133

Now take a partial molar derivative and analyze to get ∂g 1 ∂G G ∂n = 2 , (4.368) ∂ni n ∂ni − n ∂ni T,P,nj T,P,nj T,P,nj

N 1 ∂G G ∂ = 2 nk , (4.369) n ∂ni − n ∂ni T,P,nj T,P,nj k=1 ! X N 1 ∂G G ∂nk = 2 , (4.370) n ∂ni − n ∂ni T,P,nj k=1 T,P,nj X N 1 ∂G G = 2 δik, (4.371) n ∂ni − n T,P,nj k=1 X 1 ∂G G = 2 , (4.372) n ∂ni − n T,P,nj 1 g = g . (4.373) n i − n Multiplying by n and rearranging, one gets

∂g gi = g + n . (4.374) ∂ni T,P,nj

A similar result holds for other properties.

4.10 Irreversible entropy production in a closed system

Consider a multicomponent closed to mass exchanges with its sur- roundings coming into equilibrium. Allow the system to be exchanging work and heat with its surroundings. Assume the temperature difference between the system and its surround- ings is so small that both can be considered to be at temperature T . If δQ is introduced into the system, then the surroundings suffer a loss of entropy: δQ dS = . (4.375) surr − T The system’s entropy S can change via this heat transfer, as well as via other internal irreversible processes, such as internal . The second law of thermodynamics requires that the entropy change of the universe be positive semi-definite: dS + dS 0. (4.376) surr ≥ Eliminating dSsurr, one requires for the system that δQ dS . (4.377) ≥ T

CC BY-NC-ND. 28 January 2019, J. M. Powers. 134 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Consider temporarily the assumption that the work and heat transfer are both reversible. Thus, the irreversible entropy production must be associated with internal chemical reaction. Now the first law for the entire system gives

dU = δQ δW, (4.378) − = δQ P dV, (4.379) − δQ = dU + P dV. (4.380)

Note because the system is closed, there can be no species entering or exiting, and so there is no change dU attributable to dni. While within the system the dni may not be 0, the net contribution to the change in total internal energy is zero. A non-zero dni within the system simply re-partitions a fixed amount of total energy from one species to another. Substituting Eq. (4.380) into Eq. (4.377) to eliminate δQ, one gets

1 dS (dU + PdV ), (4.381) ≥ T =δQ TdS dU PdV 0, (4.382) − − ≥ | {z } dU TdS + PdV 0. (4.383) − ≤ Eq. (4.383) involves properties only and need not require assumptions of reversibility for processes in its derivation. In special cases, it reduces to simpler forms. For processes which are isentropic and isochoric, the second law expression, Eq. (4.383), reduces to dU 0. (4.384) |S,V ≤ For processes which are isoenergetic and isochoric, the second law expression, Eq. (4.383), reduces to dS 0. (4.385) |U,V ≥ Now using Eq. (4.300) to eliminate dS in Eq. (4.385), one can express the second law as

N 1 P 1 dU + dV µidni 0, (4.386) T T − T ! ≥ i=1 X =dS U,V N | 1{z } µ dn 0. (4.387) −T i i ≥ i=1 X irreversible entropy production

The irreversible entropy production| associated{z with} the internal chemical reaction must be the left side of Eq. (4.387). Often the irreversible entropy production is defined as σ, with

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.10. IRREVERSIBLE ENTROPY PRODUCTION IN A CLOSED SYSTEM 135

the second law requiring dσ 0. Equation (4.387) in terms of dσ is ≥ 1 N dσ = µ dn 0. (4.388) −T i i ≥ i=1 X Now, while most standard texts focusing on equilibrium thermodynamics go to great lengths to avoid the introduction of time, it really belongs in a discussion describing the approach to equilibrium. One can divide Eq. (4.387) by a positive time increment dt to get

1 N dn µ i 0. (4.389) −T i dt ≥ i=1 X Since T 0, one can multiply Eq. (4.389) by T to get ≥ − N dn µ i 0. (4.390) i dt ≤ i=1 X dni This will hold if a model for dt is employed which guarantees that the left side of Eq. (4.390) is negative semi-definite. One will expect then for dni/dt to be related to the chemical potentials µi. Elimination of dU in Eq. (4.383) in favor of dH from dH = dU + PdV + V dP gives

dH PdV VdP TdS + PdV 0, (4.391) − − − ≤ =dU dH VdP TdS 0. (4.392) | {z }− − ≤ Thus, one finds for isobaric, isentropic equilibration that

dH 0. (4.393) |P,S ≤ For the Helmholtz and Gibbs free energies, one analogously finds

dA 0, (4.394) |T,V ≤ dG 0. (4.395) |T,P ≤ The expression of the second law in terms of dG is especially useful as it may be easy in an experiment to control so that P and T are constant. This is especially true in an isobaric phase change, in which the temperature is guaranteed to be constant as well. Now one has

N

G = nigi, (4.396) i=1 XN

= niµi. (4.397) i=1 X CC BY-NC-ND. 28 January 2019, J. M. Powers. 136 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

One also has from Eq. (4.312): dG = SdT + V dP + N µ dn , holding T and P constant − i=1 i i that N P dG = µ dn . (4.398) |T,P i i i=1 X Here the dni are associated entirely with internal chemical reactions. Substituting Eq. (4.398) into Eq. (4.395), one gets the important version of the second law which holds that

N dG = µ dn 0. (4.399) |T,P i i ≤ i=1 X In terms of time rates of change, one can divide Eq. (4.399) by a positive time increment dt > 0 to get ∂G N dn = µ i 0. (4.400) ∂t i dt ≤ T,P i=1 X

4.11 Equilibrium in a two-component system

A major task of non-equilibrium thermodynamics is to find a functional form for dni/dt which guarantees satisfaction of the second law, Eq. (4.400) and gives predictions which agree with experiment. This will be discussed in more detail in the following chapter on thermochemistry. At this point, some simple examples will be given in which a na¨ıve but useful functional form for dni/dt is posed which leads at least to predictions of the correct equilibrium values. A much better model which gives the correct dynamics in the time domain of the system as it approaches equilibrium will be presented in the chapter on thermochemistry.

4.11.1 Phase equilibrium Here, consider two examples describing systems in phase equilibrium.

Example 4.12 ◦ Consider an equilibrium two-phase mixture of liquid and vapor H2O at T = 100 C, x =0.5. Use the steam tables to check if equilibrium properties are satisfied.

In a two-phase gas liquid mixture one can expect the following reaction:

H2O(l) ⇌ H2O(g). (4.401)

That is one mole of liquid, in the forward phase change, evaporates to form one mole of gas. In the reverse phase change, one mole of gas condenses to form one mole of liquid.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM 137

Because T is fixed at 100 ◦C and the material is a two phase mixture, the pressure is also fixed at a constant. Here there are two phases at saturation; g for gas and l for liquid. Equation (4.399) reduces to µ dn + µ dn 0. (4.402) l l g g ≤ Now for the pure H2O if a loss of moles from one phase must be compensated by the addition to another. So one must have dnl + dng =0. (4.403) Hence dn = dn . (4.404) g − l So Eq. (4.402), using Eq. (4.404) becomes

µ dn µ dn 0, (4.405) l l − g l ≤ dn (µ µ ) 0. (4.406) l l − g ≤

At this stage of the analysis, most texts, grounded in equilibrium thermodynamics, assert that µl = µg, ignoring the fact that they could be different but dnl could be zero. That approach will not be taken here. Instead divide Eq. (4.406) by a positive time increment, dt 0 to write the second law as ≥ dn l (µ µ ) 0. (4.407) dt l − g ≤ One convenient, albeit na¨ıve, way to guarantee second law satisfaction is to let dn l = κ(µ µ ), κ 0, convenient, but na¨ıve model (4.408) dt − l − g ≥ Here κ is some positive semi-definite scalar rate constant which dictates the time scale of approach to equilibrium. Note that Eq. (4.408) is just a hypothesized model. It has no experimental verification; in fact, other more complex models exist which both agree with experiment and satisfy the second law. For the purposes of the present argument, however, Eq. (4.408) will suffice. With this assumption, the second law reduces to κ(µ µ )2 0, κ 0, (4.409) − l − g ≤ ≥ which is always true. Eq. (4.408) has two important consequences: Differences in chemical potential drive changes in the number of moles. • The number of moles of liquid, n , increases when the chemical potential of the liquid is less than that • l of the gas, µl < µg. That is to say, when the liquid has a lower chemical potential than the gas, the gas is driven towards the phase with the lower potential. Because such a phase change is isobaric and isothermal, the Gibbs free energy is the appropriate variable to consider, and one takes µ = g. When this is so, the Gibbs free energy of the mixture, G = nlµl + ngµg is being driven to a lower value. So when dG = 0, the system has a minimum G. The system is in equilibrium when the chemical potentials of liquid and gas are equal: µ = µ . • l g The chemical potentials, and hence the molar specific Gibbs free energies must be the same for each constituent of the binary mixture at the phase equilibrium. That is

gl = gg. (4.410) Now since both the liquid and gas have the same molecular mass, one also has the mass specific Gibbs free energies equal at phase equilibrium:

gl = gg. (4.411)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 138 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

This can be verified from the steam tables, using the definition g = h Ts. From the tables − kJ kJ kJ g = h Ts = 419.02 ((100 + 273.15) K) 1.3068 = 68.6 , (4.412) l l − l kg − kg K − kg   kJ kJ kJ g = h Ts = 2676.05 ((100 + 273.15) K) 7.3548 = 68.4 . (4.413) g g − g kg − kg K − kg   The two values are essentially the same; the difference is likely due to table inaccuracies.

Example 4.13 This example is from BS, 16.20, p. 700. A container has liquid water at 20 ◦C, 100 kPa, in equilibrium with a mixture of water vapor and dry air, also at 20 ◦C, 100 kPa. Find the water vapor pressure and the saturated water vapor pressure.

Now at this temperature, the tables easily show that the pressure of a saturated vapor is Psat = 2.339 kPa . From the previous example, it is known that for the water liquid and vapor in equilibrium, one has gliq = gvap. (4.414) Now if both the liquid and the vapor were at the saturated state, they would be in phase equilibrium and that would be the end of the problem. But they have slight deviations from the saturated state. One can estimate these deviations with the standard formula

dg = sdT + vdP. (4.415) − The tables will be used to get values at 20 ◦C, for which one can take dT = 0. This allows the approximation of dg v dP . So for the liquid, ∼ g g = v dP v (P P ), (4.416) liq − f ∼ f − sat Z g = g + v (P P ). (4.417) liq f f − sat For the vapor, approximated here as an ideal gas, one has

g g = v dP, (4.418) vap − g Z dP = RT , (4.419) P Z P = RT ln vap , (4.420) Psat Pvap gvap = gg + RT ln . (4.421) Psat Here, once again, one allows for deviations of the pressure of the vapor from the saturation pressure. Now at equilibrium, one enforces gliq = gvap, so one has

Pvap gf + vf (P Psat) = gg + RT ln . (4.422) − Psat =gliq =gvap | {z } | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM 139

5 10

4 (Pa) 10 vap P

3 10 5 6 7 8 9 10 10 10 10 10 P (Pa)

Figure 4.1: Pressure of water vapor as a function of total pressure for example problem.

Now gf = gg, so one gets Pvap vf (P Psat)= RT ln . (4.423) − Psat

Solving for Pvap, one gets

v (P P ) P = P exp f − sat , (4.424) vap sat RT   3 m 0.001002 kg (100 kPa 2.339 kPa) = (2.339 kPa)exp − , (4.425)   kJ  0.4615 kg K (293.15 K)   = 2.3407 kP a.   (4.426)

The pressure is very near the saturation pressure. This justifies assumptions that for such mixtures, one can take the pressure of the water vapor to be that at saturation if the mixture is in equilibrium. If the pressure is higher, the pressure of the vapor becomes higher as well. Figure 4.1 shows how the pressure of the equilibrium vapor pressure varies with total pressure. Clearly, a very high total pressure, on the order of 1 GP a is needed to induce the vapor pressure to deviate significantly from the saturation value.

4.11.2 Chemical equilibrium: introduction Here consider two examples which identify the equilibrium state of a chemically reactive system.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 140 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

4.11.2.1 Isothermal, isochoric system The simplest system to consider is isothermal and isochoric. The isochoric assumption implies there is no work in coming to equilibrium.

Example 4.14 At high temperatures, between diatomic molecules induce the production of monatomic nitrogen molecules. The chemical reaction can be described by the model

N2 + N2 ⇌ 2N + N2. (4.427)

Here one of the N2 molecules behaves as an inert third body. An N2 molecule has to collide with something, to induce the reaction. Some authors leave out the third body and write instead N2 ⇌ 2N, but this does not reflect the true physics as well. The inert third body is especially important when the time scales of reaction are considered. It plays no role in . Consider 1 kmole of N2 and 0 kmole of N at a pressure of 100 kPa and a temperature of 6000 K. Using the ideal gas tables, find the equilibrium concentrations of N and N2 if the equilibration process is isothermal and isochoric. The ideal gas law can give the volume.

n RT P = N2 , (4.428) 1 V n RT V = N2 , (4.429) P1 (1 kmole) 8.314 kJ (6000 K) = kmole K , (4.430) 100 kPa  = 498.84 m3. (4.431)

Initially, the mixture is all N2, so its partial pressure is the total pressure, and the initial partial pressure of N is 0. Now every time an N2 molecule reacts and thus undergo a negative change, 2 N molecules are created and thus undergo a positive change, so 1 dn = dn . (4.432) − N2 2 N This can be parameterized by a reaction progress variable ζ, also called the degree of reaction, defined such that

dζ = dn , (4.433) − N2 1 dζ = dn . (4.434) 2 N As an aside, one can integrate this, taking ζ = 0 at the initial state to get

ζ = n n , (4.435) N2 |t=0 − N2 1 ζ = n . (4.436) 2 N Thus

n = n ζ, (4.437) N2 N2 |t=0 − nN = 2ζ. (4.438)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM 141

One can also eliminate ζ to get nN in terms of nN2 :

n = 2 (n n ) . (4.439) N N2 |t=0 − N2 Now for the reaction, one must have, for second law satisfaction, that

µ dn + µ dn 0, (4.440) N2 N2 N N ≤ µ ( dζ)+ µ (2dζ) 0, (4.441) N2 − N ≤ µ +2µ dζ 0 (4.442) − N2 N ≤ dζ µ +2µ  0. (4.443) − N2 N dt ≤  In order to satisfy the second law, one can usefully, but na¨ıvely, hypothesize that the non-equilibrium reaction kinetics are given by

dζ = k( µ +2µ ), k 0, convenient, but na¨ıve model (4.444) dt − − N2 N ≥ Note there are other ways to guarantee second law satisfaction. In fact, a more complicated model is well known to fit data well, and will be studied later. For the present purposes, this na¨ıve model will suffice. With this assumption, the second law reduces to

k( µ +2µ )2 0, k 0, (4.445) − − N2 N ≤ ≥ which is always true. Obviously, the reaction ceases when dζ/dt = 0, which holds only when

2µN = µN2 . (4.446)

Away from equilibrium, for the reaction to go forward, one must expect dζ/dt > 0, and then one must have

µ +2µ 0, (4.447) − N2 N ≤ 2µ µ . (4.448) N ≤ N2 The chemical potentials are the molar specific Gibbs free energies; thus, for the reaction to go forward, one must have 2g g . (4.449) N ≤ N2 Substituting using the definitions of Gibbs free energy, one gets

2 h T s h T s , (4.450) N − N ≤ N2 − N2 y P y P 2 h T so R ln N  h T so R ln N2 , (4.451) N − T,N − P ≤ N2 − T,N2 − P    o    o  y P y P 2 h T so h T so 2RT ln N + RT ln N2 , (4.452) N − T,N − N2 − T,N2 ≤ − P P  o   o    y P y P 2 h T so + h T so 2RT ln N RT ln N2 , (4.453) − N − T,N N2 − T,N2 ≥ P − P  o   o    2 2 o o yN P Po 2 hN T sT,N + hN2 T sT,N2 RT ln 2 , (4.454) − − − ≥ P Py 2  o N    2 o o yN P 2 hN T sT,N + hN2 T sT,N2 RT ln . (4.455) − − − ≥ y 2 P  N o    CC BY-NC-ND. 28 January 2019, J. M. Powers. 142 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

At the initial state, one has yN = 0, so the right hand side approaches , and the inequality holds. At equilibrium, one has equality. −∞

2 o o yN P 2 hN T sT,N + hN2 T sT,N2 = RT ln . (4.456) − − − y 2 P  N o    Taking numerical values from Table A.9:

kJ kJ 2 5.9727 105 (6000 K) 216.926 + − × kmole − kmole K    kJ kJ 2.05848 105 (6000 K) 292.984 × kmole − kmole K    kJ y2 P = 8.314 (6000 K) ln N , kmole K y P    N2 o  y2 P 2.87635 = ln N , (4.457) − yN2 Po ≡ln KP   y2 P 0|.0563399{z } = N , (4.458) yN2 Po ≡KP 2 | {z } nN nN +nN2 RT = (nN + nN2 ) , (4.459)  nN2  PoV nN +nN2 n2 RT  = N , (4.460) nN2 PoV 2 (2 (n 2 n 2 )) RT = N |t=0 − N , (4.461) nN2 PoV (2(1 kmole n ))2 (8.314)(6000) = − N2 .(4.462) nN2 (100)(498.84)

This is a quadratic equation for nN2 . It has two roots

nN2 =0.888154 kmole physical; nN2 =1.12593 kmole, non-physical (4.463)

The second root generates more N2 than at the start, and also yields non-physically negative nN = 0.25186 kmole. So at equilibrium, the physical root is − n = 2(1 n ) = 2(1 0.888154) = 0.223693 kmole. (4.464) N − N2 − The diatomic species is preferred.

Note in the preceding analysis, the term KP was introduced. This is the so-called equilibrium “constant” which is really a function of temperature. It will be described in more detail later, but one notes that it is commonly tabulated for some reactions. Its tabular value can be derived from the more fundamental quantities shown in this example. BS’s Table A.11 gives for this reaction at 6000 K the value of ln KP = 2.876. Note that KP is fundamentally defined in terms of thermodynamic properties for a system which− may or may not be at chemical equilibrium. Only at chemical equilibrium, can it can further be related to mole fraction and pressure ratios.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM 143

The pressure at equilibrium is

(n + n )RT P = N2 N , (4.465) 2 V (0.888154 kmole +0.223693 kmole) 8.314 kJ (6000 K) = kmole K , (4.466) 498.84  = 111.185 kP a. (4.467)

The pressure has increased because there are more molecules with the volume and temperature being equal. The molar concentrations ρi at equilibrium, are

0.223693 kmole kmole mole ρ = =4.484 10−4 = 4.484 10−7 , (4.468) N 498.84 m3 × m3 × cm3 0.888154 kmole kmole mole ρ = =1.78044 10−3 = 1.78044 10−6 . (4.469) N2 498.84 m3 × m3 × cm3

Now consider the heat transfer. One knows for the isochoric process that 1Q2 = U2 U1. The initial energy is given by −

U1 = nN2 uN2 , (4.470) = n (h RT ), (4.471) N2 N2 − kJ kJ = (1 kmole) 2.05848 105 8.314 (6000 K) , (4.472) × kmole − kmole K     = 1.555964 105 kJ. (4.473) × The energy at the final state is

U2 = nN2 uN2 + nN uN , (4.474) = n (h RT )+ n (h RT ), (4.475) N2 N2 − N N − kJ kJ = (0.888154 kmole) 2.05848 105 8.314 (6000 K) (4.476) × kmole − kmole K     kJ kJ +(0.223693 kmole) 5.9727 105 8.314 (6000 K) , (4.477) × kmole − kmole K     = 2.60966 105 kJ. (4.478) × So

Q = U U , (4.479) 1 2 2 − 1 = 2.60966 105 kJ 1.555964 105 kJ, (4.480) × − × = 1.05002 105 kJ. (4.481) × Heat needed to be added to keep the system at the constant temperature. This is because the nitrogen dissociation process is endothermic. One can check for second law satisfaction in two ways. Fundamentally, one can demand that Eq. (4.377), dS δQ/T , be satisfied for the process, giving ≥ 2 δQ S S . (4.482) 2 − 1 ≥ T Z1 CC BY-NC-ND. 28 January 2019, J. M. Powers. 144 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

For this isothermal process, this reduces to Q S S 1 2 , (4.483) 2 − 1 ≥ T (n s + n s ) N2 N2 N N |2 Q (n s + n s ) 1 2 , (4.484) − N2 N2 N N |1 ≥ T y P y P n so R ln N2 + n so R ln N N2 T,N2 − P N T,N − P    o    o  2

o yN2 P o yN P 1Q2 nN2 sT,N2 R ln + nN sT,N R ln , (4.485) − − P − P ≥ T    o    o  1

o PN2 o PN nN2 sT,N2 R ln + nN sT,N R ln − P − P    o    o  2 P P Q n so R ln N2 + n so R ln N 1 2 , (4.486) − N2 T,N2 − P N T,N − P ≥ T    o    o  1

o nN2 RT o nN RT nN2 s R ln + nN s R ln T,N2 − P V T,N − P V    o    o  2

n RT n RT Q n so R ln N2 + n so R ln N 1 2 . (4.487) −  N2 T,N2 − P V N T,N − P V  ≥ T   o    o  =0 1   |{z} Now at the initial state nN =0 kmole, and RT/Po/V has a constant value of

kJ RT 8.314 kmole K (6000 K) 1 = 3 =1 , (4.488) PoV (100 kPa)(498 .84 m ) kmole so Eq. (4.487) reduces to

o nN2 o nN nN2 sT,N2 R ln + nN sT,N R ln − 1 kmole − 1 kmole 2        o nN2 1Q2 nN2 sT,N2 R ln , − − 1 kmole 1 ≥ T     (4.489)

((0.888143) (292.984 8.314 ln (0.88143)) + (0.223714) (216.926 8.314 ln (0.223714))) − − |2 105002 ((1) (292.984 8.314 ln (1))) , − − |1 ≥ 6000 kJ kJ 19.4181 17.5004 . (4.490) K ≥ K Indeed, the second law is satisfied. Moreover the irreversible entropy production of the chemical reaction is 19.4181 17.5004 = +1.91772 kJ/K. − For the isochoric, isothermal process, it is also appropriate to use Eq. (4.394), dA T,V 0, to check for second law satisfaction. This turns out to give an identical result. Since by Eq. (4.307),| ≤A = U TS, A A = (U T S ) (U T S ). Since the process is isothermal, A A = U U T (S −S ). 2 − 1 2 − 2 2 − 1 − 1 1 2 − 1 2 − 1 − 2 − 1 For A2 A1 0, one must demand U2 U1 T (S2 S1) 0, or U2 U1 T (S2 S1), or S S −(U ≤U )/T . Since Q = U U for− this− isochoric− process,≤ one then recovers− ≤S S − Q /T. 2− 1 ≥ 2− 1 1 2 2− 1 2− 1 ≥ 1 2

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM 145

4.11.2.2 Isothermal, isobaric system

Allowing for isobaric rather than isochoric equilibration introduces small variation in the analysis.

Example 4.15 Consider the same reaction

N2 + N2 ⇌ 2N + N2. (4.491)

for an isobaric and isothermal process. That is, consider 1 kmole of N2 and 0 kmole of N at a pressure of 100 kPa and a temperature of 6000 K. Using the ideal gas tables, find the equilibrium concentrations of N and N2 if the equilibration process is isothermal and isobaric. The initial volume is the same as from the previous example:

3 V1 = 498.84 m . (4.492)

Note the volume will change in this isobaric process. Initially, the mixture is all N2, so its partial pressure is the total pressure, and the initial partial pressure of N is 0. A few other key results are identical to the previous example:

n = 2 (n n ) , (4.493) N N2 |t=0 − N2 and 2g g . (4.494) N ≤ N2 Substituting using the definitions of Gibbs free energy, one gets

2 h T s h T s , (4.495) N − N ≤ N2 − N2 y P y P 2 h T so R ln N  h T so R ln N2 , (4.496) N − T,N − P ≤ N2 − T,N2 − P    o    o  y P y P 2 h T so h T so 2RT ln N + RT ln N2 , (4.497) N − T,N − N2 − T,N2 ≤ − P P  o   o    y P y P 2 h T so + h T so 2RT ln N RT ln N2 , (4.498) − N − T,N N2 − T,N2 ≥ P − P  o   o    2 2 o o yN P Po 2 hN T sT,N + hN2 T sT,N2 RT ln 2 . (4.499) − − − ≥ P Py 2  o N    In this case Po = P , so one gets

2 o o yN 2 hN T sT,N + hN2 T sT,N2 RT ln . (4.500) − − − ≥ y 2  N    At the initial state, one has yN = 0, so the right hand side approaches , and the inequality holds. At equilibrium, one has equality. −∞

2 o o yN 2 hN T sT,N + hN2 T sT,N2 = RT ln . (4.501) − − − y 2  N    CC BY-NC-ND. 28 January 2019, J. M. Powers. 146 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Taking numerical values from Table A.9: kJ kJ 2 5.9727 105 (6000 K) 216.926 + − × kmole − kg K    kJ kJ 2.05848 105 (6000 K) 292.984 × kmole − kg K    kJ y2 = 8.314 (6000 K) ln N , kmole K y    N2  y2 2.87635 = ln N , (4.502) − yN2 ≡ln KP   y2 0|.0563399{z } = N , (4.503) yN2 ≡KP 2 | {z } nN n +n = N N2 , (4.504)  nN2  nN +nN2  n2  = N , (4.505) nN2 (nN + nN2 ) (2 (n n ))2 = N2 |t=0 − N2 , (4.506) n (2 (n n )+ n ) N2 N2 |t=0 − N2 N2 (2 (1 kmole n ))2 = − N2 . (4.507) n (2 (1 kmole n )+ n ) N2 − N2 N2

This is a quadratic equation for nN2 . It has two roots

nN2 =0.882147 kmole physical; nN2 =1.11785 kmole, non-physical (4.508)

The second root generates more N2 than at the start, and also yields non-physically negative nN = 0.235706 kmole. So at equilibrium, the physical root is − n = 2(1 n ) = 2(1 0.882147) = 0.235706 kmole. (4.509) N − N2 − Again, the diatomic species is preferred. As the temperature is raised, one could show that the monatomic species would come to dominate. The volume at equilibrium is (n + n )RT V = N2 N , (4.510) 2 P (0.882147 kmole +0.235706 kmole) 8.314 kJ (6000 K) = kmole K , (4.511) 100 kPa  = 557.630 m3. (4.512) The volume has increased because there are more molecules with the pressure and temperature being equal. The molar concentrations ρi at equilibrium, are 0.235706 kmole kmole mole ρ = =4.227 10−4 = 4.227 10−7 , (4.513) N 557.636 m3 × m3 × cm3 0.882147 kmole kmole mole ρ = =1.58196 10−3 = 1.58196 10−6 . (4.514) N2 557.636 m3 × m3 × cm3

CC BY-NC-ND. 28 January 2019, J. M. Powers. 4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM 147

The molar concentrations are a little smaller than for the isochoric case, mainly because the volume is larger at equilibrium. Now consider the heat transfer. One knows for the isobaric process that Q = H2 H1. The initial enthalpy is given by − kJ H = n h = (1 kmole) 2.05848 105 =2.05848 105 kJ. (4.515) 1 N2 N2 × kmole ×   The enthalpy at the final state is

H2 = nN2 hN2 + nN hN , (4.516) kJ kJ = (0.882147 kmole) 2.05848 105 + (0.235706 kmole) 5.9727 105 (4.517), × kmole × kmole     = 3.22368 105 kJ. (4.518) × So Q = H H =3.22389 105 kJ 2.05848 105 kJ =1.16520 105 kJ. (4.519) 1 2 2 − 1 × − × × Heat needed to be added to keep the system at the constant temperature. This is because the nitrogen dissociation process is endothermic. Relative to the isochoric process, more heat had to be added to maintain the temperature. This to counter the cooling effect of the expansion. Lastly, it is a straightforward exercise to show that the second law is satisfied for this process.

4.11.3 Equilibrium condition The results of both of the previous examples, in which a functional form of a progress variable’s time variation, dζ/dt, was postulated in order to satisfy the second law gave a condition for equilibrium. This can be generalized so as to require at equilibrium that

N

µiνi =0. (4.520) i=1 X ≡−α

Here νi represents the net number of moles| {z of} species i generated in the forward reaction. This negation of the term on the left side of Eq (4.520) is sometimes defined as the chemical affinity, α: N α µ ν . (4.521) ≡ − i i i=1 X So in the phase equilibrium example, Eq. (4.520) becomes µ ( 1) + µ (1) = 0. (4.522) l − g In the nitrogen chemistry example, Eq. (4.520) becomes µ ( 1) + µ (2) = 0. (4.523) N2 − N This will be discussed in detail in the following chapter.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 148 CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

CC BY-NC-ND. 28 January 2019, J. M. Powers. Chapter 5

Thermochemistry of a single reaction

Read BS, Chapters 13, 14 See Powers, Chapter 4 See Abbott and Van Ness, Chapter 7 See Kondepudi and Prigogine, Chapters 4, 5, 7, 9 See Turns, Chapter 2 See Kuo, Chapters 1, 2

This chapter will further develop notions associated with the thermodynamics of chemical reactions. The focus will be on single chemical reactions.

5.1 Molecular mass

The molecular mass of a molecule is a straightforward notion from chemistry. One simply sums the product of the number of atoms and each atom’s atomic mass to form the molecular mass. If one defines L as the number of elements present in species i, φli as the number of moles of atomic element l in species i, and as the atomic mass of element l, the molecular Ml mass Mi of species i L M = φ , i =1,...,N. (5.1) i Ml li Xl=1 In vector form, one would say

MT = T φ, or M = φT . (5.2) M · ·M Here M is the vector of length N containing the molecular masses, is the vector of length L containing the elemental masses, and φ is the matrix of dimensionM L N containing the × number of moles of each element in each species. Generally, φ is full rank. If N > L, φ has rank L. If N < L, φ has rank N. In any problem involving an actual chemical reaction, one will find N L, and in most cases N > L. In isolated problems not involving a reaction, ≥ 149 150 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

one may have N < L. In any case, M lies in the column space of φT , which is the row space of φ.

Example 5.1 Find the molecular mass of H2O. Here, one has two elements H and O, so L = 2, and one species, so N = 1; thus, in this isolated problem, N

φ11 M1 = 1 2 , (5.3) M M · φ21   = φ +  φ , (5.4) M1 11 M2 21 kg kg = 1 (2) + 16 (1), (5.5) kmole kmole     kg = 18 . (5.6) kmole

Example 5.2 Find the molecular masses of the two species C8H18 and CO2. Here, for practice, the vector matrix notation is exercised. In a certain sense this is overkill, but it is useful to be able to understand a general notation. One has N = 2 species, and takes i = 1 for C8H18 and i = 2 for CO2. One also has L = 3 elements and takes l = 1 for C, l = 2 for H, and l = 3 for O. Now for each element, one has 1 = 12 kg/kmole, =1 kg/kmole, = 16 kg/kmole. The molecular masses are then given by M M2 M3

φ11 φ12 M1 M2 = 1 2 3 φ21 φ22 , (5.7) M M M · φ φ    31 32 8 1  = 12 1 16 18 0 , (5.8) ·  0 2    = 114 44 . (5.9)  That is for C8H18, one has molecular mass M1 = 114 kg/kmole. For CO2, one has molecular mass M2 = 44 kg/kmole.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.2. STOICHIOMETRY 151

5.2 Stoichiometry

5.2.1 General development Stoichiometry represents a mass balance on each element in a chemical reaction. For example, in the simple global reaction 2H2 + O2 ⇌ 2H2O, (5.10) one has 4 H atoms in both the reactant and product sides and 2 O atoms in both the reactant and product sides. In this section stoichiometry will be systematized. Consider now a general reaction with N species. This reaction can be represented by

N N ′ ⇌ ′′ νiχi νi χi. (5.11) i=1 i=1 X X th ′ th Here χi is the i chemical species, νi is the forward stoichiometric coefficient of the i ′′ th ′ ′′ reaction, and νi is the reverse stoichiometric coefficient of the i reaction. Both νi and νi are to be interpreted as pure dimensionless numbers. In Equation (5.10), one has N = 3 species. One might take χ1 = H2, χ2 = O2, and χ3 = H2O. The reaction is written in more general form as

(2)χ1 + (1)χ2 + (0)χ3 ⇌ (0)χ1 + (0)χ2 + (2)χ3, (5.12)

(2)H2 + (1)O2 + (0)H2O ⇌ (0)H2 + (0)O2 + (2)H2O. (5.13) Here, one has

′ ′′ ν1 = 2, ν1 =0, (5.14) ′ ′′ ν2 = 1, ν2 =0, (5.15) ′ ′′ ν3 = 0, ν3 =2. (5.16) It is common and useful to define another pure dimensionless number, the net stoichio- metric coefficients for species i, νi. Here νi represents the net production of number if the reaction goes forward. It is given by ν = ν′′ ν′. (5.17) i i − i For the reaction 2H2 + O2 ⇌ 2H2O, one has ν = ν′′ ν′ =0 2= 2, (5.18) 1 1 − 1 − − ν = ν′′ ν′ =0 1= 1, (5.19) 2 2 − 2 − − ν = ν′′ ν′ =2 0=2. (5.20) 3 3 − 3 − With these definitions, it is possible to summarize a chemical reaction as

N

νiχi =0. (5.21) i=1 X CC BY-NC-ND. 28 January 2019, J. M. Powers. 152 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

In vector notation, one would say νT χ =0. (5.22) · For the reaction of this section, one might write the non-traditional form

2H O +2H O =0. (5.23) − 2 − 2 2 It remains to enforce a stoichiometric balance. This is achieved if, for each element, l = 1,...,L, one has the following equality:

N

φliνi =0, l =1, . . . , L. (5.24) i=1 X That is to say, for each element, the sum of the product of the net species production and the number of elements in the species must be zero. In vector notation, this becomes

φ ν = 0. (5.25) · One may recall from linear algebra that this demands that ν lie in the right null space of φ.

Example 5.3 Show stoichiometric balance is achieved for 2H O +2H O = 0. − 2 − 2 2 Here again, the number of elements L = 2, and one can take l = 1 for H and l = 2 for O. Also the number of species N = 3, and one takes i = 1 for H2, i = 2 for O2, and i = 3 for H2O. Then for element 1, H, in species 1, H2, one has

φ11 =2,H in H2. (5.26)

Similarly, one gets

φ12 = 0, H in O2, (5.27)

φ13 = 2, H in H2O, (5.28)

φ21 = 0, O in H2, (5.29)

φ22 = 2, O in O2, (5.30)

φ23 = 1, O in H2O. (5.31)

N In matrix form then, i=1 φliνi = 0 gives P ν 2 0 2 1 0 ν2 = . (5.32) 0 2 1 ·   0   ν3     This is two equations in three unknowns. Thus, it is formally under-constrained. Certainly the trivial solution ν1 = ν2 = ν3 = 0 will satisfy, but one seeks non-trivial solutions. Assume ν3 has a known value ν3 = ξ. Then, the system reduces to 2 0 ν 2ξ 1 = . (5.33) 0 2 ν − ξ    2  − 

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.2. STOICHIOMETRY 153

The inversion here is easy, and one finds ν = ξ, ν = 1 ξ. Or in vector form, 1 − 2 − 2 ν ξ 1 − ν = 1 ξ , (5.34)  2 − 2  ν3 ξ    1 −1 R1 = ξ 2 , ξ . (5.35) −1  ∈   T Again, this amounts to saying the solution vector (ν1,ν2,ν3) lies in the right null space of the coefficient matrix φli. Here ξ is any real scalar. If one takes ξ = 2, one gets

ν 2 1 − ν = 1 , (5.36)  2 −  ν3 2     This simply corresponds to 2H O +2H O =0. (5.37) − 2 − 2 2 If one takes ξ = 4, one still achieves stoichiometric balance, with −

ν1 4 ν2 = 2 , (5.38) ν   4 3 −     which corresponds to the equally valid

4H +2O 4H O =0. (5.39) 2 2 − 2 In summary, the stoichiometric coefficients are non-unique but partially constrained by mass conserva- tion. Which set is chosen is to some extent arbitrary, and often based on traditional conventions from chemistry. But others are equally valid.

There is a small issue with units here, which will be seen to be difficult to reconcile. In practice, it will have little to no importance. In the previous example, one might be tempted to ascribe units of kmoles to νi. Later, it will be seen that in classical reaction kinetics, νi is best interpreted as a pure dimensionless number, consistent with the definition of this section. So in the context of the previous example, one would then take ξ to be dimensionless as well, which is perfectly acceptable for the example. In later problems, it will be more useful to give ξ the units of kmoles. Note that multiplication of ξ by any scalar, e.g. kmole/(6.02 1026), still yields an acceptable result. ×

Example 5.4 Balance an equation for hypothesized ethane combustion

′ ′ ⇌ ′′ ′′ ν1C2H6 + ν2O2 ν3 CO2 + ν4 H2O. (5.40)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 154 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

One could also say in terms of the net stoichiometric coefficients

ν1C2H6 + ν2O2 + ν3CO2 + ν4H2O =0. (5.41)

Here one takes χ1 = C2H6, χ2 = O2, χ3 = CO2, and χ4 = H2O. So there are N = 4 species. There are also L = 3 elements: l = 1 : C, l = 2 : H, l = 3 : O. One then has

φ11 = 2, C in C2H6, (5.42)

φ12 = 0, C in O2, (5.43)

φ13 = 1, C in CO2, (5.44)

φ14 = 0, C in H2O, (5.45)

φ21 = 6, H in C2H6, (5.46)

φ22 = 0, H in O2, (5.47)

φ23 = 0, H in CO2, (5.48)

φ24 = 2, H in H2O, (5.49)

φ31 = 0, O in C2H6, (5.50)

φ32 = 2, O in O2, (5.51)

φ33 = 2, O in CO2, (5.52)

φ34 = 1, O in H2O. (5.53)

N So the stoichiometry equation, i=1 φliνi = 0, is given by P ν 2 0 10 1 0 ν 6 0 02 2 = 0 . (5.54) ν  0 2 21 3 0 ν4         Here, there are three equations in four unknowns, so the system is under-constrained. There are many ways to address this problem. Here, choose the robust way of casting the system into row echelon form. This is easily achieved by Gaussian elimination. Row echelon form seeks to have lots of zeros in the lower left part of the matrix. The lower left corner has a zero already, so that is useful. Now, multiply the top equation by 3 and subtract the result from the second to get

ν 20 1 0 1 0 ν 0 0 3 2 2 = 0 . (5.55) ν  02− 2 1 3 0 ν4         Next the last two equations to get

ν 20 1 0 1 0 ν 02 2 1 2 = 0 . (5.56) ν  0 0 3 2 3 0 − ν4         Now divide the first by 2, the second by 2 and the third by 3 to get unity in the diagonal: −

1 ν1 1 0 2 0 0 1 ν2 0 1 1 2   = 0 . (5.57)  2  ν3   0 0 1 3 0 − ν4         CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.2. STOICHIOMETRY 155

So-called bound variables have non-zero coefficients on the diagonal, so one can take the bound variables to be ν1, ν2, and ν3. The remaining variables are free variables. Here one takes the free variable to be ν4. So, set ν4 = ξ, and rewrite the system as

1 1 0 2 ν1 0 1 0 1 1 ν2 = 2 ξ . (5.58)     −2  0 0 1 ν3 3 ξ       Solving, one finds

1 1 ν1 3 ξ 3 − 7 − 7 ν2 6 ξ 6 R1   = −2  = ξ −2  , ξ . (5.59) ν3 3 ξ 3 ∈ ν4  ξ   1              Again, one finds a non-unique solution in the right null space of φ. If one chooses ξ = 6, then one gets

ν1 2 ν −7  2 = −  , (5.60) ν3 4 ν4  6          which corresponds to the stoichiometrically balanced reaction

2C2H6 +7O2 ⇌ 4CO2 +6H2O. (5.61)

In this example, ξ is dimensionless.

Example 5.5 Consider stoichiometric balance for a propane oxidation reaction which may produce carbon monox- ide and hydroxyl in addition to and water.

The hypothesized reaction takes the form

′ ′ ⇌ ′′ ′′ ′′ ′′ ν1C3H8 + ν2O2 ν3 CO2 + ν4 CO + ν5 H2O + ν6 OH. (5.62)

In terms of net stoichiometric coefficients, this becomes

ν1C3H8 + ν2O2 + ν3CO2 + ν4CO + ν5H2O + ν6OH =0. (5.63)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 156 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

There are N = 6 species and L = 3 elements. One then has

φ11 = 3, C in C3H8, (5.64)

φ12 = 0, C in O2, (5.65)

φ13 = 1, C in CO2, (5.66)

φ14 = 1, C in CO, (5.67)

φ15 = 0, C in H2O, (5.68)

φ16 = 0, C in OH, (5.69)

φ21 = 8, H in C3H8, (5.70)

φ22 = 0, H in O2, (5.71)

φ23 = 0, H in CO2, (5.72)

φ24 = 0, H in CO, (5.73)

φ25 = 2, H in H2O, (5.74)

φ26 = 1, H in OH, (5.75)

φ31 = 0, O in C3H8, (5.76)

φ32 = 2, O in O2, (5.77)

φ33 = 2, O in CO2, (5.78)

φ34 = 1, O in CO, (5.79)

φ35 = 1, O in H2O, (5.80)

φ36 = 1, O in OH. (5.81)

The equation φ ν = 0, then becomes ·

ν1 ν 301100 2 0 ν  800021 3 = 0 . (5.82)   · ν4   022111   0 ν5       ν6     Multiplying the first equation by 8/3 and adding it to the second gives −

ν1 ν 30 1 1 00  2 0 8 8 ν3 0 0 3 3 2 1 = 0 . (5.83)  − −  · ν4   02 2 1 11   0 ν5       ν   6   Trading the second and third rows gives

ν1 ν 30 1 1 00 2 0 ν  02 2 1 11 3 = 0 . (5.84)  8 8  · ν4   0 0 3 3 2 1   0 − − ν5       ν6     CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.2. STOICHIOMETRY 157

Dividing the first row by 3, the second by 2 and the third by 8/3 gives −

ν1 1 1 ν2 1 0 3 3 0 0   0 1 1 1 ν3 0 1 1 2 2 2 = 0 . (5.85)  3 3  · ν4   00 1 1 4 8   0 − − ν5       ν6     Take bound variables to be ν1, ν2, and ν3 and free variables to be ν4, ν5, and ν6. So set ν4 = ξ1, ν5 = ξ2, and ν6 = ξ3, and get

1 ξ1 1 0 3 ν1 3 0 1 1 ν = ξ1 − ξ2 ξ3 . (5.86) 2 2 2 2   ·    − −3 − 3  0 0 1 ν3 ξ + ξ + ξ − 1 4 2 8 3       Solving, one finds

1 ν1 8 ( 2ξ2 ξ3) 1 − − ν2 = 8 (4ξ1 10ξ2 7ξ3) . (5.87)    1 − −  ν3 ( 8ξ +6ξ +3ξ ) 8 − 1 2 3     For all the coefficients, one then has

1 ν1 8 ( 2ξ2 ξ3) 0 2 1 1 − − − − ν2 8 (4ξ1 10ξ2 7ξ3) 4 10 7    1 − −    −  −  ν3 ( 8ξ +6ξ +3ξ ) ξ1 8 ξ2 6 ξ3 3 = 8 − 1 2 3 = − + + . (5.88) ν4  ξ1  8  8  8  0  8  0            ν5  ξ   0   8   0     2        ν6  ξ   0   0   8     3                  Here, one finds three independent vectors in the right null space. To simplify the notation, take ξˆ1 = ξ1/8, ξˆ2 = ξ2/8, and ξˆ3 = ξ3/8. Then,

ν1 0 2 1 ν 4 −10 −7  2   −  −  ν3 8 6 3 = ξˆ1 − + ξˆ2 + ξˆ3 . (5.89) ν4  8   0   0          ν   0   8   0   5       ν6  0   0   8                  The most general reaction that can achieve a stoichiometric balance is given by

( 2ξˆ ξˆ )C H + (4ξˆ 10ξˆ 7ξˆ )O + ( 8ξˆ +6ξˆ +3ξˆ )CO +8ξˆ CO +8ξˆ H O +8ξˆ OH =0. (5.90) − 2 − 3 3 8 1 − 2 − 3 2 − 1 2 3 2 1 2 2 3 Rearranging, one gets

(2ξˆ + ξˆ )C H + ( 4ξˆ +10ξˆ +7ξˆ )O ⇌ ( 8ξˆ +6ξˆ +3ξˆ )CO +8ξˆ CO +8ξˆ H O +8ξˆ OH. (5.91) 2 3 3 8 − 1 2 3 2 − 1 2 3 2 1 2 2 3

This will be balanced for all ξˆ1, ξˆ2, and ξˆ3. The values that are actually achieved in practice depend on the thermodynamics of the problem. Stoichiometry only provides some limitations. A slightly more familiar form is found by taking ξˆ2 =1/2 and rearranging, giving

(1 + ξˆ ) C H + (5 4ξˆ +7ξˆ ) O ⇌ (3 8ξˆ +3ξˆ ) CO +4 H O +8ξˆ CO +8ξˆ OH. (5.92) 3 3 8 − 1 3 2 − 1 3 2 2 1 3

CC BY-NC-ND. 28 January 2019, J. M. Powers. 158 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

One notes that often the production of CO and OH will be small. If there is no production of CO or OH, ξˆ1 = ξˆ3 = 0 and one recovers the familiar balance of

C3H8 +5 O2 ⇌ 3 CO2 +4 H2O. (5.93)

One also notes that stoichiometry alone admits unusual solutions. For instance, if ξˆ1 = 100, ξˆ2 =1/2, and ξˆ3 = 1, one has

2 C3H8 + 794 CO2 ⇌ 388 O2 +4 H2O + 800 CO +8 OH. (5.94)

This reaction is certainly admitted by stoichiometry but is not observed in nature. To determine precisely which of the infinitely many possible final states are realized requires a consideration of the N equilibrium condition i=1 νiµi =0. Looked at in another way, we can think of three independent classes of reactions admitted by the P stoichiometry, one for each of the linearly independent null space vectors. Taking first ξˆ1 =1/4, ξˆ2 = 0, ξˆ3 = 0, one gets, after rearrangement

2CO + O2 ⇌ 2CO2, (5.95)

as one class of reaction admitted by stoichiometry. Taking next ξˆ1 = 0, ξˆ2 =1/2, ξˆ3 = 0, one gets

C3H8 +5O2 ⇌ 3CO2 +4H2O, (5.96)

as the second class admitted by stoichiometry. The third class is given by taking ξˆ1 = 0, ξˆ2 = 0, ξˆ3 = 1, and is C3H8 +7O2 ⇌ 3CO2 +8OH. (5.97) In this example, both ξ and ξˆ are dimensionless.

In general, one can expect to find the stoichiometric coefficients for N species composed of L elements to be of the following form:

N−L ν = ξ , i =1,...,N. (5.98) i Dik k Xk=1 Here is a dimensionless component of a full rank matrix of dimension N (N L) Dik × − and rank N L, and ξk is a dimensionless component of a vector of parameters of length N L. The− matrix whose components are are constructed by populating its columns − Dik with vectors which lie in the right null space of φli. Note that multiplication of ξk by any constant gives another set of νi, and mass conservation for each element is still satisfied. For later use, we associate νi with the number of moles ni, and consider a differential change in the number of moles dni from Eq. (5.98) and arrive at

N−L dn = dξ , i =1,...,N. (5.99) i Dik k Xk=1

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.2. STOICHIOMETRY 159

5.2.2 Fuel-air mixtures In combustion with air, one often models air as a simple mixture of diatomic and inert diatomic nitrogen: (O2 +3.76N2). (5.100) The air-fuel ratio, and its reciprocal, the fuel-air ratio, , can be defined on a mass A F and mole basis. mair nair mass = , mole = . (5.101) A mfuel A nfuel Via the molecular masses, one has

mair nairMair Mair mass = = = mole . (5.102) A mfuel nfuelMfuel A Mfuel If there is not enough air to burn all the fuel, the mixture is said to be rich. If there is excess air, the mixture is said to be lean. If there is just enough, the mixture is said to be stoichiometric. The equivalence ratio, Φ, is defined as the actual fuel-air ratio scaled by the stoichiometric fuel-air ratio:

Φ Factual = Astoichiometric . (5.103) ≡ Fstoichiometric Aactual The ratio Φ is the same whether ’s are taken on a mass or mole basis, because the ratio of molecular masses cancel. F

Example 5.6 Calculate the stoichiometry of the combustion of methane with air with an equivalence ratio of Φ=0.5. If the pressure is 0.1 MPa, find the dew point of the products.

First calculate the coefficients for stoichiometric combustion: ′ ′ ⇌ ′′ ′′ ′′ ν1CH4 + ν2(O2 +3.76N2) ν3 CO2 + ν4 H2O + ν5 N2, (5.104) or ν1CH4 + ν2O2 + ν3CO2 + ν4H2O + (ν5 +3.76ν2)N2 =0. (5.105) Here one has N = 5 species and L = 4 elements. Adopting a slightly more intuitive procedure for variety, one writes a conservation equation for each element to get

ν1 + ν3 = 0,C, (5.106)

4ν1 +2ν4 = 0,H, (5.107)

2ν2 +2ν3 + ν4 = 0,O, (5.108)

3.76ν2 + ν5 = 0,N. (5.109) In matrix form this becomes ν 1 0 100 1 0 ν 4 0 020  2 0   ν3 =   . (5.110) 0 2 210 · 0 ν4 0 3.76 0 0 1   0   ν5           CC BY-NC-ND. 28 January 2019, J. M. Powers. 160 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

Now, one might expect to have one free variable, since one has five unknowns in four equations. While casting the equation in row echelon form is guaranteed to yield a proper solution, one can often use intuition to get a solution more rapidly. One certainly expects that CH4 will need to be present for the reaction to take place. One might also expect to find an answer if there is one mole of CH4. So take ν1 = 1. Realize that one could also get a physically valid answer by assuming ν1 to be equal to any scalar.− With ν = 1, one gets 1 − 0 100 ν2 1 0 020 ν 4    3 =   . (5.111) 2 210 · ν4 0 3.76 0 0 1 ν5 0       One easily finds the unique inverse does exist, and that the solution is

ν2 2 ν −1  3 =   . (5.112) ν4 2 ν5 7.52     If there had been more than one free variable,  the four by four matrix would have been singular, and no unique inverse would have existed. In any case, the reaction under stoichiometric conditions is CH 2O + CO +2H O + (7.52+(3.76)( 2))N = 0, (5.113) − 4 − 2 2 2 − 2 CH4 + 2(O2 +3.76N2) ⇌ CO2 +2H2O +7.52N2. (5.114) For the stoichiometric reaction, the fuel-air ratio on a mole basis is 1 = =0.1050. (5.115) Fstoichiometric 2+2(3.76) Now Φ = 0.5, so = Φ , (5.116) Factual Fstoichiometric = (0.5)(0.1050), (5.117) = 0.0525. (5.118) By inspection, one can write the actual reaction equation as

CH4 + 4(O2 +3.76N2) ⇌ CO2 +2H2O +2O2 + 15.04N2. (5.119) Check: 1 = =0.0525. (5.120) Factual 4+4(3.76)

For the dew point of the products, one needs the partial pressure of H2O. The mole fraction of H2O is 2 y = =0.0499 (5.121) H2O 1+2+2+15.04 So the partial pressure of H2O is

PH2O = yH2OP =0.0499(100 kPa)=4.99 kP a. (5.122)

◦ From the steam tables, the saturation temperature at this pressure is Tsat = Tdew point = 32.88 C . If the products cool to this temperature in an exhaust device, the water could condense in the apparatus.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS 161

5.3 First law analysis of reacting systems

One can easily use the first law to learn much about chemically reacting systems.

5.3.1 Enthalpy of formation The enthalpy of formation is the enthalpy that is required to form a molecule from combining its constituents at P = 0.1 MP a and T = 298 K. Consider the reaction (taken here to be irreversible) C + O CO . (5.123) 2 → 2 In order to maintain the process at constant temperature, it is found that heat transfer to the volume is necessary. For the steady constant pressure process, one has

U2 U1 = 1Q2 1W2, (5.124) − − 2 = Q P dV, (5.125) 1 2 − Z1 = Q P (V V ), (5.126) 1 2 − 2 − 1 Q = U U + P (V V ), (5.127) 1 2 2 − 1 2 − 1 = H H , (5.128) 2 − 1 Q = H H . (5.129) 1 2 products − reactants So

Q = n h n h . (5.130) 1 2 i i − i i reactants productsX X In this reaction, one measures that Q = 393522 kJ for the reaction of 1 kmole of C 1 2 − and O2. That is the reaction liberates such energy to the environment. So measuring the heat transfer can give a measure of the enthalpy difference between reactants and products. Assign a value of enthalpy zero to elements in their at the reference state. Thus, C and O2 both have enthalpies of 0 kJ/kmole at T = 298 K, P = 0.1 MP a. This enthalpy is designated, for species i,

o o hf,i = h298,i, (5.131)

and is called the enthalpy of formation. So the energy balance for the products and reactants, here both at the standard state, becomes

o o o Q = n h n h n h , (5.132) 1 2 CO2 f,CO2 − C f,C − O2 f,O2 o kJ kJ 393522 kJ = (1 kmole)h (1 kmole) 0 (1 kmole) 0 . − f,CO2 − kmole − kmole     (5.133)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 162 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

o Thus, the enthalpy of formation of CO is h = 393522 kJ/kmole, since the reaction 2 f,CO2 − involved creation of 1 kmole of CO2. Often values of enthalpy are tabulated in the forms of enthalpy differences ∆hi. These are defined such that

o o h = h +(h h ), (5.134) i f,i i − f,i

=∆hi o = hf,i + ∆| hi.{z } (5.135) Lastly, one notes for an ideal gas that the enthalpy is a function of temperature only, and so does not depend on the reference pressure; hence

o o hi = hi , ∆hi = ∆hi , if ideal gas. (5.136)

Example 5.7 (BS, Ex. 15.6 pp. 629-630). Consider the heat of reaction of the following irreversible reaction in a steady state, steady flow process confined to the standard state of P =0.1 MPa, T = 298 K:

CH +2O CO +2H O(ℓ). (5.137) 4 2 → 2 2 The first law holds that

Q = n h n h . (5.138) cv i i − i i reactants productsX X All components are at their reference states. Table A.10 gives properties, and one finds

Q = n h + n h n h n h , (5.139) cv CO2 CO2 H2O H2O − CH4 CH4 − O2 O2 kJ kJ = (1 kmole) 393522 + (2 kmole) 285830 − kmole − kmole     kJ kJ (1 kmole) 74873 (2 kmole) 0 , (5.140) − − kmole − kmole     = 890309 kJ. (5.141) −

A more detailed analysis is required in the likely case in which the system is not at the reference state.

Example 5.8 (adopted from Moran and Shapiro, p. 619) A mixture of 1 kmole of gaseous methane and 2 kmole of oxygen initially at 298 K and 101.325 kPa burns completely in a closed, rigid, container. Heat transfer occurs until the final temperature is 900 K. Find the heat transfer and the final pressure.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS 163

The combustion is stoichiometric. Assume that no small concentration species are generated. The global reaction is given by CH +2O CO +2H O. (5.142) 4 2 → 2 2 The first law analysis for the closed system is slightly different:

U U = Q W . (5.143) 2 − 1 1 2 − 1 2

Since the process is isochoric, 1W2 =0. So

Q = U U , (5.144) 1 2 2 − 1 = n u + n u n u n u , (5.145) CO2 CO2 H2O H2O − CH4 CH4 − O2 O2 = n (h RT )+ n (h RT ) n (h RT ) n (h RT ), (5.146) CO2 CO2 − 2 H2O H2O − 2 − CH4 CH4 − 1 − O2 O2 − 1 = hCO2 +2hH2O hCH4 2hO2 3R(T2 T1), (5.147) o − −o − − o o = (h + ∆h )+2(h + ∆h ) (h + ∆h ) 2(h + ∆h ) CO2,f CO2 H2O,f H2O − CH4,f CH4 − O2,f O2 3R(T T ), (5.148) − 2 − 1 = ( 393522+ 28030)+ 2( 241826 + 21937) ( 74873+ 0) 2(0 + 0) − − − − − 3(8.314)(900 298), (5.149) − − = 745412 kJ. (5.150) − For the pressures, one has

P1V1 = (nCH4 + nO2 )RT1, (5.151)

(nCH4 + nO2 )RT1 V1 = , (5.152) P1 kJ (1 kmole +2 kmole) 8.314 kg K (298 K) = , (5.153) 101.325 kPa  = 73.36 m3. (5.154)

Now V2 = V1, so

(nCO2 + nH2O)RT2 P2 = , (5.155) V2 kJ (1 kmole +2 kmole) 8.314 kg K (900 K) = , (5.156) 73.36 m3  = 306.0 kP a. (5.157)

The pressure increased in the reaction. This is entirely attributable to the temperature rise, as the number of moles remained constant here.

5.3.2 Enthalpy and internal energy of combustion The enthalpy of combustion is the difference between the enthalpy of products and reactants when complete combustion occurs at a given pressure and temperature. It is also known

CC BY-NC-ND. 28 January 2019, J. M. Powers. 164 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

as the heating value or the heat of reaction. The internal energy of combustion is related and is the difference between the internal energy of products and reactants when complete combustion occurs at a given volume and temperature. The term higher heating value refers to the energy of combustion when liquid water is in the products. Lower heating value refers to the energy of combustion when water vapor is in the product.

5.3.3 Adiabatic flame temperature in isochoric stoichiometric sys- tems The adiabatic flame temperature refers to the temperature which is achieved when a fuel and oxidizer are combined with no loss of work or heat energy. Thus, it must occur in a closed, insulated, fixed volume. It is generally the highest temperature that one can expect to achieve in a combustion process. It generally requires an iterative solution. Of all mixtures, stoichiometric mixtures will yield the highest adiabatic flame temperatures because there is no need to heat the excess fuel or oxidizer. Here four examples will be presented to illustrate the following points.

The adiabatic flame temperature can be well over 5000 K for seemingly ordinary • mixtures.

Dilution of the mixture with an inert diluent lowers the adiabatic flame temperature. • The same effect would happen in rich and lean mixtures.

Preheating the mixture, such as one might find in the compression stroke of an engine, • increases the adiabatic flame temperature.

Consideration of the presence of minor species lowers the adiabatic flame temperature. • 5.3.3.1 Undiluted, cold mixture

Example 5.9 A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2 and 1 kmole of O2 at 100 kPa and 298 K. Find the adiabatic flame temperature and final pressure assuming the irreversible reaction 2H + O 2H O. (5.158) 2 2 → 2 The volume is given by

(n + n )RT V = H2 O2 1 , (5.159) P1 (2 kmole +1 kmole) 8.314 kJ (298 K) = kmole K , (5.160) 100 kPa  = 74.33 m3. (5.161)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS 165

The first law gives U U = Q W , (5.162) 2 − 1 1 2 − 1 2 U U = 0, (5.163) 2 − 1 n u n u n u = 0, (5.164) H2O H2O − H2 H2 − O2 O2 n (h RT ) n (h RT ) n (h RT ) = 0, (5.165) H2O H2O − 2 − H2 H2 − 1 − O2 O2 − 1 2h 2 h h +R( 2T +3T ) = 0, (5.166) H2O − H2 − O2 − 2 1 =0 =0 2h + (8.314) (( 2) T +(3)(298)) = 0, (5.167) H2O |{z} |{z}− 2 hH2O 8.314T2 + 3716.4 = 0, (5.168) o − h + ∆h 8.314T + 3716.4 = 0, (5.169) f,H2O H2O − 2 241826+ ∆h 8.314T + 3716.4 = 0, (5.170) − H2O − 2 238110+ ∆h 8.314T = 0. (5.171) − H2O − 2

At this point, one begins an iteration process, guessing a value of T2 and an associated ∆hH2O. When T2 is guessed at 5600 K, the left side becomes 6507.04. When T2 is guessed at 6000 K, the left side becomes 14301.4. Interpolate then to arrive at −

T2 = 5725 K. (5.172) This is an extremely high temperature. At such temperatures, in fact, one can expect other species to co-exist in the equilibrium state in large quantities. These may include H, OH, O, HO2, and H2O2, among others. The final pressure is given by n RT P = H2O 2 , (5.173) 2 V kJ (2 kmole) 8.314 kmole K (5725 K) = 3 , (5.174) 74.33 m  = 1280.71 kP a. (5.175)

The final concentration of H2O is 2 kmole kmole ρ = =2.69 10−2 . (5.176) H2O 74.33 m3 × m3

5.3.3.2 Dilute, cold mixture

Example 5.10 Consider a variant on the previous example in which the mixture is diluted with an inert, taken here to be N2. A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2, 1 kmole of O2, and 8 kmole of N2 at 100 kPa and 298 K. Find the adiabatic flame temperature and the final pressure, assuming the irreversible reaction 2H + O +8N 2H O +8N . (5.177) 2 2 2 → 2 2

CC BY-NC-ND. 28 January 2019, J. M. Powers. 166 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

The volume is given by

(n + n + n )RT V = H2 O2 N2 1 , (5.178) P1 (2 kmole +1 kmole +8 kmole) 8.314 kJ (298 K) = kmole K , (5.179) 100 kPa  = 272.533 m3. (5.180)

The first law gives

U U = Q W , 2 − 1 1 2 − 1 2 U U = 0, 2 − 1 n u n u n u + n (u u ) = 0, H2O H2O − H2 H2 − O2 O2 N2 N2 2 − N2 1 n (h RT ) n (h RT ) n (h RT )+ n ((h RT ) (h RT )) = 0, H2O H2O − 2 − H2 H2 − 1 − O2 O2 − 1 N2 N22 − 2 − N21 − 1 2h 2 h h +R( 10T + 11T )+8( h h ) = 0, H2O − H2 − O2 − 2 1 N22 − N21 =0 =0 =0 =∆hN2 |{z}2h |{z}+ (8.314)( 10T + (11)(298))|{z +} 8∆|h{z} = 0, H2O − 2 N22 2hH2O 83.14T2 + 27253.3+8∆hN22 = 0, o − 2h + 2∆h 83.14T + 27253.3+8∆h = 0, f,H2O H2O − 2 N22 2( 241826)+ 2∆h 83.14T + 27253.3+8∆h = 0, − H2O − 2 N22 456399+ 2∆h 83.14T + 8∆h = 0. − H2O − 2 N22

At this point, one begins an iteration process, guessing a value of T2 and an associated ∆hH2O. When T2 is guessed at 2000 K, the left side becomes 28006.7. When T2 is guessed at 2200 K, the left side becomes 33895.3. Interpolate then to arrive at −

T2 = 2090.5 K. (5.181)

The inert diluent significantly lowers the adiabatic flame temperature. This is because the N2 serves as a heat sink for the energy of reaction. If the mixture were at non-stoichiometric conditions, the excess species would also serve as a heat sink, and the adiabatic flame temperature would be lower than that of the stoichiometric mixture. The final pressure is given by

(n + n )RT P = H2O N2 2 , (5.182) 2 V kJ (2 kmole +8 kmole) 8.314 kmole K (2090.5 K) = 3 , (5.183) 272.533 m  = 637.74 kP a. (5.184)

The final concentrations of H2O and N2 are 2 kmole kmole ρ = =7.34 10−3 , (5.185) H2O 272.533 m3 × m3 8 kmole kmole ρ = =2.94 10−2 . (5.186) N2 272.533 m3 × m3

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS 167

5.3.3.3 Dilute, preheated mixture

Example 5.11 Consider a variant on the previous example in which the diluted mixture is preheated to 1000 K. One can achieve this via an isentropic compression of the cold mixture, such as might occur in an engine. To simplify the analysis here, the temperature of the mixture will be increased, while the pressure will be maintained. A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2, 1 kmole of O2, and 8 kmole of N2 at 100 kPa and 1000 K. Find the adiabatic flame temperature and the final pressure, assuming the irreversible reaction 2H + O +8N 2H O +8N . (5.187) 2 2 2 → 2 2 The volume is given by (n + n + n )RT V = H2 O2 N2 1 , (5.188) P1 (2 kmole +1 kmole +8 kmole) 8.314 kJ (1000 K) = kmole K , (5.189) 100 kPa  = 914.54 m3. (5.190) The first law gives U U = Q W , 2 − 1 1 2 − 1 2 U U = 0, 2 − 1 n u n u n u + n (u u ) = 0, H2O H2O − H2 H2 − O2 O2 N2 N2 2 − N2 1 n (h RT ) n (h RT ) n (h RT )+ n ((h RT ) (h RT )) = 0, H2O H2O − 2 − H2 H2 − 1 − O2 O2 − 1 N2 N22 − 2 − N21 − 1 2h 2h h + R( 10T + 11T )+8(h h ) = 0, H2O − H2 − O2 − 2 1 N22 − N21 2( 241826+ ∆h ) 2(20663) 22703 + (8.314)( 10T + (11)(1000)) + 8∆h 8(21463) = 0, − H2O − − − 2 N22 − 2∆h 83.14T 627931 + 8∆h = 0. H2O − 2 − N22

At this point, one begins an iteration process, guessing a value of T2 and an associated ∆hH2O. When T2 is guessed at 2600 K, the left side becomes 11351. When T2 is guessed at 2800 K, the left side becomes 52787. Interpolate then to arrive at −

T2 = 2635.4 K. (5.191) The preheating raised the adiabatic flame temperature. Note that the preheating was by (1000 K) (298 K) = 702 K. The new adiabatic flame temperature is only (2635.4 K) (2090.5 K) = 544.9 K− greater. − The final pressure is given by (n + n )RT P = H2O N2 2 , (5.192) 2 V kJ (2 kmole +8 kmole) 8.314 kmole K (2635.4 K) = 3 , (5.193) 914 .54 m  = 239.58 kP a. (5.194)

The final concentrations of H2O and N2 are 2 kmole kmole ρ = =2.19 10−3 , (5.195) H2O 914.54 m3 × m3 8 kmole kmole ρ = =8.75 10−3 . (5.196) N2 914.54 m3 × m3

CC BY-NC-ND. 28 January 2019, J. M. Powers. 168 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

5.3.3.4 Dilute, preheated mixture with minor species

Example 5.12 Consider a variant on the previous example. Here allow for minor species to be present at equilib- rium. A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2, 1 kmole of O2, and 8 kmole of N2 at 100 kPa and 1000 K. Find the adiabatic flame temperature and the final pressure, assuming reversible reactions. Here, the details of the analysis are postponed, but the result is given which is the consequence of a calculation involving detailed reactions rates. One can also solve an optimization problem to minimize the Gibbs free energy of a wide variety of products to get the same answer. In this case, the equilibrium temperature and pressure are found to be

T = 2484.8 K, P = 227.89 kP a. (5.197)

Equilibrium species concentrations are found to be

kmole minor product ρ = 1.3 10−4 , (5.198) H2 × m3 kmole minor product ρ = 1.9 10−5 , (5.199) H × m3 kmole minor product ρ = 5.7 10−6 , (5.200) O × m3 kmole minor product ρ = 3.6 10−5 , (5.201) O2 × m3 kmole minor product ρ = 5.9 10−5 , (5.202) OH × m3 kmole major product ρ = 2.0 10−3 , (5.203) H2O × m3 kmole trace product ρ = 1.1 10−8 , (5.204) HO2 × m3 kmole trace product ρ = 1.2 10−9 , (5.205) H2O2 × m3

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.4. CHEMICAL EQUILIBRIUM 169

kmole trace product ρ = 1.7 10−9 , (5.206) N × m3 kmole trace product ρ = 3.7 10−10 , (5.207) NH × m3 kmole trace product ρ = 1.5 10−10 , (5.208) NH2 × m3 kmole trace product ρ = 3.1 10−10 , (5.209) NH3 × m3 kmole trace product ρ = 1.0 10−10 , (5.210) NNH × m3 kmole minor product ρ = 3.1 10−6 , (5.211) NO × m3 kmole trace product ρ = 5.3 10−9 , (5.212) NO2 × m3 kmole trace product ρ = 2.6 10−9 , (5.213) N2O × m3 kmole trace product ρ = 1.7 10−9 , (5.214) HNO × m3 kmole major product ρ = 8.7 10−3 . (5.215) N2 × m3 Note that the concentrations of the major products went down when the minor species were considered. The adiabatic flame temperature also went down by a significant amount: 2635 2484.8 = 150.2 K. Some thermal energy was necessary to break the bonds which induce the presence− of minor species.

5.4 Chemical equilibrium

Often reactions are not simply unidirectional, as alluded to in the previous example. The reverse reaction, especially at high temperature, can be important. Consider the four species reaction

′ ′ ⇌ ′′ ′′ ν1χ1 + ν2χ2 ν3 χ3 + ν4 χ4. (5.216)

In terms of the net stoichiometric coefficients, this becomes

ν1χ1 + ν2χ2 + ν3χ3 + ν4χ4 =0. (5.217)

One can define a variable ζ, the reaction progress. Take the dimension of ζ to be kmoles. When t = 0, one takes ζ = 0. Now as the reaction goes forward, one takes dζ > 0. And a forward reaction will decrease the number of moles of χ1 and χ2 while increasing the number of moles of χ3 and χ4. This will occur in ratios dictated by the stoichiometric coefficients of the problem:

CC BY-NC-ND. 28 January 2019, J. M. Powers. 170 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

′ dn1 = ν1dζ, (5.218) − ′ dn2 = ν2dζ, (5.219) − ′′ dn3 = +ν3 dζ, (5.220) ′′ dn4 = +ν4 dζ. (5.221)

′ ′′ Note that if ni is taken to have units of kmoles, νi, and νi are taken as dimensionless, then ζ must have units of kmoles. In terms of the net stoichiometric coefficients, one has

dn1 = ν1dζ, (5.222)

dn2 = ν2dζ, (5.223)

dn3 = ν3dζ, (5.224)

dn4 = ν4dζ. (5.225)

Again, for argument’s sake, assume that at t = 0, one has

n = n , (5.226) 1|t=0 1o n = n , (5.227) 2|t=0 2o n = n , (5.228) 3|t=0 3o n = n . (5.229) 4|t=0 4o Then after integrating, one finds

n1 = ν1ζ + n1o, (5.230)

n2 = ν2ζ + n2o, (5.231)

n3 = ν3ζ + n3o, (5.232)

n4 = ν4ζ + n4o. (5.233)

One can also eliminate the parameter ζ in a variety of fashions and parameterize the reaction one of the species mole numbers. Choosing, for example, n1 as a parameter, one gets n n ζ = 1 − 1o . (5.234) ν1

Eliminating ζ then one finds all other mole numbers in terms of n1:

n1 n1o n2 = ν2 − + n2o, (5.235) ν1 n1 n1o n3 = ν3 − + n3o, (5.236) ν1 n1 n1o n4 = ν4 − + n4o. (5.237) ν1

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.4. CHEMICAL EQUILIBRIUM 171

Written another way, one has n n n n n n n n 1 − 1o = 2 − 2o = 3 − 3o = 4 − 4o = ζ. (5.238) ν1 ν2 ν3 ν4

N For an N-species reaction, i=1 νiχi = 0, one can generalize to say

P dni = = νidζ, (5.239)

ni = νiζ + nio, (5.240) n n i − io = ζ. (5.241) νi Note that dn i = ν . (5.242) dζ i Now, from the previous chapter, one manifestation of the second law is Eq. (4.399):

N dG = µ dn 0. (5.243) |T,P i i ≤ i=1 X Now, one can eliminate dni in Eq. (5.243) by use of Eq. (5.239) to get

N dG = µ ν dζ 0, (5.244) |T,P i i ≤ i=1 X ∂G N = µ ν 0, (5.245) ∂ζ i i ≤ T,P i=1 X = α 0. (5.246) − ≤ Then for the reaction to go forward, one must require that the affinity be positive:

α 0. (5.247) ≥ One also knows from the previous chapter that the irreversible entropy production takes the form of Eq. (4.387):

1 N µ dn 0, (5.248) −T i i ≥ i=1 X 1 N dζ µ ν 0, (5.249) −T i i ≥ i=1 X 1 dζ N µ ν 0. (5.250) −T dt i i ≥ i=1 X CC BY-NC-ND. 28 January 2019, J. M. Powers. 172 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

In terms of the chemical affinity, α = N µ ν , Eq. (5.250) can be written as − i=1 i i 1 dζ P α 0. (5.251) T dt ≥ Now one straightforward, albeit na¨ıve, way to guarantee positive semi-definiteness of the irreversible entropy production and thus satisfaction of the second law is to construct the chemical kinetic rate equation so that dζ N = k µ ν = kα, k 0, provisional, na¨ıve assumption (5.252) dt − i i ≥ i=1 X This provisional assumption of convenience will be supplanted later by a form which agrees well with experiment. Here k is a positive semi-definite scalar. In general, it is a function of temperature, k = k(T ), so that reactions proceed rapidly at high temperature and slowly at low temperature. Then certainly the reaction progress variable ζ will cease to change when the equilibrium condition N

µiνi =0, (5.253) i=1 X is met. This is equivalent to requiring α =0. (5.254) Now, while Eq. (5.253) is the most compact form of the equilibrium condition, it is not the most commonly used form. One can perform the following analysis to obtain the form in most common usage. Start by equating the chemical potential with the Gibbs free energy

per unit mole for each species i: µi = gi. Then employ the definition of Gibbs free energy for an ideal gas, and carry out a set of operations: N

giνi = 0, at equilibrium, (5.255) i=1 N X (h T s )ν = 0, at equilibrium. (5.256) i − i i i=1 X For the ideal gas, one can substitute for hi(T ) and si(T,P ) and write the equilibrium con- dition as

   N T T o c (Tˆ) y P h + c (Tˆ) dTˆ T so + Pi dTˆ R ln i  ν = 0.  298,i Pi −  298,i ˆ − P  i i=1  298  298 T o  X  Z Z     o  =so   =∆hT,i  T,i      | o {z } | {z =sT,i }   =hT,i=hT,i      | {z } | {z } (5.257)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.4. CHEMICAL EQUILIBRIUM 173

Now writing the equilibrium condition in terms of the enthalpies and entropies referred to the standard pressure, one gets

N o y P h T so R ln i ν = 0, (5.258) T,i − T,i − P i i=1    o  X N N o y P h T so ν = RT ν ln i , (5.259) T,i − T,i i − i P i=1 i=1  o   o o  X =gT,i=µT,i X | {zN } N y P νi go ν = RT ln i , (5.260) − T,i i P i=1 i=1 o X X   ≡∆Go ∆Go N y P νi | {z } = ln i , (5.261) − P RT i=1 o X   N y P νi = ln i , (5.262) P i=1 o ! Y   ∆Go N y P νi exp = i , (5.263) − P RT i=1 o   Y   ≡KP

N νi | {z } yiP K = , (5.264) P P i=1  o  Y N νi n P Pi=1 K = yνi . (5.265) P P i o i=1   Y So

N P νi K = i , at equilibrium. (5.266) P P i=1 o Y  

Here KP is what is known as the pressure-based equilibrium constant. It is dimensionless. Despite its name, it is not a constant. It is defined in terms of thermodynamic properties, and for the ideal gas is a function of T only:

∆Go KP exp , generally valid. (5.267) ≡ − RT  

Only at equilibrium does the property KP also equal the product of the partial pressures as in Eq. (5.266). The subscript P for pressure comes about because it is also related to

CC BY-NC-ND. 28 January 2019, J. M. Powers. 174 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

the product of the ratio of the partial pressure to the reference pressure raised to the net stoichiometric coefficients. Also, the net change in Gibbs free energy of the reaction at the reference pressure, ∆Go, which is a function of T only, has been defined as

N ∆Go go ν . (5.268) ≡ T,i i i=1 X The term ∆Go has units of kJ/kmole; it traditionally does not get an overbar. If ∆Go > 0, o one has 0 1, and products are favored over reactants. One can also deduce that higher pressures P push the equilibrium in such a fashion that fewer moles are present, all else being equal. One can also define ∆Go in terms of the chemical affinity, referred to the reference pressure, as ∆Go = αo. (5.269) − One can also define another convenient thermodynamic property, which for an ideal gas is a function of T alone, the equilibrium constant Kc:

N i=1 νi o Po P ∆G Kc exp , generally valid. (5.270) ≡ RT − RT     This property is dimensional, and the units depend on the stoichiometry of the reaction. N 3 νi The units of Kc will be (kmole/m )Pi=1 . The equilibrium condition, Eq. (5.266), is often written in terms of molar concentrations and Kc. This can be achieved by the operations, valid only at an equilibrium state:

ν N ρ RT i K = i , (5.271) P P i=1  o  Y N νi N o Pi=1 ∆G RT νi exp − = ρi , (5.272) RT Po     i=1 N Y νi o N Pi=1 Po ∆G νi exp − = ρi . (5.273) RT RT i=1     Y ≡Kc So | {z }

N νi Kc = ρi , at equilibrium. (5.274) i=1 Y

One must be careful to distinguish between the general definition of Kc as given in Eq. (5.270), and the fact that at equilibrium it is driven to also have the value of product of molar species concentrations, raised to the appropriate stoichiometric power, as given in Eq. (5.274).

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION 175

5.5 Chemical kinetics of a single isothermal reaction

In the same fashion in ordinary mechanics that an understanding of statics enables an under- standing of dynamics, an understanding of chemical equilibrium is necessary to understand to more challenging topic of chemical kinetics. Chemical kinetics describes the time-evolution of systems which may have an initial state far from equilibrium; it typically describes the path of such systems to an equilibrium state. Here gas phase kinetics of ideal gas mixtures that obey Dalton’s law will be studied. Important topics such as catalysis and solid or liquid reactions will not be considered. Further, this section will be restricted to strictly isothermal systems. This simplifies the analysis greatly. It is straightforward to extend the analysis of this system to non-isothermal systems. One must then make further appeal to the energy equation to get an equation for temperature evolution. The general form for evolution of species is taken to be

d ρ ω˙ i = i . (5.275) dt ρ ρ  

Multiplying both sides of Eq. (5.275) by molecular mass Mi and using the definition of mass fraction ci then gives the alternate form

dc ω˙ M i = i i . (5.276) dt ρ

5.5.1 Isochoric systems

Consider the evolution of species concentration in a system which is isothermal, isochoric and spatially homogeneous. The system is undergoing a single chemical reaction involving N species of the familiar form

N

νiχi =0. (5.277) i=1 X Because the density is constant for the isochoric system, Eq. (5.275) reduces to

dρ i =ω ˙ . (5.278) dt i

CC BY-NC-ND. 28 January 2019, J. M. Powers. 176 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

Then, experiment, as well as a more fundamental molecular theory, shows that the evolution of species concentration i is given by

≡ω˙ i

N N dρ E ν′ 1 i z β }| k  νk { = νi aT exp − ρk 1 ρk , isochoric system. (5.279) dt RT ! − Kc   Yk=1  kY=1  ≡k(T )   forward reaction  reverse reaction   | {z } | {z≡r } | {z } This relation| actually holds for isochoric,{z non-isothermal systems} as well, which will not be considered in any detail here. Here some new variables are defined as follows:

a: a kinetic rate constant called the collision frequency factor. Its units will depend • on the actual reaction and could involve various combinations of length, time, and 3 temperature. It is constructed so that dρi/dt has units of kmole/m /s; this requires it 3 (1− N ν′ ) β to have units of (kmole/m ) Pk=1 k /s/K .

β: a dimensionless parameter whose value is set by experiments, sometimes combined • with guiding theory, to account for weak temperature dependency of reaction rates.

E: the activation energy. It has units of kJ/kmole, though others are often used, and is • fit by both experiment and fundamental theory to account for the strong temperature dependency of reaction.

′ Note that in Eq. (5.279) that molar concentrations are raised to the νk and νk powers. As it does not make sense to raise a physical quantity to a power with units, one traditionally ′ ′′ interprets the values of νk, νk, as well as νk to be dimensionless pure numbers. They are also interpreted in a standard fashion: the smallest integer values that actually correspond to the underlying molecular collision which has been modeled. While stoichiometric balance can be achieved by a variety of νk values, the kinetic rates are linked to one particular set which is defined by the community. Equation (5.279) is written in such a way that the species concentration production rate increases when

The net number of moles generated in the reaction, measured by ν increases, • i The temperature increases; here, the sensitivity may be very high, as one observes in • nature,

The species concentrations of species involved in the forward reaction increase; this • embodies the principle that the collision-based reaction rates are enhanced when there are more molecules to collide,

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION 177

The species concentrations of species involved in the reverse reaction decrease. • Here, three intermediate variables which are in common usage have been defined. First one takes the reaction rate to be

N N E ν′ 1 β k  νk  r aT exp − ρk 1 ρk , (5.280) ≡ RT ! − Kc   Yk=1  Yk=1  ≡k(T )   forward reaction  reverse reaction   or | {z } | {z } | {z }

N N E ν′ 1 ν′′ β  k k  r = aT exp − ρk ρk . (5.281) RT − Kc    kY=1 kY=1  ≡k(T ), Arrhenius rate   forward reaction reverse reaction   | {z } | {zlaw} of mass| action{z } The reaction rate r has units of kmole/m| 3/s. {z } The temperature-dependency of the reaction rate is embodied in k(T ) is defined by what is known as an Arrhenius rate law:

E k(T ) aT β exp − . (5.282) ≡ RT   This equation was advocated by van’t Hoff in 1884; in 1889 Arrhenius gave a physical justifi- cation. The units of k(T ) actually depend on the reaction. This is a weakness of the theory, 3 (1− N ν′ ) and precludes a clean non-dimensionalization. The units must be (kmole/m ) Pk=1 k /s. In terms of reaction progress, one can also take 1 dζ r = . (5.283) V dt The factor of 1/V is necessary because r has units of molar concentration per time and ζ has units of kmoles. The over-riding importance of the temperature sensitivity is illustrated as part of the next example. The remainder of the expression involving the products of the species concentrations is the defining characteristic of systems which obey the law of mass action. Though the history is complex, most attribute the law of mass action to Waage and Guldberg in 1864.1 Last, the overall molar production rate of species i, is often written asω ˙ i, defined as ω˙ ν r. (5.284) i ≡ i 1P. Waage and C. M. Guldberg, 1864, “Studies Concerning Affinity, Forhandlinger: Videnskabs-Selskabet i Christiania, 35. English translation: Journal of Chemical Education, 63(12): 1044-1047.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 178 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

3 As νi is considered to be dimensionless, the units ofω ˙ i must be kmole/m /s.

Example 5.13 Study the nitrogen dissociation problem considered in an earlier example, see p. 140, in which at t = 0 s, 1 kmole of N2 exists at P = 100 kPa and T = 6000 K. Take as before the reaction to be isothermal and isochoric. Consider again the elementary nitrogen dissociation reaction

N2 + N2 ⇌ 2N + N2, (5.285) which has kinetic rate parameters of cm3 K1.6 a = 7.0 1021 , (5.286) × mole s β = 1.6, (5.287) − cal E = 224928.4 . (5.288) mole In SI units, this becomes cm3 K1.6 1 m 3 1000 mole m3 K1.6 a = 7.0 1021 =7.0 1018 , (5.289) × mole s 100 cm kmole × kmole s       cal J kJ 1000 mole kJ E = 224928.4 4.186 = 941550 . (5.290) mole cal 1000 J kmole kmole        

At the initial state, the material is all N2, so PN2 = P = 100 kPa. The ideal gas law then gives at t =0 P = P = ρ RT, (5.291) |t=0 N2 |t=0 N2 t=0 P t=0 ρN2 = | , (5.292) t=0 RT 100 kPa

= kJ , (5.293) 8.314 kmole K (6000 K) kmole = 2.00465 10−3  . (5.294) × m3 Thus, the volume, constant for all time in the isochoric process, is

n 2 1 kmole V = N |t=0 = =4.9884 102 m3. (5.295) −3 kmole ρ 2.00465 10 3 × N2 t=0 × m Now the stoichiometry of the reaction is such that

1 dn = dn , (5.296) − N2 2 N 1 (n n ) = (n n ), (5.297) − N2 − N2 |t=0 2 N − N |t=0 =1 kmole =0 n = 2(1 kmole n ), (5.298) | {z }N | −{z N}2 n 1 kmole n N = 2 N2 , (5.299) V V − V   1 kmole ρ = 2 ρ , (5.300) N 4.9884 102 m3 − N2  ×  kmole = 2 2.00465 10−3 ρ . (5.301) × m3 − N2  

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION 179

k(T) ( m 3/kmole/s)

10000

0.0001

1. x 10-12

1. x 10-20

1. x 10-28

1. x 10-36 T (K) 1000 1500 2000 3000 5000 7000 10000

Figure 5.1: k(T ) for Nitrogen dissociation example.

Now the general equation for kinetics of a single reaction, Eq. (5.279), reduces for N2 molar con- centration to

′ ′ dρ 2 E 1 N β νN2 νN νN2 νN = νN2 aT exp − (ρN2 ) (ρN ) 1 (ρN2 ) (ρN ) . (5.302) dt RT − Kc     Realizing that ν′ = 2, ν′ = 0, ν = 1, and ν = 2, one gets N2 N N2 − N 2 dρN2 β E 2 1 ρN = aT exp − ρN2 1 . (5.303) dt − RT − Kc ρ    N2  =k(T ) Examine the primary temperature dependency| {z of the} reaction E k(T ) = aT β exp − , (5.304) RT   m3K1.6 941550 kJ = 7.0 1018 T −1.6 exp − kmole , (5.305) × kmole s 8.314 kJ T   kmole K ! 7.0 1018 1.1325 105 = × exp − × . (5.306) T 1.6 T   Figure 5.1 gives a plot of k(T ) which shows its very strong dependency on temperature. For this problem, T = 6000 K, so 7.0 1018 1.1325 105 k(6000) = × exp − × , (5.307) 60001.6 6000   m3 = 40071.6 . (5.308) kmole s

Now, the equilibrium constant Kc is needed. Recall

N Pi=1 νi o Po ∆G Kc = exp − . (5.309) RT RT    

CC BY-NC-ND. 28 January 2019, J. M. Powers. 180 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

f ( ρ ) (kmole/m 3 /s) N2 unstable equilibrium 1

0.0005 0.001 0.0015 0.002 0.0025 ρ (kmole/m 3 ) N -1 2 unstable stable, equilibrium physical -2 equilibrium

Figure 5.2: Forcing function, f(ρN2 ), which drives changes of ρN2 as a function of ρN2 in isothermal, isochoric problem.

N For this system, since i=1 νi = 1, this reduces to P o o Po (2gN gN2 ) Kc = exp − − , (5.310) RT RT     o o o o P (2(h Ts ) (h 2 Ts )) = o exp − N − T,N − N − T,N2 , (5.311) RT RT   ! 100 (2(597270 (6000)216.926) (205848 (6000)292.984)) = exp − − − − ,(5.312) (8.314)(6000) (8.314)(6000)     kmole = 0.000112112 . (5.313) m3

The differential equation for N2 evolution is then given by

3 −3 kmole 2 dρN2 m 2 1 2 2.00465 10 m3 ρN2 = 40071.6 ρN2 1 kmole × − , dt − kmole − 0.000112112 3 ρ   m N2  !

≡f(ρN2 ) | {z (5.314)}

= f(ρN2 ). (5.315)

The system is at equilibrium when f(ρN2 ) = 0. This is an algebraic function of ρN2 only, and can be

plotted. Figure 5.2 gives a plot of f(ρN2 ) and shows that it has three potential equilibrium points. It is seen there are three roots. Solving for the equilibria requires solving

3 −3 kmole 2 m 2 1 2 2.00465 10 m3 ρN2 0 = 40071.6 ρN2 1 kmole × − . − kmole − 0.000112112 3 ρ   m N2   ! (5.316)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION 181

The three roots are kmole kmole kmole ρ =0 , 0.00178121 , 0.00225611 . (5.317) N2 m3 m3 m3 unstable stable unstable | {z } | {z } | {z kmole} By inspection of the topology of Fig. 5.2, the only stable root is 0.00178121 m3 . This root agrees with the equilibrium value found in an earlier example for the same problem conditions. Small perturbations from this equilibrium induce the forcing function to supply dynamics which restore the system to its orig- inal equilibrium state. Small perturbations from the unstable equilibria induce non-restoring dynamics. kmole For this root, one can then determine that the stable equilibrium value of ρN =0.000446882 m3 . eq One can examine this stability more formally. Define an equilibrium concentration ρN2 such that eq f(ρN2 )=0. (5.318) eq Now perform a Taylor series of f(ρN2 ) about ρN2 = ρN2 :

2 eq df eq 1 d f eq 2 f(ρN2 ) f(ρN2 ) + (ρN2 ρN2 )+ 2 (ρN2 ρN2 ) + ... (5.319) ∼ dρN2 eq − 2 dρN2 − ρN2 =ρN2 =0

Now the first term of| the{z Taylor} series is zero by construction. Next neglect all higher order terms as small so that the approximation becomes

df eq f(ρN2 ) (ρN2 ρN2 ). (5.320) ∼ dρN2 eq − ρN2 =ρN2

Thus, near equilibrium, one can write

dρN2 df eq (ρN2 ρN2 ). (5.321) dt ∼ dρN2 eq − ρN2 =ρN2

Since the derivative of a constant is zero, one can also write the equation as

d eq df eq (ρN2 ρN2 ) (ρN2 ρN2 ). (5.322) dt − ∼ dρN2 eq − ρN2 =ρN2

This has a solution, valid near the equilibrium point, of

eq df (ρN2 ρN2 ) = C exp t , (5.323) −  dρN2 eq  ρN2 =ρN2

 

eq df ρN2 = ρN2 + C exp t . (5.324)  dρN2 eq  ρN2 =ρN2

  Here C is some constant whose value is not important for this discussion. If the slope of f is positive, that is, df > 0, unstable, (5.325) dρN2 eq ρN2 =ρN2

the equilibrium will be unstable. That is a perturbation will grow without bound as t . If the slope is zero, → ∞ df =0, neutrally stable, (5.326) dρN2 eq ρN2 =ρN2

CC BY-NC-ND. 28 January 2019, J. M. Powers. 182 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION 3 0.002 ρ N2 0.0015

0.001

0.0005 ρ N t (s) 0.001 0.002 0.003 0.004 0.005 concentration (kmole/m )

Figure 5.3: ρN2 (t) and ρN (t) in isothermal, isochoric nitrogen dissociation problem.

the solution is stable in that there is no unbounded growth, and moreover is known as neutrally stable. If the slope is negative,

df < 0, asymptotically stable, (5.327) dρN2 eq ρN2 =ρN2

the solution is stable in that there is no unbounded growth, and moreover is known as asymptotically stable. A numerical solution via an explicit technique such as a Runge-Kutta integration is found for

Eq. (5.314). The solution for ρN2 , along with ρN is plotted in Fig. 5.3. Linearization of Eq. (5.314) about the equilibrium state gives rise to the locally linearly valid d ρ 0.00178121 = 1209.39(ρ 0.00178121) + ... (5.328) dt N2 − − N2 −  This has local asymptotically stable solution

ρ =0.00178121 + C exp( 1209.39t). (5.329) N2 − Here C is some integration constant whose value is irrelevant for this analysis. The time scale of relaxation τ is the time when the argument of the exponential is 1, which is − 1 τ = =8.27 10−4 s. (5.330) 1209.39 s−1 × One usually finds this time scale to have high sensitivity to temperature, with high temperatures giving fast time constants and thus fast reactions. The equilibrium values agree exactly with those found in the earlier example; see Eq. (4.469). Here the kinetics provide the details of how much time it takes to achieve equilibrium. This is one of the key questions of non-equilibrium thermodynamics.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION 183

5.5.2 Isobaric systems The form of the previous section is the most important as it is easily extended to a Cartesian grid with fixed volume elements in fluid flow problems. However, there is another important spatially homogeneous problem in which the formulation needs slight modification: isobaric reaction, with P equal to a constant. Again, in this section only isothermal conditions will be considered. In an isobaric problem, there can be volume change. Consider first the problem of isobaric expansion of an inert mixture. In such a mixture, the total number of moles of each species must be constant, so one gets

dn i =0, inert, isobaric mixture. (5.331) dt Now carry out the sequence of operations, realizing the the total mass m is also constant:

1 d (n ) = 0, (5.332) m dt i d n i = 0, (5.333) dt m d nV  i = 0, (5.334) dt V m   d ρ i = 0, (5.335) dt ρ   1 dρ ρ dρ i i = 0, (5.336) ρ dt − ρ2 dt dρ ρ dρ i = i . (5.337) dt ρ dt

So a global density decrease of the inert material due to volume increase of a fixed mass system induces a concentration decrease of each species. Extended to a material with a single reaction rate r, one could say either

dρ ρ dρ i = ν r + i , or (5.338) dt i ρ dt d ρ 1 i = ν r, generally valid, (5.339) dt ρ ρ i   ω˙ = i . (5.340) ρ

Equation (5.339) is consistent with Eq. (5.275) and is actually valid for general systems with variable density, temperature, and pressure. However, in this section, it is required that pressure and temperature be constant. Now

CC BY-NC-ND. 28 January 2019, J. M. Powers. 184 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

differentiate the isobaric, isothermal, ideal gas law to get the density derivative.

N

P = ρiRT, (5.341) i=1 X N dρ 0 = RT i , (5.342) dt i=1 X N dρ 0 = i , (5.343) dt i=1 X N ρ dρ 0 = ν r + i , (5.344) i ρ dt i=1 X   N 1 dρ N 0 = r ν + ρ , (5.345) i ρ dt i i=1 i=1 X X dρ r N ν = − i=1 i . (5.346) dt N ρi Pi=1 ρ ρr N ν = P i=1 i , (5.347) − N Pi=1 ρi ρr N ν = P i=1 i , (5.348) − P PRT ρRTr N ν = i=1 i , (5.349) − P P ρRTr N ν = k=1 k . (5.350) − P P

N Note that if there is no net number change in the reaction, k=1 νk = 0, the isobaric, isothermal reaction also guarantees there would be no density or volume change. It is convenient to define the net number change in the elementary reacP tion as ∆n:

N ∆n ν . (5.351) ≡ k Xk=1

Here ∆n is taken to be a dimensionless pure number. It is associated with the number change in the elementary reaction and not the actual mole change in a ; it is, however, proportional to the actual mole change.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION 185

Now use Eq. (5.350) to eliminate the density derivative in Eq. (5.338) to get

dρ ρ ρRTr N ν i = ν r i k=1 k , (5.352) dt i − ρ P P

ρ RT N = r  ν i ν  , (5.353) i − P k  k=1  reaction effect X   expansion effect    |{z}  or | {z }

dρ i = r ν y ∆n . (5.354) dt  i − i  reaction effect expansion effect   |{z} | {z } There are two terms dictating the rate change of species molar concentration. The first, a reaction effect, is precisely the same term that drove the isochoric reaction. The second is due to the fact that the volume can change if the number of moles change, and this induces an intrinsic change in concentration. Note that the term ρiRT/P = yi, the mole fraction.

Example 5.14 Study a variant of the nitrogen dissociation problem considered in an earlier example, see p. 145, in which at t = 0 s, 1 kmole of N2 exists at P = 100 kPa and T = 6000 K. In this case, take the reaction to be isothermal and isobaric. Consider again the elementary nitrogen dissociation reaction

N2 + N2 ⇌ 2N + N2, (5.355)

which has kinetic rate parameters of

cm3 K1.6 a = 7.0 1021 , (5.356) × mole s β = 1.6, (5.357) − cal E = 224928.4 . (5.358) mole

In SI units, this becomes

cm3 K1.6 1 m 3 1000 mole m3 K1.6 a = 7.0 1021 =7.0 1018 , (5.359) × mole s 100 cm kmole × kmole s       cal J kJ 1000 mole kJ E = 224928.4 4.186 = 941550 . (5.360) mole cal 1000 J kmole kmole        

At the initial state, the material is all N2, so PN2 = P = 100 kPa. The ideal gas law then gives at

CC BY-NC-ND. 28 January 2019, J. M. Powers. 186 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

t =0

P = PN2 = ρN2 RT, (5.361) P ρ = , (5.362) N2 t=0 RT 100 kPa

= kJ , (5.363) 8.314 kmole K (6000 K) kmole = 2.00465 10−3  . (5.364) × m3 Thus, the initial volume is

n 2 1 kmole V = N |t=0 = =4.9884 102 m3. (5.365) t=0 −3 kmole | ρ 2.00465 10 3 × N2 t=0 × m

In this isobaric process, one always has P = 100 kPa. Now, in general

P = RT (ρN2 + ρN ); (5.366)

therefore, one can write ρN in terms of ρN2 : P ρ = ρ , (5.367) N RT − N2 100 kPa = kJ ρN2 , (5.368) 8.314 kmole K (6000 K) − kmole = 2.00465 10−3 ρ . (5.369) × m3 − N2   Then the equations for kinetics of a single isobaric isothermal reaction, Eq. (5.353) in conjunction with Eq. (5.280), reduce for N2 molar concentration to

′ ′ dρ 2 E 1 ρ 2 RT N β νN2 νN νN2 νN N = aT exp − (ρN2 ) (ρN ) 1 (ρN2 ) (ρN ) νN2 (νN2 + νN ) . dt RT − Kc − P      ! =r (5.370) | {z } Realizing that ν′ = 2, ν′ = 0, ν = 1, and ν = 2, one gets N2 N N2 − N 2 dρN2 β E 2 1 ρN ρN2 RT = aT exp − ρN2 1 1 . (5.371) dt RT − Kc ρ − − P    N2  ! =k(T )

The temperature dependency| of the{z reaction} is unchanged from the previous reaction:

E k(T ) = aT β exp − , (5.372) RT   m3K1.6 941550 kJ = 7.0 1018 T −1.6 exp − kmole , (5.373) × kmole s 8.314 kJ T   kmole K ! 7.0 1018 1.1325 105 = × exp − × . (5.374) T 1.6 T  

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION 187

For this problem, T = 6000 K, so 7.0 1018 1.1325 105 k(6000) = × exp − × , (5.375) 60001.6 6000   m3 = 40130.2 . (5.376) kmole s

The equilibrium constant Kc is also unchanged from the previous example. Recall

N Pi=1 νi o Po ∆G Kc = exp − . (5.377) RT RT     N n For this system, since i=1 νi = ∆ = 1, this reduces to o o Po P (2gN gN2 ) Kc = exp − − , (5.378) RT RT     P (2go go ) = o exp − N − N2 , (5.379) RT RT     o o o o P (2(h Ts ) (h 2 Ts )) = o exp − N − T,N − N − T,N2 , (5.380) RT RT   ! 100 (2(597270 (6000)216.926) (205848 (6000)292.984)) = exp − − − − ,(5.381) (8.314)(6000) (8.314)(6000)     kmole = 0.000112112 . (5.382) m3

The differential equation for N2 evolution is then given by

3 −3 kmole 2 dρN2 m 2 1 2.00465 10 m3 ρN2 = 40130.2 ρN2 1 kmole × − dt kmole − 0.000112112 3 ρ   m N2   ! ρ 8.314 kJ (6000 K) 1 N2 kmole K , × − − 100 kPa ! (5.383) f(ρ ). (5.384) ≡ N2

The system is at equilibrium when f(ρN2 ) = 0. This is an algebraic function of ρN2 only, and can be

plotted. Figure 5.4 gives a plot of f(ρN2 ) and shows that it has four potential equilibrium points. It is seen there are four roots. Solving for the equilibria requires solving

3 −3 kmole 2 m 2 1 2.00465 10 m3 ρN2 0 = 40130.2 ρN2 1 kmole × − kmole − 0.000112112 3 ρ   m N2   ! ρ 8.314 kJ (6000 K) 1 N2 kmole K . × − − 100 kPa ! (5.385)

The four roots are kmole kmole kmole kmole ρ = 0.002005 , 0 , 0.001583 , 0.00254 . (5.386) N2 − m3 m3 m3 m3 stable,non−physical unstable stable,physical unstable | {z } | {z } | {z } | {z } CC BY-NC-ND. 28 January 2019, J. M. Powers. 188 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

f ( ρ ) (kmole/m 3 /s) N2 4 unstable unstable 3 equilibrium equilibrium 2

1

ρ 3 -0.002 -0.001 0.001 0.002 0.003 (kmole/m ) N2 -1 stable, stable, physical non-physical equilibrium equilibrium

Figure 5.4: Forcing function, f(ρN2 ), which drives changes of ρN2 as a function of ρN2 in isothermal, isobaric problem.

kmole By inspection of the topology of Fig. 5.2, the only stable, physical root is 0.001583 m3 . Small perturbations from this equilibrium induce the forcing function to supply dynamics which restore the system to its original equilibrium state. Small perturbations from the unstable equilibria induce non- restoring dynamics. For this root, one can then determine that the stable equilibrium value of ρN = kmole 0.000421 m3 . A numerical solution via an explicit technique such as a Runge-Kutta integration is found for

Eq. (5.385). The solution for ρN2 , along with ρN is plotted in Fig. 5.5. Linearization of Eq. (5.385) about the equilibrium state gives rise to the locally linearly valid

d ρ 0.001583 = 967.073(ρ 0.001583) + ... (5.387) dt N2 − − N2 −  This has local solution ρ =0.001583 + C exp( 967.073t) . (5.388) N2 − Again, C is an irrelevant integration constant. The time scale of relaxation τ is the time when the argument of the exponential is 1, which is − 1 τ = =1.03 10−3 s. (5.389) 967.073 s−1 × Note that the time constant for the isobaric combustion is about a factor 1.25 greater than for isochoric combustion under the otherwise identical conditions. The equilibrium values agree exactly with those found in the earlier example; see Eq. (4.514). Again, the kinetics provide the details of how much time it takes to achieve equilibrium.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS 189

0.002 3 ρ N 0.0015 2

0.001

0.0005 ρ N concentration (kmole/m ) t (s) 0.001 0.002 0.003 0.004 0.005

Figure 5.5: ρN2 (t) and ρN (t) in isobaric, isothermal nitrogen dissociation problem.

5.6 Some conservation and evolution equations

Here a few useful global conservation and evolution equations are presented for some key properties. Only some cases are considered, and one could develop more relations for other scenarios.

5.6.1 Total mass conservation: isochoric reaction

One can easily show that the isochoric reaction rate model, Eq. (5.279), satisfies the principle of mixture mass conservation. Begin with Eq. (5.279) in a compact form, using the definition of the reaction rate r, Eq. (5.281) and perform the following operations:

dρ i = ν r, (5.390) dt i d ρc i = ν r, (5.391) dt M i  i  d (ρc ) = ν M r, (5.392) dt i i i d L (ρc ) = ν φ r, (5.393) dt i i Ml li Xl=1 =Mi L d | {z } (ρc ) = φ ν r, (5.394) dt i Ml li i Xl=1 CC BY-NC-ND. 28 January 2019, J. M. Powers. 190 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

N d N L (ρc ) = φ ν r, (5.395) dt i Ml li i i=1 i=1 X X Xl=1

d N L N ρ c  = φ ν r, (5.396) dt i Ml li i  i=1  l=1 i=1  X  X X  =1    | {z } dρ L N = r φ ν . (5.397) dt Ml li i i=1 Xl=1 X =0

Therefore, we get | {z }

dρ =0. (5.398) dt

N Note the term i=1 φliνi = 0 because of stoichiometry, Eq. (5.24). P 5.6.2 Element mass conservation: isochoric reaction Through a similar series of operations, one can show that the mass of each element, l = 1,...,L, in conserved in this reaction, which is chemical, not nuclear. Once again, begin with Eq. (5.281) and perform a set of operations,

dρ i = ν r, (5.399) dt i dρ φ i = φ ν r, l =1, . . . , L, (5.400) li dt li i d (φ ρ ) = rφ ν , l =1, . . . , L, (5.401) dt li i li i N d N (φ ρ ) = rφ ν , l =1, . . . , L, (5.402) dt li i li i i=1 i=1 X X d N N φ ρ = r φ ν , l =1, . . . , L, (5.403) dt li i li i i=1 ! i=1 X X =0 d N φ ρ = 0,| {z l =1} , . . . , L. (5.404) dt li i i=1 ! X CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS 191

N The term i=1 φliρi represents the number of moles of element l per unit volume, by the following analysis P N N moles element l moles species i moles element l φ ρ = = ρ e. (5.405) li i moles species i volume volume ≡ l i=1 i=1 X X e Here the elemental mole density, ρl , for element l has been defined. So the element concen- tration for each element remains constant in a constant volume reaction process:

dρ e l =0, l =1, . . . , L. (5.406) dt One can also multiply by the elemental mass, to get the elemental mass density, ρe: Ml l ρe ρ e, l =1, . . . , L. (5.407) l ≡Ml l Since is a constant, one can incorporate this definition into Eq. (5.406) to get Ml dρe l =0, l =1, . . . , L. (5.408) dt The element mass density remains constant in the constant volume reaction. One could also simply say since the elements’ density is constant, and the mixture is simply a sum of the elements, that the mixture density is conserved as well.

5.6.3 Energy conservation: adiabatic, isochoric reaction Consider a simple application of the first law of thermodynamics to reaction kinetics: that of a closed, adiabatic, isochoric combustion process in a mixture of ideal gases. One may be interested in the rate of temperature change. First, because the system is closed, there can be no mass change, and because the system is isochoric, the total volume is a non-zero constant; hence, dm = 0, (5.409) dt d (ρV ) = 0, (5.410) dt dρ V = 0, (5.411) dt dρ = 0. (5.412) dt For such a process, the first law of thermodynamics is dU = Q˙ W.˙ (5.413) dt −

CC BY-NC-ND. 28 January 2019, J. M. Powers. 192 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

But there is no heat transfer or work in the adiabatic isochoric process, so one gets dU = 0, (5.414) dt d (mu) = 0, (5.415) dt du dm m + u = 0, (5.416) dt dt =0 du |{z} = 0. (5.417) dt

Thus for the mixture of ideal gases, u(T, ρ1,..., ρN ) = uo. One can see how reaction rates affect temperature changes by expanding the derivative in Eq. (5.417)

d N c u = 0, (5.418) dt i i i=1 ! X N d (c u ) = 0, (5.419) dt i i i=1 X N du dc c i + u i = 0, (5.420) i dt i dt i=1 X   N du dT dc c i + u i = 0, (5.421) i dT dt i dt i=1 X   N dT dc c c + u i = 0, (5.422) i vi dt i dt i=1 X   dT N N dc c c = u i , (5.423) dt i vi − i dt i=1 i=1 X X =cv dT N d M ρ | c{z } = u i i , (5.424) v dt − i dt ρ i=1 X   dT N dρ ρc = u M i , (5.425) v dt − i i dt i=1 X dT N ρc = u M ν r, (5.426) v dt − i i i i=1 X dT r N ν u = i=1 i i . (5.427) dt − ρc P v CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS 193

If one defines the net energy change of the reaction as

N ∆U ν u , (5.428) ≡ i i i=1 X one then gets

dT r∆U = . (5.429) dt − ρcv

The rate of temperature change is dependent on the absolute energies, not the energy dif- ferences. If the reaction is going forward, so r > 0, and that is a direction in which the net molar energy change is negative, then the temperature will rise.

5.6.4 Energy conservation: adiabatic, isobaric reaction

Solving for the reaction dynamics in an adiabatic isobaric system requires some non-obvious manipulations. First, the first law of thermodynamics says dU = dQ dW . Since the process is adiabatic, one has dQ = 0, so dU + PdV = 0. Since it is isobaric, one− gets d(U + PV ) = 0, or dH = 0. So the total enthalpy is constant. Then

d H = 0, (5.430) dt d (mh) = 0, (5.431) dt dh = 0, (5.432) dt d N c h = 0, (5.433) dt i i i=1 ! X N d (c h ) = 0, (5.434) dt i i i=1 X N dh dc c i + h i = 0, (5.435) i dt i dt i=1 X N dh dT dc c i + h i = 0, (5.436) i dT dt i dt i=1 X CC BY-NC-ND. 28 January 2019, J. M. Powers. 194 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

N dT N dc c c + h i = 0, (5.437) i Pi dt i dt i=1 i=1 X X dT N N dc c c + h i = 0, (5.438) dt i Pi i dt i=1 i=1 X X =cP dT N d ρ M c |+ {z h} i i = 0, (5.439) P dt i dt ρ i=1 X   dT N d ρ c + h M i = 0. (5.440) P dt i i dt ρ i=1 X   Now use Eq (5.339) to eliminate the term in Eq. (5.440) involving molar concentration derivatives to get

dT N ν r c + h i = 0, (5.441) P dt i ρ i=1 X dT r N h ν = i=1 i i . (5.442) dt − ρc P P So the temperature derivative is known as an algebraic function. If one defines the net enthalpy change as N ∆H h ν , (5.443) ≡ i i i=1 X one gets dT r∆H = . (5.444) dt − ρcP Now differentiate the isobaric ideal gas law to get the density derivative.

N

P = ρiRT, (5.445) i=1 X N dT N dρ 0 = ρ R + RT i , (5.446) i dt dt i=1 i=1 X X dT N N ρ dρ 0 = ρ + T ν r + i , (5.447) dt i i ρ dt i=1 i=1 X X   1 dT N N 1 dρ N 0 = ρ + r ν + ρ . (5.448) T dt i i ρ dt i i=1 i=1 i=1 X X X CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS 195

Solving, we get dρ 1 dT N ρ r N ν = − T dt i=1 i − i=1 i . (5.449) dt N ρi P i=1 ρ P One takes dT/dt from Eq. (5.442) to getP N 1 r i=1 hiνi N N dρ P ρi r νi = T ρcP i=1 − i=1 . (5.450) dt N ρi P i=1 ρ P

Now recall that ρ = ρ/M and cP = cP M, soPρ cP = ρcP . Then Equation (5.450) can be reduced slightly: =1

N N i=1 hiνi ρi N P νi cP T z }|ρ{ − i=1 dρ i=1 = rρ X , (5.451) dt N P i=1 ρi

N hiνi N νi i=1 cP TP− i=1 = rρ N , (5.452) P i=1 ρPi

N hi i=1 νPi c T 1 = rρ P − , (5.453) P  P RT ρRT N h = r ν i 1 , (5.454) P i c T − i=1 P X   N h = rM ν i 1 , (5.455) i c T − i=1 P X   N where M is the mean molecular mass. Note for exothermic reaction i=1 νihi < 0, so exothermic reaction induces a density decrease as the increased temperature at constant pressure causes the volume to increase. P Then using Eq. (5.455) to eliminate the density derivative in Eq. (5.338), and changing the dummy index from i to k, one gets an explicit expression for concentration evolution: N dρi ρi hk = νir + rM νk 1 , (5.456) dt ρ cP T − Xk=1   N ρi hk = r νi + M νk 1  , (5.457) ρ cP T − k=1    X   =yi   N  |{z} hk = r νi + yi νk 1 . (5.458) cP T − ! Xk=1   CC BY-NC-ND. 28 January 2019, J. M. Powers. 196 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

Defining the change of enthalpy of the reaction as ∆H N ν h , and the change of ≡ k=1 k k number of the reaction as ∆n N ν , one can also say ≡ k=1 k P P

dρ ∆H i = r ν + y ∆n . (5.459) dt i i c T −   P 

Exothermic reaction, ∆H < 0, and net number increases, ∆n > 0, both tend to decrease the molar concentrations of the species in the isobaric reaction.

Lastly, the evolution of the adiabatic, isobaric system, can be described by the simul- taneous, coupled ordinary differential equations: Eqs. (5.442, 5.450, 5.458). These require numerical solution in general. Note also that one could also employ a more fundamental N treatment as a differential algebraic system involving H = H1, P = P1 = RT i=1 ρi and Eq. (5.338). P

5.6.5 Entropy evolution: Clausius-Duhem relation

Now consider whether the kinetics law that has been posed actually satisfies the second law of thermodynamics. Consider again Eq. (4.387). There is an algebraic relation on the right side. If it can be shown that this algebraic relation is positive semi-definite, then the second law is satisfied, and the algebraic relation is known as a Clausius-Duhem relation.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS 197

Now take Eq. (4.387) and perform some straightforward operations on it:

1 N dS = µ dn 0, (5.460) |U,V −T i i ≥ i=1 X irreversible entropy production N dS V | {zdni 1 } = µ 0, (5.461) dt − T i dt V ≥ U,V i=1 X V N dρ = µ i 0, (5.462) − T i dt ≥ i=1 X

N N N V E ν′ 1 = µ ν aT β exp − ρ k 1 ρ νk  0,(5.463) − T i i k − K k ≥ i=1 RT ! c X   kY=1  kY=1  ≡k(T )   forward reaction  reverse reaction   | {z } | {z≡r } | {z } N N N ′ V | νk 1{z νk } = µiνik(T ) ρk 1 ρk 0, (5.464) − T − Kc ≥ i=1 k=1 ! k=1 ! X N Y N Y N V ν′ 1 = k(T ) ρ k 1 ρ νk µ ν 0, (5.465) − T k − K k i i ≥ ! c ! i=1 ! kY=1 Yk=1 X =−α Change the dummy index from k back to i, | {z } N N V ν′ 1 = k(T ) ρ i 1 ρ νi α 0, (5.466) T i − K i ≥ i=1 ! c i=1 ! Y Y V = rα, (5.467) T α dζ = . (5.468) T dt

Consider now the affinity α term in Eq. (5.466) and expand it so that it has a more useful form:

N N α = µ ν = g ν , (5.469) − i i − i i i=1 i=1 X X N P = go + RT ln i ν , (5.470) − T,i P i i=1 o X    CC BY-NC-ND. 28 January 2019, J. M. Powers. 198 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

N N P νi = go ν RT ln i , (5.471) − T,i i − P i=1 i=1 o X X   =∆Go

| {z } o N νi ∆G Pi = RT − ln  , (5.472) RT − Po i=1    X   =ln KP   N νi  | {z } Pi = RT ln K ln , (5.473) P − P i=1 o ! Y   1 N P νi = RT ln + ln i , (5.474) − K P P i=1 o ! Y   1 N P νi = RT ln i , (5.475) − K P P i=1  o  ! YN νi N ν Po Pi=1 ρ RT i = RT ln RT i , (5.476) −  K P   c i=1 o Y    N  1 = RT ln ρ νi . (5.477) − K i c i=1 ! Y Equation (5.477) is the common definition of affinity. Another form can be found by em- ploying the definition of Kc from Eq. (5.270) to get

N − νi o N P Pi=1 ∆G α = RT ln o exp ρ νi , (5.478) − RT RT i     i=1 ! N Y o − νi N ∆G P Pi=1 = RT + ln o ρ νi , (5.479) − RT RT i   i=1 !! N Y − νi N P Pi=1 = ∆Go RT ln o ρ νi . (5.480) − − i RT i=1 !   Y To see clearly that the entropy production rate is positive semi-definite, substitute Eq. (5.477) into Eq. (5.466) to get

N N N dS V ν′ 1 1 = k(T ) ρ i 1 ρ νi RT ln ρ νi 0, dt T i − K i − K i ≥ U,V i=1 ! c i=1 ! c i=1 !! Y Y Y (5.481)

N N N ν′ 1 1 = RVk(T ) ρ i 1 ρ νi ln ρ νi 0. (5.482) − i − K i K i ≥ i=1 ! c i=1 ! c i=1 ! Y Y Y CC BY-NC-ND. 28 January 2019, J. M. Powers. 5.7. SIMPLE ONE-STEP KINETICS 199

Define forward and reverse reaction coefficients, ′, and ′′, respectively, as R R N ′ ′ k(T ) ρ νi , (5.483) R ≡ i i=1 YN k(T ) ′′ ′′ ρ νi . (5.484) R ≡ K i c i=1 Y Both ′ and ′′ have units of kmole/m3/s. It is easy to see that R R r = ′ ′′. (5.485) R − R Note that since k(T ) > 0, K > 0, and ρ 0, that both ′ 0 and ′′ 0. Since c i ≥ R ≥ R ≥ ν = ν′′ ν′, one finds that i i − i N N ′′ 1 1 k(T ) ν′′−ν′ ρ νi = ρ i i = R . (5.486) K i K k(T ) i ′ c i=1 c i=1 Y Y R Then Eq. (5.482) reduces to dS ′′ ′′ = RV ′ 1 R ln R 0. (5.487) dt − R − ′ ′ ≥ U,V     R R

Finally, we get dS ′ = RV ( ′ ′′) ln R 0. (5.488) dt R − R ′′ ≥ U,V   R

Obviously, if the forward rate is greater than the reverse rate ′ ′′ > 0, ln( ′/ ′′) > 0, R − R R R and the entropy production is positive. If the forward rate is less than the reverse rate, ′ ′′ < 0, ln( ′/ ′′) < 0, and the entropy production is still positive. The production R − R R R rate is zero when ′ = ′′. Note that theR affinityRα can be written as

α = RT ln R′′ . (5.489) R  And so when the forward reaction rate exceeds the reverse, the affinity is positive. It is zero at equilibrium, when the forward reaction rate equals the reverse.

5.7 Simple one-step kinetics

A common model in theoretical combustion is that of so-called simple one-step kinetics. Such a model, in which the molecular mass does not change, is quantitatively appropriate only

CC BY-NC-ND. 28 January 2019, J. M. Powers. 200 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

for isomerization reactions. However, as a pedagogical tool as well as a qualitative model for real chemistry, it can be valuable. Consider the reversible reaction A ⇌ B. (5.490) where chemical species A and B have identical molecular masses MA = MB = M. Consider further the case in which at the initial state, no moles of A only are present. Also take the reaction to be isochoric and isothermal. These assumptions can easily be relaxed for more general cases. Specializing then Eq. (5.240) for this case, one has

nA = νA ζ + nAo , (5.491)

=−1 =no nB = νB ζ + nB . (5.492) |{z} |{z}o =1 =0 Thus |{z} |{z} nA = ζ + n , (5.493) − o nB = ζ. (5.494)

Now no is constant throughout the reaction. Scale by this and define the dimensionless reaction progress as ζˆ ζ/n to get ≡ o nA = ζˆ +1, (5.495) no −

=yA nB |{z} = ζ.ˆ (5.496) no

=yB In terms of the mole fractions then, one|{z} has yA = 1 ζ,ˆ (5.497) − yB = ζ.ˆ (5.498) The reaction kinetics for each species reduce to

dρA no = r, ρA(0) = ρ , (5.499) dt − V ≡ o dρB = r, ρB(0) = 0. (5.500) dt Addition of Eqs. (5.499) and (5.500) gives d (ρA + ρB) = 0, (5.501) dt ρA + ρB = ρo, (5.502) ρA ρB + = 1. (5.503) ρo ρo

=yA =yB

CC BY-NC-ND. 28 January 2019, J. M. Powers.|{z} |{z} 5.7. SIMPLE ONE-STEP KINETICS 201

In terms of the mole fractions yi, one then has

yA + yB =1. (5.504)

The reaction rate r is then

1 ρB r = kρA 1 , (5.505) − K ρA  c  ρA 1 ρB/ρo = kρo 1 , (5.506) ρ − K ρA/ρ o  c o  1 yB = kρoyA 1 , (5.507) − K yA  c  ˆ ˆ 1 ζ = kρo(1 ζ) 1 . (5.508) − − Kc 1 ζˆ! − ˆ ˆ ˆ Now r = (1/V )dζ/dt = (1/V )d(noζ)/dt = (no/V )d(ζ)/dt = ρodζ/dt. So the reaction dynamics can be described by a single ordinary differential equation in a single unknown:

ˆ ˆ dζ ˆ 1 ζ ρo = kρo(1 ζ) 1 , (5.509) dt − − Kc 1 ζˆ! − dζˆ 1 ζˆ = k(1 ζˆ) 1 . (5.510) dt − − Kc 1 ζˆ! − Equation (5.510) is in equilibrium when

ˆ 1 1 ζ = 1 1 + ... (5.511) 1+ ∼ − Kc Kc

As K , the equilibrium value of ζˆ 1. In this limit, the reaction is irreversible. That c →∞ → is, the species B is preferred over A. Equation (5.510) has exact solution

1 1 exp k 1+ K t ζˆ = − − c . (5.512) 1+1   Kc

For k > 0, Kc > 0, the equilibrium is stable. The time constant of relaxation τ is 1 τ = . (5.513) k 1+ 1 Kc   For the isothermal, isochoric system, one should consider the second law in terms of the Helmholtz free energy. Combine then Eq. (4.394), dA 0, with Eq. (4.311), dA = |T,V ≤

CC BY-NC-ND. 28 January 2019, J. M. Powers. 202 CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

SdT PdV + N µ dn and taking time derivatives, one finds − − i=1 i i P N dA T,V = SdT PdV + µidni 0, (5.514) | − − ! ≤ i=1 T,V X dA N dn = µ i 0, (5.515) dt i dt ≤ T,V i=1 X 1 dA V N dρ = µ i 0. (5.516) −T dt − T i dt ≥ i=1 X This is exactly the same form as Eq. (5.482), which can be directly substituted into Eq. (5.516) to give

N N N 1 dA ν′ 1 1 = RVk(T ) ρ i 1 ρ νi ln ρ νi 0, −T dt − i − K i K i ≥ T,V i=1 ! c i=1 ! c i=1 ! Y Y Y (5.517) N N N dA ν′ 1 1 = RVTk(T ) ρ i 1 ρ νi ln ρ νi 0. dt i − K i K i ≤ T,V i=1 ! c i=1 ! c i=1 ! Y Y Y (5.518)

For the assumptions of this section, Eq. (5.518) reduces to

dA 1 ζˆ 1 ζˆ = RTkρ V (1 ζˆ) 1 ln 0, (5.519) dt o − − K ˆ K ˆ ≤ T,V c 1 ζ ! c 1 ζ ! − − ˆ ˆ 1 ζ 1 ζ = knoRT (1 ζˆ) 1 ln 0. (5.520) − − Kc 1 ζˆ! Kc 1 ζˆ! ≤ − − Since the present analysis is nothing more than a special case of the previous section, Eq. (5.520) certainly holds. One questions however the behavior in the irreversible limit, 1/K 0. Evaluating this limit, one finds c →

dA ˆ 1 ˆ ˆ ˆ ˆ lim = knoRT (1 ζ) ln +(1 ζ) ln ζ (1 ζ) ln(1 ζ)+ ...  0. 1/Kc→0 dt − K − − − − ≤ T,V  c   >0   →−∞  | {z } (5.521) Now, performing the distinguished limit| {z as }ζˆ 1; that is the reaction goes to completion, → one notes that all terms are driven to zero for small 1/Kc. Recall that 1 ζˆ goes to zero faster than ln(1 ζˆ) goes to . Note that the entropy inequality is ill-defined− for a formally − −∞ irreversible reaction with 1/Kc = 0.

CC BY-NC-ND. 28 January 2019, J. M. Powers. Chapter 6

Thermochemistry of multiple reactions

See Powers, Chapter 5 See Turns, Chapter 4, 5, 6 See Kuo, Chapters 1, 2 See Kondepudi and Prigogine, Chapter 16, 19

This chapter will extend notions associated with the thermodynamics of a single chemical reactions to systems in which many reactions occur simultaneously.

6.1 Summary of multiple reaction extensions

Consider now the reaction of N species, composed of L elements, in J reactions. This section will focus on the most common case in which J (N L), which is usually the case in large chemical kinetic systems in use in engineering models.≥ − While much of the analysis will only require J > 0, certain results will depend on J (N L). It is not difficult to study the complementary case where 0

L M = φ , i =1,...,N. (6.1) i Ml li Xl=1 However, each reaction has a stoichiometric coefficient. The jth reaction can be summarized in the following ways:

N N ′ ⇌ ′′ χiνij χiνij, j =1, . . . , J, (6.2) i=1 i=1 XN X

χiνij = 0, j =1, . . . , J. (6.3) i=1 X 203 204 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

Stoichiometry for the jth reaction and lth element is given by the extension of Eq. (5.24):

N

φliνij =0, l =1,...,L, j =1, . . . , J. (6.4) i=1 X The net change in Gibbs free energy and equilibrium constants of the jth reaction are defined by the extensions of Eqs. (5.268, 5.267, 5.270):

N ∆Go go ν , j =1, . . . , J, (6.5) j ≡ T,i ij i=1 X o ∆Gj KP,j exp − , j =1, . . . , J, (6.6) ≡ RT   N i=1 νij o Po P ∆Gj Kc,j exp − , j =1, . . . , J. (6.7) ≡ RT RT     The equilibrium of the jth reaction is given by the extension of Eq. (5.253):

N

µiνij =0, j =1, . . . , J, (6.8) i=1 X or the extension of Eq. (5.255):

N

giνij = 0, j =1, . . . , J. (6.9) i=1 X The multi-reaction extension of Eq. (4.521) for affinity is

N α = µ ν , j =1, . . . , J. (6.10) j − i ij i=1 X In terms of the chemical affinity of each reaction, the equilibrium condition is simply the extension of Eq. (5.254): αj =0, j =1, . . . , J. (6.11) At equilibrium, then the equilibrium constraints can be shown to reduce to the extension of Eq. (5.266):

N P νij K = i , j =1, . . . , J, (6.12) P,j P i=1 o Y  

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.1. SUMMARY OF MULTIPLE REACTION EXTENSIONS 205

or the extension of Eq. (5.274):

N νij Kc,j = ρi , j =1, . . . , J. (6.13) i=1 Y For isochoric reaction, the evolution of species concentration i due to the combined effect of J reactions is given by the extension of Eq. (5.279):

≡ω˙ i

J N ′ N dρi z β Ej }| νkj 1 νkj { = ν a T j exp − ρ 1 ρ , i =1,...,N. dt ij j k − K k j=1 RT ! c,j X   kY=1  Yk=1  ≡k (T )   j forward reaction  reverse reaction    | {z } |≡rj =(1{z/V )dζj}/dt | {z } | {z } (6.14) The extension to isobaric reactions is straightforward, and follows the same analysis as for a single reaction. Again, three intermediate variables which are in common usage have been defined. First one takes the reaction rate of the jth reaction to be the extension of Eq. (5.280)

N ′ N Ej ν 1 ν βj kj  kj  rj ajT exp − ρk 1 ρk , j =1, . . . , J, (6.15) ≡ RT ! − Kc,j   kY=1  Yk=1  ≡k (T )   j forward reaction  reverse reaction    or the extension| of Eq.{z (5.281)} | {z } | {z }

N ′ N ′′ Ej ν 1 ν βj  kj kj  rj = ajT exp − ρk ρk , j =1, . . . , J, (6.16) RT − Kc,j    kY=1 kY=1  ≡k (T ), Arrhenius rate   j forward reaction reverse reaction    | {z } | {zlaw} of mass| action{z } | {z } 1 dζ r = j . (6.17) j V dt th Here ζj is the reaction progress variable for the j reaction. Each reaction has a temperature-dependent rate function kj(T ), which is an extension of Eq. (5.282):

βj Ej kj(T ) ajT exp − , j =1, . . . , J. (6.18) ≡ RT  

CC BY-NC-ND. 28 January 2019, J. M. Powers. 206 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

The evolution rate of each species is given byω ˙ i, defined now as an extension of Eq. (5.284):

J ω˙ ν r , i =1,...,N. (6.19) i ≡ ij j j=1 X The multi-reaction extension of Eq. (5.239) for mole change in terms of progress variables is J

dni = νijdζj, i =1,...,N. (6.20) j=1 X One also has Eq. (5.243):

N dG = µ dn , (6.21) |T,P i i i=1 XN J

= µi νikdζk, (6.22) i=1 k=1 XN XJ ∂G ∂ζk = µi νik , (6.23) ∂ζj ∂ζj ζp i=1 k=1 X X N J

= µi νikδkj, (6.24) i=1 j=1 X X N

= µiνij, (6.25) i=1 X = α , j =1, . . . , J. (6.26) − j For a set of adiabatic, isochoric reactions, one can show the extension of Eq. (5.429) is

J dT rj∆Uj = j=1 , (6.27) dt −P ρcv where the energy change for a reaction ∆Uj is defined as the extension of Eq. (5.428):

N

∆Uj = uiνij, j =1, . . . , J. (6.28) i=1 X Similarly for a set of adiabatic, isobaric reactions, one can show the extension of Eq. (5.444):

J dT rj∆Hj = j=1 , (6.29) dt −P ρcP

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.1. SUMMARY OF MULTIPLE REACTION EXTENSIONS 207

where the enthalpy change for a reaction ∆Hj is defined as the extension of Eq. (5.443):

N

∆Hj = hiνij, j =1, . . . , J. (6.30) i=1 X

Moreover, the density and species concentration derivatives for an adiabatic, isobaric set can be shown to be extensions of Eqs. (5.455, 5.459):

dρ J N h = M r ν i 1 , (6.31) dt j ij c T − j=1 i=1 P X X   dρ J ∆H i = r ν + y j ∆n , (6.32) dt j ij i c T − j j=1 P X   

where

N

∆nj = νkj. (6.33) Xk=1

In a similar fashion to that shown for a single reaction, one can further sum over all reactions and prove that mixture mass is conserved, element mass and number are conserved.

Example 6.1

Show that element mass and number are conserved for the multi-reaction formulation.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 208 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

Start with Eq. (6.14) and expand as follows:

J dρ i = ν r , (6.34) dt ij j j=1 X dρ J φ i = φ ν r , (6.35) li dt li ij j j=1 X J d (φ ρ ) = φ ν r , (6.36) dt li i li ij j j=1 X N d N J (φ ρ ) = φ ν r , (6.37) dt li i li ij j i=1 i=1 j=1 X X X N J N d φ ρ = φ ν r , (6.38) dt li i li ij j i=1 ! j=1 i=1 X X X e =ρl | {z dρ e} J N l = r φ ν , (6.39) dt j li ij j=1 i=1 X X =0 dρ e l = 0,| l =1{z,...,L,} (6.40) dt d ( ρ e) = 0, l =1,...,L, (6.41) dt Ml l dρ e l = 0, l =1,...,L. (6.42) dt It is also straightforward to show that the mixture density is conserved for the multi-reaction, multi- component mixture: dρ =0. (6.43) dt

The proof of the Clausius-Duhem relationship for the second law is an extension of the single reaction result. Start with Eq. (5.460) and operate much as for a single reaction model.

1 N dS = µ dn 0, (6.44) |U,V −T i i ≥ i=1 X irreversible entropy production N dS V | {zdni 1 } = µ 0, (6.45) dt − T i dt V ≥ U,V i=1 X V N dρ = µ i 0, (6.46) − T i dt ≥ i=1 X CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.1. SUMMARY OF MULTIPLE REACTION EXTENSIONS 209

V N J = µ ν r 0, (6.47) − T i ij j ≥ i=1 j=1 X X V N J = µ ν r 0, (6.48) − T i ij j ≥ i=1 j=1 X X V J N = r µ ν 0, (6.49) − T j i ij ≥ j=1 i=1 X X J N N N V ν′ 1 = k ρ ij 1 ρνij µ ν 0, (6.50) − T j i − K i i ij ≥ j=1 i=1 c,j i=1 ! i=1 X Y Y X J N N N V ν′ 1 1 = k ρ ij 1 ρνij RT ln ρνij 0, − T j i − K i K i ≥ j=1 i=1 c,j i=1 ! c,j i=1 !! X Y Y Y (6.51) J N N N ν′ 1 1 = RV k ρ ij 1 ρνij ln ρνij 0. (6.52) − j i − K i K i ≥ j=1 i=1 c,j i=1 ! c,j i=1 ! X Y Y Y Note that Eq. (6.49) can also be written in terms of the affinities (see Eq. (6.10)) and reaction progress variables (see Eq. (6.17) as dS 1 J dζ = α j 0. (6.53) dt T j dt ≥ U,V j=1 X

Similar to the argument for a single reaction, if one defines extensions of Eqs. (5.483, 5.484) as N ν′ ′ = k ρ ij , (6.54) Rj j i i=1 YN k ν′′ ′′ = j ρ ij , (6.55) Rj K i c,j i=1 Y then it is easy to show that r = ′ ′′, (6.56) j Rj − Rj and we get the equivalent of Eq. (5.488):

dS J ′ = RV ′ ′′ ln Rj 0. (6.57) dt Rj − Rj ′′ ≥ U,V j=1 Rj  X 

Since kj(T ) > 0, R> 0, and V 0, and each term in the summation combines to be positive semi-definite, one sees that the≥ Clausius-Duhem inequality is guaranteed to be satisfied for multi-component reactions.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 210 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

6.2 Equilibrium conditions

For multicomponent mixtures undergoing multiple reactions, determining the equilibrium condition is more difficult. There are two primary approaches, both of which are essentially equivalent. The most straightforward method requires formal minimization of the Gibbs free energy of the mixture. It can be shown that this actually finds the equilibrium associated with all possible reactions.

6.2.1 Minimization of G via Lagrange multipliers

Recall Eq. (4.395), dG 0. Recall also Eq. (4.397), G = N g n . Since µ = g = |T,P ≤ i=1 i i i i ∂G N N , one also has G = µ ni. From Eq. (4.398), dG T,P = µ dni. Now ∂ni i=1 i P i=1 i P,T,nj |

one must also demand for a systemP coming to equilibrium that the elemePnt numbers are conserved. This can be achieved by requiring

N φ (n n )=0, l =1, . . . , L. (6.58) li io − i i=1 X

Here recall nio is the initial number of moles of species i in the mixture, and φli is the number of moles of element l in species i. If one interprets n n as ν , the negative of the net io − i − ij mole change, Eq. (6.58) becomes N φ ν = 0, equivalent to Eq. (6.4). − i=1 li ij One can now use the method of constrainedP optimization given by the method of Lagrange multipliers to extremize G subject to the constraints of element conservation. The extremum will be a minimum; this will not be proved, but it will be demonstrated. Define a set of L ∗ Lagrange multipliers λl. Next define an augmented Gibbs free energy function G , which is simply G plus the product of the Lagrange multipliers and the constraints:

L N G∗ = G + λ φ (n n ). (6.59) l lk ko − k Xl=1 Xk=1

Now when the constraints are satisfied, one has G∗ = G, so assuming the constraints can be satisfied, extremizing G is equivalent to extremizing G∗. To extremize G∗, take its differential

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.2. EQUILIBRIUM CONDITIONS 211

with respect to ni, with P , T and nj constant and set it to zero for each species:

∂G∗ ∂G ∂ L N = + λl φlk(nko nk) =0, i =1,...,N, ∂ni ∂ni ∂ni − T,P,nj T,P,nj T,P,nj l=1 k=1 ! X X = µi

| {z } (6.60) L N ∂nk = µi λl φlk =0, i =1,...,N, (6.61) − ∂ni l=1 k=1 T,P,nj X X δ ki

L N | {z } = µ λ φ δ =0, i =1,...,N, (6.62) i − l lk ki l=1 k=1 XL X = µ λ φ =0, i =1,...,N. (6.63) i − l li Xl=1 Next, for an ideal gas, one can expand the chemical potential so as to get

L o Pi µT,i + RT ln λlφli =0, i =1,...,N, (6.64) Po −   Xl=1 =µi | {z } L o  niP 1  µT,i + RT ln N λlφli =0, i =1,...,N. (6.65) n Po −  k=1 k !  l=1   X  =P   P i    N | {z } Recalling that k=1 nk = n, in summary then, one has N + L equations

P L o ni P µT,i + RT ln λlφli = 0, i =1,...,N, (6.66) n Po −   l=1 N X φ (n n ) = 0, l =1, . . . , L. (6.67) li io − i i=1 X

in N + L unknowns: ni, i =1,...,N, λl,l =1,...,L.

Example 6.2 Consider a previous example problem, see p. 145, in which

N2 + N2 ⇌ 2N + N2. (6.68)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 212 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

Take the reaction to be isothermal and isobaric with T = 6000 K and P = 100 kPa. Initially one has 1 kmole of N2 and 0 kmole of N. Use the extremization of Gibbs free energy to find the equilibrium composition. First find the chemical potentials at the reference pressure of each of the possible constituents. o o o µo = go = h T so = h + ∆h T so. (6.69) T,i i i − i 298,i i − i For each species, one then finds

kJ µo = 0+205848 (6000)(292.984) = 1552056 , (6.70) N2 − − kmole kJ µo = 472680+124590 (6000)(216.926) = 704286 . (6.71) N − − kmole To each of these one must add n P RT ln i , nP  o  to get the full chemical potential. Now P = Po = 100 kPa for this problem, so one only must consider RT =8.314(6000) = 49884 kJ/kmole. So, the chemical potentials are

nN2 µN2 = 1552056 + 49884ln , (6.72) − n + n 2  N N  nN µN = 704286 + 49884ln . (6.73) − n + n 2  N N  Then one adds on the Lagrange multiplier and then considers element conservation to get the following coupled set of nonlinear algebraic equations:

nN2 1552056 + 49884ln 2λN = 0, (6.74) − n + n 2 −  N N  nN 704286 + 49884ln λN = 0, (6.75) − n + n 2 −  N N  nN +2nN2 = 2. (6.76)

These non-linear equations are solved numerically to get

nN2 = 0.88214 kmole, (6.77)

nN = 0.2357 kmole, (6.78) kJ λ = 781934 . (6.79) N − kmole These agree with results found in the earlier example problem, see p. 145.

Example 6.3 Consider a mixture of 2 kmole of H2 and 1 kmole of O2 at T = 3000 K and P = 100 kPa. Assuming an isobaric and isothermal equilibration process with the products consisting of H2, O2, H2O, OH, H, and O, find the equilibrium concentrations. Consider the same mixture at T = 298 K and T = 1000 K.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.2. EQUILIBRIUM CONDITIONS 213

The first task is to find the chemical potentials of each species at the reference pressure and T = 3000 K. Here one can use the standard tables along with the general equation

o o o µo = go = h T so = h + ∆h T so. (6.80) T,i i i − i 298,i i − i

For each species, one then finds

kJ µo = 0+88724 3000(202.989) = 520242 , (6.81) H2 − − kmole kJ µo = 0+98013 3000(284.466) = 755385 , (6.82) O2 − − kmole kJ µo = 241826 + 126548 3000(286.504) = 974790 , (6.83) H2O − − − kmole kJ µo = 38987+89585 3000(256.825) = 641903 , (6.84) OH − − kmole kJ µo = 217999+56161 3000(162.707) = 213961 , (6.85) H − − kmole kJ µo = 249170+56574 3000(209.705) = 323371 . (6.86) O − − kmole

To each of these one must add

n P RT ln i , nP  o 

to get the full chemical potential. Now P = Po = 100 kPa for this problem, so one must only consider RT =8.314(3000) = 24942 kJ/kmole. So, the chemical potentials are

nH2 µH2 = 520243 + 24942ln , (6.87) − n 2 + n 2 + n 2 + n + n + n  H O H O OH H O  nO2 µO2 = 755385 + 24942ln , (6.88) − n 2 + n 2 + n 2 + n + n + n  H O H O OH H O  nH2O µH2O = 974790 + 24942ln , (6.89) − n 2 + n 2 + n 2 + n + n + n  H O H O OH H O  nOH µOH = 641903 + 24942ln , (6.90) − n 2 + n 2 + n 2 + n + n + n  H O H O OH H O  nH µH = 213961 + 24942ln , (6.91) − n 2 + n 2 + n 2 + n + n + n  H O H O OH H O  nO µO = 323371 + 24942ln . (6.92) − n 2 + n 2 + n 2 + n + n + n  H O H O OH H O 

Then one adds on the Lagrange multipliers and then considers element conservation to get the following

CC BY-NC-ND. 28 January 2019, J. M. Powers. 214 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

coupled set of nonlinear equations:

nH2 520243 + 24942ln 2λH = 0, (6.93) − n 2 + n 2 + n 2 + n + n + n −  H O H O OH H O  nO2 755385 + 24942ln 2λO = 0, (6.94) − n 2 + n 2 + n 2 + n + n + n −  H O H O OH H O  nH2O 974790 + 24942ln 2λH λO = 0, (6.95) − n 2 + n 2 + n 2 + n + n + n − −  H O H O OH H O  nOH 641903 + 24942ln λH λO = 0, (6.96) − n 2 + n 2 + n 2 + n + n + n − −  H O H O OH H O  nH 213961 + 24942ln λH = 0, (6.97) − n 2 + n 2 + n 2 + n + n + n −  H O H O OH H O  nO 323371 + 24942ln λO = 0, (6.98) − n 2 + n 2 + n 2 + n + n + n −  H O H O OH H O  2nH2 +2nH2O + nOH + nH = 4, (6.99)

2nO2 + nH2O + nOH + nO = 2. (6.100)

These non-linear algebraic equations can be solved numerically via a Newton-Raphson technique. The equations are sensitive to the initial guess, and one can use ones intuition to help guide the

selection. For example, one might expect to have nH2O somewhere near 2 kmole. Application of the Newton-Raphson iteration yields

n = 3.19 10−1 kmole, (6.101) H2 × n = 1.10 10−1 kmole, (6.102) O2 × n = 1.50 100 kmole, (6.103) H2O × n = 2.20 10−1 kmole, (6.104) OH × n = 1.36 10−1 kmole, (6.105) H × n = 5.74 10−2 kmole, (6.106) O × kJ λ = 2.85 105 , (6.107) H − × kmole kJ λ = 4.16 105 . (6.108) O − × kmole

At this relatively high value of temperature, all species considered have a relatively major presence. That is, there are no truly minor species.

Unless a very good guess is provided, it may be difficult to find a solution for this set of non- linear equations. Straightforward algebra allows the equations to be recast in a form which sometimes

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.2. EQUILIBRIUM CONDITIONS 215

converges more rapidly:

n 520243 λ 2 H2 = exp exp H , (6.109) n + n + n + n + n + n 24942 24942 H2 O2 H2O OH H O      n 755385 λ 2 O2 = exp exp O , (6.110) n + n + n + n + n + n 24942 24942 H2 O2 H2O OH H O      n 974790 λ λ 2 H2O = exp exp O exp H , (6.111) n + n + n + n + n + n 24942 24942 24942 H2 O2 H2O OH H O        n 641903 λ λ OH = exp exp O exp H , (6.112) n + n + n + n + n + n 24942 24942 24942 H2 O2 H2O OH H O       n 213961 λ H = exp exp H , (6.113) n + n + n + n + n + n 24942 24942 H2 O2 H2O OH H O     n 323371 λ O = exp exp O , (6.114) n + n + n + n + n + n 24942 24942 H2 O2 H2O OH H O     2nH2 +2nH2O + nOH + nH = 4, (6.115)

2nO2 + nH2O + nOH + nO = 2. (6.116)

Then solve these considering ni, exp(λO/24942), and exp(λH /24942) as unknowns. The same result is recovered, but a broader range of initial guesses converge to the correct solution. One can verify that this choice extremizes G by direct computation; moreover, this will show that the extremum is actually a minimum. In so doing, one must exercise care to see that element

conservation is retained. As an example, perturb the equilibrium solution above for nH2 and nH such that

n = 3.19 10−1 + ξ, (6.117) H2 × n = 1.36 10−1 2ξ. (6.118) H × − Leave all other species mole numbers the same. In this way, when ξ = 0, one has the original equilibrium solution. For ξ = 0, the solution moves off the equilibrium value in such a way that elements are 6 N conserved. Then one has G = i=1 µini = G(ξ). The difference G(ξ) G(0) is plotted in Fig. 6.1. When ξ = 0, there is no deviation from the value predicted by the Newton-Raphson− P iteration. Clearly when ξ = 0, G(ξ) G(0), takes on a minimum value, and so then does G(ξ). So the procedure works. − At the lower temperature, T = 298 K, application of the same procedure yields very different results:

n = 4.88 10−27 kmole, (6.119) H2 × n = 2.44 10−27 kmole, (6.120) O2 × n = 2.00 100 kmole, (6.121) H2O × n = 2.22 10−29 kmole, (6.122) OH × n = 2.29 10−49 kmole, (6.123) H × n = 1.67 10−54 kmole, (6.124) O × kJ λ = 9.54 104 , (6.125) H − × kmole kJ λ = 1.07 105 . (6.126) O − × kmole At the intermediate temperature, T = 1000 K, application of the same procedure shows the minor

CC BY-NC-ND. 28 January 2019, J. M. Powers. 216 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

G(ξ) - G(0) (kJ)

40

30

20

10

-0.01 -0.005 0.005 0.01 ξ (kmole)

Figure 6.1: Gibbs free energy variation as mixture composition is varied maintaining element conservation for mixture of H2, O2, H2O, OH, H, and O at T = 3000 K, P = 100 kP a.

species become slightly more prominent:

n = 4.99 10−7 kmole, (6.127) H2 × n = 2.44 10−7 kmole, (6.128) O2 × n = 2.00 100 kmole, (6.129) H2O × n = 2.09 10−8 kmole, (6.130) OH × n = 2.26 10−12 kmole, (6.131) H × n = 1.10 10−13 kmole, (6.132) O × kJ λ = 1.36 105 , (6.133) H − × kmole kJ λ = 1.77 105 . (6.134) O − × kmole

6.2.2 Equilibration of all reactions

In another equivalent method, if one commences with a multi-reaction model, one can require each reaction to be in equilibrium. This leads to a set of algebraic equations for rj = 0, which

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.2. EQUILIBRIUM CONDITIONS 217 from Eq. (6.16) leads to

N i=1 νij o N Po P ∆Gj νkj Kc,j = exp − = ρ , j =1, . . . , J. (6.135) RT RT k     kY=1 With some effort it can be shown that not all of the J equations are linearly independent. Moreover, they do not possess a unique solution. However, for closed systems, only one of the solutions is physical, as will be shown in the following section. The others typically involve non-physical, negative concentrations. Nevertheless, Eqs. (6.135) are entirely consistent with the predictions of the N + L equa- tions which arise from extremization of Gibbs free energy while enforcing element number constraints. This can be shown by beginning with Eq. (6.65), rewritten in terms of molar concentrations, and performing the following sequence of operations:

L o ni/V P µT,i + RT ln N λlφli = 0, i =1,...,N, (6.136) Po − k=1 nk/V ! l=1 XL o P ρi P µT,i + RT ln N λlφli = 0, i =1,...,N, (6.137) Po − k=1 ρk ! l=1 XL o P ρi P µT,i + RT ln λlφli = 0, i =1,...,N, (6.138) ρ Po −   l=1 XL o RT µT,i + RT ln ρi λlφli = 0, i =1,...,N, (6.139) Po −   l=1 XL o RT νijµT,i + νijRT ln ρi νij λlφli = 0, i =1,...,N, Po −   Xl=1 j =1, . . . , J, (6.140) N N RT N L ν µo + ν RT ln ρ ν λ φ = 0, j =1, . . . , J, (6.141) ij T,i ij i P − ij l li i=1 i=1 o i=1 X X   X Xl=1 o =∆Gj | {z } N RT L N ∆Go + RT ν ln ρ λ φ ν = 0, j =1, . . . , J, (6.142) j ij i P − l li ij i=1 o i=1 X   Xl=1 X =0 N RT ∆Go + RT ν ln | ρ {z } = 0, j =1, . . . , J. (6.143) j ij i P i=1 o X   Here, the stoichiometry for each reaction has been employed to remove the Lagrange multi-

CC BY-NC-ND. 28 January 2019, J. M. Powers. 218 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

pliers. Continue to find

N νij o RT ∆Gj ln ρi = , j =1, . . . , J, (6.144) Po − RT i=1   X ν N RT ij ∆Go exp ln ρ = exp j , j =1, . . . , J, (6.145) i P − i=1 o ! RT X     N νij o RT ∆Gj ρi = exp , j =1, . . . , J, (6.146) Po − RT i=1     YN νij N Pi=1 ∆Go RT νij j ρi = exp , j =1, . . . , J, (6.147) Po − RT   i=1   Y N N i=1 νij o Po P ∆Gj ρ νij = exp , j =1, . . . , J,(6.148) i − i=1 RT RT Y     =Kc,j

In summary, | {z }

N νij ρi = Kc,j, j =1, . . . , J, (6.149) i=1 Y which is identical to Eq. (6.135), obtained by equilibrating each of the J reactions. Thus, extremization of Gibbs free energy is consistent with equilibrating each of the J reactions.

6.3 Concise reaction rate law formulations

One can additional analysis to obtain a more efficient representation of the reaction rate law for multiple reactions. There are two important cases: 1) J (N L); this is most common ≥ − for large chemical kinetic systems, and 2) J < (N L); this is common for simple chemistry models. − The species production rate is given by Eq. (6.14), which reduces to

dρ 1 J dζ i = ν j , i =1,...,N. (6.150) dt V ij dt j=1 X Now recalling Eq. (5.99), one has

N−L dn = dξ , i =1,...,N. (6.151) i Dik k Xk=1 CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.3. CONCISE REACTION RATE LAW FORMULATIONS 219

Comparing then Eq. (6.151) to Eq. (6.20), one sees that

J N−L ν dζ = dξ , i =1,...,N, (6.152) ij j Dik k j=1 X Xk=1 − 1 J 1 N L ν dζ = dξ , i =1,...,N. (6.153) V ij j V Dik k j=1 X Xk=1 6.3.1 Reaction dominant: J (N L) ≥ − Consider first the most common case in which J (N L). One can say the species production rate is given ≥ −

− dρ 1 N L dξ J i = k = ν r , i =1,...,N. (6.154) dt V Dik dt ij j j=1 Xk=1 X

One would like to invert and solve directly for dξk/dt. However, ik is non-square and has no inverse. But since N φ = 0, and N φ ν = 0, L of theD equations N equations i=1 liDip i=1 li ij in Eq. (6.154) are redundant. At this point, it isP more convenient to goP to a Gibbs vector notation, where there is an obvious correspondence between the bold vectors and the indicial counterparts: dρ 1 dξ = D = ν r, (6.155) dt V · dt · dξ DT D = V DT ν r, (6.156) · · dt · · dξ = V (DT D)−1 DT ν r. (6.157) dt · · · · Because of the L linear dependencies, there is no loss of information in this matrix projection. This system of N L equations is the smallest number of differential equations that can be − solved for a general system in which J > (N L). Lastly, one recovers the original system when− forming dξ D = V D (DT D)−1 DT ν r. (6.158) · dt · · · · · =P Here, the N N projection matrix P is| symmetric,{z has norm} of unity, has rank of N L, has N L eigenvalues× of value unity, and L eigenvalues of value zero. And, while application− of a general− projection matrix to ν r filters some of the information in ν r, because the · · N (N L) matrix D spans the same column space as the N J matrix ν, no information is lost× in− Eq. (6.158) relative to the original Eq. (6.155). Mathematica× lly, one can say

P ν = ν. (6.159) ·

CC BY-NC-ND. 28 January 2019, J. M. Powers. 220 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

6.3.2 Species dominant: J < (N L) − Next consider the case in which J < (N L). This often arises in models of simple chemistry, for example one- or two-step kinetics. − The fundamental reaction dynamics are most concisely governed by the J equations which form

1 dζ = r. (6.160) V dt

However, r is a function of the concentrations; one must therefore recover ρ as a function of reaction progress ζ. In vector form, Eq. (6.150) is written as dρ 1 dζ = ν . (6.161) dt V · dt Take as an initial condition that the reaction progress is zero at t = 0 and that there are an appropriate set of initial conditions on the species concentrations ρ:

ζ = 0, t =0, (6.162)

ρ = ρo, t =0. (6.163)

Then, since ν is a constant, Eq. (6.161) is easily integrated. After applying the initial conditions, Eq. (6.163), one gets 1 ρ = ρ + ν ζ. (6.164) o V · Last, if J =(N L), either approach yields the same number of equations, and is equally concise. −

6.4 Adiabatic, isochoric kinetics

Here an example which uses multiple reactions for an adiabatic isochoric system is given.

Example 6.4 Consider the full time-dependency of a problem considered in a previous example in which the equilibrium state was found; see Sec. 5.3.3.4,. A closed, fixed, adiabatic volume contains at t = 0 s a stoichiometric mixture of 2 kmole of H2, 1 kmole of O2, and 8 kmole of N2 at 100 kPa and 1000 K. Find the reaction dynamics as the system proceeds from its initial state to its final state. This problem requires a detailed numerical solution. Such a solution was performed by solving Eq. (6.14) along with the associated calorically imperfect species state equations for a mixture of eighteen interacting species: H2, H, O, O2, OH, H2O, HO2, H2O2, N, NH2, NH3, N2H, NO, NO2, N2O, HNO, and N2. The equilibrium values were reported in a previous example. The dynamics of the reaction process are reflected in Fig. 6.2. At early time, t < 4 10−4 s, the × pressure, temperature, and major reactant species concentrations (H2, O2, N2) are nearly constant.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 6.4. ADIABATIC, ISOCHORIC KINETICS 221

0 H 10 2 H O 2 N O2 H 2 −5 2 O NO 10 - 2 −10 NO2 3 Oxygen-Related 10 N N 2O OH Species H −10 O 3 10 Evolution −20 HO 2 10 Nitrogen H O H O 2 2 (kmole/m ) i 2 NO and Nitrogen

ρ −15 −30

(kmole/m ) Oxides Species 10 i 10 ρ Evolution a H2 O 2 b −20 −40 10 10 −6 −4 −2 −6 −4 −2 10 10 10 10 10 10 t (s) t (s)

3000 250

2500 200 2000 T (K) P (kPa) 150 1500 induction time c d 1000 100 −6 −4 −2 −6 −4 −2 10 10 10 10 10 10 t (s) t (s)

Figure 6.2: Plot of a) ρH2 (t), ρH (t), ρO(t), ρO2 (t), ρOH (t), ρH2O(t), ρHO2 (t), ρH2O2 (t), b) ρN (t),

ρNO(t), ρNO2(t), ρN2O(t), ρN2 (t), c) T (t), and d) P (t) for adiabatic, isochoric combustion of a mixture of 2H2 + O2 +8N2 initially at T = 1000 K, P = 100 kP a.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 222 CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

However, the minor species, e.g. OH, NO, HO2 and the major product, H2O, are undergoing very rapid growth, albeit concentrations whose value remains small. In this period, the material is in what is known as the induction period. After a certain critical mass of minor species has accumulated, exothermic recombination of these minor species to form the major product H2O induces the temperature to rise, which accelerates further the reaction rates. This is manifested in a thermal explosion. A common definition of the end of the induction period is the induction time, t = tind, the time when dT/dt goes through a maximum. Here one finds t =4.53 10−4 s. (6.165) ind × At the end of the induction zone, there is a final relaxation to equilibrium.

CC BY-NC-ND. 28 January 2019, J. M. Powers. Chapter 7

Kinetics in some more detail

See Powers, Chapter 1

Here we give further details of kinetics. These notes are also used to introduce a separate combustion course and have some overlap with previous chapters. Let us consider the reaction of N molecular chemical species composed of L elements via J chemical reactions. Let us assume the gas is an ideal mixture of ideal gases that satisfies Dalton’s law of partial pressures. The reaction will be considered to be driven by molecular collisions. We will not model individual collisions, but instead attempt to capture their collective effect. An example of a model of such a reaction is listed in Table 7.1. There we find a N =9 species, J = 37 step irreversible reaction mechanism for an L = 3 hydrogen-oxygen-argon 1 2 mixture from Maas and Warnatz, with corrected fH2 from Maas and Pope. The model has 3 also been utilized by Fedkiw, et al. We need not worry yet about fH2 , which is known as a collision efficiency factor. The one-sided arrows indicate that each individual reaction is considered to be irreversible. Note that for nearly each reaction, a separate reverse reaction is listed; thus, pairs of irreversible reactions can in some sense be considered to model reversible reactions. In this model a set of elementary reactions are hypothesized. For the jth reaction we have the collision frequency factor aj, the temperature-dependency exponent βj and the activation energy Ej. These will be explained in short order. Other common forms exist. Often reactions systems are described as being composed of reversible reactions. Such reactions are usually notated by two sided arrows. One such system is reported by Powers and Paolucci4 reported here in Table 7.2. Both overall models are complicated.

1Maas, U., and Warnatz, J., 1988, “Ignition Processes in Hydrogen-Oxygen Mixtures,” Combustion and Flame, 74(1): 53-69. 2Maas, U., and Pope, S. B., 1992, “Simplifying Chemical Kinetics: Intrinsic Low-Dimensional Manifolds in Composition Space,” Combustion and Flame, 88(3-4): 239-264. 3Fedkiw, R. P., Merriman, B., and Osher, S., 1997, “High Accuracy Numerical Methods for Thermally Flows with Chemistry,” Journal of Computational Physics, 132(2): 175-190. 4Powers, J. M., and Paolucci, S., 2005, “Accurate Spatial Resolution Estimates for Reactive Supersonic

223 224 CHAPTER 7. KINETICS IN SOME MORE DETAIL

− ′ − N ′ 3 (1 νM,j Pi=1 νij ) j Reaction a (mol/cm ) β E kJ j s Kβj j j mole  14  1 O2 + H OH + O 2.00 10 0.00 70.30  → × 13 2 OH + O O2 + H 1.46 10 0.00 2.08 → × 4 3 H2 + O OH + H 5.06 10 2.67 26.30 → × 4 4 OH + H H2 + O 2.24 10 2.67 18.40 → × 8 5 H2 + OH H2O + H 1.00 10 1.60 13.80 → × 8 6 H2O + H H2 + OH 4.45 10 1.60 77.13 → × 9 7 OH + OH H2O + O 1.50 10 1.14 0.42 → × 10 8 H2O + O OH + OH 1.51 10 1.14 71.64 → × 18 9 H + H + M H2 + M 1.80 10 1.00 0.00 → × 18 − 10 H2 + M H + H + M 6.99 10 1.00 436.08 → × 22 − 11 H + OH + M H2O + M 2.20 10 2.00 0.00 → × 23 − 12 H2O + M H + OH + M 3.80 10 2.00 499.41 → × 17 − 13 O + O + M O2 + M 2.90 10 1.00 0.00 → × 18 − 14 O2 + M O + O + M 6.81 10 1.00 496.41 → × 18 − 15 H + O2 + M HO2 + M 2.30 10 0.80 0.00 → × 18 − 16 HO2 + M H + O2 + M 3.26 10 0.80 195.88 → × 14 − 17 HO2 + H OH + OH 1.50 10 0.00 4.20 → × 13 18 OH + OH HO2 + H 1.33 10 0.00 168.30 → × 13 19 HO2 + H H2 + O2 2.50 10 0.00 2.90 → × 13 20 H2 + O2 HO2 + H 6.84 10 0.00 243.10 → × 13 21 HO2 + H H2O + O 3.00 10 0.00 7.20 → × 13 22 H2O + O HO2 + H 2.67 10 0.00 242.52 → × 13 23 HO2 + O OH + O2 1.80 10 0.00 1.70 → × 13 − 24 OH + O2 HO2 + O 2.18 10 0.00 230.61 → × 13 25 HO2 + OH H2O + O2 6.00 10 0.00 0.00 → × 14 26 H2O + O2 HO2 + OH 7.31 10 0.00 303.53 → × 11 27 HO2 + HO2 H2O2 + O2 2.50 10 0.00 5.20 → × 22 − 28 OH + OH + M H2O2 + M 3.25 10 2.00 0.00 → × 24 − 29 H2O2 + M OH + OH + M 2.10 10 2.00 206.80 → × 12 − 30 H2O2 + H H2 + HO2 1.70 10 0.00 15.70 → × 12 31 H2 + HO2 H2O2 + H 1.15 10 0.00 80.88 → × 13 32 H2O2 + H H2O + OH 1.00 10 0.00 15.00 → × 12 33 H2O + OH H2O2 + H 2.67 10 0.00 307.51 → × 13 34 H2O2 + O OH + HO2 2.80 10 0.00 26.80 → × 12 35 OH + HO2 H2O2 + O 8.40 10 0.00 84.09 36 H O + OH→ H O + HO 5.40 × 1012 0.00 4.20 2 2 → 2 2 × 37 H O + HO H O + OH 1.63 1013 0.00 132.71 2 2 → 2 2 ×

Table 7.1: Third body collision efficiencies with M are fH2 = 1.00, fO2 = 0.35, and fH2O =6.5. CC BY-NC-ND. 28 January 2019, J. M. Powers. 225

− ′ − N ′ 3 (1 νM,j Pi=1 νij ) j Reaction a (mol/cm ) β E cal j s Kβj j j mole  13  1 H2 + O2 ⇋ OH + OH 1.70 10 0.00 47780 × 9 2 OH + H2 ⇋ H2O + H 1.17 10 1.30 3626 × 16 3 H + O2 ⇋ OH + O 5.13 10 0.82 16507 × 10 − 4 O + H2 ⇋ OH + H 1.80 10 1.00 8826 × 18 5 H + O2 + M ⇋ HO2 + M 2.10 10 1.00 0 × 19 − 6 H + O2 + O2 ⇋ HO2 + O2 6.70 10 1.42 0 × 19 − 7 H + O2 + N2 ⇋ HO2 + N2 6.70 10 1.42 0 × 13 − 8 OH + HO2 ⇋ H2O + O2 5.00 10 0.00 1000 × 14 9 H + HO2 ⇋ OH + OH 2.50 10 0.00 1900 × 13 10 O + HO2 ⇋ O2 + OH 4.80 10 0.00 1000 × 8 11 OH + OH ⇋ O + H2O 6.00 10 1.30 0 × 12 12 H2 + M ⇋ H + H + M 2.23 10 0.50 92600 × 11 13 O2 + M ⇋ O + O + M 1.85 10 0.50 95560 × 23 14 H + OH + M ⇋ H2O + M 7.50 10 2.60 0 × 13 − 15 H + HO2 ⇋ H2 + O2 2.50 10 0.00 700 × 12 16 HO2 + HO2 ⇋ H2O2 + O2 2.00 10 0.00 0 × 17 17 H2O2 + M ⇋ OH + OH + M 1.30 10 0.00 45500 18 H O + H ⇋ HO + H 1.60 × 1012 0.00 3800 2 2 2 2 × 19 H O + OH ⇋ H O + HO 1.00 1013 0.00 1800 2 2 2 2 ×

Table 7.2: Nine species, nineteen step reversible reaction mechanism for an H2/O2/N2 mix- ture. Third body collision efficiencies with M are f5(H2O) = 21, f5(H2)=3.3, f12(H2O) = 6, f12(H) = 2, f12(H2) = 3, f14(H2O) = 20.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 226 CHAPTER 7. KINETICS IN SOME MORE DETAIL

7.1 Isothermal, isochoric kinetics

For simplicity, we will first focus attention on cases in which the temperature T and vol- ume V are both constant. Such assumptions are known as “isothermal” and “isochoric,” respectively. A nice fundamental treatment of elementary reactions of this type is given by Vincenti and Kruger in their detailed monograph.5

7.1.1 O O dissociation − 2 One of the simplest physical examples is provided by the dissociation of O2 into its atomic component O.

7.1.1.1 Pair of irreversible reactions To get started, let us focus for now only on reactions 13 and 14 from Table 7.1 in the limiting case in which temperature T and volume V are constant.

7.1.1.1.1 Mathematical model The reactions describe oxygen dissociation and recom- bination in a pair of irreversible reactions:

13 : O + O + M O + M, (7.1) → 2 14 : O + M O + O + M, (7.2) 2 → with mole −2 K kJ a =2.90 1017 , β = 1.00, E =0 (7.3) 13 × cm3 s 13 − 13 mole   mole −1 K kJ a =6.81 1018 , β = 1.00, E = 496.41 . (7.4) 14 × cm3 s 14 − 14 mole   The irreversibility is indicated by the one-sided arrow. Though they participate in the overall hydrogen oxidation problem, these two reactions are in fact self-contained as well. So let us just consider that we have only oxygen in our box with N = 2 species, O2 and O, J = 2 reactions (those being 13 and 14), and L = 1 element, that being O. Recall that in the cgs system, common in thermochemistry, that 1 erg = 1 dyne cm = 10−7 J = 10−10 kJ. Recall also that the cgs unit of force is the dyne and that 1 dyne = 1 g cm/s2 = 10−5 N. So for cgs we have

erg kJ 1010 erg erg E =0 , E = 496.41 =4.96 1012 . (7.5) 13 mole 14 mole kJ × mole   Flow with Detailed Chemistry,” AIAA Journal, 43(5): 1088-1099. 5W. G. Vincenti and C. H. Kruger, 1965, Introduction to Physical Gas Dynamics, Wiley, New York.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 227

The standard model for chemical reaction induces the following two ordinary differential equations for the evolution of O and O2 molar concentrations:

dρO β13 E13 β14 E14 = 2 a13T exp − ρ ρ ρ +2 a14T exp − ρ ρ , (7.6) dt − RT O O M RT O2 M     =k13(T ) =k14(T )

| {z =r13 } | {z =r14 } dρ O2 β13 E13 β14 E14 = a13T| exp − {z ρ ρ ρ }a14T| exp − {z ρ ρ . } (7.7) dt RT O O M − RT O2 M     =k13(T ) =k14(T )

| {z =r13 } | {z =r14 }

Here we use the notation| ρi as the{z molar concentration} | of species{zi. Also a common} usage for

molar concentration is given by square brackets, e.g. ρO2 = [O2]. The symbol M represents an arbitrary third body and is an inert participant in the reaction. We also use the common notation of a temperature-dependent portion of the reaction rate for reaction j, kj(T ), where

βj Ej kj(T )= ajT exp . (7.8) RT   The reaction rates for reactions 13 and 14 are defined as

r13 = k13ρOρOρM , (7.9)

r14 = k14ρO2 ρM . (7.10) We will give details of how to generalize this form later. The system Eq. (7.6-7.7) can be written simply as dρ O = 2r +2r , (7.11) dt − 13 14 dρ O2 = r r . (7.12) dt 13 − 14 Even more simply, in vector form, Eqs. (7.11-7.12) can be written as dρ = ν r. (7.13) dt · Here we have taken ρ ρ = O , (7.14) ρ  O2  2 2 ν = , (7.15) −1 1  −  r r = 13 . (7.16) r  14 CC BY-NC-ND. 28 January 2019, J. M. Powers. 228 CHAPTER 7. KINETICS IN SOME MORE DETAIL

In general, we will have ρ be a column vector of dimension N 1, ν will be a rectangular × matrix of dimension N J of rank R, and r will be a column vector of length J 1. So Eqs. (7.11-7.12) take the× form × d ρ 2 2 r O = − 13 . (7.17) dt ρ 1 1 r14  O2   −    Note here that the rank R of ν is R = L = 1. Let us also define a stoichiometric matrix φ of dimension L N. The component of φ, φli represents the number of element l in species i. Generally φ will× be full rank, which will vary since we can have L < N, L = N, or L > N. Here we have L < N and φ is of dimension 1 2: × φ = 1 2 . (7.18) Element conservation is guaranteed by insisting  that ν be constructed such that φ ν = 0. (7.19) · So we can say that each of the column vectors of ν lies in the right null space of φ. For our example, we see that Eq. (7.19) holds: 2 2 φ ν = 1 2 − = 0 0 . (7.20) · · 1 1  −    The symbol R is the universal gas constant, where J 107 erg erg R =8.31441 =8.31441 107 . (7.21) mole K J × mole K   Let us take as initial conditions

ρO(t =0)= ρO, ρO2 (t =0)= ρO2 . (7.22) Now M represents an arbitrary third body, so here b b

ρM = ρO2 + ρO. (7.23) Thus, the ordinary differential equations of the reaction dynamics reduce to

dρO β13 E13 = 2a13T exp − ρ ρ ρ + ρ dt − RT O O O2 O    β14 E14 +2a14T exp − ρ ρ + ρ , (7.24) RT O2 O2 O   dρ  O2 β13 E13 = a13T exp − ρ ρ ρ + ρ dt RT O O O2 O    β14 E14 a14T exp − ρ ρ + ρ . (7.25) − RT O2 O2 O    CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 229

Equations (7.24-7.25) with Eqs. (7.22) represent two non-linear ordinary differential equa-

tions with initial conditions in two unknowns ρO and ρO2 . We seek the behavior of these two species concentrations as a function of time. Systems of non-linear equations are generally difficult to integrate analytically and gen- erally require numerical solution. Before embarking on a numerical solution, we simplify as much as we can. Note that dρ dρ O +2 O2 = 0, (7.26) dt dt d ρ +2ρ = 0. (7.27) dt O O2 We can integrate and apply the initial conditions (7.22) to get

ρO +2ρO2 = ρO +2ρO2 = constant. (7.28) The fact that this algebraic constraint exists for all time is a consequence of the conservation b b of mass of each O element. It can also be thought of as the conservation of number of O atoms. Such notions always hold for chemical reactions. They do not hold for nuclear reactions. Standard linear algebra provides a robust way to find the constraint of Eq. (7.28). We can use elementary row operations to cast Eq. (7.16) into a row-echelon form. Here our goal is to get a linear combination which on the right side has an upper triangular form. To achieve this add twice the second equation with the first to form a new equation to replace the second equation. This gives d ρ 2 2 r O = − 13 . (7.29) dt ρ +2ρ 0 0 r14  O O2     

Obviously the second equation is one we obtained earlier, d/dt(ρO +2ρO2 ) = 0, and this induces our algebraic constraint. We also note the system can be recast as 1 0 d ρ 2 2 r O = − 13 . (7.30) 1 2 dt ρ 0 0 r14    O2      This is of the matrix form dρ L−1 P = U r. (7.31) · · dt · Here L and L−1 are N N lower triangular matrices of full rank N, and thus invertible. The matrix U is upper triangular× of dimension N J and with the same rank as ν, R L. × ≥ The matrix P is a permutation matrix of dimension N N. It is never singular and thus always invertable. It is used to effect possible row exchanges× to achieve the desired form; often row exchanges are not necessary, in which case P = I, the N N identity matrix. × Equation (7.31) can be manipulated to form the original equation via dρ = P−1 L U r. (7.32) dt · · · =ν

| {z CC} BY-NC-ND. 28 January 2019, J. M. Powers. 230 CHAPTER 7. KINETICS IN SOME MORE DETAIL

What we have done is the standard linear algebra decomposition of ν = P−1 L U. · · We can also decompose the algebraic constraint, Eq. (7.28), in a non-obvious way that is more readily useful for larger systems. We can write 1 ρ = ρ ρ ρ . (7.33) O2 O2 − 2 O − O   Defining now ξ = ρ ρ , we can sayb b O O − O ρ ρ 1 b O = O + ξ . (7.34) ρ ρ 1 O  O2   O2  − 2  b =D =ξ =ρ =ρ b b |{z} | {z } | {z } | {z } This gives the dependent variables in terms of a smaller number of transformed dependent variables in a way which satisfies the linear constraints. In vector form, the equation becomes

ρ = ρ + D ξ. (7.35) · Here D is a full rank matrix which spansb the same column space as does ν. Note that ν may or may not be full rank. Since D spans the same column space as does ν, we must also have in general

φ D = 0. (7.36) · We see here this is true: 1 1 2 = (0). (7.37) · 1 − 2   We also note that the term exp( Ej/RT ) is a modulating factor to the dynamics. Let us see how this behaves for high and− low temperatures. First for low temperature, we have

E lim exp − j =0. (7.38) T →0 RT   At high temperature, we have

E lim exp − j =1. (7.39) T →∞ RT   And lastly, at intermediate temperature, we have

E E exp − j (1) when T = j . (7.40) RT ∼ O O R     A sketch of this modulating factor is given in Figure 7.1. Note

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 231

E exp(- j /(R T)) 1

E T j / R

Figure 7.1: Plot of exp( E /R/T ) versus T ; transition occurs at T E /R. − j ∼ j

for small T , the modulation is extreme, and the reaction rate is very small, • for T E /R, the reaction rate is extremely sensitive to temperature, and • ∼ j for T , the modulation is unity, and the reaction rate is limited only by molecular • collision→∞ frequency.

3 Now ρO and ρO2 represent molar concentrations which have standard units of mole/cm . So the reaction rates dρ dρ O and O2 dt dt have units of mole/cm3/s. Note that after conversion of Ej from kJ/mole to erg/mole we find the units of the argument of the exponential to be unitless. That is

E erg mole K 1 j = dimensionless. (7.41) RT mole erg K ⇒   Here the brackets denote the units of a quantity, and not molar concentration. Let us get units for the collision frequency factor of reaction 13, a13. We know the units of the rate 3 (mole/cm /s). Reaction 13 involves three molar species. Since β13 = 1, it also has an extra temperature dependency. The exponential of a unitless number is unitless,− so we need not worry about that. For units to match, we must have

mole mole mole mole = [a ] K−1. (7.42) cm3 s 13 cm3 cm3 cm3         CC BY-NC-ND. 28 January 2019, J. M. Powers. 232 CHAPTER 7. KINETICS IN SOME MORE DETAIL

So the units of a13 are mole −2 K [a ]= . (7.43) 13 cm3 s   For a14 we find a different set of units! Following the same procedure, we get mole mole mole = [a ] K−1. (7.44) cm3 s 14 cm3 cm3       So the units of a14 are mole −1 K [a ]= . (7.45) 14 cm3 s   This discrepancy in the units of aj the molecular collision frequency factor is a burden of tra- ditional chemical kinetics, and causes many difficulties when classical non-dimensionalization is performed. With much effort, a cleaner theory could be formulated; however, this would require significant work to re-cast the now-standard aj values for literally thousands of re- actions which are well established in the literature.

7.1.1.1.2 Example calculation Let us consider an example problem. Let us take T = 3 3 5000 K, and initial conditions ρO = 0.001 mole/cm and ρO2 = 0.001 mole/cm . The initial temperature is very hot, and is near the temperature of the surface of the sun. This is also realizable in laboratory conditions,b but uncommon inb most combustion engineering environments. We can solve these in a variety of ways. I chose here to solve both Eqs. (7.24-7.25) without the reduction provided by Eq. (7.28). However, we can check after numerical solution to see if Eq. (7.28) is actually satisfied. Substituting numerical values for all the constants to get −2 β13 E13 7 mole K −1 2a13T exp − = 2 2.9 10 (5000 K) exp(0), − RT − × cm3 s     ! mole −2 1 = 1.16 1014 , (7.46) − × cm3 s   −1 β14 E14 18 mole K −1 2a14T exp − = 2 6.81 10 (5000 K) RT × cm3 s     ! 4.96 1012 erg exp − × mole , × 8.31441 107 erg (5000 K)  × mole K  mole −1 1 = 1.77548 1010 , (7.47) × cm3 s  −2 β13 E13 13 mole 1 a13T exp − = 5.80 10 , (7.48) RT × cm3 s     −1 β14 E14 9 mole 1 a14T exp − = 8.8774 10 . (7.49) − RT − × cm3 s     CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 233

ρ , ρ 3 O O2 (mole/cm ) 0.00150

O2 0.00100

0.00070

0.00050

O

t HsL 10 -11 10 -10 10 -9 10 -8 10 -7 10 -6

Figure 7.2: Molar concentrations versus time for oxygen dissociation problem.

Then the differential equation system becomes dρ O = (1.16 1014)ρ2 (ρ + ρ )+(1.77548 1010)ρ (ρ + ρ ), (7.50) dt − × O O O2 × O2 O O2 dρ O2 = (5.80 1013)ρ2 (ρ + ρ ) (8.8774 109)ρ (ρ + ρ ), (7.51) dt × O O O2 − × O2 O O2 mole ρ (0) = 0.001 , (7.52) O cm3 mole ρ (0) = 0.001 . (7.53) O2 cm3 These non-linear ordinary differential equations are in a standard form for a wide variety of numerical software tools. Solution of such equations are not the topic of these notes.

7.1.1.1.2.1 Species concentration versus time A solution was obtained numeri- cally, and a plot of ρO(t) and ρO2 (t) is given in Figure 7.2. Note that significant reaction does not commence until t 10−10 s. This can be shown to be very close to the time between ∼ molecular collisions. For 10−9 s 10−7 s, e the reaction appears to be equilibrated. The calculation gives the equilibrium values ρO and e ρO2 , as

e mole lim ρO = ρO =0.0004424 , (7.54) t→∞ cm3 e mole lim ρO = ρO =0.00127 , (7.55) t→∞ 2 2 cm3

Note that at this high temperature, O2 is preferred over O, but there are definitely O molecules present at equilibrium.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 234 CHAPTER 7. KINETICS IN SOME MORE DETAIL

r

- 1.5´ 10 16

- 1.0´ 10 16

- 7.0´ 10 17

- 5.0´ 10 17

- 3.0´ 10 17

t HsL - - - - - 10 10 10 9 10 8 10 7 10 6

Figure 7.3: Dimensionless residual numerical error r in satisfying the element conservation constraint in the oxygen dissociation example.

We can check how well the numerical solution satisfied the algebraic constraint of element conservation by plotting the dimensionless residual error r

ρ +2ρ ρ 2ρ r = O O2 − O − O2 , (7.56) ρ +2ρ O O2 b b as a function of time. If the constraint is exactlyb b satisfied, we will have r = 0. Any non-zero r will be related to the numerical method we have chosen. It may contain roundoff error and have a sporadic nature. A plot of r(t) is given in Figure 7.3. Clearly the error is small, and has the character of a roundoff error. In fact it is possible to drive r to be smaller by controlling the error tolerance in the numerical method.

7.1.1.1.2.2 Pressure versus time We can use the ideal gas law to calculate the pressure. Recall that the ideal gas law for molecular species i is

PiV = niRT. (7.57)

Here Pi is the partial pressure of molecular species i, and ni is the number of moles of molecular species i. Note that we also have n P = i RT. (7.58) i V Note that by our definition of molecular species concentration that n ρ = i . (7.59) i V

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 235

So we also have the ideal gas law as

Pi = ρiRT. (7.60) Now in the Dalton mixture model, all species share the same T and V . So the mixture temperature and volume are the same for each species Vi = V , Ti = T . But the mixture pressure is taken to be the sum of the partial pressures:

N

P = Pi. (7.61) i=1 X Substituting from Eq. (7.60) into Eq. (7.61), we get

N N

P = ρiRT = RT ρi. (7.62) i=1 i=1 X X For our example, we only have two species, so

P = RT (ρO + ρO2 ). (7.63) The pressure at the initial state t = 0 is

P (t =0) = RT (ρO + ρO2 ), (7.64) erg mole mole = 8.31441 107 (5000 K) 0.001 +0.001 , (7.65) b ×b mole K cm3 cm3    dyne  = 8.31441 108 , (7.66) × cm2 = 8.31441 102 bar. (7.67) × This pressure is over 800 . It is actually a little too high for good experimental correlation with the underlying data, but we will neglect that for this exercise. At the equilibrium state we have more O2 and less O. And we have a different number of molecules, so we expect the pressure to be different. At equilibrium, the pressure is

P (t ) = lim RT (ρO + ρO ), (7.68) →∞ t→∞ 2 erg mole mole = 8.31441 107 (5000 K) 0.0004424 +0.00127 , × mole K cm3 cm3     (7.69) dyne = 7.15 108 , (7.70) × cm2 = 7.15 102 bar. (7.71) × The pressure has dropped because much of the O has recombined to form O2. Thus there are fewer molecules at equilibrium. The temperature and volume have remained the same. A plot of P (t) is given in Figure 7.4.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 236 CHAPTER 7. KINETICS IN SOME MORE DETAIL

P (dyne/cm 2) 8.4´ 10 8

8.2´ 10 8

8. ´ 10 8

7.8´ 10 8

7.6´ 10 8

7.4´ 10 8

7.2´ 10 8

t HsL ------10 11 10 10 10 9 10 8 10 7 10 6

Figure 7.4: Pressure versus time for oxygen dissociation example.

7.1.1.1.2.3 Dynamical system form Now Eqs. (7.50-7.51) are of the standard form for an autonomous dynamical system:

dy = f(y). (7.72) dt

y T f Here is the vector of state variables (ρO, ρO2 ) . And is an algebraic function of the state variables. For the isothermal system, the algebraic function is in fact a polynomial. Equilibrium

The dynamical system is in equilibrium when

f(y)= 0. (7.73)

This non-linear set of algebraic equations can be difficult to solve for large systems. For common chemical kinetics systems, such as the one we are dealing with, there is a guarantee of a unique equilibrium for which all state variables are physical. There are certainly other equilibria for which at least one of the state variables is non-physical. Such equilibria can be quite mathematically complicated. Solving Eq. (7.73) for our oxygen dissociation problem gives us symbolically from Eq. (7.6-

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 237

7.7)

E13 e e e β13 E14 e e β14 2a13 exp − ρ ρ ρ T +2a14 exp − ρ ρ T = 0, (7.74) − RT O O M RT O2 M    

β13 E13 e e e β14 E14 e e a13T exp − ρ ρ ρ a14T exp − ρ ρ = 0. (7.75) RT O O M − RT O2 M     e We notice that ρM cancels. This so-called third body will in fact never affect the equilib- e rium state. It will however influence the dynamics. Removing ρM and slightly rearranging Eqs. (7.74-7.75) gives

β13 E13 e e β14 E14 e a13T exp − ρ ρ = a14T exp − ρ , (7.76) RT O O RT O2    

β13 E13 e e β14 E14 e a13T exp − ρ ρ = a14T exp − ρ . (7.77) RT O O RT O2     e These are the same equations! So we really have two unknowns for the equilibrium state ρO e and ρO2 but seemingly only one equation. Note that rearranging either Eq. (7.76) or (7.77) gives the result

β14 −E14 e e a14T exp ρOρO RT e = = K(T ). (7.78)  −E13  ρO2 β13 a13T exp RT   That is, for the net reaction (excluding the inert third body), O2 O+O, at equilibrium the product of the concentrations of the products divided by the pro→duct of the concentrations of the reactants is a function of temperature T . And for constant T , this is the so-called equilibrium constant. This is a famous result from basic chemistry. It is actually not complete yet, as we have not taken advantage of a connection with thermodynamics. But for now, it will suffice. We still have a problem: Eq. (7.78) is still one equation for two unknowns. We solve this be recalling we have not yet taken advantage of our algebraic constraint of element e e conservation, Eq. (7.28). Let us use the equation to eliminate ρO2 in favor of ρO: 1 ρe = ρ ρe + ρ . (7.79) O2 2 O − O O2   So Eq. (7.76) reduces to b b

β13 E13 e e β14 E14 1 e a13T exp − ρ ρ = a14T exp − ρ ρ + ρ . (7.80) RT O O RT 2 O − O O2         =ρe b O2 b Equation (7.80) is one algebraic equation in one unknown. Its| solution gives{z the equilibrium} e e value ρO. It is a quadratic equation for ρO. Of its two roots, one will be physical. We note

CC BY-NC-ND. 28 January 2019, J. M. Powers. 238 CHAPTER 7. KINETICS IN SOME MORE DETAIL

3 f( ρO) (mole/cm /s)

300000

200000

100000

3 -0.004 -0.003 -0.002 -0.001 0.001 ρO (mole/cm ) -100000

-200000

Figure 7.5: Equilibria for oxygen dissociation example.

that the equilibrium state will be a function of the initial conditions. Mathematically this is because our system is really best posed as a system of differential-algebraic equations. Systems which are purely differential equations will have equilibria which are independent of their initial conditions. Most of the literature of mathematical physics focuses on such systems of those. One of the foundational complications of chemical dynamics is the the equilibria is a function of the initial conditions, and this renders many common mathematical notions from traditional dynamic system theory to be invalid Fortunately, after one accounts for the linear constraints of element conservation, one can return to classical notions from traditional dynamic .

Consider the dynamics of Eq. (7.24) for the evolution of ρO. Equilibrating the right hand

side of this equation, gives Eq. (7.74). Eliminating ρM and then ρO2 in Eq. (7.74) then substituting in numerical parameters gives the cubic algebraic equation 33948.3 (1.78439 1011)(ρ )2 (5.8 1013)(ρ )3 = f(ρ )=0. (7.81) − × O − × O O

This equation is cubic because we did not remove the effect of ρM . This will not affect the equilibrium, but will affect the dynamics. We can get an idea of where the roots are

by plotting f(ρO) as seen in Figure 7.5. Zero crossings of f(ρO) in Figure 7.5 represent e e equilibria of the system, ρO, f(ρO) = 0. The cubic equation has three roots mole ρe = 0.003 , non-physical, (7.82) O − cm3 mole ρe = 0.000518944 , non-physical, (7.83) O − cm3 mole ρe = 0.000442414 , physical. (7.84) O cm3 Note the physical root found by our algebraic analysis is identical to that which was identified by our numerical integration of the ordinary differential equations of reaction kinetics. Stability of equilibria

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 239

We can get a simple estimate of the stability of the equilibria by considering the slope of f near f = 0. Our dynamic system is of the form dρ O = f(ρ ). (7.85) dt O Near the first non-physical root at ρe = 0.003, a positive perturbation from equi- • O − librium induces f < 0, which induces dρO/dt < 0, so ρO returns to its equilibrium. Similarly, a negative perturbation from equilibrium induces dρO/dt > 0, so the system returns to equilibrium. This non-physical equilibrium point is stable. Note stability does not imply physicality!

e Perform the same exercise for the non-physical root at ρO = 0.000518944. We find • this root is unstable. −

e Perform the same exercise for the physical root at ρO = 0.000442414. We find this • root is stable. In general if f crosses zero with a positive slope, the equilibrium is unstable. Otherwise, it is stable. Consider a formal Taylor series expansion of Eq. (7.85) in the neighborhood of an equi- 3 librium point ρO:

d e e df e (ρO ρO)= f(ρO) + (ρO ρO)+ ... (7.86) dt − dρO ρ =ρe − O O =0

We find df/dρO by differentiating| Eq.{z (7.81)} to get

df 11 14 2 = (3.56877 10 )ρO (1.74 10 )ρO. (7.87) dρO − × − ×

We evaluate df/dρO near the physical equilibrium point at ρO =0.0004442414 to get df = (3.56877 1011)(0.0004442414) (1.74 1014)(0.0004442414)2, dρO − × − × 1 = 1.91945 108 . (7.88) − × s Thus the Taylor series expansion of Eq. (7.24) in the neighborhood of the physical equi- librium gives the local kinetics to be driven by d (ρ 0.000442414) = (1.91945 108)(ρ 0.0004442414) + .... (7.89) dt O − − × O − So in the neighborhood of the physical equilibrium we have

ρ =0.0004442414 + A exp 1.91945 108t . (7.90) O − ×  CC BY-NC-ND. 28 January 2019, J. M. Powers. 240 CHAPTER 7. KINETICS IN SOME MORE DETAIL

Here A is an arbitrary constant of integration. The local time constant which governs the times scales of local evolution is τ where 1 τ = =5.20983 10−9 s. (7.91) 1.91945 108 × × This nano-second time scale is very fast. It can be shown to be correlated with the mean time between collisions of molecules.

7.1.1.1.3 Effect of temperature Let us perform four case studies to see the effect of T on the system’s equilibria and it dynamics near equilibrium.

T = 3000 K. Here we have significantly reduced the temperature, but it is still higher • than typically found in ordinary combustion engineering environments. Here we find

mole ρe = 8.9371 10−6 , (7.92) O × cm3 τ = 1.92059 10−7 s. (7.93) × The equilibrium concentration of O dropped by two orders of magnitude relative to T = 5000 K, and the time scale of the dynamics near equilibrium slowed by two orders of magnitude.

T = 1000 K. Here we reduce the temperature more. This temperature is common in • combustion engineering environments. We find

mole ρe = 2.0356 10−14 , (7.94) O × cm3 τ = 2.82331 101 s. (7.95) × The O concentration at equilibrium is greatly diminished to the point of being difficult to detect by standard measurement techniques. And the time scale of combustion has significantly slowed.

T = 300 K. This is obviously near . We find • mole ρe = 1.14199 10−44 , (7.96) O × cm3 τ = 1.50977 1031 s. (7.97) × The O concentration is effectively zero at room temperature, and the relaxation time is effectively infinite. As the oldest star in our galaxy has an age of 4.4 1017 s, we see × that at this temperature, our mathematical model cannot be experimentally validated, so it loses its meaning. At such a low temperature, the theory becomes qualitatively correct, but not quantitatively predictive.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 241

T = 10000 K. Such high temperature could be achieved in an atmospheric re-entry • environment. mole ρe = 2.74807 10−3 , (7.98) O × cm3 τ = 1.69119 10−10 s. (7.99) × At this high temperature, O become preferred over O2, and the time scales of reaction become extremely small, under a nanosecond.

7.1.1.2 Single reversible reaction The two irreversible reactions studied in the previous section are of a class that is common in combustion modeling. However, the model suffers a defect in that its link to classical equilibrium thermodynamics is missing. A better way to model essentially the same physics and guarantee consistency with classical equilibrium thermodynamics is to model the process as a single reversible reaction, with a suitably modified reaction rate term.

7.1.1.2.1 Mathematical model

7.1.1.2.1.1 Kinetics For the reversible O O2 reaction, let us only consider reaction 13 from Table 7.2 for which −

13 : O2 + M ⇌ O + O + M. (7.100) For this system, we have N = 2 molecular species in L = 1 elements reacting in J = 1 reaction. Here mole −1 cal a =1.85 1011 (K)−0.5, β =0.5, E = 95560 . (7.101) 13 × cm3 13 13 mole   Units of cal are common in chemistry, but we need to convert to erg, which is achieved via cal 4.186 J 107 erg erg E = 95560 =4.00014 1012 . (7.102) 13 mole cal J × mole       For this reversible reaction, we slightly modify the kinetics equations to

dρO β13 E13 1 = 2 a13T exp − ρO2 ρM ρOρOρM , (7.103) dt RT − Kc,13     =k13(T )

| {z } =r13 dρ E 1 O2 | β13 13 {z } = a13T exp − ρO2 ρM ρOρOρM . (7.104) dt − RT − Kc,13     =k13(T )

| {z } =r13

| CC{z BY-NC-ND. 28 January} 2019, J. M. Powers. 242 CHAPTER 7. KINETICS IN SOME MORE DETAIL

Here we have used equivalent definitions for k13(T ) and r13, so that Eqs. (7.103-7.104) can be written compactly as dρ O = 2r , (7.105) dt 13 dρ O2 = r . (7.106) dt − 13 In matrix form, we can simplify to

d ρ 2 O = (r ). (7.107) dt ρ 1 13  O2  −  =ν Here the N J or 2 1 matrix ν is | {z } × × 2 ν = . (7.108) 1 −  Performing row operations, the matrix form reduces to

d ρ 2 O = (r ), (7.109) dt ρ +2ρ 0 13  O O2    or 1 0 d ρ 2 O = (r ). (7.110) 1 2 dt ρ 0 13    O2    So here the N N or 2 2 matrix L−1 is × × 1 0 L−1 = . (7.111) 1 2   The N N or 2 2 permutation matrix P is the identity matrix. And the N J or 2 1 upper triangular× × matrix U is × × 2 U = . (7.112) 0   Note that ν = L U or equivalently L−1 ν = U: · · 1 0 2 2 = . (7.113) 1 2 · 1 0   −    =L−1 =ν =U Once again the stoichiometric matrix| {z φ} is| {z } |{z}

φ = 1 2 . (7.114)  CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 243

And we see that φ ν = 0 is satisfied: · 2 1 2 = 0 . (7.115) · 1 −    As for the irreversible reactions, the reversible reaction rates are constructed to conserve O atoms. We have d ρ +2ρ =0. (7.116) dt O O2  Thus, we once again find

ρO +2ρO2 = ρO +2ρO2 = constant. (7.117)

As before, we can say b b ρ ρ 1 O = O + ξ . (7.118) ρ ρ 1 O  O2   O2  − 2  b =D =ξ =ρ =ρ b b |{z} | {z } | {z } | {z } This gives the dependent variables in terms of a smaller number of transformed dependent variables in a way which satisfies the linear constraints. In vector form, the equation becomes

ρ = ρ + D ξ. (7.119) · Once again φ D = 0. · b 7.1.1.2.1.2 Thermodynamics Equations (7.103-7.104) are supplemented by an ex- pression for the thermodynamics-based equilibrium constant Kc,13 which is:

o Po ∆G13 Kc,13 = exp − . (7.120) RT RT   6 2 Here Po = 1.01326 10 dyne/cm = 1 atm is the reference pressure. The net change of × o Gibbs free energy at the reference pressure for reaction 13, ∆G13 is defined as

∆Go =2go go . (7.121) 13 O − O2 We further recall that the Gibbs free energy for species i at the reference pressure is defined in terms of the enthalpy and entropy as

o go = h T so. (7.122) i i − i o o It is common to find hi and si in thermodynamic tables tabulated as functions of T .

CC BY-NC-ND. 28 January 2019, J. M. Powers. 244 CHAPTER 7. KINETICS IN SOME MORE DETAIL

We further note that both Eqs. (7.103) and (7.104) are in equilibrium when

e e 1 e e e ρO2 ρM = ρOρOρM . (7.123) Kc,13 We rearrange Eq. (7.123) to find the familiar ρe ρe [products] K = O O = . (7.124) c,13 ρe [reactants] O2 Q If Kc,13 > 1, the products are preferred. If Kc,Q13 < 1, the reactants are preferred. Now, Kc,13 is a function of T only, so it is known. But Eq. (7.124) once again is one equation in two unknowns. We can use the element conservation constraint, Eq. (7.117) to reduce to one equation and one unknown, valid at equilibrium: e e ρOρO Kc,13 = (7.125) ρ + 1 (ρ ρe ) O2 2 O − O Using the element constraint, Eq. (7.117), we can recast the dynamics of our system by modifying Eq. (7.103) into one equationb in oneb unknown:

dρO β13 E13 = 2a13T exp − dt RT   1 1 1 1 (ρO2 + (ρO ρO)) (ρO2 + (ρO ρO)+ ρO) ρOρO (ρO2 + (ρO ρO)+ ρO) , × 2 − 2 − −Kc,13 2 −  =ρ =ρ =ρ   b bO2 b b M b b M    | {z } | {z } | {z (7.126)}

7.1.1.2.2 Example calculation Let us consider the same example as the previous sec- tion with T = 5000 K. We need numbers for all of the parameters of Eq. (7.126). For O, we find at T = 5000 K that o erg h = 3.48382 1012 , (7.127) O × mole erg so = 2.20458 109 . (7.128) O × mole K So erg erg go = 3.48382 1012 (5000 K) 2.20458 109 , O × mole − × mole K  erg    = 7.53908 1012 . (7.129) − × mole

For O2, we find at T = 5000 K that o erg h = 1.80749 1012 , (7.130) O2 × mole erg so = 3.05406 109 . (7.131) O2 × mole K

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 245

So erg erg go = 1.80749 1012 (5000 K) 3.05406 109 O2 × mole − × mole K  erg    = 1.34628 1013 . (7.132) − × mole

Thus, by Eq. (7.121), we have erg ∆Go = 2( 7.53908 1012) ( 1.34628 1013)= 1.61536 1012 . (7.133) 13 − × − − × − × mole

Thus, by Eq. (7.120) we get for our system

1.01326 106 dyne K = × cm2 c,13 8.31441 107 erg (5000 K) × mole K erg 1.61536 1012 exp − × mole , (7.134) × − 8.31441 107 erg (5000 K) × mole K !! mole = 1.187 10−4 .  (7.135) × cm3

Substitution of all numerical parameters into Eq. (7.126) and expansion yields the fol- lowing

dρ O = 3899.47 (2.23342 1010)ρ2 (7.3003 1012)ρ3 = f(ρ ), ρ (0) = 0.001.(7.136) dt − × O − × O O O

A plot of the time-dependent behavior of ρO and ρO2 from solution of Eq. (7.136) is given in Figure 7.6. The behavior is similar to the predictions given by the pair of irreversible reactions in Fig. 7.1. Here direct calculation of the equilibrium from time integration reveals

mole ρe =0.000393328 . (7.137) O cm3

Using Eq. (7.117) we find this corresponds to

mole ρe =0.00130334 . (7.138) O2 cm3

We note the system begins significant reaction for t 10−9 s and is equilibrated for t ∼ ∼ 10−7 s. The equilibrium is verified by solving the algebraic equation

f(ρ ) = 3899.47 (2.23342 1010)ρ2 (7.3003 1012)ρ3 =0. (7.139) O − × O − × O

CC BY-NC-ND. 28 January 2019, J. M. Powers. 246 CHAPTER 7. KINETICS IN SOME MORE DETAIL

3 ρO, ρO2 (mole/cm ) 0.00150

O2 0.00100

0.00070

0.00050

O 0.00030

t HsL 10 -11 10 -9 10 -7 10 -5 0.001

Figure 7.6: Plot of ρO(t) and ρO2 (t) for oxygen dissociation with reversible reaction.

This yields three roots: mole ρe = 0.003 , non-physical, (7.140) O − cm3 mole ρe = 0.000452678 , non-physical, (7.141) O − cm3 mole ρe = 0.000393328 , physical, (7.142) O cm3 (7.143) is given in Figure 7.6. Linearizing Eq. (7.136) in the neighborhood of the physical equilibrium yields the equa- tion d (ρ 0.000393328) = (2.09575 107)(ρ 0.000393328) + ... (7.144) dt O − − × O − This has solution ρ =0.000393328 + A exp 2.09575 107t . (7.145) O − × Again, A is an arbitrary constant. Obviously the equilibrium is stable. Moreover the time constant of relaxation to equilibrium is 1 τ = =4.77156 10−8 s. (7.146) 2.09575 107 × × CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 247

3 f(ρ O) (mole/cm /s)

40000

30000

20000

10000

3 -0.004 -0.003 -0.002 -0.001 0.001 ρO (mole/cm ) -10000

-20000

Figure 7.7: Plot of f(ρO) versus ρO for oxygen dissociation with reversible reaction.

This is consistent with the time scale to equilibrium which comes from integrating the full equation.

7.1.2 Zel’dovich mechanism of NO production

Let us consider next a more complicated reaction system: that of NO production known as the Zel’dovich6 mechanism. This is an important model for the production of a major pollutant from combustion processes. It is most important for high temperature applications.

7.1.2.1 Mathematical model

The model has several versions. One is

1 : N + NO ⇌ N2 + O, (7.147)

2 : N + O2 ⇌ NO + O. (7.148)

similar to our results for O2 dissociation, N2 and O2 are preferred at low temperature. As the temperature rises N and O begin to appear, and it is possible when they are mixed for NO to appear as a product.

6Yakov Borisovich Zel’dovich, 1915-1987, prolific Soviet physicist and father of thermonuclear weapons.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 248 CHAPTER 7. KINETICS IN SOME MORE DETAIL

7.1.2.1.1 Standard model form Here we have the reaction of N = 5 molecular species with

ρNO ρ  N  ρ = ρN2 . (7.149)  ρ   O   ρ   O2    We have L = 2 with N and O as the 2 elements. The stoichiometric matrix φ of dimension L N =2 5 is × × 11200 φ = . (7.150) 10012   The first row of φ is for the N atom; the second row is for the O atom. And we have J = 2 reactions. The reaction vector of length J = 2 is

β1 Ta,1 1 r a1T exp T ρN ρNO K ρN2 ρO r = 1 = − − c,1 , (7.151) r β2 Ta,2 1 2 a2T exp   ρ ρ ρ ρ    − T N O2 − Kc,2 NO O  1     k1 ρN ρNO ρN ρO = − Kc,1 2 . (7.152) 1 k2 ρ ρ ρ ρ  N O2 − Kc,2 NO O    Here, we have

T k = a T β1 exp a,1 , (7.153) 1 1 − T   Ta,2 k = a T β2 exp . (7.154) 2 2 − T   In matrix form, the model can be written as

ρ 1 1 NO − ρN 1 1 d   − −  r1 ρN2 = 1 0 . (7.155) dt r2  ρ   1 1     O     ρ   0 1  O2   −     =ν  Here the matrix ν has dimension N J which| {z is 5 } 2. The model is of our general form × × dρ = ν r. (7.156) dt ·

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 249

Note that our stoichiometric constraint on element conservation for each reaction φ ν = 0 · holds here: 1 1 − 1 1 11200 − −  0 0 φ ν = 1 0 = . (7.157) · 10012 · 0 0    1 1       0 1  −  We get 4 zeros because there are 2 reactions each with 2 element constraints.

7.1.2.1.2 Reduced form Here we describe non-traditional, but useful reductions, using standard techniques from linear algebra to bring the model equations into a reduced form in which all of the linear constraints have been explicitly removed. Let us perform a series of row operations to find all of the linear dependencies. Our aim is to convert the ν matrix into an upper triangular form. The lower left corner of ν already has a zero, so there is no need to worry about it. Let us add the first and fourth equations to eliminate the 1 in the 4, 1 slot. This gives ρ 1 1 NO − ρN 1 1 d   − −  r1 ρN2 = 1 0 . (7.158) dt r2 ρ + ρ   0 2     NO O    ρ   0 1  O2   −  Next, add the first and third equations to get  ρ 1 1 NO − ρN 1 1 d   − −  r1 ρNO + ρN2 = 0 1 . (7.159) dt r2  ρ + ρ   0 2     NO O     ρ   0 1  O2   −  Now multiply the first equation by 1 and add it to the second to get − ρ 1 1 NO − ρNO + ρN 0 2 d −   −  r1 ρNO + ρN2 = 0 1 . (7.160) dt r2  ρ + ρ   0 2     NO O     ρ   0 1  O2   −  Next multiply the fifth equation by 2 and add it to the second to get − ρ 1 1 NO − ρNO + ρN 0 2 d  −   −  r1 ρNO + ρN2 = 0 1 . (7.161) dt r2  ρ + ρ   0 2     NO O     ρ + ρ 2ρ   0 0  − NO N − O2        CC BY-NC-ND. 28 January 2019, J. M. Powers. 250 CHAPTER 7. KINETICS IN SOME MORE DETAIL

Next add the second and fourth equations to get

ρ 1 1 NO − ρNO + ρN 0 2 d  −   −  r1 ρNO + ρN2 = 0 1 . (7.162) dt r2  ρ + ρ   0 0     N O     ρ + ρ 2ρ   0 0  − NO N − O2        Next multiply the third equation by 2 and add it to the second to get

ρ 1 1 NO − ρNO + ρN 0 2 d  −   −  r1 ρNO + ρN +2ρN2 = 0 0 . (7.163) dt r2  ρ + ρ   0 0     N O     ρ + ρ 2ρ   0 0  − NO N − O2        Rewritten, this becomes

1 000 0 ρ 1 1 NO − 1100 0 ρN 0 2 −  d    −  r1 1 120 0 ρN2 = 0 0 . (7.164) dt r2  0 101 0   ρ   0 0       O     1100 2  ρ   0 0  − −   O2     =L−1     =U 

A way to think of| this type{z of row echelon} form is that| it{z defines} two free variables, those associated with the non-zero pivots of U: ρNO and ρN . The remain three variables ρN2 , ρO and ρO2 are bound variables which can be expressed in terms of the free variables. The last three of the ordinary differential equations are homogeneous and can be inte- grated to form

ρNO + ρN +2ρN2 = C1, (7.165)

ρN + ρO = C2, (7.166) ρ + ρ 2ρ = C . (7.167) − NO N − O2 3

The constants C1, C2 and C3 are determined from the initial conditions on all five state variables. In matrix form, we can say

ρNO 1 120 0 ρ C  N  1 0 101 0 ρ = C . (7.168)   N2  2 1100 2  ρ  C − −  O  3    ρ     O2    CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 251

Considering the free variables, ρNO and ρN , to be known, we move them to the right side to get

20 0 ρ C ρ ρ N2 1 − NO − N 01 0 ρO = C2 ρN . (7.169) 0 0 2 ρ  C + ρ− ρ  − O2 3 NO − N       Solving, for the bound variables, we find

ρ 1 C 1 ρ 1 ρ N2 2 1 − 2 NO − 2 N ρ = C ρ . (7.170)  O   2 N  ρ 1 C 1−ρ + 1 ρ O2 − 2 3 − 2 NO 2 N     We can rewrite this as

1 1 1 ρN2 2 C1 2 2 − − ρNO ρO = C2 + 0 1 . (7.171)    1   1 −1  ρN ρ C3   O2 − 2 − 2 2       We can get a more elegant form by defining ξNO = ρNO and ξN = ρN . Thus we can say our state variables have the form

ρNO 0 1 0 ρN 0 0 1    1   1 1  ξNO ρN2 = 2 C1 + 2 2 . (7.172) − − ξN  ρ   C2   0 1    O     −   ρ   1 C   1 1   O2  − 2 3 − 2 2        By translating via ξNO = ξNO + ρNO and ξN = ξN + ρN and choosing the constants C1, C2 and C3 appropriately, we can arrive at b b ρNO ρNO 1 0 ρ ρ 0 1 N  N  ξ   b  1 1  NO ρN2 = ρN2 + 2 2 . (7.173)   − − ξN  ρO   bρO   0 1        1 −1   ρO  bρ   =ξ  2   O2  − 2 2        =ρ  b  =D | {z } =ρ b b | {z } | {z } | {z } This takes the form of

ρ = ρ + D ξ. (7.174) · Here the matrix D is of dimension N Rb, which here is 5 2. It spans the same column space as does the N J matrix ν which× is of rank R. Here in× fact R = J = 2, so D has the × same dimension as ν. In general it will not. If c1 and c2 are the column vectors of D, we

CC BY-NC-ND. 28 January 2019, J. M. Powers. 252 CHAPTER 7. KINETICS IN SOME MORE DETAIL

see that c c forms the first column vector of ν and c c forms the second column − 1 − 2 1 − 2 vector of ν. Note that φ D = 0: · 1 0 0 1 11200   0 0 φ D = 1 1 = . (7.175) · 10012 · − 2 − 2 0 0    0 1    −   1 1  − 2 2    Equations (7.165-7.167) can also be linearly combined in a way which has strong phys- ical relevance. We rewrite the system as three equations in which the first is identical to Eq. (7.165); the second is the difference of Eqs. (7.166) and (7.167); and the third is half of Eq. (7.165) minus half of Eq. (7.167) plus Eq. (7.166):

ρNO + ρN +2ρN2 = C1, (7.176) ρ + ρ +2ρ = C C , (7.177) O NO O2 2 − 3 1 ρ + ρ + ρ + ρ + ρ = (C C )+ C . (7.178) NO N N2 O O2 2 1 − 3 2 Equation (7.176) insists that the number of nitrogen elements be constant; Eq. (7.177) demands the number of oxygen elements be constant; and Eq. (7.178) requires the number of moles of molecular species be constant. For general reactions, including the earlier studied oxygen dissociation problem, the number of moles of molecular species will not be constant. Here because each reaction considered has two molecules reacting to form two molecules, we are guaranteed the number of moles will be constant. Hence, we get an additional linear constraint beyond the two for element conservation. Note that since our reaction is isothermal, isochoric and mole-preserving, it will also be isobaric.

7.1.2.1.3 Example calculation Let us consider an isothermal reaction at T = 6000 K. (7.179) The high temperature is useful in generating results which are easily visualized. It insures that there will be significant concentrations of all molecular species. Let us also take as an initial condition mole ρ = ρ = ρ = ρ = ρ =1 10−6 . (7.180) NO N N2 O O2 × cm3 For this temperatureb and concentrations,b b b theb pressure, which will remain constant through the reaction, is P =2.4942 106 dyne/cm2. This is a little greater than atmospheric. Kinetic data for this reaction× is adopted from Baulch, et al.7 The data for reaction 1 is mole −1 1 a =2.107 1013 , β =0, T =0 K. (7.181) 1 × cm3 s 1 a1   7Baulch, et al., 2005, “Evaluated Kinetic Data for Combustion Modeling: Supplement II,” Journal of Physical and Chemical Reference Data, 34(3): 757-1397.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 253

For reaction 2, we have

mole −1 1 a =5.8394 109 , β =1.01, T = 3120 K. (7.182) 2 × cm3 K1.01 s 2 a2  

Here the so-called activation temperature Ta,j for reaction j is really the activation energy scaled by the universal gas constant:

Ej Ta,j = . (7.183) R Substituting numbers we obtain for the reaction rates

0 mole −1 1 k = (2.107 1013)(6000)0 exp − =2.107 1013 , (7.184) 1 × 6000 × cm3 s     3120 mole −1 1 k = (5.8394 109)(6000)1.01 exp − =2.27231 1013 . 2 × 6000 × cm3 s     (7.185)

We will also need thermodynamic data. The data here will be taken from the Chemkin database.8 Thermodynamic data for common materials is also found in most thermodynamic texts. For our system at 6000 K, we find erg go = 1.58757 1013 , (7.186) NO − × mole erg go = 7.04286 1012 , (7.187) N − × mole erg go = 1.55206 1013 , (7.188) N2 − × mole erg go = 9.77148 1012 , (7.189) O − × mole erg go = 1.65653 1013 . (7.190) O2 − × mole o Thus for each reaction, we find ∆Gj : ∆Go = go + go go go , (7.191) 1 N2 O − N − NO = 1.55206 1013 9.77148 1012 +7.04286 1012 +1.58757 1013,(7.192) − × −erg × × × = 2.37351 1012 , (7.193) − × mole ∆Go = go + go go go , (7.194) 2 NO O − N − O2 = 1.58757 1013 9.77148 1012 +7.04286 1012 +1.65653 1013,(7.195) − × −erg × × × = 2.03897 1012 . (7.196) − × mole 8R. J. Kee, et al., 2000, “The Chemkin Thermodynamic Data Base,” part of the Chemkin Collection Release 3.6, Reaction Design, San Diego, CA.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 254 CHAPTER 7. KINETICS IN SOME MORE DETAIL

At 6000 K, we find the equilibrium constants for the J = 2 reactions are

o ∆G1 Kc,1 = exp − , (7.197) RT   2.37351 1012 = exp × , (7.198) (8.314 107)(6000)  ×  = 116.52, (7.199) o ∆G2 Kc,2 = exp − , (7.200) RT   2.03897 1012 = exp × , (7.201) (8.314 107)(6000)  ×  = 59.5861. (7.202)

Again, omitting details, we find the two differential equations governing the evolution of the free variables are

dρ NO = 0.723+2.22 107ρ +1.15 1013ρ2 9.44 105ρ 3.20 1013ρ ρ , dt × N × N − × NO − × N NO (7.203) dρ N = 0.723 2.33 107ρ 1.13 1013ρ2 +5.82 105ρ 1.00 1013ρ ρ . dt − × N − × N × NO − × N NO (7.204)

Solving numerically, we obtain a solution shown in Fig. 7.8. The numerics show a relaxation to final concentrations of

−7 mole lim ρNO = 7.336 10 , (7.205) t→∞ × cm3 −8 mole lim ρN = 3.708 10 . (7.206) t→∞ × cm3

Equations (7.203-7.204) are of the form

dρ NO = f (ρ , ρ ), (7.207) dt NO NO N dρ N = f (ρ , ρ ). (7.208) dt N NO N

At equilibrium, we must have

fNO(ρNO, ρN ) = 0, (7.209)

fN (ρNO, ρN ) = 0. (7.210)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 255

3 ρN, ρ NO (mole/cm ) - 1 ´ 10 5 - 5 ´ 10 6

- 1 ´ 10 6 - 5 ´ 10 7 NO

- 1 ´ 10 7 - 5 ´ 10 8 N

t s ------10 10 10 9 10 8 10 7 10 6 10 5

Figure 7.8: NO and N concentrations versus time for T = 6000 K, P = 2.4942 × 106 dyne/cm2 Zel’dovich mechanism.

We find three finite roots to this problem:

mole 1 : (ρ , ρ )=( 1.605 10−6, 3.060 10−8) , non-physical, (7.211) NO N − × − × cm3 mole 2 : (ρ , ρ )=( 5.173 10−8, 2.048 10−6) , non-physical, (7.212) NO N − × − × cm3 mole 3 : (ρ , ρ )=(7.336 10−7, 3.708 10−8) , physical. (7.213) NO N × × cm3

Obviously, because of negative concentrations, roots 1 and 2 are non-physical. Root 3 however is physical; moreover, it agrees with the equilibrium we found by direct numerical integration of the full non-linear equations. We can use local linear analysis in the neighborhood of each equilibria to rigorously ascertain the stability of each root. Taylor series expansion of Eqs. (7.207-7.208) in the neighborhood of an equilibrium point yields

d ∂f ∂f (ρ ρe ) = f + NO (ρ ρe )+ NO (ρ ρe )+ ..., dt NO − NO NO|e ∂ρ NO − NO ∂ρ N − N NO e N e =0 (7.214) | {z } d ∂f ∂f (ρ ρe ) = f + N (ρ ρe )+ N (ρ ρe )+ .... (7.215) dt N − N N |e ∂ρ NO − NO ∂ρ N − N NO e N e =0

|{z} CC BY-NC-ND. 28 January 2019, J. M. Powers. 256 CHAPTER 7. KINETICS IN SOME MORE DETAIL

Evaluation of Eqs. (7.214-7.215) near the physical root, root 3, yields the system d ρ 7.336 10−7 2.129 106 4.155 105 ρ 7.336 10−7 NO − × = − × − × NO − × . dt ρ 3.708 10−8 2.111 105 3.144 107 ρ 3.708 10−8  N − ×   × − ×   N − ×  f =J= ∂ ∂ρ |e | {z } (7.216) This is of the form d ∂f (ρ ρe)= (ρ ρe)= J (ρ ρe) . (7.217) dt − ∂ρ · − · − e

It is the eigenvalues of the Jacobian9 matrix J that give the time scales of evolution of the concentrations as well as determine the stability of the local equilibrium point. Recall that we can usually decompose square matrices via the diagonalization J = S Λ S−1. (7.218) · · Here, S is the matrix whose columns are composed of the right eigenvectors of J, and Λ is the diagonal matrix whose diagonal is populated by the eigenvalues of J. For some matrices (typically not those encountered after our removal of linear dependencies), diagonalization is not possible, and one must resort to the so-called near-diagonal Jordan form. This will not be relevant to our discussion, but could be easily handled if necessary. We also recall the eigenvector matrix and eigenvalue matrix are defined by the standard eigenvalue problem J S = S Λ. (7.219) · · We also recall that the components λ of Λ are found by solving the characteristic polynomial which arises from the equation det (J λI)=0, (7.220) − where I is the identity matrix. Defining z such that S z ρ ρe, (7.221) · ≡ − and using the decomposition Eq. (7.218), Eq. (7.217) can be rewritten to form d (S z) = S Λ S−1 (S z), (7.222) dt · · J· · · ρ−ρe dz | {z } S = S Λ z, | {z } (7.223) · dt · · dz S−1 S = S−1 S Λ z, (7.224) · · dt · · · dz = Λ z. (7.225) dt · 9after Carl Gustav Jacob Jacobi, 1804-1851, German mathematician.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 257

Eq. (7.225) reduces to the diagonal form dz = Λ z. (7.226) dt · This has solution for each component of z of

z1 = C1 exp(λ1t), (7.227)

z2 = C2 exp(λ2t), (7.228) . . (7.229)

Here, our matrix J, see Eq. (7.216), has two real, negative eigenvalues in the neighborhood of the physical root 3: 1 λ = 3.143 107 , (7.230) 1 − × s 1 λ = 2.132 106 . (7.231) 2 − × s Thus we can conclude that the physical equilibrium is linearly stable. The local time con- stants near equilibrium are given by the reciprocal of the magnitude of the eigenvalues. These are

τ = 1/ λ =3.181 10−8 s, (7.232) 1 | 1| × τ = 1/ λ =4.691 10−7 s. (7.233) 2 | 2| × Evolution on these two time scales is predicted in Fig. 7.8. This in fact a multiscale problem. One of the major difficulties in the numerical simulation of combustion problems comes in the effort to capture the effects at all relevant scales. The problem is made more difficult as the breadth of the scales expands. In this problem, the breadth of scales is not particularly challenging. Near equilibrium the ratio of the slowest to the fastest time scale, the stiffness ratio κ, is τ 4.691 10−7 s κ = 2 = × = 14.75. (7.234) τ 3.181 10−8 s 1 × Many combustion problems can have stiffness ratios over 106. This is more prevalent at lower temperatures. We can do a similar linearization near the initial state, find the local eigenvalues, and the the local time scales. At the initial state here, we find those local time scales are

τ = 2.403 10−8 s, (7.235) 1 × τ = 2.123 10−8 s. (7.236) 2 × So initially the stiffness, κ = (2.403 10−8 s)/(2.123 10−8 s)=1.13 is much less, but the × × time scale itself is small. It is seen from Fig. 7.8 that this initial time scale of 10−8 s well

CC BY-NC-ND. 28 January 2019, J. M. Powers. 258 CHAPTER 7. KINETICS IN SOME MORE DETAIL

predicts where significant evolution of species concentrations commences. For t< 10−8 s, the model predicts essentially no activity. This can be correlated with the mean time between molecular collisions–the theory on which estimates of the collision frequency factors aj are obtained. We briefly consider the non-physical roots, 1 and 2. A similar eigenvalue analysis of root 1 reveals that the eigenvalues of its local Jacobian matrix are 1 λ = 1.193 107 , (7.237) 1 − × s 1 λ = 5.434 106 . (7.238) 2 × s Thus root 1 is a saddle and is unstable. For root 2, we find 1 λ = 4.397 107 + i7.997 106 , (7.239) 1 × × s 1 λ = 4.397 107 i7.997 106 . (7.240) 2 × − × s The eigenvalues are complex with a positive real part. This indicates the root is an unstable spiral source. A detailed phase portrait is shown in Fig. 7.9. Here we see all three roots. Their local character of sink, saddle, or spiral source is clearly displayed. We see that trajectories are attracted to a curve labeled SIM for “Slow Invariant Manifold.” A part of the SIM is constructed by the trajectory which originates at root 1 and travels to root 3. The other part is constructed by connecting an equilibrium point at infinity into root 3. Details are omitted here.

7.1.2.2 Stiffness, time scales, and numerics One of the key challenges in computational chemistry is accurately predicting species concen- tration evolution with time. The problem is made difficult because of the common presence of physical phenomena which evolve on a widely disparate set of time scales. Systems which evolve on a wide range of scales are known as stiff, recognizing a motivating example in mass-spring-damper systems with stiff springs. Here we will examine the effect of tempera- ture and pressure on time scales and stiffness. We shall also look simplistically how different numerical approximation methods respond to stiffness.

7.1.2.2.1 Effect of temperature Let us see how the same Zel’dovich mechanism be- haves at lower temperature, T = 1500 K; all other parameters, including the initial species concentrations are the same as the previous high temperature example. The pressure how- ever, lowers, and here is P =6.23550 105 dyne/cm2, which is close to atmospheric pressure. × For this case, a plot of species concentrations versus time is given in Figure 7.10.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 259

-7 x 10 5

saddle 3 sink 0 SIM

) 1

3 SIM -5

(mole/cm -10 N ρ

-15

spiral -20 source 2

-4 -3 -2 -1 0 1 2 -6 x 10 ρ 3 NO (mole/cm )

Figure 7.9: NO and N phase portraits for T = 6000 K, P = 2.4942 106 dyne/cm2 Zel’dovich mechanism. ×

CC BY-NC-ND. 28 January 2019, J. M. Powers. 260 CHAPTER 7. KINETICS IN SOME MORE DETAIL

ρN, ρNO (mole/cm 3) - 10 6 NO

- 10 8

- 10 10

- 10 12 N

t s - - - - 10 11 10 9 10 7 10 5 0.001 0.1 10

Figure 7.10: ρNO and ρN versus time for Zel’dovich mechanism at T = 1500 K, P = 6.23550 105 dyne/cm2. ×

At T = 1500 K, we notice some dramatic differences relative to the earlier studied T = 6000 K. First, we see the reaction commences in around the same time, t 10−8 s. For t 10−6 s, there is a temporary cessation of significant reaction. We notice a∼long plateau ∼ in which species concentrations do not change over several decades of time. This is actually a pseudo-equilibrium. Significant reaction recommences for t 0.1 s. Only around t 1 s does the system approach final equilibrium. We can perform∼ an eigenvalue analysis∼ both at the initial state and at the equilibrium state to estimate the time scales of reaction. For this dynamical system which is two ordinary differential equations in two unknowns, we will always find two eigenvalues, and thus two time scales. Let us call them τ1 and τ2. Both these scales will evolve with t. At the initial state, we find

τ = 2.37 10−8 s, (7.241) 1 × τ = 4.25 10−7 s. (7.242) 2 ×

The onset of significant reaction is consistent with the prediction given by τ1 at the initial state. Moreover, initially, the reaction is not very stiff; the stiffness ratio is κ = 17.9. At equilibrium, we find

−9 mole lim ρNO = 4.6 10 , (7.243) t→∞ × cm3 −14 mole lim ρN = 4.2 10 , (7.244) t→∞ × cm3

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 261

and

τ = 7.86 10−7 s, (7.245) 1 × τ = 3.02 10−1 s. (7.246) 2 × The slowest time scale near equilibrium is an excellent indicator of how long the system takes to relax to its final state. Note also that near equilibrium, the stiffness ratio is large, κ = τ /τ 3.8 105. This is known as the stiffness ratio. When it is large, the scales in 2 1 ∼ × the problem are widely disparate and accurate numerical solution becomes challenging. In summary, we find the effect of lowering temperature while leaving initial concentrations constant lowers the pressure somewhat, slightly slowing down the collision time, and slightly • slowing the fastest time scales, and slows the slowest time scales many orders of magnitude, stiffening the system signifi- • cantly, since collisions may not induce reaction with their lower collision speed.

7.1.2.2.2 Effect of initial pressure Let us maintain the initial temperature at T = 1500 K, but drop the initial concentration of each species to mole ρ = ρ = ρ = ρ = ρ = 10−8 . (7.247) NO N N2 O2 O cm3 With this decrease in numberb b of moles,b theb pressureb now is dyne P =6.23550 103 . (7.248) × cm2 This pressure is two orders of magnitude lower than atmospheric. We solve for the species concentration profiles and show the results of numerical prediction in Figure 7.11 Relative to the high pressure P =6.2355 105 dyne/cm2, T = 1500 K case, we notice some similarities and dramatic differences. The× overall shape of the time-profiles of concentration variation is similar. But, we see the reaction commences at a much later time, t 10−6 s. For ∼ t 10−4 s, there is a temporary cessation of significant reaction. We notice a long plateau in∼ which species concentrations do not change over several decades of time. This is again actually a pseudo-equilibrium. Significant reaction recommences for t 10 s. Only around ∼ t 100 s does the system approach final equilibrium. We can perform an eigenvalue analysis both∼ at the initial state and at the equilibrium state to estimate the time scales of reaction. At the initial state, we find

τ = 2.37 10−6 s, (7.249) 1 × τ = 4.25 10−5 s. (7.250) 2 ×

The onset of significant reaction is consistent with the prediction given by τ1 at the initial state. Moreover, initially, the reaction is not very stiff; the stiffness ratio is κ = 17.9.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 262 CHAPTER 7. KINETICS IN SOME MORE DETAIL

ρN, ρNO (mole/cm 3) - 10 8

NO - 10 10

- 10 12

- 10 14 N - 10 16

- - - t s 10 8 10 6 10 4 0.01 1 100

Figure 7.11: ρNO and ρN versus time for Zel’dovich mechanism at T = 1500 K, P = 6.2355 103 dyne/cm2. ×

Interestingly, by decreasing the initial pressure by a factor of 102, we increased the initial time scales by a complementary factor of 102; moreover, we did not alter the stiffness. At equilibrium, we find

−11 mole lim ρNO = 4.6 10 , (7.251) t→∞ × cm3 −16 mole lim ρN = 4.2 10 , (7.252) t→∞ × cm3 (7.253) and τ = 7.86 10−5 s, (7.254) 1 × τ = 3.02 101 s. (7.255) 2 × By decreasing the initial pressure by a factor of 102, we decreased the equilibrium concentra- tions by a factor of 102 and increased the time scales by a factor of 102, leaving the stiffness ratio unchanged. In summary, we find the effect of lowering the initial concentrations significantly while leaving temperature constant lowers the pressure significantly, proportionally slowing down the collision time, as well • as the fastest and slowest time scales, and does not affect the stiffness of the system. •

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.1. ISOTHERMAL, ISOCHORIC KINETICS 263

7.1.2.2.3 Stiffness and numerics The issue of how to simulate stiff systems of ordinary differential equations, such as presented by our Zel’dovich mechanism, is challenging. Here a brief summary of some of the issues will be presented. The interested reader should consult the numerical literature for a full discussion. See for example the excellent text of Iserles.10 We have seen throughout this section that there are two time scales at work, and they are often disparate. The species evolution is generally characterized by an initial fast transient, followed by a long plateau, then a final relaxation to equilibrium. We noted from the phase plane of Fig. 7.9 that the final relaxation to equilibrium (shown along the green line labeled “SIM”) is an attracting manifold for a wide variety of initial conditions. The relaxation onto the SIM is fast, and the motion on the SIM to equilibrium is relatively slow. Use of common numerical techniques can often mask or obscure the actual dynamics. Numerical methods to solve systems of ordinary differential equations can be broadly cat- egorized as explicit or implicit. We give a brief synopsis of each class of method. We cast each as a method to solve a system of the form dρ = f(ρ). (7.256) dt Explicit: The simplest of these, the forward Euler method, discretizes Eq. (7.256) as • follows: ρ ρ n+1 − n = f(ρ ), (7.257) ∆t n so that

ρn+1 = ρn + ∆t f(ρn). (7.258) Explicit methods are summarized as – easy to program, since Eq. (7.258) can be solved explicitly to predict the new

value ρn+1 in terms of the old values at step n.

– need to have ∆t < τfastest in order to remain numerically stable, – able to capture all physics and all time scales at great computational expense for stiff problems, – requiring much computational effort for little payoff in the SIM region of the phase plane, and thus – inefficient for some portions of stiff calculations. Implicit: The simplest of these methods, the backward Euler method, discretizes • Eq. (7.256) as follows: ρ ρ n+1 − n = f(ρ ), (7.259) ∆t n+1 10A. Iserles, 2008, A First Course in the Numerical Analysis of Differential Equations, Cambridge Uni- versity Press, Cambridge, UK.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 264 CHAPTER 7. KINETICS IN SOME MORE DETAIL

so that

ρn+1 = ρn + ∆t f(ρn+1). (7.260) Implicit methods are summarized as – more difficult to program since a non-linear set of algebraic equations, Eq. (7.260), must be solved at every time step with no guarantee of solution, – requiring potentially significant computational time to advance each time step, – capable of using very large time steps and remaining numerically stable, – suspect to missing physics that occur on small time scales τ < ∆t, – in general better performers than explicit methods. A wide variety of software tools exist to solve systems of ordinary differential equations. Most of them use more sophisticated techniques than simple forward and backward Euler methods. One of the most powerful techniques is the use of error control. Here the user specifies how far in time to advance and the error that is able to be tolerated. The algorithm, which is complicated, selects then internal time steps, for either explicit or implicit methods, to achieve a solution within the error tolerance at the specified output time. A well known public domain algorithm with error control is provided by lsode.f, which can be found in the netlib repository.11 Let us exercise the Zel’dovich mechanism under the conditions simulated in Fig. 7.11, T = 1500 K, P = 6.2355 103 dyne/cm2. Recall in this case the fastest time scale near equilibrium is τ =7.86 10× −5 s 10−4 s at the initial state, and the slowest time scale is 1 × ∼ τ =3.02 101 s at the final state. Let us solve for these conditions using dlsode.f, which uses internal× time stepping for error control, in both an explicit and implicit mode. We specify a variety of values of ∆t and report typical values of number of internal time steps selected by dlsode.f, and the corresponding effective time step ∆teff used for the problem, for both explicit and implicit methods, as reported in Table 7.3. Obviously if output is requested using ∆t > 10−4 s, the early time dynamics near t 10−4 s will be missed. For physically stable systems, codes such as dlsode.f will still provide∼ a correct solution at the later times. For physically unstable systems, such as might occur in turbulent flames, it is not clear that one can use large time steps and expect to have fidelity to the underlying equations. The reason is the physical instabilities may evolve on the same time scale as the fine scales which are overlooked by large ∆t.

7.2 Adiabatic, isochoric kinetics

It is more practical to allow for temperature variation within a combustor. The best model for this is adiabatic kinetics. Here we will restrict our attention to isochoric problems. 11Hindmarsh, A. C., 1983,“ODEPACK, a Systematized Collection of ODE Solvers,” Scientific Com- puting, edited by R. S. Stepleman, et al., North-Holland, Amsterdam, pp. 55-64. Source code at http://www.netlib.org/alliant/ode/prog/lsode.f

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 265

Explicit Explicit Implicit Implicit ∆t (s) Ninternal ∆teff (s) Ninternal ∆teff (s) 102 106 10−4 100 102 101 105 10−4 100 101 100 104 10−4 100 100 10−1 103 10−4 100 10−1 10−2 102 10−4 100 10−2 10−3 101 10−4 100 10−3 10−4 100 10−4 100 10−4 10−5 100 10−5 100 10−5 10−6 100 10−6 100 10−6

Table 7.3: Results from computing Zel’dovich NO production using implicit and explicit methods with error control in dlsode.f. 7.2.1 Thermal explosion theory There is a simple description known as thermal explosion theory which provides a good explanation for how initially slow exothermic reaction induces a sudden temperature rise accompanied by a final relaxation to equilibrium. Let us consider a simple isomerization reaction in a closed volume A ⇌ B. (7.261) Let us take A and B to both be calorically perfect ideal gases with identical molecular masses MA = MB = M and identical specific heats, cvA = cvB = cv; cPA = cP B = cP . We can consider A and B to be isomers of an identical molecular species. So we have N = 2 species reacting in J = 1 reactions. The number of elements L here is irrelevant.

7.2.1.1 One-step reversible kinetics Let us insist our reaction process be isochoric and adiabatic, and commence with only A present. The reaction kinetics with β = 0 are

dρA E 1 = a exp − ρA ρB , (7.262) dt − RT − Kc     =k | {z }=r dρB E 1 = a exp| − ρ{zA ρB , } (7.263) dt RT − Kc     =k | {z }=r ρ (0) = ρ , (7.264) A |A {z } ρB(0) = 0. (7.265) b CC BY-NC-ND. 28 January 2019, J. M. Powers. 266 CHAPTER 7. KINETICS IN SOME MORE DETAIL

For our alternate compact linear algebra based form, we note that E 1 r = a exp − ρA ρB , (7.266) RT − Kc     and that d ρ 1 A = − (r). (7.267) dt ρ 1  B   Performing the decomposition yields d ρ 1 A = − (r). (7.268) dt ρ + ρ 0  A B   Expanded, this is 1 0 d ρ 1 A = − (r). (7.269) 1 1 dt ρ 0    B   Combining Eqs. (7.262-7.263) and integrating yields d (ρ + ρ ) = 0, (7.270) dt A B

ρA + ρB = ρA, (7.271)

ρB = ρA ρA. (7.272) b − Thus, Eq. (7.262) reduces to b

dρA E 1 = a exp − ρA ρA ρA . (7.273) dt − RT − Kc −      Scaling, Eq. (7.273) can be rewritten as b

d ρ E 1 ρ 1 ρ A = exp A 1 A . (7.274) d(at) − −RT T/To − Kc − ρA   o  ρA  ρA 

7.2.1.2 First lawb of thermodynamics b b Recall the first law of thermodynamics and neglecting potential and kinetic energy changes: dU = Q˙ W.˙ (7.275) dt − Here U is the total internal energy. Because we insist the problem is adiabatic Q˙ = 0. Because we insist the problem is isochoric, there is no work done, so W˙ = 0. Thus we have dU =0. (7.276) dt

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 267

Thus, we find

U = Uo. (7.277)

Recall the total internal energy for a mixture of two calorically perfect ideal gases is

U = nAuA + nBuB, (7.278) n n = V A u + B u , (7.279) V A V B = V (ρAuA + ρBuB) , (7.280) P P = V ρ h A + ρ h B , (7.281) A A − ρ B B − ρ   A   B  = V ρ h RT + ρ h RT , (7.282) A A − B B − o o = V ρ cP (T To)+ h RT + ρ cP (T To)+ h RT ,(7.283) A − To,A − B − To,B −    o  o  = V (ρ + ρ )(c (T T ) RT )+ ρ h + ρ h , (7.284) A B P − o − A To,A B To,B  o o  = V (ρ + ρ )((c R)T c T )+ ρ h + ρ h , (7.285) A B P − − P o A To,A B To,B  o  o = V (ρ + ρ )((c R)T (c R + R)T )+ ρ h + ρ h , (7.286) A B P − − P − o A To,A B To,B  o o  = V (ρ + ρ )(c T (c + R)T )+ ρ h + ρ h , (7.287) A B v − v o A To,A B To,B  o o  = V (ρ + ρ )c (T T )+ ρ (h RT )+ ρ (h RT ) , (7.288) A B v − o A To,A − o B To,B − o = V (ρ + ρ )c (T T )+ ρ uo + ρ uo .  (7.289) A B v − o A To,A B To,B

Now at the initial state, we have T = To, so 

o o Uo = V ρAuTo,A + ρBuTo,B . (7.290)   So, we can say our caloric equation of stateb is b

U U = V (ρ + ρ )c (T T )+(ρ ρ )uo +(ρ ρ )uo , (7.291) − o A B v − o A − A To,A B − B To,B   = V (ρ + ρ )c (T T )+(ρ ρ )uo +(ρ ρ )uo . (7.292) A B v − o A − bA To,A B − bB To,B   As an aside, on a molar basis,b web scale Eq. (7.292) tob get b u u = c (T T )+(y y )uo +(y y )uo . (7.293) − o v − o A − Ao To,A B − Bo To,B

And because we have assumed the molecular masses are the same, MA = MB, the mole fractions are the mass fractions, and we can write on a mass basis

u u = c (T T )+(c c )uo +(c c )uo . (7.294) − o v − o A − Ao To,A B − Bo To,B

CC BY-NC-ND. 28 January 2019, J. M. Powers. 268 CHAPTER 7. KINETICS IN SOME MORE DETAIL

Returning to Eq. (7.292), our energy conservation relation, Eq. (7.277), becomes

0 = V (ρ + ρ )c (T T )+(ρ ρ )uo +(ρ ρ )uo . (7.295) A B v − o A − A To,A B − B To,B   Now we solve for T b b b b 0 = (ρ + ρ )c (T T )+(ρ ρ )uo +(ρ ρ )uo , (7.296) A B v − o A − A To,A B − B To,B ρA ρA o ρB ρB o 0 = cbv(T bTo)+ − uTo,A + b − uTo,B, b (7.297) − ρA + ρB ρA + ρB o b o b ρA ρA uTo,A ρB ρB uTo,B T = To + − b b+ − b b. (7.298) ρA + ρB cv ρA + ρB cv b b Now we impose our assumption that ρ = 0, giving also ρ = ρ ρ , b b B b b B A − A o o ρA ρA uTo,A ρB uTo,B T = To +b − b, (7.299) ρA cv − ρA cv b o o ρA ρA uTo,A uTo,B = To + b− − b . (7.300) ρA cv o bo In summary, realizing that h h = uo uo we can write T as a function of To,A − To,Bb To,A − To,B ρA:

(ρA ρA) o o T = To + − (hTo,A hTo,B). (7.301) ρAcv − o o b We see then that if h > h , that as ρ decreases from its initial value of ρ that T To,A To,B b A A will increase. We can scale Eq. (7.301) to form

o o b T ρ h h =1+ 1 A To,A − To,B . (7.302) T − c T  o   ρA  v o !

We also note that our caloric state equation,b Eq. (7.293) can, for yAo = 1, yBo =0 as u u = c (T T )+(y 1)uo + y uo , (7.303) − o v − o A − To,A B To,B = c (T T )+((1 y ) 1)uo + y uo , (7.304) v − o − B − To,A B To,B = c (T T ) y (uo uo ). (7.305) v − o − B To,A − To,B Similarly, on a mass basis, we can say, u u = c (T T ) c (uo uo ). (7.306) − o v − o − B To,A − To,B For this problem, we also have ∆Go Kc = exp − , (7.307) RT   CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 269

with

o o o ∆G = gB gA, (7.308) o − o o o = hB T sB (hA T sA), (7.309) o − o − o− o = (hB hA) T (sB sA), (7.310) o − o− − = (h h ) T (so so ). (7.311) To,B − To,A − To,B − To,A So

o o o o hTo,A hTo,B T (sTo,A sTo,B) Kc = exp − − − , (7.312) RT ! o o c T h h T (so so ) = exp v o To,A − To,B − To,A − To,B , (7.313) RT cvTo !! o o o o 1 1 hTo,A hTo,B T (sTo,A sTo,B) = exp T − − . (7.314) k 1 cvTo − To cv − To !!

Here we have used the definition of the ratio of specific heats, k = cP /cv along with R = cP cv. So we can solve Eq. (7.273) by first using Eq. (7.314) to eliminate Kc and then Eq.− (7.301) to eliminate T .

7.2.1.3 Dimensionless form Let us try writing dimensionless variables so that our system can be written in a compact dimensionless form. First lets take dimensionless time τ to be τ = at. (7.315) Let us take dimensionless species concentration to be z with ρ z = A . (7.316) ρA Let us take dimensionless temperature to be θ with b T θ = . (7.317) To Let us take dimensionless heat release to be q with o o h h q = To,A − To,B . (7.318) cvTo Let us take dimensionless activation energy to be Θ with E Θ= . (7.319) RTo

CC BY-NC-ND. 28 January 2019, J. M. Powers. 270 CHAPTER 7. KINETICS IN SOME MORE DETAIL

And let us take the dimensionless entropy change to be σ with (so so ) σ = To,A − To,B . (7.320) cv So our equations become dz Θ 1 = exp z (1 z) , (7.321) dτ − − θ − K −    c  θ = 1+(1 z)q, (7.322) − 1 1 K = exp (q θσ) . (7.323) c k 1 θ −  −  It is more common to consider the products. Let us define for general problems ρ ρ λ = B = B . (7.324) ρA + ρB ρA + ρB

Thus λ is the mass fraction of product. For ourb problem,b ρB =0 so ρ ρ ρ λ = B = A − A . b (7.325) ρA ρA c Thus, b b λ =1 z. (7.326) − We can think of λ as a reaction progress variable as well. When λ = 0, we have τ = 0, and the reaction has not begun. Thus, we get dλ Θ 1 = exp (1 λ) λ , (7.327) dτ − θ − − K    c  θ = 1+ qλ, (7.328) 1 1 K = exp (q θσ) . (7.329) c k 1 θ −  −  7.2.1.4 Example calculation Let us choose some values for the dimensionless parameters: 7 Θ=20, σ =0, q = 10, k = . (7.330) 5 With these choices, our kinetics equations reduce to dλ 20 25 = exp − (1 λ) λ exp − , λ(0) = 0. (7.331) dτ 1+10λ − − 1+10λ      CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 271

Λ 1.0

0.8

0.6

0.4

0.2

Τ 0 1. ´ 106 2. ´ 106 3. ´ 106 4. ´ 106 5. ´ 106

Figure 7.12: Dimensionless plot of reaction product concentration λ versus time τ for adia- batic isochoric combustion with simple reversible kinetics.

The right side of Eq. (7.331) is at equilibrium for values of λ which drive it to zero. Numerical root finding methods show this to occur at λ 0.920539. Near this root, Taylor series expansion shows the dynamics are approximated by∼ d (λ 0.920539) = 0.17993(λ 0.920539) + ... (7.332) dτ − − − Thus the local behavior near equilibrium is given by

λ =0.920539 + C exp ( 0.17993 τ) . (7.333) − Here C is some arbitrary constant. Clearly the equilibrium is stable, with a time constant of 1/0.17993 = 5.55773. Numerical solution shows the full behavior of the dimensionless species concentration λ(τ); see Figure 7.12. Clearly the product concentration λ is small for some long period of time. At a critical time near τ = 2.7 106, there is a so-called thermal explosion with a rapid increase in λ. Note that the estimate× of the time constant near equilibrium is orders of magnitude less than the explosion time, 5.55773 << 2.7 106. Thus, linear analysis × here is a poor tool to estimate an important physical quantity, the ignition time. Once the ignition period is over, there is a rapid equilibration to the final state. The dimensionless temperature plot is shown in Figure 7.13. The temperature plot is similar in behavior to the species concentration plot. At early time, the temperature is cool. At a critical time, the thermal explosion time, the temperature rapidly rises. This rapid rise, coupled with the exponential sensitivity of reaction rate to temperature, accelerates the formation of product. This process continues until the reverse reaction is activated to the extent it prevents further creation of product.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 272 CHAPTER 7. KINETICS IN SOME MORE DETAIL

Θ

10

8

6

4

2

Τ 1. ´ 106 2. ´ 106 3. ´ 106 4. ´ 106 5. ´ 106

Figure 7.13: Dimensionless plot of temperature θ versus time τ for adiabatic, isochoric combustion with simple reversible kinetics.

7.2.1.5 High activation energy asymptotics

Let us see if we can get an analytic prediction of the thermal explosion time, τ 2.7 106. ∼ × Such a prediction would be valuable to see how long a slowing reacting material might take to ignite. Our analysis is similar to that given by Buckmaster and Ludford in their Chapter 1.12 For convenience let us restrict ourselves to σ = 0. In this limit, Eqs. (7.327-7.329) reduce to

dλ Θ q = exp (1 λ) λ exp − , (7.334) dτ −1+ qλ − − (k 1)(1 + qλ)     −  with λ(0) = 0. The key trouble in getting an analytic solution to Eq. (7.334) is the presence of λ in the denominator of an exponential term. We need to find a way to move it to the numerator. Asymptotic methods provide one such way. Now we recall for early time λ << 1. Let us assume λ takes the form

2 3 λ = ǫλ1 + ǫ λ2 + ǫ λ3 + ... (7.335)

Here we will assume 0 <ǫ<< 1 and that λ1(τ) (1), λ2(τ) (1), ... , and will define ǫ in terms of physical parameters shortly. Now with∼ O this assumption,∼ O we have

1 1 = 2 3 . (7.336) 1+ qλ 1+ ǫqλ1 + ǫ qλ2 + ǫ qλ3 + ...

12J. D. Buckmaster and G. S. S. Ludford, 1983, Lectures on Mathematical Combustion, SIAM, Philadelphia.

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 273

Long division of the term on the right side yields the approximation 1 = = 1 ǫqλ + ǫ2(q2λ2 qλ )+ ..., (7.337) 1+ qλ − 1 1 − 2 = 1 ǫqλ + (ǫ2). (7.338) − 1 O So Θ exp exp Θ(1 ǫqλ + (ǫ2)) , (7.339) −1+ qλ ∼ − − 1 O   e−Θ exp ǫqΘλ + (ǫ2) , (7.340) ∼ 1 O We have moved λ from the denominator to the numerator of the most important exponential term. Now, let us take the limit of high activation energy by defining ǫ to be 1 ǫ . (7.341) ≡ Θ Let us let the assume the remaining parameters, q and k are both (1) constants. When Θ is large, ǫ will be small. With this definition, Eq. (7.340) becomes O Θ exp e−1/ǫ exp qλ + (ǫ2) . (7.342) −1+ qλ ∼ 1 O    With these assumptions and approximations, Eq. (7.334) can be written as d (ǫλ + ... ) = e−1/ǫ exp qλ + (ǫ2) dτ 1 1 O  q (1 ǫλ ... ) (ǫλ + ... ) exp − . × − 1 − − 1 (k 1)(1 + qǫλ + ... )   − 1  (7.343) Now let us rescale time via 1 τ = e−1/ǫτ. (7.344) ∗ ǫ With this transformation, the chain rule shows how derivatives transform:

d dτ∗ d 1 d = = 1/ǫ . (7.345) dτ dτ dτ∗ ǫe dτ∗ With this transformation, Eq. (7.343) becomes

1 d 1 2 1/ǫ (ǫλ1 + ... ) = 1/ǫ exp qλ1 + (ǫ ) ǫe dτ∗ e O  q (1 ǫλ ... ) (ǫλ + ... ) exp − . × − 1 − − 1 (k 1)(1 + qǫλ + ... )   − 1  (7.346)

CC BY-NC-ND. 28 January 2019, J. M. Powers. 274 CHAPTER 7. KINETICS IN SOME MORE DETAIL

This simplifies to

d 2 (λ1 + ... ) = exp qλ1 + (ǫ ) dτ∗ O  q (1 ǫλ ... ) (ǫλ + ... ) exp − . × − 1 − − 1 (k 1)(1 + qǫλ + ... )   − 1  (7.347)

Retaining only (1) terms in Eq. (7.347), we get O

dλ1 = exp (qλ1) . (7.348) dτ∗

This is supplemented by the initial condition λ1(0) = 0. Separating variables and solving, we get

exp( qλ )dλ = dτ , (7.349) − 1 1 ∗ 1 exp( qλ ) = τ + C. (7.350) −q − 1 ∗

Applying the initial condition gives 1 exp( q(0)) = C, (7.351) −q − 1 = C. (7.352) −q So 1 1 exp( qλ ) = τ , (7.353) −q − 1 ∗ − q exp( qλ ) = qτ +1, (7.354) − 1 − ∗ 1 exp( qλ ) = q τ , (7.355) − 1 − ∗ − q   1 qλ = ln q τ , (7.356) − 1 − ∗ − q    1 1 λ = ln q τ . (7.357) 1 −q − ∗ − q   

For q = 10, a plot of λ1(τ∗) is shown in Fig. 7.14. We note at a finite τ∗ that λ1 begins to exhibit unbounded growth. In fact, it is obvious from Eq. (7.348) that as

1 τ , ∗ → q

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 275

λ1 1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.00 0.02 0.04 0.06 0.08 0.10 0.12 τ*

Figure 7.14: λ1 versus τ∗ for ignition problem.

that λ . 1 →∞ That is there exists a finite time for which λ1 violates the assumptions of our asymptotic theory which assumes λ = (1). We associate this time with the ignition time, τ : 1 O ∗i 1 τ = . (7.358) ∗i q Let us return this to more primitive variables: 1 1 1 exp − τ = , (7.359) ǫ ǫ i q   1 ǫ exp ǫ τi = , (7.360) q  exp Θ τ = . (7.361) i Θq For our system with Θ = 20 and q = 10, we estimate the dimensionless ignition time as exp 20 τ = =2.42583 106. (7.362) i (20)(10) × This is a surprisingly good estimate, given the complexity of the problem. Recall the nu- merical solution showed ignition for τ 2.7 106. ∼ ×

CC BY-NC-ND. 28 January 2019, J. M. Powers. 276 CHAPTER 7. KINETICS IN SOME MORE DETAIL

In terms of dimensional time, ignition time prediction becomes exp Θ t = , (7.363) i aΘq

1 RTo cvTo E = o o exp . (7.364) a E hT ,A hT ,B ! RTo   o − o   Note the ignition is suppressed if the ignition time is lengthened, which happens when

he activation energy E is increased, since the exponential sensitivity is stronger than • the algebraic sensitivity,

o o the energy of combustion (hTo,A hTo,B) is decreased because it takes longer to react • to drive the temperature to a critical− value to induce ignition, and

the collision frequency factor a is decreased, which suppresses reaction. • 7.2.2 Detailed H O N kinetics 2 − 2 − 2 Here is an example which uses multiple reactions for an adiabatic isothermal system is given. Consider the full time-dependency of a problem similar to the thermal explosion problem just considered. A closed, fixed, adiabatic volume, V =0.3061251 cm3, contains at t =0 s a stoichiometric −5 −5 −5 hydrogen-air mixture of 2 10 mole of H2, 1 10 mole of O2, and 3.76 10 mole of N at P = 2.83230 106 ×P a and T = 1542.7 ×K.13 Thus the initial molar× concentrations 2 o × o are ρ =6.533 10−5 mole/cm3, H2 × ρ =3.267 10−5 mole/cm3, O2 × ρ =1.228 10−4 mole/cm3. H2 × The initial mass fractions are calculated via ci = Miρi/ρ. They are

cH2 =0.0285,

cO2 =0.226,

cN2 =0.745. To avoid issues associated with numerical roundoff errors at very early time for species with very small compositions, the minor species were initialized at a small non-zero value near precision; each was assigned a value of 10−15 mole. The minor species all have ρ =1.803 10−16 mole/cm3. They have correspondingly small initial mass fractions. i × 13 This temperature and pressure correspond to that of the same ambient mixture of H2, O2 and N2 which was shocked from 1.01325 105 Pa, 298 K, to a value associated with a freely propagating detonation. ×

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 277

0 H 10 O H 2 O 2 OH H 2 O −5 H 2 O 2 10 HO 2 N 2

ci

−10 10

−15 10

−10 −5 0 10 10 10 t (s)

Figure 7.15: Plot of cH2 (t), cH (t), cO(t), cO2 (t), cOH (t), cH2O(t), cHO2 (t), cH2O2 (t), cN2 (t), for adiabatic, isochoric combustion of a mixture of 2H2 +O2 +3.76N2 initially at To = 1542.7 K, P =2.8323 106 P a. o ×

We seek the reaction dynamics as the system proceeds from its initial state to its final state. We use the reversible detailed kinetics mechanism of Table 7.2. This problem requires a detailed numerical solution. Such a solution was performed by solving the appropriate equations for a mixture of nine interacting species: H2, H, O, O2, OH, H2O, HO2, H2O2, and N2. The dynamics of the reaction process are reflected in Figs. 7.15-7.17. At early time, t < 10−7 s, the pressure, temperature, and major reactant species con- centrations (H2,O2, N2) are nearly constant. However, the minor species, e.g. OH, HO2, and the major product, H2O, are undergoing very rapid growth, albeit with math fractions whose value remains small. In this period, the material is in what is known as the induction period. After a certain critical mass of minor species has accumulated, exothermic recombination of these minor species to form the major product H2O induces the temperature to rise, which accelerates further the reaction rates. This is manifested in a thermal explosion. A common definition of the end of the induction period is the induction time, t = tind, the time when dT/dt goes through a maximum. Here one finds

t =6.6 10−7 s. (7.365) ind ×

A close-up view of the species concentration profiles is given in Fig. 7.18. At the end of the induction zone, there is a final relaxation to equilibrium. The equilib-

CC BY-NC-ND. 28 January 2019, J. M. Powers. 278 CHAPTER 7. KINETICS IN SOME MORE DETAIL

4000

3500

3000

2500

2000 T (K) 1500

1000

500

0 −10 −5 10 10 t (s)

Figure 7.16: Plot of T (t), for adiabatic, isochoric combustion of a mixture of 2H2+O2+3.76N2 initially at T = 1542.7 K, P =2.8323 106 P a. o o ×

7 10 P (Pa)

6 10 −10 −5 10 10 t (s)

Figure 7.17: Plot of P (t), for adiabatic, isochoric combustion of a mixture of 2H2 + O2 + 3.76N initially at T = 1542.7 K, P =2.8323 106 P a. 2 o o ×

CC BY-NC-ND. 28 January 2019, J. M. Powers. 7.2. ADIABATIC, ISOCHORIC KINETICS 279

0 10

H O H −5 2 c 10 i O 2 OH

H 2 O

H 2 O 2 −10 HO 10 2 N 2

2 4 6 8 10 −7 t (s) x 10

Figure 7.18: Plot near thermal explosion time of cH2 (t), cH (t), cO(t), cO2 (t), cOH (t), cH2O(t), cHO2 (t), cH2O2 (t), cN2 (t), for adiabatic, isochoric combustion of a mixture of 2H2+O2+3.76N2 initially at at T = 1542.7 K, P =2.8323 106 P a. o o × rium mass fractions of each species are

c = 1.85 10−2, (7.366) O2 × c = 5.41 10−4, (7.367) H × c = 2.45 10−2, (7.368) OH × c = 3.88 10−3, (7.369) O × c = 3.75 10−3, (7.370) H2 × c = 2.04 10−1, (7.371) H2O × c = 6.84 10−5, (7.372) HO2 × c = 1.04 10−5, (7.373) H2O2 × c = 7.45 10−1. (7.374) N2 ×

We note that because our model takes N2 to be inert that its value remains unchanged. Other than N2, the final products are dominated by H2O. The equilibrium temperature is 3382.3 K and 5.53 106 P a. ×

CC BY-NC-ND. 28 January 2019, J. M. Powers. 280 CHAPTER 7. KINETICS IN SOME MORE DETAIL

CC BY-NC-ND. 28 January 2019, J. M. Powers. Bibliography

M. M. Abbott and H. C. van Ness, 1989, Thermodynamics with Chemical Applications, Second Edition, Schaum’s Outline Series in Engineering, McGraw-Hill, New York.

First published in 1972, this is written in the style of all the Schaum’s series. It has extensive solved problems and a crisp rigorous style that is readable by undergraduate engineers. It has a chemical engineering emphasis, but is also useful for all engineers.

P. Atkins, 2010, The : a Very Short Introduction, Oxford Univer- sity Press, Oxford.

This is a short book by an eminent chemist summarizing the foundations of thermodynamics for an interested general reader.

A. Bejan, 2006, Advanced Engineering Thermodynamics, Third Edition, John Wiley, Hobo- ken, New Jersey.

This is an advanced undergraduate text first published in 1988. It gives a modern treatment of the science of classical thermodynamics. It does not confine its attention to traditional engineering problems, and considers applications across biology and earth sciences as well; some readers will find parts of the discussion to be provocative, as received wisdom is occasionally challenged. The thermodynamics of irreversible processes are discussed in detail.

R. S. Berry, S. A. Rice, and J. Ross, 2000, Physical Chemistry, Second Edition, Oxford University Press, Oxford.

This is a rigorous general text in physical chemistry at a senior or first year graduate level. First appearing in 1980, it has a full treatment of classical and statistical thermodynamics as well as .

L. Boltzmann, 1995, Lectures on Gas Theory, Dover, New York.

This is a detailed monograph by the founding father of statistical thermodynamics. This edition is a translation of the original German version, Gastheorie, from 1896-1898.

C. Borgnakke and R. E. Sonntag, 2009, Fundamentals of Thermodynamics, Seventh Edition, John Wiley, New York.

281 This classic and popular undergraduate mechanical engineering text, which in earlier editions was authored by J. G. van Wylen and Sonntag, has stood the test of time and has a full treatment of most classical problems. Its first edition appeared in 1965.

M. Born, 1949, Natural Philosophy of Cause and Chance, Clarendon Press, Oxford.

This monograph is a summary of the Waynflete Lectures delivered at Oxford University in 1948 by the author, the winner of the 1954 Nobel Prize in physics. The lectures consider various topics, and include an important chapter on thermodynamics with the author’s earlier defense of the approach of Carath`eodory playing a prominent role.

R. Boyle, 2003, The Sceptical Chymist, Dover, New York.

This is a reprint of the original 1661 work of the famous early figure of the scientific revolution.

P. W. Bridgman, 1943, The Nature of Thermodynamics, Harvard, Cambridge.

This short monograph by the winner of the 1946 Nobel Prize in physics gives a prosaic introduction to thermodynamics which is directed at a scientifically literate audience who are not interested in detailed mathematical exposition.

H. B. Callen, 1985, Thermodynamics and an Introduction to Thermostatistics, Second Edi- tion, John Wiley, New York.

This advanced undergraduate text, an update of the 1960 original, has an emphasis on classical physics applied to thermodynamics, with a few chapters devoted to quantum and statistical foundations.

S. Carnot, 2005, Reflections on the Motive Power of Fire, Dover, New York.

This is a translation of the author’s foundational 1824 work on the heat engines, R´eflexions sur la Puissance Motrice du Feu et sur les propres `ad´evelopper cette Puissance (“Reflections on the Motive Power of Fire and on Machines Fitted to Develop that Power”). Also included is a paper of Clausius.

Y. A. C¸engel and M. A. Boles, 2010, Thermodynamics: An Engineering Approach, Seventh Edition, McGraw-Hill, Boston.

This popular undergraduate mechanical engineering text which first appeared in 1989 has most of the features expected in a modern book intended for a large and varied audience.

S. Chandrasekhar, 2010, An Introduction to the Study of , Dover, New York.

This monograph, first published in 1939, on astrophysics is by the winner of the 1983 Nobel Prize in physics. It has large sections devoted to rigorous axiomatic classical thermodynamics in the style of Carath`eodory, highly accessible to engineering students, which show how thermodynamics plays a critical role in understanding the physics of the heavens.

282 R. J. E. Clausius, 2008, The Mechanical Theory of Heat, Kessinger, Whitefish, Montana. This is a reprint of the 1879 translation of the 1850 German publication of the great German scientist who in many ways founded classical thermodynamics.

S. R. de Groot and P. Mazur, 1984, Non-Equilibrium Thermodynamics, Dover, New York. This is an influential monograph, first published in 1962, which summarizes much of the work of the famous Belgian school of thermodynamics. It is written at a graduate level and has a strong link to fluid mechanics and chemical reactions.

E. Fermi, 1936, Thermodynamics, Dover, New York. This short 160 page classic clearly and efficiently summarizes the fundamentals of thermodynamics. It is based on a series of lectures given by this winner of the 1938 Nobel Prize in physics. The book is highly recommended, and the reader can benefit from multiple readings.

R. P. Feynman, R. B. Leighton, and M. Sands, 1963, The Feynman Lectures on Physics, Volume 1, Addison-Wesley, Reading, Massachusetts. This famous series documents the introductory undergraduate physics lectures given at the California Institute of Technology by the lead author, the 1965 Nobel laureate in physics. Famous for their clarity, depth, and notable forays into challenging material, they treat a wide range of topics from classical to modern physics. Volume 1 contains chapters relevant to classical and modern thermodynamics.

J. B. J. Fourier, 2009, The Analytical Theory of Heat, Cambridge, Cambridge. This reprint of the 1878 English translation of the 1822 French original, Th´eorie Analytique de la Chaleur is a tour de force of science, engineering, and mathematics. It predates Carnot and the development of the first and second laws of thermodynamics, but nevertheless successfully develops a theory of non-equilibrium thermodynamics fully consistent with the first and second laws. In so doing, the author makes key advances in the formulation of partial differential equations and their solution by what are now known as Fourier series.

J. W. Gibbs, 1957, The Collected Works of J. Willard Gibbs, Yale U. Press, New Haven. This compendium gives a complete reproduction of the published work of the monumental American engineer and scientist of the nineteenth century, including his seminal work on classical and statistical thermodynamics.

W. T. Grandy, 2008, Entropy and the Time Evolution of Macroscopic Systems, Oxford, Oxford. This monograph gives an enlightening description, fully informed by both historical and modern interpretations, of entropy and its evolution in the context of both continuum and statistical theories.

E. A. Guggenheim, 1933, Modern Thermodynamics by the Methods of Willard Gibbs, Methuen, London. This graduate-level monograph was important in bringing the work of Gibbs to a wider audience.

283 E. P. Gyftopoulos and G. P. Beretta, 1991, Thermodynamics: Foundations and Applica- tions, Macmillan, New York. This beginning graduate text has a rigorous development of classical thermodynamics.

C. S. Helrich, 2009, Modern Thermodynamics with , Springer, Berlin. This is an advanced undergraduate text aimed mainly at the physics community. The author includes a full treatment of classical thermodynamics and moves easily into statistical mechanics. The author’s undergraduate training in engineering is evident in some of the style of the text which should be readable by most undergraduate engineers after a first class in thermodynamics.

J. O. Hirschfelder, C. F. Curtis, and R. B. Bird, 1954, Molecular Theory of Gases and , Wiley, New York. This comprehensive tome is a valuable addition to any library of thermal science. Its wide ranging text covers equations of state, molecular collision theory, reactive hydrodynamics, reaction kinetics, and many other topics all from the point of view of careful physical chemistry. Much of the work remains original.

J. R. Howell and R. O. Buckius, 1992, Fundamentals of Engineering Thermodynamics, Second Edition, McGraw-Hill, New York. This is a good undergraduate text for mechanical engineers; it first appeared in 1987.

W. F. Hughes and J. A. Brighton, 1999, Dynamics, Third Edition, Schaum’s Outline Series, McGraw-Hill, New York. Written in the standard student-friendly style of the Schaum’s series, this discussion of fluid mechanics includes two chapters on compressible flow which bring together fluid mechanics and thermodynamics. It first appeared in 1967.

F. P. Incropera, D. P. DeWitt, T. L. Bergman, and A. S. Lavine, 2006, Fundamentals of Heat and Mass Transfer, Sixth Edition, John Wiley, New York. This has evolved since its introduction in 1981 into the standard undergraduate text in heat transfer. The interested thermodynamics student will find the subject of heat transfer builds in many ways on a classical thermodynamics foundation. Especially relevant to thermodynamics are chapters on boiling and , heat exchangers, as well as extensive tables of thermal properties of real materials.

J. H. Keenan, F. G. Keyes, P. G. Hill, and J. G. Moore, 1985, Steam Tables: Thermodynamic Including Vapor, Liquid, and Solid Phases, John Wiley, New York. This is a version of the original 1936 data set which is the standard for water’s thermodynamic properties.

J. Kestin, 1966, A Course in Thermodynamics, Blaisdell, Waltham, Massachusetts.

284 This is a foundational textbook that can be read on many levels. All first principles are reported in a readable fashion. In addition the author makes a great effort to expose the underlying mathematical foundations of thermodynamics.

R. J. Kee, M. E. Coltrin, and P. Glarborg, 2003, Chemically Reacting Flow: Theory and Practice, John Wiley, New York. This comprehensive text gives an introductory graduate level discussion of fluid mechanics, thermo- chemistry, and finite rate chemical kinetics. The focus is on low reacting flows, and there is significant discussion of how to achieve computational solutions.

C. Kittel and H. Kroemer, 1980, , Second Edition, Freeman, San Francisco.

This is a classic undergraduate text, first introduced in 1970; it is mainly aimed at physics students. It has a good introduction to statistical thermodynamics and a short effective chapter on classical thermodynamics. Engineers seeking to broaden their skill set for new technologies relying on micro- scale thermal phenomena can use this text as a starting point.

D. Kondepudi and I. Prigogine, 1998, Modern Thermodynamics: From Heat Engines to Dissipative Structures, John Wiley, New York.

This is a detailed modern exposition which exploits the authors’ unique vision of thermodynamics with both a science and engineering flavor. The authors, the second of whom is one of the few engineers who was awarded the Nobel Prize (chemistry 1977, for the work summarized in this text), often challenge the standard approach to teaching thermodynamics, and make the case that the approach they advocate, with an emphasis on non-equilibrium thermodynamics, is better suited to describe natural phenomena and practical devices than the present approach, which is generally restricted to equilibrium states.

K. K. Kuo, 2005, Principles of Combustion, Second Edition, John Wiley, New York.

This is a readable graduate level engineering text for combustion fundamentals. First published in 1986, it includes a full treatment of reacting thermodynamics as well as discussion of links to fluid mechanics.

K. J. Laidler, 1987, Chemical Kinetics, Third Edition, Prentice-Hall, Upper Saddle River, NJ.

This is a standard advanced undergraduate chemistry text on the dynamics of chemical reactions. It first appeared in 1965.

L. D. Landau and E. M. Lifshitz, 2000, Statistical Physics, Part 1, Volume 5 of the Course of Theoretical Physics, Third Edition, Butterworth-Heinemann, Oxford.

This book, part of the monumental series of graduate level Russian physics texts, first published in English in 1951 from Statisticheskaya fizika, gives a fine introduction to classical thermodynamics as a prelude to its main topic, statistical thermodynamics in the spirit of Gibbs.

285 B. H. Lavenda, 1978, Thermodynamics of Irreversible Processes, John Wiley, New York.

This is a lively and opinionated monograph describing and commenting on irreversible thermodynam- ics. The author is especially critical of the Prigogine school of thought on entropy production rate minimization.

A. Lavoisier, 1984, Elements of Chemistry, Dover, New York.

This is the classic treatise by the man known as the father of modern chemistry, translated from the 1789 Trait´e Elementaire´ de Chimie, which gives the first explicit statement of mass conservation in chemical reactions.

G. N. Lewis and M. Randall, 1961, Thermodynamics and the Free Energy of Chemical Substances, Second Edition, McGraw-Hill, New York.

This book, first published in 1923, was for many years a standard reference text of physical chemistry.

H. W. Liepmann and A. Roshko, 2002, Elements of Gasdynamics, Dover, New York.

This is an influential text in compressible aerodynamics that is appropriate for seniors or beginning graduate students. First published in 1957, it has a strong treatment of the physics and thermody- namics of compressible flow along with elegant and efficient text. Its treatment of both experiment and the underlying theory is outstanding, and in many ways is representative of the approach to engineering sciences fostered at the California Institute of Technology, the authors’ home institution.

J. C. Maxwell, 2001, Theory of Heat, Dover, New York.

This is a short readable book by the nineteenth century master, first published in 1871. Here the mathematics is minimized in favor of more words of explanation.

M. J. Moran and H. N. Shapiro, 2007, Fundamentals of Engineering Thermodynamics, Sixth Edition, John Wiley, New York.

This is a standard undergraduate engineering thermodynamics text, and one of the more popular. First published in 1988, it has much to recommend it including good example problems, attention to detail, good graphics, and a level of rigor appropriate for good undergraduate students.

P. M. Morse, 1969, Thermal Physics, Second Edition, Benjamin, New York.

This is a good undergraduate book on thermodynamics from a physics perspective. It covers classical theory well in its first sections, then goes on to treat kinetic theory and statistical mechanics. It first appeared in 1964.

I. M¨uller and T. Ruggeri, 1998, Rational Extended Thermodynamics, Springer-Verlag, New York.

This modern, erudite monograph gives a rigorous treatment of some of the key issues at the frontier of modern continuum thermodynamics.

286 I. M¨uller and W. Weiss, 2005, Entropy and Energy, Springer-Verlag, New York.

This is a unique treatise on fundamental concepts in thermodynamics. The author provide mathe- matical rigor, historical perspective, and examples from a diverse set of scientific fields.

I. M¨uller, 2007, A History of Thermodynamics: the Doctrine of Energy and Entropy, Springer-Verlag, Berlin.

The author gives a readable text at an advanced undergraduate level which highlights some of the many controversies of thermodynamics, both ancient and modern.

I. M¨uller and W. H. M¨uller, 2009, Fundamentals of Thermodynamics and Applications: with Historical Annotations and Many Citations from Avogadro to Zermelo, Springer, Berlin.

The author presents an eclectic view of classical thermodynamics with much discussion of its history. The text is aimed at a curious undergraduate who is unsatisfied with industrial-strength yet narrow and intellectually vapid textbooks.

W. Nernst, 1969, The New Heat Theorem: Its Foundation in Theory and Experiment, Dover, New York.

This monograph gives the author’s exposition of the development of the third law of thermodynamics. It first appeared in English translation in 1917 and was originally published in German.

W. Pauli, 2000, Thermodynamics and the , Dover, New York.

This is a monograph on statistical thermodynamics by the winner of the 1945 Nobel Prize in physics. It is actually derived from his lecture course notes given at ETH Zurich, as compiled by a student in his class, E. Jucker, published in 1952.

M. Planck, 1990, Treatise on Thermodynamics, Dover, New York.

This brief book, which originally appeared in German in 1897, gives many unique insights from the great scientist who was the winner of the 1918 Nobel Prize in physics. It is rigorous, but readable by an interested undergraduate student.

H. Poincar´e, 1892, Thermodynamique: Le¸cons Profess`ees Pendant le Premier Semestre 1888-89, Georges Carr´e, Paris.

This text of classical undergraduate thermodynamics has been prepared by one of the premier math- ematicians of the nineteenth century.

J. M. Powers, 2016, Combustion Thermodynamics and Dynamics, Cambridge, New York.

This is a graduate level text on combustion and its interplay with both thermodynamics and dynamics. It uses a large portion of the present course notes as a foundation for its discussion of thermodynamics of reacting mixtures.

287 I. Prigogine, 1967, Introduction to Thermodynamics of Irreversible Processes, Third Edition, Interscience, New York. This is a famous book that summarizes the essence of the work of the Belgian school for which the author was awarded the 1977 Nobel Prize in chemistry. This book first appeared in 1955.

W. J. M. Rankine, 1908, A Manual of the and Other Prime Movers, Seven- teenth Edition, Griffin, London. This in an accessible undergraduate text for mechanical engineers of the nineteenth century. It first appeared in 1859. It contains much practical information on a variety of devices, including steam engines.

L. E. Reichl, 1998, A Modern Course in Statistical Physics, John Wiley, New York. This full service graduate text has a good summary of key concepts of classical thermodynamics and a strong development of modern statistical thermodynamics.

O. Reynolds, 1903, Papers on Mechanical and Physical Subjects, Volume III, The Sub- Mechanics of the Universe, Cambridge, Cambridge. This volume compiles various otherwise unpublished notes of Reynolds and includes his detailed deriva- tions of general equations of conservation of mass, , and energy employing his transport theorem.

W. C. Reynolds, 1968, Thermodynamics, Second Edition, McGraw-Hill, New York. This is an unusually good undergraduate text written for mechanical engineers. The author has wonderful qualitative problems in addition to the usual topics in such texts. A good introduction to statistical mechanics is included as well. This particular edition is highly recommended; the first edition appeared in 1965.

S. I. Sandler, 1998, Chemical and Engineering Thermodynamics, Third Edition, John Wiley, New York. This is an advanced undergraduate text in thermodynamics from a chemical engineering perspective with a good mathematical treatment. It first appeared in 1977.

E. Schr¨odinger, 1989, Statistical Thermodynamics, Dover, New York. This is a short monograph written by the one of the pioneers of quantum physics, the co-winner of the 1933 Nobel Prize in Physics. It is based on a set of lectures delivered to the Dublin Institute for Advanced Studies in 1944, and was first published in 1946.

A. H. Shapiro, 1953, The Dynamics and Thermodynamics of Compressible Fluid Flow, Vols. I and II, John Wiley, New York. This classic two volume set has a comprehensive treatment of the subject of its title. It has numerous worked example problems, and is written from a careful engineer’s perspective.

288 J. M. Smith, H. C. van Ness, and M. Abbott, 2004, Introduction to Chemical Engineering Thermodynamics, Seventh Edition, McGraw-Hill, New York. This is probably the most common undergraduate text in thermodynamics in chemical engineering. It first appeared in 1959. It is rigorous and has went through many revisions.

A. Sommerfeld, 1956, Thermodynamics and Statistical Mechanics, Lectures on Theoretical Physics, Vol. V, Academic Press, New York. This is a compilation of the author’s lecture notes on this subject. The book reflects the author’s stature of a leader of theoretical physics of the twentieth century who trained a generation of students (e.g. Nobel laureates Heisenberg, Pauli, Debye and Bethe). The book gives a fine description of classical thermodynamics with a seamless transition into quantum and statistical mechanics.

J. W. Tester and M. Modell, 1997, Thermodynamics and Its Applications, Third Edition, Prentice Hall, Upper Saddle River, New Jersey. First appearing in 1974, this entry level graduate text in thermodynamics is written from a chemical engineer’s perspective. It has a strong mathematical development for both classical and statistical thermodynamics.

C. A. Truesdell, 1980, The Tragicomic History of Thermodynamics, 1822-1854, Springer Verlag, New York. This idiosyncratic monograph has a lucid description of the history of nineteenth century thermal science. It is written in an erudite fashion, and the reader who is willing to dive into a difficult subject will be rewarded for diligence by gain of many new insights.

C. A. Truesdell, 1984, Rational Thermodynamics, Second Edition, Springer-Verlag, New York. This is an update on the evolution of classical thermodynamics in the twentieth century. The book itself first appeared in 1969. The second edition includes additional contributions by some contempo- raneous leaders of the field.

S. R. Turns, 2000, An Introduction to Combustion, Second Edition, McGraw-Hill, Boston. This is a senior-level undergraduate text on combustion which uses many notions from thermodynam- ics of mixtures. It first appeared in 1996.

H. C. van Ness, 1983, Understanding Thermodynamics, Dover, New York. This is a short readable monograph from a chemical engineering perspective. It first appeared in 1969.

W. G. Vincenti and C. H. Kruger, 1976, Introduction to Physical Gas Dynamics, Krieger, Malabar, Florida. This graduate text on high speed non-equilibrium flows contains a good description of the interplay of classical and statistical mechanics. There is an emphasis on aerospace science and fundamental engineering applications. It first appeared in 1965.

289 F. M. White, 2010, Fluid Mechanics, Seventh Edition, McGraw-Hill, Boston. This standard undergraduate fluid text draws on thermodynamics in its presentation of the first law and in its treatment of compressible flows. It first appeared in 1979.

B. Woodcraft, 1851, The of Hero of Alexandria, London, Taylor Walton and Maberly. This is a translation and compilation from the ancient Greek of the work of Hero (10 A.D.-70 A.D.). The discussion contains descriptions of the engineering of a variety of technological devices including a primitive steam engine. Other devices which convert heat into work are described as well.

L. C. Woods, 1975, The Thermodynamics of Fluid Systems, Clarendon, Oxford. This graduate text gives a good, detailed survey of the thermodynamics of irreversible processes, especially related to fluid systems in which and diffusion play important roles.

290